You are on page 1of 23

Road Materials and Pavement Design

ISSN: 1468-0629 (Print) 2164-7402 (Online) Journal homepage: http://www.tandfonline.com/loi/trmp20

Effect of blending conditions on nano-clay


bitumen nanocomposite properties

Stephen A. Bagshaw, Tim Kemmitt, Mark Waterland & Sam Brooke

To cite this article: Stephen A. Bagshaw, Tim Kemmitt, Mark Waterland & Sam Brooke (2018):
Effect of blending conditions on nano-clay bitumen nanocomposite properties, Road Materials and
Pavement Design

To link to this article: https://doi.org/10.1080/14680629.2018.1468802

Published online: 07 May 2018.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=trmp20
Road Materials and Pavement Design, 2018
https://doi.org/10.1080/14680629.2018.1468802

Effect of blending conditions on nano-clay bitumen nanocomposite


properties
Stephen A. Bagshawa∗ , Tim Kemmittb , Mark Waterlandc and Sam Brookec
a WSP Opus Research, WSP Opus International Consultants Ltd, Petone, New Zealand b Callaghan Innova-
tion, Lower Hutt, New Zealand c Institute of Fundamental Sciences, College of Science, Massey University,
Palmerston North, New Zealand

(Received 29 August 2017; accepted 12 April 2018 )

Nanoclay/bitumen nanocomposites have been prepared using a commercial organically mod-


ified nanoclay (Nanomer), and two different penetration grade bitumens under a range of
different blending conditions. Simple paddle stirring and high-shear mixing for different times
at different shear speeds were investigated in order to determine whether expensive and
time-consuming blending was necessary to achieve significant changes in the properties of
the bitumen composites. Longer blending times and/or higher shearing speeds resulted in
increased dispersion and exfoliation of the Nanomer particles, but this was not necessar-
ily beneficial to the physico-chemical properties of the final material. Rheological analyses
showed insignificant differences in properties between the samples blended for different times
and speeds of shear mixing. Simple paddle mixing was sufficient to disperse the Nanomer to
generate nanocomposites with improved physical properties. Dispersion of Nanomer into the
softer of the two bitumen grades caused larger apparent changes than blending into the stiffer
grade. These results suggest that the use of simple, low-cost blending techniques should be
possible to manufacture effective clay-bitumen nanocomposites.
Keywords: bitumen; nanoclay; organoclay; nanocomposite; blending; rheology

1. Introduction
Over 80% of the sealed road network in New Zealand consists of chipseals (sprayseals) of various
kinds constructed over unbound granular pavements (UGPs) (NZTA, 2005a). In New Zealand,
these chipseals are used as the primary wearing course on urban, peri-urban and inter-city traf-
fic routes (NZTA, 2005a). Internationally, however, chipseals are more likely to be applied as
surface dressings on worn or damaged asphalt concrete (AC) surfaces for life-span extension
(Gransberg & James, 2005; Howard, Cox, Alvarado, & Jordan, 2017). Chipseals in New Zealand
are also being subjected to ever-increasing loads and stresses; therefore, the materials used to
prepare these chipseals must be able to withstand these increasing performance needs. Unfortu-
nately, chipseals can fail to meet these performance requirements, experiencing different failure
mechanisms that manifest variously in the short, medium and longer terms. They include friction
loss (polishing), cracking, texture loss (flushing), potholing and chip loss (NZTA, 2005b).
While some failure mechanisms are due to degradation of the chip, other failure mecha-
nisms can occur because of the bitumen binder not meeting the increasing performance demands
being placed upon it in areas such as temperature stability, shear/creep, adhesion and long-term

*Corresponding author. Email: steve.bagshaw@opus.co.nz

© 2018 Informa UK Limited, trading as Taylor & Francis Group


2 S.A. Bagshaw et al.

aging (Brown, Henning, Herrington, & Wu, 2015; Herrington, Kodippily, & Henning, 2015).
Ultimately, these failure mechanisms will result in the road requiring early resurfacing, rehabili-
tation (overlays) or complete pavement reconstruction (Neaylon & Harrow, 2017; NZTA, 2005b;
Rahman, Shahidul Islam, Musty, & Hossain, 2012). While the total amount of bitumen binder
that is present in a unit volume of chipseal is somewhat lower than that of an AC surface, it is
also likely that the film thickness of binder associated with any given aggregate chip is some-
what greater than that in an AC. Therefore, the in-service mobility of the chipseal binder may be
somewhat higher, thereby placing significant emphasis on improving the rheological properties
of the binder.
In a wider research programme undertaken here, new technologies and methods are being
sought to improve chipseal performance through bitumen binder modification (New Zealand
Ministry of Business, Innovation and Employment (MBIE), 2016). Primary among the perfor-
mance criteria that might offer improvements in the chipsealing context are those of flow or
shear at elevated temperatures (summer conditions), cracking at reduced temperatures (winter
conditions), chip adhesion (wet and cold conditions) and aging/embrittlement over time (oxida-
tion/composition change). While various means are available to modify binder properties, such
as the addition of polymers, often the materials properties offered by those modifications do not
lead to desirable outcomes, they are often expensive and they can introduce new complications
(Galooyak, Dabir, Nazarbeygi, Moeini, & Berahman, 2011; Kök & Akpolat, 2015; Moghadas
Nejad, Azorhoosh, & Hamedi, 2014).
The incorporation of nanoclays into polymeric, elastomeric and gel materials (plastics, resins,
rubbers, hydrogels), forming so-called clay/polymer nanocomposites, has been demonstrated to
mitigate many of the shortcomings that have been inherent in the industrial use of polymers
and has provided many new opportunities (Arora & Padua, 2010; Hu, Onyebueke, & Abatan,
2010; Olad, 2011; Osman, Mittal, & Suter, 2007; Paul & Robeson, 2008). Many of the failure
issues implicated in polymer engineering are similar to those identified in bitumen failure mech-
anisms such as load induced creep, insufficient softening/melting temperatures, adhesion, fatigue
resistance and gas permeability. It seems logical to suggest, therefore, that the formation of bitu-
men/clay nanocomposites is possible and that such materials may go some way to mitigating
some of the aforementioned failure issues (Fang, Yu, Liu, & Li, 2013; Ghile, 2006; Jahromi &
Khodaii, 2009; You et al., 2011).

2. Background
The chemical composition of bitumen can be described in terms of the SARA fractions,
namely Saturates, Aromatics, Resins (Maltenes) and Asphaltenes. Saturates and aromatics are
solvent-like molecules, resins are heavier compounds and oligomeric species and asphaltenes
are poly-cyclic aromatic compounds with planar structures and large aspect ratios. Bitumens also
include numbers of other species such as heteroatomic species and metal compounds (Read &
Whiteoak, 2003; Weigel & Stephan, 2017). The leading model that describes the arrangements
of those molecules into a structure is the so-called sol–gel model. This describes the structure
of bitumen as a dispersion of asphaltene compounds that are stabilised by resins and in turn
dispersed in the continuous phase of mixed molecular weight aromatics and saturates (Lesueur,
2009). Soft, or more fluid, bitumens with low asphaltene contents ( < 18%) are referred to as ‘sol’
bitumens where the rheological properties are dominated by the ‘S, A and R’ components. Stiffer,
or more viscous, bitumens have higher asphaltene contents ( > 25%) and are referred to as ‘gel’
bitumens and the rheological properties are more determined by the ‘R’ and asphaltene (‘A’)
components. The prevailing structural model as described above suggests that asphaltenes are
Road Materials and Pavement Design 3

planar poly-cyclic aromatic hydrocarbon molecules that bear structural resemblance to graphene
fragments (Redelius, 2006). In that work, it was proposed that those planar asphaltenes can form
stacks which are solubilised by the surrounding resin phase into the maltene continuous phase.
Redelius, however, challenged the so-called sol–gel model suggesting that emulsified micelles
do not exist in bitumen. He rather suggested that bitumen is a continuum of species of different
molecular weights, polarities and polarisabilites, which is necessary to retain complete stability
of the matrix. It may be possible to invoke either structural model to explain the possible sol-
ubilisation or stabilisation of organoclay particles into bitumen continuous phases, where they
are either stabilised by the maltene/ resin phase, or they possess suitable low polarity to become
solublised within the maltene/resin phase.
Montmorillonite clays (MMTs) are members of the smectite, ‘swelling clay’ family of phyl-
losilicates and form two-dimensional plate-like structures, which stack into ‘pack-of-cards’ type
tactoids. Charge mismatch between the aluminate and silicate layers causes the 2:1 layers to
be negatively charged and this charge is balanced by exchangeable cations, typically hydrated
Na+ or Ca2+ (Figure 1(a)). The platelet layers are 1–2 nm thick with 2D aspect ratios of approx-
imately 100:1, and the platelets stack into tactoids of several platelets and tactoids assemble
into particles (Anthony, Bideaux, Bladh, & Nichols, 1995). The nanometre dimension of the
platelets, therefore, loosely permits modified clays to be referred to as nanoclays when applied
in non-monolithic applications.
The surfaces of clay platelets however are hydrophilic and therefore chemically incompati-
ble with the organic bituminous matrix. To rectify this, the ion-exchange capability of the MMT
nanoclay is exploited and the interlayer Na+ cations are replaced by long-chain alkyl-ammonium
cations, to form organo-nanoclay (OMMT). This renders the material organophilic and therefore
chemically compatible with the bitumen continuous phase (Liu, Wu, van de Ven, Yu, & Mole-
naar, 2010). In an aqueous suspension, a hydrophilic clay exists as predominantly exfoliated
(widely separated) platelets. In an organic suspension, however, it is possible to form different
structures depending on the blending regime, the continuous phase chemistry and its rheology:
Phase-separated, Intercalated and Exfoliated (Figure 1(b)). Each describes a different degree of
dispersion and separation of the clay particles into the continuous phase (e.g. bitumen) and each
phase may impart different rheology modification upon the matrix.
From the rheology perspective, the non-Newtonian nature of bitumen means that the rheologi-
cal responses are different at low, intermediate and high temperatures, and under different loading
rates. The formation of the bitumen–clay nanocomposite is aimed at augmenting the binder per-
formance in road surfacing, in all temperature regimes and load conditions. When very cold, a
standard bitumen might behave as a brittle material, while a modified bitumen may reinforce
the matrix and reduce cracking rates. At cool, to moderate temperatures, the linear elastic nature
of the material predominates where organoclay addition might result in matrix reinforcement
with improved elasticity and fatigue characteristics. At higher temperatures, where the bitumen
behaves as a Newtonian viscous fluid, the flow dynamics of the fluid may be modified by added
organoclay, causing increased viscosities and thereby reduced mobilities.
The addition of nanoclays into polymeric and gel-like matrices is believed to promote
supramolecular reorganisation within the polymeric continuous phases (Stefanescu, 2008). Many
different types of these supramolecular structures might be formed depending on the concentra-
tion or type of clay and the particular matrix, all of which might be precursors for the more
widespread hierarchical structuring of the matrix. This kind of molecular rearrangement might
be especially interesting in visco-elastic materials, such as bitumen, that are especially sensitive
to external stimuli including shear stress and temperature (Nazzal, Kaya, Gunay, & Ahmedzade,
2013). The improvement of bitumen properties to meet ever-increasing performance demands in
road surfacing is an intense area of research and development. To this end, the incorporation of
4 S.A. Bagshaw et al.

(a)

(b)

Figure 1. (a) Layered montmorillonite organoclay (Cavalcanti et al., 2012). (b) Stages of nanocomposite
formation (Source: https://www.intechopen.com/source/html/42654/media/image1_w.jpg)

various polymer, micro and nanomaterials into bitumen has been studied increasingly (Ashish,
Singh, & Bohm, 2017; Hamedi, Moghadas Nejad, & Oveisi, 2015; Jahromi, Andalibizade, &
Vossough, 2010; Kattak, Khattab, Rizvi, & Zhang, 2012). The incorporation of these materi-
als into bitumen binders almost invariably demonstrates a modification to the rheological and
engineering properties of the binder (Ghile, 2006).
In the chipsealing context, of interest to practitioners in New Zealand, at moderate ‘in-service’
temperatures, primary among the engineering challenges of bitumen binders that require solving
are those of shear and/or creep, long-term aging/cracking and bitumen/chip adhesion (Brown
et al., 2015; Herrington et al., 2015). The proposition offered here is that the formation of bitu-
men/clay nanocomposites will mitigate these challenges by making fundamental changes to the
structure and nature of the bitumen binder. While the body of research to date indicates that it
Road Materials and Pavement Design 5

can now be reasonably asserted that bitumen/clay nanocomposites are advantageous in the AC
context, there is scant evidence to suggest the same will be so in the chipsealing context.
Should such materials progress into the field, an important aspect of their commercial devel-
opment will be to understand what processing conditions will optimise the nanocomposite
performance and optimise the commercial value proposition. To that end, in this work we have
examined nanoclay/bitumen blending regimes, in the laboratory context, with regards to mixing
type, shear mixing speed and time (Karşal, Tanoğlu, Odabaş, Ersoy, & Karakaya, 2009). The
effects of different blending regimes on structure, rheology and basic empirical performance val-
ues of bitumen/clay nanocomposites were examined using a range of instrumental and physical
probe methods and are described. Subsequent work will describe the improvements or otherwise
with respect to larger-scale chipsealing outcomes.

3. Materials and methods


3.1. Organo-nanoclay and bitumen materials
Organo-functionalised montmorillonite nanoclay materials, ‘Nanomer 1.44P’, hereafter
‘Nanomer’, are manufactured by and were kindly supplied by Nanocor Inc, IL, USA. The mate-
rials properties of Nanomer are proprietary but the Nanomer I.44P technical data provided are
presented in Table 1. Two different penetration grade bitumen binders were investigated in
this work. Typically, the prevailing climatic conditions in New Zealand allow specification of
the high penetration, ‘sol-type’, 180–200 grade binder, while in hotter or higher load areas a
‘gel-type’ 80–100 grade can be specified.
While the use of bitumen emulsions for chipsealing is increasing in New Zealand, the use of
hot cut-back binders still predominates. The two binders used here were supplied by The NZ
Refining Company Ltd and were obtained from different batches prepared contemporaneously
and their properties are given in Table 2.

3.2. Sample preparation


Bitumen samples of approximately 400 g were pre-heated to 160 ± 10°C for 2–3 h, in steel cans
contained in sand baths (to control and maintain temperatures). Under mechanical paddle stirring
with a flat, semi-circular shaped paddle, either 2 or 6 wt% of Nanomer was added slowly into
the vortex to ensure good initial dispersion. Upon completion of Nanomer addition, when shear
mixing was required, the hot sample was transferred to a Silverson L5M-A high-speed shear
mixer equipped with a 50 mm vane, covered and blended at either 2000 or 4000 rpm for 15, 30
or 60 mins at 160 ± 10°C. Hot samples for the various analyses were taken and the remainder
was sealed and stored for future use. During blending, all nanocomposite samples changed from
being viscous but flowing fluids, to being much more viscous, or into weak gels that flowed only

Table 1. Nanomer I.44P product data.

Surface modifier Di-methyl, di-hydrogenated tallow ammonium


Appearance Off white free flowing powder
Surface modifier concentration 34–36 wt%
Bulk density 250–300 kg/m3
Particle size (mean) 14–18 micron
Specific gravity 1.8 g/cm3
XRD (d001 ) 2.4–2.6 nm
6 S.A. Bagshaw et al.

Table 2. Physical properties of base bitumen and paddle-mixed nanocomposite samples.


180–200 180–200
180–200 80–100 2% 6% 80–100 2% 80–100 6%
Propertya grade grade OMMT OMMT OMMT OMMT

Density (25°C) 1.033 1.040 1.039 1.042 1.046 1.057


(g cm−3 )
Penetration (25°C) 170 84 122 105 74 54
Softening point (°C) 40.3 47.2 43.7 50.2 49.6 56.7
Ductility (cm) > 100 > 100 > 100 > 100 > 100 > 100
Rotational viscosity 0.275 0.418 0.471 1.378 0.756 1.813
(135°C) (Pa s)
b G* 5°C (MPa) 16.8 44.8 22.1 28.7 54.2 59
b Phase angle 5°C (°) 43.8 37.7 57.3 42.7 44.1 36.5
b G* 55°C (MPa) 1.58 19.10 8.10 17.06 26.95 42.18
b Phase angle 55°C (°) 79.4 79.3 77.7 77.1 79.3 74.3
a Properties determined from base and paddle-mixed samples.
b G* and phase angle determined at 9.042 Hz for base and paddle-mixed samples

slowly at 160°C. For simplicity, the following discussion will be limited to the samples prepared
with 6 wt% added Nanomer.

3.3. Physical properties


Density, penetration, softening point and ductility were all measured using standard ASTM meth-
ods, details of which can be found elsewhere (ASTM D113-17, 2017; ASTM D36/D36M-14e1,
2014; ASTM D5/D5M, 2013; ASTM D70-17, 2017). Changes in rotational viscosity following
the different blending regimes were identified by measuring viscosity at 135°C with a Brookfield
digital viscometer and Thermocell, using SC4-# 21 spindle at 45 rpm over approximately 6000 s.
Results were obtained from three replicates which showed low variability of less than 10%
between samples. Storage stability was examined using a technique applied to polymer-modified
binders as described in the standard (AG:PT/T108, 2006).

3.4. X-ray diffraction


Dispersion and intercalation/exfoliation of the Nanomer particles in the blended bitumen
were followed using X-ray diffraction (XRD) to measure the apparent dispersion of the clay
and the distances between clay layers. The diffractometer used was a Philips PW1700-series
Bragg Brentano diffractometer, using Co Kα radiation with a weighted average wavelength of
0.179026 nm, from a long fine focus tube running at 40 kV and 35 mA. The diffractometer was
equipped with mechanical variable divergence slit; diffracted beam graphite monochromator;
goniometer radius of 173 mm and Xenon gas proportional counter detector. The receiving slit
was a 0.2 mm aperture. Soller slits for axial collimation were fitted to both incident and diffracted
beams, while no anti-scatter slit was fitted.
It is assumed here that the nanoclay distribution was sufficiently homogeneous to provide for
representative sampling. Increasing dispersion of the nanoclay particles, therefore, will result
in more particles being present in the beam and therefore, increasing the intensity of the X-ray
patterns obtained. Simultaneously, increasing intercalation and exfoliation of the nanoclay will
be observed as a shift of d00l reflections to lower angles, which indicate wider layer spacings. The
distance between repeated atomic layers of the clay mineral, i.e. the basal spacing, which changes
Road Materials and Pavement Design 7

depending on the degree of intercalation and/or exfoliation, is determined by Braggs Law:

d = nλ/2sinθ, (1)

where n is an integer representing the number of repeat layers, λ is the incident X-ray wavelength
and θ is the angle of the diffracted X-rays. In the case of 2D layers materials, the reflections are
assigned d00l where l = n. Therefore, the lower the angle at which the reflection is detected, the
further apart are the repeating layers that create that reflection.
Hot and stirred bitumen samples were cast into a 25 mm diameter, 2 mm deep aluminium
sample holder, allowed to cool, then levelled by cutting with a hot metal knife to render the
sample flat and level with the top of the sample holder. Diffractograms were obtained from 1° to
20° 2θ, collecting for 1 s at 0.5° 2θ steps.

3.5. Atomic force microscopy


Dispersion of clay particles into the bitumen continuous phase was also analysed by Atomic
Force Microscopy (AFM). Micrographs were obtained with a NanoSurf easyScan 2 in tap-
ping mode, at normal room temperature and pressure, with regular rectangular silicon can-
tilevers, using scan parameters: Time/line: 0.6 s; Points/line: 512; Rotation: 0°; Setpoint: 50%
(Tapping mode amplitude); P-gain: 1500–2500 (sample-dependent); I-Gain: 750–1500 (sample-
dependent); D-gain: 0. Images were manipulated as follows: Backgrounds were removed and
deflection and phase data were filtered at high 1D Fourier modulus density to smooth feedback
artefacts. Samples were presented as 5 mm diameter plugs in silicone moulds. Samples were
poured hot into the moulds, allowed to cool and then cut level with the top surface of the mould
using a hot knife. The top surface was then annealed briefly with a hot-air gun to remove surface
irregularities.

3.6. Fourier transform infrared spectroscopy


Changes to the molecular nature of the materials following blending were followed with Fourier
Transform Infra-Red spectroscopy (FTIR). Samples were prepared as thin films smeared onto a
KBr plate and spectra obtained with a Perkin-Elmer System 2000 FTIR spectrometer in transmis-
sion mode at 2 cm−1 resolution, from 450 to 4500 cm−1 over 100 scans. To mitigate differences
due to sampling inconsistencies, spectra were normalised to the bitumen aromatic structural
mode at 812 cm−1 .

3.7. Dynamic shear rheology


Rheological properties were followed with an AR 2000EX Dynamic Shear Rheometer using the
Linear Sweep technique where the responses of the different samples at three different temper-
atures (5°C, 20°C and 55°C) were observed over a fixed frequency range of 0.09042–90.42 Hz
and deformation strain of 1%. Samples were 8 or 25 mm diameter depending on the test temper-
ature, and all samples were analysed as 1-mm-thick films. Various rheological properties were
determined from the data obtained.
The complex modulus (G*) represents the stiffness, or resilience to deformation, of the binder,
with higher values indicating greater stiffness. The phase angle, δ, represents the elasticity, or
deformation recovery of the binder, with lower phase angles representing the more elastic char-
acter and hence greater degree of recovery from deformation (McGennis, Shuler, & Bahia, 1994).
Values for G* and phase angle (δ) were taken at 9.042 Hz and plotted against the temperature of
the measurement, namely 5°C, 20°C and 55°C.
8 S.A. Bagshaw et al.

The formula G*/sin δ can be used to describe the resistance to permanent deformation at ele-
vated temperatures. The Superpave standard for AC requires that G*/sin δ = 1 kPa as a minimum
for the rutting parameter of an unaged sample (D’Angelo, 2002). A suitable value for chipsealing
binders may be different.
The failure temperature of bitumen is defined by the Superpave specification as the point at
which the rutting parameter (G*/sin δ) falls below 1.0 kPa. This property is used to identify the
performance grade of AC binders (McGennis et al., 1994). A suitable value for chipseal binders
might be different.
G*.sinδ is used to describe the fatigue resistance of a bitumen binder at intermediate tempera-
tures. Superpave provides for a maximum fatigue parameter of 5000 kPa (McGennis et al., 1994)
in AC. A suitable value for chipsealing binders may be different.

4. Results and discussion


4.1. Physical properties
The data presented in Table 2 show very clear changes in the values of the physical proper-
ties of the nanocomposites over those of the base binders. The densities of the nanocomposites
all increased in an approximately linear fashion, suggesting that the nanoclays have a signif-
icant effect on the molecular structuring within the bitumen. Penetration values all decreased
and will be discussed more deeply in a later section. The softening points of all the samples
increased significantly after Nanomer blending. It was observed that the softening point of the
6 wt% OMMT/180–200 grade bitumen was higher than that of the 80–100 grade base bitumen.
Indeed, the softening point of the 6 wt% 80–100 grade sample was found to be similar to that
of a 40–50 Pen grade bitumen, indicating that the effect was maintained across grades. Soften-
ing and increased mobility of chipsealing binder is a major challenge in the chipsealing sector;
therefore, this is a potentially very significant finding. The ductility measurements indicated that
incorporation of the Nanomer into the bitumen did not affect the low-rate, controlled-temperature
elasticity of the bitumen.

4.2. Nanomer dispersion


Current understanding suggests that intercalation or exfoliation of the nanoclay particles within
the bitumen matrix is important for creating the largest effect on the materials properties of
the composite binder (Ashish et al., 2017). However, it may indeed be possible that modifying
the degree of dispersion will offer opportunities for tuning of the bitumen composite prop-
erties. If it is assumed that the dispersion is homogeneous, then XRD from a cast, then cut
surface should reasonably probe the dispersion of clay in the bitumen. XRD patterns presented
in Figure 2 describe the range of structural changes that were induced by different blending
regimes. The weak XRD pattern of the base bitumen showed no evidence of crystalline com-
ponents, exhibiting only the broad amorphous halo that is centred around 25° 2θ , that is typical
of such non-crystalline materials. Conversely, all of the bitumen/clay composite samples exhib-
ited diffraction patterns due to OMMT clay superimposed on the weak bitumen pattern. These
patterns changed depending on the blending regime.
Powder XRD patterns of non-oriented dry Na-MMT clays (Figure 2(a)) are often broad and
exhibit d001 and d002 reflections at around 7.5° 2θ , representing layer spacings of approxi-
mately 0.70 nm (Anthony et al., 1995). The XRD pattern of a non-oriented powder of Nanomer
(Figure 2(b)), in which Na+ cations have been ion-exchanged for long-chain alkyl-ammonium
cations, displays two broad d001 and d002 reflections at 4.0° and 8.1° 2θ , respectively. These
Road Materials and Pavement Design 9

(a) (b)

(c) (d)

(e) (f)

Figure 2. XRD patterns of 6 wt% Nanomer/bitumen samples after shear mixing at 150°C. (a) Na-MMT,
(b) Nanomer I.44P OMMT, (c) 180–200, 2000 rpm; (d) 180–200, 4000 rpm; (e) 80–100, 2000 rpm and (f)
80–100, 4000 rpm.

reflections correspond to layer spacings that have expanded to approximately 1.3 nm. In sharp
contrast, the XRD patterns of all of the nanocomposite materials exhibited three to five sharp d00l
reflections with varying intensities, depending on the mixing regime of the sample. Sharp reflec-
tions indicate that particles with multiple platelets were present, while the multiple reflections
indicate that the clay particles appear to have been preferentially oriented during either mixing
or sample preparation. Furthermore, the reflections all shift to even lower angle, relative to the
starting Nanomer, with d001 reflections being found at approximately 2.3° 2θ. This describes
layer spacings that have increased to approximately 2.4 nm, which must be due to a high degree
of intercalation of bitumen into the organophilic interlayer space.

4.2.1. 180–200 Grade bitumen


For the more fluid 180–200 grade bitumen, at low shearing speed (2000 rpm, Figure 2(c)), the
intensities of the reflections tended to increase with mixing time, but there was no further signifi-
cant shift of the reflections to lower angles. At 2000 rpm, all of the d001 reflections were observed
10 S.A. Bagshaw et al.

at approximately 2.25° 2θ . This suggests that intercalation of bitumen into the interlayer space
of the Nanomer occurred rapidly after addition to the hot bitumen, but that little extra swelling
of the layers occurred. Simultaneously, preferential orientation of the clay particles occurred at
a slow rate. Maximum intensities, i.e. greatest dispersion, were observed after 60 min mixing at
2000 rpm.
At higher shearing speed (4000 rpm, Figure 2(d)), the intensities of reflections changed much
less with increasing mixing time and the d001 reflection was found at approximately 2.3° 2θ
for the paddle mixed and 15 min samples. However, after shear mixing at 30 and 60 min, a
new reflection at approximately 1° 2θ became apparent, and the first clear reflection shifted to
approximately 2.4° 2θ. The reflections below 2.0° 2θ, and the 2.4° 2θ reflection are therefore
reassigned to d001 and d002 respectively and are likely due to layer spacings greater than 5 nm.
This observation is commensurate with the formation of an exfoliated phase (Figure 1(b)) where
the clay platelets are further solubilised by the bitumen components and pushed further apart.
These observations indicate that the organoclay Nanomer is dispersed rapidly and widely into
the solvent dominated, sol-type 180–200 grade bitumen, even with the low-energy input of the
paddle mixer. They also indicate that shear mixing, over even short times, can cause preferential
alignment of the plate-like organoclay particles within the bitumen, and that shear mixing at
high speed (4000 rpm) and longer times (30 and 60 min) induces exfoliation of the organoclay
structure. However, the clay layers remain coherent with one another. This might perhaps be
better described as extended intercalation rather than true exfoliation where platelets are expected
to lose a degree of coherency.

4.2.2. 80–100 Grade bitumen


For the stiffer 80–100 grade bitumen, the XRD results reflected the more viscous nature of the
stiffer binder and suggested that dispersion and exfoliation of the nanoclay required somewhat
more work. The reflections after shear mixing at low speed (2000 rpm, Figure 2(e)) were sharp
with low intensities and reflections assignable out to d004 were observed. This suggested that
the Nanomer clay particles were not quite as well dispersed or aligned as seen in the 180–200
grade bitumen, but that preferential orientation did still occur. The d001 reflection was observed
at approximately 2.2° 2θ in all samples, suggesting that the degree of intercalation was similar
to that in the 180–200 grade bitumen samples.
After mixing at the higher speed however (4000 rpm, Figure 2(f)), the intensities of the reflec-
tions increased sharply over those of the low-speed samples, and at up to 30 min mixing the
patterns remained approximately the same. This indicated better dispersion of the nanoclay par-
ticles throughout the bitumen than that obtained with low-speed mixing, continued preferential
orientation and similar intercalation. Mixing at 4000 rpm for 60 min appeared to induce sig-
nificantly more exfoliation of the clay layers with the appearance of new, weak reflections at
approximately 1.5° and 3.0° 2θ. This would suggest the formation of a co-existing phase with a
layer spacing of approximately 3.5 nm which is indicative of a highly intercalated organoclay.
The Nanomer organoclay dispersed into the more viscous 80–100 grade bitumen, but more
slowly than into the more fluid 180–200 grade bitumen, but the clay particles were also pref-
erentially oriented. The degree of exfoliation observed was lower than that seen in the more
fluid bitumen, but a new, highly intercalated phase was formed after high-speed mixing for 60
min. If the sol–gel model of bitumen is assumed, this difference in intercalation potential can
be explained by the difference in the interactions of the organoclay with the low molecular mass
solvent-like resin/maltene components of the bitumen. These solvent-like molecules might effec-
tively dissolve the organo-clays into the matrix and are thereafter able to penetrate the interlayer
space and act to exfoliate the clay layers. Lower ratios of solvent-like molecules to asphaltene
Road Materials and Pavement Design 11

molecules in the 80–100 grade bitumen manifest in lower rates of intercalation of those sol-
vent molecules into the interlayer spaces of the nanoclay. In the more fluid 180–200, the higher
composition of solvent molecules makes the opposite true.

4.3. AFM
AFM was employed in an effort to image the nanoclay particles and their dispersion throughout
the bitumen phases and indeed, how that dispersion might affect the microstructure of the bitumen
itself. It has been proposed by various authors (Aguiar-Moya et al., 2015; Das, Baaj, Tighe, &
Kringos, 2016; Jahangir, Little, & Bhasin, 2015; Rebelo et al., 2014) that AFM is capable of
probing not only the surface structure, but also the microrheological properties of the material
(Allen, Little, Bhasin, & Lytton, 2013). AFM investigations by Masson, Leblond, and Margeson
(2006) suggested four different structural phases that contain the various chemical phases, i.e. the
SARA fractions, within the bitumen; namely the sal-, para-, per- and catana phases. However, no
direct correlations were identified between the AFM morphologies and the chemical components.
In this paper, our aim was simply to establish if it is possible to identify the dispersed nanoclay
particles and any differences in the physical states of the nanocomposite materials that were
prepared by different blending methods. Phase mode has been selected whereby components of
different stiffness or hardness can be distinguished.

4.3.1. 180–200 Grade bitumen


Phase images of 180–200 grade bitumen/nanoclay composites blended at 4000 rpm for different
times, as indicated, are provided in Figure 3. In the image of the base binder (Figure 3(a)), it is
possible to make out the widely dispersed, so-called bee or catana phase, along with the so-called
per-phase that disperses the catana phase, as the darker material surrounding each ‘bee’. The
per-phase is then observed to be supported by the lighter coloured para- or sal-phase, which are
considered to be the least stiff, or continuous solvent phases (Lyne, Wallqvist, Rutland, Claes-
son, & Birgisson, 2013; McCarron, Yu, Tao, & Burnham, 2012). The maximium phase angle
observed here was approximately 90° (Pauli, Grimes, Beemer, Turner, & Branthaver, 2011). It
was interesting to note, that in the 180–200 grade samples seen in Figure 3, this phase angle
remained reasonably consistent across the samples with different blending treatments. As the
degree of blending increased, it was possible to identify islands of stiffer or harder material that
became more widely dispersed. In Figure 3(c) and 3(d), the light patches with phase angles
around 85° and 95° might be ascribed to dispersed nanoclay particles that are emulsified by the
bitumen per-phase and that catana phases have been dispersed into the continuous phase. The
homogeneity of the para- and sal-phases appears to have been improved by the shear mixing.
The image in Figure 3(e) was somewhat less clear, however might be possible to suggest that the
stiffer particles are further dispersed and showing formation of new catana phase at the surface.
While it seems clear that AFM analysis did not show unambiguously the presence of nanoclay
particles in the surface of the bitumen, it is nevertheless possible to suggest that the structure of
the bitumen was modified by blending of nanoclay into the bitumen.

4.3.2. 80–100 Grade bitumen


The AFM Phase images of the 80–100 grade bitumen samples with nanoclay blended for differ-
ent times are presented in Figure 4. The images show different behaviour to that of the 180–200
grade material. The base binder (Figure 4(a)) shows small numbers of large catana phase (some-
what distorted in the image), with a narrow spread of phase angle 9.5–37°. This suggests that
the surface of the 80–100 grade base binder is homogeneous with reasonably uniform stiffness.
12 S.A. Bagshaw et al.

(a) (b)

(c) (d)

(e)

Figure 3. AFM micrographs of Nanomer/bitumen samples, phase mode. (a) Base 180–200 bitumen; (b)
paddle mix 15 min; (c) shear blended 15 min, 4000 rpm; (d) shear blended 30 min, 4000 rpm and (e) shear
blended 60 min, 4000 rpm.
Note: Due to an experimental artefact, the phase angle scale in 3A is the inverse of the other images.

Paddle mixing of nanoclay into the bitumen (Figure 4(b)) modified the stiffness of the sample sig-
nificantly, increasing the maximum to 129°. This might suggest, therefore, that nanoclay particles
were found throughout the sample, but that they were large and not well dispersed. The remaining
figures (Figure 4(c)–(e)) appear to show the dispersion and breaking down of the nanoclay par-
ticles within the para- and sal-phases of the 80–100 grade bitumen. The maximum phase angles
and ranges decreased in an approximately linear fashion and the domains of the stiffer/harder
components reduced to small points dispersed within the continuous bitumen phase.
Overall, it is possible to suggest that AFM imaging can identify modifications to the bitumen
structure caused by blending of nanoclay into the bitumen. It is, however, possibly less rigorous
to suggest that the imaging of individual, or dispersed, clay particles is possible.
Road Materials and Pavement Design 13

(a) (b)

(c) (d)

(e)

Figure 4. AFM micrographs of Nanomer/bitumen samples, phase mode. (a) Base 80–100 bitumen; (b)
paddle mix 15 min; (C) shear blended 15 min, 4000 rpm; (d) shear blended 30 min, 4000 rpm and (e) shear
blended 60 min, 4000 rpm.

4.4. FTIR
FTIR spectra of the variously blended samples were analysed to see if it is possible to identify
structures that change due to the blending regimes. One of the distinct challenges of probing
the structures of bituminous materials to which have been added organo-modified materials is
that the vibrational bands associated with the organic functions in the bitumen are, in the main
part, coincident with the bands of the organic function of the Nanomer material. The vibrational
bands that are not coincident are those associated with the clay silicate material itself, which are
perhaps less likely to be modified by bitumen blending.
Following blending of the alkyl-ammonium functionalised Nanomer into the bitumen, it was
interesting to observe the presence of one new vibration at 1080 cm−1 . This band was not
14 S.A. Bagshaw et al.

(a) (b)

Figure 5. FTIR spectra of 6 wt% Nanomer/80–100 grade binder after shear mixing at 150°C, 4000 rpm.
(a) CH2 and CH3 stretching region and (b) C–C aromatic and silicate structural region.

observed in either the original Nanomer or the base bitumen. The band might be due either to
the creation of oxidised bitumen species or to an interaction between the nanoclay and the bitu-
men. The intensity of the new band did not change with blending conditions, but experiments
with different loadings of Nanomer (not presented here) indicated that the intensity of the band
increased with increased loading of Nanomer. The band might, therefore, be assigned tentatively
to an interaction between the nanoclay and the bitumen, the nature of which remains unclear at
this time.
Any other changes to spectra due to the different blending regimes were very subtle. While
no notable shifts of band positions were observed, some changes in the intensities of bands were
observed. The bands in the CH2 and CH3 stretching region from 2700–3100 cm−1 were all seen
to increase in intensity following blending of 6 wt% Nanomer into the base bitumen (Figure 5(a)).
It is possible that the changes in absorbance may be due to the increased concentration of long-
chain alkyl species (i.e. CH2 and CH3 species) that occurs upon blending of the Nanomer into
the bitumen. More widely dispersed Nanomer particles might contribute more strongly to this
portion of the spectrum given that more clay surface modifier species will absorb. This is perhaps
what is observed in Figure 5(a), but the non-linearity of the response might suggest that other
factors such as sub-sampling issues are present. In the 450–1600 cm−1 region, where the organic
and inorganic structural deformation bands are found, the Si–O and Al–O vibrations due to the
alumino-silicate clay structure at 960–1140 cm−1 and 450–650 cm−1 were added to the spectrum
of the bitumen (Figure 5(b)). However, no other significant differences were observed in any of
the spectra, other than some small intensity differences.
The FTIR spectra, therefore, very clearly identified the addition of Nanomer to the bitumen.
However, owing to the coincident natures of the spectra of the bitumen and of the intercalated
organic function of the Nanomer, it was possible to identify only one new vibration suggesting
some form of interaction between the bitumen and the Nanomer, the nature of which is unclear.

4.5. Penetration – stiffness modification


The penetration test remains a common method with which to probe the relative stiffness of
bitumen samples, lower penetration corresponds to higher stiffness. Penetration was examined
to see if the different Nanomer blending regimes would influence the penetration behaviour of
the bitumen. All penetration measurements were done in triplicate following the ASTM standard
method (ASTM, D5, 2013).
The penetration values for 180–200 and 80–100 base binders presented in Figure 6 were mea-
sured following paddle stirring of those materials without the addition of nanoclay. This resulted
in penetration values that were very slightly below the nominal standard values, presumably due
Road Materials and Pavement Design 15

(a) (b)

(c) (d)

Figure 6. 25°C penetration test values of 6 wt% Nanomer/bitumen after shear mixing at 150°C. (a)
180–200, 2000 rpm; (b) 180–200, 4000 rpm; (c) 80–100, 2000 rpm and (d) 80–100, 4000 rpm.

to slight aging that possibly occurred during the high temperature stirring treatment. In contrast,
blending of Nanomer into the bitumens, either by paddle mixing or by shear mixing, induced
significantly reduced penetrations in all of the nanocomposite samples. Penetration into the flex-
ible 180–200 grade binder (Figure 6(a) and 6(b)) was reduced by 28 ± 3%, while penetration
into the stiffer 80–100 grade (Figure 6(c) and 6(d)) was reduced by a smaller 16 ± 5%.
The penetration results indicated real and significant changes to the stiffness of the bitumens
following nanoclay blending. However, the magnitudes of the differences between samples after
different blending times and speeds were small and irregular, 3.5–5.5%. While this result supports
the XRD analysis, that nanoclay dispersion into the bitumen is rapid, the result also appears
to suggest that increasing the dispersion or intercalation of the nanoclay was not reflected in
additional stiffness over dispersion of the nanoclay by simple paddle mixing.
The ability of the Nanomer to remain suspended in the bitumen during storage or transport is a
very important property of the material. Segregation stability was probed by penetration testing
after static heating at 160°C for 60 h. For the 180–200 grade bitumen, the average penetration dif-
ference was 4.6 ± 0.4%. For the 80–100 grade bitumen, the average penetration difference was
3.3 ± 0.8%. These results suggest, therefore, that the Nanomer blended bitumens were generally
stable towards segregation during storage.

4.6. Rotational viscosity


Viscosity is an important property in terms of both high-temperature delivery and flow at ele-
vated in-service temperatures. Addition of organo-nanoclay particles to the bitumen is expected
to increase the amount of energy required for bitumen molecules to flow at any given rate, i.e.
to increase the viscosity (Symon, 1971). It is, however, not clear how different mixing regimes
or degrees of dispersion will impact the bitumen viscosity. It might be assumed that if the nan-
oclay particles are more widely dispersed they ought to contribute more strongly to increase the
resistance to molecular shear flow.
16 S.A. Bagshaw et al.

(a) (b)

(c) (d)

Figure 7. Rotational viscosity values of 6 wt% Nanomer/bitumen after shear mixing at 150°C. (a)
180–200, 2000 rpm; (b) 180–200, 4000 rpm; (c) 80–100, 2000 rpm and (d) 80–100, 4000 rpm.

Results of rotational viscosity measurements at 135°C are presented in Figure 7. The results
indicate that addition of Nanomer increased dramatically the viscosity of the bitumen. An
increase in viscosity, over that of the base material, was attained in all samples regardless
of the mixing conditions; however, subtle differences were observed with different mixing
conditions. Paddle mixing of Nanomer into both grade bitumens increased the viscosity by
approximately 100%. The viscosities of the 180–200 samples blended by high-speed shear
mixing increased by between 400 and 500 ± 20% (Figure 7(a) and 7(b)), while the increases
in the 80–100 samples blended by high-speed shear mixing were more modest at 150–
250 ± 20% (Figure 7(c) and 7(d)). Observationally, the addition of 6 wt% Nanomer to both
the 180–200 and 80–100 materials transformed what were viscous fluids into weak gel-like
materials, indicating the creation of some form of extended macro-structure within the bitumen
matrix.
In the 180–200 grade material, the viscosities of all of the 2000 rpm shear mixed samples
were approximately the same, while those of the samples shear mixed at 4000 rpm fell as the
mixing time increased. For the 80–100 grade material, however, the trend was reversed with the
viscosities falling with longer shearing time at 2000 rpm and remaining approximately the same
at 4000 rpm. In contrast to the penetration results, shear mixing clearly increased the viscosities
over those of paddle-mixed samples, but the observed values showed more scatter. It might be
argued that the differences in measured values of the shear mixed samples were not greatly sig-
nificant and that perhaps a similar argument might be made that shear mixing beyond 2000 rpm
and 15 min does not produce significantly different viscosities.
It would appear therefore that it was not necessary to blend the nanoclay for either long periods
of time or at high speeds to achieve dispersions that resulted in increased bitumen viscosities.
Paddle mixing achieved approximately 100% increase in viscosity of both 180–200 and 80–
100 grade bitumens, while blending at 2000 rpm for 15 min was sufficient to achieve viscosity
increases up to 500%.
Road Materials and Pavement Design 17

(a) (b)

Figure 8. Isochronal plots of complex modulus vs. temperature for 6 wt% Nanomer/bitumen nanocom-
posites tested: (a) 180–200 grade binder and (b) 80–100 grade binder.

(a) (b)

Figure 9. Phase angle values of 6 wt% Nanomer/bitumen samples: (a) 180–200 grade binder and (b)
80–100 grade binder.

4.7. Dynamic shear rheometry


Results from analyses of the DSR measurements are presented in Figures 8 and 9.

4.7.1. Complex modulus – |G*|


For the 180–200 grade composites (Figure 8(a)), the values for complex modulus showed a
linear reduction over the temperature range investigated and all of the nanocomposite samples
exhibited increased modulus over the base binder sample. However, at each temperature, the
results for the differently blended materials were generally very much the same. This suggested
that different blending regimes appeared to have little effect on the moduli of the materials. In
addition to the base binder, the sample blended at 4000 rpm for 60 min exhibited lower complex
moduli. This result is perhaps in line with the viscosity results, and suggests that blending of
the Nanomer into the bitumen at the highest speed for the longest time appears to reduce the
influence of the Nanomer on the rheological properties of the bitumen.
For the 80–100 grade bitumen (Figure 8(b)), the result was more straightforward. All the
nanocomposite samples exhibited higher complex moduli than the base binder and were similar
at each temperature. This appears to more closely reflect the rotational viscosity measurements.
Overall, the complex modulus measurements indicated that the nanocomposite bitumens were
more resistant to deformation than the base binders, across the temperature range investigated
here. However, they also suggested that there were only minor differences between the differently
blended samples.

4.7.2. Phase angle – δ


The phase angle responses followed similar trends to those of the complex moduli, with all
nanocomposite samples exhibiting lower phase angles than the base binders and hence higher
18 S.A. Bagshaw et al.

(a) (b)

Figure 10. Rutting parameters and failure temperatures of 6 wt% Nanomer/bitumen samples: (a) 180–200
grade and (b) 80–100 grade binders. Black and Red dashed lines represent the data extrapolated to 1.0 kPa
for the two separate groups of samples.

elasticity. Overall, the phase angles increased with increasing temperature as would be expected,
but the addition of Nanomer shifted the phase angle responses such that the materials behaved
more elastically across the temperature range. For the 180–200 grade binder (Figure 9(a)), the
sample blended at 4000 rpm for 60 min exhibited a higher phase angle across the temperature
range, while the base binder only appeared to separate from the group at 55°C. All the other
samples fell into the group at each temperature with the very narrow spread.
For the stiffer 80–100 grade binder (Figure 9(b)), all of the Nanomer modified binders exhib-
ited phase angles that were very clearly separate from that of the base binder, across the entire
temperature range, and the separation became wider as the temperature increased.

4.7.3. Rutting parameter and failure temperature


Rutting parameter, G*/sinδ and failure temperature values are presented in Figure 10. The results
show the formation of the nanocomposites increased the resistance to permanent deformation. At
the lowest temperature of 5°C, the rutting parameters of all samples, including the base binders,
were approximately the same, but not identical. As the temperature increased, the rutting param-
eters of the nanocomposite samples diverged to higher values from those of the base binders,
which indicated that the formation of nanocomposites produced materials with increased resis-
tance to rutting over the base binders. This result matched data obtained by other workers for
other nanocomposite materials (Yang & Tighe, 2013; Yao et al., 2013; Zhang, van de Ven, &
Wu, 2012; Zhang, Zhu, Yan, & Yu, 2015). But as with much of the other data reported here, there
appears to be minimal difference between materials prepared under different blending conditions.
The highest temperature investigated here was 55°C; therefore, the failure temperature has
been interpolated at 1 kPa. For the 180–200 grade base bitumen, the failure temperature occurred
at approximately 68°C, while the failure temperature for all of the blended nanocomposite
samples occurred at approximately 75°C. The stiffer 80–100 grade binder exhibited a slightly
higher failure temperature, with the base binder failure temperature at approximately 74°C and
the nanocomposites failure temperature at approximately 80°C. The results show clearly that
formation of nanocomposites increases the failure temperature by around 6°C. This may be a
significant difference at in-service temperatures. However, the results obtained here also suggest
that different blending regimes had no appreciable further effect on the failure temperature.

4.7.4. Fatigue parameter


Fatigue parameters obtained at 5°C, 20°C and 55°C are presented in Figure 11, but unlike the
Superpave method (1.59 Hz), the fatigue parameters were calculated under the same constant
Road Materials and Pavement Design 19

(a) (b)

Figure 11. Fatigue parameters of 6 wt% Nanomer/bitumen samples: (a) 180–200 grade and (b) 80–100
grade binders. Black dashed lines represent the fatigue parameters at 10°C. Red dashed lines represent the
temperatures at 5000 kPa.

loading frequency of 9.042 Hz as for the other rheology data presented here. The results indi-
cate that 180–200 grade samples possess fatigue parameters greater than 5000 kPa until they
attain approximately 17°C, while for the stiffer 80–100 grade samples this value is attained
at approximately 25°C. The fatigue parameters for all samples declined, and the values for
the nanocomposite samples diverged from the values for the base binders, as the temperatures
increased. In some nanocomposite materials reported elsewhere, the fatigue parameter at low
temperatures was lower than that of the base binder; however, that was not the case in this work
(Ali et al., 2017).
In line with the other rheology results obtained here, there appears to be no significant dif-
ferences between the fatigue parameter values for the nanocomposites prepared under different
blending regimes.

5. Conclusions
Should clay-bitumen nanocomposites reach the stage of deployment into the roading industry,
it will be important to optimise the cost/performance ratio of their manufacture. In this work,
we have sought to understand if significant differences can be identified between bitumen–clay
nanocomposites that are prepared under different blending conditions, namely simple paddle
mixing, and high-shear mixing, for different times and speeds. While this work was performed
with one commercial organoclay product and two base bitumen binders, there are of course other
products in the marketplace which may perform differently. The following points summarise the
conclusions from this study:

• Good dispersion of Nanomer into ‘sol’ and ‘gel’ type bitumens is achieved by simple
paddle stirring over a short time.
• This dispersion significantly modifies the physical and rheological properties of the two
different grades of bitumen. The softer ‘sol’ type binder was affected more than the harder
‘gel’ type binder.
• Addition of Nanomer causes obvious long-range structuring of the bitumen through gel
formation.
• Increased dispersion via high-energy shear mixing for short times increases the degree
of dispersion and intercalation of the Nanomer particles and increases the effect on some
properties (e.g. penetration, rotational viscosity) over that achieved by paddle mixing.
• There appears to be little further change in the properties of the nanocomposites subjected
to extended or higher speed shear blending over those prepared at low speeds and short
times.
20 S.A. Bagshaw et al.

• There appears to be little difference in the rheological performances between samples


blended by high shear over those blended by paddle stirring.
• It would appear therefore, that tuning of the physical and rheological properties of
bitumen–clay nanocomposite structures may be possible with different blending regimes.
High-energy, high-cost shear mixing may not necessary for the manufacture of bitu-
men/organo nanoclay nanocomposites with significantly improved properties, simple
low-energy mixing procedures may be sufficient.

The following tentative hypothesis is introduced to explain these observations. The planar
and organophilic organoclay has a physical and chemical presentation that very loosely resem-
bles that of asphaltene compounds and stacks thereof. Given these loose similarities, it might
be expected that organoclay platelets or tactoids might induce physical properties into a bitumen
continuous phase similar to an increase in asphaltene content. The high molecular weight asphal-
tene composition of a bitumen is the primary determinant of the rheological properties of that
bitumen, so that dispersion of organoclay into a bitumen acts in a similar manner to that of an
increase in the asphaltene content. It is therefore possible to modify the ‘grade’ of a given base
bitumen through the addition, via low-energy blending techniques, of organoclay additives. In
this way, it may be possible to tune the properties of a bitumen to provide very exact intermediate
grades and bitumens that behave differently in different environments.

Acknowledgements
The authors acknowledge Nanocor Inc. for the supply of organoclay samples used in this work. We also
appreciate the assistance provided by technical staff at Callaghan Innovation.

Disclosure statement
No potential conflict of interest was reported by the authors.

Funding
The authors acknowledge the Ministry of Business, Innovation and Employment (MBIE), New Zealand for
research funding through project OPSX1501 – Waterproof Roads.

References
Aguiar-Moya, J. P., Salazar-Delgado, J., Bonilla-Mora, V., Rodriguez-Castro, E., Leiva-Villacorta, F., &
Loria-Salazar, L. (2015). Morphological analysis of bitumen phases using atomic force microscopy.
Road Materials and Pavement Design, 16(S1), 138–152.
Ali, A. I. A., Ismail, A., Karim, M. R., Yusoff, N. I. M., Al-Mansob, R. A., & Aburkaba, E. (2017). Per-
formance evaluation of Al2 O3 nanoparticle-modified asphalt binder. Road Materials and Pavement
Design, 18(6), 1251–1268.
Allen, R. G., Little, D. N., Bhasin, A., & Lytton, R. L. (2013). Identification of the composite relaxation
modulus of asphalt binder using AFM nanoindentation. Journal of Materials in Civil Engineering,
25(4), 530–539.
Anthony, J. W., Bideaux, R. A., Bladh, K. W., & Nichols, M. C. (Eds.). (1995). Montmorillonite. Handbook
of mineralogy II (silica, silicates). Chantilly, VA: Mineralogical Society of America.
Arora, A., & Padua, G. W. (2010). Review: Nanocomposites in food packaging. Journal of Food Science,
75(1), R43–R49.
Ashish, P. K., Singh, D., & Bohm, S. (2017). Investigation on influence of nanoclay addition on rheological
performance of asphalt binder. Road Materials and Pavement Design, 18(5), 1007–1026.
ASTM D113-17. (2017). Standard test method for ductility of asphalt materials. West Conshohocken, PA:
ASTM International.
Road Materials and Pavement Design 21

ASTM D36/D36M-14e1. (2014). Standard test method for softening point of bitumen (ring-and-ball
apparatus). West Conshohocken, PA: ASTM International.
ASTM D5/D5M. (2013). Standard test method for penetration of bituminous materials. West Con-
shohocken, PA: ASTM International.
ASTM D70-17. (2017). Standard test method for density of semi-solid bituminous materials (Pycnometer
method). West Conshohocken, PA: ASTM International.
Brown, D. N., Henning, T. F. P., Herrington, P. R., & Wu, J. (2015). Chipseal cracking. NZTA Research
Report 579.
Cavalcanti, J. V. F. L., Abreu, C. A. M., Carvalho, M. N., Motta Sobrinho, M. A., Benachour, M.,
& Baraúna, O. S. (2012). Removal of effluent from petrochemical wastewater by adsorption using
organoclay. In V. Patel (Ed.), Petrochemicals (pp. 277–294). Rijeka: InTech. doi:10.5772/37200
D’Angelo, T. (2002). Performance specifications for asphalt bitumens in the US. Plenary lecture at the 39th
Annual Peterson asphalt research conference, Laramie, Wyoming, July 15–17.
Das, P. K., Baaj, H., Tighe, S., & Kringos, N. (2016). Atomic force microscopy to investigate asphalt
binders: A state-of-the-art review. Road Materials and Pavement Design, 17(3), 693–718.
Fang, C., Yu, R., Liu, S., & Li, Y. (2013). Nanomaterials applied in asphalt modification: A review. Journal
of Material Science and Technology, 29(7), 589–594.
Galooyak, S. S., Dabir, B., Nazarbeygi, A. E., Moeini, A. E., & Berahman, B. (2011). The effect of nan-
oclay on rheological properties and storage stability of SBS-modified bitumen. Petroleum Science and
Technology, 29(8), 850–859.
Ghile, D. B. (2006). Effects of nanoclay modification on rheology of bitumen and on performance of asphalt
mixtures. MSc Thesis, Delft University of Technology.
Gransberg, D., & James, D. M. B. (2005). Chipseal best practices. NCHRP synthesis 342. Washington, DC:
National Cooperative Highways Research Programme.
Hamedi, G. H., Moghadas Nejad, F., & Oveisi, K. (2015). Investigating the effects of using nanomaterials
on moisture damage of HMA. Road Materials and Pavement Design, 66, 51–59.
Herrington, P. R., Kodippily, S., & Henning, T. F. P. (2015). Flushing in Chipseals. NZTA Research Report
576.
Howard, I. L., Cox, B. C., Alvarado, A., & Jordan, W. S. (2017). Material selection and traffic open-
ing guidance for chip or scrub seals. Road Materials and Pavement Design, 1436(3), 2017, 1–23.
doi:10.1080/14680629.2017.1340327
Hu, H., Onyebueke, L., & Abatan, A. (2010). Characterising and modelling mechanical properties
of nanocomposites-review and evaluation. Journal of Minerals & Material Characterisation &
Engineering, 9(4), 275–319.
Jahangir, R., Little, D., & Bhasin, A. (2015). Evolution of asphalt binder microstructure due to tensile
loading determined using AFM and image analysis techniques. International Journal of Pavement
Engineering, 16(4), 337–349.
Jahromi, S. G., Andalibizade, B., & Vossough, S. (2010). Engineering properties of nanoclay modified
asphalt concrete mixtures. Arabian Journal of Science and Engineering, 35(1B), 89–103.
Jahromi, S. G., & Khodaii, A. (2009). Effects of nanoclay on the rheological properties of bitumen binder.
Construction and Building Materials, 23, 2894–2904.
Karşal, C., Tanoğlu, M., Odabaş, S., Ersoy, O. G., & Karakaya, N. (2009). Effect of blending conditions on
the properties of EPDM/organoclay nanocomposites. 17th conference on composite materials, ICCM-
17, Edinburgh.
Kattak, M. J., Khattab, A., Rizvi, H. R., & Zhang, P. (2012). The impact of carbon nano-fibre modification
on asphalt binder rheology. Construction and Building Materials, 30, 257–264.
Kök, B. V., & Akpolat, M. (2015). Effects of using Sasobit and SBS on the engineering properties of
bitumen and stone mastic asphalt. Journal of Materials in Civil Engineering, 27(10), 04105006.
Lesueur, D. (2009). The colloidal structure of bitumen: Consequences on the rheology and on the
mechanisms of bitumen modification. Advanced in Colloidal and Interface Science, 145(1-2), 42–82.
Liu, G., Wu, S., van de Ven, M., Yu, J., & Molenaar, A. (2010). Influence of sodium and organo-
montmorillonites on bitumen. Applied Clay Science, 49, 69–73.
Lyne, A. L., Wallqvist, V., Rutland, M. W., Claesson, P., & Birgisson, B. (2013). Surface wrinkling: The
phenomenon causing bees in bitumen. Journal of Materials Science, 48(20), 6970–6976.
Masson, J. F., Leblond, V., & Margeson, J. (2006). Bitumen morphologies by phase-detection atomic force
microscopy. Journal of Microscopy, 182(1), 32–39.
McCarron, B., Yu, X., Tao, M., & Burnham, N. (2012). The investigation of ‘Bee-structures’ in asphalt
binders. Worcester, MA: Qualifying Project.
22 S.A. Bagshaw et al.

McGennis, R. B., Shuler, S., & Bahia, H. U. (1994). Background of superpave asphalt binder test methods.
Lexington, KY: Federal Highway Administration, Office of Technology Applications.
Moghadas Nejad, F., Azorhoosh, A., & Hamedi, G. H. (2014). Effect of high density polyethylene on the
fatigue and rutting performance of hot mix asphalt-a laboratory study. Road Materials and Pavement
Design, 15(3), 746–756.
Nazzal, M. D., Kaya, S., Gunay, T., & Ahmedzade, P. (2013). Fundamental characterisation of asphalt clay
nanocomposites. Journal of Nanomechanics and Micromechanics, 3, 1–8.
Neaylon, K. L., & Harrow, L. (2017). Lessons to be learned from 15 year old second coat seals and reseals.
NZTA Research Report 612.
New Zealand Ministry of Business, Innovation and Employment (MBIE). (2016). New Zealand for research
funding through project OPSX1501 – Waterproof Roads.
New Zealand Transport Agency. (2005a). Chipsealing in New Zealand. Transit New Zealand. Wellington:
Author.
New Zealand Transport Agency. (2005b). Chipsealing in New Zealand. Transit New Zealand. Wellington:
Author.
Olad, A. (2011). Polymer/clay nanocomposites. In B. Reddy (Ed.), Advances in diverse industrial
applications of nanocomposites (pp. 113–138). Rijeka: InTech.
Osman, M. A., Mittal, V., & Suter, U. W. (2007). Poly(propylene)-layered silicate nanocomposite: Gas
permeation properties and clay exfoliation. Macromolecular Chemistry and Physics, 208(1), 68–75.
Paul, D. R., & Robeson, L. M. (2008). Polymer nanotechnology: Nanocomposites. Polymer, 49(15), 3187–
3204.
Pauli, A. T., Grimes, R. W., Beemer, A. G., Turner, T. F., & Branthaver, J. F. (2011). Morphology of asphalts,
asphalt fractions and model wax-doped asphalts studied by atomic force microscopy. International
Journal of Pavement Engineering, 12(4), 291–309.
Rahman, F., Shahidul Islam, M., Musty, H., & Hossain, M. (2012). Aggregate retention in chipseal.
Transportation Research Record: Journal of Transportation Research Board, 2267, 56–64.
Read, J., & Whiteoak, D. (2003). The Shell bitumen handbook. London: Thomas Telford.
Rebelo, L. M., Cavalcante, P. N., de Sousa, J. S., Mendes Fliho, J., Soares, S. A., & Soares, J. B.
(2014). Micromorphology and microrheology of modified bitumen by atomic force microscopy. Road
Materials and Pavement Design, 15(2), 300–311.
Redelius, P. G. (2006). The structure of asphaltenes in bitumen. Road Materials and Pavement Design,
7(Suppl. 1), 143–162.
Stefanescu, E. A. (2008). A study of rheological and thermodynamic properties of polymer-clay gels and
multi-layered films. PhD thesis, Louisiana State University.
Symon, K. (1971). Mechanics. 3rd ed. Boston: Addison-Wesley.
Weigel, S., & Stephan, D. (2017). Relationships between the chemistry and physical properties of bitumen.
Road Materials and Pavement Design, 466(1), 1–15. doi:10.1080/14680629.2017.1338189
Yang, J., & Tighe, S. (2013). A review of advances of nanotechnology in asphalt mixtures. Procedia-Social
and Behavioural Sciences, 96, 1269–1276.
Yao, H., You, Z., Li, L., Lee, C. H., Wingard, D., Yap, Y. K., . . . Goh, S. W. (2013). Rheological properties
and chemical bonding of asphalt modified with nanosilica. Journal of Materials in Civil Engineering,
25, 1619–1630.
You, Z., Mills-Beale, J., Foley, J. M., Roy, S., Odegard, G. M., Dai, Q., & Goh, S. W. (2011). Nanoclay-
modified asphalt materials: Preparation and characterisation. Construction and Building Materials, 25,
1072–1078.
Zhang, J., van de Ven, M., & Wu, S. (2012). Morphology and rheological analysis of nanoclay in polymer
modified bitumen. Key Engineering Materials, 509, 155–161.
Zhang, H., Zhu, C., Yan, K., & Yu, J. (2015). Effect of rectorite and its organic modification on properties
of bitumen. Journal of Materials in Civil Engineering, 27(8), C4014002.

You might also like