Вы находитесь на странице: 1из 94

Progress in Materials Science 67 (2015) 194

Contents lists available at ScienceDirect

Progress in Materials Science


journal homepage: www.elsevier.com/locate/pmatsci

Nanoindentation in polymer nanocomposites


Ana M. Dez-Pascual a,, Marin A. Gmez-Fatou a, Fernando Ania b,
Araceli Flores b,*
a
Department of Polymer Physics, Elastomers and Energy Applications, Institute of Polymer Science and Technology (ICTP-CSIC),
Juan de la Cierva 3, 28006 Madrid, Spain
b
Department of Macromolecular Physics, Institute for Structure of Matter (IEM-CSIC), Serrano 119, 28006 Madrid, Spain

a r t i c l e i n f o a b s t r a c t

Article history: This article reviews recent literature on polymer nanocomposites


Received 12 June 2014 using advanced indentation techniques to evaluate the surface
Accepted 14 June 2014 mechanical properties down to the nanoscale level. Special empha-
Available online 17 July 2014
sis is placed on nanocomposites incorporating carbon-based
(nanotubes, graphene, nanodiamond) or inorganic (nanoclays,
Keywords:
spherical nanoparticles) nanollers. The current literature on
Nanoindentation
Polymer nanocomposites instrumented indentation provides apparently conicting informa-
Modulus tion on the synergistic effect of polymer nanocomposites on
Hardness mechanical properties. An effort has been done to gather informa-
Filler reinforcement tion from different sources to offer a clear picture of the state-of-
the-art in the eld. Nanoindentation is a most valuable tool for
the evaluation of the modulus, hardness and creep enhancements
upon incorporation of the ller. It is shown that thermoset, glassy
and semicrystalline matrices can exhibit distinct reinforcing mech-
anisms. The improvement of mechanical properties is found to
mainly depend on the nature of the ller and the dispersion and
interaction with the matrix. Other factors such as shape, dimen-
sions and degree of orientation of the nanoller, as well as matrix
morphology are discussed. A comparison between nanoindenta-
tion results and macroscopic properties is offered. Finally, indenta-
tion size effects are also critically examined. Challenges and future
perspectives in the application of depth-sensing instrumentation
to characterize mechanical properties of polymer nanocomposite
materials are suggested.
2014 Elsevier Ltd. All rights reserved.

Corresponding authors.
E-mail addresses: adiez@ictp.csic.es (A.M. Dez-Pascual), emaraceli@iem.cfmac.csic.es (A. Flores).

http://dx.doi.org/10.1016/j.pmatsci.2014.06.002
0079-6425/ 2014 Elsevier Ltd. All rights reserved.
2 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

Abbreviations

AAER acrylic electrophoretic resin


AFM atomic force microscopy
APTES aminopropyltriethoxysilane
CB carbon black
CCNT coiled carbon nanotube
CEAR cathodic electrophoretic acrylic resin
CF carbon bre
CNC cellulose nanocrystals
CNF carbon nanobre
COF coefcient of friction
CSH calcium silicate hydrate
CSM continuous stiffness measurement
CVD chemical vapour deposition
deox-ND decarboxylated nanodiamond
DMA dynamic mechanical analysis
DSC differential scanning calorimetry
DSI depth-sensing indentation
EG exfoliated graphite
EGO exfoliated graphene oxide
EGS exfoliated graphene sheets
EOGCNF exfoliated oxidized graphitized carbon nanobre
EPDM ethylene-propylene-diene rubber
FP-POSS uoropropyl polyhedral oligomeric silsequioxane
FG few-layer graphene
FGS functionalized graphene sheets
GBL gamma-butyrolactone
GF glass bre
Glymo glycidoxy-propyltrimethoxysilane
GNP graphene nanoplatelets
GO graphene oxide
GP graphite platelets
GPTS glycidoxypropyltrimethoxysilane
GS graphene sheets
HA hydroxyapatite
HDDA 1,6-hexanediol diacrylate
HDPE high density polyethylene
HEMA 2-hydroxyethyl methacrylate
IBMA isobornyl methacrylate
IF inorganic fullerene-like
ISE indentation size effect
KF kenaf bre
LbL layer-by-layer
MBZ methylbenzoate
MDMA monomeric dimethacrylates
MEH-PPV poly[2-methoxy-5-2(2-ethylhexyloxy-p-phenylenevinylene)]
MEMO (3-(trimethoxysilyl) propyl methacrylate
MMT montmorillonite
MPTES methacryloxypropyltriethoxysilane
MPTMS methacryloxypropyltrimethoxysilane
MWCNT multi-walled carbon nanotubes
NA non available
Na-MMT sodium montmorillonite
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 3

ND nanodiamond
ND-NH2 aminated nanodiamond
OC-POSS octa[(epoxycyclohexylethyl) dimethylsilyloxy]silsesquioxane
ODA octadecylamine
OH-POSS octa[(epoxyhexyl) dimethylsilyloxy]silsesquioxane
Oib-POSS octaisobutyl POSS
OM optical microscopy
OMMT organically modied montmorillonite
OP Oliver and Pharr
OTES octyltriethoxysilane
ox- oxidized
PA polyamide
PAA polyacrylic acid
PAH polyallylamine hydrochloride
PAM polyacryl amide
PAN polyacrylonitrile
PC polycarbonate
PDDA poly(dimethyldiallylammonium chloride)
PDMS polydimethylsiloxane
PE polyethylene
PEDOT-PSS poly(3,4-ethylenedioxythiophene)-poly(styrene sulphonate)
PEEK poly(ether ether ketone)
PEES poly(ether ether sulphone)
PEI polyethyleneimine
PEN polyethylene naphthalate
PEO polyethylene oxide
PET polyethylene terephthalate
PHB poly(3-hydroxybutyrate)
PHBV poly(3-hydroxybutyrate-co-3-hydroxyvalerate)
PHEA poly(2-hydroxyethyl acrylate)
PHO poly(3-hydroxyoctanoate)
PHPS perhydropolysilazane
PI polyimide
PLA polylactic acid
PLLA poly(L-lactic acid)
PLS polymer/layered silicate
PMA polymethacrylate
PMMA poly(methyl methacrylate)
PNIPAm poly(N-isopropylacrilamide)
POSS polyhedral oligomeric silsequioxane
POSS-OH allyl alcohol terminated POSS
PP polypropylene
PPC poly(propylene carbonate)
PP-g-MA polypropylene grafted-maleic anhydride
PPS polyphenylene sulphide
PPy polypyrrole
PS polystyrene
PSS polystyrene sulphonate
PSBA poly(styrene-co-butyl acrylate)
PU polyurethane
PUF poly(ureaformaldehyde)
PVA polyvinyl alcohol
PVC polyvinyl chloride
PVDF polyvinylidene uoride
4 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

PVK poly(9-vinyl carbazole)


QD quantum dot
rGO reduced graphene oxide
RT room temperature
S sonication
SAXS small angle X-ray scattering
SB sonication and ball milling
SC super critical solvent
SCCO2 supercritical carbon dioxide
SCF short carbon bre
SEM scanning electron microscopy
SEBS styreneethylene/butylenestyrene
(SiOEt)3 triethoxysilyl
SSE single-screw extrusion
SWCNT single-walled carbon nanotubes
TEM transmission electron microscopy
TEOS tetraethyl orthosilicate (or tetraetoxysilane)
TEG thermally expanded graphite
TGA thermogravimetric analysis
TMOS tetramethyl orthosilicate
TSE twin-screw extrusion
Tsp-POSS trisilanolphenyl POSS
UHMWPE ultra-high molecular weight polyethylene
UPE unsaturated polyester resin
UV ultra violet radiation
VGCNFs vapour-grown carbon nanobres
WAXS wide angle X-ray scattering

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2. Recent advances on instrumented indentation applied to polymer materials: methods and
developments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1. Depth-sensing indentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1.1. Determination of elastic modulus and hardness. Standard methods of analysis applied to
polymer materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1.2. DSI time-dependent properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.3. Indentation size effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2. Correlation mechanical properties-nanostructure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3. Comparison of indentation studies with bulk techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3. General aspects on polymer nanocomposites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4. Polymer nanocomposites incorporating carbon nanofillers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.1. Carbon nanotubes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.1.1. Synthesis methods and types of CNTs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.1.2. Preparation of polymer/CNT nanocomposites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.1.3. Modulus changes in polymer/CNT nanocomposites: comparison of thermoplastic and
thermoset matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.1.4. Hardness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.1.5. Indentation size effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.1.6. Creep . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.2. Graphene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.2.1. Modulus enhancements in graphene/polymer nanocomposites . . . . . . . . . . . . . . . . . . . 49
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 5

4.2.2. Comparison with macroscopic mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 50


4.2.3. Indentation size effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.2.4. Creep . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.3. Other organic nanofillers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.3.1. Modulus enhancements in organic/polymer nanocomposites . . . . . . . . . . . . . . . . . . . . . 53
4.3.2. Comparison with macroscopic mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.3.3. Indentation size effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.3.4. Creep . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.4. Main features on organic fillers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.4.1. Carbon nanotubes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.4.2. Graphene. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.4.3. Other organic fillers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5. Polymer nanocomposites incorporating inorganic nanofillers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.1. Layered silicates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.1.1. Characteristics, preparation methods and types of nanocomposites. . . . . . . . . . . . . . . . 59
5.1.2. Modulus and hardness changes in polymer/layered silicate nanocomposites . . . . . . . . 60
5.1.3. Other layered silicate/polymer systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.1.4. Comparison with macroscopic properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.1.5. Indentation size effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.1.6. Creep properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.2. Spherical inorganic nanoparticles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.2.1. Characteristics, preparation methods and types of nanocomposites. . . . . . . . . . . . . . . . 66
5.2.2. Modulus changes in nanocomposites containing spherical inorganic nanoparticles . . . 68
5.2.3. Indentation size effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.2.4. Creep properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.2.5. Comparison with macroscopic mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.3. Other inorganic nanofillers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5.4. Main features on inorganic nanofillers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.4.1. Layered silicates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.4.2. Spherical inorganic nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6. Comparison of the reinforcement effect of different nanofillers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.1. Effect of the type of nanofiller . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.2. Influence of the geometry of the filler and orientation effects. . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
6.3. Comparison of properties at the macro and the nanoscale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
7. Conclusions and future perspectives. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

1. Introduction

Depth-sensing indentation (DSI) represents nowadays one of the principal techniques for the mechan-
ical characterization of materials. The method monitors the penetration of an indenter into the material
surface during the application and release of a load [1]. At the present time, the most advanced DSI
instruments can produce indentations with depths of only a few tens of nanometers, most of them also
offering the possibility of approaching an upper limit in the micron regime [2]. The terminology DSI
denes the principle of measurement to attain indentation data. Usually, this technique is also referred
to as nanoindentation, even when penetration depths of a few microns are involved.
A typical DSI test includes a loading-hold-unloading cycle (see Fig. 1). Elastic contact consider-
ations are usually adopted to analyse load-depth curves. Most commonly, hardness, H, and quasi-sta-
tic elastic modulus, E, values are derived assuming linear elastic behaviour at the onset of unloading
[3,4]. Because the new generation of DSI testers involves the most advanced engineering approaches
to accurately record penetration depths in the nanoscale range, special care should be paid to calibra-
tion procedures. These are usually friendly integrated in the software developed for collecting the
indentation data. In addition, as DSI devices have sophisticated over the years to widen the range of
applications, important considerations related to instrumental artefacts and to data analysis are mat-
6 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

1.0

0.8
Pmax

s
Load [mN] 0.6

0.4

0.2

hp hmax
0.0
0 100 200 300 400 500
Displacement [nm]

Fig. 1. Load as a function of indentation depth for a semicrystalline PEEK sample (degree of crystallinity of 40%). Unpublished
results.

ters of active discussion within the indentation community. This is reected in the number of papers
fully or partially dedicated to the problems encountered in obtaining meaningful DSI data [57].
The application of DSI to polymers raises additional concerns [6,813]. On the one hand, linear elas-
ticity cannot be straightforward assumed and particular strategies need to be taken into account. On
the other hand, the calibration procedures should be carefully examined because some of them
require pure elasticplastic response. Finally, a comprehensive characterization of polymer materials
by means of DSI requires specic approaches to extract information on their time dependence.
A step forward in this direction has been produced with the introduction of dynamic DSI, or continu-
ous stiffness measurements (CSM) [14]. CSM superimposes an oscillatory force to the quasi-static
loading opening up the possibility of using DSI devices as micro or nanoscale dynamic mechanical
analysers.
DSI data are most frequently acquired using purpose-designed instrumentation. Alternatively,
Atomic Force Microscopy (AFM) can be used as a depth-sensing instrument provided that careful cal-
ibrations to achieve meaningful load and depth data are carried out [15,16]. The advantage of using
atomic force microscopy for indentation purposes is the higher displacement sensitivity and superior
imaging facilities. In contrast, compared to instrumented indentation, the technique is more liable to
instrumental errors and miscalibrations, especially in force modulation experiments [16].
DSI is being progressively incorporated as a routine technique for mechanical characterization of
polymer materials and yet a number of decisions concerning the device calibration, the test procedure
and the test parameters are of great importance and should be critically examined by an experienced
researcher. The attractiveness of instrumented indentation relies on the ability to extract mechanical
properties from a small local deformation. This is extremely valuable for systems that are only avail-
able in small amounts, or those with limited dimensionality (thin lms and coatings). In addition, DSI
allows spatially resolving the mechanical properties and this is of great importance for heterogeneous
materials such as polymer nanocomposites.
Polymer-based composites incorporating nanoscaled llers have attracted much attention over
recent years owing to their unique mechanical, thermal and electrical properties that are difcult to
achieve using conventional llers [17,18]. This superior performance combined with their low density
make them suitable for a broad range of technological sectors such as telecommunications, electronics
and transport industries, especially for aeronautic and aerospace applications where the reduction of
weight is crucial to reduce fuel consumption [19]. The eld of polymer nanocomposites has evolved
very quickly since the last decade being layered silicates, carbon nanotubes (CNTs) and very recently
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 7

graphene the most widely used and highly successful examples of nanollers incorporated into poly-
mer matrices. However, nanoller aggregation has been found to hamper mechanical property
improvements. Consequently, a great effort has been devoted to establish the most suitable conditions
for the transfer of mechanical load from the matrix to the nano-reinforcement. A prerequisite for such
an endeavour is the homogeneous distribution of the nanollers and the establishment of a strong
afnity with the surrounding polymer matrix.
The application of DSI to polymer nanocomposites has gained increasing interest in recent years.
The technique has been proved to be sensitive to ller content, ller dispersion, as well as to the inter-
facial nanoller-matrix adhesion [2022]. Information on heterogeneities of the composite material,
either across the thickness or along the surface arising as a consequence of changes in the matrix mor-
phology or uneven distribution of the ller, can be readily detected by means of DSI [23,24].
The present article reviews the most relevant contributions concerning the application of DSI to
polymer-based nanocomposites, with special emphasis in those containing carbon-based (e.g. nano-
tubes, graphene, nanodiamond) or inorganic (e.g. nanoclays, spherical nanoparticles) nanollers.
The rst part of the review introduces the DSI method and summarizes the most important
approaches developed to derive mechanical data in case of materials with time dependent properties
such as polymers. A number of specic recommendations for extracting meaningful mechanical data
will be offered to the reader. The central part of the review presents the state-of-the-art in the appli-
cation of DSI to polymer nanocomposites. This technique has been frequently used to evaluate the
reinforcing effect of different kinds of nanollers. Results from different laboratories in similar rein-
forced polymers often seem to be at variance. An effort has been done to present a comprehensive
understanding of the inuence of multiple factors to the mechanical enhancement. Finally, challenges
and future perspectives in the application of DSI to polymer nanocomposites will be addressed.

2. Recent advances on instrumented indentation applied to polymer materials: methods and


developments

2.1. Depth-sensing indentation

The usual aim of DSI experiments is to determine the material mechanical properties from load-
indentation depth data. A characteristic DSI test includes a load-hold-unload sequence as that shown
in Fig. 1. However, other methods including partial unloading, reloading, etc. are sometimes
employed. In order to optimize the data collection, certain test parameters such as the approach dis-
tance, the allowable drift rate or the criteria to decide whether the indenter has contacted the surface
should be chosen. Pyramid indenters are most commonly adopted for DSI among the broad variety of
indenter tips because of their geometrical self-similarity and improved spatial resolution. Berkovich
indenters are preferred to the Vickers geometry because the latter is prone to undesired roof tip
imperfections. Spheres and at-ended cylindrical punches are particularly useful when the viscoelas-
tic character of the material needs to be emphasized (see Section 2.1.2 on time-dependent properties).
The use of at punches offers additional advantages because this geometry eliminates uncertainties in
the calculation of the contact area.

2.1.1. Determination of elastic modulus and hardness. Standard methods of analysis applied to polymer
materials
2.1.1.1. The compliance method. The methods commonly adopted to analyse load-depth data from DSI
are based on elastic contact theories, rst developed for spherical indenters by Hertz, extended to
other indenter geometries by Sneddon and nally applied to the analysis of DSI curves by Doerner
and Nix and later improved by Oliver and Pharr [24]. Linear elastic response is a major requirement
to allow using any of these fundamental equations or protocols. For shallow indents, meaningful mod-
ulus data have been achieved assuming linear elasticity during loading [15]. However, the procedure
most widely employed relies on the analysis of the initial portion of the unloading curve and is com-
monly known as the compliance method [3,4]. In this procedure, the following relationship between
the contact stiffness S, dP/dh, the reduced modulus Er, and the projected contact area A, is proposed:
8 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

p
p S
Er p 1
2b A
where
 
1 1  m2 1  m2i
2
Er E Ei
E and m are the elastic modulus and Poissons ratio of the material respectively and Ei and mi those of
the indenter. In Doerner and Nix analysis, linear unloading is assumed and S is calculated from the t
of a tangent to the unloading curve at maximum load Pmax (see Fig. 1). The extrapolation to zero load is
taken as the plastic depth, hp, i.e. the depth of the indenter in contact with the sample at the onset of
2
unloading. In case of a Berkovich indenter, A is related to hp through: A 24:5 hp . Oliver and Pharr
extended this procedure to account for the curvature in the unloading data. A power law was used
to t the unloading curve and the derivative at maximum load was used to achieve S. The projected
2
contact area A was obtained using A 24:5 hc , with:
hc hmax  Pmax =S 3
where hmax is the maximum displacement (see Fig. 1). The constant is often taken as = 0.75 for a
Berkovich indenter [25] although a rigorous study would require selecting the value as a function
of the power law exponent deduced from the analysis of the unloading curve [26]. The b parameter
in Eq. (1) is a correction factor. A thorough discussion on the preferred b values that should be adopted
for each type of the indenter is given in Refs. [7,25]. In case of a Berkovich indenter, any value in the
range 1.02261.085 can be taken, probably being b = 1.05 the best choice. Once the contact area is
determined, the hardness can be derived using:
Pmax
H 4
A
Oliver and Pharrs method (OP) is nowadays extensively used for the analysis of DSI data and most
commercial instruments have incorporated the procedure into the software. In elasticplastic materi-
als with limited pile-up, the method has been shown to be most successful provided the appropriate
calibrations are carried out (mainly machine compliance, tip area function and thermal drift).

2.1.1.2. Calibrations procedures. The most widespread method for calibration of machine compliance
and tip area function was initially proposed and further implemented by Oliver and Pharr [4,25].
The methodology uses the indentation data of a reference sample that does not pile up (for example,
fused silica) and relies on the assumption that H and E are independent of the penetration depth.
Alternatively, the indenter shape area function can be achieved using material-independent methods
such as AFM imaging [6]. Because the material-dependent procedure for tip calibration is usually
friendly integrated in the DSI software, this methodology is usually preferred to the material-indepen-
dent ones. Some authors have suggested that the tip calibration constructed upon indentation on a
hard material is not appropriate for soft materials such as polymers. Briscoe et al. [13] suggested that
Hertzian-like deformations of the indenter tip when contacting hard materials may not resemble the
contact situation with softer surfaces and proposed that the indenter tip area shape for polymer mate-
rials should be calibrated against the properties of a polymer reference sample. The concept that some
interaction between the tip and the indenter and/or an inadequate detection of the polymer surface
produce an apparent increase of E and H values lies behind the work of Loubet et al. [8] who developed
an alternative approach to calibrate the indenter-polymer contact area at shallow penetrations by
introducing an apparent tip-defect. Flores and Balt-Calleja [10] adopted a similar assumption of con-
stant E, H values as a function of indentation depth for a glassy polymer and used an alternative
approach to correct for the tip defect by extrapolating the shape of the load-depth curves at large
penetrations (>2 lm) to shallow penetrations. The consequence of this correction is clearly appreci-
ated in Fig. 2 where the hardness values of glassy poly(ethylene terephthalate) (PET), before and after
the tip defect correction, are represented as a function of indentation depth. In this work, the OP
method was used on the derivation of H. Nowadays, a common procedure adopted when polymer
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 9

210

H [MPa] 180

150

0 2 4 6 8 10
h [ m]

Fig. 2. Hardness values for glassy PET as a function of indenter displacement. The OP method was used to determine H values
with ( ) and without ( ) application of the tip defect correction. Adapted from Ref. [10].

materials are tested at shallow depths is to calibrate for the tip defect using Oliver and Pharrs
approach [4,25] on a polymer standard material such as polycarbonate (PC) or poly(methyl methac-
rylate) (PMMA) [27,28].
The thermal expansion of both, the material and the equipment, produces an increase in displace-
ment during the course of the indentation test that overlaps with the genuine response of the material.
To correct for this thermal drift, a new hold period nearly at the end of the unloading (typically 10% of
Pmax) that monitors the changes in depth is usually introduced in the indentation test [29]. This hold
period is illustrated in the load-depth plot of Fig. 1 for a poly(ether ether ketone) (PEEK) sample
(P  0.08 mN). Conversely to elasticplastic materials, the indentation depth signicantly decreases
while maintaining the small load nearly at the end of the test. This is due to the characteristic visco-
elastic nature of polymer materials. It is clear that the contribution of thermal drift to the measured
displacement cannot be easily evaluated for materials with time-dependent properties. A criterion to
achieve negligible thermal drift effects on the calculate modulus based on the duration of the unload-
ing period in relation to the unloading rate, stiffness and contact depth has been proposed [30]. In
practice, the procedure most commonly adopted is to minimize thermal drift effects by choosing short
indentation cycles and allowing thermal equilibrium before the experiment.

2.1.1.3. Application of the OP method to polymers. The application of the OP method to polymer mate-
rials encounters certain difculties as was soon realized by Oliver and Pharr themselves who stated in
their original paper that displacements recovered during rst unloading may not be entirely elastic,
and because of this, the use of the rst unloading curves in the analysis of elastic properties can some-
times lead to inaccuracies. Indeed, this is nowadays widely accepted in the indentation community
[6,813]. However, it is frequently considered that the OP method can be applied if there is no increase
in displacement upon load release in the form of a nose. This nose is usually avoided by introducing
a hold period at peak load in combination with a rapid unloading rate. However, whether creep inu-
ences the condition of pure elastic unloading even in the case that long holding times and high unload-
ing rates are used is still a matter of concern [10,11]. In addition, it has been shown that the criterion
adopted for the analysis of the unloading data (the portion of the unloading curve that is used for the
curve tting) can also inuence the stiffness values [6]. As an example, Fig. 3 shows the variation of
the elastic modulus values of a glassy polymer sample as a function of the loading/unloading time
for two different holding periods. It is clearly seen that the E values increase with increasing unloading
time. This effect is due to creep contributing to the total displacement, especially at long loading/
10 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

3.5

3.0
E [GPa] thold = 6 s

thold = 200 s
2.5

2.0
0 2 4 6
tload [min]

Fig. 3. Variation of the elastic modulus values of a glassy PET sample as a function of the loading time to Pmax for two different
holding periods: ( ) 6 s, ( ) 200 s. Adapted from Ref. [10].

unloading times and short holding periods. To minimize the effect of creep on the calculation of the
unloading stiffness, Loubet et al. [8] proposed that elastic unloading can be assumed if the strain rate
at the beginning of unloading is 10 times greater than the strain rate measured at the end of the hold-
ing time. A more elaborated approach was developed by Feng and Ngan [30] who proposed a proce-
dure to correct the apparent contact stiffness data for creep and thermal drift based on the
measurement of the total displacement rate at the end of the hold period.

2.1.2. DSI time-dependent properties


2.1.2.1. Creep analysis. Worthy efforts have been done to characterize the indentation response of
polymers with time using creep or load relaxation studies [2,3138]. Creep compliance (or relaxation
modulus) has been frequently employed as a quantitative measure of the capacity of the material to
ow. Elasticviscoelastic correspondence has been commonly adopted to generate solutions for the
creep compliance using the governing equations of a rigid body into an elastic medium [3138]. Some
of these studies describe the creep indentation behaviour in terms of a spring-dashpot mechanical
model with the ultimate goal of tting the model to the experimental data and relate the tted param-
eters to the material properties (typically two elastic moduli and a viscosity coefcient) [31,3438].
Because the above approaches assume linear viscoelasticity, maintaining the indentation response
within this regime has been a major concern [3234]. In this respect, spherical or cylindrical tip geom-
etries are preferred to conical or pyramidal ones [32,39]. A generalized model of viscoelasticplastic
behaviour has been proposed by introducing an additional plastic element in the dashpot-spring
model of viscoelasticity [40,41].
Creep by DSI has also been computed using a steady-state approach equivalent to that employed in
uniaxial testing describing the strain rate as a power law of the stress. The method makes use of:

e_ / Hn 5
Here, e_ is the strain rate dened as the instantaneous displacement rate divided by the instantaneous
displacement and n is the creep exponent. Strain rate and hardness values are usually recorded either
in the hold period at maximum load or by performing a series of loading ramps at a constant strain
rate [42,43]. The two methods differ in an important question: in the former procedure a wealth of
data is computed during the hold period where the hardness and the strain rate are continuously
changing as the test proceeds, while in the constant e_ ramping, one H value is obtained in steady-state
conditions for each test. In turn, n values can be achieved in a loglog plot of e_ vs. H. Constant
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 11

strain-rate ramping yields higher n values than those achieved from the hold experiments, being the
former in close agreement with those derived from uniaxial testing [42].
Alternatively, a simple way to compare the time-dependent deformation of different polymer
materials is to directly t the variation of depth as a function of time during the hold period to a simple
equation of the form [44,45]:
h A
lnBt 1 6
h0 h0
The h(0) value denes the initial depth at the beginning of the hold period. The dimensionless param-
eter A/h(0) is known as the creep strain rate sensitivity and denes the extent of the time-dependent
deformation while the B parameter in Eq. (6) is interpreted as the rate term.

2.1.2.2. Dynamic indentation testing. Dynamic indentation testing, also known as continuous stiffness
measurements (CSM) [4,14], represents an alternative route to study the time dependent behaviour of
polymers. In addition, CSM represents an improved method for surface detection, which is fairly crit-
ical for polymer materials. In dynamic DSI, a sinusoidal force is superimposed to the quasi-static load-
ing allowing for a direct measurement of the harmonic contact stiffness as a continuous function of
depth. This raises two important consequences. Firstly, a prole of the mechanical properties as a
function of indentation displacement in a single loading cycle can be achieved. Secondly, dynamic
DSI can operate as a nanodynamic mechanical analyser (nanoDMA) yielding values of the storage
modulus E0 and loss modulus E00 as a function of the frequency of the harmonic force. It is well known
that for a viscoelastic material the complex modulus:
00
E E0 iE 7
where

E0 E cos d 8
and

E00 E sin d 9
being d the phase shift between the complex stress and the strain and E the usual quasi-static
modulus.
It is important to note that often, in the DSI literature, no clear distinction is made between quasi-
static indentation modulus E and dynamic indentation modulus E0 , being both represented by E. In this
review, the symbol EI will describe a general indentation modulus, either static or dynamic, while E
and E0 will keep their exact meaning.
In order to extract contact stiffness measurements from dynamic DSI, the instrument-sample sys-
tem is usually represented as a mechanical model of springs and dashpots [2,4651]. Appropriate
elements are introduced in the mechanical model depending on the characteristics of the indenta-
tion instrument and the tip-sample interaction [2]. The latter is usually modelled as a simple Voigt
solid of a spring and dashpot in parallel [4649]. In this case, the materials dynamic stiffness K and
contact damping Ds can be directly related to the storage modulus and loss modulus values
following:
p
p K
E0r p 10
2b
A
p
00 p xDs
Er p 11
2b A
where x is the angular excitation frequency. K and Ds can be calculated from experimental parameters
obtained from dynamic indentation testing, after subtracting the contribution of the instrument to the
total measured response [2,47,49]. The projected area of contact A for a Berkovich indenter is usually
determined in a similar manner as that for the quasi-static experiments, i.e., using the hc values
dened in Eq. (3) where S is now the dynamic contact stiffness K. Cylindrical punches are
12 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

4
E [GPa]

400

300
H [MPa]

200

100

0
0 500 1000 1500 2000
Displacement, h [nm]
0
Fig. 4. E and H values as a function of displacement into surface for a semicrystalline PEEK sample (degree of crystallinity of
40%). Unpublished results.

recommended in this type of experiments because having a constant contact area value at all depths
(A = pa2 where a is the radius of the punch), the lack of accuracy introduced in the determination of A
is excluded. The area of contact can also be used to derive hardness values following Eq. (4). Fig. 4
shows, as an example, the plot of elastic modulus (E0 ) and hardness values obtained following the pro-
cedure above described as a function of displacement into surface for a PEEK sample using a Berkovich
indenter.
Valuable work has been done to develop more sophisticated models that incorporate the material
as a standard linear solid [50,51]. An additional spring is introduced into the Voigt model to capture
instantaneous elastic behaviour. The solutions are more complex than for the above conventional
analysis; however, once the elements of the standard linear solid system are determined, the sample
storage and loss moduli can be predicted as a function of frequency. In addition, more accurate values
of the sample stiffnesses are claimed after careful subtraction of the instrument contribution than can
be relevant in the case of stiff materials [51].
A recent paper has reported a comparison between storage and loss modulus values obtained
by means of dynamic DSI and indentation creep using sharp indenters [36]. Interestingly, the E0
values were in close agreement with each other for small applied forces and were also found to be
in the range of those determined by means of DMA, E0DMA . However, the E00 values obtained by means
of the different methods exhibited clear disparities [36]. The authors addressed an important
observation, i.e. dynamic DSI measures recoverable viscoelastic deformation whereas indentation
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 13

creep experiments relies on the measurement of the total displacement that generally includes plas-
ticity, viscoplasticity, as well as elasticity and viscoelasticity. Hence, in the latter case it is especially
difcult to maintain the conditions of testing within the linear viscoelastic regime and the use of
geometries of the indenter that can maintain the elastic contact over a wide force range is especially
recommended.

2.1.3. Indentation size effect


2.1.3.1. Bulk materials. Early research in metals and ceramics using geometrically self-similar pyramid
indenters in conventional hardness testers reported an increase in hardness with decreasing penetra-
tion depth [52]. This observation is commonly known as indentation size effect (ISE) and has been
extensively investigated since. The ISE has been attributed to diverse sources, the most common ones
related to the indenter-sample contact (frictional forces at the interface) [52] and to mechanisms asso-
ciated to fundamental material processes [53]. With the advent of DSI testers, the study of the inden-
tation size effect has been extended to shallower penetration depths. Because this modern
instrumentation requires a careful calibration and the data can be inuenced by many factors, special
care should be taken to separate experimental artefacts or effects arising from the testing conditions
from an ISE associated to the genuine material properties [53].
Among the number of artefacts that could introduce an apparent increase of hardness with
decreasing penetration depth, an inappropriate area function calibration is the most recognized
one. In this case, not only H but also EI values rise at shallow penetrations. If the rounding of the inden-
ter tip is underestimated, the calculated contact area is smaller than the real one and an ISE will be
apparent [53]. Tip blunting can produce an additional depth-dependence due to the transition from
the deformation mode of a sphere to that of a pyramid [7,53]. In addition, an incorrect determination
of the point of initial contact towards higher displacement values can also lead to an ISE [53]. In gen-
eral, for a correct evaluation of the mechanical properties at shallow depths, a critical examination of
all the possible effects that can introduce inaccuracies in the analysis of EI and H is needed. An example
is the case of surface roughness that becomes especially signicant in polymer materials due to the
difculty in producing a surface nish with low roughness [27].
A number of models based on strain gradient plasticity have been proposed to explain the hardness
increase with decreasing indentation size [54,55]. Nanoindentation data in two polymer glasses were
found to be well described by a strain gradient plasticity model based on a molecular kinking mech-
anism [54]. Gao et al. [55] extended the classical expanding cavity model to account for strain gradient
effects. The model incorporates strain hardening characteristics of the material and arrived to an
indentation size dependence of hardness that was validated against data for a soft metal.
In spite of the well-known hardness dependence on strain rate for polymer materials [13], this is
frequently overlooked in the ISE interpretation. It is of great importance to highlight that indentation
experiments using a constant loading rate (trapezoidal load function) will normally result in an ISE
because the indentation strain rate during the loading segment continuously decreases with increas-
ing displacement [8,42,56]. The study of the near-surface mechanical properties in polymer materials
can only be appropriately approached if the same constant indentation strain rate is applied at all pen-
etration depths. For pyramidal indenters, this can be achieved by controlling that the instantaneous
loading rate, P0 = dP/dt, divided by the instantaneous load, P, remains constant [56].
Most of commercial advanced DSI devices have incorporated the possibility of applying a constant
strain rate during the loading cycle. The combination of this option with the application of a small
oscillating force provides a rational route to obtain H (and E0 ) values at all penetration depths. Finally,
it is noteworthy that a recent comprehensive study of the depth dependence of indentation data using
a dynamic method (CSM) has revealed that this methodology can induce a signicant error in H and E0
at small depths [5] in such a way that it has even not been recommended for the characterization of
the ISE [53]. In the case of a soft metal such as copper (Er/H  225), misleading hardness and modulus
values can be found for depths below 300 nm [5]. Reducing the amplitude of the displacement oscil-
lation shifted the limit for reliable data to shallower depths. Although this source of error is minimized
for materials with lower Er/H ratio (polymers typically exhibit Er/H  10), the study of ISE in polymers
using a dynamic method should be carried out with some caution, using small amplitudes of oscilla-
tion and allowing them to vary in order to discard possible artefacts.
14 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

All the above considerations help to understand why detecting a genuine gradient of mechanical
properties across the sample thickness of a polymer or identifying a size-dependent mechanism of
deformation is quite a difcult task. Indeed, most indentation studies do not elucidate whether ISE
is related to instrumental conditions or to the distinct properties of the surface. In spite of all these
problems, there are some useful hints on the indentation size effect in polymers that can provide valu-
able information on the variation of the mechanical properties across the polymer sample thickness
related to morphological or structural changes in the material. For example, localized oxidation, age-
ing or modication of the near-to-surface layers during fabrication have been suggested as mecha-
nisms yielding an increase in H and EI at shallow depths [13]. Higher degrees of crystallinity at the
near-surface region have been also proposed to account for the enhanced hardness and modulus val-
ues for low displacements found for semicrystalline polymer samples [11]. In the case of composite
materials, constant H and EI values found with increasing indentation depth have been interpreted
in terms of a good dispersion of the ller in the matrix [57]. In injection-moulded parts, an initial hard-
ness decrease with increasing indentation depth followed by a hardness increase above 1 lm has
been discussed on the basis of uncertainties at the very surface layer and an inhomogeneous distribu-
tion of the crystalline morphology, respectively [58].

2.1.3.2. Thin lms. In the particular case of thin lms, it should be borne in mind the possible inuence
of the substrate on depth sensing indentation results. This inuence often takes the form of an ISE
(normal o reverse) with increasing penetration depth. To minimize the problem, a rule of thumb com-
ing from the early days of microindentation and without any sound physical basis, proposes a limit of
10% of the lm thickness for the penetration depth. From this point of view, the use of blunt indenters
such as at-ended cylindrical punches is in principle recommended. However, as lm thickness
becomes increasingly thin and/or for a large elastic modulus mismatch, substrate effects cannot be
avoided and a number of models have been developed to extract the intrinsic elastic modulus of thin
lms from the composite lm/substrate modulus value [3,5965].
Doerner and Nix in 1986 [3] were the rst to empirically model the inuence of the substrate on
the elastic measurements assuming a rigid at-ended cylindrical punch that indents a homogeneous
half space. They proposed an equation for the compliance (reciprocal of S in Eq. (1)), adding two expo-
nential terms depending on the relative indentation depth (depth divided by lm thickness) which
include a constant to be determined empirically for each particular case. Soon afterwards, numerical
calculations performed by King were able to give values for the constant for different lm-substrate
systems and at-ended indenter geometries [59]. The method of King was further extended by Saha
and Nix to a pyramidal indenter, by assuming that a at punch is located at the tip of a Berkovich
indenter [60]. Gao et al. also modelled the unloading process of a rigid indenter penetrating a thin
lm/substrate by a rigid cylindrical punch in frictionless contact with a layered elastic half space
[61]. However, their analysis differs from previous studies in conceiving a rst order rigorous mod-
uli-perturbation method to derive a closed-form solution for the contact compliance of the composite
system. The effective compliance is expressed as a function of lm and substrate Poissons ratios and
shear moduli. The method was proved to be valid for a limited range of shear moduli ratios between
lm and substrate and Poissons ratios. The SongPharrs model can be considered as an extension of
the work of Gao et al. modied to increase the applicable range of modulus mismatch between the
lm and substrate [62]. By using a reduced lm thickness, subtracting the contact depth to the lm
thickness, a good agreement between the model and experimental and numerical data was found.
Menck et al. also proposed some modications to the Gaos approach and compared it with other dif-
ferent approximation functions [63]. Bec et al. proposed an analytical model based on the indentation,
by a rigid cylindrical punch, of a homogeneous lm deposited onto a substrate considered as a semi-
innite half space [64]. The composite system was simply modelled by two springs connected in ser-
ies. The lm reduced modulus can be calculated from the apparent modulus if the reduced Youngs
modulus of the substrate and the lm thickness are known.
According to Hay and Crawford [65], some of the previous models work reasonably well for com-
pliant lms on stiff substrates but are not so adequate if the lm modulus is more than twice the sub-
strate modulus. They put forward an elastic lm/substrate model where the lm is allowed to act in
series and in parallel with the substrate. The model has been found to work well for both compliant
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 15

lms on stiff substrates and vice versa and yields accurate modulus values for indentation depths up
to 25% of lm thickness.
From a practical point of view, an alternate approach to the above mentioned models is described
in ISO standard 14577-4 [66]. It species a nanoindentation testing method which should be suitable
for thin coatings of all kinds of materials. The substrate independent elastic modulus should be deter-
mined by measuring the elastic modulus according to the OP method as a function of penetration
depth and then linearly extrapolating the data to zero displacement, where the inuence of the sub-
strate should be negligible.
Compared to the elastic moduli, hardness data of a thin lm are less affected by the substrate. This
is to be expected because the elastic eld under the indenter is a long-range eld that extends well
beyond the plastic one [60]. The HanYuVlassak method [67] has been recently reported as an
improved method for determining the hardness of thin lm/substrate systems based on measure-
ments of the contact stiffness as a function of indentation depth, provided that the elastic properties
of the lm and substrate are known. The method overcomes the limitations, at large indentation
depths, of the Saha and Nix model arising from the at punch assumption. To calculate the lm mod-
ulus, the OP analysis, followed by extrapolating down to zero indentation depth by means of a poly-
nomial t, is used. This procedure, in addition to eliminating substrate effects, helps to minimize
problems arising from surface roughness and/or pile-up effects.

2.2. Correlation mechanical properties-nanostructure

It is well known that the mechanical properties of a polymer material can be tuned by changing the
internal nanostructure. Establishing correlations between indentation magnitudes (typically H and EI)
and characteristic nanostructural parameters of a polymer material (such as, for example, crystallinity
and crystal thickness) is of great importance to control the material properties. DSI studies on poly-
mers have been mainly focused on the method of analysis of load-depth data and there is limited lit-
erature available on the mechanical properties-nanostructure correlations. However, there are well-
established empirical models to explain the hardness dependence on the polymer nanostructure;
the hardness numbers traditionally arise from conventional indentation testing based on the optical
measurement of the residual indentation. These models have been topic of research for many decades.
The additivity law describing the hardness of a semicrystalline polymer material in terms of that of the
crystalline and of the amorphous regions, Hc and Ha respectively, has been extensively validated and is
expressed as [68,69]:

H Hc v c Ha 1  v c 12

where vc represents the volume fraction of crystalline material. An analogous additivity law for the
modulus values extracted from DSI has not been sufciently contrasted and only some indications that
EI could follow a parallel model with vc, as H does, have been proposed [70]. In addition, the Hc values
have been also shown to be inuenced by the crystal thickness, lc, through:

H1
c
Hc 13
1 lbc

Here, H1c is the hardness of an innitely thick crystal and b is a parameter related to the ratio between
the surface free energy of the crystals, re, and the energy required to plastically deform them through
a number of shearing planes, Dh (b = 2re/Dh). H1 c is mainly related to the chain packing density within
the crystals and represents the upper limit to the hardness of a polymer material.
As an example, Fig. 5 shows the hardness dependence with vc and lc for PEEK [71]. The glassy mate-
rial displays the lowest H value (135 MPa) while the semicrystalline samples exhibit H values that
tend to increase with increasing degree of crystallinity. However, the H changes cannot be only attrib-
uted to variations in vc because lc is concurrently changing. A colour code has been used in Fig. 5 to
identify each group of data associated to a range of lc values. The straight lines follow the additive
model of Eq. (12) with Ha = 135 MPa. Increasing crystal thickness values yield straight lines with
increasing slope (and hence, higher right-hand y-axis intercepts, Hc). One can now clearly perceive
16 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

nm
400
- 4n m
=3

6
2n

5-
lc
lc =

=
lc
H [MPa]
300

200

Glass
100
0.0 0.2 0.4 0.6 0.8 1.0
vc

Fig. 5. Hardness dependence with vc for PEEK. The straight lines follow the hardness additivity law (Eq. (12)) with
Ha = 135 MPa. Each colour denes a group of data with similar crystal thickness values, lc. Note that increasing the crystal
thickness yields higher Hc values (right-hand y-axis intercepts). Data taken from Ref. [71], copyright 1991, with permission from
Springer Science and Business Media.

the separate inuence of vc and lc on the H values. Moving along a straight line denes the hardness
increase due to changes in the degree of crystallinity while jumping from one straight line to
another one of higher slope explains the hardness increase due to crystal thickening.
The hardness additivity law of Eq. (12) has been proved to be successful in describing the resistance
to plastic deformation of a polymer material in terms of the separate contribution of stiff and soft com-
pliant elements. At the heart of this success is the connectivity between the crystalline and amorphous
regions through tie molecules, interconnecting chains, etc.
Application of this concept to multicomponent materials such as polymer composites would
depend to a large extent on the connectivity between the different phases. In reinforced polymer
matrices, the ller-matrix interface that is responsible of the adequate load transfer across the mate-
rial is a critical issue that will certainly inuence the validity of any model predicting the properties of
the composite in terms of those of the constituents.
Over one hundred years ago, there was a raising need of estimating the effective elastic moduli of
heterogeneous materials. The VoigtReuss rule of mixtures established upper and lower bounds for
the modulus of a composite material; pure elastic behaviour was assumed at constant strain (Voigt)
or stress (Reuss) under uniaxial loading. In case of a two-phase composite material, both approaches
can be expressed according to:

Ec v m Em v f Ef Voigt : upper bound; parallel model 14


E1
c v 1
m Em v 1
f Ef Reuss : lower bound; normal model 15

where v represents volume fractions and the subscripts c, m and f correspond to composite, matrix and
ller, respectively.
Using the elasticviscoelastic correspondence principle, the expressions (14) and (15) could also
hold for viscoelastic materials by changing the modulus values by their complex counterparts,
Ec ; Em and Ef [72]. Other micromechanics additivity models are: the HashinShtrikman formulae
[73] and the HalpinTsai model [74]. The former were derived by applying variational principles of
linear elasticity and offer tighter bounds for the effective modulus values of isotropic composites than
those of the VoigtReuss limits. The HalpinTsai equations are based on a rigorous self-consistent
method which models the composite as a single bre, surrounded by a cylinder of matrix and the
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 17

ensemble embedded in an unbounded homogeneous medium. These equations account for a wide
variety of reinforcement geometries through an empirically adjustable parameter. The above models
[73,74] have been used in a number of occasions to describe the modulus values of polymer-based
composites obtained by means of indentation techniques [7578].

2.3. Comparison of indentation studies with bulk techniques

Comparing modulus data obtained from DSI with those achieved by conventional macroscopic
techniques is quite a challenging task. In addition to the sources of error that can be introduced
in DSI experiments concerning instrument calibration, other factors such as strain hardening, inter-
action forces between the indenter and the sample, ISE or material heterogeneities could contribute
to the discrepancies found between nanoindentation and traditional mechanical methods. Also
important are the differences in the principle of measurement and the testing geometry, some
of them are: (i) in DSI, the volume of deformation is continuously changing, as it is the inhomo-
geneous distribution of stresses and strains generated beneath the indenter; this clearly differenti-
ates DSI from other macroscopic techniques such as tensile testing. Indeed, the concept of an
effective indentation strain rate enables application of well-established uniaxial tensile analyses
such as power-law creep [42]. (ii) The volume loaded is signicantly smaller in DSI. In addition,
for dynamic testing, the different amplitudes of oscillation between DMA and DSI (a few microm-
eters in DMA vs. a few nm in DSI) has been suggested to give rise to different strain levels that
could activate distinct mechanisms of deformation [36]; and (iii) the load direction is believed
to evolve radially from the point of rst contact in case of indentation testing while for macro-
scopic tensile testing it is unidirectional. In the latter technique, mainly tensile stresses are applied
while in DSI a combination of compressive, tensile and shear forces are exerted on the material. For
nanoller-reinforced systems, the disparity between results from the different techniques is
expected to be even larger because in uniaxial testing and in contrast to DSI, the longer ller
dimension would preferentially orient along the loading direction enhancing the overall material
reinforcement.
In spite of the above described limitations, it is rather frequent to attempt a comparison between
DSI and conventional macroscopic data. However, one should bear in mind the phenomenological
character of such correspondence. In the early stages of application of dynamic DSI (nanoDMA)
to polymers, DMA data were frequently taken as a reference to validate dynamic indentation mea-
surements because the latter method assumes certain models and analogies (such as linear visco-
elasticity and elasticviscoelastic correspondence) that are not straightforward for a polymer
material under standard testing conditions. A number of reports suggest that the modulus values
E0 obtained from depth-sensing instrumentation are in fair agreement with those obtained by
DMA or uniaxial compression with harmonic oscillation [47,48,79,80]. However, a signicant dis-
parity between E0 and E0DMA data has also been reported [20,22,76,81]. In many cases, the modulus
values measured by dynamic nanoindentation have been found to be higher than those obtained
from DMA. The origin of the discrepancy between dynamic indentation and conventional bulk mea-
surements with harmonic oscillation is still open to debate. In addition to the differences in the test
geometry and operating mode mentioned above, it seems that comparison of data from the differ-
ent techniques is only valid if indentation data remain in the linear viscoelastic regime. However,
this is often overlooked and still indentation measurements have been found to be in reasonable
agreement with DMA. Interestingly, VanLandingham et al. [33] investigated a number of polymer
materials by means of traditional dynamic mechanical analysis and dynamic nanoindentation and
found an excellent agreement for the stiffer polymers while divergent results were found for the
more compliant polymer samples. The authors argued that the current DSI calculation methods
might be appropriate for materials where energy storage dominates the dynamic response but
might not be able to capture the response of materials with a pronounced viscous or rubbery ow
character.
In summary, although there is a technical similarity between dynamic DSI and DMA, neither the
stresses applied and strain levels achieved, nor the volume and modes of deformation, amongst other
factors, are identical and hence a perfect match of results cannot be expected.
18 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

3. General aspects on polymer nanocomposites

The optimization of the properties of polymer nanocomposites depends to a great extent on the
interaction between the matrix and the ller [18]. The state of dispersion of the ller and the nature
of the interface/interphase with the host matrix are the main two factors accounting for the interac-
tion between ller and matrix, and in turn, compromising the performance of the nanocomposite.
Other important parameters such as the aspect ratio and orientation of the ller or the changes trig-
gered in the matrix morphology upon addition of the reinforcement can signicantly inuence the
composite properties. Microscopic techniques such as scanning and transmission electron micros-
copies (SEM and TEM) can provide some direct information about the degree of orientation and dis-
tribution of the nanollers, or the matrix-ller interactions. Wide angle X-ray scattering (WAXS)
and differential scanning calorimetry (DSC) experiments are frequently performed to assess the crys-
talline structure, degree of crystallinity and crystal orientation of the matrix in the composite.
The following sections collect and summarize published data on nanoller-reinforced polymer
nanocomposites obtained via nanoindentation, in order to extract some general conclusions about
the inuence of the abovementioned parameters on the mechanical performance of the different
materials. Tables 16 present a comprehensive collection of room temperature (RT) DSI results
reported to date in this eld. 
The absolute
 values of EI and H and the percentage of change compared
to those of the matrix, DEI EI  E0I =E0I  100 and DH = (H  H0)/H0  100 being E0I and H0 the val-
ues of the neat material, have been included in the tables. Indentation modulus values are specied
whether obtained using a quasi-static (E) or a dynamic method (E0 ). Berkovich indenter tips are always
used unless otherwise stated. The nanoller type, characteristics, concentration and state of dispersion
as well as the composite processing method and the degree of crystallinity of the matrix are included
in the cases that this information is available. Furthermore, modulus data obtained by conventional
bulk techniques such as tensile testing, Y, or DMA are also listed in order to establish a correlation
between nanoindentation results and macroscopic properties. Because the selection of the method
of analysis is relevant to achieve meaningful indentation data as discussed in Section 2, this informa-
tion is offered to the reader in the different Tables. Finally, other aspects such as the variation of
mechanical properties with the indentation depth (ISE) are also reported.
Those indentation modulus data possibly affected by a large inaccuracy, due to the experimental
procedure followed, are commented in the footnotes of the corresponding Tables. In addition, red
symbols are used for such data when appearing in the Figures. For example, it is well known that
at the beginning of an indentation test, the detection of the point of initial contact encounters impor-
tant difculties if the material exhibits low modulus [78,82]. This is especially critical under two cir-
cumstances: (i) when sharp indenters such as Berkovich are used because the contact area between
the sample and the indenter is small; (ii) when the dynamic option enjoying improved resolution
to detect changes in stiffness is not used. Hence, from the above considerations, it is questionable
the signicance of modulus values and increments on very low-modulus materials (E < 0.15 GPa)
derived from quasi-static measurements using a sharp indenter. Other sources of error are the exis-
tence of a clear nose effect at the onset of the unloading curves and the inadequate correction of
the load-depth data by thermal drift (Section 2.1.1).

4. Polymer nanocomposites incorporating carbon nanollers

4.1. Carbon nanotubes

4.1.1. Synthesis methods and types of CNTs


One of the most efcient nanollers to reinforce polymer matrices are carbon nanotubes (CNTs),
one dimensional carbon-based nanomaterials discovered by Iijima in 1991 [83] that possess extraor-
dinary high Youngs modulus (up to 1.2 TPa) and tensile strength (ca. 50200 GPa), combined with
very large aspect ratio (>1000), high exibility and low density (1.8 g/cm3) [84,85]. There are two
main types of CNTs: those consisting of a single graphite sheet wrapped into a cylindrical tube with
a diameter in the range of 0.73 nm, the single-walled carbon nanotubes (SWCNTs), and those
Table 1
DSI modulus (either quasi-static E or dynamic E0 ) and hardness data for CNT-reinforced polymer nanocomposites. Type of CNT, composite preparation method, morphological information on
the state of aggregation of the ller and/or extent of the ller-matrix interaction and crystallinity values are also included. Values of storage modulus from DMA, E0DMA , and Youngs modulus
data from tensile experiments, Y, are incorporated in the table when appropriate. The gures in parenthesis indicate the percentage of property increase in comparison to that of the neat
matrix. The method of analysis of the DSI data is included in the table with the following abbreviations: From unloading (OP) indicates that a single reading of E and H is determined from the
onset of unloading using the OP method described in Eqs. (1)(4); Dynamic is based on the derivation of the harmonic contact stiffness where E0 , E00 and H are calculated following Eqs. (10),
(11) and (4) respectively. In the dynamic method, the contact depth values are determined using the OP procedure by substituting the quasi-static contact stiffness for the harmonic one (Eq.
(3)). Usually, this option is accompanied by a continuous measurement of the material properties. The variation of mechanical properties with indentation depth is also included in the table
when this information available. Berkovich indenter tips are used unless otherwise stated. NA indicates that the information is non-available.

Matrix Nanoller type Processing Nanoller Quasi- Dynamic Hardness Morphology Crystallinity E0 DMA Y (GPa) Method Depth Ref.

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


method content static DSI H (GPa) (Dispersion, (%) (GPa) (DY, %) of analysis variation
(wt%) DSI E E0 (DH, %) interaction) (% variation) (DE0 DMA, %) (range)
(GPa) (GPa)
(DE, %) (DE0 , %)

Epoxy- Raw MWCNTs Sonication/ 0a 1.81 0.075 SEM (aggregates NA ISE (0.155 lm) [108]
acrylate mixing at 0.75 wt%)
photocuring
0.25a 2.20(22) 0.098(30)
0.75a 2.60(42) 0.133(77)
0b 2.20 0.114
0.25b 2.50(14) 0.137(20)
0.75b 2.05(7) 0.094(9)
Epoxy CVD raw Sonication/ 0 4.22 0.325 SEM 2.00 From unloading (OP) [91,95]
MWCNTs mixing Curing in 0.1 4.70(11) 0.388(19) (agglomerates) 2.20(10)
oven
Epoxy Raw CVD Sonication/ 0 4.53 0.183 SEM From unloading (OP) [109]
MWCNTs mixing Curing in 3.0 4.75(5) 0.187(2) (agglomerates)
oven
Epoxy Raw MWCNTs Sonication/ 0 3.16 0.138 SEM (good 3.18 From unloading (OP)o [110]
mixing Curing in 0.1 3.19(1) 0.141(2) dispersion up to 3.24(2)
oven 0.5 3.35(6) 0.150(9) 0.5 wt%) 3.50(10)
1.0 3.44(9) 0.157(14) 3.68(16)
Epoxy Raw CVD Stirring/mixing 0 3.70 SEM, TEM (good Dynamic ISE (0.21 lm) [111]
MWCNTs Curing in oven 5.0c 7.40(100) dispersion)
5.0d 4.62(25)
Epoxy Raw CVD Stirring/mixing 0 2.26 0.125 2.14 From unloading (OP) ISE (0.10.4 lm) [112]
MWCNTs Curing in hot- 1.0 3.10(37) 0.187(50) 2.50(17)
press 2.0 3.64(61) 0.293(85) 3.09(44)
Epoxy Raw SWCNTs Stirring/casting 0 3.0 SEM, AFM (a lot of From unloading (OP) ISE (0.150.3 lm) [113]
Curing in oven 1.0 3.0(0) aggregates)
Epoxy Acid- Stirring/mixing 0 3.70 SEM, TEM (good Dynamic ISE (0.21 lm) [111]
functionalized Curing in oven 5.0c 7.55(104) dispersion)
CVD MWCNTs 5.0d 4.25(15)
Epoxy Fluorinated- Stirring/casting 0 3.0 SEM, AFM (many From unloading (OP) ISE (0.150.3 lm) [113]
SWCNTs Curing in oven 1.0 3.0(0) aggregates)
Epoxy Silane Stirring/casting 0 3.0 SEM, AFM (few From unloading (OP) ISE (0.150.3 lm) [113]
functionalized- Curing in oven 1.0 3.0(0) aggregates)
SWCNTs

19
(continued on next page)
Table 1 (continued)
20
Matrix Nanoller type Processing Nanoller Quasi- Dynamic Hardness Morphology Crystallinity E0 DMA Y (GPa) Method Depth Ref.
method content static DSI H (GPa) (Dispersion, (%) (GPa) (DY, %) of analysis variation
0
(wt%) DSI E E0 (DH, %) interaction) (% variation) (DE DMA, %) (range)
(GPa) (GPa)
(DE, %) (DE0 , %)

Epoxy Plasma Sonication/ 0 4.22 0.235 SEM (good 2.00 From unloading (OP) [91,95]
functionalized mixing Curing in 0.1 5.05(19) 0.456(40) dispersion) 2.40(20)
CVD MWCNTse oven
Epoxy MBZ Solution mixing 0 4.0 4.0 0.34 SEM (uniform From unloading (OP) H and E constant [57]
functionalized Curing in oven 1.0 4.4(10) 5.0(24) 0.37(9) distribution up to and Dynamic (0.21 lm)
Arc-discharge 3.0 5.6(40) 5.9(48) 0.41(20) 1 wt%)
SWCNTs 5.0 7.0(75) 6.7(67) 0.44(30)
Epoxy Puried laser- Sonication/ 0 3.90 0.245 SEM (poor 5.46 Dynamic H and E0 constant [24,114]
grown mixing Curing in 0.1 3.78(3) 0.254(4) dispersion) 4.78(12) (24.5 lm)
SWCNTs oven 0.5 3.98(2) 0.262(7) 4.96(9)
1.0 4.06(4) 0.265(8) 6.01(10)
3.0 4.25(9)
5.0 4.56(17)
Epoxy Aligned CVD Mixing/curing 0 3.6 0.157 SEM,TEM (quite NA [115]
MWCNT 0.5 4.0(11) 0.177(13) agglomerated)
forests
Epoxy Aligned CVD In situ wetting/ 0 3.6 SEM (dense CNT From unloading (OP)o [116]
MWCNT Realign/curing NA 6.4(78) array)
forests NA 9.6(166)
Epoxy Aligned CVD Mechanical 0 4.8 OM, SEM From unloading (OP) [117]
MWCNT densication/ 0.75 4.9(2) (densication)
forests Capillary- 5.5 6.0(25)
induced wetting/ 14.2 8.8(85)
curing
Epoxy CVD MWCNT Shear mixing/ 0 4.8 OM, SEM (good From unloading (OP) [117]
forests curing 0.75 4.6(4) dispersion)
3.0 5.1(6)
Epoxy Coiled CVD Sonication/ 0 4.90 0.240 SEM (disentangled From unloading (OP) H and E constant [118]
CNTs mixing Curing in 1.0 5.60(14)f 0.271(13)f and evenly (1.011.03 lm)
oven 3.0 6.00(23)f 0.315(31)f spaced)
5.0 7.00(42)f 0.350(46)f
Epoxy CVD CNTs Curing 0 2.23 0.13 Raman, AFM From unloading (OP) NA [119]
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

grafted silica NA 2.77(24) 0.17(28)


bre
Chitosan Acid- Solution-casting 0 0.119 SEM, TEM (good 1.08 Dynamic ISE (<0.3 lm) [99]
functionalized 0.2 0.128(8) dispersion up to 1.33(23) Reverse ISE
CVD MWCNTs 0.4 0.163(37) 2.0 wt%) 1.92(77) (15 lm)
0.8 0.150(26) 2.08(93)
2.0 2.15(99)
Chitosan PEDOT-PSS Solution mixing/ 0 2.30 0.130 SEM (good 3.05 Dynamic ISE (<0.2 lm) [98]
functionalized casting 0.1 2.60(13) 0.146(12) dispersion) 3.54(16)
MWCNTs 0.2 3.15(37) 0.155(19) 3.75(23)
0.5 3.50(52) 0.223(31) 4.09(34)
Table 1 (continued)

Matrix Nanoller type Processing Nanoller Quasi- Dynamic Hardness Morphology Crystallinity E0 DMA Y (GPa) Method Depth Ref.
method content static DSI H (GPa) (Dispersion, (%) (GPa) (DY, %) of analysis variation
0
(wt%) DSI E E0 (DH, %) interaction) (% variation) (DE DMA, %) (range)
(GPa) (GPa)
(DE, %) (DE0 , %)

PHB Raw SWCNTs Solution-casting 0 5.66 0.31 OM (poor Increase in From unloading (OP) [97]
1.0 7.62(35) 0.33(6) dispersion; large crystallinity,
10.0 11.74(107) 0.35(12) aggregates at decrease in
10 wt%) crystallite size
PHO Raw SWCNTs Solution-casting 0 0.12g 0.0056 From unloading (OP) [97]
1.0 0.15(25) 0.0077(37)
10.0 0.53(341) 0.0137(145)
PP Raw MWCNTs In situ 0 1.45 0.065h From unloading (OP)o [106]
polymerization 0.4 2.22(53) 0.092(41)h
Compression 1.0 2.23(54) 0.100(54)h
moulding 1.3 1.58(9) 0.081(24)h
1.6 2.64(82) 0.120(85)h
2.2 2.02(39) 0.093(42)h
2.4 2.00(38) 0.091(40)h
PA-6 Acid-puried Melt-blending/ 0 1.18 0.06 SEM (uniform CNTs act as a- 2.08 0.40 Dynamic ISE (<0.2 lm) [20]
CVD MWCNTs Compression - 0.2 1.33(13) 0.08(33) dispersion) nucleating 2.55(23) 0.68(72)
moulding 0.5 1.60(36) 0.10(66) agents 2.57(24) 0.77(93)
1.0 2.02(71) 0.11(83) 3.14(51) 0.85(113)
PA-6i Acid- Solution mixing/ 0 0.95 0.240 SEM, TEM (good 33 0.90 1.20 From unloading (OP)p [100]
functionalized electrospinning 1.0 1.56(64) 0.252(5) dispersion) 36(9) 2.00(70)
MWCNTsj 2.5 2.14(125) 0.270(13) NA 1.50(67) 2.40(100)
5.0 2.38(150) 0.274(14) NA 2.60(120)
7.5 2.66(180) 0.280(17) NA 3.30(267) 3.00(150)
PLLA Acid-puried Sonication/ 0 3.6 SEM (good From unloading (OP) [101]
Arc-discharge Solution-casting 1.25 4.4(22) dispersion)
MWCNTs 2.5 5.8(61)
3.75 6.6(83)
5.0 7.5(108)
6.25 7.8(117)
UHMWPE Raw MWCNTs Milling/ 0 2.02 0.104 SEM (quite 55 0.60 From unloading (OP) [105]
Electrostatic 5.0 2.23(10) 0.116(12) agglomerated) 43(22) 1.28(115)
spraying
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

UHMWPE Plasma Solution mixing/ 0 2.75 0.062 SEM (good Dynamic [96]
functionalized casting 0.05 3.1(12) 0.820(32) dispersion) XPS
Arc-discharge 0.1 4.26(55) 0.100(62) (strong interfacial
SWCNTs 0.2 4.35(58) 0.103(66) adhesion)
UHMWPE Acid- Ball milling/ 0 0.617 0.261 SEM (good 52 0.027l From unloading (OP)k [104]
functionalized Compression- 0.5 0.728(18) 0.282(8) dispersion) 55(5) 0.034(26)l (13)
MWCNTs moulding 1.0 0.808(31) 0.381(46) Raman (strong 58(10) 0.23(763)l (26)
1.5 0.925(50) 0.412(58) interfacial 61(15) 0.3(1000)l (38)
2.0 1.67(170) 0.457 (75) adhesion) 64(23) 0.4(1380)l (44)
2.5 1.75(183) 0.483 (85) 66(27) 0.5(1750)l (40)
5.0 1.89(206) 0.564(116) 71(37) 0.6(2120)l (24)

(continued on next page)


21
22
Table 1 (continued)

Matrix Nanoller type Processing Nanoller Quasi- Dynamic Hardness Morphology Crystallinity E0 DMA Y (GPa) Method Depth Ref.
method content static DSI H (GPa) (Dispersion, (%) (GPa) (DY, %) of analysis variation
(wt%) DSI E E0 (DH, %) interaction) (% variation) (DE0 DMA, %) (range)
(GPa) (GPa)
(DE, %) (DE0 , %)

PVDFi Raw MWCNTsj Near-eld 0 0.26 SEM (uniform NA [120]


electrospinning 0.03 0.31(19) dispersion) (25)m
PVA Raw arc- Solution mixing 0 7.0 0.30 TEM (strong 14 NA [103]
discharge 0.1 8.4(20) 0.30(0) interfacial 24(71)
MWCNTs 0.25 9.4(35) 0.40(30) bonding) 25(75)
0.5 10.4(48) 0.42(38) 26(80)
1.0 12.5(79) 0.47(57) 27(93)
PVA Acid- Solution casting 0 0.66 0.038 42 From unloading (OP) [121]
functionalize 0.2 6.93(950) 0.278(630) 52(24)
Arc-discharge 0.4 7.30(1000) 0.280(637) 55(30)
SWCNTs 0.6 7.80(1180) 0.290(663) 64(51)
PEEK/GF Arc-puried Melt-blending/ 0 3.9 0.198 SEM (few 39 16.7 15.6 Dynamic H and E0 constant [122]
SWCNTs Hot-compression 1.0 4.25(9) 0.218(10) agglomerates) 38(2) 16(3) (0.10.4
lm)64(51)
PEEK/GF PEES-wrapped Melt-blending/ 0 3.9 0.198 SEM (good 39 16.7 15.6 Dynamic H and E0 constant [122]
Arc-puried Hot-compression 1.0 4.60(18) 0.238(20) dispersion) 37(5) 17.8(14) (0.10.4 lm)
SWCNTs
PDMS Raw MWCNTs Ultrasonication/ 0 0.0067g 0.0022 SEM 0.0015 0.165 From unloading (OP) [124]
curing 4.0 0.0089(32) 0.0027(23) (agglomerates) 0.0028(88) 0.234(42)
n
PC CVD raw Melt-extrusion 0 4.0 0.38 SEM From unloading (NA) Reverse ISE [125]
MWCNTs 2.0 4.2(5) 0.40(4) (homogeneous (<1 lm)
4.0 4.4 (10) 0.41(8) dispersion)
6.0 4.5(12) 0.43(13)
15.0 5.2(30) 0.50(32)
PI Raw Arc- Sonication/spin- 0 5.23 0.43 AFM, SEM, OM Dynamic ISE (<1 lm) [126]
discharge coating 0.05 8.41(59) 0.72(67) (some
SWCNTs agglomerates,
good interfacial
bonding)
PVK Raw arc- Solution mixing 0 2.0 0.30 NA [103]
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

discharge 0.5 2.24(12) 0.33(10)


MWCNTs 1.0 3.10(55) 0.44(48)
2.0 2.56(28) 0.50(68)
4.0 2.96(48) 0.63(110)
8.0 6.0(200) 0.60(100)
PMMA Amide- Solution mixing/ 0 3.6 0.05 SEM (non uniform From unloading (OP) E constant ISE in [28,102]
functionalized spin-coating 1.0 3.38(6) 0.048(4) dispersion) H (0.10.5 lm)
CVD MWCNTs 3.0 3.67(2) 0.051(2)
5.0 3.53(2) 0.05(0)
Table 1 (continued)

Matrix Nanoller type Processing Nanoller Quasi- Dynamic Hardness Morphology Crystallinity E0 DMA Y Method Depth Ref.
method content static DSI H (GPa) (Dispersion, (%) (GPa) (GPa) of analysis variation
(wt%) DSI E E0 (DH, %) interaction) (% (DE0 DMA, (DY, (range)
(GPa) (GPa) variation) %) %)
0
(DE, %) (DE , %)

PMMA Silica-coated CVD Solution 0 3.6 0.05 SEM (aggregates From unloading (OP) E constant ISE in [28,102]
MWCNTs mixing/spin- 1.0 3.24(10) 0.05(0) increasing with CNT H (0.10.5 lm)
coating 2.0 4.93(37) 0.085(70) content)
3.0 6.66(85) 0.10(100)

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


4.0 8.28(130) 0.11(120)
P(MMA-co- APTES functionalized Solgel/spin- 0 5.0 0.107 OM (homogeneous From unloading (OP) Reverse ISE [127]
MPTMS) acid-treated MWCNTs coating/curing 2.0 6.3(26) 0.209(95) dispersion) (0.170.95 lm)
P(MMA-co- TEOS functionalized Solgel/spin- 0 5.0 0.107 OM (aggregates) Reverse ISE [127]
MPTMS) PAH-dispersed coating/curing 2.0 4.9(2) 0.189(76) (0.140.8 lm)
MWCNTs

(See abbreviations in the appendix).


a
Cured for 12 h.
b
Cured for 24 h.
c
MWCNTs predispersed in acetone.
d
MWCNTs predispersed in gamma-butyrolactone (GBL).
e
Amine functionalized CNTs grafted to the matrix.
f
Results showing nose effect at the beginning of unloading.
g
Materials with matrix modulus <0.15 GPa, possibly affected by a large uncertainty in the detection of the point of initial contact.
h
Martens hardness [2].
i
Composite bre.
j
Aligned along the bre axis.
k
Inadequate thermal drift correction.
l
Data from rheological measurements.
m
Crystallinity of b phase.
n
Experiments carried out using cycling loading.
o
ISO 14577 test.
p
Spherical indenter.

23
Table 2

24
Nanomechanical and macroscopic properties for graphene-reinforced polymer nanocomposites. Columns as indicated in Table 1.

Matrix Nanoller type (thickness) Processing Nanoller Quasi- Dynamic Hardness Morphology Crystallinity E0DMA Y (GPa) Method of analysis Depth Ref.
method content static DSI E0 H (GPa) (Dispersion, (%) (GPa) (D, Y %) variation
(wt%) DSI E (GPa) (DH, %) interaction) (% (DE0DMA , (range)
0
(GPa) (DE , %) variation) %)
(DE, %)

Epoxy GP (350550 nm) Sonication/ 0 2.5 2.8 AFM (random and 2.8 2.70 From unloading ISE (0.5 [166,167]
mixing/ 1.0 3.2(28)a 3.5(25) homogeneous 3.5(25) 3.41(26) (OP) and Dynamic 1.5 lm)
curing 5.8 3.3(32)a 3.8(35) dispersion) 3.8(36)
Epoxy GNP (68 nm) Mixing/ 0 3.7 0.130 From unloading [168]

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


curing 0.05 4.0(8) 0.165(26) (OP)i
0.1 4.1(11) 0.175(35)
0.25 4.3(17) 0.175(35)
0.5 4.6(25) 0.175(35)
Epoxy GNP (7 nm) Mixing/ 0 3.61 0.26 SEM (homogeneous 2.72 Dynamic ISE (<0.2 lm) [169]
curing in 1.0 3.68(2) 0.26(0) and random 2.82(4) H and E
oven 2.0 3.90(8) 0.25(2) dispersion of the 2.92(8) constant
3.0 3.93(9) 0.25(2) blocks) 3.04(12) (>0.2 lm)
4.0 3.97(10) 0.24(4) 3.15(16)
5.0 4.30(19) 0.26(0) 3.26(20)
6.0 4.70(30) 0.26(0) 3.36(30)
Epoxy GNP Mixing/ 0.1 (0)b NA AFM (graphene From unloading [170]
stirring clusters) (NA)
curing
Epoxy EG Sonication/ 0 0.32 SEM (micrometer 3.0 From unloading [171]
photocuring 1.0 0.34(6) agglomerates) 1.8 (NA)
(40)
Epoxy FGS (stacks up to 7 sheets) Sonication/ 0 0.32 SEM (stacked groups 3.0 From unloading [171]
photocuring 1.0 0.34(6) of sheets) 3.0 (0) (NA)
Epoxy TEG Functionalized nitric Mixing/ 0 3.27 0.396 SEM (homogeneous From unloading [172]
and sulphuric acid (Blocks curing 1.0 3.82(17)b 0.386(2) dispersion of the (OP)d
of graphene nanosheets) 2.0 4.00(22)b 0.400(1) blocks)
Chitosan Arc-discharge N-doped GS Solution 0 3.2 0.30 Raman/TEM From unloading [156]
(Few layered graphene casting 0.1 6.7(110)b 0.42(40) (homogeneous (NA)
sheets) 0.2 7.3(128)b 0.42(40) dispersion of the
0.3 7.8(144)b 0.39(30) sheets)
0.6 6.6(106)b 0.40(35)
2.3 6.7(110)b 0.47(57)
PET CVD GSc (12 layers) Solution 0 3.7 0.15 From unloading [157]
mixing/ NA 4.3(16) 0.30(100) (OP)
coating
PP GO (monolayer, bilayer (2 Melt- 0 0.459 0.00661 SEM (agglomerates) From unloading [154]
3 nm), multilayer) blending/ 0.1 0.539(15) 0.00715(8) (OP)d
hot- 0.5 0.592(28) 0.0080(21)
pressing 1.0 0.666(45) 0.0087(30)
PP OTES-modied GO Melt- 0 0.459 0.00661 SEM (homogeneous From unloading [154]
(monolayer, bilayer (2 blending/ 0.1 0.613(33) 0.0218(228) dispersion) (OP)d
3 nm), multilayer) hot- 0.5 0.651(42) 0.0254(283)
pressing 1.0 0.743(62) 0.0256(286)
Table 2 (continued)

Matrix Nanoller type (thickness) Processing Nanoller Quasi- Dynamic Hardness Morphology Crystallinity E0DMA Y (GPa) Method of analysis Depth Ref.
method content static DSI E0 H (GPa) (Dispersion, (%) (GPa) (D, Y %) variation
(wt%) DSI E (GPa) (DH, %) interaction) (% (DE0DMA , (range)
0
(GPa) (DE , %) variation) %)
(DE, %)

PVA Acid-functionalized FG Solution 0 0.657 0.038 Raman (good 42 From unloading [21]
(few layer) casting 0.2 0.688(5) 0.039(3) interaction GS- 42.8(2) (OP)
0.4 0.734(12) 0.0441(15) matrix) 46.2(10)
0.6 0.885(35) 0.0557(47) 47.5(13)
PVA GNP (EOGCNF) Partially Solution 0 0.230 0.0129 From unloading [160]

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


oxidized (monolayer) casting 0.1 0.257(4) 0.0135(5) (10) (OP)d
0.5 0.388(20) 0.0153(19) (18)
1.0 0.349(52) 0.0144(12) (2)
2.0 0.651(80) 0.0183(42) (4)
4.0 0.854(160) 0.0211(62) (5)
f,g
PVA GO Layer-by- 0 8.88 0.34 AFM (homogeneous From unloading [173]
layer 17.64(99) 1.15(240) dispersion) (OP)
PMMA Amide-functionalized FG Solution 0 2.12 0.140 From unloading [21]
(few layer) casting 0.2 3.36(58) 0.1449(4) (OP)
0.4 3.61(70) 0.1456(4)
0.6 3.65(72) 0.1531(9)
PMMA EG Melt- 0 4.3 0.42 SEM (regular From unloading [155]
blending 0.25 4.5(5)b 0.45(8) dispersion, (OP)
0.5 4.7(10)b 0.51(23) agglomerates at
1.0 4.7(10)b 0.49(16) 1.0 wt%)
PAA PEI-modied GO Layer-by- 0 1.3 0.160 SEM, AFM, IR Dynamic (NA)j ISE (<0.2 lm) [174]
layer NA 1.9(46) 0.240(50) (increase in
roughness)
PU Highly aligned rGOg,h Solution 0 0.4 0.025 SEM (homogeneous 0.3 From unloading [158]
(1 nm) casting 1.0 1.0(150) 0.035(40) dispersion) 1.25(320) (NA)
2.0 3.1(675) 0.050(100) 3.00(920)
5.0 7.4(1750) 0.820(228) 6.30(2000)
PU GOh (<50 nm) Solution 0 0.022e 0.061 SEM (homogeneous 0.011 NA [159]
casting 1.0 0.027(23) 0.078(27) dispersion) 0.031(180)
4.0 0.061(175) 0.115(88) 0.092(735)

a
Data for an average platelet length of 1 lm.
b
Results showing nose effect (too short holding time [155,170]; no holding time [156,172]).
c
Graphene overlayer onto PET substrate.
d
AFM nanoindentation with a silicon tip.
e
Materials with matrix modulus <0.15 GPa, possibly affected by a large uncertainty in the detection of the point of initial contact.
f
Planar orientation.
g
Partially reduced graphene oxide.
h
Nanoller covalently bonded to the polymer matrix.
i
ISO 14577 test.
j
Lack of experimental details, indenter geometry not reported.

25
26
Table 3
Nanomechanical and macroscopic properties for polymer nanocomposites reinforced with different types of organic nanomaterials: carbon nanobres (CNFs), nanodiamond (ND), carbon
black (CB), fullerenes, cellulose nanocrystals (CNC), and hybrids with two carbon-based nanollers. Columns as indicated in Table 1.

Matrix Nanoller Processing Nanoller Quasi- Dynamic Hardness Morphology (Dispersion, Crystallinity E0DMA Y (GPa) Method of Depth Ref.
type method content static DSI E0 H (GPa) interaction) [%] (GPa) (DY, %) analysis variation
(wt%) DSI E (GPa) (DH, %) (% DE0DMA , (range)
(GPa) (DE0 , %) variation) %)
(DE, %)

Epoxy VGCNF Calendaring/ 0 4.50 0.375 TEM (Agglomerates 2.18 From unloading [197]
curing 1.5 4.05(10) 0.310(17) increasing with nanobre 2.23(2) (OP)

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


3.0 4.50(0) 0.370(2) content. Good bre 2.36(8)
5.0 4.70(27) 0.432(15) wetting) 2.82(30)
Epoxy AA-coated Mixing/curing 0 3.60 0.21 SEM (agglomerates, bre 3.57 Dynamic ISE [198]
CNF 0.5 3.59(0) 0.21(0) pull-out) 3.58(0) (<0.3 lm)
Epoxy ND  NH2 a Stirring/curing 0b 1.90 TEM (some agglomerates, Dynamic j
[199]
0c 2.50 particularly at high
0d 2.30 loadings)
0e 1.91
0f 2.22
9.1g 1.84(0)
10.0d 2.41(5)
14.5e 2.43(30)
16.7c 3.08(20)
40.0f,i 3.94(78)
53.3h,i 12.1(426)
Epoxy ox-ND Stirring/curing 0b 1.90 TEM (agglomerates) Dynamicj [199]
0d 2.30
8.8b 2.22(15)
8.8d 2.50(9)
Epoxy CB Ultrasonic 0 6.0 (<10) SEM (uniform dispersion) From unloading Reverse ISE [193]
mixing/curing 1.4 6.24(4) (OP) (0.2
2.8 6.72(12) 0.25 lm)
3.4 7.38(23)
4.6 8.4(40)
5.8 6.48(8)
Epoxy Mixed Mixing/curing 0 3.6 0.157 NA [115]
fullerenes 10.0 1.8(50) 0.058(63)
PVA (90)/PAA CNC Sonication/ 0 2.20 AFM (some aggregates) 1.50 Single readingk [200]
(10) Solution casting 10 3.10(40) 2.10(40) (Herzt)l
20 2.10(5) 2.05(37)
PVA CNC Sonication/ 0 2.0 AFM (heterogeneous 1.20 Single readingk [200]
Solution casting 15 3.0(50) dispersion) 1.40(17) (Herzt)l
m n
PVA CNC Mixing/ 0 2.10n Single reading [201]
electrospinning 1.0 2.31(10)n (OP)j
3.0 2.72(30)n
5.0 3.16(50)n
10.0 4.26(105)n
15.0 5.88(180)n
20.0 7.58(262)n
Table 3 (continued)

Matrix Nanoller Processing Nanoller Quasi- Dynamic Hardness Morphology (Dispersion, Crystallinity E0DMA Y (GPa) Method of Depth Ref.
type method content static DSI E0 H (GPa) interaction) [%] (GPa) (DY, %) analysis variation
(wt%) DSI E (GPa) (DH, %) (% DE0DMA , (range)
(GPa) (DE0 , %) variation) %)
(DE, %)

PVA ox-ND Sonication/ 0 0.67 0.0383 TEM (good dispersion) 42 From unloading [121,202]
solution 0.2 0.87(30) 0.0437(14) 52.6(25) (OP)
casting/ 0.4 0.96(43) 0.0528(38) 55.0(31)
0.6 1.33(99) 0.0684(80) 56.6(35)
PVAc CNC Solution 0 3.00 0.060 AFM (aggregates at From unloading [203]
casting/ 1.0 4.20(40) 0.102(70) contents >1 wt%) (OP)
compression 2.0 4.30(43) 0.093(55)
3.0 5.40(80) 0.156(160)
PLLA as Solution casting 0 2.6 0.05 TEM (poor dispersion) Dynamic (Herzt)j,l [187]
received- 1.0 2.8(9) 0.12(140)
ND 3.0 2.7(5) 0.11(120)
PLLA ODA- Solution casting 0 2.6 0.05 TEM (good dispersion) Increase in Dynamic (Herzt)j,l [187]
modied 1.0 5.3(107) 0.21(320) NMR: good interaction crystallinity
ND 3.0 5.5(113) 0.25(400) with matrix
5.0 5.9(129) 0.26(420)
7.0 6.8(165) 0.31(520)
10.0 7.9(206) 046(820)
PA-11m ox-ND Sonication/ 0 2.0 0.075 SEM (increase in surface Dynamicj [204]
electrospinning 2.5 3.5(75) 0.082(9) roughness with increasing
10.0 3.5(75) 0.094(25) loading)
20.0 8.0(300) 0.14(86)
PA-11 ox-ND Milling/thermal 0 1.80n 0.150 SEM (poor distribution) Dynamic [205]
spray 7.0 1.62(10)n 0.146(2)
deposition 13.0 1.70(5)n 0.143(4)
25.0 1.70(5)n 0.140(6)
PS-basedo CB Ultrasonication/ 0 1.2 0.045 SEM (uniform dispersion) 0.240 From unloading [206]
curing 1.0 1.50 (25) 0.054(20) 0.265(10) (OP)
2.0 2.16 (80) 0.061(38) 0.300(25)
3.5 3.2 (170) 0.101(125) 0.336(40)
5.5 5.4(350) 0.205(350) 0.430(82)
PS PS-coated Mixing/curing 0 3.60 SEM (uniform dispersion) 2.10 Dynamic ISE [198]
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

CNF 1.0 3.70(3) 2.15(2) (<0.3 lm)


2.0 3.73(4) 2.10(0) Reverse ISE
3.0 3.90(9) 2.30(10) (0.34 lm)
5.0 3.80(6) 2.27(8)
PMMAm CNCn Solution 0 5.30 SEM (good dispersion at Dynamicj E0 constant [207]
mixing/ 5 5.94(12) low loadings) (0.1
electrospining 9 6.15(16) 0.25 lm)
17 6.25(18)
23 6.15(16)
33 6.90(30)
41 6.25(18)

(continued on next page)


27
28
Table 3 (continued)

Matrix Nanoller Processing Nanoller Quasi- Dynamic Hardness Morphology (Dispersion, Crystallinity E0DMA Y (GPa) Method of Depth Ref.
type method content static DSI E0 H (GPa) interaction) [%] (GPa) (DY, %) analysis variation
(wt%) DSI E (GPa) (DH, %) (% DE0DMA , (range)
(GPa) (DE0 , %) variation) %)
(DE, %)

PAN ox-ND Solution casting 0 5.8 TEM (poor dispersion) 2.0 NA [208]
5.0 6.32(9) 2.32(16)
10.0 7.08(22) 2.50(25)
20.0 9.0(55) 2.50(25)

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


30.0 13.0(125) 2.45(22)
PAN ox-ND (de- Bead milling/ 0 5.8 TEM (good dispersion) 2.0 NA [208]
aggregated) Solution casting 5.0 6.13(10) 2.30(15)
10.0 6.96(20) 2.58(28)
20.0 7.37(27) 3.00(50)
30.0 11.14(92) 3.60(80)
PAN deox-ND Solution casting 0 5.8 TEM (poor dispersion) NA [208]
5.0 5.51(5)
10.0 4.52(22)
20.0 6.96(20)
30.0 7.83(35)
PVA ox- Solution casting 0 0.66 0.038 42.0 From unloading [121]
SWCNTs/ 0.4/0.2 9.3(1300) 0.366(865) 56.5(35) (OP)
ox-FG 0.2/0.4 8.6(1200) 0.337(787) 57.5(37)
PVA ox- Solution casting 0 0.66 0.038 42.0 From unloading [121]
SWCNTs/ 0.4/0.2 7.5(1036) 0.314(726) 55.1(31) (OP)
ox-ND 0.2/0.4 9.3(1300) 0.353(829) 57.2(37)
PVA ox-FG/ox- Solution casting 0 0.66 0.038 42.0 From unloading [121]
ND 0.4/0.2 1.6(150) 0.066(75) 55.1(31) (OP)
0.2/0.4 1.3(97) 0.061(61) 54.8(30)

a
Nanoller grafted to the polymer matrix.
b
r (overall amine groups/epoxide groups including curing agent and ND-NH2 when appropriate) = 0.39.
c
r (overall amine groups/epoxide groups including curing agent and ND-NH2 when appropriate) = 0.57.
d
r (overall amine groups/epoxide groups including curing agent and ND-NH2 when appropriate) = 1.
e
r (overall amine groups/epoxide groups including curing agent and ND-NH2 when appropriate) = 1.54.
f
r (overall amine groups/epoxide groups including curing agent and ND-NH2 when appropriate) = 0.64.
g
r (overall amine groups/epoxide groups including curing agent and ND-NH2 when appropriate) = 0.34.
h
r (overall amine groups/epoxide groups including curing agent and ND-NH2 when appropriate) = 1.18.
i
Without curing agent.
j
spherical indenter [199,187], spheroconical [204,205], conical [207] and cube-corner [201].
k
AFM nanoindentation.
l
Hertz model used ([200] applies Hertz to loading and unloading curves, values in the table corresponding to the load tting).
m
Composite bre; naligned along the bre axis.
n
Results showing nose effect at the beginning of unloading.
o
Shape memory polymer (PS: 30 wt%, styrene: 55 wt%, vinyl neodecanoate: 10 wt%).
Table 4
Nanomechanical and macroscopic properties for polymer layered silicates nanocomposites. Columns as indicated in Table 1.

Matrix Nanoller type Processing method Nanoller Quasi- Dynamic Hardness Morphology (Dispersion, Crystallinity E0 DMA (GPa) Y (GPa) Method of Depth Ref.
content static DSI E0 H(GPa) interaction) (%) (% (DE0 DMA, %) (DY, %) analysis variation
(wt%) DSIE(GPa) (GPa) (DH, %) variation) (range)
(DE, %) (DE0 ,%)

PA-6 Organically Melt blending/ 0 1.06 0.054 TEM (agglomeration at 10 wt%) 15 1.50 From [223]
modied clay injection moulding 2.5 1.84(74) 0.100(85) SAXS (Exfol 65 wt% Inter/ 13.75(8) 2.25(50) unloading
(I30TC) 5 1.97(86) 0.118(119) exf > 5 wt%) 12(20) (OP)
7.5 2.3(117) 0.138(156) 10.5(30) 2.66(77)
10 2.42(128) 0.141(161) 10(33)

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


PA-6 CoAl-LHD In situ 0 1.13 0.063 TEM, X-ray (homogeneous 1.0 Dynamic ISE [224]
(organo-modied polymerization/ 0.5 1.46(29) 0.077(22) dispersion. Exfoliation) 1.6(60) (<1 lm)
layered double compression 1 1.49(32) 0.079(25) 2.0(100)
hydroxides) moulding 2 1.50(35) 0.085(35) 2.1(110)
PA-6 Organically Solution blending/ 0 3.352 3.482 0.121 X-ray (intercalation) 27.6 Dynamic [225]
modied clay compression L-3 4.431(32) 4.392(26) 0.194(60) 26.4(4) and from
(OMMT-L, D,DD) moulding L-6 4.748(42) 4.855(39) 0.200(65) 26.2(5) unloading
L-9 5.460(63) 5.445(56) 0.170(41) 23.1(16) (OP)
D-9 4.766(42) 4.392(26) 0.168(39)
DD-9 4.417(32) 0.172(42) 25.5(8)
PA-6 Surface modied Melt blending/ 0 1.621 0.179 X-ray (clay platelets uniformly 1.376 From [226]
MMT (1.34TCN) injection moulding 4 2.265(40) 0.198(11) dispersed. Exfoliation) 1.469(7) unloading
6 3.376(108) 2.136(55) (OP)
PA-6,6 Surface modied Melt blending/ 0 2.345 0.193 X-ray (clay platelets uniformly 1.667 From [226]
MMT (1.34TCN) injection moulding 4 2.917(24) 0.204(6) dispersed. Exfoliation) 1.800(8) unloading
6 2.909(24) 0.290(50) 1.819(9) (OP)
PA6/PA6,6 Surface modied Melt blending/ 0 2.370 0.197 X-ray (clay platelets uniformly 1.636 From [226]
(30/70) MMT (1.34TCN) injection moulding 4 2.875(21) 0.233(13) dispersed. Exfoliation) 1.648(0.7) unloading
6 2.763(17) 0.222(13) 1.705(4) (OP)
PA6/PA6,6 Surface modied Melt blending/ 0 1.938 0.172 X-ray (clay platelets uniformly 1.599 From [226]
(50/50) MMT (1.34TCN) injection moulding 2 1.746(10) 0.163(5) dispersed Exfoliation) 1.566(2) unloading
4 3.004(55) 0.365(112) 1.686(5) (OP)
6 3.149(63) 0.352(105) 1.704(7)
PA6/PA6,6 Surface modied Melt blending/ 0 2.190 0.201 X-ray (clay platelets uniformly 1.618 From [226]
(70/30) MMT (1.34TCN) injection moulding 4 2.881(32) 0.283(41) dispersed. Exfoliation) 1.713(6) unloading
6 3.274(50) 0.341(70) 1.742(8) (OP)
PA-6,6 Surface modied Melt blending/ 0 2.30a 0.098a TEM, X-ray (homogeneous 52c 1.89 3.04 Dynamic ISE [227
MMT (1.34TCN) injection moulding 0 2.27b 0.099b dispersion. Exfoliation) (<0.4 lm) 229]
1 2.39(4)a 0.105(7)a 42(19)c 2.22(18) 3.19(5)
1 2.34(3)b 0.098(1)b
2 2.57(12)a 0.116(18)a 41.5(20)c 2.24(19) 3.41(12)
2 2.42(7)b 0.106(7)b
5 2.72(18)a 0.125(28)a 38(27)c 2.36(25) 3.91(29)
5 2.44(8)b 0.106(7)b
PA-6,6 Organically Melt blending/ 0 1.6d,e 0.0875d Fractography 2.6d NA [230]
modied injection moulding 1 2(25)d,e 0.103(18)d 3.5(35)d
Hectorite 5 2.5(56)d,e 0.119(36)d 4(54)d

(continued on next page)

29
Table 4 (continued)
30
Matrix Nanoller type Processing method Nanoller Quasi- Dynamic Hardness Morphology (Dispersion, Crystallinity E0 DMA (GPa) Y (GPa) Method of Depth Ref.
content static DSI E0 H(GPa) interaction) (%) (% (DE0 DMA, %) (DY, %) analysis variation
(wt%) DSIE(GPa) (GPa) (DH, %) variation) (range)
(DE, %) (DE0 ,%)

PA-6,6/SEBS- Organically Melt blending/ 0 1.35 0.110 TEM, AFM (good dispersion. 2.95 From [231]
g-MA modied MMT injection moulding 5 1.82(35)f 0.140(27)f Exfoliation) 2.65(10)f unloading
(Cl 30B) (OP)
PA-11 Surface modied Melt blending/ 0 1.35 0.086 1.08 0.61 Dynamic ISE [232]
MMT (1.34TCN) compression 1 1.58(17) 0.100(16) 1.31(21) 0.62(2) (<1.0 lm)
moulding 2 1.65(22) 0.102(19) 1.41(31) 0.75(23)
5 1.78(32) 0.109(27) 1.84(70) 0.86(41)
PA-12 Surface modied Melt blending/ 0 1.03 0.06 TEM (homogeneous dispersion. 1.40 1.03 Dynamic ISE [233]
MMT (1.34TCN) injection moulding 1 1.39(35) 0.09(50) Exfoliation) 1.42(1) 1.10(7) (<0.3 lm)
2 1.44(40) 0.09(50) 1.45(4) 1.12(9)
5 1.56(52) 0.10(67) 1.49(6) 1.49(45)
PA-12 Layered silicate Melt blending/ 0 2.212g 0.156 SEM, TEM (Intercalation/ From [234]
nanoclay (LK-PA- injection moulding 1 2.340(6)g 0.161(3) Exfoliation) unloading
CR1) 3 2.900(31)g 0.202(30) (OP)
5 3.835(73)g 0.338(117)
PLA Organically Melt blending/ 0 4.360 0.219 TEM (clay nely distributed 40 3.401 Dynamic [235]
modied MMT compression 1 4.380(0.5) 0.223(2) Intercalation/Exfoliation) 40(0) 3.914(15)
(Cl 30B) moulding 3 4.970(14) 0.288(32) 40(0) 4.901(44)
5 4.990(15) 0.298(36) 40(0) 5.577(64)
PLA-KF bre Organically Melt blending/ 0 6.25 0.25 From [236]
modied MMT injection moulding 3 6.21(1) 0.26(4) unloading
(OP)
PHBV Organically Solution 0 0.761 0.046 Homogeneous dispersion 0.633 Dynamic ISE [237]
modied MMT intercalation/Solvent 1 0.881(16) 0.051(11) 1.043(65) (<2.0 lm)
(Cl 30A) casting 2.5 1.270(67) 0.077(67) 1.311(107)
5 1.546(103) 1.677(165)
PP/PP-g-MA Organically Single-screw 0 2.18 TEM (different degree of Dynamic [218]
(90/10) modied MMT extrusion with in-line SSE-3 2.29(5) dispersion for TSE and SSE
(Cl 15A) SCCO2 (SSE) Twin- TSE-3 2.33(6) Intercalation/Exfoliation)
screw extrusion (TSE) SSE-5 2.22(2)
TSE-5 2.50(13)
PP-g-MA Organically Melt blending/ 0 TEM Mixture of highly Dynamic ISE [219]
modied MMT injection moulding 2.5 (68) intercalated and well exfoliated (12) (<0.2 lm)
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

(I31PS) 5 (64) (20)


7.5 (76) (24)
10 (98) (32)
HDPE/HDPE- Organically Melt blending/ 0 1.07 0.050 SEM Agglomeration/well 67.1 0.793 0.663 From [238]
g-MA modied MMT compression 15A-1 1.11(4) 0.050(0) dispersed depending on type and 63.6(5) 0.738(7) 0.617(7) unloading
(98/2) (Cl 15A, N1.44P) moulding 15A-2.5 1.19(11) 0.065(30) composition Intercalation/ 63.7(5) 0.756(5) 0.670(1) (OP)
15A-5 1.18(10) 0.071(42) Exfoliation 60.4(10) 0.864(8) 0.581(12)
N1.44P-1 1.07(0) 0.048(4) 62.4(7) 0.461(42) 0.569(14)
N1.44P- 1.06(1) 0.057(14) 62.9(6) 0.912(15) 0.621(6)
2.5
N1.44P-5 1.20(12) 0.061(22) 63.9(5) 0.750(5) 0.664(0.2)
Table 4 (continued)

Matrix Nanoller type Processing method Nanoller Quasi- Dynamic Hardness Morphology (Dispersion, Crystallinity E0 DMA (GPa) Y (GPa) Method of Depth Ref.
content static DSI E0 H(GPa) interaction) (%) (% (DE0 DMA, %) (DY, %) analysis variation
(wt%) DSIE(GPa) (GPa) (DH, %) variation) (range)
(DE, %) (DE0 ,%)

PEN Organically Melt blending/ 0 5.02h TEM (intercalation) 29.1h 2.83h Dynamic [239]
modied MMT compression 2 5.39(7)h 28.8(1)h 3.53(25)h
moulding
PEO Na+MMT (G105) Solution (S) Melt S-0 0.621 0.030 X-ray (intercalation) From ISE (1 [240]
Organo-clay(I28) blending (M) S-G105-5 0.421(32) 0.025(17) unloading 5.5 lm)
S-G105-20 1.694(173) 0.054(80) (OP)
S-G105-50 2.279(267) 0.088(193)
S-G105-80 1.619(161) 0.104(247)
S-I28-5 0.751(21)
S-I28-20 1.162(87) 0.058(93)
M-0 0.821 0.043
M-G105-5 0.966(18) 00.36(16)
M-G105- 1.544(88) 0.054(26)
20
PPy Organically Electro-deposition 0 2.3 0.125 SEM (homogeneous dispersion) NA [241]
modied MMT 0.01 3.3(44) 0.17(36)
Layer structure 0.05 2.7(17) 0.18(44)
0.1 2.5(9) 0.125(0)
0.2 1.3(44) 0.03(76)
0.5 1.2(48) 0.05(60)
PPC Organically Solution intercalation 0 1.414 0.105 TEM (aggregation/well dispersed Dynamic ISE [242]
modied MMT Interc 1.950(38) 0.114(9) Intercalation/Exfoliation) (<0.2 lm)
(Cl 20B) Exf 2.340(66) 0.121(15)
PS Organically Melt blending/ 0 5.23 0.361 Intercalation Dynamic NA [243]
modied MMT compression 0.5 5.51(5) 0.381(6)
(Cl 15A) moulding
PS Surface modied Melt blending/spin 0 3.6 0.290 Dynamic ISE [244]
MMT (Cl 20A) coated 10 5.1(42) 0.390(35) (<0.4 lm)
PDDA Na+MMT Layer-by-layer 50 layer 0.42 AFM (smooth surface Dynamic Reverse [245]
Multilayer lm deposition process 100 layer 9.5 0.32 Intercalation MMT layers) ISE (0.2
2 lm)
Chitosan/HA Na+MMT Solution intercalation Chi-0 5.29 0.16 AFM (well distributed NA [246]
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

(60/40) Chi-10 8.28(57) 0.28(75) nanoparticles) FTIR (strong


Chi/HAP-0 7.02(33) 0.22(38) interfacial interaction.
Chi/HAP- 9.15(73) 0.28(75) Intercalation)
10
Epoxy Organically Stirring + sonication/ 0 2.87 0.209 TEM, X-ray (agglomeration/well 2.37 From ISE [247]
modied MMT casting Curing in 1 3.08(7) 0.212(1) dispersed depending on 2.68(13) unloading (<2.0 lm)
(Cl 93A) oven 2.5 3.22(12) 0.216(3) composition. Exfoliation 2.73(15) (OP)
5 3.30(15) 0.218(4) <2.5 wt% Intercalation >2.5;wt%) 2.79(18)
7.5 3.45(20) 0.220(5) 2.85(20)

(continued on next page)


31
32
Table 4 (continued)

Matrix Nanoller type Processing method Nanoller Quasi- Dynamic Hardness Morphology (Dispersion, Crystallinity E0 DMA (GPa) Y (GPa) Method of Depth Ref.
content static DSI E0 H(GPa) interaction) (%) (% (DE0 DMA, %) (DY, %) analysis variation
(wt%) DSIE(GPa) (GPa) (DH, %) variation) (range)
(DE, %) (DE0 ,%)

Epoxy Na+MMT Sonication (S) 0 4.01 0.145 SEM (clay aggregates. 1.67 From [248]
Silylated MMT Sonication + ball Na-S-3 4.10(2) 0.162(12) Intercalation) 1.79(7) unloading
(MMT-A1100, milling (SB)/casting A1100-S-3 4.30(7) 0.195(35) 1.9(16) (OP)
MMT-A1120) Curing in oven A1120-S-3 4.34(8) 0.220(52) 1.94(16)
Na-SB-3 3.70(8) 1.83(10)
A1100-SB- 4.20(5) 0.178(23) 1.84(10)

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


3
A1120-SB- 4.29(7) 0.157(8) 1.87(12)
3
CEAR Na+ MMT Stirring/ 0 3.6 0.172 TEM (clay randomly dispersed. 1.1 NA [249]
(NMMT) electrodeposition 1 4.0(11) 0.175(2) Intercalation/Exfoliation) 1.18(7)
Curing in oven 2 4.1(14) 0.175(2) 1.2(9)
4 4.2(17) 0.188(9) 1.2(9)
5 4.25(18) 0.188(9) 1.4(27)
AAER Na+ MMT Stirring/ NMMT 2.9 X-ray (intercalation) Dynamic [250]
(NMMT) electrodeposition lm 5.0 (72)
Hydrothermal Curing in oven HMMT
method lm (8%
Laminates polymer)
(HMMT)
UPE Organically Stirring/casting 0 5.39 0.301 X-ray (intercalation) From [251]
modied layered Curing in oven 1 6.19(15) 0.387(29) unloading
silicate 3 6.07(13) 0.372(24) (OP)
5 6.65(23) 0.343(14)
PU Organically In situ 0 0.014i 0.0017 X-ray (intercalation/Exfoliation) From [252]
modied MMT polymerization/ C20-1 0.017(21) 0.0018(6) unloading
(C20, C30) casting Curing in C20-3 0.021(50) 0.0020(18) (OP)
oven C20-5 0.023(64) 0.0024(41)
C30-1 0.021(50) 0.0022(29)
C30-3 0.025(79) 0.0026(53)
C30-5 0.029(107) 0.0028(65)

a
Polished samples.
b
Unpolished samples.
c
Measured by WAXS.
d
Dried as moulded samples (DAM [230]).
e
Results showing nose effect at the beginning of unloading.
f
PA-6,6 reinforced with OMMT rst and then blended with SEBS-g-MA (N3 in [231]).
g
Reduced modulus.
h
Annealed samples.
i
Material with matrix modulus < 0.15 GPa, where it is difcult to determine the initial point of contact.
Table 5
Nanomechanical and macroscopic properties for polymer nanocomposites incorporating inorganic spherical nanollers. Columns as indicated in Table 1.

Matrix Nanoller type Processing method Nanoller Quasi- Dynamic Hardness H Morphology (Dispersion, Crystallinity E0DMA (GPa) Y (GPa) Method of Depth Ref.
content static DSI DSI E0 (GPa) interaction) (%) (% (DE0DMA , %) (DY, %) analysis variation
(wt%) E (GPa) (GPa) (DH, %) variation) (range)
(DE, %) (DE0 , %)

Epoxy Nanosilica 25 nm Mechanical mixing 4(40 C) 2.35.5a SEM (temp. dependent From [262]
preheating 4(80 C) 7.4a cluster size) unloading
ultrasonication 4(100 C) 8.8a (OP)
Epoxy Fumed silica 20 nm Ultrasonication 0 3.39 0.318 SEM, Optical microscope 3.3 From [263]
forming 200 mechanical mixing 0.5 3.33(2) 0.299(6) (clusters) 3.25(2) unloading
300 nm clusters curing RT + 90 C 1 3.00(12) 0.295(7) 2.90(12) (OP)

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


2 3.05(10) 0.273(14) 2.96(10)
3 2.55(25) 0.256(19) 2.48(25)
Epoxy Nanosilica 10 nm Ultrasonication 0 3.30a 0.143 TEM (good dispersion) 1.46c From [264]
mechanical mixing 1 3.58(8)a 0.160(12) 1.77(21)c unloading
curing RT 3 3.87(17)a 0.177(24) 2.20(51)c (OP)
5 4.00(21)a 0.180(26) 2.86(96)c
Epoxy Surface-modied Solgel 0 2.51 0.190 3.27 From [76,267]
nanosilica 20 nm 5.3 2.66(6) 0.210(11) 3.60(10) unloading
10.3 2.72(8) 0.215(13) 3.71(13) (OP)d
13.3 2.96(18) 0.222(17)
16.7 3.07(22) 0.225(18) 3.87(18)
24.2 3.51(40) 0.250(32)
Epoxy Surface-modied Solgel 0 5.88e 2.90e Dynamic ISE (<200 lN) [75,267]
nanosilica 25 nm 1.8 6.08(5)e 3.13(8)e
5.3 6.69(14)e 3.38(17)e
10.3 6.74(15)e 3.55(22)e
16.7 6.94(18)e 3.71(28)e
22.7 8.59(46)e 4.07(40)e
Epoxy PMMA + fumed Blending PMMA + SiO2 0 3.80a 0.143f SEM, AFM (good dispersion of From [268]
nanosilica 14 nm mechanical mixing 0.1 3.74(2)a 0.136(5)f PMMA domains and SiO2 unloading
curing 50 C + 150 C 0.2 3.75(1)a 0.138(3)f particles in the domains)
0.3 3.96(4)a 0.147(3)f
0.4 3.96(4)a 0.146(2)f
0.5 3.96(4)a 0.146(2)f
Poly- Nanosilica 8 nm Spin coating thermal 0 3.84 0.283 TEM (local aggregations) Dynamic [77]
phenylene curing 5.2 4.02(5) 0.285(1)
SiLKI 12.7 4.50(17) 0.315(11)
21.4 4.32(13) 0.320(13)
31.9 4.66(21) 0.330(17)
42 5.77(50)
Poly- Nanosilica 8 nm Spin coating thermal 0 3.26 0.240 TEM (local aggregations) Dynamic [77]
phenylene curing 10.5 3.45(6) 0.250(4)
SiLKD 16.2 3.74(15) 0.251(5)
21.4 4.01(23) 0.268(12)
26.7 4.81(47) 0.308(28)
MDMA Nanosilica 10 Ultrasonication UV 0 1.91 0.078 0.650 Dynamic ISE (<0.2 lm) [210]
15 nm curing 1 3.09(62) 0.131(68) 1.20(85)
2 3.62(89) 0.203(160) 1.45(123)

33
(continued on next page)
Table 5 (continued)
34
Matrix Nanoller type Processing method Nanoller Quasi- Dynamic Hardness H Morphology (Dispersion, Crystallinity E0DMA (GPa) Y (GPa) Method of Depth Ref.
content static DSI DSI E0 (GPa) interaction) (%) (% (DE0DMA , %) (DY, %) analysis variation
(wt%) E (GPa) (GPa) (DH, %) variation) (range)
(DE, %) (DE0 , %)

3 3.78(98) 0.234(200) 1.44(123)


5 4.11(115) 0.250(221)
PMA Nanosilica 5 UV curing 0 1.85 0.057 TEM (good dispersion below From ISE [270]
15 nm silanized 19.6 3.64(97) 0.262(360) 35 wt%) unloading
grafted 35.5 4.41(138) 0.360(532) (OP)
48.5 5.39(191) 0.492(763)
PVA Nanosilica Solution mixing 0 0.3 SEM (some aggregation Gradual 0.3 0.3 From E constant [78]
functionalized 28.1 3.4(1000) prevented by HA coating) decrease 3.22(970) 3.0(900) unloading (0.22.0 lm)
80 nm 37.5 4.8(1500) with ller 4.87(1500) 5.0(1600) (OP)
44.0 6.8(2200) content 6.65(2100) 6.5(2100)
54.6 8.1(2600) 7.20(2300) 8.0(2600)
54.6 + HA 11.8(3800) 11.5(3700) 13.6(4400)
PEEK Surface-treated Melt mixing 400 C 0g 10.7a 0.631 SEM, AFM (aggregates) From [272]
fumed silica 13 nm 1 11.2(5)a 0.657(4) unloading
+ 25 wt% CF 1.5 14.6(36)a 0.727(15) (OP)
2 15.4(44)a 0.768(22)
PMMA Nanosilica 3 nm Silanization of SiO2 0 3.84 0.243 SEM (good deagglomeration 1.627 From Reverse ISE [273]
forming aggregates sonication in situ 3 3.88(1) 0.258(6) by SC) 1.885(16) unloading (<10 mN)
polymerization thermal 3 (SC gel) 3.88(1) 0.276(14) 2.384(47) (OP)
curing 3 (SC sol) 4.56(19) 0.326(34) 2.502(54)
PS Silica nanodomains Block copolymerization 0 5.1a 0.28 TEM (vinyl phenol favours NA ISE (<0.2 lm) [274]
3545 nm microphase separation 17.2 7.2(47)a 0.40(43) microphase separation) Reverse ISE
calcination 20.8 5.5(8)a 0.34(21) (>0.2 lm)
PS Silica nanodomains Block copolymerization 0 7.4a,g 0.25g TEM (vinyl phenol favours NA ISE (<0.2 lm) [274]
3545 nm microphase separation 17.2 13.5(82)a,g 0.34(36)g microphase separation) Reverse ISE
calcination 20.8 10.8(46)a,g 0.37(48)g (>0.2 lm)
Poly-siloxane Nanosilica Spin coating curing 0 3.0 TEM (some aggregation) From [276]
functionalized 100 C 5 vol% 3.6(20) unloading
15 nm grafted 10 vol% 4.3(43) (OP)
15 vol% 5.1(70)
Poly-siloxane Nanosilica Spin coating curing 0 3.0 TEM (some aggregation) From [276]
functionalized 100 C 5 vol% 3.6(20) unloading
35 nm grafted 10 vol% 4.0(33) (OP)
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

15 vol% 4.3(43)
PMMA Al2O3 39 nm Vortex mixing in 0 3.3 SEM (clusters) 1.8 2.3 Dynamic [278]
Fe3O4 90 nm solution 1.5 Al2O3 1.7(48) 0.9(50) 2.1(9)
1.2 Fe3O4 1.3(61) 1.3(28) 2.0(13)
PS Al2O3 39 nm Vortex mixing in 0 4.3 SEM (clusters) 2.1 2.8 Dynamic [278]
Fe3O4 90 nm solution 1.4 Al2O3 0.2(95) 2.0(5) 2.7(4)
1.1 Fe3O4 3.8(12) 1.8(14) 2.5(11)
PDMS c-Al functionalized Sonication 0.21 0.0043b SEM From [279]
grafted Hydrosylilation 0.36 0.0029b unloading
0.56 0.0021b (OP)
0.96 0.0022b
1.11 0.0024b
Table 5 (continued)

Matrix Nanoller type Processing method Nanoller Quasi- Dynamic Hardness H Morphology (Dispersion, Crystallinity E0DMA (GPa) Y (GPa) Method of Depth Ref.
content static DSI DSI E0 (GPa) interaction) (%) (% (DE0DMA , %) (DY, %) analysis variation
(wt%) E (GPa) (GPa) (DH, %) variation) (range)
(DE, %) (DE0 , %)

iPP IF-WS2 80 nm Melt-blending/ 0 2.70 0.153(0) 0.604 1.5 1.27 Dynamic E0 and H [22]
Hot-compression 0.05 2.41(11) 0.147(4) 0.582(4) 1.29(2) constant
0.1 2.70(0) 0.152(1) 0.598(1) 1.6(7) 1.38(9) (P1.5 lm)
0.25 3.00(11) 0.175(14) 0.626(4) 1.53(21)
0.5 2.95(9) 0.172(12) 0.619(2) 1.57(24)
1.0 2.93(9) 0.165(8) 0.634(5) 1.75(38)
2.0 3.10(15) 0.182(19) 0.649(7) 1.71(35)
4.0 3.13(16) 0.175(14) 0.641(6) 1.8(20) 1.81(43)

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


8.0 3.19(18) 0.180(18) 0.657(9) 2.0(33) 1.86(46)
PPS IF-WS2 80 nm Melt-blending/Hot- 0 3.70 0.242 0.519 1.8 2.18 Dynamic E0 and [22]
compression 0.05 3.70(0) 0.251(4) 0.473(9) 2.3(28) 2.15(1) Hconstant
0.1 4.00(7) 0.285(18) 0.521(0) 2.1(17) 2.35(8) (P1.5 lm)
0.5 4.70(26) 0.359(48) 0.596(15) 3.25(81) 2.77(27)
1.0 4.50(20) 0.336(39) 0.582(12) 3.20(78) 2.84(30)
2.0 4.95(34) 0.382(58) 0.607(17) 2.75(53) 2.58(18)
4.0 4.80(29) 0.350(45) 0.603(16) 2.70(50) 2.86(32)
8.0 5.00(35) 0.390(61) 0.619(19) 3.25(81) 2.97(36)
PEEK IF-WS2 80220 nm Aerosol-assisted 0 6.7 0.33 SEM (some aggregation) Possible NA [282]
deposition 1 7.5(12) 0.32(3) increase due
2.5 7.2(8) 0.35(6) to ller
5 7.9(18) 0.43(30)
10 10.1(51) 0.48(45)
20 10.7(60) 0.55(67)
Collagen NanoHA 100 Electrostatic co- 0 0.2 0.026 SEM (aggregations at higher 0.0062 Dynamic [286]
150 nm spinning 10 0.5(150) 0.060(130) load) 0.23(3600)
20 0.6(200) 0.100(290)
Collagen NanoHA 100 Electrostatic co- 0 0.52 0.068 SEM (aggregations at higher Dynamic [286]
crosslinked 150 nm spinning 10 0.96(85) 0.100(47) load)
20 1.10(110) 0.118(74)
MEH-PPV Cd Se Solution mixing 0 9.6h 0.23h TEM (better dispersion at From [287]
ultrasonication spin 75 24.3(150)h 0.70(200)h higher loads) unloading
casting annealing 95 38.4(300)h 0.83(260)h (OP)
120 C
MEH-PPV Cd Se Solution mixing 0 11.7i 0.23i TEM (better dispersion at From ISE (<40 [287]
ultrasonication spin 75 27.5(140)i 0.43(87)i higher loads) unloading 80 nm)
casting annealing 95 40.2(240)i 0.79(240)i (OP) Reverse ISE
120 C (>80 nm)

a
Results showing nose effect at the beginning of unloading.
b
Materials with modulus <0.15 GPa, where it is difcult to determine the initial point of contact.
c
Measured at 0 C.
d
Conical indenter.
e
Dynamic 10 Hz.
f
Martens hardness.
g
Composites on glass.
h
Constant load 10 lN/s.

35
i
Constant load 100 lN/s.
Table 6

36
Nanomechanical and macroscopic properties for miscellaneous inorganic nanoparticle-reinforced polymer composites. Columns as indicated in Table 1.

Matrix Nanoller type Processing method Nanoller Quasi-static Dynamic Hardness Morphology (Dispersion, Crystallinity E0 DMA Y (GPa) Method of Depth Ref.
content DSI E (GPa), DSI E0 (GPa), H (GPa) interaction) [%] (% (GPa) (DY, %) analysis variation
0
(wt%) (DE, %) (DE , %) (DH, %) variation) (DE0 DMA, (range)
%)

Epoxy IF-WS2 nanotubes Solution mixing/ 0 3.47a 0.192 SEM/TEM (good dispersion) From [289]
60 nm curing 3.0 4.44(28)a 0.190(1) unloading
(OP)b
PA-6 Oib-POSS Melt-blending 0 0.137a,c,d 2.2 0.024 AFM/TEM (large irregular 24.4 1.34 From ISE [292]
5.0 0.776(466)a,c 4.0(82) 0.075(212) aggregates) 29.2(20) 1.77(32) unloading (<0.3 lm)

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


a,c
10.0 2.23(1527) 4.8(118) 0.166(590) 35.3(45) 1.81(35) (OP) and
dynamic
PA-6 Tsp-POSS Melt-blending 0 0.137a,c,d 2.2 0.024 AFM/TEM (small regularly 24.4 1.34 From ISE [292]
5.0 0.987(620)a 4.3(95) 0.066(175) dispersed POSS domains) 27.4(12) 1.91(42) unloading (<0.3 lm)
10.0 2.380(1637)a 5.0(127) 0.157(554) 29.4(21) 1.96(46) (OP) and
dynamic
PVDF FP-POSS Solution casting 0 2.61e 0.174 OM (POSS aggregates) gradual 1.80 1.71 From E,H [293]
3.0 2.77(6)e 0.177(2) increase with 2.28(27) 2.00(17) unloading constant
5.0 2.33(11)e 0.158(9) ller content 2.13(18) 1.59(7) (OP). (0.5
8.0 2.21(15)e 0.130(25) 1.86(3) 1.17(32) 2 lm)
PU GPTS-modied Solgel/Spin 0 8.98 SEM (homogeneous NA [290]
AlOOH nanorodf 10 coating/thermal 40.0 8.86(1) 0.83 dispersion)
20 nm curing
PU GPTS-modied Solution mixing/ 0 8.98 SEM (homogeneous NA [290]
AlOOH nanorodf 10 casting/UV curing 5.0 9.22(3) dispersion)
20 nm
PU GPTS-modied Solgel/Spin 0 8.98 SEM (isotropic dispersion) NA [290]
AlOOH nanoparticle coating/thermal 40.0 9.88(10) 0.98
curing
PU GPTS-modied Solution mixing/ 0 8.98 SEM (isotropic dispersion) NA [290]
AlOOH nanoparticle casting/UV curing 5.0 9.43(5)
PU Silica network Solgel/UV curing 0 2.75 0.13 OM (homogenous From [298]
generatedin situfrom 20 3.28(19) 0.16(23) dispersion) unloading
TEOSg 50 3.73(36) 0.18(38) (OP)
PMMA Eu:Gd2O3 Solution casting 0 4.28a 0.263 SEM (agglomerates) 2.15 From [300]
nanoparticles 40 nm 3.0 4.82(13)a 0.268(1) 2.60(21) unloading
(OP)
PMMA APTES-modied Solution casting 0 4.28a 0.263 SEM (good dispersion) 2.15 From [300]
Eu:Gd2O3 3.0 5.60(31)a 0.293(11) 3.05(42) unloading
(OP)
PMMA Trioctylamine- Solution mixing/ 0 2.20 0.137 From [288]
functionalized annealing 3.0h 4.81(118) 0.285(108) unloading
graphene-like BN 3.0i 5.11(132) 0.310(126) (OP)
PMMA-co- Silica network In situ 0 4.10 0.25 SEM (transparent, no phase From Reverse [296]
MPTES generated in situ from polymerization/Sol 22 7.60(85) 0.50(100) separation) unloading ISE
TEOSg gel/Spin coating/ 47 6.60(61) 0.54(116) (OP)
curing 72 9.50(132) 0.85(240)
Table 6 (continued)

Matrix Nanoller type Processing method Nanoller Quasi-static Dynamic Hardness Morphology (Dispersion, Crystallinity E0 DMA Y (GPa) Method of Depth Ref.
content DSI E (GPa), DSI E0 (GPa), H (GPa) interaction) [%] (% (GPa) (DY, %) analysis variation
0
(wt%) (DE, %) (DE , %) (DH, %) variation) (DE0 DMA, (range)
%)

PMMA Silica network In situ 0 6.90 0.32 SEM (transparent, no phase From Reverse [297]
generated in situ from polymerization Sol 17 6.30(9) 0.45(41) separation) unloading ISE
TEOS gel/Spin coating/ 43 8.20(19) 0.56(75) (OP)
curing
PDDAj CSH nanoparticles Stirring/ 4.44(83)k 0.11(80)k TEM (high packing density) From E,H [291]
precipitation 100 26.8 1.09 unloading constant

A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194


(OP) (1512
mN)
PI POSS-OH Solution casting/ 0 2.9 0.22 SEM (homogenous NA [294]
thermal imidization 3.0 3.0(3) 0.22(0) dispersion. Aggregates at
5.0 3.2(10) 0.23(5) 10 wt%; strong H-bonding)
10.0 3.5(17) 0.24(9)
PI OC-POSS Solution casting 0 4.80 0.48 3.30 From [295]
7.8l 4.70(2) 0.48(0) 3.10(6) unloading
15.4l 4.70(2) 0.44(8) 3.00(10) (OP)
30.2l 4.80(0) 0.40(17) 3.30(0)
44.5l 5.20(8) 0.40(17) 3.00(10)
PI OH-POSS Solution casting 0 4.80 0.48 3.30 From [295]
7.0l 4.75(1) 0.45(6) 3.40(3) unloading
13.9l 4.60(4) 0.40(16) 3.70(12) (OP)
27.9l 4.65(3) 0.38(21) 3.20(3)
41.8l 4.20(12) 0.34(29) 3.00(10)

a
Reduced modulus.
b
Spherical indenter.
c
Results showing nose effect at the beginning of unloading due to short holding time.
d
Materials with matrix modulus <0.15 GPa, where it is difcult to determine the initial point of contact.
e
Results affected by thermal drift correction.
f
Aligned along the surface direction.
g
Covalently grafted to the matrix.
h
DMF solvent.
i
Chloroform solvent.
j
Intercalated between the nanoparticle lamellae.
k
Percentage of variation compared to the pure nanoparticles (cement).
l
Calculated theoretically.

37
38 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

composed of more than two coaxial cylinders, each rolled out of single sheets, with diameters between
2 and 40 nm, the multi-walled carbon nanotubes (MWCNTs). Both types of CNTs can be synthesized
using different routes [86], including high-temperature evaporation using arc-discharge or laser abla-
tion, and gas-phase catalytic growth from carbon monoxide or chemical vapour deposition (CVD) from
hydrocarbons. CVD materials generally contain residual catalyst particles, while the main contami-
nants in the products of the high-temperature reactions are carbonaceous impurities. The laser pro-
cess generates CNTs of the highest quality (few defects, high crystallinity, extremely high aspect
ratio) compared to the other synthesis methods [87], while nanotubes produced by the CVD technique
present a larger number of defects as a result of the lower growth temperature. The purity, quality,
aspect ratio and nature of impurities, hence the source of the CNTs, can have strong inuence on
the nal properties of CNT-reinforced composites.
In order to fabricate polymer/CNT composites with enhanced mechanical performance, two main
challenges need to be addressed. Firstly, it is very difcult to attain homogenous CNT dispersion since
these nanollers have a strong tendency to gather and form bundles owing to strong van der Waals
interactions. Moreover, their large aspect ratio causes extremely high viscosity in the polymer-CNT
blend, which in turn affects their uniform distribution. Secondly, the inert nature of the CNT surface
results in poor interfacial adhesion with the host matrix, which prevents efcient stress transfer dur-
ing loading [88]. To solve these issues, several methods have been reported such as mechanical disper-
sion (i.e. ultrasonication [89], ball-milling [90]), plasma treatment [91] and chemical modication
involving either non-covalent or covalent bonding between nanotubes and polymers [92]. The non-
covalent approach consists in the physical adsorption and/or wrapping of polymers to the CNT surface.
The graphitic CNT sidewalls can interact with conjugated polymers via p  p stacking as well as with
those containing heteroatoms with free electron pairs. This route preserves the nanotube integrity and
properties, since it does not destroy the conjugated system of the CNT sidewalls. The covalent method
involves the chemical bonding (grafting) of polymer chains to CNTs, and can be performed via graft-
ing to or grafting from strategies. The rst approach is based on the synthesis of a polymer deriv-
ative that can react with the surface of pristine, oxidized or functionalized CNTs [93]. However, the
oxidation/functionalization treatments performed in acid media generally shorten the CNTs and can
bring signicant damage such as sidewall opening or tube breakage [94], introducing defects in the
tubular framework that can adversely impact their mechanical properties. Another disadvantage of
this method is that the grafted polymer content is restricted due to the low reactivity and high steric
hindrance of the polymer chains. In the grafting from strategy the polymer is grown from the CNT
surface via in situ polymerization of monomers initiated by chemical species immobilized on the CNT
sidewalls and edges. The high reactivity of monomers makes the grafting from process efcient and
controllable, enabling the preparation of nanocomposites with high degree of grafting. However, this
method requires strict control of the amounts of each reactant and the polymerization conditions.
Alternatively, CNT functionalization can be performed by plasma treatment [91,95,96], a time saving
and environmentally friendly technique for modifying CNTs by directly introducing a high density of
functional groups. The CNT surfaces can readily change from hydrophobic to hydrophilic with plasma
processing, thus facilitating the dispersion within the polymer matrix. Furthermore, this treatment
does not alter the intrinsic mechanical properties of the tubes.

4.1.2. Preparation of polymer/CNT nanocomposites


Several methods have been reported for the preparation of polymer/CNT nanocomposites, includ-
ing solution mixing [28,97103], milling [104,105], melt-compounding [20] and in situ polymerization
[106,107]. In the solution processing route, the CNT powder is typically dispersed in a liquid medium
by stirring and/or sonication and then mixed with the polymer solution, followed by evapouration of
the solvent. Ball milling is a method employed to reduce particle size and, in turn, alter surface prop-
erties of materials by modifying surface area, porosity and shape. This technique can be used to pre-
pare a ne polymer/CNT power that is subsequently melt-mixed with a twin-roll or extruder. Melt-
processing is the most widespread technique for the fabrication of CNT-reinforced thermoplastic com-
posites, and is also suitable for polymers with low solubility in common solvents. It consists in the
blending of the polymer melt with the CNT material by application of intense shear forces, and is often
combined with a shape-forming step like injection moulding, spinning or hot-compression. The in situ
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 39

polymerization of monomers in the presence of CNTs has been extensively explored for the prepara-
tion of polymergrafted nanotube composites. This technique is particularly important for insoluble
and thermally unstable matrices, and produces homogenous composites with high CNT weight
fraction.

4.1.3. Modulus changes in polymer/CNT nanocomposites: comparison of thermoplastic and thermoset


matrices
Table 1 gathers nanoindentation results collected from the literature on polymer/CNT nanocom-
posites [20,24,91,95106,108127]. The percentage of change in the elastic modulus DEI (either
quasi-static E or dynamic E0 ) of thermoplastic and thermoset matrices as a function of the nanotube
loading is represented in Fig. 6a and b, respectively. These gures show that in general, CNTs have

1200
(a) Chitosan/MWCNT [98]
PHB/SWCNT [97]
1000 PP/MWCNT [106]
PA-6/MWCNT [20]
PA-6/MWCNT [100]
PLLA/MWCNT [101]
UHMWPE/SWCNT [96]
300
UHMWPE/MWCNT [105]
E [%]

UHMWPE/MWCNT [104]
PVDF/MWCNT [120]
PVA/MWCNT [103]
200
PVA/SWCNT [121]
PHO/SWCNT [97]
PVK/MWCNT [103]
PMMA/MWCNT [28]
100 PMMA/MWCNT [28]
PMMA/MWCNT [127]
PMMA/MWCNT [127]
PI/SWCNT [126]
0 PDMS/MWCNT [124]

0 2 4 6 8 10
CNT [wt%]

120
(b) [115]
100 [91]
[91]
[57]
80 [24]
[109]
[110]
E [%]

60 [111]
[111]
[117]
[117]
40
[112]
[113]
[113]
20
[108]

0 1 2 3 4 5 6
CNT [wt%]

Fig. 6. Percentage of change in the elastic modulus (DEI) of carbon nanotube-reinforced polymer nanocomposites as a function
of the CNT loading: (a) thermoplastic and (b) thermoset matrices. Solid and open symbols correspond to raw and functionalized
CNTs, respectively; s and h in Fig. 6a represent TEOS and APTES functionalized MWCNTs, respectively; indicates CNTs grafted
to the polymer matrix, + and j denote aligned MWCNTs and  designates silica-coated CNTs. Symbols in red highlight the data
that should be taken with caution (see Table 1 for details).
40 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

a positive effect on EI, reinforcing the polymer matrix already at very low loadings. The maximum CNT
content incorporated is 10 wt% due to the nanotube tendency to agglomerate into bundles. It is note-
worthy that the maximum ller concentration is signicantly lower for the thermoset matrices
(5 wt%) compared to the thermoplastic ones (10 wt%). As will be seen below, this is a direct con-
sequence of the commonly encountered difculties in producing epoxy-based composites with low
agglomeration and high degree of polymer bonding. In the case of thermoplastic matrices, signicant
DEI are achieved at high loadings, reaching 200% increase for ller loadings of 8 wt% (see Fig. 6a), while
much more limited EI improvements are found for the thermoset matrices.

4.1.3.1. Thermoplastic matrices. A number of papers dealing with thermoplastic-based composites


[20,9698,100,101,103,104,121] have reported a systematic increase of EIwith increasing CNT content
up to around 12 wt% loading. Higher ller loadings are frequently not considered because agglomer-
ation of the nanotubes into bundles starts to play a signicant role and mechanical enhancements
level-off or even decrease. Fig. 6a shows that for low nanotube contents (<2 wt%), the stiffness incre-
ments for composites incorporating raw [106,120] or functionalized CNTs [20,96,98,100,101,104] lie
within the same range; the diversity of methodologies and indentation testing conditions employed
in the analysis of DSI values can induce some data scattering that hides any signicant difference. Sim-
ilarly, it is not clear that composites incorporating SWCNTs exhibit larger modulus improvements
than those using MWCNTs in spite of the superior modulus of the former [128]. The most noticeable
EI increments for low CNT contents (see Fig. 6a) have been explained as summarized below
[96,98,103,106,121,126].
Satyanarayana et al. [126] found a noticeable increase in the dynamic modulus of glassy polyimide
(PI) (59%) upon addition of only 0.05 wt% raw arc-discharge SWCNTs albeit the existence of small
agglomerates was inferred from AFM analysis. The strong nanotube-matrix interfacial bonding
attained via p  p stacking interactions between the aromatic rings of PI and the sp2 hexagonal net-
works of the nanotubes was discussed by the authors of this work to be responsible for the notewor-
thy improvement at such low loadings. Moreover, a partial alignment of the CNTs that could have been
attained during the spin-coating process is also suggested to inuence the measured data.
Interestingly, Cadek et al. [103] found an 80% increment in the modulus of polyvinyl alcohol (PVA)
with the incorporation of 1.0 wt% raw arc-discharge MWCNTs, value above those reported for similar
amounts of acid-functionalized MWCNTs embedded in polyamide 6 (PA-6) [20,100] or poly(L-Lactide)
(PLLA) [101] (Fig. 6a). Such signicant improvement is ascribed to the drastic increase in the crystal-
linity of PVA (93%) upon addition of the MWCNTs. The benets of such outstanding crystallinity
increase are twofold: on the one hand, the mechanical properties of the matrix are enhanced; on
the other hand, the nucleation of PVA crystals resulted in an extremely strong CNT-matrix interfacial
interaction, as demonstrated by TEM observations, which showed that the matrix did not fail at the
polymernanotube interface. Surprisingly, Prasad et al. [121] reported a DE of 1180% for PVA upon
addition of only 0.6 wt% acid-functionalized SWCNTs, attributed to the strong bonding between the
functionalized CNTs and the matrix combined with the very high SWCNT aspect ratio. Nevertheless,
this result could be inaccurate since the matrix modulus is quite low (E = 0.7 GPa) and hence prone
to a difcult determination of the point of initial contact, as discussed in Section 3. In addition, the role
of water content on the indentation property improvement of this type of hydrophilic matrices has not
been conveniently explored and could certainly shed some light on the mechanisms for such extraor-
dinary mechanical enhancement.
A very efcient non-covalent functionalization approach was reported by Wu et al. [98] for chito-
san based nanocomposites. MWCNTs were wrapped by poly(3,4-ethylenedioxythiophene)-poly(sty-
rene sulphonate) (PEDOT-PSS), a surfactant that improved the nanoller-matrix compatibility. The
sulphonic acid groups of PSS enabled ionic linkage with positively charged amine groups of chitosan,
while PEDOT was bound to the MWCNT surface through p  p and hydrophobic interactions. As a
result, a 52% improvement in modulus was attained at 0.5 wt% loading.
On the other hand, Samad and Sinha [96] reported a D E0 of 58% for ultra-high molecular weight
polyethylene (UHMWPE) coatings upon addition of only 0.2 wt% plasma-treated SWCNTs (see
Fig. 6a). The functionalization via plasma treatment resulted in a large number of oxygen containing
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 41

groups on the nanotube surface that are envisaged to promote bundle disentanglement and improve
the ller interaction with the surrounding matrix, leading to a very uniform dispersion.
Focusing on nanotube concentrations P2 wt%, semicrystalline matrices incorporating acid-func-
tionalized MWCNTs [100,101,104] are found to exhibit enhanced stiffness compared to those rein-
forced with pristine CNTs (Fig. 6a). As mentioned earlier, the oxidation in acid medium leads to a
shortening of the tubes and frequently induces some damage on their sidewalls, drawbacks that seem
to be overweighed by the more homogenous nanotube distribution and stronger CNT-matrix interfa-
cial adhesion attained through this functionalization process, which is reected in larger modulus
increments. As an example, Baji et al. [100] reported remarkable E enhancements (180%) for an array
of randomly electrospun PA-6 bres containing 7.5 wt% of acid-functionalized MWCNTs aligned along
the bre axis. The connement effect of the nanotubes on the mobility of PA-6 chains, together with
the increase in crystallinity and the strong interactions between the carboxylic groups of the MWCNTs
and the amide moieties of the matrix, were suggested to give rise to the noticeable stiffening effect.
However, the work does not address the inuence of orientation effects on the mechanical properties.
Sreekanth and Kanagaraj [104] described one of the largest DE values in thermoplastic matrices. About
206% increase in the modulus was reported for UHMWPE reinforced with 5.0 wt% oxidized MWCNTs
(Fig. 6a). A very homogenous dispersion, combined with a large increase in crystallinity of almost 40%,
are believed to be the reasons for such strong mechanical enhancement. These increments however,
should be contemplated with care because they can be affected by the low modulus of the neat matrix,
in addition to the inadequate correction of the load-depth data for thermal drift (see Section 2.1.1 for a
detailed discussion). Contrary to the aforementioned results, very small E increase (10%) was found
upon addition of 5.0 wt% raw MWCNTs to UHMWPE [105], ascribed to the poor nanoller-matrix
interface, the presence of voids and the nanotube waviness that limits the efciency of the
reinforcement.
Other functionalization approaches like amidation reactions have been employed to improve the
dispersion of large amounts of MWCNTs (>2 wt%) within thermoplastic polymers such as amorphous
poly(methyl methacrylate) (PMMA) [28,102]. However, SEM images revealed the presences of some
aggregates, and hardly increase in modulus was attained (see Fig. 6a). It was suggested that the ex-
ibility of CNTs, their curved morphology and inferior bending properties may contribute to the
reduced reinforcement action. In order to overcome these drawbacks, MWCNTs were coated with a
silica shell encasing the nanotubes. The strategy was found to be most successful yielding DE values
of 130% for 4.0 wt% loading (see Fig. 6a). Mammeri et al. [127] described two novel strategies to func-
tionalize MWCNTs in order to incorporate them in poly(methacrylic acid-co-methacryloxypropyltri-
methoxysilane) P(MMA-co-MPTMS): the covalent grafting of hydrolyzable triethoxysilyl (SiOEt)3
groups on oxidized MWCNTs and the non-covalent adsorption of a polycation, polyallylamine hydro-
chloride (PAH), on pristine CNTs. While the rst approach resulted in very homogenous composites
with moderate improvement in the matrix modulus (26% at 2.0 wt% loading), the second procedure
was insufcient to disaggregate the nanotubes, and no increase in modulus was attained.

4.1.3.2. Thermoset matrices. With regard to nanocomposites based on thermoset matrices, as men-
tioned above, DEI values are in general lower than those found for the thermoplastic counterparts,
by up to 100% in the concentration range of 0.15.5 wt% (Fig. 6b) [57,91,95,108113,115,118], and
even some decrements in stiffness are reported [24,114,117]. This discrepancy is due to the poor
nanoller dispersion in epoxy composites (Table 1), which could be related to diffusion and re-aggre-
gation of CNTs during the curing process [129]. Furthermore, contrary to the tendency observed for
thermoplastic/CNT samples, no signicant stiffness difference is found for composites reinforced with
pristine or functionalized CNTs in all the range of concentrations tested. Results hint at an important
modulus increment for plasma functionalized MWCNTs grafted to the epoxy matrix yielding a
DE = 20% for a loading of only 0.1 wt% [91]. The plasma method enabled incorporating a high density
of primary amine groups onto the nanotube surface that allowed direct covalent bond to the epoxide
groups, thus resulting in improved ller-matrix adhesion and nanotube dispersion, as revealed by SEM
micrographs (Fig. 7). Surprisingly, similar moderate EI increments, with a tendency to increase at
higher ller loadings, have been reported for composites incorporating raw MWCNTs [108,112],
despite their poor dispersion.
42 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

Fig. 7. (a) Representative DSI loaddisplacement curves on neat epoxy and 0.1 wt% unfunctionalized and plasma functionalized
MWCNT-reinforced nanocomposites. (b) and (c) SEM images of the fractured surfaces of the nanocomposites with
unfunctionalized and functionalized CNTs, respectively, showing the difference in the state of nanotube dispersion. Adapted
from Ref. [91], copyright 2013, with permission from Elsevier.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 43

Clear examples on the insignicant moduli differences between epoxy composites lled with either
raw or functionalized CNTs were provided by Lagoudas et al. [113] and Mionic et al. [111]. Interest-
ingly, Mionic demonstrated that the improvements in the dynamic modulus were dependent on the
solvent used for the CNT pre-dispersion and not on the functionalization approach, indicating the lack
of formation of chemical bonds between the constituents. The largest stiffness increase (about 104% at
5.0 wt% acid-treated MWCNTs) was found for acetone, a highly volatile solvent that would evaporate
rapidly enabling a fast polymerization of the matrix. On the other hand, the smallest increment (on
average 20% for the same concentration) was observed for gamma-butyrolactone (GBL), which copo-
lymerizes with the epoxy and acts as a plasticizer, weakening the mechanical properties of the mate-
rial. Li et al. [57] also found relevant DEI of 75% with the incorporation of 5.0 wt% methylbenzoate
functionalized arc-discharge SWCNTs (see Fig. 6b). The functionalized tubes were easily dissolved
in polar solvents and integrated in the matrix via solution processing, resulting in relatively good dis-
persion for all the samples. Intercalation of the polymer into the bundles was proposed as the key rein-
forcing mechanism for these nanocomposites.
In contrast, only small dynamic modulus improvements (around 17% at 5.0 wt% loading) were
reported for epoxy nanocomposites incorporating puried laser-grown SWCNTs [24,114] (Fig. 6b),
regardless of the high quality and aspect ratio of these nanollers, attributed to their curved morphol-
ogy, the presence of a large number of aggregates, the sliding between individual nanotubes within a
bundle and the composite microvoids. Even more striking is the limited DE found when CNT forests
are incorporated to the epoxy matrix. CNT forests are dense, millimeter-tall and vertically aligned
structures grown by water-assisted CVD process and incorporated into the epoxy matrices [115
117]. These nanocomposites are expected to exhibit enhanced mechanical performance when the load
is applied along the vertical direction of the aligned CNTs. However, only modest EI improvements
have been reported [115,117] and, also surprisingly, no signicant differences were observed com-
pared to samples lled with the random counterparts [117]. CNT waviness was suggested to be the
major factor accounting for the deviations of the modulus values of the composites from the rule of
mixtures (Eq. (14)). The authors conclude that a ne control of this parameter would be highly desir-
able to reach a signicant stiffness contribution from the CNTs.
Finally, it is worth mentioning that coiled carbon nanotubes (CCNTs) represent an interesting alter-
native to traditional straight nanotubes for reinforcement of polymer matrices [130]. They exhibit a
Youngs modulus of 0.7 TPa, and their coiled conguration enables an improved bonding to the sur-
rounding matrix, resulting in an efcient reinforcement. Li et al. [118] investigated a series of CCNT-
reinforced epoxy composites with different weight percentage of nanotubes by means of DSI. How-
ever, no conclusion can be drawn from this work because modulus values derived from the onset of
unloading present the well-known nose effect, a fact that critically compromises the validity of
the stiffness data (see Section 2.1.1).

4.1.3.3. Other CNT-reinforced polymer systems. Recently, the layer-by-layer (LbL) assembly of oppo-
sitely charged polyelectrolytes on a charged surface has emerged as an alternative method for inte-
grating functionalized CNTs into polymers [131135]. It is an environmentally sound and cost-
effective approach that can result in nanocomposites with high nanoller weight fraction and con-
trolled internal structure, having the potential to reach the desired mechanical properties through tai-
lored design, and with eventual applications in the development of a new generation of scaffolds. For
instance, in poly(dimethyldiallylammonium chloride) (PDDA)/poly(acrylic acid) (PAA) multilayers
prepared via LbL, the addition of 4.7 wt% acid-treated SWCNTs to PAA led to 120% increment in the
in-plane modulus of the polyelectrolyte lm [131,132], attributed to a homogenous nanotube distri-
bution combined with their alignment parallel to the lm plane. Contrary to the aforementioned
results, Firkowska et al. [133] and Pavoor et al. [135] reported that the presence of acid-treated
MWCNTs in polyethyleneimine (PEI) or PAH lms was detrimental to the mechanical performance
of LbL multilayers of polystyrene sulphonate (PSS)/PEI or PAH/PAA, respectively. The striking fact
about these results is that the modulus values were signicantly lower than those of the LbL lms
made solely of polyelectrolytes. This effect was explained in terms of the lubricant role of the CNTs,
easily rolling or sliding between the different layers of the assembly. It is noteworthy that CNT
44 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

bending, waviness, curved morphology or sliding ability have been posed to explain the limited rein-
forcing effect of the nanotubes in epoxy matrices [24,114,117].
On the other hand, about a 40-fold enhancement in modulus has been reported for isocyanate
microcapsules, with poly(ureaformaldehyde) (PUF) as a shell, upon incorporation of only 1.1 wt%
oxygen plasma functionalized MWCNTs [107]. These novel materials were prepared by in situ poly-
merization, which enabled a very uniform dispersion of the nanotubes within the microcapsule shell.
In addition, the authors argued that oxygen groups from the nanotube surface could strongly interact
with the amide groups of PUF, resulting in extraordinary improved mechanical properties. The inden-
tation data analysis is, however, a matter of concern because modulus values are too low (54 MPa) to
allow for a precise determination of the point of initial contact.
CNTs have also been incorporated into conventional bre-reinforced polymers to fabricate multi-
scale (also named hierarchical) composites with improved mechanical performance [136]. Two alter-
native strategies for the preparation of CNT-based hierarchical composites have been reported: the
dispersion of CNTs into the matrix [122] or their direct anchoring onto the bre surface [119]. Nano-
indentation in these systems is usually directed to produce a map of local mechanical properties aim-
ing at distinguishing between the polymer matrix, the bres and the interface/interphase. For
example, Qian et al. [119] combined nanoindentation studies with AFM and Raman spectroscopy to
probe the local mechanical properties of epoxy composites incorporating MWCNTs grafted onto silica
bres via CVD process. The authors demonstrated that the grafting of MWCNTs to the bres increased
the modulus of the epoxy matrix in the bre surroundings. More recently, the local mechanical prop-
erties of SWCNT-reinforced PEEK/glass bre (GF) laminates have been investigated using nanoinden-
tation [122]. About 9% improvement in the PEEK matrix modulus was attained upon addition of
1.0 wt% puried arc-SWCNTs, increment that was doubled with the incorporation of the same amount
of SWCNTs wrapped in poly(ether ether sulphone) (PEES). More important, the paper demonstrates
that the polymerbre interface can be easily detected using the dynamic technique, and provides
experimental evidence that the extent of the interphase increases if the matrix is reinforced with
SWCNTs, particularly those wrapped in the compatibilizing agent.

4.1.3.4. Comparison with macroscopic properties. The data of Table 1 show that the increments in inden-
tation quasi-static modulus (calculated from OP analysis) increase with increasing ller content fol-
lowing a similar trend than those of tensile or exural moduli [91,100,104,110,112], although the
comparison of the increment values exhibit some disparity [104,110,112]. Y increments have been
found to be higher [105,110], lower [104,112] or in the range [100] of those for E. The differences
between these numbers sometimes approach one order of magnitude, in one or the other direction
[104,105]. The larger DE values obtained by means of indentation have been occasionally attributed
to an indentation size effect due to agglomeration of CNT bundles at the surface [112]. Most surprising
is the fact that modulus increments for indentation and tensile testing were quite close only in the
case of electrospun mats [100]. In this work, molecular chains were aligned along the loading direction
for tensile testing and perpendicularly for indentation.
Although it seems that comparison between E and E0DMA does not have a reasonable ground, some-
times it yields an excellent match. For example, Chen et al. highlighted that the large E increase found
upon addition of functionalized MWCNT to an epoxy resin was in agreement with DMA results (see
Table 1) [91]. In contrast, Wu et al. reported substantial differences between the modulus increments
obtained by dynamic DSI and tensile testing (see Table 1) [98]. The authors of this work highlighted
the consistency of both techniques in revealing a rising trend of modulus with increasing MWCNTs
without further discussion on the relative increments. This limited analysis seems quite judicious in
view of the large differences between both mechanical indentation tests.
Table 1 only offers one example in which the E0 and E0DMA can be contrasted [20]. It is found that the
modulus increments exhibit the same trend but different values. This discrepancy was discussed by the
authors in terms of the different frequencies employed in DMA and DSI (1 Hz vs. 45 Hz, respectively) and
the different loading directions with respect to the plane of the compression moulding lms.
In summary, both indentation analysis and traditional macroscopic mechanical techniques reveal
similar rising trends of the modulus with increasing CNT content. However, the increment values do
not exhibit a clear correspondence.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 45

4.1.4. Hardness
A limited amount of H data, compared to EI, is available for CNT/polymer nanocomposites. This is
especially apparent for CNT-reinforced epoxies (see Table 1). The table shows that the H behaviour
with ller content is qualitatively similar to that discussed previously for the modulus. In the follow-
ing sections, it will be found that this is usually also the case for other nanoller-reinforced polymer
systems. Hence, in order to avoid unnecessary repetitions, the main discussion will be focused on the
modulus data and only some hints of the differences between the elastic and the plastic responses are
given below and also remarked in other sections when appropriate.
As an example, Fig. 8 illustrates the percentage of change in indentation hardness of thermoplastic
matrices as a function of the nanotube loading. The gure shows that the H increments are in general
lower than those of EI (see Fig. 6a), suggesting that CNTs are more effective in enhancing the stiffness
than the hardness of thermoplastic matrices. Mammeri et al. [127] observed that the elastic modulus
is more sensitive than the hardness to the type of surface modication of the CNTs. While similar H
enhancements were recorded by incorporating 2.0 wt% of MWCNTPAHSiO2 or MWCNTCOOH
APTES to P(MMA-co-MPTMS), a higher elastic modulus increment was determined when using the
second type of nanoller, which exhibited improved dispersion. Results suggest that the interfaces
and the degree of dispersion of CNTs inuence in a different manner the elastic and plastic deforma-
tion processes. These ndings are in agreement with the general trend found for larger EI enhance-
ments compared to those of H. Fig. 8a also offers some examples of hardness enhancements
modulated by changes in the degree of crystallinity and crystal size [97,100,103,104], analogous to
those discussed in Section 4.1.3.1 for the modulus. In short, CNTs act as nucleating agents at low load-
ings generally enhancing the matrix crystallinity and decreasing the crystal size. Such structural
changes contribute to the H variations with ller content in terms of the models described in Eqs.
(12) and (13). At large CNT contents, agglomeration has a dual detrimental effect: on the one hand,
the CNT-matrix interfacial area decreases; on the other hand, the reduced efciency of the CNTs as
nucleating agents limits the amount of crystallinity in the polymer matrix.

4.1.5. Indentation size effect


A wide number of studies on CNT/polymer composites have reported a dependence of EI or H on
indentation depth [20,28,98,99,108,111113,125127]. The difculty in discerning whether an ISE
is attributed to genuine material properties or is it due to instrumental effects was pointed out in Sec-
tion 2.1.3. This is especially critical at penetrations below a few hundred nanometers.

650 Chitosan/MWCNT [98]


Chitosan/MWCNT [99]
PHB/SWCNT [97]
PP/MWCNT [106]
150 PA-6/MWCNT [20]
PA-6/MWCNT [100]
UHMWPE/SWCNT [96]
UHMWPE/MWCNT [105]
100 UHMWPE/MWCNT [104]
H [%]

PVA/MWCNT [103]
PVA/SWCNT [121]
PHO/SWCNT [97]
PVK/MWCNT [103]
50 PMMA/MWCNT [28]
PMMA/MWCNT [28]
PMMA/MWCNT [127]
PMMA/MWCNT [127]
PI/SWCNT [126]
0 PDMS/MWCNT [124]

0 2 4 6 8 10
CNT [wt%]

Fig. 8. Percentage of change in the indentation hardness (DH) of carbon nanotube-reinforced thermoplastic polymer
nanocomposites as a function of the CNT loading. Symbols as indicated in Fig. 6.
46 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

As an example, Fig. 9a illustrates the variation of hardness as a function of indentation depth for a
chitosan sample and a MWCNT-reinforced one [99]. A conspicuous increase of H below penetration
depths of 200 nm is clearly seen for both materials. Analogous EI and H enhancements have also
being observed in other CNT-reinforced chitosan composites [98], polyamide/CNT materials [20]
and in the 100/0 PI/SWCNT composite [126]. All this research makes use of the dynamic method in
the determination of E0 and H as a function of indentation depth [20,98,99,126]. A recent paper sug-
gests that such a method may produce an apparent reverse ISE at shallow depths [5]; a hint of such
effect is given in Ref. [126]. Depth dependences at small penetrations have also been reported for
epoxy/CNT systems although signicantly lower E or H enhancements were found [112,113]. Inden-
tation size effects at shallow depths are most commonly attributed to inaccuracies in the determina-
tion of the point of contact and of the tip area function.
In contrast to the above studies, Olek et al. [28] reported constant E values in the range 75500 nm
(in contact depth) for PMMA and CNT-reinforced PMMA composites. It should be noticed that the
authors used a soft polymer material for calibration of the tip area function. Indeed, this could be
at the heart of the E tendency to remain constant at such small penetration depths. In contrast, the

Fig. 9. (a) Variation of H with displacement into surface for neat chitosan (solid symbols) and chitosan/MWCNT (0.4 wt%)
nanocomposite (open symbols). Adapted from Ref. [99], copyright 2005, with permission from the American Chemical Society.
(b) Dynamic modulus E0 as a function of indentation depth for epoxy composites reinforced with nonfunctionalized (blue) and
functionalized (red) MWCNTs prepared in acetone as solvent. Adapted from Ref. [111], copyright 2010, with permission from
John Wiley & Sons, Inc.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 47

H values were found to diminish in the same depth interval. This was explained in terms of surface
roughness, time-dependent plastic deformation or other nanostructural features in the spin-coated
lms. However, it is not clear why these phenomena would inuence the depth dependence of H
and not that of E.
Most interesting are the studies on CNT-reinforced polymer composites reporting a variation of EI
and H as a function of indenter displacement that approaches the micrometer scale [108,111,126]. In
this range, the sources of error inuence the indentation data to a lesser extent and the ISE is most
likely associated to the genuine properties of the material. As an example, Fig. 9b shows the dynamic
modulus E0 of epoxy-based samples with pristine and functionalized MWCNTs as a function of the
indenter displacement. The continuous decrease of modulus values above h  200 nm has been dis-
cussed as a consequence of a morphology gradient across the sample thickness arising from solvent
molecules trapped in the reinforced matrix [111]. In other CNT-reinforced polymer composites,
because appropriate information on the experimental procedure and data analysis are not given, it
is difcult to evaluate whether the ISE is associated to a gradient of material properties [108,126].
For example, it seems that Dos Santos et al. have used a trapezoidal load function for the loading/
unloading cycles [108]. As indicated in Section 2.1.3, because the strain rate is continuously changing
as the test proceeds, this will normally result in an ISE. Other authors have found that EI and H values
remain constant in a wide range of indentation depths [57,114] and have interpreted this result as
indicative of the good level of CNT dispersion throughout the matrix [114].
Fig. 9a also shows that an H increase is found with increasing displacement for h > 500 nm [99].
This behaviour represents a clear example of reverse ISE, attributed in this case to substrate effects.
The lm properties are expected to be represented by the values shown at h  500 nm-1000 nm, that
is, in the small interval between the normal ISE and the reverse one. A detailed analysis of the reverse
ISE due to substrate effects is given by Mammeri et al. [127] on the basis of indentation studies on
CNT-PMMA based hybrid coatings. It was clearly shown that the ISE inuences the elastic modulus
values to a larger extent than the hardness ones because the elastic eld of deformation encompasses
that associated to plasticity and hence, the former would feel the substrate at an earlier stage (see
Section 2.1.3.1).

4.1.6. Creep
The incorporation of CNTs into polymer matrices is found to have a benecial effect on the creep
behaviour [24,109]. The time-dependent deformation of epoxy/SWCNT nanocomposites with different
ller content has been analysed in terms of the power law relationship between the strain rate and the
hardness through the creep exponent n (see Eq. (5)) [24]. Experimental data during the hold period at
peak load were employed. The double logarithmic plot of H and e_ was found to exhibit a linear trend in
almost all the range of strain rates (Fig. 10). However, two observations are worth commenting. The
rst one is that, in contrast to the ndings of Lucas and Oliver [42], no transient period at the begin-
ning of the holding time is observed. The second observation is that the power law t does not hold for
low H and e_ values. The n values are presumably obtained in the linear part of the log Hlog e_ plot.
Eventually, the work reports increasing values of n(reduced creep deformation) with increasing
amount of CNT incorporated into the epoxy matrix.
Tehrani et al. [109] analysed the creep behaviour of epoxy/MWCNT nanocomposites with 3 wt% of
ller using two different approaches. The rst one described creep experiments in terms of the creep
strain rate sensitivity parameter dened in Eq. (6). The second method monitored the time-dependent
deformation from the measurement of the creep compliance. A general tendency for the strain rate
sensitivity to decrease when CNTs are incorporated into the epoxy matrix was found. This was
explained by the authors in terms of several factors, the most important one possibly being CNTs act-
ing as blocking sites hindering the displacement of the amorphous epoxy chains subjected to the eld
of deformation. In addition, the work also analysed the increased strain rate sensitivity found at
increasing temperatures above RT for both, the neat polymer and the reinforced composites. This
result was discussed in terms of increasing free volume as the temperature rose towards the glass
transition. The authors concluded that strain rate sensitivity was a more useful parameter in the study
of creep deformation than creep compliance studies.
48 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

Fig. 10. Variation of indentation strain rate with hardness for epoxy composites with different SWCNT content. Reprinted from
Ref. [24], copyright 2004, with permission from Cambridge University Press. Values of the g parameter included in brackets
refer to the creep exponent n dened in Eq. (5).

4.2. Graphene

Graphene is one of the most exciting nanomaterials being currently investigated both for funda-
mental studies and potential applications. It consists in a at, atomically thick, two-dimensional sheet
composed of sp2 carbon atoms arranged in a honeycomb structure, and it is the basic building block
for graphitic materials of all other dimensionalities. Graphene sheets offer extraordinary electronic,
thermal and mechanical properties [137142], superior electron mobility (15,000 cm2/V s), excellent
thermal conductivity (5000 W/m K) [140], very large surface area (2630 m2/g) and the highest
electrical conductivity known at RT (6000 S/cm) [140]. Moreover, graphene is one of the strongest
materials on earth, the elastic modulus of a suspended monolayer measured by nanoindentation
was found to be 1 TPa [142]. In addition, an ultimate strength of 130 GPa and a breaking strength
200 times greater than that of steel were reported [143]. These unique properties make graphene
an excellent additive for reinforcing polymer matrices, and the resulting nanocomposites have the
potential to rival or even surpass the performance of their carbon nanotube-based counterparts pro-
vided that cheap, large-scale production and processing methods for graphene become available
[141,144147]. Graphene has already been used in a variety of applications such as sensors, batteries,
supercapacitors, hydrogen storage devices and so forth. The rst isolation of free-standing single-layer
graphene from micromechanical cleavage of highly oriented pyrolytic graphite was reported by Nov-
oselov et al. [138] in 2004. To date, several methods have been reported for the preparation of graph-
ene sheets such as epitaxial growth on SiC wafers [148], chemical vapour deposition of hydrocarbons
on metal surfaces [149], thermal exfoliation of graphite oxide [150], chemical reduction of graphite
oxide [151] or electrochemical intercalation [152]. Among them, the thermal and chemical exfoliation
of graphite oxide are effective ways to prepare graphene in relatively large quantities at a low cost.
Graphite oxide is a layered material produced by the treatment of graphite using strong mineral acids
and oxidizing agents. Each layer consists of covalently attached oxygen-containing groups such as
hydroxyl, epoxy and carboxylic acid. The reduction of highly oxidized graphene oxide (GO) sheets
from the exfoliated graphite oxide leads to reduced GO (rGO) or functionalized graphene sheets (FGS).
Three main approaches have been reported for the preparation of polymer/graphene nanocompos-
ites: melt-blending, solution intercalation and in situ polymerization of monomers [153]. In the rst
approach, the matrix is mechanically mixed with graphene at elevated temperature via extrusion or
injection moulding. However, because of thermal instability of most chemically modied graphene,
the use of melt blending has been limited to a few studies [141,154,155]. The solution technique
involves the dispersion of graphene in an appropriate solvent, the adsorption of the polymer onto
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 49

delaminated graphene sheets in solution and the elimination of the solvent, resulting in sandwich-like
nanocomposites [21,156160]. In the in situ polymerization strategy, graphene is rst swollen within
the liquid monomer, the initiator is then added and subsequently the polymerization is started either
by heat or radiation [161]. To attain stable dispersions of graphene and control the microstructure of
the resulting nanocomposites, polymer-covalent functionalization of graphene is highly desired [153].
The direct modication of graphene akes with polymer matrices by using the oxygenated groups (i.e.
hydroxyl, epoxy and carboxyl) in graphene oxide (GO) represents an interesting strategy. In particular,
thermoplastic polymers such as PVA [162], polyvinyl chloride (PVC) [163] or PMMA [161] and thermo-
sets like polyurethane (PU) [158,159] have been covalently bonded to the GO surface. Moreover, click
chemistry reactions have been recently successfully used to modify graphene for its incorporation in
polymer nanocomposites [164,165]. An alternative approach is the non-covalent modication, which
enables the attachment of molecules through p  p stacking or hydrophobic (van der Waals) interac-
tions, preserving the intrinsic electronic properties of graphene [154].

4.2.1. Modulus enhancements in graphene/polymer nanocomposites


Nanoindentation has been successfully applied to study the reinforcing effect of graphene-like ll-
ers in polymer matrices. Table 2 collects literature data published to date on this topic [21,154
160,166174]. Similarly to Tables 1 and 2 includes relevant information for the discussion, e.g. sample
morphology, state of aggregation of the ller or methodology employed for analysis of the DSI data.
Fig. 11 illustrates DEI values for different polymer matrices as a function of graphene content. It is
worth remarking that the term graphene encompasses a wide range of graphene-like structures
including graphite platelets (GP), graphene nanoplatelets (GNP) and different types of functionalized
graphene sheets such as GO, exfoliated graphite (EG), thermally expanded graphite (TEG) and rGO. In
addition, the graphene-like stacks can adopt different average sizes. Fig. 11 shows that, similarly to
CNT-reinforced materials, most of the data apply to low ller contents (<1 wt%) and there are only
a few examples at higher loadings. A gradual reinforcement with increasing ller concentration is
observed for different glassy and semicrystalline thermoplastic matrices such as PMMA [21], PP
[154] and PVA [160]. In the case of PVA, the trend surpasses the low concentration range approaching
graphene levels of 4 wt%. This behaviour contrasts with that of epoxy matrices where the modulus
hardly increases with the addition of graphene [166169], especially at high ller loadings
[166,167,169]. It is noteworthy that in all cases graphene sheets have been introduced in epoxy

2000

1000
Epoxy/GP [166]
Epoxy/GNP [168]
Epoxy/GNP [169]
Epoxy/GNP [170]
150 Epoxy/TEG [172]
E [%]

Chitosan/GS [156]
PP/GO [154]
PP/GO [154]
100 PVA/FG [21]
PVA/GNP [160]
PMMA/FG [21]
PMMA/EG [157]
50 PU/rGO [158]
PU/GO [159]

0 1 2 3 4 5 6
Graphene [wt%]

Fig. 11. Percentage of increase in the elastic modulus (DEI) for different graphene-reinforced polymer nanocomposites as a
function of the nanoller content. Symbols in red indicate data that should be taken with caution (see Table 2 for specic
details), and those in blue correspond to thermoset matrices. Solid and open symbols designate raw and functionalized
nanollers, respectively; M and s correspond to GO and OTES-modied GO, respectively; indicates GO covalently bonded to
the polymer matrix and + denotes aligned rGO anchored to the matrix.
50 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

matrices without functionalization. This route seems to give rise to micrometer or sub-micrometer
blocks that are homogenously distributed in the matrix [166,169]. Hence, the limited EI (and H, see
Table 2) increments detected in case of epoxy matrices must be related to the aggregation of graphene
into stacks and the poor matrix-ller interaction.
Fig. 11 clearly shows that the largest DEI (1800%) has been attained for amorphous PU composites
incorporating 5 wt% of well aligned, ultralarge-size rGO layers being the direction of the load perpen-
dicular to the graphene sheets [158]. Although this DE value was calculated relative to E = 0.4 GPa of
neat PU and hence can be affected by some inaccuracies, it is reasonable to suggest that the orientation
and size of the graphene sheets would play an important role. These rGO sheets exhibit mean lateral
size of more than 10 lm that minimizes the occurrence of matrix-ller interfaces. Most important,
graphene sheets are self-aligned in the polymer matrix with their basal plane perpendicular to the
lm thickness. Indentation studies on suspended graphene or reduced graphene oxide monolayers
have yielded modulus values of 1 TPa [142] and 0.25 TPa [175] respectively, while only 36.5 GPa
has been found for the modulus of exfoliation of graphene sheets [169]. Hence, orientation should
be a major contribution to the outstanding reinforcement of PU/graphene composites.
Concerning thermoplastic matrices, the largest EI improvements are found for PMMA-based nano-
composites approaching a modulus increase of 70% with the incorporation of only 0.6 wt% amide-
functionalized few-layer graphene (FG) [21]. This functionalization route seems to be most effective
in providing a good load transfer across the polymermatrix interface. The same authors reported
lower increments (DE = 35%) but still signicant when incorporating the same amount of FG function-
alized with COOH and OH groups to PVA [21]. Fig. 12a shows the plot of the DSI load-depth data for
this series of PVA/FG nanocomposites incorporating different percentages of ller. It can be clearly
observed that the stiffness (slope at the onset of unloading) rises with increasing graphene loading.
The modulus increments were partially attributed to an increase in the degree of crystallinity of the
matrix, as illustrated in Fig. 12b. Crystalline phases are suggested to nucleate at the interphase
between the ller and the matrix resulting in a strong polymermatrix interaction. Other authors have
reported similar modulus enhancements for PVA with the incorporation of partially oxidized graphene
(see Fig. 11 and Table 2) [160]. Interestingly, the linear increment in modulus found with increasing
graphene loading was persistent up to the high loading of 4 wt% [160]. Enhanced crystallinity values
were suggested to contribute to the modulus improvement for low ller loadings, in agreement with
the work of Das et al. [21]. However, for ller contents above 1 wt%, the crystallinity decreased due to
physical hindrance of graphene to the crystal growth, and the modulus increments were attributed
solely to the reinforcing effect of the nanoller, its homogenous dispersion and the strong interfacial
adhesion between the two nanocomposite phases. Comparable increments to those of PVA have been
reported for semicrystalline PP based nanocomposites loaded with 1.0 wt% GO [154]. Surface modi-
cation of GO with a coupling agent (octyltriethoxysilane, OTES) was found to further enhance E of PP
nanocomposites (see Fig. 11) due to increased reactive sites on the GO surface that enhanced the inter-
facial interaction with the matrix and hence promoted mechanical interlocking [154].
Planar orientation of the graphene sheets is also expected to play a fundamental role in achieving
enhanced mechanical properties in lms of graphene/polymer bilayer and multilayer assemblies. In
fact, a recent study on ultrathin PVA/GO multilayer lms fabricated by the LbL technique has attrib-
uted the signicant enhancement found for Er and H, 99% and 240% respectively, to the high degree of
planar orientation of the graphene-like nanosheets [173]. The LbL technique has also been used to
sequentially assemble a number of layers to produce polyelectrolyte/GO membranes. Incorporation
of GO was found to be benecial for the indentation mechanical properties [174]. The deposition of
a top graphene layer onto a exible polymer substrate is possibly the simplest layer assembly in which
graphene has been found to enhance the indentation mechanical properties of the lm [157]. A mod-
erate DEI of 16% was reported while a 97% H improvement was attained [157]. The authors discussed
the limited increment found for EI as a consequence of the far-eld nature of the elastic stress eld
compared to the plastic deformation zone.

4.2.2. Comparison with macroscopic mechanical properties


In spite of the a priori difculties in comparing nanoindentation and mechanical data from
conventional macroscopic techniques due to technical dissimilarities, different scale of the volume
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 51

Fig. 12. (a) Representative loaddisplacement curves for PVA with different contents of acid functionalized FG. (b) Change in
crsytallinity as a function of FG content. Adapted from Ref. [21], copyright 2009, with permission from IOP publishing.

of deformation, diverse load directionality and distribution of stresses, amongst other important fac-
tors discussed in detailed in preceding sections, Table 2 shows an outstanding correspondence
between the increments in modulus derived from indentation and macroscopic techniques. Compar-
ison of the Y and EI values [158,169] shows a reasonable agreement, the latter values being slightly
higher. Most interesting is the excellent agreement found in the only example where modulus values
from dynamic indentation and DMA are compared, an epoxy matrix reinforced with GP [166,167]. In
this work, special care was paid to carry out a comprehensive comparison between nano and macro
techniques. In the rst place, three-point bending DMA and indentation measurements were com-
pared at the same applied frequency. In addition, it was demonstrated that, for 1 lm-size GP, the vol-
ume of deformation in DSI encompassed a platelet distribution that was representative of that of the
bulk material. In fact, increasing the size of the graphite platelets to 15 lm produced divergent results
between DSI and DMA because the local nanoindentation measurements were found to be no more
representative of the macroscopic properties.
52 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

4.2.3. Indentation size effect


Because the dimensions of the graphene-like platelets typically approach a few micrometers in lat-
eral size, the depth dependence of modulus and hardness in graphene/polymer composites is espe-
cially sensitive to the distribution of platelets throughout the sample. Dynamic DSI studies on
epoxy-based composites with 5.8 wt% GP (1 lm of mean lateral size) reported different E0 proles
depending on the location of the initial contact that nally converged to the same modulus value at
large penetration depths (h  1500 nm) [166]. In contrast, similar experiments using 15 lm size GP
yielded E0 proles vs. indentation depth that exhibited a number of undulations, which were associ-
ated with the occurrence of new platelets in the immediacies of the indenter tip [166]. In this case,
the depth dependence did not exhibit a nal plateau because the volume of deformation is not a good
representation of the composite composition at any stage of the indentation test. In contrast to these
studies, King et al. [169] reported constant E0 and H values with penetration depth above h = 200 nm
for epoxy matrices reinforced with 5 wt% of graphene nanoplatelets with average particle diameter of
15 lm. Results seem to be at variance with the E0 depth dependence discussed above for similar epoxy
reinforced matrices [166]. The apparent disagreement could be due to the size of the graphite blocks
employed. For shallow penetration depths (h < 200 nm), a conspicuous increase of E0 and H with
decreasing depth has been observed in a number of studies, possibly associated to an incorrect tip cal-
ibration or uncertainties in the determination of the point of initial contact [169,174].

4.2.4. Creep
The incorporation of graphene-like nanollers is generally found to enhance the creep resistance of
polymer matrices [21,157,169,173,176]. This is immediately detected in a standard load-hold-unload
cycle. During the hold period, depth increases at a lower rate when the nanoller is present
[21,157,173]; an illustration of such effect can be seen in the load-depth curves of Fig. 12 on PVA/
FG composites. A thorough analysis of the creep behaviour has only been recently offered
[145,169,176].
Chen et al. [157] analysed in detail the creep behaviour of a bilayer lm of graphene on top of PET.
The creep strain, dened as the indenter displacement at any point during the hold period divided by
the displacement at the beginning of the hold period, was found to be substantially smaller for the
PET/graphene nanocomposite compared to that of the neat polymer. Analysis of the creep compliance
also revealed that the graphene overlayer signicantly improved the creep resistance of PET. In the
case of epoxy matrices, the incorporation of 0.1 wt% GNP signicantly decreased the creep displace-
ment at peak load [176]. It was found that the creep displacement vs. creep time curves obtained with
a cylindrical punch are well described by a power law, the exponent values being smaller for the nano-
composite compared to the neat polymer. In contrast, King et al. [169] only found a slightly lower
creep compliance for GNP/epoxy composites than that of neat epoxy for relatively low loadings
(2 mN) while the same creep compliance was found for both lled and pristine material for applied
loads in the range 1045 mN.

4.3. Other organic nanollers

In addition to CNTs and graphene, other carbon nanomaterials such as carbon nanobres (CNFs),
nanodiamond (ND), fullerenes, carbon black (CB) and so forth have been successfully used for the
development of polymer nanocomposites. CNFs are intimately related to CNTs regarding their struc-
ture and properties. They exhibit a tubular hollow conguration, with length in the range of 50 to
1000 lm and outer diameter of 40150 nm, smaller than that of conventional CFs (510 lm) but con-
siderably larger than that of CNTs (140 nm). CNFs possess very high thermal conductivity (1900 W/
m K [177]), electrical conductivity close to that of graphite (2  104 S/cm [178]) and superior Youngs
modulus (in the range of 300600 GPa [179]) and tensile strength (8.7 GPa [180]). They are generally
synthesized by CVD process [181] in which the bre grows around particles of a transition metal cat-
alyst. This procedure involves several stages including gas decomposition, carbon deposition, bre
growth, bre thickening, graphitization, and purication. CNFs produced by this method are known
as vapour-grown carbon nanobres (VGCNFs).
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 53

Among diamond-based nanomaterials, detonation nanodiamond (ND), also named ultradispersed


diamond or ultrananocrystalline diamond, is most attractive for nanotechnological applications. Det-
onation nanodiamond can be easily suspended in water, has moderate cost and is non-cytotoxic and
biocompatible, hence, highly suitable for the biomedical eld. It is produced by detonation of carbo-
naceous explosives in a closed chamber either in a gaseous atmosphere, e.g. CO2 (dry method), or in
water (wet method) [182], and can also be synthesized from a suspension of graphite in organic liquid
at atmospheric pressure and RT using ultrasonic cavitation [183]. ND is composed of crystalline nano-
particles with typical diameter of 36 nm, specic surface area in the range of 300500 m2/g, thermal
conductivity of 1200 W/m K [184], hardness of 300 GPa [185] and modulus close to 500 GPa [186].
When mixed with a polymer matrix, non-functionalized diamond nanoparticles exhibit poor disper-
sion and tend to aggregate resulting in irrelevant property improvement. Therefore, a major issue
in manufacturing ND-reinforced composites is to improve ller dispersion. To achieve this goal, var-
ious approaches similar to those described previously for CNTs have been utilized, including high
shear mixing, extrusion, in situ polymerization, wrapping the nanoparticles with surfactants, and
covalent chemical functionalization [187].
Fullerenes are hollow molecules in the form of spheres or ellipsoids composed entirely of sp2 car-
bon atoms. They are similar in structure to graphite, but they also contain pentagonal and/or heptag-
onal rings [188]. The simplest and most symmetric fullerene (named buckminsterfullerene C60), is
composed of 20 hexagons and 12 pentagons, and was discovered in 1985 by Curl, Kroto and Smalley
[189]. Unlike other allotropes of carbon, fullerenes are superconductors and exhibit low thermal con-
ductivity (0.4 W/m K [188]). This material is harder than graphite but softer than diamond, showing
a hardness of around 170 GPa [185].
Carbon black (CB) is an amorphous form of carbon produced by the incomplete combustion of
heavy petroleum products. It is composed of roughly spherical particles with diameters between 10
and 400 nm that are fused together into aggregates [190]. CB is extensively employed for polymer
reinforcement due to its high electrical (102 S/cm1 [191]) and thermal conductivity (300700 W/
m K [192]), abundant availability, low density and low cost. Compared to the above mentioned nanof-
illers, CB exhibits modest mechanical properties, modulus values approaching 15 GPa [193].
Recently, nanollers based on natural polysaccharides like cellulose, starch or chitin have been
used for polymer reinforcement. Cellulose nanocrystals (CNC) can be obtained from native cellulose
bres by an acid hydrolysis [194], giving rise to highly crystalline (63%) and rigid nanoparticles that
are approximately 100300 nm long and 315 nm wide, referred to as nanowhiskers. CNC extracted
from wood pulp display has good mechanical properties for use in material applications: elastic mod-
ulus in the range of 100150 GPa and tensile strength of 7.5 GPa [195], comparable with those of
Kevlar and superior to those of GF. The most important restraint to the use of CNC as reinforcement
in nanocomposites is the poor compatibility between the hydrophilic polysaccharides and the typi-
cally hydrophobic polymer matrices. In order to prevent nanocrystals from aggregation and to
improve their adhesion with the matrix, surface modication is required via graft polymerization, sily-
lation of the OH group or the use of surfactants [196].
Table 3 presents DSI data for polymer nanocomposites reinforced with different types of organic
nanomaterials [115,121,187,193,197208], together with other relevant information. The increments
in modulus DEI are plotted as a function of the nanoller content in Fig. 13.

4.3.1. Modulus enhancements in organic/polymer nanocomposites


Fig. 13 shows that limited DEI values are reported for epoxy matrices upon addition of a number of
llers including carbon nanobres, either vapour grown or coated, nanodiamond, fullerene and carbon
black. Remarkable increments are only found for very high ller loadings (50 wt%) as commented
below. Analogous moderate improvements, or even decrements, are observed for the hardness values
(see Table 3). This behaviour resembles that observed for CNT and graphene-reinforced epoxies (see
Sections 4.1 and 4.2) and similarly to these latter studies, poor nanoller dispersion and weak l-
ler-matrix bonding are envisaged to be the main factors limiting the mechanical enhancement of car-
bon llers/epoxy nanocomposites [197,198]. Interestingly, Snchez et al. [197] suggested that the
presence of CNFs hinders the curing reaction of the epoxy matrix giving rise to reduced crosslinking
density which results in lower stiffness and hardness. This effect explains the high nanoller content
54 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

500
Epoxy/CNF [197]
Epoxy/CNF [198]
400 Epoxy/ND [199]
Epoxy/ND [199]
Epoxy/fuller. [115]
Epoxy/CB [193]
300 PLLA/ND [187]
PLLA/ND [187]
E [%]

PA-11/ND [205]
200 PA-11/ND [204]
PVA/ND [202]
PVA/CNC [201]
PVA/CNC [200]
100 PVAc/CNC [203]
PS/CNF [198]
PS/CB [206]
0 PAN/ND [208]
PMMA/CNC [207]

0 10 20 30 40 50
nanofiller [wt%]

Fig. 13. Percentage of change in the elastic modulus (DEI) for different polymer nanocomposites reinforced with carbon
nanobres (CNF), nanodiamond (ND), carbon black (CB), fullerenes (fuller.) or cellulose nanocrystals (CNC), as a function of the
nanoller content. The open symbols in Refs. [199,202,204,205,208] correspond to ox-ND. indicates nanoller grafted to the
polymer matrix, + and j denote aligned nanoller. Red symbols refer to data possibly affected by inaccuracies (see Table 3).

(5 wt%) that is necessary to achieve an apparent improvement of mechanical properties. The authors
reported AFM studies that provided evidence of limited pile-up for indentations on the neat resin and
the CNF/epoxy composites with low nanoller content while signicant pile-up was detected for
higher ller loadings of 3 wt% presumably due to an incomplete curing process (see Fig. 14). The
pile-up effect was reverted for higher ller loadings of 5 wt% due to enhanced stiffness as a conse-
quence of the incorporation of CNFs.
Most interesting is the work of Neitze et al. that reports a route to covalently bond nanodiamond to
the epoxy network [199,209]. Amino groups were incorporated into the surface of nanodiamond par-
ticles and participated actively in the curing chemistry of the epoxy resin [199,209]. Modulus and
hardness values of an aminated ND reinforced epoxy were found to be signicantly higher (50%
and 200% respectively) than those of a carboxylated ND/epoxy material with the same ller content
[209]. It was found that amino groups of the nanodiamond particles compete with those of the curing
agent in such a way that they can compromise the mechanical properties of the composite [199].
Table 3 collects the modulus increments for some of these aminated ND/epoxy nanocomposites. Out-
standing DE values are found at high loadings achieving 400% upon incorporation of 50 wt% of ami-
nated nanodiamond (see Fig. 13). Such remarkable increment was explained as due to the formation of
a ND network crosslinked by epoxy. Other amine-ND reinforced polymers have been prepared using
octadecylamine (ODA) covalently attached to ND [187]. A signicant increase of mechanical properties
in amine-ND/PLLA composites was achieved at ller concentrations up to 10 wt% (see Fig. 13). In con-
trast, addition of the as-received ND nanoller terminated with oxygen containing functional groups
produced irrelevant changes in modulus and reduced increments in hardness (see Table 3). The good
afnity between ND-ODA and the PLLA matrix together with the hydrophobicity of the ller provided
the framework for enhanced ller dispersion. TEM studies showed the occurrence of small agglomer-
ates, tens of nanometers in size, at all ller concentrations that turned out to interconnect at high
loadings. Uniform ller distribution, enhanced ller-matrix interaction and increased crystallinity val-
ues account for the remarkable enhancement of mechanical properties.
The addition of ND to other polymer matrices has proved most successful in enhancing the
mechanical performance when signicant functional groups at the surface of ND are promoted
[121,202,204,208]. Indeed, the increased surface reactivity of oxidized ND (ox-ND) with respect to
decarboxylated ND (deox-ND) has been shown to be critical for the improvement of the elastic mod-
ulus values of polyacrylonitrile (PAN) [208]. Ox-ND has also been shown to produce marked modulus
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 55

Fig. 14. AFM image of indentations produced on: (a) neat epoxy and (b) epoxy/CNF (3.0 wt%). (c) Line proles for neat epoxy
and the nanocomposites. (d) Comparison of the elastic modulus of epoxy and CNF/epoxy nanocomposites measured by DMA
(black symbols) and nanoindentation tests at two different maximum loads. Adapted from Ref. [197], copyright 2011, with
permission from Elsevier.

and hardness increments when incorporated to PVA at low ller loadings (80100% increase in E and
H for only 0.6% ND, see Table 3 and Fig. 13) [121,202]. It was demonstrated that part of this improve-
ment is due to the nucleating effect of ND producing signicant matrix crystallinity enhancements and
favouring a strong matrix-ller interaction [121,202]. Most signicant is the modulus increase of PA-
11 (300%) with addition of 20 % oxidized ND, although the authors do not describe the procedure
employed for the DSI data analysis and hence, results should be taken with caution [204]. Neverthe-
less, this work proposes an interesting approach to achieve uniform dispersion of ND that consists in
the use of electrospinning as a vehicle to deposit nanodiamond with minimal agglomeration followed
by heating the electrospun nanobres to produce ND reinforced thermoplastic coatings.
Concerning CNC-reinforced polymers, the study of the nanomechanical properties reported so far
reveals moderate improvements [200,203,207]. Composites in the form of bres or lms, as well as
glassy and semicrystalline polymer matrices have been investigated. Loadings as high as 41 wt% have
been added to electrospun PMMA bres, and yet they produce mechanical enhancements of no more
than 30% as revealed by nanoDMA [207]. The authors argued that these limited increments could
arise from the anisotropy of the mechanical properties. CNC were aligned along the PMMA bre axis,
while indentations were produced in the perpendicular direction and hence, more signicant
increases should be expected along the bre direction. In CNC-thermoplastic lms, nanoller agglom-
eration is found to be the main limitation for mechanical improvement. This is indeed behind the AFM
nanoindentation studies on PVA/CNC and PVA/PAA/CNC (CNC wt% 620) [200] and polyvinyl acetate
(PVAc)/CNC (CNC wt% 63) [203]. As an example, Fig. 15a shows the inhomogeneous distribution of
56 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

Fig. 15. (a) Topographic images of the PVA/CNC (15 wt%) nanocomposite showing the heterogeneous CNT dispersion. (b)
Comparison of the elastic modulus obtained by tensile testing with those derived from the application of Hertzs model to the
loading and unloading indentation curves on (A) neat PVA; (B) PVA 90 wt%/PAA 10 wt%; (C) PVA 85 wt%/CNC 15 wt%; (D) PVA
80 wt%/PAA 10 wt%/CNC 10 wt%; (E) PVA 70 wt%/PAA 10 wt%/CNC 20 wt%. Adapted from Ref. [200], copyright 2012, with
permission from IOP publishing.

CNC in a PVA/CNC 85/15 composite [200]. Interestingly, indentations produced at different locations,
on a CNC agglomerate and on the matrix, revealed only small differences in the load-depth curves in
terms of maximum penetration depth and slope because the polymer underneath CNC contributes to
the recorded deformation behaviour. This work also highlights an important consideration for the
analysis of load-depth curves: E values derived from the application of Hertzs model to the loading
and unloading seem to be at variance (see Fig. 15b) due to the deviation of the unloading behaviour
from the condition of full elasticity.
Finally and most surprising is the remarkable modulus enhancement shown in Fig. 13 (up to 350%
for 5 wt%) for a PS-based shape memory nanocomposites incorporating 15.5 wt% of carbon black par-
ticles [206]. CB particles were incorporated as received in the crosslinked system, and nevertheless the
reinforcement seems to be most efcient. This result contrasts with the modest increments attained
for other thermoset materials incorporating similar amounts of CB (see Fig. 13). At the heart of this
inconsistency could be the fact that low storage modulus values have been determined by means of
DMA for all the PS-based shape memory nanocomposites (see Table 3). Indentation testing on compli-
ant materials is known to display additional concerns in the detection of the initial contact that can
introduce some inaccuracy in the modulus data (see Section 3).
Recently, novel PVA-based hybrid nanocomposites incorporating two acid-functionalized carbon
nanollers with different dimensionalities have been successfully developed [121]. Results demon-
strated extraordinary mechanical improvements due to synergistic effects (Table 3). In particular,
DE values in the range 10001300% were found upon addition of binary combinations of SWCNTs/
FG or SWCNTs/ND in small amounts (0.2 or 0.4 wt% each). Surprisingly, almost no change in the
matrix crystallinity was detected, suggesting that an increase in this parameter did not contribute
to the observed synergy. The reader should be here aware that, as repeatedly indicated in this review,
the low E values reported for the PVA matrix (E = 0.7 GPa) are prone to some inaccuracy.

4.3.2. Comparison with macroscopic mechanical properties


The interest in correlating indentation modulus values with those determined from macroscopic
techniques is manifested in the signicant number of papers that approach this issue
[197,198,200,206,208].
The reinforcing effect of the ller is usually detected in a similar manner by DSI and conventional
macroscopic techniques [197,206,208] although some discrepancies have also been reported
[200,208]. For PS/CNF composites, Y values have been successfully used as a reference to correct for
pile-up effects in the determination of the contact area necessary for DSI analysis [198]. However, it
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 57

is a most common nding that considerably higher modulus values are achieved by means of inden-
tation testing [197,200,206,208]. Snchez et al. [197] attributed these differences to indentation size
effects and the occurrence of local higher loadings. Pakzad et al. [200] argued in terms of less available
free volume in DSI, occurrence of defects and microscopic cracks in macroscopic testing, differences
between compression (identied with DSI) and tensile testing and inuence of dissimilar strain rates
in nano and macro techniques.

4.3.3. Indentation size effect


Few studies have approached the depth-dependence of EI and H and only some examples on poly-
mer nanocomposites containing CNF, CB and CNC are available [193,198,207]. A conspicuous ISE
below h  300 nm and a minor reverse ISE above this penetration depth have been reported for PS/
CNF and epoxy/CNF composites [198]. The marked E0 and H rise at small displacements was similar
to that reported in the preceding sections for other carbon-based reinforced polymers and can be asso-
ciated to an underestimation of the contact area and other sources of error at shallow penetration
depths. In carbon black/epoxy composites, a gradual rise of modulus values at h > 200 nm has been
associated to the inuence of the stiffer silicon substrate [193]. NanoDMA storage modulus data have
been reported for PMMA and CNC/PMMA electrospun bres of 500 nm average diameter over a nar-
row range of indentation depths covering 2030% of the relative depth (indenter depth/bre diameter)
[207]. The constant E0 values encountered within this range of relative depths was interpreted by the
authors as a means of reassuring that issues arising from probe shape and sample depth were not
inuencing the genuine values of the mechanical properties.

4.3.4. Creep
The creep behaviour of CNF, ND, CB and CNC polymer nanocomposites by means of indentation
testing has been little explored so far, and yet a quick look to the data during the hold segment at max-
imum load can provide worthy information on the creep properties. For example, the introduction of
CB and amine functionalized-ND in PS and PLLA respectively has been shown to signicantly reduce
creep [187,206]. Some analysis of the creep behaviour is only offered by Kaboorani et al. [203] using
the relative change of the indentation depth at the peak load for PVAc composites incorporating CNC.
A progressive reduction of creep was found upon addition of the nanoller reaching prominent values
of 55% decrease with only 3% of CNC.

4.4. Main features on organic llers

4.4.1. Carbon nanotubes


A wealth of data on CNT-reinforced polymers is available although with signicant data scatter.
This is possibly due to the different experimental procedures and criteria adopted for the load-depth
analysis. However, and overall, some conclusions can be drawn that teaches us important aspects on
the reinforcement mechanisms.
Reinforcement of thermoset matrices with CNTs is quite a difcult task. Moderate increments of
modulus or even some decrements are usually found. A poor dispersion of CNTs is possibly the main
factor limiting the effective mechanical enhancement. Functionalization approaches do not seem to be
successful; a few examples reveal that similar results are obtained for functionalized and raw CNTs.
Perhaps the most successful route is the use of plasma functionalized MWCNTs grafted to the matrix,
although more research is needed to extrapolate this behaviour to higher loadings.
In case of thermoplastic matrices, a fair dispersion of the ller, either raw or functionalized, is usu-
ally achieved at low CNT loadings (<2%). Hence, this factor does not seem to be critical in obtaining
moderate EI and H increments that rise with increasing ller content. An optimum CNT-matrix inter-
action seems to play here a relevant role. It has also been shown that the changes in crystallinity trig-
gered by the CNT on the polymer matrix can signicantly contribute to the EI and H enhancement.
Results for high CNT loadings (>2%) suggest that in addition to the ller-matrix interface adhesion, l-
ler dispersion becomes now an important factor. In this respect, acid-functionalization of CNTs seems
to be the most efcient approach.
58 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

Interestingly, both in epoxy and thermoplastic matrices, some authors have explained the poor
mechanical enhancement found in certain composites in the light of the specic properties of CNTs:
waviness, curved morphology, inferior bending properties or sliding ability.
Finally, creep studies on CNT-reinforced polymers are quite limited but results published suggest
that the incorporation of CNTs into polymer matrices can diminish the time-dependent deformation.

4.4.2. Graphene
Nanoindentation studies on epoxy/graphene composites reveal irrelevant mechanical improve-
ments. It is noteworthy that in this case the ller has always been introduced without functionaliza-
tion. In contrast, the incorporation of graphene-like llers in thermoplastic matrices has been proved
to be most successful always following a functionalization route. Occasionally, in semicrystalline
matrices, the modulus increments have been partially attributed to the nucleating effect of graphene
that, on the one hand, increases the degree of crystallinity of the matrix and on the other, nucleates
crystalline phases at the interphase between the ller and the matrix, giving rise to a strong poly-
mermatrix interaction.
Overall, the factor that seems to produce the most signicant enhancement of mechanical proper-
ties is the orientation of the graphene sheets. Incorporation of well-oriented micrometer-sized graph-
ene sheets into a polymer matrix produces a remarkable increment of indentation modulus in the
direction perpendicular to the plane of the graphene.
For this type of composites, indentation size effects can be expected because ller size is of the
same order as indentation displacement. It has been shown that the modulus of epoxy/graphene com-
posites at large indentation depths is only characteristic of that of the bulk material when the volume
of deformation encompasses a representative number of platelets. In this latter case, three point bend-
ing DMA studies and dynamic nanoindentation measurements, both carried out at the same fre-
quency, were found to be in excellent agreement.
Finally, the incorporation of graphene-like nanollers has been shown to enhance the creep resis-
tance of thermoplastic matrices and to a lesser extent of epoxy resins. Moreover, a graphene overlayer
has been also found to reduce the creep strain of a exible PET lm. Detailed creep analysis has only
been offered recently and it is expected that the eld progresses in the next years.

4.4.3. Other organic llers


Similarly to CNT and graphenereinforced epoxies, the incorporation of other carbon-based nanof-
illers (carbon nanobres, nanodiamond, fullerenes and carbon black) to epoxy polymers produces low
or moderate modulus and hardness increments, probably due to the poor ller dispersion and weak
matrixller interaction. In addition, it has been suggested that the nanoller can hinder the curing
reaction of the epoxy giving rise to a matrix with reduced stiffness and hardness. Unprecedented
mechanical increments have been found in one example of aminated-ND/epoxy composite where
the ller was covalently attached to the epoxy matrix (DE0  400% for a loading of 50 wt%).
Nanodiamond seems to be the most effective reinforcing ller for thermoplastic matrices, amine
functionalization being the most successful route for optimum mechanical performance. Improved l-
ler dispersion, enhanced llermatrix interaction and enhanced levels of crystallinity in case of semi-
crystalline matrices account for the remarkable increase of mechanical properties.
Analysis of the indentation data at peak load reveals that the incorporation of CB, ND and CNC to
thermoplastic matrices signicantly enhances the creep resistance.
Recent indentation results on PVA reinforced with binary combinations of carbon nanollers (CNT,
graphene and ND) point towards signicant synergistic effects when two of these nanollers are
employed. Further research in this direction should be of great value.

5. Polymer nanocomposites incorporating inorganic nanollers

Inorganicorganic nanocomposites generally refer to polymer composites composed by nanoscale


inorganic building blocks and a polymer matrix. These building blocks include: layered silicates (e.g.,
montmorillonite, hectorite, saponite), metal nanoparticles (e.g., Au, Ag), oxides (e.g., SiO2, TiO2, Al2O3),
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 59

semiconductors (e.g., PbS, CdS) and so forth. They aim to combine the characteristic properties of poly-
mers with those of an inorganic material. The small size of the nanollers yields a very large interfacial
area, which may give rise to a signicant relative amount of interphases with properties different from
the bulk. This fact opens the possibility to have synergistic phenomena that produce effects greater
than the sum of the individual components. On the contrary, the very high surface energy of the
nanollers favours the agglomeration into larger particles. The consequence is poor nanoparticle dis-
persion within the nanocomposite that usually leads to degradation of the properties and, for this rea-
son, the preparation process becomes critical [210,211].
As previously done in Section 4, dedicated to organic based nanollers, this part of the review col-
lects most of the published data obtained by DSI techniques on polymer nanocomposites with inor-
ganic nanollers. The main aim is, likewise, to understand the inuence of the different parameters
that characterize the indentation response of these materials and to compare the results with those
obtained by macroscale conventional techniques. The discussion on different types of inorganic nanof-
illers will be subdivided in three sections: layered silicates or nanoclays (5.1), spherical nanoparticles
(5.2) and other inorganic nanollers (5.3).

5.1. Layered silicates

5.1.1. Characteristics, preparation methods and types of nanocomposites


Polymer/layered silicate (PLS) nanocomposites have attracted great scientic and industrial inter-
est due to the extraordinary improvements in properties attained at very low ller contents. Enhanced
strength, modulus and toughness, increased tear, radiation and re resistances and reduced thermal
expansion and permeability to gases have been reported in these nanocomposites making them ideal
materials for applications in food packaging, structural automotive components and electronics
among others [212217].
Montmorillonite (MMT) and hectorite are the layered silicates most commonly used for the
preparation of the nanocomposites. These clays exhibit a very high aspect ratio with layer thick-
nesses of 1 nm and lateral dimensions that may vary from a few hundred nanometers to microns.
Their structure include charge compensating counter-ions such as Na+ located in the inter-layer
space which can be replaced by organic ammonium and phosphonium cations with long alkyl
chains leading to organically modied clays having better compatibility with polymers. In fact,
the ability of layered silicates to tune their surface chemistry through ion exchange reactions with
organic and inorganic cations, together with the possibility to disperse them into individual layers,
are the two main characteristics that determine their adequate incorporation in polymer nanocom-
posites [213,214].
Depending on the strength of the interfacial interactions between the polymer matrix and the
clay, two main types of PLS nanocomposites can be obtained: intercalated and exfoliated struc-
tures. The rst type is formed when polymer chains are inserted between the layers of the clay
and a separation of several nanometers between the platelets is observed, while the second one
is obtained when the layers of the clay are completely separated and dispersed throughout the
polymer matrix. The latter structure usually leads to better mechanical properties than the inter-
calated one because it maximizes the polymerclay interactions facilitating the stress transfer.
However, complete exfoliation of the clays is difcult to achieve and most of the polymer nano-
composites reported have mixed intercalated and exfoliated structures [212214]. Moreover, for
highly non-polar polymers such as polyolens, intercalation is thermodynamically unfavourable
and it becomes necessary to incorporate a compatibilizer, usually a functionalized polymer or
block copolymer, in order to favour interactions with the clay and miscibility with the matrix
[218222].
The ability to promote organoclay exfoliation is strongly inuenced by the preparation of the PLS
nanocomposites. Four main preparation strategies have been developed: in situ template synthesis
(solgel technology), in situ intercalative polymerization, solution intercalation and melt intercalation.
The last method presents environmental advantages and is most commonly used in the industry. A
detailed description of the methods for producing PLS nanocomposites can be found in a number of
reviews [212214].
60 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

5.1.2. Modulus and hardness changes in polymer/layered silicate nanocomposites


Regarding the mechanical performance of PLS nanocomposites, signicant improvements in
Youngs modulus and tensile strength have been observed for very low clay concentrations (25%)
[212215]. In addition to the preparation method, the nature of the matrix and the content and degree
of dispersion of the ller, other factors such as surface treatment of the clay, orientation of the dis-
persed platelets or the inuence of the nanoller on the development of crystallinity in thermoplastic
based PLS nanocomposites need to be considered when analysing the mechanical properties of this
type of materials.
DSI results on PLS nanocomposites are collected in Table 4 and in Fig. 16 [218,219,223252].
Youngs modulus data determined by tensile testing and DMA storage modulus values are also
included in the table. Similarly to preceding tables, other characteristics such as nanoller type and
content, composite processing method and morphological features related to the ller dispersion
and ller-matrix interaction have also been included for the sake of the discussion.

5.1.2.1. Polymer matrix inuence. Fig. 16 shows that the largest DEI values have been reported for nano-
composites with semicrystalline polymer matrices, mostly polyamides and polyesters. A similar trend
has been observed for hardness data (Table 4). Enhancements of 130% in E and 160% in H have been
reported for a PA-6 nanocomposite containing 10 wt% of an organically modied clay [223], while val-
ues of DE = 110% and DH = 75% were found with addition of 6 wt% of a surface modied MMT [226],
both nanocomposites prepared by melt blending. In addition, there are many examples of polyamide
nanocomposites (PA-6, PA-6,6, PA-11 and PA-12) with signicant DEI and DH at low ller contents
[224,226234] regardless the method of preparation (see Table 4 and Fig. 16). In the case of polyesters
such as poly (3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV), the addition of 5 wt% organoclay by a
solution intercalation method resulted in E0 and H enhancements of 100% and 90%, respectively [237].
Outstanding DE and DH values have been reported for poly(ethylene oxide) (PEO)/clay nanocompos-
ites with very high loading (270% and 250% respectively at 50 wt% clay content), however at low con-
centrations the improvements are not signicant [240]. One of the reasons for the large improvements
described in both mechanical properties for polyamide and polyester PLS nanocomposites is the
strong interaction between the matrix and the silicate layers that in some cases has been attributed

Epoxy/Cl 93A [247]


Epoxy/A1100 [248]
140 Epoxy/A1120 [248]
PA-6/I30TC [223]
PA-6/CoAl-LHD [224]
120 PA-6/1.34TCN [226]
PA-66/1.34TCN [226]
PA-66/1.34TCN [227]
100 PA-66/1.34TCN [227]
PA-66/Hectorite [230]
PA-11/1.34TCN [232]
80 PA-12/1.34TCN [233]
E [%]

PA-12/layered silicate [234]


PLA/Cl 30B [235]
60 PHBV/Cl 30A [237]
PP/PP-g-MA/Cl 15A [218]
PP/PP-g-MA/Cl 15A [218]
40 PP-g-MA/I31PS [219]
HDPE/HDPE-g-MA/Cl 15A [238]
HDPE/HDPE-g-MA/N1.44P [238]
20 PS/Cl 15A [243]
PS/Cl 20A [244]
CEAR/NMMT [249]
0 UPE/layered silicate [251]
PU/C20 [252]
PU/C30 [252]
0 2 4 6 8 10
clay [wt%]

Fig. 16. Changes in DSI modulus DEI as a function of nanoclay loading for different polymer/layered silicate nanocomposites.
Organically-modied clays have been used in all cases except in Ref. [249]. Red symbols highlight data that should be taken
with caution (see Table 4 for details). Blue symbols have been employed for epoxy-based composites. Symbols h and M
correspond to values for polished and unpolished samples, respectively;  and  denote composites prepared by SSE and TSE,
respectively.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 61

to the formation of hydrogen bonds [213,226]. In contrast, PLS nanocomposites based on semicrystal-
line non-polar polymers such as polyolens, exhibit small DEI and DH values even with the incorpo-
ration of compatibilizers due to poor clay-matrix interactions (see Fig. 16 and Table 4) [218,238].
Indeed, the enhancements in modulus measured in PP/organically modied MMT (OMMT) nanocom-
posites with maleic anhydride functionalized polypropylene (PP-g-MA) as a compatibilizer [218] and
in high density polyethylene (HDPE)/OMMT nanocomposites compatibilized with polyethylene-
grafted maleic anhydride (PE-g-MA) [238] are not higher than 13% for clay contents 65 wt%. These
values can be raised to DE0  100% at 10 wt% of clay using PP-g-MA as the polymer matrix because
more MA functional groups are incorporated [219].
Fig. 16 and Table 4 show moderate H and E0 increments for PLS nanocomposites based on amor-
phous polymers [242244]. On the other hand, DEI and DH for thermoset PLS nanocomposites are
usually lower than those of thermoplastic based ones [247249,251], as already remarked for organic
llers. Indeed, the largest DE reported for epoxy/OMMT nanocomposites is 20% for a clay content of
7.5 wt% [247] and 23% for a nanoclay reinforced unsaturated polyester (UPE) with 5 wt% ller
[251]. Analogously, the strongest rise in H found for epoxy/OMMT nanocomposites is 50% for a clay
content of 3 wt% [248]. One exception to this trend is found in PU/OMMT nanocomposites with exfo-
liated structures [252], where very large improvements in modulus and hardness (107% and 65%
respectively at 5 wt% ller content) are reported. The consistency of these results will be commented
below. Functionalized organic cations are commonly employed to enhance the thermoset/clay inter-
action. However, it is worth mentioning that some organic modiers, such as alkyl ammonium chains,
may lower the matrix cross-linking density near the clay surfaces [247]. Therefore, the degree of cross-
linking is affected by the degree of interaction and has a competitive effect on the clay reinforcement.
In epoxy/OMMT nanocomposites the smaller increments in properties at higher loadings were attrib-
uted to the reduced crosslinking density of the epoxy matrix due to the higher presence of surfactant,
together with the increased population of intercalated clay clusters [247].
At this point, it becomes apparent that only some general trends regarding EI and H values of PLS
nanocomposites in relation to the nature of the polymer matrix (semicrystalline, amorphous or ther-
moset) can be drawn. To fully understand the differences observed in these properties other crucial
factors must be taken into account.

5.1.2.2. Clay dispersion and type of structure. One of the most important parameters that has to be con-
sidered when analysing the mechanical performance of PLS nanocomposites is the degree of disper-
sion of the clay and, in particular the formation of intercalated and/or exfoliated structures.
Exfoliated structures have been reported to display higher EI and H values than the intercalated ones
[218,242,252]. Yusoh and Song [252] found signicant differences in the surface mechanical proper-
ties of exfoliated and intercalated PU/organoclay nanocomposites thin lms. However, these results
should be taken with caution: in the rst place, the load-depth curves of the neat polymer exhibit a
clear artefact at the beginning of unloading (most probably associated to the test exceeding the max-
imum penetration allowed); in addition, E and H values are remarkably small and possibly affected by
an incorrect determination of the point of initial contact. Higher E0 and H values for the exfoliated sys-
tem compared with the intercalated one have also been reported for poly(propylene carbonate) (PPC)/
organoclay composites prepared by solution intercalation [242]. Treece and Oberhauser [218]
reported the inuence of two different melt-blending strategies (conventional twin-screw extrusion
(TSE) and single-screw extrusion (SSE) with in line supercritical carbon dioxide (SCCO2) feed) for
the preparation of compatibilized PP/PP-g-MA/OMMT nanocomposites. Different degrees of clay exfo-
liation and dispersion were observed by TEM. The high shear of the TSE was very effective in exfoli-
ating and dispersing the clay than the SSE method. DSI experiments were most worthy in evaluating
the effect of processing, dispersion and exfoliation on the modulus. A signicant mechanical enhance-
ment was only found for nanocomposites prepared by the TSE process (see Table 4). In PA-6/clay
nanocomposites E and H increased with clay loading (110 wt%), as expected, but a slowdown of
the growth was observed above 5 wt% clay content [223]. These results were related to the change
in the clay morphology observed when loading exceeded 5 wt%, from exfoliation-dominated to inter-
calation/exfoliation mixture as evidenced by TEM and X-ray diffraction. The same trend was observed
in epoxy/organoclay nanocomposites [247], the mechanical properties exhibiting smaller increments
62 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

above 2.5 wt% clay content due to the formation of intercalated structures besides the exfoliated ones.
Piscitelli et al. [248] investigated epoxy based nanocomposites with sodium montmorillonite (Na-
MMT) untreated and modied with two different aminosilanes prepared by two different dispersion
methods: sonication (S) and combination of sonication and ball milling (SB). Hardness and modulus
increased with the addition of the clay but the enhancement was found to depend on the dispersion
method and the functionalization of the layered silicate (see Table 4). Sonication allowed better nano-
ller dispersion and better interface interaction between the clay tactoids and the epoxy matrix. How-
ever, the combination of sonication and ball milling increased the aggregation of clay tactoids
reducing the interfacial interactions. Consequently, all the nanocomposites prepared by the combined
method showed lower E and H than their counterparts prepared by simple sonication. On the other
hand, those nanocomposites with silylated MMT showed higher values for both properties than the
nanocomposite with Na-MMT. The authors concluded that the interfacial interactions between the
amine moieties anchored to the clay surface and the epoxy matrix play a more signicant role than
does the clay morphology over the composite properties.
From the examples discussed above it can be concluded that dispersion of the clay nanoplatelets in
the polymer matrix is one of the most important factors controlling the EI and H improvement in PLS
nanocomposites independently of the nature of the matrix. Exfoliated structures make the entire sur-
face of clay layers available for the polymer and maximize polymerclay interactions, boosting the
reinforcement. Other factors playing a role in enhancing the nanocomposite mechanical properties
include the organic modication of the clay and the addition of compatibilizers to the polymer matrix.

5.1.2.3. Crystallinity of the matrix. Another parameter that needs to be taken into account is the inu-
ence of the nanoller on the development of crystallinity in PLS nanocomposites based on semicrys-
talline polymers and its impact on the nanoindentation measurements. It is well established that clay
platelets can affect the crystalline structure and crystallization behaviour of the polymer nanocompos-
ites but controversial results have been reported [212]. A number of DSI studies have determined that
the incorporation of clay nanollers in systems with a semicrystalline matrix results in an increase in
EI and H with clay content but simultaneously reduces the degree of crystallinity [223,225,227
229,238]. Between the two competing factors affecting the nanocomposites mechanical behaviour
with increasing clay loading, i.e., the reinforcing effect of the nanoclay and the decrease in crystallinity
of the polymer matrix, the rst one seems to dominate. Sikdar et al. clearly observed this behaviour in
PA-6/OMMT using three organic modiers with the same composition and different end functional
groups [225]. Moduli and H were found to increase as the levels of crystallinities were reduced upon
addition of the OMMTs (Fig. 17). Results were explained invoking different clay-polymer interactions

Fig. 17. Effect of organic modiers on the DSI modulus of PA6/OMMT nanocomposites. Reprinted from Ref. [225], copyright
2007, with permission from John Wiley & Sons, Inc.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 63

for each OMMT that give rise to distinct polymer chain entanglements and interphases. It is important
to point out that in this study the nanoclay ller also introduced a polymorphic change in the crystal-
line structure of the matrix, from the a phase in pure PA-6 to the c phase in the nanocomposites. A
similar H and EI rise concurrent to a crystallinity decrease was observed in HDPE nanocomposites with
organically modied clays and a compatibilizer (PE-g-MA) [238] and in PA-6,6/OMMT nanocompos-
ites [227]. It has been suggested that the presence of the clay connes polymer chains and segmental
mobility hindering rearrangement during crystallization and restricting the formation of crystals.
However, there are also examples in which the enhancements in mechanical properties were not
accompanied by any modication in the matrix crystallinity and thus, were attributed exclusively
to the reinforcing effect of the stiff clay nanoller such as PLA/OMMT nanocomposites prepared by
melt intercalation [235].

5.1.2.4. Processing method. Nanoindentation can also be very useful to explore mechanical anisotropy
induced by the processing method. Injection moulded composites of crystallisable polymer matrices
represent a clear example in which valuable information on the local mechanical properties can be
achieved by means of DSI. Shen et al. [223] studied E and H of injection moulded PA-6/clay nanocom-
posites and found that both properties adopted higher values along the injection direction than per-
pendicularly to the melt ow and increased from the surface to the core of the samples. The difference
in experimental values across the samples was found to be as large as 20%. The results were explained
as due to a parallel rise of clay content and crystallinity from the outer to the inner region of the sam-
ples and of clay orientation along the injection direction. Crystallinity changes were associated to the
temperature gradient effect induced by the injection moulding process. Fig. 18 illustrates the variation
of the mechanical properties at different locations in the moulded bars, exposed by polishing, from the
near surface (A), through the intermediate zone (B), to nally reach the core position (C). Uneven dis-
tribution of the clay has been also put forward to explain the enhanced E0 and H values found in the
inner regions of PA-6,6/clay of the injection moulded bars [229].

Fig. 18. DSI (a) modulus and (b) hardness for three exposed surfaces at 0.5 cm (A), 3 cm (B) and 5.5 cm (C) from the outer
surface of the injection-moulded bars for neat PA6 and its nanocomposites with 5 and 10 wt% clay. Reprinted from Ref. [223],
copyright 2005, with permission from Elsevier.
64 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

Preferred orientation of the clay along the injection direction in poly(ethylene naphtalate) (PEN)/
OMMT nanocomposites has been revealed by small angle X-ray scattering (SAXS) [239]. SAXS results
also showed that, upon annealing, this preferred orientation induced an anisotropic crystalline mor-
phology in the PEN matrix with more secondary lamellae aligned parallel to the clay basal plane.
The higher modulus values measured by DSI in the ow direction were then attributed to the addi-
tional reinforcing effect of the anisotropic arrangement of the crystalline lamellae together with the
clay orientation in this direction. Here, in annealed samples, the degree of crystallinity of the nano-
composites was similar to that of the matrix.
Other effects, such as multiple reprocessing cycles, were investigated in clay/nanocomposites by
DSI tests and correlated with changes in their nanostructure. In polystyrene (PS)/clay nanocomposites
[243], X-ray diffraction experiments determined that the state of intercalation of the clay was
improved after eight processing cycles using TSE. H values of the nanocomposites increased by 10%
after reprocessing (from 380 to 420 MPa) and by 16% compared to that of neat PS (360 MPa). Results
were attributed to better dispersion of the clay in the matrix. In contrast, H of the neat polymer was
slightly reduced to 350 MPa after 8 cycles due to a reduction in molecular weight caused by thermal
degradation. The same behaviour was observed for E0 which increased by 5% with clay incorporation
and by an additional 3% after reprocessing the nanocomposites. However, E0 of neat PS decreased pro-
gressively with the reprocessing cycles due to chain scission.
The effect of gamma irradiation and natural weathering (up to 130 days) on the degradation behav-
iour of PLA/OMMT nanocomposites and its impact on the local mechanical properties were also stud-
ied [253,254]. While neat PLA was strongly degraded by gamma irradiation the corresponding
nanocomposites were less affected. The steric interactions of the clay layers were modied by irradi-
ation which promoted a better dispersion of the clay within the matrix. As a consequence, E0 and H
values of the irradiated nanocomposites showed certain improvement [253]. A slight increase in these
properties, depending on clay content and exposure time, was also observed in the nanocomposites
under natural weathering [254].
As a nal point, the inuence of strain rate on the mechanical properties of PLS nanocomposites has
been investigated by dynamic DSI [228]. In exfoliated PA-6,6/organoclay nanocomposites it was
observed that E0 was not affected by the strain rate whilst H increased with increasing strain rate
for both neat PA-6,6 and the nanocomposite samples. It was suggested that the former elastic
response is insensitive to the strain rate because the motion of the amorphous chains in semicrystal-
line materials is frozen in the glassy state.

5.1.3. Other layered silicate/polymer systems


PLS nanocomposites using polymer blends as matrices have been studied by DSI. Jarrar et al.
reported signicant enhancements in the reduced elastic modulus and hardness of PA-6/PA-6,6 blends
with the incorporation of an organically modied clay, the greatest augment being for the 50/50 com-
position [226]. The enhancements in mechanical properties were supported by FTIR spectroscopy
results which showed the formation of hydrogen bonding and possible formation of ionic bonds
between the polymers and the nanoclays.
Additionally, the inuence of the combination of clays with other micro or nanollers on the nano-
mechanical properties of PLS nanocomposites has also been investigated [246,236]. Biopolymer based
chitosan/MMT/hydroxyapatite (HA) nanocomposite exhibited signicant E and H enhancements as
compared to pure chitosan as well as to chitosan/MMT and chitosan/HA composites [246] (see
Table 4). The improvements were attributed to better nanoparticle dispersion as observed by AFM
and X-ray diffraction and stronger interactions between the three components as evidenced by FTIR
studies. In PLA/Kenaf bre (KF)/OMMT nanocomposites H was improved as compared with PLA/
OMMT and PLA/KF due to good interfacial bonding [236].
Finally, several authors have investigated laminated or multilayer PLS nanocomposites by DSI
[241,245,250]. E and H values of layer-structured polypyrrole (PPy)/MMT nanocomposite lms pre-
pared by electrodeposition increased at very low clay loadings, but above 0.01 wt% both properties
decreased due to MMT agglomerates [241]. Another organicinorganic nanocomposite with laminated
structure was prepared by a hydrothermal-electrophoretic method [250]. In this assembly an acrylic
anodic electrophoretic resin (AAER) was intercalated into the interlayer space of Na+MMT (NMMT) by
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 65

the hydrothermal process (HMMT lms). A remarkable DE0 value (70%) was observed for the HMMT
lms compared with neat NMMT considering that the polymer content in the former is very low
(8 wt% as determined by TGA). This improvement was attributed to the laminated structure that plays
a crucial role in absorbing energy during the elastic deformation of the nanocomposite. The last exam-
ple of PLS layered assemblies is constituted by PDDA/MMMT thin lms prepared by stepwise alternat-
ing polyelectrolyte and clay deposition from solution [245]. The hardness H of the multilayer thin lm
was 0.42 GPa, which is the highest value reported for PLS nanocomposites (Table 4) and is comparable
to that of a soft metal such as copper (0.46 GPa). The well-ordered anisotropic structure also yielded a
high modulus value of E0 = 9.5 GPa.

5.1.4. Comparison with macroscopic properties


Most of the DSI studies on PLS nanocomposites offer, as well, the Youngs modulus values mea-
sured by tensile testing (Table 4). Comparison with the corresponding DSI moduli, reveals that a sim-
ilar trend with clay content and material modication is found although, in general, the absolute
values derived from both techniques are different. EI values have been found to be higher [235], or
lower [226,227] than Y. The discrepancies have been ascribed to scale factors and size effects, differ-
ences in the loading direction and local variations of crystallinity and crosslinking density. Interest-
ingly, injection moulded nanocomposites having a processing-induced mechanical anisotropy
exhibited EI values parallel to the injection ow that were comparable with tensile data. In both cases,
the loading direction was applied parallel to the ow direction [223]. A detailed study of local vs. glo-
bal stiffness has been reported for PP-g-MA/MMT nanocomposites in which DSI was performed in the
clay-polymer intercalated region, the boundary region between clay aggregate and matrix and the
matrix alone [219]. Fig. 19 shows that the local stiffness enhancements determined for indentations
at a depth of 500 and 1000 nm are considerably higher than results from tension and compression.
This is partially due to the increment in local ller content. In addition, other factors such as polymer
chain connement and topological constraints at the nanometer scale need to be considered.
The storage modulus obtained by DMA has also been compared with the storage modulus mea-
sured by dynamic DSI (nanoDMA) and differences have been reported [227,232,233,238,239]. The ori-
gin of the discrepancies was explained again based on size effects and differences in loading directions

Fig. 19. Relative increase in stiffness for PP-g-MA/MMT nanocomposites vs. clay content. Stiffness values were measured by DSI
at two different indenter displacements (500 nm and 1000 nm) and by tensile and compressive testing. Reprinted from Ref.
[219], copyright 2006, with permission from Elsevier.
66 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

and load frequency used. It is worth mentioning that in one DSI study [225] both, E and E0 , were deter-
mined. It was found that E0 followed the same trend as E as a function of clay content and of the dif-
ferent organic modiers of the clay. The absolute values from both methods were in this case very
close to each other (see Table 4).

5.1.5. Indentation size effect


A variation of EI and H as a function of indentation depth has been reported in a number of DSI
studies on PLS nanocomposites [219,224,227229,232,233,240,242,247]. Most of them used the
dynamic method [219,224,227229,232,233,242] and found an increase of the mechanical properties
with decreasing penetration depth. At shallow depths (300 nm), the behaviour was attributed to a
number of factors, as have been explained in preceding sections of this review (see Sections 2.1.3,
4.1.5, 4.2.3 and 4.3.3). At higher indentation depths, in the range 10006000 nm, Beake et al. [240]
investigated PEO/clay nanocomposites by means of a number of load-partial unload experiments.
The authors observed a continuous decrease of E and H values with increasing penetration depth in
the whole interval studied, the variation being more pronounced in the solution processed samples
than in the melt processed ones. Results were partially associated to a strain rate effect arising from
a trapezoidal load function and not to a real gradient in mechanical properties (see Section 2.1.3 for a
detailed explanation).

5.1.6. Creep properties


The creep behaviour of PLS nanocomposites has also been investigated by DSI
[223,224,238,240,244,247,255]. In general, a decrease in creep displacement and rate has been
observed in PLS nanocomposites with increasing clay content showing that the incorporation of rigid
clay nanollers improves the creep resistance [223,238,244,255]. In injection moulded PA6/MMT
nanocomposites the creep behaviour was investigated on the skin and core regions of the samples
and compared with creep measured by DMA cantilever-bending [255]. Both techniques showed a
decrease in creep compliance with the addition of the clay but, only the cantilever measurements
revealed an additional improvement in the time-dependent compliance behaviour. The authors
explained this behaviour considering the higher capacity of the organoclay to restrict the molecular
mobility in the bulk compared to the surface. In other systems, factors related to the nanostructure
of the polymer matrix seem to have an inuence on the creep behaviour. In PA66/MMT nanocompos-
ites, an unexpected increase in creep displacement was observed with increasing clay content [227].
This reduction of creep resistance was attributed to the decrease of crystal size and crystallinity in the
nanocomposites as revealed by X-ray diffraction. In epoxy/OMMT nanocomposites the creep resis-
tance increased when clay loading was less than 2.5 wt% and the opposite behaviour was found at
higher loadings [247]. This result was discussed on the basis of a reduced crosslinking density near
the clay surface.

5.2. Spherical inorganic nanoparticles

The high aspect ratio of the nanollers in most of the aforementioned polymer nanocomposites
(e.g., carbon nanotubes, carbon nanobres, graphenes or nanoclays) often results in a large increase
of the nanocomposite melt viscosity as llers resist shear. The melt viscosity rise can cause undesir-
able slower production rates and higher processing costs. The use of spherical nanoparticles opens the
possibility of achieving both a viscosity reduction together with a reinforcement of the tensile mod-
ulus. It has been shown that the size and shape of nanoparticles can be a key parameter to modify rhe-
ological and mechanical properties of polymer nanocomposites [256].

5.2.1. Characteristics, preparation methods and types of nanocomposites


The most common inorganic nanollers, generally adopting a spherical shape, can be classied in
three main groups: metallic nanoparticles, oxide nanoparticles and other miscellaneous spherical
llers.
Currently, nano-sized metal particles are a focus of interest in biomedical sciences, engineering and
many other elds. These materials can be synthesized and modied with various chemical functional
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 67

groups which allow them to be conjugated with antibodies, ligands, and drugs, opening a wide range
of potential applications in biotechnology, drug delivery, and diagnostic imaging [257]. Other signif-
icant applications result from their optical and electromagnetic properties, especially those related
to quantum-size effects such as photo and thermo-luminescence, dichroism, size-dependent ferro-
magnetism, superparamagnetism, electromagnetic wave absorption, and interference shielding or
super-catalytic activity [258]. The difcult handling of metallic nanoparticles makes embedding into
polymer matrices a valid solution for many applications because most of the mentioned properties
remain to a certain extent in the nanocomposites. Concerning the mechanical properties of polymer
nanocomposites including metallic nanollers, the literature search performed for the present survey
did not yield any DSI study, representing an area of great potential for near future investigations.
Inorganic oxide nanoparticles share many of the properties of the metallic nanollers. They are also
essential in modern technologies such as microelectronics (silica, high-k dielectrics), sensors (tin
oxide), corrosion protection (alumina, zirconia), optoelectronics (transparent conductors), energy pro-
duction and storage (titania, zeolites), heterogeneous and environmental catalysis (transition metal
oxides), drug carriers or biocompatible materials (HA). Common methods for nanocomposite prepara-
tion are solution blending techniques, graft polymerization, solgel methods and surface modication
of the nanoparticles [259].
Among the numerous inorganicorganic nanocomposites, polymer composites reinforced with
nanosilica (SiO2) are the most commonly reported in the literature, having attracted substantial aca-
demic and industrial interest and being employed in a variety of applications. Silica nanoparticles
present several advantages, for instance ease of preparation at a relatively low cost, possibility of per-
forming surface modications with different functional groups and acceptable biocompability. These
nanocomposites can be prepared following various synthetic routes, according to the way that each
phase is introduced. The organic polymer matrix can be introduced as a precursor (monomer or oli-
gomer), as a preformed linear polymer (in molten, solution, or emulsion states), or as a polymer net-
work, physically or chemically cross-linked. The nanoller, in turn, can be introduced as pre-existing
nanoparticles or as precursors such as tetraethyl orthosilicate (TEOS), tetramethyl orthosilicate
(TMOS) or perhydropolysilazane (PHPS). This leads to three general methods for the preparation of
polymer/silica nanocomposites according to the starting materials and processing techniques: blend-
ing, solgel processes, and in situ polymerization (Fig. 20) [211]. Blending involves simple mixing of
the polymer matrix and the silica llers, and agglomeration of nanoparticles is a usual drawback.
The second method is based on direct mixing of monomers with silica particles followed by a poly-
merization process. In situ polymerization begins with a mixture of both silica precursors and polymer
monomers, followed by the polymerization of precursors and monomers [78].
Because of the strong tendency towards aggregation among inorganic nanoparticles, which may
depress the overall physicalchemical properties, considerable efforts have been devoted to optimize
the interfacial interaction between the two nanocomposite phases, i.e. to enhance the compatibility
between the polymer (hydrophobic) and the nanosilica. A common procedure is to modify the surface
of the silica nanoparticles (especially for the blending and in situtechniques), which simultaneously

Fig. 20. Scheme showing the three general approaches to prepare polymer/silica nanocomposites. Reprinted from Ref. [211],
copyright 2008, with permission from the American Chemical Society.
68 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

can enhance the dispersion of nanosilica in the polymer matrix [260]. The surface modication of
nanosilica can be carried out by either chemical or physical methods. The former involve modication
either with coupling agents or by grafting polymers while physical methods use surfactants or mac-
romolecules adsorbed onto the surface of the silica nanoparticles. Composites with nanoparticles con-
taining an inorganic core and an organic shell or surface modied nanoparticles can be considered a
special type of nanocomposites and they are often designated as hybrid nanoparticles [261]. For fur-
ther details, the reader is referred to the recent and comprehensive review by Zou et al. on polymer/
silica nanocomposites [211].

5.2.2. Modulus changes in nanocomposites containing spherical inorganic nanoparticles


The DSI results published for the nanocomposites containing spherical inorganic nanoparticles are
summarized in Fig. 21 and Table 5. The gure summarizes the percent variation, in relation to the
value of the corresponding neat polymer matrix, of DSI elastic modulus data for different polymer
matrices reinforced with spherical nanoparticles as a function of the nanoller content in weight per-
cent. A general look at the gure shows that the proportion of loading is usually much higher than
those reported for other types of nanollers, probably due to a comparatively lower cost, but also
because of the especial applications in which they are involved. In the following, a discussion on
the effect of the type of matrix on the mechanical properties of polymer/silica nanocomposites will
be presented, together with a description of nanocomposites including different oxide nanoparticles
and other miscellaneous spherical shaped llers.

5.2.2.1. Polymer/silica nanocomposites. (a) Thermoset matrices


Several studies have reported the characterization of epoxy/nanosilica nanocomposites by DSI
[75,76,262264,268]. Lam and Lau [262] emphasized the importance of studying the localized elastic
modulus by means of DSI to understand the distribution of nanoparticle agglomerations in epoxy/
nanosilica composites. The nanocomposites were prepared by preheating and ultrasonication at dif-
ferent temperatures followed by in situ polymerization. DSI experiments on the cross-section of these
composites revealed that at lower sonication temperatures (40 C) E decreased as the indenter moved
from the bottom to the top sample surface. This result was attributed to the slow curing time which
favours the gravitational effect of the nanoclay. At higher temperatures, however, local mechanical
properties of the composites were evenly distributed throughout the entire sample. At a ller concen-
tration of 4 wt%, EI was found to increase from 2.3 to 5.5 GPa across the cross-section of the sample for

Fig. 21. Changes in DSI modulus DEI as a function of nanoller loading for different polymer nanocomposites incorporating
inorganic spherical nanollers. See Table 5 for comments on the data in red colour. The symbols M, O, s and } correspond to
functionalized nanoparticles. indicates nanosilica grafted to the polymer matrix.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 69

a sonication temperature of 40 C, and up to a constant value of 8.8 GPa when the temperature was
elevated to 100 C. It must be noted that the remarkably high absolute values provided (see Table 5)
should be taken with caution because of the notorious nose effect observed in the experimental load-
depth curves. Very recently the work by Tzetsis et al. [263] illustrated the negative effects of clustering
on the mechanical properties of epoxy nanocomposites reinforced with fumed silica. Both, uniaxial
tensile tests and DSI results, showed that Y, E and H of the nanocomposites steadily decreased with
increasing fumed silica content. The decrease in E was around 25% for a composite with 3 wt% SiO2,
attributed to the strong tendency to cluster formation of fumed silica as also revealed by optical
microscopy (OM) and SEM analyses. In this case, the sonication procedure and the high mechanical
mixing followed by curing at RT and postcuring at 100 C for 2 h were not sufcient to yield a good
dispersion. A discussion on different possible sources of error concerning E data was also given, being
one of them that the tip area function can be material-dependent (see Section 2.1.1 and [12]). Direct
incorporation of dehydrated nanosilica with 10 nm average diameter in an epoxy matrix, by simply
mechanical mixing and ultrasound waves, was also used by Allahverdi et al. [264]. This study however
found signicant increments in E and H obtained by DSI with the addition of the nanoller to the poly-
mer. A nanocomposite with 5 wt% of well dispersed nanosilica showed DE and DH values of 21% and
26%, respectively, as compared with the neat epoxy. It is noteworthy that the DSI experimental pro-
cedure used does not seem to contemplate a hold period at peak load which might cause inaccurate
results. The reinforcement effect of the silica nanoparticles was also found to improve the storage
modulus E0DMA of the composite in both glassy and rubbery regions. The surface modication of SiO2
has been shown to more effectively enhance the tensile elastic modulus and yield strength of an epoxy
resin [265,266]. Following this route, a solgel technique was employed by Wang et al. [76] and Zhang
et al. [75] to prepare epoxy based nanocomposites containing surface-modied nanosilica particles
with average size of 20 and 25 nm, respectively. This procedure resulted in homogeneously dispersed
nanoparticles even at high ller contents [267] (Fig. 22, left). In the former study [76], the authors per-
formed micro and nanoscale indentation and scratch tests on the nanocomposites to determine the
inuence of the load content on the mechanical and tribological properties. Indentation results
showed that H and E of the composites increased monotonously with the particle content reaching
increments of 33% and 40%, respectively, by the addition of 24 wt% nanosilica. The experimental val-
ues were well tted using the HashinShtrikman lower bound model [73], which is situated slightly
above the Reuss limit. In addition, scratch tests also revealed an effective improvement in the tribo-
logical properties of the epoxy with the addition of proper amounts of SiO2. Zhang et al. [75], in turn,
reported the viscoelastic properties of similar nanocomposites by dynamic nanoindentation and

Fig. 22. (a) TEM picture of silica/epoxy nanocomposite with 5 wt% SiO2; (b) Dynamic storage modulus vs. indentation load
using a frequency of 75 Hz for the different composites investigated (from top to bottom: 0, 2, 5, 10, 17, 23 wt% silica). Adapted
from Ref. [75], copyright 2009, with permission from John Wiley & Sons, Inc.
70 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

derived their dynamic storage modulus E0 and loss tangent. E0 was found to increase with the force
frequency, denoting a loading rate effect and also with the nanosilica volume fraction (Fig. 22, right).
Using the proper indentation loads, DE0 yielded a value of 46% with the addition of 23 wt% nanosilica.
Castrillo et al. [268] used PMMA microparticles lled with nanosilica (10 wt%) to modify an epoxy
matrix. An addition of 5% microparticles resulted in a nanosilica content in the nanocomposite of only
0.5%. This nal small content of ller was not intended to achieve high mechanical reinforcement, but
to enhance other properties such as absorption of ultraviolet light, electric or magnetic shielding, con-
ductivity or dielectric behaviour. However, the good dispersion of the silica nanoparticles in the
PMMA domains, as well as that of the silica-lled PMMA microparticles in the epoxy matrix, main-
tained or even slightly increased Martens hardness (2%) and modulus (4%) as compared to those of
the neat epoxy.
Apart from epoxy resins, other DSI studies on thermosets such as crosslinked polyphenylenes or
polyacrylates with nanosilica as reinforcement have been found in the literature [77,210,269,270].
In the case of polyacrylate matrices, Fig. 21 shows that the enhancement in mechanical properties
is much higher than the one found for epoxy thermosets. Lin et al. [77] reported the fabrication of
low dielectric constant (low-k) nanocomposite thin lms to be used in small semiconductor devices.
The lms were prepared by spin coating and thermal curing of solution mixtures of low-k thermoset
prepolymers (highly crosslinked polyphenylenes) and silica nanoparticles with an average diameter of
8 nm. The mechanical properties of these nanocomposites were characterized by means of DSI using a
dynamic method at very small loads. It was found that the addition of the nanoller enhanced both E0
and H (around 50% and 40%, respectively, for 40 wt% SiO2). DE0 was smaller than that predicted by the
HalpinTsai equations [74]. This result was tentatively explained due to poor interfacial adhesion and/
or aggregation of the hydrophilic silica nanoparticles in the hydrophobic thermoset matrices. The
addition of SiO2, however, had strong negative effects on the insulating properties of the nanocompos-
ites which were attributed to the presence of impurities in the silica nanoparticle solution. The DSI
dynamic method was also used to test acrylate-based nanocomposites prepared by UV in situ poly-
merization [210,271]. The results showed strong E0 and H increases, by more than 100% and 200%,
respectively (Table 5) for a nanocomposite with 5 wt% SiO2. Devaprakasam et al. [269] aimed to
understand the different tribo-mechanical performance of a micro and a nanosilica reinforced polymer
composite produced by visible light curing. In both cases the load content was 56 vol% in matrices of
monomeric dimethacrylates. The nanocomposite included 4070 nm diameter nanosilica and the pri-
mary particles in the microcomposite presented a bimodal size distribution ranging from 200 to
500 nm and 1 to 4 lm, respectively. DSI results showed that E and H of the nanocomposite were
homogeneous throughout the sample surface yielding extremely high absolute values (around 15
and 0.7 GPa, respectively). On the other hand, results for the microcomposite were found to be highly
heterogeneous. The sharp microparticle edges increased friction and wear, giving also rise to a high-
energy dissipation. Such behaviour was only locally found in the nanocomposite in the presence of
nanoparticle clusters and aggregates. Soloukhin et al. [270] reported the preparation of hybrid nano-
sized silica particles (515 nm) by mixing a colloidal silica dispersion in a water-containing organic
medium with a silane-coupling agent, 3-(trimethoxysilyl) propyl methacrylate (MEMO) and subse-
quently with different methacrylic monomers and a UV photoinitiator. The obtained cross-linked sil-
ica-methacrylate hybrid coatings were deposited on PC substrates and the mechanical properties were
determined using DSI. They found a modulus E increase of almost 200% for a silica content of 48 wt%.
The mechanical properties attained were found to be dependent on the chemical nature of the mono-
mer used that considerably inuenced the degree of crosslinking density. The authors concluded that,
in general, the properties of hybrid coatings cannot be simply discussed in the light of ller content,
but other factors such as the mutual inuence of ller content on cross-link density, coating thickness
and chemical composition have to be taken into account.
(b) Thermoplastic matrices
Bhattacharya and Chaudhry [78] tested silica-reinforced PVA biocompatible nanocomposites with
the nal aim of mimicking the properties of human bone. A homogeneous silica dispersion in water
containing PAH was added to a water solution of PVA and cast on glass, forming 40 lm thick lms.
The indentation modulus E increased from 300 MPa for neat PVA to 8.1 GPa for the nanocomposite
with 55 wt% nanosilica (2600% increase), as shown in Fig. 21. The results from this work were reported
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 71

by the authors as one of the largest increases in modulus found in the literature, being certainly the
largest in this review. However, the reader should be aware of the very low modulus values found
for the neat polymer which can affect the obtained D E values. Results were explained as a conse-
quence of the attachment of the molecular chains of PVA to the surface of the functionalized silica
nanoparticles through hydrogen-bonding interactions between the hydroxyl groups of PVA and the
amide groups of PAH. The good dispersion in the PVA matrix and the absence of particle agglomera-
tion also contributed to the superior mechanical properties. A further factor that should be taken into
account for a semicrystalline matrix is the crystallinity variation in the presence of the ller. Here, as a
gradual decrease of polymer crystallinity with increasing ller content was observed, it would have
contributed to a reduction in the modulus of the matrix. Hence, the increase in mechanical properties
due to ller reinforcement was even more remarkable. A similar behaviour was observed for polymer/
layered nanocomposites as discussed in Section 5.1. Moreover, by coating the surface of the silica
nanoparticles with calcium HA, the resulting nanocomposite (55 wt% of silica) further increased the
modulus to 11 GPa, a value that is close to that of cortical bone.
As a new example of hierarchical polymer composites (see Section 4.1.3.3) that are reinforced by
more than one reinforcing agent, Molazemhosseini et al. [272] reported the fabrication by the melt-
mixing process at 400 C of PEEK based hybrid composites reinforced with short carbon bres (SCFs)
and nano-SiO2 particles. Nanoindentation and nanoscratch methods were used to evaluate the nano-
mechanical phases, namely the bulk PEEK matrix, the SCFs and the interphase region. Results revealed
an instantaneous elastic recovery in the SCFs while a time-dependent contribution was found in the
neat matrix denoted by a nose in the unloading cycle. Such an effect could be favoured by the absence
of a holding time at the end of the loading cycle and as a consequence the accuracy of the experimen-
tal results can be questioned. It was found that the incorporation of 25 wt% SCFs into neat PEEK
yielded a remarkable DE of 140%. In addition, the joint presence of nano-SiO2 particles in the conven-
tional composite was shown to effectively contribute to additional E and H improvements of 44% and
22% for a nanosilica content of 2 wt%, respectively.
Stojanovic et al. [273] reported a different route for the surface modication of nanosilica, namely
using SCCO2 as an environmentally friendly solvent to improve the mechanical properties of PMMA. A
MEMO-type coupling agent was also used. A comparison of conventional and supercritical coating
methods revealed enhanced properties for the nanocomposites obtained through the latter processes
regarding particle size distribution, amount of coated silane, and homogeneous dispersion in the
PMMA matrix. Thus, the nanocomposite containing 3 wt% of SiO2 obtained by supercritical processing
of the nanosilica sol showed an increase in E and H of 19% and 34%, respectively.
Low temperature solgel methods, although frequently used to produce nanosilica composites,
have two main drawbacks: silica synthesized contains many lattice defects and the composites are
often contaminated by catalysts. For this reason, Saito et al. [274] synthesized PS/silica nanocompos-
ites from a silica precursor such as PHPS that could overcome both difculties. A blended organic solu-
tion of PHPS and PS derivatives with hydroxyl groups gave rise to grafted block copolymers containing
PHPS branches. Solution casting and subsequent calcination at around 100 C produced composite
lms with microphase separated domains of PHPS and PS. Clear microphase separation was observed
for the composites when the degree of grafting was close to 100% and in some cases silica spheres of
40 nm of diameter surrounded by the organic matrix were obtained. EI and H were found to increase
with the silica content, although one can notice a nose effect at the beginning of unloading; no sig-
nicant differences were observed among the different morphologies. Organic/silica nanocomposites
prepared with PHPS could then be used as hard coatings, lighter than the pure silica counterparts. The
same authors [275], prepared nanocomposites of PMMA-rich spheres in a silica-rich matrix following
a similar PHPS route to provide a convenient hard coating for neat PC lms. They obtained a maximum
surface hardness for the composites on the substrate of 1.07 GPa.
Douce et al. [276] synthesized silica nanoparticles from TEOS, a frequent precursor. The llers had
different sizes (from 15 to 60 nm diameter) and two types of surface modication in order to promote
or minimize interactions with the matrix. The obtained nanosilica were prepared as colloids and intro-
duced at different concentrations in glycidoxy-propyltrimethoxysilane (Glymo) type coatings. Such
coatings are meant to protect transparent polymer lenses which usually have a poor scratch resis-
tance. Both, indentation and scratch experiments were used to characterize the mechanical properties
72 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

of the lms. The moduli E gradually increased with the addition of llers, being this effect larger (70%
increase at 15 vol% SiO2) for the ller with smaller size (15 nm). On the other hand, the weakness of
matrix-ller interactions had no signicant effect on the modulus of the Glymo lms, while coreshell
llers improved it slightly.
Polymer/silica nanocomposites based on poly(2-hydroxyethyl acrylate) (PHEA) were prepared by
Rodrguez Hernndez et al. [277] by the simultaneous polymerization of the organic and the silica
phases in a solgel process using TEOS as silica precursor. The structure of this system was investi-
gated using AFM in the tapping mode and DSI experiments. AFM nanoindentation explores the visco-
elastic behaviour up to 30 nm depth from the surface, but the obtained stiffness also depends on the
structural characteristics of the surrounding material. The stiffness increase was quite signicant for
silica contents lower than 10 wt% (from around 0.08 N/m for the PHEA homopolymer to 0.27 N/m for
the composite containing 10 wt% of silica, i.e. more than 200% increase in stiffness). However, there
was a further qualitative change in stiffness for silica fractions above 15 wt%, resulting in highly scat-
tered stiffness values. The explanation offered was that for a silica content lower than 15 wt%, the sys-
tem consisted of isolated silica aggregates dispersed in the organic matrix, while above that
concentration the structure became co-continuous with that of the organic matrix. Hence, silica llers
behaved as an inorganic scaffold improving the mechanical properties of the nanocomposite.

5.2.2.2. Other oxides. This subsection includes a few examples of other metal oxides (e.g., ZnO, Al2O3,
Fe3O4) frequently used as nanollers and investigated by means of DSI. Ciprari et al. reported the char-
acterization of the mechanical properties of four polymer nanocomposite systems by DSI tests and
DMA and investigated the role of the interphase structure on their behaviour [278]. Alumina
(Al2O3) and magnetite (Fe3O4) nanoparticles were dispersed in PS and PMMA matrices, which were
chosen on the basis of their differing reactivity with metal oxides. The structure of the interphase
was investigated and correlated with the mechanical properties of the composites. The results indi-
cated that Al2O3 nanoparticles were more reactive with the polymer matrix than Fe3O4 counterparts,
but neither showed strong interactions with the matrix as compared to other studies. The low inter-
phase density around the high number of nanoparticles resulted in a decrease of the tensile modulus
of the nanocomposites compared with those of the neat polymers.
Conducting polymer materials used for instance in photovoltaic devices are easily degraded by the
combined action of oxygen and moisture. Polymer nanocomposites are suitable materials for the
encapsulation and protection of such devices. Gupta et al. [279] reported a technique based on grafting
hydride-terminated PDMS to allyltrimethoxysilane functionalized c-alumina nanoparticles to make
nanocomposites that can be used for the abovementioned purpose. The DSI technique was used to
analyse the mechanical characteristics of the composites. In all cases E values were extremely low
(see Table 5) but, from the unloading curves, it was clear that a large elastic recovery took place after
the removal of the load which is an important characteristic of an encapsulant. The work by Chakr-
aborty et al. [280] aimed to compare the scratch hardness with the indentation hardness of PMMA/
ZnO nanocomposites produced by solution blending and in situ spin coating. A good correlation
was found between both properties at low indentation depths, where a linear relationship between
hardness and the ZnO content was observed. Ramezanzadeh and Attar [281] studied the corrosion
resistance of polymer composites to be used as coatings to protect metal parts. For this purpose epoxy
nanocomposites containing different contents of nanozirconia particles were prepared by in situ poly-
merization. The nanocomposites were exposed to 3.5 wt% NaCl solution up to 60 days, and their
mechanical properties (before and after exposure to NaCl solution) were studied by DMA and DSI
techniques. Results showed that the blank sample was severely deteriorated after exposure to the cor-
rosive electrolyte due to a signicant decrease of cross-linking density with a concurrent diminution
of indentation hardness. It was shown that these properties improved with the addition of nanopar-
ticles giving rise to nanocomposites with better corrosion resistance.

5.2.2.3. Miscellaneous spherical shaped llers. A few studies [22,282] have been devoted to investigate
the nanomechanical properties of polymer composites reinforced with nanoparticles of layered metal
dichalcogenides like WS2and MoS2. These non-carbon materials possess similar structure to fuller-
enes, and are named as inorganic fullerene-like (IF) nanoparticles. They have exceptional properties
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 73

such as very high stiffness and strength, attributed to their small size (typically in the range of 40
180 nm); closed-cage layered structure and excellent solid lubricant behaviour. Moreover, because
of their chemical inertness and low agglomerating tendency, they can be homogenously dispersed
within polymer matrices via simple melt-blending technique [283]. The reinforcing effect of IF-WS2
nanoparticles on three different thermoplastic matrices, PP and polyphenylene sulphide (PPS) [22]
and PEEK [282] has been recently investigated by DSI (Table 5). In the rst study [22,284,285],
dynamic DSI was used to obtain the mechanical properties of both polymer composites and they were
found to initially rise with increasing ller content and then level-off at loadings of 0.5 wt%. The
mechanical improvement was discussed in terms of the inuence of the nanoparticles on the nano-
structure of the polymer matrices and the reinforcement due to intrinsic properties of the ller,
including matrix-particle load transfer. The analysis of the H data suggested that the nanostructural
changes induced in the polymer matrices can contribute up to DH  10%, while the ller reinforce-
ment can reach DH  50%. Remarkable DE0 and DH values were found for PPS based nanocomposites
(by up to 35% and 60%, respectively, at 8 wt% IF-WS2), while only moderate enhancements were
observed for the PP counterparts (around 18% for the same concentration). This discrepancy was
attributed to the different nature of the amorphous phase for both nanocomposites.
In the second study [282], IF-WS2 nanoparticles were incorporated into PEEK coatings using an aer-
osol-assisted deposition process with the aim of reducing the coefcient of friction (COF) and improv-
ing the wear resistance. Higher proportions of IF-WS2than the abovementioned ones were added and
the results showed signicant enhancements in the mechanical properties of the nanocomposite coat-
ings. In particular, E and H continuously increased up to 60% at 20 wt% IF-WS2 concentration. The
inuence of possible crystallinity changes was not discussed. On the other hand, the COF was reduced
by 70% upon incorporation of 2.5 wt% of the nanoller, but a further increase in the IF-WS2 content did
not produce any additional reduction in the COF of the coatings.
Nanosized spherical HA has also been used to improve the mechanical properties of biopolymers. A
study by Thomas et al. reported for the rst time the electrostatic cospinning of collagen (type I) and
HA (Fig. 23a and b), representing a potential nanobrous osteoconductive and bioactive nanobiocom-
posite scaffolds [286]. The HA nanoparticles exhibited a mean diameter size of 100150 nm (Fig. 23c).
SEM analyses showed a well interconnected pore network structure with nanobrous morphology of
randomly oriented bres in the diameter range of 500700 nm, depending on the composition
(Fig. 23a and b). The bre diameter increased with increasing HA content. Tensile testing and dynamic
DSI were used for the mechanical characterization (Fig. 23d). The neat collagen brous material
showed a dynamic modulus E0 of 0.2 GPa, value that increased to 0.6 GPa (200% difference) with the
addition of 20 wt% HA. This strong rise was attributed to an increase in rigidity of the polymer due
to the addition of HA and the strong adhesion between the two composite components. It is worth
mentioning that the small modulus values measured in this work bring up serious experimental prob-
lems, especially in the detection of the point of initial contact (see Section 3), a fact that was mini-
mized by using the dynamic DSI option. Nevertheless, the signicant enhancement of mechanical
behaviour achieved by the incorporation of HA can be clearly discerned from a quick inspection to
the load-depth curves of Fig. 23d. As a nal point it was established that the chemical crosslinking
of collagen (x-collagen) further increased the absolute values of the mechanical properties of the
nanobiocomposites.
Organic electronics based on conjugated polymers constitute a promising route for low cost and
exible optoelectronic applications. A method to increase device efciency is to incorporate semicon-
ducting nanostructures such as quantum dots (QDs) into the polymer. QDs are nanoscale particles
with tunable optical and electrical properties. McCumiskey et al. [287] reported the preparation
and mechanical characterization of nanocomposite thin lms consisting of CdSe QDs and the electro-
luminescent polymer poly[2-methoxy-5-2(20 -ethylhexyloxy-p-phenylenevinylene)] (MEH-PPV). The
composite lms were prepared by solution blending and spin casting on glass substrates, using differ-
ent concentrations of QDs. QD dispersion was assessed by means of TEM. The QDs were found to be
more homogeneously dispersed at 90 wt% than at 50 wt% (Fig. 24a and b). The former concentration is
more typically employed for QD polymer devices. The modulus values E of the polymer matrix linearly
increase with increasing QD volume content (Fig. 24c and d). This is a common result among polymer
nanocomposites, however, it is quite exceptional to keep a linear increase up to very high nanoller
74 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

Fig. 23. SEM micrographs of electrospun scaffolds: (a) collagen bres and (b) collagen +10% nanoHA (uncrosslinked), (c) TEM
image of nanoHA powders and (d) Nanoindentation loaddisplacement curves of collagen/nanoHA biocomposites. As nanoHA
content in the bre increases, the maximum load also increases. Adapted from Ref. [286], copyright 2007, with permission from
Elsevier.

contents (i.e. 95 vol%). In this case, a DEI of 240% was observed for a QDs content of 95% (from 11.7 to
41.2 GPa).

5.2.3. Indentation size effect


The EI and/or H variation with penetration depth have been observed in a number of papers
[75,210,270,273,274,287]. For silica nanocomposites and using a dynamic DSI method, an ISE was
found for penetration depths below 200 nm [210]. This range was considered of little value to derive
any material information. Some authors have recommended the application of indentation loads as
large as possible (over 200 lN) to get constant and reliable properties [75].
Soloukhin et al. investigated the typical response of a coating-substrate system where the substrate
(PC) is softer than the polymer methacrylate/silica nanocomposite [270]. Thus the strong decrease in E
with increasing indentation depth, using loads as high as 1 N, was attributed to the inuence of the
softer PC substrate. On the other hand, the absence of a constant E plateau at shallow indentation
depths where the inuence of the substrate was negligible, was tentatively explained as due to a com-
bination of sink-in and pile-up effects. The effect of a harder glass substrate, typically producing a
reverse ISE which often comes together with a normal ISE at smaller penetration depths, has been also
reported [274,287]. The inuence of a glass substrate on the H variation with indentation depth of PS/
silica nanocomposites [274] and on the E dependence for MEH-PPV/CdSe nanocomposites was studied
[287]. In both cases, a discussion on which penetration was appropriate for mechanical characteriza-
tion of the material was carried out. In the rst case, H values were taken at 20% displacement [274],
while McCumiskey et al. determined the average modulus E and hardness values for each sample
within an optimum range, in between the observed ISE and reverse ISE effect (Fig. 24c) [287]. Finally
it is worth mentioning that a reverse ISE was also observed for indentation loads >10 mN in PMMA/
nanocomposites [273].
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 75

Fig. 24. Plan-view TEM images showing CdSe QD dispersion at (a) 50 wt% in MEH-PPV, and (b) 90 wt% in MEH-PPV, drop-cast
from pyridinechloroform binary solvent mixtures. (c) Elastic modulus as a function of contact depth and (d) as a function of QD
loading (wt% and vol%) in MEH-PPV (error bars represent values within one standard deviation of the mean). Adapted from Ref.
[287] with permission from IOP Publishing.

5.2.4. Creep properties


The only study involving creep behaviour concerns MEH-PPV/CdSe lms [287]. The indentation tip
displacement was monitored as a function of time during the 30 s hold period at maximum load. It
was shown that the amount of viscoelastic creep decreased with increasing CdSe QD loading. Creep
results also determined that high QD contents suppressed the viscoelastic behaviour of the matrix.
One could expect that it was a consequence of a granular character of the lms due to insufcient
polymerpolymer bonding that would lead to a stiffness decrease. However, the actual stiffness
increase had to be explained by a different bonding mechanism which could be envisaged as sintering.

5.2.5. Comparison with macroscopic mechanical properties


Dynamic DSI (nanoDMA) and DMA data, concerning nanocomposites incorporating spherical nano-
particles, have been compared in a number of studies [22,75,278]. In most cases, although the relative
variations DE0 and DE0DMA are quite similar, the absolute modulus values usually are higher for dynamic
indentation, E0 > E0DMA . For epoxy/silica nanocomposites [75], E0 and E0DMA results were also compared
with quasi-static E data [76], Youngs modulus values Y and with theoretical results using microme-
chanical models (VoigtReuss and HashinShtrikman [73]). It was shown that E0  2E0DMA using the
same force frequency for both techniques. In turn, E and Y were found to be more similar to E0DMA than
to E0 . Flores et al. [22] discussed the higher E0 values obtained by DSI in PP/IF-WS2 and PPS/IF-WS2
nanocomposites as compared with those obtained by DMA in the light of the different directionality
of the stresses applied in both methods. Quasi-static DSI moduli E have also been found to be higher
76 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

than E0DMA in a number of cases [264,273]. On the other hand, two studies [210,286] reported higher E0
values than the corresponding tensile properties. In the second one, concerning porous electrospun
materials [286], the difference was attributed to the fact that tensile properties, in bulk, mainly
depend on their porosity, total bre content, and orientation of bres whereas nanomechanical prop-
erties should be closer to those of the individual electrospun bres. The weakness of this argument is
that using penetration depths of 500 nm with a Berkovich indenter, and being the bre diameter of the
same order, it is difcult to gure out a deformation eld not including a good number of bres, pores
and different orientations. Tzetsis et al. [263], working on fumed silica/epoxy systems showed that E
and Y modulus values practically coincided when using a modied area function related to the neat
epoxy polymer as a reference. It was considered that the typical calibration procedure which involves
calibration on a hard reference material such as fused silica may not resemble the contact situation
with softer surfaces. Finally, and in contrast with other studies, Bhattacharya et al. for PVA/silica nano-
composites obtained modulus values derived from DMA and tensile tests which were comparable
with those of quasi-static DSI and seemed to agree well with the ones obtained using the HalpinTsai
equations [78].

5.3. Other inorganic nanollers

In addition to layered silicates and spherical nanoparticles, other inorganic nanollers such as two-
dimensional graphene-like boron nitride (BN) [288], inorganic fullerene-like tungsten disulphide IF-
WS2 nanotubes [289], AlOOH boehmite nanorods [290], calcium silicate hydrate (CSH) [291], polyhe-
dral oligomeric silsequioxane (POSS) [292295], to mention but a few, have been incorporated into
polymer matrices. Amongst them, POSS has received extensive research attention for the preparation
of hybrid nanocomposites. It has a nanometer-sized cage-like structure composed of a cubic octameric
molecule with an inner inorganic silicon and oxygen framework that is externally surrounded by
organic functions. It can be depicted by the formula (RSiO1.5)n where n is an even number and
R = H, Cl, or a variety of organic groups. POSS exhibits superior thermomechanical properties in terms
of wearability, thermal stability, oxidation resistance and mechanical strength. The dispersion behav-
iour of POSS molecules in a polymer matrix, and consequently the mechanical properties of the result-
ing nanocomposites depend strongly on the nature of the R groups. On the other hand, organic
inorganic hybrid thin lms based on methacrylates or thermosets reinforced with metal oxides
obtained from different precursors via solgel processes have been recently developed [296298],
and their mechanical properties have also been characterized by DSI. The results published on the
nanomechanical properties of different inorganic hybrid nanocomposites are summarized in Table 6.
Very large quasi-static E improvements, up to 16-fold increase at 10 wt% loading, were reported for
PA-6 nanocomposites incorporating homogenously dispersed polar trisilanolphenyl (Tsp)-POSS [292]
(Table 6). These results, however, should be contemplated with caution due to the extremely low
reduced modulus obtained for the neat PA-6 (0.137 GPa) as compared with typical values that are
in the range 13 GPa (Table 4) [223,225,226]. Dynamic storage modulus data provided by the same
authors were E0 = 1.34 GPa for PA-6 and DE0 = 130% for the nanocomposite with a 10 wt% Tsp-POSS
content. The increment was ascribed to the strong stiffening effect of the robust POSS cages, their
homogenous dispersion within the matrix and the efcient role played as nucleating agents. Interest-
ingly, Tsp-POSS yielded greater improvements in mechanical properties than non-polar octaisobutyl
(Oib)-POSS, since the hydroxyl groups of the former nanoller allowed hydrogen bonding interactions
with the amide linkages of PA6. On the other hand, only 3040% values were observed for DE0DMA deter-
mined using conventional DMA tests, probably due to the preferential location of the llers on the
nanocomposite surface. This morphology could also explain the observed decrease in E0 with increas-
ing indentation depth below 300 nm. Further, this ISE effect could also be related to the different cool-
ing rate between the nanocomposite bulk and surface during the melt-blending process. In fact, the
fastest cooling at the surface should lead to restrained chain mobility and, hence, to a higher modulus.
Strong E improvements have also been observed in PMMA nanocomposites lled with 3.0 wt%
graphene-like BN [288], showing a maximum increase of 130% for samples prepared using chloroform
as solvent. Moreover, Mammeri et al. found noticeable E enhancements in hybrid PMMA-SiO2 thin
lms prepared by solgel processes and exhibiting different interfaces between the polymer and
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 77

the silica network [296,297]. In order to extract the mechanical properties, the authors estimated the
creep rate of the material that was used to correct the contact stiffness following Tang and Ngan equa-
tion [299]. It was found that the viscoelastic effects inuenced the modulus to an extent lower than
10%. Hybrid materials, where the organic and inorganic components were linked through covalent
chemical bonds (class II) [296], showed better mechanical properties than those where the two com-
ponents exchange only weak interactions (class I) [297]. Class I and II materials were synthesized by
in situ polymerization of TEOS with PMMA monomers, which were free of or functionalized with tri-
ethoxysilane groups, respectively. Hybrid materials of class II were, in a nal step, reinforced with SiO2
nanoparticles [297]. It was concluded that the mechanical response was mainly determined by the
size of the hybrid interface, since the lowest mechanical properties corresponded to materials formed
from silica nanoparticles which exhibited a more dened interface.
On the other hand, moderate DE values (30%) have been reported for epoxy composites incorporat-
ing 3.0 wt% IF-WS2 nanotubes [289], being superior to those found for equivalent SWCNT-reinforced
nanocomposites. This result is highly interesting since the SWCNTs possess higher strength and elastic
properties than the IF-WS2, and was attributed to the improved dispersion of the inorganic llers
within the matrix. In addition, the IF-WS2 nanotubes (70 nm diameter) are expected to restrict the
movement of advancing cracks during the deformation process, thus enhancing the composite
strength and toughness, while the SWCNTs (12 nm in diameter) are much smaller than the crack-
opening displacement and, hence, they are unlikely to cause crack pinning. Chen et al. [290], dealing
with PU composite coatings reinforced with GPTS-modied AlOOH nanorods or nanoparticles only
found small modulus increments compared to that of the neat resin, despite the formation of a strong
ller-matrix interface in the presence of the compatibilizing agent. Further, E values for nanorod-lled
coatings were slightly lower than the nanoparticle-lled counterparts, ascribed to the different ller
orientation in the composites since nanorods were horizontally aligned and thus offered lower resis-
tance to the normal force applied during the indentation tests, while the isotropically distributed
nanoparticles were more suitable to bear the load under compression. Interestingly, different defor-
mation morphologies were observed in the indents made on these samples (Fig. 25). Large cracks,
located along or at an angle to the indenters edges, were detected in the nanoparticle-reinforced coat-
ing (Fig. 25a), while the nanorod lled counterpart exhibited no sign of chipping (Fig. 25)b, and only
small radial cracks were found between contact edges together with minuscule cracks perpendicular
to the radial ones (see the arrows marked on the image). Such features could be due to the release of
pressure in the lm as the indenter was removed, which allowed relaxation of the deformation caused
by buckling in the coating layer under indentation and should lead to signicantly higher fracture
toughness in the nanorod-lled coating. Slight stiffness improvements were also reported for other
amorphous polymers like PMMA upon incorporation of Eu:Gd2O3 nanoparticles (Table 6); the modi-
cation of the llers with a silane coupling agent enhanced their dispersion within the polymer,
resulting in somewhat increased modulus, in good agreement with data derived from DMA tests [300].

Fig. 25. Indents made of PU-based coating samples lled with GPTS-modied AlOOH nanoparticles (a) and nanorods (b).
Adapted from Ref. [290], copyright 2009, with permission from Elsevier.
78 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

Contrary to the aforementioned results, some authors have reported a diminution in stiffness upon
addition of inorganic llers. For example, for PVDF/uoropropyl (FP)-POSS composites, it was shown
that small contents of ller caused an incomplete nucleation of PVDF and generated big spherical par-
ticles that slightly reinforced the matrix, while the incorporation of higher concentrations provoked a
complete nucleation and formed separated and low-crosslinked particles that behaved as soft llers,
reducing the matrix modulus [293]. DSI results conrmed that DE was slightly positive (6%) for a
3 wt% loading while at higher loadings, 5.0 and 8.0 wt%, the modulus decrease was about 11 and
15%, respectively [293]. Tensile data followed similar trend to that found from nanoindentation. In
parallel, however, DEDMA showed increases for all FP-POSS loadings although after a 27% increase at
a ller content of 3%, there was a decreasing tendency. DMA results also revealed that the viscosity
of the nanocomposites was much higher than that of the neat polymer and, together with the
observed increasing fracture toughness, made these materials to favour plastic ow and become more
efcient on energy dissipation with increasing addition of FP-POSS.
Analogously, Huang et al. [295] found a decrease in E and H of neat PI with the addition of epoxide-
modied POSS, despite the signicant increase in crosslinking density due to the reaction between ter-
minal amine groups of the matrix and epoxy groups of POSS. This behaviour was attributed to the for-
mation of a soft interphase around the POSS molecules in the nanocomposites due to the presence of
eight exible aliphatic substituents attached to the OH cage. Pelisser et al. [291], in the search of Port-
land cement-based materials, studied PDDA/CSH nanocomposites and reported a decrease in E and H
of around 80%, with respect to the CSH values, due to the intercalation of the polymer between the
CSH lamellae that resulted in lower nanoparticle packing density.

5.4. Main features on inorganic nanollers

5.4.1. Layered silicates


DSI experiments on polymer/layered silicates suggest that the largest mechanical enhancements
are associated to nanocomposites with semicrystalline thermoplastic matrices. In contrast to organic
nanollers, nanoclays hinder the development of crystallinity and in spite of this EI and H increase
with increasing ller load. Matrix crystallinity can also change due to temperature gradient effects
during processing. In addition, local orientation of the nanoller and/or the matrix can be generated
giving rise to mechanical anisotropy effects that are readily detected by nanoindentation. Other fac-
tors such as the modication of the clay and the interfacial interactions between the polymer matrix
and the layered silicates produce a strong effect on the degree of dispersion of the ller and on the
structure of the nal layered structures, being the exfoliated ones those exhibiting the best perfor-
mance as measured by DSI.

5.4.2. Spherical inorganic nanoparticles


Research efforts on DSI results concerning inorganic spherical nanoparticles are mainly devoted to
silica reinforced composites. In most cases a continuous increase of the mechanical properties with
the nanosilica content is observed, in spite of the typical high loadings used. Several authors have
compared the experimental results of the modulus with values obtained from different models
[7578]. The lower bound of the HashinShtrikman model [73] and the HalpinTsai equations [74]
seem to offer the best approximations. Polyacrylate matrices reinforced with silica seem to yield
the most remarkable properties [210,270] (see Fig. 21), taking into account that the highest mechan-
ical reinforcement found for PVA may be an issue for further assessment due to the low modulus of the
matrix [78]. Collagen with HA nanollers also yields remarkable mechanical properties (see Fig. 21).

6. Comparison of the reinforcement effect of different nanollers

6.1. Effect of the type of nanoller

Figs. 2628 compare the reinforcing effect of different carbon and inorganic nanollers on the DSI
modulus of epoxy, glassy and semicrystalline matrices, respectively. To provide a clearer comparison,
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 79

100 MWCNT [117]


MWCNT [112]
MWCNT [111]
80 MWCNT [91]
MWCNT [91]
MWCNT [108]
SWCNT [24]
60
SWCNT [57]
GNP [168]
E [%]

GP [166]
40 ND [199]
Cl 93 A [247]
A1 100 [248]
20 A1 200 [248]
fumed SiO2 [263]
SiO2 [76]
0 SiO2 [75]

-20
0 2 4 6 8 10
nanofiller [wt%]

Fig. 26. Comparison of the reinforcing effect of different carbon and inorganic nanollers on the DSI modulus of thermoset
matrices. Data for composites incorporating the same ller have been represented using identical colour. Open and solid
symbols refer to functionalized and raw llers respectively. indicates nanoller grafted to the polymer matrix and j denotes
aligned nanoller.

140

120

100
PMMA/MWCNT [28]
PMMA/MWCNT [28]
80 PMMA/FG [21]
E [%]

PS/CNF [186]
60 PMMA/CNC [195]
PS/Cl 15A [231]
PS/Cl 20A [232]
40 PMMA/SiO2 [261]

20

-20
0 2 4 6 8 10
nanofiller [wt%]

Fig. 27. Comparison of DEI values for glassy matrices incorporating different nanollers. Colour code as in Fig. 26. Open symbols
refer to functionalized llers.  refers to CNTs functionalized and additionally coated with silica; + denotes electrospun bres.

Fig. 26 only includes the most representative data regarding epoxy/CNT composites, covering the
whole range of modulus increments reported. The gure shows that limited EI improvements are
observed at all nanoller contents, most of them lying in the range 040%. Especially low are the
DEI values at large ller loadings (above 5 wt%). Analogous moderate improvements, or even decre-
ments, are observed for the hardness values (see Tables 16). Surprisingly, the alignment of nanollers
with high aspect ratio such as CNTs in the direction of the loading [115] or the grafting of crystalline
nanoparticles to the polymer matrix [199] do not result in higher modulus improvements compared
to their random or non-grafted lled epoxy counterparts respectively. These limited mechanical
enhancements have been frequently associated, as discussed in preceding sections, to a poor disper-
sion of the nanoller including aggregation into micro or submicro-domains and/or inhomogeneous
distribution of the blocks, together with weak matrixller interaction. These features are commonly
80 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

200 PP/MWCNT [106]


PA-6/MWCNT [100]
PA-6/MWCNT [20]
PVA/MWCNT [103]
150 PLLA/MWCNT [101]
PP/GO [154]
PVA/GNP [160]
PVA/FG [21]
E [%]

100 PVA/ND [202]


PLLA/ND [187]
PP/PP-g-MA/Cl 15A [218]
PP/PP-g-MA/Cl 15A [218]
50 PA-6/I30TC [223]
PA-6/1.34TCN [226]
PLA/Cl 30 B [235]
PP/IF-WS2 [22]
0

0.0 2.5 5.0 7.5 10.0


nanofiller [wt%]

Fig. 28. Effect of carbon and inorganic nanollers on EI of different semicrystalline thermoplastic matrices. Colour code as in
Fig. 26. Solid symbols designate raw llers. The rest of symbols refer to functionalized llers.  and  denote composites with
OMMT prepared by SSE and TSE, respectively.

encountered in all types of lled epoxy materials and possibly arise as a consequence of an inadequate
dispersion of the ller in the blend and the latter interfering in the curing process of the epoxy matrix.
Some authors have suggested that the nanoller inuences the cross-linking density in the epoxy
[197,270], especially in the surroundings of the matrix-ller interface [247], hence, affecting the stiff-
ness and hardness of the matrix. From the above considerations it seems that the intrinsic properties
of the ller are not fully exploited and this turns up in similar matrix reinforcements for llers that
signicantly differ in mechanical properties. For example, the addition of 2D graphene platelets
[166,168] results in analogous modulus enhancements compared to the incorporation of similar
amounts of layered silicates [247,248] (see Fig. 26). It can also be observed that the largest mechanical
increments are reported for CNT-reinforced epoxies. However, it is not clear whether this observation
is related to the development of improved strategies for ller dispersion and llermatrix interaction.
In fact, the largest mechanical increment in Fig. 26 (DEI = 100% for 5 wt%) has been explained as aris-
ing from a minimum interference of the ller with the curing process of the epoxy through an ade-
quate selection of the solvent employed for CNT pre-dispersion [111].
Fig. 27, in turn, gathers DEI values of glassy matrices incorporating different nanollers. Repre-
sented data are restricted to PMMA and PS because these two matrices are the only examples in
the literature offering indentation data on nanocomposites using more than one type of ller. In case
of PS, the addition of clays or carbon nanobres produces small increments of the mechanical proper-
ties possibly due to agglomeration of the ller [198,243,244]. Most interesting are the data on PMMA
where a wider collection of nanollers can be compared. As shown in Fig. 27, few-layer graphene and
silica-coated CNT seem to be most effective. In both cases, the nanoller is functionalized to improve
the ller-matrix interaction. Compared to CNTs, graphene exhibits a large surface area that promotes
the interaction between the oxide groups of the acid-functionalized graphene sheets and those of the
glassy matrix [21]. In fact, functionalized CNTs have been shown to produce quite limited reinforce-
ment on PMMA (see Fig. 27). It has been suggested that the curved morphology and inferior bending
properties of CNTs could account for the reduced reinforcement action [28]. Only when CNTs are
coated with silica shell a signicant reinforcing effect comparable to that of graphene is detected. Nev-
ertheless, it is noteworthy that lower amounts of FG are needed with respect to silica-coated CNT to
produce a similar reinforcement effect on PMMA.
Finally, Fig. 28 illustrates the effect of carbon and inorganic nanollers on EI of different semicrys-
talline thermoplastic matrices. Only those matrices using different types of ller are included in the
gure. It is noteworthy that functionalization of the ller is a common route for all nanocomposites
represented. In the glassy matrices of Fig. 27, as discussed above, the intrinsic properties of the ller
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 81

and the distribution and interaction with the matrix have been demonstrated to be the key features to
achieve enhanced properties. In the case of semicrystalline matrices, the nucleating effect of the ller
is additionally expected to contribute to the mechanical enhancement through the changes triggered
in the matrix nanostructure and via the nucleation of crystals at the interface. To what extent each
aspect contributes to the overall mechanical enhancement is usually not approached in most inden-
tation studies, however, some valid analyses have been done in this direction [22].
In spite of the difculties in comparing indentation data from different sources, clear trends can be
found in Fig. 28. First, one can see that DEI conspicuously increases with increasing ller content up to
1 wt% where the rate of increase diminishes or levels off. It is also clearly seen that carbon-based ll-
ers such as CNTs, graphene and nanodiamond tend to lie in the upper part of the plot, especially for
ller contents above 1 wt%; hereafter, D EI still rises at a signicant rate [20,100,101,160,187]. In con-
trast, data for clays and inorganic particles are mainly located in the lower part of the plot and exhibit
a tendency to remain constant above 1 wt% [22,218,235] although there are also examples where EI
further increases at higher loadings [223,226].
All carbon-based llers seem to act as reinforcement to approximately the same extent within the
limitations inherent to the comparison of data from different sources. For instance, DEI values for PVA
reinforced with CNTs, graphene-like llers and ND lie within the same range [21,103,160,202]. More-
over, PA6, PLLA and PP provide examples in which the reinforcing effect of carbon-based llers and
inorganic particles can be clearly discerned. In case of PA6, the reinforcing effect of MWCNT is signif-
icantly higher at all ller loadings than that of clays (see Fig. 28 [20,100,223,226]). Moreover, incorpo-
ration of MWCNT or ND to PLLA produces a marked reinforcement compared to the one originated in
PLA with the addition of clay [101,187,235]. For PP, higher enhancements are achieved when graphene
oxide is used with respect to organically modied MMT and inorganic fullerene-like nanoparticles
[22,154,218]. Data for PP/MWCNT composites exhibit too large scatter (see Fig. 28) and hence, have
been not taken into account in the discussion. A detailed study on PP/IF-WS2 composites has shown
that enhancements associated to the ller itself contribute more signicantly to the overall reinforce-
ment than those related to changes induced in the PP morphology. The latter changes, however,
account for the initial drop of mechanical properties at the lowest ller content [22].

6.2. Inuence of the geometry of the ller and orientation effects

The aspect ratio of the nanoller itself does not seem to introduce an additional value to the
mechanical properties of reinforced polymers by means of indentation. This appears to be quite rea-
sonable for polymer matrices incorporating randomly distributed nanollers taking into account that
the applied stress during an indentation test evolves in a triaxial manner. Anisotropy of the llers is
not exploited because the force is not applied along the direction of higher stiffness. Indeed, there
is no evidence that carbon-based llers with different geometries such as ND, CNTs and graphene
exhibit different reinforcing effects that can be attributed to the geometry when they are randomly
distributed in a polymer matrix. On the contrary, in case of CNTs, it has been argued that the cylindri-
cal shape can be detrimental for the indentation properties due to the curved morphology and inferior
bending properties of the nanotubes [28]. However, the anisotropic shape of the ller can be used to
enhance the llermatrix interaction. For example, the large surface area of graphene sheets can be
used to create a large number of active sites to promote the interaction with the matrix [21].
In the case that the anisotropic ller exhibits some preferential orientation in the composite mate-
rial, mechanical properties are expected to be dependent on the relative orientation of the composite
with respect to the indentation loading direction. The application of the load perpendicularly to the
direction of highest stiffness in whiskers of cellulose nanocrystals has been proposed as one of the pos-
sible reasons for the minor mechanical enhancements encountered in CNC-reinforced PMMA bres
[207]. Most strikingly are the indentation results reported in epoxy composites including CNT for-
ests where application of the load along the vertical direction of the aligned CNTs only revealed mod-
est modulus improvements [115117]. Again, CNT waviness was suggested as the major factor
accounting for the limited stiffness enhancements.
Graphene represents the best example where mechanical anisotropy seems to be fully exploited.
The planar orientation of graphene sheets in a composite material has been proved to be a most
82 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

successful route in providing enhanced mechanical properties [158,173]. Indentation studies on com-
posites of self-aligned ultralarge-size rGO sheets and PU, and graphene/polymer multilayer assemblies
have demonstrated that modulus and hardness values can signicantly benet from the planar orien-
tation of graphene. However and as several times stated throughout this review, the reader should be
aware that such achievements could be questionable due to the very low modulus values reported for
the neat PU matrix.

6.3. Comparison of properties at the macro and the nanoscale

Fig. 29 illustrates the plot of the modulus values obtained by means of indentation, EI (E or E0 ), as a
function of those determined by means of macroscopic techniques, Emacro (Y or E0 DMA). A colour code
has been used to distinguish between dynamic DSI or quasi-static DSI data vs. macroscopic tensile
or DMA measurements. Some correlation between EI and Emacro can be distinguished that allows estab-
lishing clear limits to the EI/Emacro ratios most commonly encountered. This observation is purely phe-
nomenological, as extensively discussed in Section 2.3. The gure shows that most data lie within the
area dened by two dotted lines, EI = 2Emacro and EI = Emacro/2, that are drawn as a guide to the eye in
the gure. Only two series of EI  Emacro data lie outside this area. One of the series is associated to PS-
based shape memory polymers reinforced with carbon black [206], exhibiting very low E0DMA values, in
fact the lowest of all the Emacro data included in Fig. 29. As discussed in Section 4, this can produce rel-
evant consequences in the determination of the point of initial contact during an indentation test
introducing important inaccuracies in the indentation data. On the other hand, it is difcult to ascer-
tain the possible reason for the signicant deviation of the series of PAN/ND data from the general
trend because only limited information on the indentation experimental procedure and method of
analysis is available [208].
It can also be observed from Fig. 29 that the correspondence between E from quasi-static indenta-
tion and Y from tensile testing (blue symbols) is fairly good, most of the data lying in the line EI = Emacro
(also drawn in the gure). All the nanocomposites in this group of data include randomly distributed
nanollers except the one constituted by self-aligned ultralarge-size rGO layers incorporated into PU.
In the latter case, it is interesting to note that the modulus of indentation achieved with the graphene
sheets perpendicular to the load direction seems to be in agreement with tensile testing data obtained

Epoxy/MWCNT [91]
Epoxy/MWCNT [110]
Epoxy/MWCNT [112]
Epoxy/SWCNT [24]
Chitosan/MWCNT [98]
PA-6/MWCNT [100]
8 PA-6/MWCNT [20]
UHMWPE/MWCNT [105]
Epoxy/GP [166]
Epoxy/GNP [169]
Pu/rGO [158]
6 Epoxy/CNF [197]
Epoxy/CNF [198]
PS/CNF [198]
E [GPa]

PVA/CNC [200]
PS/CB [206]
4 PAN/ND [208]
PA-6/I30TC [223]
PA-6/CoAl-LHD [224]
PA-6/1.34TCN [226]
PLA/Cl 30B [235]
2 PHBV/Cl 30A [237]
Epoxy/Cl 93A [247]
Epoxy/fumed SiO 2[263]
Epoxy/SiO 2 [75]
PMMA/SiO 2 [273]
0 PP/IF-WS 2 [22]
0 2 4 6 PPS/IF-WS 2 [22]

Emacro [GPa]

Fig. 29. DSI moduli EI (E or E0 ) vs. values determined by means of macroscopic techniques Emacro (Y or E0DMA ) for polymer
nanocomposites reinforced with different nanollers. Open and solid symbols refer to functionalized and raw llers
respectively. + as in Fig. 11. The dashed lines represent EI = 2Emacro, and EI = Emacro/2, while the dotted-dashed line corresponds to
EI = Emacro. Symbols in blue, green, magenta and black correspond to E vs. Y, E vs. E0DMA , E0 vs. Y and E0 vs. E0 DMA, respectively.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 83

directing the load along the graphene plane [158]. The majority of the EI data in blue colour have been
analysed using Oliver and Pharrs method [100,110,112,223,226,247,263]. Only one paper uses a dif-
ferent method of analysis consisting in applying Hertz to the loading curve [200]. It seems that the
latter procedure yields somehow higher EI/Emacro values approaching 2. According to the available
data, it is then suggested that OP analysis offers modulus values in the range of those determined from
tensile testing. This contention is most probably mediated by the fact that all the polymer matrices
included in this group of data (epoxies, PA6, PVA, PAN) exhibit E > 1 GPa and a limited viscous char-
acter. Indeed, a recent paper examined the validity of Oliver and Pharrs analysis to polymer materials
and suggested that this method can yield consistent results for polymers with glass transition temper-
ature well above the temperature of measurement [301].
Fig. 29 additionally demonstrates that in spite of the large scatter detected in the plot of E0 vs. Y
(magenta symbols), data are found to uctuate around the line dened by EI = Emacro. This is quite sur-
prising taking into account the technical dissimilarities between dynamic indentation and quasi-static
tensile testing. The reason for this behaviour is probably grounded on a similar argument to that pre-
viously discussed: the limited rubbery ow character at room temperature of most polymers matrices
represented. Accordingly, plotting E0 vs. Y or E vs. Y should look quite similar.
At the present stage, no clear correspondence is found between modulus from indentation and
those of DMA. The large scatter of data in the representation of E0 vs. E0DMA (black symbols in
Fig. 29) does not allow drawing any signicant tendency other than that the EI/Emacro ratio can take
any value between 0.5 and 2. On the other hand, modulus values from quasi-static DSI vs. those of
DMA (green symbols) tend to lie along the line EI = 2Emacro. However, the amount of data is so limited
that this contention should be only taken as a hint for further research.

7. Conclusions and future perspectives

An effort has been done to gather indentation data on polymer nanocomposites. This is quite a
challenging task. On the one hand, DSI devices are too often employed as routine instrumentation
without a basic knowledge of the physics underlying. Moreover, the methods developed for the deter-
mination of the mechanical properties from DSI curves rely on a number of assumptions that are often
overlooked. As a consequence, large errors in the determination of the modulus and hardness can be
introduced. This has been known for over more than two decades and yet, too many published papers
report values for mechanical properties that at rst sight are recognized to be incorrect. On the other
hand, we have frequently observed that many authors offer little experimental information, in such a
way that an adequate evaluation of the relevance of the results is not feasible. It is noteworthy that the
most outstanding reinforcing effects are commonly related to matrices with very low modulus values.
This observation applies to quite different llers and polymer matrices, e.g., carbon nanotubes and
PHO (E = 0.12 GPa) [97] or UHMWPE (E = 0.6 GPa) [104]; graphene-like llers and PU (E = 0.4 GPa)
[158] or PVA (E = 0.2 GPa) [160]; a combination of different carbon llers and PVA (E = 0.7 GPa)
[121]; clays and PU (E = 0.01 GPa) [252]; spherical nanoparticles such as nanoHA and collagen
(E0 = 0.20 GPa) [286] or silica and PVA (E = 0.30 GPa) [78]. Results on low modulus materials need a care-
ful consideration, in view of the well-recognized technical difculties. Some useful recommendations
are given in Refs. [302,303].
In spite of all the limitations outlined, we think that the following general conclusions can be
drawn from the present review.
Indentation data on epoxies clearly reveal quite limited modulus and hardness increments with
respect to thermoplastic materials. Functionalization does not seem to play here a relevant role. It
has been suggested that the limited nanoller dispersion in the blend, together with the interference
of the ller with the curing chemistry are key parameters affecting the nal properties. In thermoplas-
tic materials, conspicuous mechanical property enhancements with increasing ller content followed
by a levelling-off at loadings of 12% have been observed as a general performance for most nanol-
lers. The state of distribution and dispersion of the ller is usually the basis of this behaviour. At this
point, a signicant difference is found between raw and functionalized llers, the latter usually exhib-
iting larger enhancements and increasing their positive difference with the raw ones at high loadings.
84 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

In case of semicrystalline thermoplastic polymers, the inuence of the ller on the nanostructure of
the matrix has also been shown to be of great importance for a full understanding of the reinforce-
ment. Creep results, that can be easily evaluated using depth-sensing instrumentation, provide an
additional source of information on the time-dependent mechanical behaviour of nanocomposites.
Most indentation studies show an enhanced creep resistance with the incorporation of the ller.
Hitherto, the shape of the nanoller does not seem to play a clear role on mechanical improvement,
as measured by means of DSI. However, it should be noted that a limited amount of data are available.
One should recall that the compressive stress eld developed underneath the indenter is approxi-
mately radial from the point of rst contact. This setup does not favour orientation of anisotropic
nanollers in the direction that yields their highest stiffness, as could be the case in uniaxial tensile
testing, thus limiting the potential benets of nanoller anisotropy. Nevertheless, in some cases the
anisotropic shape of the ller promotes mechanical reinforcement because an enlarged surface area
can accommodate a higher number of interaction sites with the matrix.
Most important, nanoindentation provides worthy information on the local mechanical properties
at the submicroscale. Uneven distribution of the ller or the matrix morphology can be readily
detected by means of DSI [111,223,229,262]. At shallow depths (below 200300 nm), a large number
of studies report remarkable indentation size effects that are commonly discussed in the light of
instrumental issues such as: area calibration, tip blunting, frictional forces between the indenter
and sample or determination of the point of initial contact [20,98,99,126,169,174]. For indentation
depths exceeding h  200 nm, the continuous decrease of modulus and/or hardness values has been
discussed as a consequence of a morphology gradient across the sample thickness [111,229]. On the
other hand, the constancy of EI and H values over a broad range of indentation depths has been inter-
preted as indicative of the good dispersion of the ller throughout the matrix [114]. For nanollers
such as graphene where at least one dimension approaches a few micrometers, DSI can clearly mon-
itor the continuous change of the mechanical properties as the indenter penetrates into the surface
and nally converge towards the mean value of the material at large penetration depths [166].
Finally, a phenomenological correspondence has been found between indentation modulus (either
quasi-static or dynamic) and Youngs modulus values from tensile testing. Such a fair correlation is
somewhat unexpected due to evident experimental differences (e.g., loading direction, scale of volume
deformed, occurrence of material heterogeneities). A larger data scattering is observed when compar-
ing EI with DMA measurements. In this case further research is needed taking into account the limited
data available.
What should we expect for the next decade? It is clear that the eld of polymer nanocomposites
will continue growing with the incorporation of new nanollers and the development of complex
hybrid and hierarchical materials to be applied as thin lms, coatings, multilayers or bulk. The DSI
technique, although still in its infancy, has been shown to be an effective tool to characterize the
mechanical properties of polymer nanocomposites, particularly when there are limitations in size
or arrangement of the specimen nanocomposite, or when a local probe of the mechanical properties
is required. These features will permit to address the new challenges presented by innovative mate-
rials, especially from the point of view of understanding the nanocomposite heterogeneity, on the one
side, and the macroscopic mechanical properties on the other.
We have noticed that little research has been devoted to the correlation between nanoindentation
results and structural information that can be extracted by other techniques. Usually, DSI data of poly-
mer nanocomposites are correlated to microscopic studies but no further structural analysis is offered.
This is indeed an area that still needs to be developed, as it is the analysis of creep or, in general, the
time-dependent behaviour. The integration of nanoindentation with real-time techniques such as
electron imaging, X-ray scattering, electrical and thermal characterization among others, represents
one of the most attractive challenges for the coming years in polymer nanocomposites and materials
science [304]. The spread use of dynamic indentation, i.e. nanoDMA, in a wider range of frequencies
and temperatures should also be encouraged. We can anticipate that at short term, dynamic DSI will
become a convenient technique for recording information on the relaxation modes of polymer spec-
imens in the nanoscale range. Some research has already been done in this direction [305]. It is also
noteworthy that few studies employ indentation testing to assess the mechanical properties of inter-
phases. Improving the force and depth resolution of the current depth-sensing instrumentation will
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 85

indeed foster the research in this area. In parallel to the rising concern on environmental problems,
bionanocomposites are likely to be increasingly used and the investigation of their mechanical perfor-
mance in comparison with traditional non-biodegradable composites will certainly be of interest
[306]. Taking a step forward, the application of DSI to biological samples, that are hierarchical nano-
composites by nature, is also an area of great potential. Hard biological materials such as bone struc-
tures and dentine are easy to measure by indentation techniques because modulus values lie in the
GPa range [307309]. On the contrary, indentation measurements on soft polymer biomaterials and
tissues, with modulus values in the MPa or kPa range represent quite a challenge [310,311]. Nanoin-
dentation on soft materials has been frequently carried out by means of AFM because of the ease of
use of uid cells for experiments with hydrated specimens, together with superior displacement
and force sensitivity [312314]. Nevertheless, the analysis of AFM data faces severe difculties related
to detection of initial contact, force-depth calibration and tip geometry [315]. With respect to AFM,
DSI represents a robust method arriving from the submicron scale with well-established calibration
methods that provide reliable mechanical properties. In addition, some DSI instruments have been
claimed to account for the necessary stiffness sensitivity to properly measure the properties of mate-
rials as soft as 10 kPa in modulus [316]. Other commercial nanoindenters incorporate a force feedback
control that ensures an accurate measurement of the load (and not of the stiffness) as a means of
improving the surface detection. Initial DSI studies on hydrated biological samples, with modulus val-
ues in the kPa range, highlight the potential of the technique to the study of biological nanocomposites
[316,317]. The precise knowledge of the viscoelastic properties of biological systems in a controlled
test environment reproducing in vivo conditions can help, for instance, to understand the mechanisms
involved behind a pathological behaviour or to fabricate biological replacement parts [311]. To prop-
erly extract the true elastic modulus of very soft materials, such as biological systems, data analyses
capable of separating rst the inuence of external effects and then the non-linear viscous compo-
nents from the linear ones should be employed [318].
An effective use of DSI for soft biological materials will require exploiting both the full potential of
the new generation of nanoindenters with improved depth and force resolution and the most
advanced analysis routes.

Acknowledgments

The authors wish to thank the MICINN (Ministerio de Ciencia e Innovacin), Spain, for nancial
support under the Grants MAT2010-21070-C02-01 and FIS2010-18069. AD would like to thank the
CSIC for a JAE postdoctoral contract conanced by the EU.

References

[1] Pethica JB. Microhardness tests with penetration depths less than ion implanted layer thickness. In: Ashworth V, Grant W,
Proctor R, editors. Ion implantation into metals. Oxford: Pergamon Press; 1982. p. 14757.
[2] Fischer-Cripps AC. Nanoindentation. 2nd ed. New York: Springer; 2004.
[3] Doerner MF, Nix WD. A method for interpreting the data from depth-sensing indentation instruments. J Mater Res
1986;1(4):6019.
[4] Oliver WC, Pharr GM. An improved technique for determining hardness and elastic modulus using load displacement
sensing indentation experiments. J Mater Res 1992;7:156483.
[5] Pharr GM, Strader JH, Oliver WC. Critical issues in making small-depth mechanical property measurements by
nanoindentation with continuous stiffness measurement. J Mater Res 2009;24:65366.
[6] VanLandingham MR, Villarubia JS, Guthrie WF, Meyers GF. Nanoindentation of polymers: an overview. Macromol Symp
2001;167:1543.
[7] Fischer-Cripps AC. Critical review of analysis and interpretation of nanoindentation test data. Surf Coat Technol
2006;200:415365.
[8] Hochstetter G, Jimenez A, Loubet JL. Strain-rate effects on hardness of glassy polymers in the nanoscale range. Comparison
between quasi-static and continuous stiffness measurements. J Macromol Sci Phys 1999;38:68192.
[9] Turnbull A, White D. Nanoindentation and microindentation of weathered unplasticised poly-vinyl chloride (UPVC). J
Mater Sci 1996;31:418998.
[10] Flores A, Balt-Calleja FJ. Mechanical properties of poly(ethylene terephthalate) at the near surface form depth-sensing
experiments. Phil Mag 1998;78:128397.
[11] Beake BD, Leggett GJ. Nanoindentation and nanoscratch testing of uniaxially and biaxially drawn poly(ethylene
terephthalate) lm. Polymer 2002;43:31927.
86 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

[12] Tranchida D, Piccarolo S, Loos J, Alexeev A. Mechanical characterization of polymers on a nanometer scale through
nanoindentation. A study on pile-up and viscoelasticity. Macromolecules 2007;40:125967.
[13] Briscoe BJ, Fiori L, Pelillo E. Nano-indentation of polymeric surfaces. J Phys D: Appl Phys 1998;31:2395405.
[14] Pethica JB, Oliver WC. Mechanical properties of nanometer volumes of material: use of the elastic response of small area
indentations. In: Bravman JC, Nix WD, Barnett DM, Smith DA, editors. Thin lms: Stresses and mechanical properties.
Mater res soc symp proc, vol. 130. Pittsburgh (PA): Materials Research Society; 1989. p. 1323.
[15] Tranchida D, Piccarolo S, Soliman M. Nanoscale mechanical characterization of polymers by AFM nanoindentations:
critical approach to the elastic characterization. Macromolecules 2006;39:454756.
[16] Cohen SR, Kalfon-Cohen E. Dynamic nanoindentation by instrumented nanoindentation and force microscopy: a
comparative review. Beilstein J Nanotechnol 2013;4:81533.
[17] Manias E. Nanocomposites: stiffer by design. Nat Mater 2007;6:911.
[18] Kickelbick G. Concepts for the incorporation of inorganic building blocks into organic polymers on a nanoscale. Prog
Polym Sci 2003;28:83114.
[19] Rahman A, Ali I, Al Zahrani SM, Eleithy RH. A review of the applications of nanocarbon polymer composites. Nano
2011;6:185203.
[20] Liu T, Phang IY, Shen L, Chow SY, Zhang W-D. Morphology and mechanical properties of multiwalled carbon nanotubes
reinforced nylon-6 composites. Macromolecules 2004;37:721422.
[21] Das B, Prasad KE, Ramamurty U, Rao CNR. Nano-indentation studies on polymer matrix composites reinforced by few-
layer graphene. Nanotechnology 2009;20:12570510.
[22] Flores A, Naffakh M, Dez-Pascual AM, Ania F, Gmez-Fatou MA. Evaluating the reinforcement of inorganic fullerene-like
nanoparticles in thermoplastic matrices by depth-sensing indentation. J Phys Chem C 2013;117:2093643.
[23] Shen L, Phang IY, Chen L, Liu T, Zeng K. Nanoindentation and morphological studies on nylon-66 nanocomposites.
Polymer 2004;45:33419.
[24] Dutta AK, Penumadu D, Files B. Nanoindentation testing for evaluating modulus and hardness of single-walled carbon
nanotubereinforced epoxy composites. J Mater Res 2004;19:15864.
[25] Oliver WC, Pharr GM. Measurements of hardness and elastic modulus by instrumented indentation: advances in
understanding and renements to methodology. J Mater Res 2004;19:320.
[26] Martin M, Troyon M. Fundamental relations used in nanoindentation: critical examination on experimental
measurements. J Mater Res 2002;17:222734.
[27] Klapperich C, Komvopoulos K, Pruitt L. Nanomechanical properties of polymers determined from nanoindentation
experiments. J Tribol 2001;123:62431.
[28] Olek M, Kempa K, Jurga S, Giersig M. Nanomechanical properties of silica-coated multiwall carbon nanotubes-
poly(methyl methacrylate) composites. Langmuir 2005;21:314652.
[29] Hay JL, Pharr GM. Instrumented indentation testing. In: Kuhn H, Medlin D, editors. ASM handbook: mechanical testing
and evaluation, 10th ed., vol. 8. Ohio: ASM International, Materials Park; 2000. p. 23243.
[30] Feng G, Ngan AHW. Effects of creep and thermal drift on modulus measurement using depth-sensing indentation. J Mater
Res 2002;17:6608.
[31] Cheng L, Xia X, Yu W, Scriven LE, Gerberich WW. Flat-punch indentation of viscoelastic material. J Polym Sci B: Polym
Phys 2000;38:1022.
[32] Tweedie CA, Van Vliet KJ. Contact creep compliance of viscoelastic materials via nanoindentation. J Mater Res
2006;21:157689.
[33] VanLandingham MR, Chang N-K, Drzal PL, White CC, Chang S-H. Viscoelastic characterization of polymers using
instrumented indentation. I. Quasi-static testing. J Polym Sci B: Polym Phys 2005;43:1794811.
[34] Oyen ML. Spherical indentation creep following ramp loading. J Mater Res 2005;20:2094100.
[35] Fischer-Cripps AC. A simple phenomenological approach to nanoindentation creep. Mat Sci Eng A 2004;385:7482.
[36] Monclus MA, Jennett NM. In search of validated measurements of the properties of viscoelastic materials by indentation
with sharp indenters. Phil Mag 2011;91:130828.
[37] Zhang CY, Zhang YW, Zeng KY. Extracting the mechanical properties of a viscoelastic polymeric lm on a hard elastic
substrate. J Mater Res 2004;19:305361.
[38] Cheng L, Xia X, Scriven LE, Gerberich WW. Spherical-tip indentation of viscoelastic material. Mech Mater
2005;37:21326.
[39] Oyen ML. Analytical techniques for indentation of viscoelastic materials. Phil Mag 2006;86:562541.
[40] Oyen ML, Cook RF. Loaddisplacement behavior during sharp indentation of viscouselasticplastic materials. J Mater Res
2003;18:13950.
[41] Mencik J, He LH, Swain MV. Determination of viscoelasticplastic material parameters of biomaterials by instrumented
indentation. J Mech Behav Biomed Mater 2009;2:31825.
[42] Lucas BN, Oliver WC. Indentation power-law creep of high-purity indium. Metall Mat Trans A 1999;30:60110.
[43] Maier V, Durst K, Mueller J, Backes B, Hppel HW, Gken M. Nanoindentation strain-rate jump tests for determining the
local strain-rate sensitivity in nanocrystalline Ni and ultrane-grained Al. J Mater Res 2011;26:142130.
[44] Beake B. Modelling indentation creep of polymers: a phenomenological approach. J Phys D: Appl Phys 2006;39:447885.
[45] Berthoud P, GSell C, Hiver JM. Elasticplastic indentation creep of glassy poly(methyl methacrylate) and polystyrene:
characterization using uniaxial compression and indentation tests. J Phys D: Appl Phys 1999;32:292332.
[46] White CC, Vanlandingham MR, Drzal PL, Chang N-K, Chang S-H. Viscoelastic characterization of polymers using
instrumented indentation. II. Dynamic testing. J Polym Sci: B: Polym Phys 2005;43:181224.
[47] Herbert EG, Oliver WC, Pharr GM. Nanoindentation and the characterization of viscoelastic solids. J Phys D: Appl Phys
2008;41:740219.
[48] Odegard GM, Gates TS, Herring HM. Characterization of viscoelastic properties of polymeric materials through
nanoindentation. Exp Mech 2005;45:1306.
[49] Hay JL, Agee P, Herbert E. Continuous stiffness measurements during instrumented indentation testing. Exp Tech
2010;34(3):8694.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 87

[50] Wright WJ, Maloney AR, Nix WD. An improved analysis for viscoelastic damping in dynamic nanoindentation. Int J Surf
Sci Eng 2007;1(2/3):27492.
[51] Wright WJ, Nix WD. Storage and loss stiffnesses and moduli as determined by dynamic. J Mater Res 2009;24(3):86371.
[52] Gilman JJ. In: Randall MG, editor. Chemistry and physics of mechanical hardness. John Wiley & Sons; 2009. p. 20.
[53] Pharr GM, Herbert EG, Gao Y. The indentation size effect: a critical examination of experimental observations and
mechanistic interpretations. Annu Rev Mater Res 2010;40:27192.
[54] Chong ACM, Lam DCC. Strain gradient plasticity effect in indentation hardness of polymers. J Mater Res 1999;14:410310.
[55] Gao X-L. New expanding cavity model for indentation hardness including strain-hardening and indentation size effects. J
Mater Res 2006;21:131726.
[56] Lucas BN, Oliver WC, Pharr GM, Loubet JL. Time-dependent deformation during indentation testing. Mat Res Soc Symp
Proc 1997;436:2338.
[57] Li X, Gao H, Scrivens WA, Fei D, Xu X, Sutton MA, et al. Nanomechanical characterization of single-walled carbon
nanotube reinforced epoxy composites. Nanotechnology 2004;15:141623.
[58] Shen L, Phang IY, Liu T, Zeng K. Nanoindentation and morphological studies on nylon 66/organoclaynanocomposites. II.
Effect of strain rate. Polymer 2004;45:82219.
[59] King RB. Elastic analysis of some punch problems for a layered medium. Int J Solids Struct 1987;23(12):165764.
[60] Saha R, Nix WD. Effects of the substrate on the determination of thin lm mechanical properties by nanoindentation. Acta
Mater 2002;50(1):238.
[61] Gao H, Chiu C-H, Lee J. Elastic contact versus indentation modelling of multi-layered materials. Int J Solids Struct
1992;29(20):247192.
[62] Rar A, Song H, Pharr GM. Assessment of new relation for the elastic compliance of a lm-substrate system. MRS Proc
2001;695:L10.
[63] Menck J, Munz D, Quandt E, Weppelmann ER, Swain MV. Determination of elastic modulus of thin layers using
nanoindentation. J Mater Res 1997;12:247584.
[64] Bec S, Tonck A, Loubet JL. A simple guide to determine elastic properties of lms on substrate from nanoindentation
experiments. Phil Mag 2006;86(33):534758.
[65] Hay J, Crawford B. Measuring substrate-independent modulus of thin lms. J Mater Res 2011;26(6):72738.
[66] ISO 14577-4:2007. Metallic materials instrumented indentation test for hardness and materials parameters. Part 4: Test
method for metallic and non-metallic coatings.
[67] Han SM, Guyer EP, Nix WD. Extracting thin lm hardness of extremely compliant lms on stiff substrates. Thin Solid Films
2011;519:32214.
[68] Balt-Calleja FJ, Fakirov S. The microhardness of polymers. Cambridge: Cambridge University Press; 2000. p. 90-106.
[69] Flores A, Ania F, Balt-Calleja FJ. From the glassy state to ordered polymer structures: a microhardness study. Polymer
2009;50:72946.
[70] Flores A, Balt-Calleja FJ, Asano T. Creep behavior and elastic properties of annealed cold-drawn poly(ethylene
terephthalate): the role of the smectic structure as a precursor of crystallization. J Appl Phys 2001;90:600610.
[71] Deslandes Y, Alva Rosa E, Brisse F, Meneghini T. Correlation of microhardness and morphology of poly(ether-ether-
ketone) lms. J Mater Sci 1991;26:276977.
[72] Lakes RS. Viscoelastic materials. Cambridge: Cambridge University Press; 2009. p. 344353.
[73] Hashin Z, Shtrikman S. A variational approach to the theory of the elastic behaviour of multiphase materials. J Mech Phys
Solids 1963;11:12740.
[74] Halpin JC, Kardos JL. The HalpinTsai equations: a review. Polym Eng Sci 1976;16:34452.
[75] Zhang Y-F, Bai S-L, Li X-K, Zhang Z. Viscoelastic properties of nanosilica-lled epoxy composites investigated by dynamic
nanoindentation. J Polym Sci Polym Phys 2009;47:10308.
[76] Wang ZZ, Gu P, Zhang Z, Gu L, Xu YZ. Mechanical and tribological behavior of epoxy/silica nanocomposites at the micro/
nano scale. Tribol Lett 2011;42:18591.
[77] Lin Q, Cohen SA, Gignac L, Herbst B, Klaus D, Simonyi E, et al. Low dielectric constant nanocomposite thin lms based on
silica nanoparticle and organic thermosets. J Polym Sci B: Polym Phys 2007;45:148293.
[78] Bhattacharya M, Chaudhry S. High-performance silica nanoparticle reinforced poly (vinyl alcohol) as templates for
bioactive nanocomposites. Mater Sci Eng C 2013;33:260110.
[79] Herbert EG, Oliver WC, Lumsdaine A, Pharr GM. Measuring the constitutive behavior of viscoelastic solids in the time and
frequency domain using at punch nanoindentation. J Mater Res 2009;24:62637.
[80] Hayes SA, Goruppa AA, Jones FR. Dynamic nanoindentation as a tool for the examination of polymeric materials. J Mater
Res 2004;19(11):3298306.
[81] Stojanovic D, Orlovic A, Markovic S, Radmilovic V, Uskokovic PS, Aleksic R. Nanosilica/PMMA composites obtained by the
modicationof silica nanoparticles in a supercritical carbon dioxideethanol mixture. J Mater Sci 2009;44:622332.
[82] Guillonneau G, Kermouche G, Bec S, Loubet JL. Extraction of mechanical properties with second harmonic detection for
dynamic nanoindentation testing. Exp Mech 2012;52:93344.
[83] Iijima S. Helical microtubules of graphitic carbon. Nature 1991;354:568.
[84] Popov VN. Carbon nanotubes: properties and application. Mater Sci Eng R 2004;43:61102.
[85] Thostenson ET, Ren Z, Chou T-W. Advances in the science and technology of carbon nanotubes and their composites: a
review. Compos Sci Technol 2001;61:1899912.
[86] Dervishi E, Li Z, Xu Y, Saini V, Biris AR, Lupu D, et al. Carbon nanotubes: synthesis, properties, and applications. Particulate
Sci Technol 2009;27:10725.
[87] Dez-Pascual AM, Naffakh M, Gmez MA, Marco C, Ellis G, Martnez MT, et al. Development and characterization of PEEK/
carbon nanotube composites. Carbon 2009;47:307990.
[88] Gonzlez-Domnguez JM, Anson-Casaos A, Dez-Pascual AM, Ashra B, Naffakh M, Backman D, et al. Solvent-free
preparation of high-toughness epoxy-SWNT composite materials. ACS Appl Mater Interfaces 2011;3:144150.
[89] Ma PC, Siddiqui NA, Marom G, Kim JK. Dispersion and functionalization of carbon nanotubes for polymer-based
nanocomposites: a review. Composites Part A 2010;41:134567.
88 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

[90] Li YB, Wei BQ, Liang J, Yu Q, Wu DH. Transformation of carbon nanotubes to nanoparticles by ball milling process. Carbon
1999;37:4937.
[91] Chen Z, Dai XJ, Magniez K, Lamb PR, de Celis Leal DR, Fox BL, et al. Improving the mechanical properties of epoxy using
multiwalled carbon nanotubes functionalized by a novel plasma treatment. Composites Part A 2013;45:14552.
[92] Karousis N, Tagmatarchis N, Tasis D. Current progress on the chemical modication of carbon nanotubes. Chem Rev
2010;110:536697.
[93] Dez-Pascual AM, Naffakh M. Grafting of an aminated poly(phenylene sulphide) derivative to functionalized single-walled
carbon nanotubes. Carbon 2012;50:85768.
[94] Huang W, Lin Y, Taylor S, Gaillard J, Rao AM, Sun YP. Sonication-assisted functionalization and solubilization of carbon
nanotubes. Nano Lett 2002;2:2314.
[95] Chen Z, Dai XJ, Lamb PR, de Celis Leal DR, Fox BL, Chen Y, et al. Practical amine functionalization of multi-walled carbon
nanotubes for effective interfacial bonding. Plasma Process Polym 2012;9:73341.
[96] Samad MA, Sinha SK. Mechanical, thermal and tribological characterization of a UHMWPE lm reinforced with carbon
nanotubes coated on steel. Tribol Int 2011;44:193241.
[97] Yun SI, Gadd GE, Latella BA, Lo V, Russell RA, Holden PJ. Mechanical properties of biodegradable polyhydroxyalkanoates/
single wall carbon nanotube nanocomposite lms. Polym Bull 2008;61:26775.
[98] Wu T, Pan Y, Bao H, Li L. Preparation and properties of chitosan nanocomposite lms reinforced by poly(3,4-
ethylenedioxythiophene)-poly(styrenesulfonate) treated carbon nanotubes. Mater Chem Phys 2011;129:9328.
[99] Wang S-F, Shen L, Zhang W-D, Tong Y-J. Preparation and mechanical properties of chitosan/carbon nanotubes composites.
Biomacromolecules 2005;6:306772.
[100] Baji A, Mai Y-W, Wong S-C, Abtahi M, Dua X. Mechanical behavior of self-assembled carbon nanotube reinforced nylon
6,6 bers. Compos Sci Technol 2010;70:14019.
[101] Fung PY. Poly(L-Lactide)/multi walled carbon nanotube composites: Mechanical properties and interaction with
osteoblast-like cells in vitro. Ph.D. Thesis. Hong Kong: Polytechnic University; 2007.
[102] Olek M, Kempa K, Jurga S, Giersig M. Multiwall carbon nanotubes-based composites-mechanical characterization using
the nanoindentation technique. Int J Mater Res 2006;97:12358.
[103] Cadek M, Coleman JN, Barron V, Hedicke K, Blau WJ. Morphological and mechanical properties of carbon-nanotube-
reinforced semicrystalline and amorphous polymer composites. Appl Phys Lett 2002;81:51235.
[104] Sreekanth PSR, Kanagaraj S. Assessment of bulk and surface properties of medical grade UHMWPE based nanocomposites
using nanoindentation and microtensile testing. J Mech Behav Biomed Mater 2013;18:14051.
[105] Bakshi SR, Balani K, Laha T, Tercero J, Agarwal A. The nanomechanical and nanoscratchproperties of MWNT reinforced
ultrahigh-molecular-weight polyethylene coatings. J Miner Met Mater Soc 2007;59:503.
[106] Bhandari NL, Lach R, Grellmann W, Adhikari R. Depth-dependent indentation microhardness studies of different polymer
nanocomposites. Macromol Symp 2012;315:4451.
[107] Wang W, Xu L, Liu F, Li X, Xing L. Synthesis of isocyanate microcapsules and micromechanical behavior improvement of
microcapsule shells by oxygen plasma treated carbon nanotubes. J Mater Chem A 2013;1:77682.
[108] Dos Santos MN, Opelt CV, Lafratta FH, Lepienski CM, Pezzin SH, Coelho LAF. Thermal and mechanical properties of a
nanocomposite of a photocurable epoxy-acrylate resin and multiwalled carbon nanotubes. Mater Sci Eng A
2011;528:431824.
[109] Tehrani M, Safdari M, Al-Haik MS. Nanocharacterization of creep behavior of multiwall carbon nanotubes/epoxy
nanocomposite. Int J Plasticity 2011;27:887901.
[110] Ayatollahi MR, Doagou-Rad S, Shadlou S. Nano-/microscale investigation of tribological and mechanical properties of
epoxy/MWNT nanocomposites. Macromol Mater Eng 2012;297:689701.
[111] Mionic M, Jiguet S, Judelewicz M, Karimi A, Forro L, Magrez A. Study of the mechanical response of carbon nanotubes-SU8
composites by nanoindentation. Phys Status Solidi B 2010;247:30725.
[112] Her Z-C, Shiu T-Y. Carbon nanotubes reinforced nanocomposite lms by nanoindentation technique. Adv Sci Lett
2012;13:54750.
[113] Lagoudas DC, Thakre PR, Benzerga AA. Nanoindentation on CNT reinforced epoxy nanocomposites. In: Proceedings of the
16th European conference on fracture (ECF16). Alexandroupolis (Greece); 2006.
[114] Penumadu D, Dutta A, Pharr GM, Files B. Mechanical properties of blended single-wall carbon nanotube composites. J
Mater Res 2003;18:184953.
[115] Wang H, Feng J, Hu X, Ng KM. Tribological behaviors of aligned carbon nanotube/fullerene-epoxy nanocomposites. Polym
Eng Sci 2008;48(8):146775.
[116] Jiang L, Spearing SM, Monclus MA, Jennett NM. Formation and mechanical characterisation of SU8 composite lms
reinforced with horizontally aligned and high volume fraction CNTs. Compos Sci Technol 2011;71:13018.
[117] Cebeci H, Guzman de Villoria R, Hart AJ, Wardle BL. Multifunctional properties of high volume fraction aligned carbon
nanotube polymer composites with controlled morphology. Compos Sci Technol 2009;69:264956.
[118] Li X-F, Lau K-T, Yin Y-S. Mechanical properties of epoxy-based composites using coiled carbon nanotubes. Compos Sci
Technol 2008;68:287681.
[119] Qian H, Kalinka G, Chan KL, Kazarian SG, Greenhalgh ES, Bismarck A, et al. Mapping local microstructure and mechanical
performance around carbon nanotube grafted silica bres: methodologies for hierarchical composites. Nanoscale
2011;3:475967.
[120] Liu ZH, Pan CT, Lin LW, Lai HW. Piezoelectric properties of PVDF/MWCNT nanober using near-eld electrospinning. Sens
Actuators A 2013;193:1324.
[121] Prasad KE, Das B, Maitra U, Ramamurty U, Rao CNR. Extraordinary synergy in the mechanical properties of polymer
matrix composites reinforced with 2 nanocarbons. Proc Natl Acad Sci 2009;106:131869.
[122] Dez-Pascual AM, Gmez-Fatou MA, Ania F, Flores A. Nanoindentation assessment of the interphase in carbon nanotube-
based hierarchical composites. J Phys Chem C 2012;116:24193200.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 89

[123] Dez-Pascual AM, Ashra B, Naffakh M, Gonzlez-Domnguez JM, Johnston A, Simard B, et al. Inuence of carbon
nanotubes on the thermal, electrical and mechanical properties of poly(ether ether ketone)/glass ber laminates. Carbon
2011;49:281733.
[124] Wu C-L, Lin H-C, Hsu J-S, Yip M-C, Fang W. Static and dynamic mechanical properties of polydimethylsiloxane/carbon
nanotube nanocomposites. Thin Solid Films 2009;517:4895901.
[125] Nagy PM, Aranyi D, Horvth P, Ptschke P, Pegel S, Klmn E. Nanoindentation investigation of carbon nanotubepolymer
composites. Internet Electron J Mol Des 2006;5:13543.
[126] Satyanarayana N, Rajan KSS, Sinha SK, Shen L. Carbon nanotube reinforced polyimide thin-lm for high wear durability.
Tribol Lett 2007;27:1818.
[127] Mammeri F, Teyssandier J, Connan C, Le Bourhisb E, Chehimi MM. Mechanical properties of carbon nanotubePMMA
based hybrid coatings: the importance of surface chemistry. RSC Adv 2012;2:24628.
[128] Salvetat JP, Bonard JM, Thomson NB, Kulik AJ, Forr L, Benoit W, et al. Mechanical properties of carbon nanotubes. Appl
Phys A: Mater Sci Process 1999;69:25560.
[129] Bauhofer W, Kovacs JC. A review and analysis of electrical percolation in carbon nanotube polymer composites. Compos
Sci Technol 2009;69:148698.
[130] Lau KT, Lu M, Hui D. Coiled carbon nanotubes: synthesis and their potential applications in advanced composite
structures. Compos Part B 2006;37:43748.
[131] Lu H, Huang G, Wang B, Mamedov A, Gupta S. Characterization of the linear viscoelastic behavior of single-wall carbon
nanotube/polyelectrolyte multilayer nanocomposite lm using nanoindentation. Thin Solid Films 2006;500:197202.
[132] Huang G, Wang B, Lu H, Mamedov A, Gupta S. Material characterization and modeling of single-wall carbon nanotube/
polyelectrolyte multilayer nanocomposites. J Appl Mech 2006;73:73744.
[133] Firkowska I, Olek M, Pazos-Perez N, Rojas-Chapana J, Giersig M. Highly ordered MWNT-based matrixes: topography at
the nanoscale conceived for tissue engineering. Langmuir 2006;22:542734.
[134] Lee D, Cui T. Suspended carbon nanotube nanocomposite beams with a high mechanical strength via layer-by-layer
nano-self-assembly. Nanotechnology 2011;22:1656019.
[135] Pavoor P, Gearing BP, Gorga RE, Bellare A, Cohen RE. Engineering the friction-and-wear behavior of polyelectrolyte
multilayer nanoassembliesthrough block copolymer surface capping, metallic nanoparticles, and multiwall carbon
nanotubes. J Appl Polym Sci 2004;92:43948.
[136] Dez-Pascual AM, Naffakh M, Marco C, Gmez-Fatou MA, Ellis GJ. Multiscale ber-reinforced thermoplastic composites
incorporating carbon nanotubes: a review. Current Opinion Solid State Mater Sci. http://dx.doi.org/10.1016/
j.cossms.2013.06.003
[137] Geim AK, Novoselov KS. The rise of graphene. Nat Mater 2007;6:18391.
[138] Novoselov KS, Geim AK, Morozov SV, Jiang D, Zhang Y, Dubonos SV, et al. Electric eld in atomically thin carbon lms.
Science 2004;306:6669.
[139] Novoselov KS, Falko VI, Colombo L, Gellert PR, Schwab MG, Kim K. A roadmap for graphene. Nature 2012;490:192200.
[140] Balandin AA, Ghosh S, Bao W, Calizo I, Teweldebrhan D, Miao F, et al. Superior thermal conductivity of single-layer
graphene. Nano Lett 2008;8:9027.
[141] Kim H, Abdala AA, Macosko CW. Graphene/polymer nanocomposites. Macromolecules 2010;43:651530.
[142] Lee C, Wei X, Kysar JW, Hone J. Measurement of the elastic properties and intrinsic strength of monolayer graphene.
Science 2008;321:3858.
[143] Warner JH, Schaffel F, Rummeli M, Bachmatiuk A. Graphene: fundamentals and emergent applications. Oxford: Elsevier;
2013. p. 109 [chapter 3].
[144] Potts JR, Dreyer DR, Bielawski CW, Ruoff RS. Graphene-based polymer nancomposites. Polymer 2011;52:525.
[145] Huang X, Qi X, Boey F, Zhang H. Graphene-based composites. Chem Soc Rev 2012;41:66686.
[146] Das TK, Prusty S. Graphene-based polymer composites and their applications. Polym Plast Tech Eng 2013;52:31931.
[147] Salavagione HJ, Martnez G, Ellis G. Recent advances in the covalent modication of graphene with polymers. Macromol
Rapid Commun 2011;32(22):177189.
[148] Sutter P. Epitaxial graphene: how silicon leaves the scene. Nat Mater 2009;8:1712.
[149] Yuan GD, Zhang WJ, Yang Y, Tang YB, Li YQ, Wang JX, et al. Graphene sheets via microwave chemical vapor deposition.
Chem Phys Lett 2009;467:3614.
[150] McAllister MJ, Li J-L, Adamson DH, Schniepp HC, Abdala AA, Liu J, et al. Single sheet functionalized graphene by oxidation
and thermal expansion of graphite. Chem Mater 2007;19:4396404.
[151] Stankovich S, Dikin DA, Piner RD, Kohlhaas KA, Kleinhammes A, Jia YY, et al. Synthesis of graphene-based nanosheets via
chemical reduction of exfoliated graphite oxide. Carbon 2007;45:155865.
[152] Morales GM, Schifani P, Ellis G, Ballesteros C, Martnez G, Barbero C, et al. High-quality few layer graphene produced by
electrochemical intercalation and microwave-assisted expansion of graphite. Carbon 2011;49:280916.
[153] Kuilla T, Bhadra S, Yao D, Kim NH, Bose S, Lee JH. Recent advances in graphene based polymer composites. Prog Polym Sci
2010;35:135075.
[154] Shin K-Y, Hong J-Y, Lee S, Jang J. Evaluation of anti-scratch properties of graphene oxide/polypropylene nanocomposites. J
Mater Chem 2012;22:78719.
[155] Chakraborty H, Sinha A, Mukherjee N, Chattopadhyay PP. Exfoliated graphite reinforced PMMA composite: a study on
nanoindentation and scratch behavior. J Nanotechnology 2012:5. Article ID 940516.
[156] Fan H, Wang L, Zhao K, Li N, Shi Z, Ge Z, et al. Fabrication, mechanical properties, and biocompatibility of graphene-
reinforced chitosan composites. Biomacromolecules 2010;11:234551.
[157] Chen J, Guo X, Tang Q, Zhuang C, Liu J, Wu S, et al. Nanomechanical properties of graphene on poly(ethylene
terephthalate) substrate. Carbon 2013;55:14450.
[158] Youse N, Gudarzi MM, Zheng Q, Lin X, Shen X, Jia J, et al. Highly aligned, ultralarge-size reduced graphene oxide/
polyurethane nanocomposites: mechanical properties and moisture permeability. Compos A 2013;49:4250.
[159] Cai D, Jin J, Yusoh K, Raq R, Song M. High performance polyurethane/functionalized graphene nanocomposites with
improved mechanical and thermal properties. Compos Sci Technol 2012;72:7027.
90 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

[160] Lee S, Hong J-Y, Jang J. The effect of graphene nanoller on the crystallization behavior and mechanical properties of
poly(vinyl alcohol). Polym Int 2013;62:9018.
[161] Goncalves G, Marques PAAP, Barros-Timmons A, Bdkin I, Singh MK, Emamic N, et al. Graphene oxide modied with
PMMA via ATRP as a reinforcement ller. J Mater Chem 2010;20:992734.
[162] Salavagione HJ, Martnez G, Gmez MA. Polymeric modication of graphene through esterication of graphite oxide and
poly(vinyl alcohol). Macromolecules 2009;42:63314.
[163] Salavagione HJ, Martinez G, Ellis G. Graphene-based polymer nanocomposites. In: Mikhailov S, editor. Physics and
applications of graphene experiments, vol. 1. Rijeka (Croatia): Intech; 2010. p. 16992.
[164] Castelain M, Martnez G, Merino P, Martn-Gago JA, Segura JL, Ellis G, et al. Graphene functionalization with a conjugated
poly(uorene) via click coupling. Striking electronic properties in solution. Chem A Eur J 2012;18:496573.
[165] Castelan M, Martnez G, Ellis G, Salavagione HJ. Versatile chemical tool for the preparation of conductive graphene-based
polymer nanocomposites. Chem Comm 2013;49:89679.
[166] Chasiotis I, Chen Q, Odegard GM, Gates TS. Structureproperty relationships in polymer composites with micrometer and
submicrometer graphite platelets. Exp Mech 2005;45:50716.
[167] Odegard GM, Gates TS. Modeling and testing of the viscoelastic properties of a graphite nanoplatelet/epoxy composite. J
Intell Mater Syst Struct 2006;17:23946.
[168] Shokrieh MM, Hosseinkhani MR, Naimi-Jamal MR, Tourani H. Nanoindentation and nanoscratch investigations on
graphene-based nanocomposites. Polym Test 2013;32:4551.
[169] King JA, Klimek DR, Miskioglu I, Odegard GM. Mechanical properties of graphene nanoplatelet/epoxy composites. J Appl
Polym Sci 2013;128:421723.
[170] Zandiatashbar A, Picu CR, Koratkar N. Depth sensing indentation on nanoscale graphene platelets in nanocomposite thin
lms. Mater Res Soc Symp 2011;1312:37984.
[171] Martin-Gallego M, Hernndez M, Lorenzo V, Verdejo R, Lpez-Manchado MA, Sangermano M. Cationic photocured epoxy
nanocomposites lled with different carbon llers. Polymer 2012;53:18318.
[172] Neto AS, Lopes da Cruz DT, vila AF. Nano-modied adhesive by graphene: the single lap-joint case. Mater Res
2013;16:5926.
[173] Zhao X, Zhang Q, Hao Y, Li Y, Fang Y, Chen D. Alternate multilayer lms of poly(vinyl alcohol) and exfoliated graphene
oxide fabricated via a facial layer-by-layer assembly. Macromolecules 2010;43:94116.
[174] Wang N, Ji S, Zhang G, Li J, Wang L. Self-assembly of graphene oxide and polyelectrolyte complex nanohybrid membranes
for nanoltration and pervaporation. Chem Eng J 2012;213:31829.
[175] Gmez-Navarro C, Burghard M, Kern K. Elastic properties of chemically derived single graphene sheets. Nano Lett
2008;8:20459.
[176] Zandiatashbar A, Picu CR, Koratkar N. Control of epoxy creep using graphene. Small 2012;8:167682.
[177] Li J, Luo R, Yan Y. Effect of carbon nanobers on the inltration and thermal conductivity of carbon/carbon composites.
Mater Res Bull 2011;46:143742.
[178] Endo M, Koyama T, Hishiyama Y. Structure improvement of carbon bers prepared from benzene. Jpn J Appl Phys
1976;15:20736.
[179] Zhou Z, Lai C, Zhang L, Qian Y, Hou H, Reneker DH, et al. Development of carbon nanobers from aligned electrospun
polyacrylonitrile nanober bundles and characterization of their microstructural, electrical, and mechanical properties.
Polymer 2009;50:29993006.
[180] Ozkan T, Naraghi M, Polycarpou AA, Chasiotis I. Mechanical strength of pyrolytically stripped and functionalized heat
treated vapor grown carbon nanobers. In: 11th International congress and exhibition on experimental and applied
mechanics 2008, Orlando, FL; 2008. p. 110813.
[181] Xia W, Su D, Birkner A, Ruppel L, Wang Y, Wll C, et al. Chemical vapor deposition and synthesis on carbon nanobers:
sintering of ferrocene-derived supported iron nanoparticles and the catalytic growth of secondary carbon nanobers.
Chem Mater 2005;17:573742.
[182] Iakoubovskii K, Baidakova MV, Wouters BH, Stesmans A, Adriaenssens GJ, Vul AY, et al. Structure and defects of
detonation synthesis nanodiamond. Diamond Relat Mater 2000;9:8615.
[183] Khachatryan AK, Aloyan SG, May PW, Sargsyan R, Khachatryan VA, Baghdasaryan VS. Graphite-to-diamond
transformation induced by ultrasonic cavitation. Diamond Relat Mater 2008;17:9316.
[184] Kidalov SV, Shakhov FM, Ya A, Vul AY. Thermal conductivity of nanocomposites based on diamonds and nanodiamonds.
Diamond Relat Mater 2007;16:20636.
[185] Blank V, Popov M, Pivovarov G, Lvova N, Gogolinsky K, Reshetov V. Ultrahard and superhard phases of fullerite C60:
Comparison with diamond on hardness and wear. Diamond Relat Mater 1998;7:42731.
[186] Dubrovinskaia N, Dubrovinsky L, Crichton W, Langenhorst F, Richter A. Aggregated diamond nanorods, the densest and
least compressible form of carbon. Appl Phys Lett 2005;87:831069.
[187] Zhang Q, Mochalin VN, Neitzel I, Knoke IY, Han J, Klug CA, et al. Fluorescent PLLA-nanodiamond composites for bone
tissue engineering. Biomaterials 2011;32:8794.
[188] Murayama H, Tomonoh S, Alford JM, Karpuk ME. Fullerene production in tons and more: from science to industry.
Fullerenes Nanotubes Carbon Nanostruct 2004;12:19.
[189] Kroto W, Heath JR, OBrien SC, Curl RF, Smalley RE. C60: buckminsterfullerene. Nature 1985;318:1623.
[190] Donnet JB, Voet A. Carbon black physics, chemistry and elastomeric reinforcement. New York: Dekker; 1976.
[191] Barton RL, Keith JM, King JA. Development and modelling of electrically conductive carbon lled liquid crystalline
polymer composites for fuel cell bipolar plate applications. J New Mater Electrochem Syst 2007;10:2259.
[192] Khizhnyak PE, Chechetkin AV, Glybin AP. Thermal conductivity of carbon black. J Eng Phys 1979;37:10735.
[193] Seena V, Fernandes A, Pant P, Mukherji S, Rao VR. Polymer nanocomposite nanomechanical cantilever sensors: material
characterization, device development and application in explosive vapour detection. Nanotechnology
2011;22:29550112.
[194] Peng BL, Dhar N, Liu HL, Tam KC. Chemistry and applications of nanocrystalline cellulose and its derivatives: a
nanotechnology perspective. Can Journal Chem Eng 2001;89:1191206.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 91

[195] Berglund L. Cellulose-based nanocomposites. In: Mohanty AK, Misra M, Drzal L, editors. Natural bers, biopolymers and
biocomposites. Boca Raton, Florida: CRC Press; 2005. p. 80732.
[196] Habibi Y, Gofn A-L, Schiltz N, Duquesne E, Dubois P, Dufresne A. Bionanocomposites based on poly(e-caprolactone)-
grafted cellulose nanocrystals by ring-opening polymerization. J Mater Chem 2008;18:500210.
[197] Snchez M, Rams J, Campo M, Jmenez-Suarez A, Urea A. Characterization of carbon nanober/epoxy nanocomposites by
the nanoindentation technique. Composites Part B 2011;42:63844.
[198] Lee H, Mall S, He P, Shi D, Narasimhadevara S, Yun Y-H, et al. Characterization of carbon nanotube/nanober-reinforced
polymer composites using an instrumented indentation technique. Composites Part B 2007;38:5865.
[199] Neitzel I, Mochalin VN, Niu J, Cuadra J, Kontsos A, Palmese GR, et al. Maximizing Youngs modulus of aminated
nanodiamond-epoxy composites measured in compression. Polymer 2012;53:596571.
[200] Pakzad A, Simonsen J, Yassar RS. Elastic properties of thin poly(vinylalcohol)-cellulose nanocrystal membranes.
Nanotechnology 2012;23:8570617.
[201] Lee J, Deng Y. Nanoindentation study of individual cellulose nanowhisker-reinforced PVA electrospun ber. Polym Bull
2013;70:120519.
[202] Maitra U, Prasad KE, Ramamurtyb U, Rao CNR. Mechanical properties of nanodiamond-reinforced polymermatrix
composites. Sol State Commun 2009;149:16937.
[203] Kaboorani A, Riedl B, Blanchet P, Fellin M, Hosseinaei O, Wang S. Nanocrystalline cellulose (NCC): a renewable nano-
material for polyvinyl acetate (PVA) adhesive. Eur Polym J 2012;48:182937.
[204] Behler KD, Stravato A, Mochalin V, Korneva G, Yushin G, Gogotsi Y. Nanodiamond-polymer composite bers and coatings.
ACS Nano 2009;3:3639.
[205] Stravato A, Knight R, Mochalin V, Picardi SC. HVOF-sprayed nylon-11 + nanodiamond composite coatings: production &
characterization. J Therm Spray Technol 2008;17:8127.
[206] Xu B, Zhang L, Pei YT, Luo JK, Tao SW, De Hosson JThM, et al. Electro-responsive polystyrene shape memory polymer
nanocomposites. Nanosci Nanotechnol Lett 2012;4:81420.
[207] Dong H, Strawhecker KE, Snyder JF, Orlicki JA, Reiner RS, Rudie AW. Cellulose nanocrystals as a reinforcing material for
electrospun poly(methylmethacrylate) bers: Formation, properties and nanomechanical characterization. Carbohydr
Polym 2012;87:248895.
[208] Branson BT, Seif MA, Davidson JL, Lukehart CM. Fabrication and macro/nanoscale characterization of aggregated and
highly de-aggregated nanodiamond/polyacrylonitrile composite thick lms. J Mater Chem 2011;21:188329.
[209] Mochalin VN, Neitzel I, Etzold BJM, Peterson A, Palmese P, Gogotsi Y. Covalent incorporation of aminated nanodiamond
into an epoxy polymer network. ACS Nano 2011;5:7494502.
[210] Xu GC, Li AY, Zhang LD, Yu XY, Xie T, Wu GS. Nanomechanic properties of polymer-based nanocomposites with nanosilica
by nanoindentation. J Reinf Plast Compos 2004;23:136572.
[211] Zou H, Wu S, Shen J. Polymer/silica nanocomposites: preparation, characterization, properties, and applications. Chem
Rev 2008;108:3893957.
[212] Pavlidou S, Papaspyrides CD. A review on polymer-layered silicate nanocomposites. Prog Polym Sci 2008;33:111998.
[213] Ray SS, Okamoto M. Polymer/layered silicate nanocomposite: a review from preparation to processing. Prog Polym Sci
2003;28:1539641.
[214] Alexandre M, Dubois P. Polymer/layered silicate nanocomposite: preparation, properties and uses of a new class of
materials. Mater Sci Eng R 2000;28:163.
[215] Kojima J, Usuki A, Kawasumi M, Okada A, Kurauchi T, Kamigaito O. One-pot synthesis of nylon 6-clay hybrid. J Polym Sci
Polym Chem Ed 1993;31:17558.
[216] Snchez C, Julin B, Belleville P, Polpall M. Applications of hybrid organic-inorganic nanocomposites. J Mater Chem
2005;28:355992.
[217] Snchez C, Belleville P, Polpall M, Nicole L. Applications of advanced hybrid organic-inorganic nanomaterials: from
laboratory to market. Chem Soc Rev 2011;40:696753.
[218] Treece MA, Oberhauser JP. Processing of polypropylene-clay nanocomposites: single-screw extrusion with in-line
supercritical carbon dioxide feed versus twin-screw extrusion. J Appl Polym Sci 2007;103:88492.
[219] Wong SC, Lee H, Qu S, Mall S, Chen L. A study of the global vs local properties for maleic anhydride modied
polypropylene nanocomposites. Polymer 2006;47:747784.
[220] Martn Z, Jimnez I, Gmez MA, Ade H, Kilkoyne DA, Hernndez-Cruz D. Spectromicroscopy study of intercalation and
exfoliation in polypropylene/montmorillonite nanocomposites. J Phys Chem B 2009;113:111605.
[221] Martn Z, Jimnez I, Gmez MA, Ade H, Kilkoyne DA. Interfacial interactions in PP/MMT/SEBS nanocomposites.
Macromolecules 2010;43:44853.
[222] Martn Z, Jimnez I, Gmez-Fatou MA, West M, Hitchcock AP. Interfacial interactions in polypropylene-organoclay-
elastomer nanocomposites: inuence of polar modications on the location of the clay. Macromolecules
2011;44:217989.
[223] Shen L, Tjiu WC, Liu T. Nanoindentation and morphological studies on injection-molded nylon-6 nanocomposites.
Polymer 2005;46:1196977.
[224] Peng H, Tjiu WC, Shen L, Huang S, He C, Liu T. Preparation and mechanical properties of exfoliated CoAl layered double
hydroxide (LHD)/polyamide 6 nanocomposites by in situ polymerization. Compos Sci Technol 2009;69:9916.
[225] Sikdar D, Katti D, Katti K, Mohanty B. Effect of organic modiers on dynamic and static nanomechanical properties and
crystallinity of intercalated clay-polycaprolactam nanocomposites. J Appl Polym Sci 2007;105:790802.
[226] Jarrar R, Mohsin MA, Haik Y. Alteration of the mechanical and thermal properties of Nylon 6/Nylon 6,6 blends by
nanoclay. J Appl Polym Sci 2012;124:188090.
[227] Shen L, Phang IY, Chen L, Liu T, Zeng K. Nanoindentation and morphological studies on nylon 66 nanocomposites. I. Effect
of clay loading. Polymer 2004;45:33419.
[228] Shen L, Phang IY, Chen L, Liu T, Zeng K. Nanoindentation and morphological studies on nylon 66 nanocomposites. II. Effect
of strain rate. Polymer 2004;45:82219.
92 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

[229] Shen L, Liu T, Lv P. Polishing effect on nanoindentation behavior of nylon 66 and its nanocomposites. Polym Test
2005;24:7469.
[230] Timmaraju MV, Gnanamoorthy R, Kannan K. Inuence of imbibed moisture and organoclay on tensile and indentation
behavior of polyamide 66/hectorite nanocomposites. Composites Part B 2011;42:46672.
[231] Dasari A, Yu ZZ, Mai YW. Nanoscratching of nylon 66-based ternary nanocomposites. Acta Mater 2007;55:63546.
[232] Hu Y, Shen L, Yang H, Wang M, Liu T, Liang T, et al. Nanoindentation studies on nylon 11/clay nanocomposites. Polym Test
2006;25:4927.
[233] Phang IY, Liu T, Mohamed A, Pramoda KP, Chen L, Shen L, et al. Morphology, thermal and mechanical properties of nylon
12/organoclay nanocomposites prepared by melt compounding. Polym Int 2005;54:45664.
[234] Aldousiri B, Dhakal HN, Onuh S, Zhang ZY, Bennett N. Nanoindentation behavior of layered silicate lled spent polyamide-
12 nanocomposites. Polym Test 2011;30:68892.
[235] Zaidi L, Bruzaud S, Bourmaud A, Mdric P, Kaci M, Grohens Y. Relationship between structure and rheological,
mechanical and thermal properties of polylactide/cloisite 30B nanocomposites. J Appl Polym Sci 2010;116:135765.
[236] Lutpi HA, Anuar H, Samat N, Surip SN, Bonnia NN. Evaluation of elastic modulus and hardness of polylactic acid-based
biocomposite by nano-indentation. Adv Mater Res 2012;576:4469.
[237] Bruzaud S, Bourmaud A. Thermal degradation and (nano)mechanical behavior of layered silicate reinforced poly(3-
hydroxybutyrate-co-3-hydroxyvalerate) nanocomposites. Polym Test 2007;26:6529.
[238] Adewole JK, Al-Mubaiyedh UA, Ul-Hamid A, Al-Juhani AA, Hussein IA. Bulk and surface mechanical properties of clay
modied HDPE used in liner applications. Can J Chem Eng 2012;90:106678.
[239] Chua YCh, Chen L, Lu X. Oriented clay-induced anisotropic crystalline morphology in poly(ethylene naphthalate)/clay
nanocomposites and its impact on mechanical properties. Composites Part A 2009;40:42330.
[240] Beake BD, Chen S, Hull JB, Gao F. Nanoindentation behavior of clay/poly(ethylene oxide) nanocomposites. J Nanosci
Nanotech 2002;2:739.
[241] Wang CA, Chen K, Huang Y, Lee H. Electrochemical synthesis and properties of layer-structured polypyrrole/
montmorillonite nanocomposite lms. J Mater Res 2010;25:65864.
[242] Khan SB, Akhtar K, Seo J, Han H, Rub MA. Effect of nano-ller dispersion on the thermal, mechanical and water sorption
properties of green environmental polymer. Chin J Polym Sci 2012;30:73543.
[243] Remili Ch, Kaci M, Benhamida A, Bruzaud S, Grohens Y. The effects of reprocessing cycles on the structure and properties
of polystyrene/cloisite15A nanocomposites. Polym Degrad Stab 2011;96:148996.
[244] Palacio M, Bhushan B, Ferrell N, Hansford D. Nanomechanical characterization of polymer beam structures for BioMEMS
applications. Sens Actuators A 2007;135:63750.
[245] Fan X, Park M, Xia C, Advincula R. Surface structural characterization and mechanical testing by nanoindentation
measurements of hybrid polymer/clay nanostructured multilayer lms. J Mater Res 2002;17:162233.
[246] Katti KS, Katti DR, Dash R. Synthesis and characterization of a novel chitosan/montmorillonite/hydroxyapatite
nanocomposite for bone tissue engineering. Biomed Mater 2008;3:03412234.
[247] Shen L, Wang L, Liu T, He Ch. Nanoindentation and morphological studies of epoxy nanocomposites. Macromol Mater Eng
2006;291:135866.
[248] Piscitelli F, Scamardella AM, Romeo V, Lavorgna M, Barra G, Amendola E. Epoxy composites based on amino-silylated
MMT: the role of interfaces and clay morphology. J Appl Polym Sci 2012;124:61628.
[249] Chen KY, Wang C, Huang Y, Lin W. Preparation and characterization of polymerclay nanocomposite lms. Sci China Ser B
2009;52:23238.
[250] Lin W, Wang C, Le H, Long B, Huang Y. Special assembly of laminated nanocomposite that mimics nacre. Mater Sci Eng C
2008;28(7):10317.
[251] Dhakal HN, Zhang ZY, Richardson MOW. Nanoindentation behavior of layered silicate reinforced unsaturated polyester
nanocomposites. Polym Test 2006;25:84652.
[252] Yusoh K, Song JJM. Subsurface mechanical properties of polyurethane/organoclay nanocomposite thin ms studied by
nanoindentation. Prog Org Coat 2010;67:2204.
[253] Zaidi L, Bruzaud S, Kaci M, Bourmaud A, Gautier N, Grohens Y. The effects of gamma irradiation on the morphology and
properties of polylactide/cloisite 30B nanocomposites. Polym Degrad Stab 2013;98:34855.
[254] Zaidi L, Kaci M, Bruzaud S, Bourmaud A, Grohens Y. Effect of natural weather on the structure and properties of
polylactide/cloisite 30B nanocomposites. Polym Degrad Stab 2010;95:17518.
[255] Seltzer R, Mai YW, Frontini PM. Creep behavior of injection moulded polyamide 6/organoclay nanocomposites by
nanoindentation and cantilever-bending. Composites Part B 2012;43:839.
[256] Filippone G, Acierno D. Nanoparticle dynamics in polymer melts. In: Hashim A, editor. Nanotechnology and
nanomaterials: smart nanoparticles technology; 2012. p. 387406. [chapter 18] ISBN:978-953-51-0500-8 http://
dx.doi.org/10.5772/33647.
[257] Mody VV, Siwale R, Singh A, Mody HR. Introduction to metallic nanoparticles. J Pharm Bioall Sci 2010;2(4):2829.
[258] Li S, Lin MM, Toprak MS, Kim DK, Muhammed M. Nanocomposites of polymer and inorganic nanoparticles for optical and
magnetic applications. Nano Reviews 2010;1:5214. http://dx.doi.org/10.3402/nano.v1i0.5214. 19pp.
[259] Rodrguez JA, Fernndez-Garca M. The world of oxide nanoparticles. In: Rodrguez JA, Fernndez-Garca M, editors.
Synthesis, properties and applications of oxide nanoparticles. New Jersey: Wiley; 2007. p. 15. http://dx.doi.org/10.1002/
0470108975.
[260] Kango S, Kaliab S, Celli A, Njugunad J, Habibie Y, Kumar R. Surface modication of inorganic nanoparticles for
development of organic-inorganic nanocomposites: a review. Prog Polym Sci 2013;38:123261.
[261] Hanemann T, Szab DV. Polymernanoparticle composites: from synthesis to modern applications. Materials
2010;3(6):3468517.
[262] Lam CK, Lau KT. Localized elastic modulus distribution of nanoclay/epoxy composites by using nanoindentation. Compos
Struct 2006;75:5538.
[263] Tzetsis D, Mansour G, Tsias I, Pavlidou E. Nanoindentation measurements of fumed silica epoxy reinforced
nanocomposites. J Reinf Plast Compos 2013;32(3):16073.
A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194 93

[264] Allahverdi A, Ehsanib M, Janpoura H, Ahmadi S. The effect of nanosilica on mechanical, thermal and morphological
properties of epoxy coating. Prog Org Coat 2012;75:5438.
[265] Zhang G, Rasheva Z, Karger-Kocsis J, Burkhart T. Synergetic role of nanoparticles and micro-scale short carbon bers on
the mechanical proles of epoxy resin. eXPRESS Polym Lett 2011;5(10):85972.
[266] Zheng Y, Zheng Y, Ning R. Effects of nanoparticle SiO2 on the performance of nanocomposites. Mater Lett
2003;57:29404.
[267] Zhang H, Zhang Z, Friedrich K, Eger C. Property improvements of in situ epoxy nanocomposites with reduced interparticle
distance at high nanosilica content. Acta Mater 2006;54:183342.
[268] Castrillo PD, Olmos D, Gonzlez-Benito J. Novel polymer composites based on a mixture of preformed nanosilica-lled
poly(methyl methacrylate) particles and a diepoxy/diamine thermoset system. J App Polym Sci 2009;111:206270.
[269] Devaprakasam D, Hatton PV, Mbus G, Inkson BJ. Nanoscale tribology, energy dissipation and failure mechanisms of
nano- and micro-silica particle-lled polymer composites. Tribol Lett 2009;34:119.
[270] Soloukhin VA, Posthumus W, Brokken-Zijp JCM, Loos J, de With G. Mechanical properties of silica-(meth)acrylate hybrid
coatings on polycarbonate substrate. Polymer 2002;43:616981.
[271] Xu GC, Li AY, Zhang LD, Wu GS, Yuan XY, Xie T. Synthesis and characterization of silica nanocomposite in situ
photopolymerization. J Appl Polym Sci 2003;90:83740.
[272] Molazemhosseini A, Tourani H, Naimi-Jamal MR, Khavandi A. Nanoindentation and nanoscratching responses of PEEK
based hybrid composites reinforced with short carbon bers and nano-silica. Polym Test 2013;32:52534.
[273] Stojanovic D, Orlovic A, Markovic S, Radmilovic V, Uskokovic PS, Aleksic R. Nanosilica/PMMA composites obtained by the
modication of silica nanoparticles in a supercritical carbon dioxideethanol mixture. J Mater Sci 2009;44:622332.
[274] Saito R, Kobayashi S-i, Hosoya T. Effects of hydroxyl groups and architecture of organic polymers on polystyrene/silica
nanocomposites. J Appl Polym Sci 2005;97:183547.
[275] Saito R, Kobayashi S-i, Hosoya T. Effect of solvent composition on transparency and surface hardness of poly(methyl
methacrylate)-silica nanocomposites provided on polycarbonate. J Appl Polym Sci 2008;109:1498504.
[276] Douce J, Boilot JP, Biteau J, Scodellaro L, Jimnez A. Effect of ller size and surface condition of nano-sized silica particles
in polysiloxane coatings. Thin Solid Films 2004;466:11422.
[277] Rodrguez Hernndez JC, Salmern Snchez M, Gmez Ribelles JL, Monlen Pradas M. Polymersilica nanocomposites
prepared by solgel technique: nanoindentation and tapping mode AFM studies. Eur Polym J 2007;43:277583.
[278] Ciprari D, Jacob K, Tannenbaum R. Characterization of polymer nanocomposite interphase and its impact on mechanical
properties. Macromolecules 2006;39:656573.
[279] Gupta S, Ramamurthy PC, Madras G. Covalent grafting of polydimethylsiloxane over surface-modied alumina
nanoparticles. Ind Eng Chem Res 2011;50:658593.
[280] Chakraborty H, Sinha A, Mukherjee N, Ray D, Chattopadhyay PP. A study on nanoindentation and tribological behaviour of
multifunctional ZnO/PMMA nanocomposite. Mater Lett 2013;93:13740.
[281] Ramezanzadeh B, Attar MM. Studying the corrosion resistance and hydrolytic degradation of an epoxy coating containing
ZnO nanoparticles. Mater Chem Phys 2011;130:120819.
[282] Hou X, Shan CX, Choy K-L. Microstructures and tribological properties of PEEK-based nanocomposite coatings
incorporating inorganic fullerene-like nanoparticles. Surf Coat Technol 2008;202:228791.
[283] Naffakh M, Dez-Pascual AM, Marco C, Ellis G, Gmez-Fatou MA. Opportunities and challenges in the use of inorganic
fullerene-like nanoparticles to produce advanced polymer nanocomposites. Prog Polym Sci 2013;38:1163231.
[284] Naffakh M, Martn Z, Fanegas N, Marco C, Gmez MA, Jimnez I. Inuence of inorganic fullerene-like WS2 nanoparticles
on the thermal behavior of isotactic polypropylene. J Polym Sci Part B: Polym Phys 2007;45:230921.
[285] Naffakh M, Marco C, Gmez MA, Gmez-Herrero J, Jimnez I. Use of inorganic fullerene-like WS2 to produce new high-
performance polyphenylene sulde nanocomposites: role of the nanoparticle concentration. J Phys Chem B
2009;113:1010411.
[286] Thomas V, Dean DR, Jose MV, Mathew B, Chowdhury S, Vohra YK. Nanostructured biocomposite scaffolds based on
collagen coelectrospun with nanohydroxyapatite. Biomacromolecules 2007;8(2):6317.
[287] McCumiskey EJ, Chandrasekhar N, Taylor CR. Nanomechanics of CdSe quantum dot-polymer nanocomposite lms.
Nanotechnology 2010;21:225703. 7pp.
[288] Kiran MSRN, Raidongia K, Ramamurty U, Rao CNR. Improved mechanical properties of polymer nanocomposites
incorporating graphene-like BN: dependence on the number of BN layers. Script Mater 2011;64(6):5925.
[289] Tehrani M, Luhrs CC, Al-Haik MS, Trevino J, Zea H. Synthesis of WS2 nanostructures from the reaction of WO3 with CS2
and mechanical characterization of WS2 nanotube composites. Nanotechnology 2011;22(28):28571424.
[290] Chen Q, Udomsangpetch C, Shen SC, Liu YC, Chen Z, Zheng XT. The effect of AlOOH boehmite nanorods on mechanical
property of hybrid composite coatings. Thin Solid Films 2009;517(17):48714.
[291] Pelisser F, Gleize PJP, Mikowski A. Effect of poly(diallyldimethylammonium chloride) on nanostructure and mechanical
properties of calcium silicate hydrate. Mater Sci Eng A 2010;527(26):70459.
[292] Misra R, Fu BS, Plagge A, Morgan SE. POSS-nylon 6 nanocomposites: inuence of POSS structure on surface and bulk
properties. J Polym Sci B: Polym Phys 2009;47(11):1088102.
[293] Zeng F, Liu Y, Sun Y, Hu E, Zhou Y. Nanoindentation, nanoscratch, and nanotensile testing of poly(vinylidene uoride)-
polyhedral oligomeric silsesquioxane nanocomposites. J Polym Sci B: Polym Phys 2012;50(23):1597611.
[294] Wahab MA, Mya KI, He C. Synthesis, morphology, and properties of hydroxyl terminated-POSS/polyimide low-k
nanocomposite lms. J Polym Sci A: Polym Chem 2008;46(17):588796.
[295] Huang J, Xiao Y, Mya KY, Liu X, He C, Dai J, et al. Thermomechanical properties of polyimide-epoxy nanocomposites from
cubic silsesquioxane epoxides. J Mater Chem 2004;14(19):285863.
[296] Mammeri F, Le Bourhis E, Rozes L, Sanchez C. Elaboration and mechanical characterization of nanocomposites thin lms
part I: determination of the mechanical properties of thin lms prepared by in situ polymerisation of tetraethoxysilane in
poly(methyl methacrylate). J Eur Ceram Soc 2006;26(3):25966.
94 A.M. Dez-Pascual et al. / Progress in Materials Science 67 (2015) 194

[297] Mammeri F, Le Bourhis E, Rozes L, Sanchez C. Elaboration and mechanical characterization of nanocomposites thin lms
part II: correlation between structure and mechanical properties of SiO2-PMMA hybrid materials. J Eur Ceram Soc
2006;26(3):26772.
[298] Yahyaei H, Mohseni M. Use of nanoindentation and nanoscratch experiments to reveal the mechanical behavior of solgel
prepared nanocomposite lms on polycarbonate. Tribol Int 2013;57:14755.
[299] Ngan AHW, Tang B. Viscoelastic effects during unloading in depth-sensing indentation. J Mater Res 2003;18(5):11418.
[300] Musbah SS, Radojevic V, Radovic I, Uskokovic PS, Stojanovic DB, Dramicanin M, et al. Preparation, characterization and
mechanical properties of rare-earth-based nanocomposites. J Min Metall B 2012;48(1):30918.
[301] Gir-Paloma J, Roa JJ, Dez-Pascual AM, Rayn E, Flores A, Martnez M, et al. Depth-sensing indentation applied to
polymers: a comparison between standard methods of analysis in relation to the nature of the materials. Eur Poym J
2013;49:404753.
[302] Deuschle J, Enders S, Arzt E. Surface detection in nanoindentation of soft polymers. J Mater Res 2007;22(11):310719.
[303] Piccarolo S, Falsone A, Poulose AM. Improving surface detection on nanoindentation of compliant materials. Meas Sci
Technol 2010;21(6):065701. 7pp.
[304] Nili H, Kalantar-zadeh K, Bhaskaran M, Sriram S. In situ nanoindentation: probing nanoscale multifunctionality. Prog Mat
Sci 2013;58:129.
[305] Hayes SA, Goruppa AA, Jones FR. Dynamic nanoindentation as a tool for the examination of polymeric materials. J Mater
Res 2004;19(11):3298306.
[306] Rhim J-W, Park H-M, Ha C-S. Bio-nanocomposites for food packaging applications. Prog Polym Sci 2013;38(10
11):162952.
[307] Ebenstein DM, Pruitt LA. Nanoindentation of biological materials. Nanotoday 2006;1(3):2633.
[308] Shepherd TN, Zhang J, Ovaert TC, Roeder RK, Niebur GL. Direct comparison of nanoindentation and macroscopic
measurements of bone viscoelasticity. J Mech Behavior Biomed Mater 2011;4:205562.
[309] Zysset PK. Indentation of bone tissue: a short review. Osteoporos Int 2009;20:104955.
[310] McKee CT, Last JA, Russell P, Murphy CJ. Indentation versus tensile measurements of Youngs modulus for soft biological
tissues. Tissue Eng B 2011;17(3):15564.
[311] Tang B, Ngan AHW. A rate-jump method for characterization of soft tissues using nanoindentation techniques. Soft
Matter 2012;8(22):59749.
[312] Balooch M, Habelitz S, Kinney JH, Marshall SJ, Marshall GW. Mechanical properties of mineralized collagen brils as
inuenced by demineralization. J Struct Biol 2008;162:40410.
[313] Notbohm J, Poon B, Ravichandran G. Analysis of nanoindentation of soft materials with an atomic force microscope. J
Mater Res 2012;27(1):22937.
[314] Ren J, Yu S, Gao N, Zou Q. Indentation quantication for in-liquid nanomechanical measurement of soft material using an
atomic force microscope: rate-dependent elastic modulus of live cells. Phys Rev E 2013;88:052711. 12 pp.
[315] Crick SL, Yin FC-P. Assessing micromechanical properties of cells with atomic force microscopy: importance of the contact
point. Biomechan Model Mechanobiol 2007;6:199210.
[316] Constantinides G, Kalcioglu ZI, McFarland M, Smith JF, VanVliet KJ. Probing mechanical properties of fully hydrated gels
and biological tissues. J Biomech 2008;41:32859.
[317] Ebenstein DM, Pruitt LA. Nanoindentation of soft hydrated materials for application to vascular tissues. J Biomed Mater
Res A 2004;69(2):22232.
[318] Ngan AHW, Tang B. Response of power-law-viscoelastic and time-dependent materials to rate jumps. J Mater Res
2009;24(3):85362.

Вам также может понравиться