You are on page 1of 31

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

net/publication/254314270

Glass Transition and Phase Stability in Asphalt Binders

Article  in  Road Materials and Pavement Design · January 2008


DOI: 10.1080/14680629.2008.9690158

CITATIONS READS

27 1,020

3 authors, including:

Pavel Kriz Jiri Stastna


Imperial Oil, Canada The University of Calgary
13 PUBLICATIONS   236 CITATIONS    85 PUBLICATIONS   1,344 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

otakar D(t) View project

Reducing variability of DSR test View project

All content following this page was uploaded by Pavel Kriz on 11 February 2014.

The user has requested enhancement of the downloaded file.


Glass Transition and Phase Stability in
Asphalt Binders

Pavel Kriz* — Jiri Stastna* — Ludo Zanzotto*

*Bituminous Materials Chair


Schulich School of Engineering
University of Calgary
2500 University Dr. NW
Calgary, AB, T2N 1N4
Canada
pkriz@ucalgary.ca
stastna@ucalgary.ca
zanzotto@ucalgary.ca

ABSTRACT. Major aspects of the glass transition of asphalt binders are described and an
extensive literature review of the phenomenon and its relation to chemical composition is
presented.
The glass transition of asphalt binders was studied by modulated differential scanning
calorimetry and also via dynamic mechanical analysis. A certain analogy between the glass
transition of amorphous polymers and asphalts is suggested. The overall transition was found
to be very broad on the temperature scale.
The effects of evaporation of light-end components and oxidation on asphalt phase stability
and glass transition were studied. It was suggested that phase incompatibility may exist in
asphalts; however, the phase separation is observable after long-term isothermal
conditioning at a temperature within the glass transition range.
Based on the presented results, it is suggested that phase incompatibility develops if there is a
discontinuity in the molecular distribution. Such discontinuity may be present in some neat
binders as well as in severely oxidized or aged asphalt binders.
KEYWORDS: glass transition, modulated differential scanning calorimetry, asphalt binder,
molecular mobility, physical aging, oxidative aging.

Road Materials and Pavements Design. Volume X – No X/2007, pages 1 to n


2 Road Materials and Pavements Design. Volume X – No X/2007

1. Introduction

The glass transition is, perhaps, the most important factor that determines the
viscoelastic properties of amorphous material (Williams et al., 1955, Ngai, 2004). It
is a reversible change in an amorphous domain from a viscous or rubbery state to a
hard and relatively brittle glassy state, and vice versa. The glass transition occurs
when the characteristic time of molecular motions responsible for structural
rearrangements becomes longer than the timescale of the experiment. The timescale
for structural relaxation increases rapidly with decreasing temperature (Moynihan,
1994). Therefore, the glass transition is usually considered as a predominantly rate
and non-equilibrium phenomenon.
The transition to a glassy state is accompanied by a sudden change in the
mechanical, optical and thermodynamic properties of the material. The increase in
viscosity is enormous (up to 1013 Pa.s). The material becomes glossy in appearance
and extremely brittle and rigid. The thermodynamic change during the glass
transition upon cooling is recognized as the stepwise drop in heat capacity and loss
in free volume (Ngai, 2004). Note, that the free volume is not a real volume
measurable in cm3/g, but an operational quantity dealing with the molecular mobility
(Struik, 1991).
The glass transition temperature (Tg) is the temperature arbitrarily chosen to
represent the temperature range over which the glass transition takes place: it is
usually assigned to the temperature where half of the sample is already vitrified or
devitrified (Wunderlich, 1994).
In the current study, the glass transition of asphalt is thoroughly described and
investigated by two techniques – modulated differential scanning calorimetry
(MDSC) and dynamic mechanical analysis (DMA). The similarities between the
glass transition of amorphous polymer and asphalt are suggested.

2. Background

The glass transition is a function of molecular mobility. Upon cooling, molecular


mobility decreases. Molecules are considered “glassy”, if their molecular motion has
appeared to cease with respect to observation time. Different molecules have
different degrees of motion at particular temperatures. The molecular motion is a
function of molecular weight and structure, as well as of inter- and intra-molecular
interactions. Before the glass transition is studied, the asphalt composition and inter-
molecular relations should be understood.
Glass Transition in Asphalt Binders 3

2.1. Molecular Continuity

Asphalt consists of tens of thousands different chemical species (Speight, 1999).


The whole scale of molecules ranges from non-polar fully saturated linear alkanes to
highly polar polycyclic porphyrins or hetero-hydrocarbons. Each molecular
functional group is essential to the phase stability of the asphalt system. It is,
therefore, necessary to view the material as a continuum of molecules with a gradual
transition in polarity, molecular weight and functionality.
It is a generally accepted fact that the molecular functional groups overlap and,
thus, are impossible to separate (Andersen et al., 2001). It has been reported
(Carlson et al., 1958) that there is no sharp transition between asphaltenes and
aromates. It has also been shown that part of the asphaltene fraction (based on the
solubility parameter) may cosolubilize the rest of the fraction and, thus, behave as a
resin (Heithaus, 1962, Andersen, 1995). There is no sudden transition in the
molecular structural properties between resins and asphaltenes (Groenzin et al.,
1999). Due to the complexity of the material and the interactions between high
molecular weight molecules, the commonly used methods for fraction separation of
asphalt cannot provide well-defined chemical fractions (Selucky et al., 1981). If part
of the material is removed from the system during the separation, the molecular
distribution and interaction are affected. The solubility of a substance is not a
function of its general hydrocarbon skeleton and its chemical functionality alone, but
also depends on interactions with other substances that act as co-solvents. Whether a
given compound appears, for instance, as a resin or an asphaltene depends upon the
presence or absence of other substances (Farcasiu, 1977).
Separation into fractions has not proved to be a reliable predictor of asphalt
performance on test roads (Goodrich et al., 1986). Indeed, asphalt refining is based
on another type of fraction separation – distillation (Corbett, 1984). For the sake of
simplicity, we may still use the terms “asphaltenes”, “resins”, and so on; but instead
of relating these terms to solubility, we relate them to their functionality in the
system. The asphaltenes represent the polyaromatic fraction, which tends to be
insoluble in the continuous phase at certain conditions and needs to be stabilized
(cosolubilized). This fraction consists of different molecules at different conditions
(temperature, pressure, change in the continuous phase composition). Resins are
unique aromates that are not qualitatively different from asphaltenes, yet restrict the
asphaltenes from further association and, due to complex interaction, keep them in a
single phase providing the molecular transition between the polar and non-polar
ends. And finally, the continuous (or oil) phase, with composition gradually ranging
from semi-polar resins to non-polar paraffins, is present.
4 Road Materials and Pavements Design. Volume X – No X/2007

2.2. Asphalt Structure

The concept of asphalt being a colloidal system was proposed by Nellensteyn


(Nellensteyn, 1938). A colloidal system is an intermediate between a homogenous
solution and a heterogenous system, with the particle size ranging from 1 nanometer
to 1 micrometer (Hiemenz et al., 1997). There is strong experimental evidence that
the biggest single molecules in asphalt do not exceed the molecular weight of 1000
g/mol (McKay et al., 1978, Groenzin et al., 2001), which corresponds to molecular
diameters in the range of 1-2 nm (Katz et al., 1945, Groenzin et al., 1999). The most
polar molecules usually associate into bigger particles. The size of these particles
rarely exceeds 6 nm (van der Hart et al., 1990); however, if the system is
destabilized, they may grow bigger and eventually separate from the system. The
presence of molecules and aggregates bigger than 1 nm disqualify asphalt from
being a homogenous solution. The upper limit of a colloidal system (1 μm) is
generally met in asphalts, even though recent atomic microscopy experiments
revealed different domains of a size exceeding 1μm (Masson et al., 2006a).
The molecular interactions (especially between asphaltenes and resins) and the
whole structure of the system are still topics for discussion. The lyophobic models,
where the insoluble asphaltenes are peptized by the resins that provide a protective
sheet, were actually motivated by early observations of the solubilizing power of
resins (Mack, 1932, Nellensteyn, 1938, Eilers, 1949, Koots et al., 1975). This is
inconsistent with the fact that the asphaltene precipitates can be re-dissolved in a
good solvent, even in the absence of resins (Andersen et al., 1990, Andersen et al.,
1991). Also, the asphaltene molecules are not many times larger than the size of the
resin molecules, as they would logically have to be in order for many resin
molecules to be able to fit around them and form the protective layer, similar to the
water-surfactant-oil emulsion. The asphaltene and resin molecules are, in fact, not
much different in size (Storm et al., 1995, Porte et al., 2003). Whether a particular
molecule behaves as an asphaltene or resin always depends on actual conditions
(temperature, pressure, surrounding molecules, etc.).

2.2.1. The Molecular Forces


The interaction between the asphaltene molecules and/or resin molecules is
crucial in order to keep the colloidal character of the material and prevent further
aggregation towards bigger agglomerates, which would eventually lead to instability
and potential phase separation. There is no reason to assume that the inter-molecular
forces acting between asphaltenes, resins and the oil phase should be in any way
different from those existing between other well-known organic molecules
containing the same atomic groups. Inter- and intra-molecular forces affecting the
phase stability and molecular structure are charge transfer forces, electrostatic
interactions, van der Waals interaction, exchange-repulsion interaction, forces
arising from induced dipole, and hydrogen bonding (Murgich, 2002).
Glass Transition in Asphalt Binders 5

2.2.2. Thermodynamics of the Micelle Formation and Further Association


The formation and stability of molecular aggregates, such as micelles or disperse
solids, present in asphalt are determined by the changes in the total free energy of
the system. The binding of molecules is always an unfavorable entropic process
(reduction in degrees of freedom) (Vinter, 1996). This implies that the change in
enthalpy must be negative (Murgich, 2002), since only processes with negative
change in free energy are spontaneous. This was proven experimentally, and the
number of interaction sites on an asphaltene molecule was estimated to be between 1
and 2 and the heat of association in the range of 2-7 kJ/mol (Merino-Garcia et al.,
2004a, Merino-Garcia et al., 2004b). The association of heavy aromatics into bigger
entities, or even more pronounced structures, is an exothermic process. The
association becomes more pronounced with decreasing temperature. At low
temperatures, however, the increasing viscosity of the continuous phase impedes the
molecular diffusion and, therefore, the association.
The aggregates or micelles may further associate into more complex structures.
In the case of a sufficiently high dissolving power of the medium, phase separation
does not set in; and, with the accumulation of asphaltenes, internal structures
(networks) may form (Rogacheva et al., 1980). The very complex distribution of the
molecular aggregates (Sheu et al., 1992) suggests that asphalt cannot be described
simply as a pure sol or gel system (Saal et al., 1940, Eilers, 1949) formed by solid
asphaltene particles dispersed by resins, or as a simple and uniform micellar system
similar to those found in aqueous solutions of pure surfactants. Diversity of asphalt
molecules (in size, shape, polarity, etc.) disallows the organized structure (crystal) to
be built, and most of the molecules conform in an amorphous organization (Murgich
et al., 2001).

2.3. Parameters Affecting the Glass Transition

2.3.1. Molecular Weight


The molecular weight of the glass forming molecules is an important factor in
glass transition. The effect is not straightforward, and it is also dependent on the
molecular structure. For linear molecules, the chain ends exhibit greater mobility
over the repeat inner units, because they are bonded only on the one side. The
shorter molecules have more mobile chain ends per volume unit than longer
molecules and, thus, exhibit greater molecular mobility. Molecules with higher
mobility (shorter chains) become motionless at a lower temperature than those with
lower mobility (longer chains). This means that Tg increases with the increasing
molecular weight (Fox et al., 1955). Contrarily, the limited data for small polymer
rings surprisingly predicted the increase in Tg with decreasing molecular weight
(Clarson et al., 1985). The free volume theory (Fox et al., 1955) is inconsistent with
these results; however, one cannot omit the simple explanation that the free volume
6 Road Materials and Pavements Design. Volume X – No X/2007

decreases with decreasing ring size (molecular weight), as the rings become tighter
and tighter.

2.3.2. Functional and Structural Groups


The effect of various functional and structural groups on Tg may be very
complex and difficult to estimate. In polymers, for instance, introduction of para-
phenyl rings into the linear chain significantly increases Tg, due to fact that para-
phenyl restricts the chain backbone rotation. On the other hand, introduction of a
methylene group or oxygen (ether bond, C–O–C) into the backbone lowers Tg, due
to an increase in flexibility of the chain. Introduction of side alkyl chains also
significantly decreases Tg. It has been shown (Weyland et al., 1970, Krevelen et al.,
1976) that Tg can be successfully predicted by summing the partial Tg (Tgi) for
particular functional groups. Tgi is dependent on molecular weight and structure.
Generally, saturated and linear molecules possess higher molecular mobility due to
the rotational degree of freedom of the single bond. Therefore, saturated molecules
vitrify at much lower temperatures than unsaturated molecules. Rings and multiple
bonds, as well as heteroatoms (especially carbonyl groups), introduce a certain
rigidity into the molecule and reduce the molecular mobility. For instance, Tgi of the
carbonyl group is 964 K, while Tgi of the ether group is 232 K. The presence of
sulfur in the chain (C–S–C, Tgi=234 K) may lower the overall Tg, while Tgi of the
SO2 group (905 K) may do the opposite. Carbonyl and sulfonyl groups are often
considered as products of oxidative aging of asphalts, and their contribution to Tg
may be important in oxidized asphalts. The prediction of Tg using the method
developed by Krevelen may be difficult for asphalt, due to the vast number of
functional groups. It has been shown, however, that reasonable estimates of Tg can
be obtained (Huynh et al., 1978).

2.3.3. Cross-linking and Molecular Interactions


It is known in polymer science that introduction of cross-links into a polymer
strongly affects the local segmental relaxation and, hence, the glass transition. The
cross-links reduce the configurational degree of freedom, thereby increasing Tg. The
increase in Tg can be understood by a possible decrease in the molecular motion,
because the van der Waals interactions are replaced by shorter and firmer covalent
bonds (Chang, 1992). In asphalt, no formation of new inter-molecular covalent
bonds is expected at ambient conditions. Major interactions and associations
between asphalt molecules are due to weak van der Waals forces, hydrogen bonds
and other forces. The weak inter-molecular interactions in asphalt may reduce the
configurational degree of freedom, at least to some extent, as the covalent cross-
links do, possibly contributing to an increase in Tg. One should not omit the fact that
the association energies between the most polar molecules are in the order of 5
kJ/mol (Merino-Garcia et al., 2004a), while single covalent bonds (cross-links in
polymers) are much stronger (approximately in the order of 180-450 kJ/mol
(Israelachvili, 1985)).
Glass Transition in Asphalt Binders 7

2.3.4. Crystallinity
Some asphalts may contain a certain amount of crystalline phase (Claudy et al.,
1992); therefore, the effect of crystallinity on Tg should be also considered. Tg of
amorphous phase may increase as well as decrease with the degree of crystallinity,
depending on the relative density of the crystalline and amorphous phases. In most
cases, the ordered crystalline phase possesses higher density; and, the molecular
chains of the amorphous phase are entrapped in the crystalline lattice. Their mobility
is reduced, and Tg is increased (Flocke, 1962, Bair, 1994). The effect of crystalline
phase on the glass transition of asphalts was found to be significant (Kriz et al.,
2007a).

2.3.5. Dilution
The solvent power and Tg of the oily phase in asphalt are important factors in the
overall Tg. In polymer science, the effect of diluents to lower or increase Tg is well
known. Dilution with a solvent that has a lower Tg than the original material
decreases Tg of the system. Essentially, solvent with a higher Tg than the original
material increases the resulting Tg (Plazek et al., 1991). Furthermore, there are
solutions in which the mobility of the solvent is increased by the presence of a
polymer whose undiluted Tg is higher than that of the solvent. Free volume theory
cannot explain these unpredictable effects, and the additional concept of inter-
molecular coupling was introduced (Santangelo et al., 1994).

2.3.6. Other Effects


The effect of pressure on Tg is typically in the order of an increase by 20°C per
100 MPa of pressure (Ngai, 2004). Since the pressure applied on pavement by traffic
is in the order of several hundreds of kPa or a few Mpa, the increase of Tg due to
pressure would be in the order of several tenths °C.
It has also been shown that, for thin films of polystyrene and other polymers on
aggregate (silicone wafers), Tg of a particular polymer is strongly dependent on the
film thickness (Fryer et al., 2000). However, this effect was observable for film
thicknesses of 50 nm and thinner. Since the average binder film thickness in asphalt
mix is in the order of μm (Frolov et al., 1983), the effect of the film thickness on Tg
is probably negligible.

3. Experiment and Methods

3.1. Instruments

A TA Instruments Q100 differential scanning calorimeter (DSC), equipped with


the modulated temperature setup and liquid nitrogen cooling system, was used in
8 Road Materials and Pavements Design. Volume X – No X/2007

this study. Ultra pure helium (99.999 percent) was used to purge the experimental
cell. The rate was 50 mL/min. Standard hermetic aluminum pans were used for all
experiments. The sample mass was in the range of 7-10 mg. Sample pans were
sealed under nitrogen. The thermal history of the sample, if not stated otherwise,
was deleted prior the experiment. The procedure was as follows. The sample was
heated up to 150°C and underwent isotherm for 15 minutes. The sample was then
quenched to −100°C at a rate of 10°C/minute. The MDSC setup was developed and
evaluated in our other papers (Kriz et al., 2007b, Kriz et al., 2007c). The modulation
amplitude was 2°C; the modulation period, 60 seconds; and, the linear heating rate,
2°C/minute. The typical range was from −100 to 100 or 150°C. The glass transition
temperature was assigned to a temperature where the reversible heat capacity
reached half of the overall change.

Table 1. Basic properties of asphalt binders used in this study

Asphalt A Asphalt B AAC-11 AAV1


Ural Alaska
Crude oil source Cold Lake Red-water
Russia North Slope
Performance Grade (PG) PG 58-282 PG 52-282 PG 58-16 PG 52-22

Penetration 25°C [dmm] 1623 1863 133 121

Tg (MDSC) [°C] −25.82 −18.90 −23.59 −20.78

Wax (% wt) 0.704 4.624 5.06 3.13

1
data obtained from the SHRP Materials Reference Library (Jones, 1993), except for
the Tg determination.
2
AASHTO (American Association of State Highway and Transportation Officials)
MP320 (Asphalt Institute, 1994).
2,3
data acquired from Bituminous Materials Chair at the University of Calgary.
4
UOP (Universal Oil Products) 46-85 method (UOP Inc., 1985).

A Rheometric Scientific strain-controlled rheometer (ARES) was used for the


dynamic mechanical analysis. The linear viscoelastic (LVE) region was determined
as a linear region in the G ′ , G ′′ versus testing strain (Figure not presented). The
dynamic frequency sweeps were run (0.1 to 100 rad/s) at a constant temperature and
a constant strain within the LVE region.
A TA Instruments Q500 thermogravimetric analyzer (TGA) was used. About 40
mg of binder was placed on standard platinum pan prior to the experiment. Each
sample was first melted, and small droplets were prepared. These droplets were
Glass Transition in Asphalt Binders 9

stored in a freezer to minimize the evaporation and to ensure the same number of re-
heatings for each sample. The purging gas was either nitrogen (non-oxidative
evaporation) or air (oxidation) at a flow rate of 60 mL/min. The heating rate was
20°C/min (from ambient to test temperature), and test temperatures varied between
60 and 200°C, depending on the experiment. The sample was held at the test
temperature for 50-1440 minutes.

3.2. Asphalt Samples

Four different asphalts were used in this study. Asphalt A was an asphalt from
heavy Alberta crude oil (Cold Lake) of penetration grade 150/200, supplied by
Husky Energy Inc. It represented a high-quality binder for low-temperature
application. Asphalt B was an asphalt from Ural crude oil of penetration grade
150/200, supplied by SLOVNAFT Plc., Slovakia. It represented an asphalt binder
produced from the crude oil typically processed in Central Europe. The other two
binders were Asphalts AAC-1 (Redwater) and AAV (Alaska North Slope) from the
original Strategic Highway Research Program (SHRP) Material Reference Library.
These three binders had been stored in bulk (15kg) at room temperature for 15 years.

4.Results and Discussion

4.1. Dependence of the Glass Transition on Observation Time

According to the definition of a glassy state, the domain is considered glassy if


the molecular motion appears as nonexistent within the observation time. This
means that the glass transition is a kinetic process, and it is dependent on the
particular experiment’s observation time.
Figure 1 presents a master curve constructed from several isothermal dynamic
data measured for Asphalt A. The time-temperature superposition principle (tTs)
was applied (Ferry, 1961), and horizontal shift factors ( aT ) were determined. Shift
factors were fitted with the Williams-Landel-Ferry equation (WLF):

log aT = − c1 (T − Tr ) / (c 2 + T − Tr ) [1]

where, c1 and c2 are the empirical fitting parameters, T is the temperature, and Tr is
an arbitrarily selected reference temperature. The frequency where the loss modulus
( G ′′ ) attains its maximum value is usually assigned as the glass transition frequency
(Wada et al., 1959). With the use of tTs, the frequency can be converted to
temperature:
10 Road Materials and Pavements Design. Volume X – No X/2007

ω = aT ω [2]

where, ω is the reduced frequency, and ω is the testing frequency. The


combination of Equations 1 and 2 yields:

Tg = Tr +
(
c1 log ω g / ω ) [3]
(
c 2 + log ω / ω g )
where, ω g is the frequency at the maximum of G ′′ , and Tg is the glass transition
temperature. Using Equation 3, the glass transition temperature of the material can
be calculated as a function of testing frequency (Kriz et al., 2008). Results for
Asphalt A are presented in Figure 2.

Figure 1. Asphalt A – master curve, time-temperature superposition, test


temperatures from −30 to 80°C, G ′ (○), G ′′ (●), tan δ (−). Constant strains within
the linear viscoelastic region (LVE) were used. Geometry varied according to the
viscosity of the specimen: 50mm cone and plate (T>50°C), 25mm cone and plate
(T~20-50°C), 10mm plate-plate (T~0-20°C) and rectangular torsion bar (T<0°C).

It is apparent from Figure 2 that the glass transition temperature is strongly


dependent on the testing frequency. A single oscillation (period) refers to the
observation time. The frequency of 108 rad/s refers to the observation time in the
order of nanoseconds. The time allowed for the response to deformation is extremely
short at high frequencies; therefore, the response of relatively soft asphalt (at
temperatures above 0°C) is glassy. On the other hand, a low testing frequency (e.g.
10-7 rad/s) corresponds to extremely long observation times (2 years). Therefore, a
specimen has sufficient time to respond to deformation, and the asphalt appears as
non-glassy, even at very low temperatures – below −50°C.
Glass Transition in Asphalt Binders 11

10

-10

-20
Tg [°C]
-30

-40

-50

-60
-10 -5 0 5 10
log ω [rad/s]

Figure 2. Dependence of glass transition temperature (Tg) on testing frequency (ω)


in Asphalt A.

Figure 3. Effect of the heating rate on Tg (DSC experiment, Asphalt A).

A similar effect can be observed in a calorimetric experiment. In this case, we


can vary the heating or cooling rate. Examples of the effect of the heating rate on the
glass transition of Asphalt A is presented in Figures 3 and 4.
The data indicates that the glass transition temperature is strongly dependent on
the experimental conditions. Therefore, the glass transition temperature is valid only
for a particular set of conditions. Relating viscoelastic or even performance
parameters of the asphalt to a single glass transition temperature without further
12 Road Materials and Pavements Design. Volume X – No X/2007

specification of observation time, transition range, thermal history, etc. is


impractical.

Heating rate [°C/min]

1 10 100
-20

-24
Tg [°C]

-28

-32 y = 2.24Ln(x) - 33.48


R2 = 0.98
-36

Figure 4. Tg as a function of the heating rate in DSC experiment – Asphalt A.

4.2. Phase Compatibility

The major difference between the modulated differential scanning calorimetry


(MDSC) and the standard differential scanning calorimetry (DSC) is the thermal
program. In MDSC, the conventional linear heating program is modulated by
superimposing a sine wave (or other periodic waveform) of small amplitude on the
linear temperature ramp. Portions of each cycle then involve heating, while other
portions involve cooling. The overall trend, however, remains a linear change in
average temperature with time. The resultant heat flow signal is analyzed to separate
the response to the perturbation from the response to the underlying heating
program. Thus, the equilibrium (e.g. change of state) and kinetic (e.g. glass
transition) thermal events can be separated into two independent signals (Baur et al.,
1998, Judovits et al., 1998). Those two signals are the irreversible heat flow (change
of state) and the reversible heat flow (glass transition). The use of MDSC in asphalt
research was pioneered by Memom and Chollar (Memon et al., 1997) and by J.-F.
Masson (Masson et al., 2001, Masson et al., 2002, Masson et al., 2005a, Masson et
al., 2005b, Masson et al., 2006b).
The reversible signal can be used to describe the kinetic processes associated
with the amorphous phase, i.e. the glass transition. For polymers, it has already been
shown that the derivative of reversible heat capacity, with respect to temperature
signal, is very useful in the determination of the compatibility of the amorphous
phase (Song et al., 1995, Song et al., 1998a, Song et al., 1998b). The signal itself
represents the rate at which the molecules pass through the glass transition. If the
amorphous phase is fully compatible and uniform, the shape of the signal should
Glass Transition in Asphalt Binders 13

correspond to a Gaussian type distribution with a single maximum (Song et al.,


1998b). For incompatible blends of polymers, however, multiple peaks were
observed suggesting that there were actually two or more separate amorphous
domains, each of which passes through its own glass transition at different
temperatures (Song et al., 1998b). Therefore, the signal can be used to study the
compatibility of the amorphous domain and also evaluate the effect of various
processes on the phase stability.

Figure 5. Compatible amorphous phase in Asphalt A (c) and incompatible


amorphous phase in Asphalt AAV (a) and Asphalt AAC-1 (b). No separate glass
transition outside the main glass transition was observed in these asphalts upon
heating up to 150°C. Curves are vertically shifted.

Examples of compatible and incompatible amorphous phases are presented in


Figure 5. It has been shown that the multiphase system in asphalt also evolves with
time in isothermal conditions (Masson et al., 2001, Masson et al., 2002). The
compatibility and evolution of amorphous domain were found to be strongly
dependent on the presence of crystalline phase (Kriz et al., 2007a). In the case of
semi-crystalline asphalt, the multiphase system has to be considered. At least three
separate components exist in asphalt based on its physical state: crystalline, glassy
amorphous and non-glassy amorphous phases. Part of the glassy amorphous domain
may be considered as rigid amorphous phase (RAPh), which has been found in
semi-crystalline polymers and is also indicated in asphalt (Kriz et al., 2008). The
physical state of a particular molecule, i.e. crystalline, glassy/non-glassy amorphous,
depends on molecular weight and structure, the rest of the molecules in the sample
(interaction, solubility, spatial interference), thermodynamic parameters
(temperature, pressure) and time/thermal history. Each domain consists of molecules
in the same state. The crystallization and vitrification of asphalt molecules define the
physical state of the domains present in asphalt at low temperatures, i.e. at
temperatures within the glass transition and/or melting range. These processes are
fully reversible upon heating (above the temperature range of glass transition and/or
14 Road Materials and Pavements Design. Volume X – No X/2007

melting). The following section discusses the effect of irreversible change in


composition – oxidation – on the phase compatibility and glass transition.

4.3. Effect of Oxidative Aging on Phase Compatibility and Glass Transition

The effect of oxidative aging on asphalt phase compatibility and the glass
transition temperature was studied. Asphalt A, which has a negligible crystalline
domain, was chosen for this experiment. The processes associated with the
crystalline phase may not only overlap with the glass transition on the calorimetric
signal, but also contribute to the incompatibility of the amorphous phase (RAPh)
(Kriz et al., 2007a). The absence of the crystalline phase in Asphalt A permitted the
evaluation of the effect of oxidation on phase stability and glass transition. The
reversible calorimetric signal of neat Asphalt A (Figure 5) shows no phase
separation and a uniform amorphous phase. The peak shape suggests a Gaussian
molecular distribution. Each molecule contributes to the glass transition at a
particular temperature based on the molecular weight, structure and inter-molecular
interaction. An irreversible calorimetric signal (not presented) showed a very small
crystalline domain (heat of melting in the order of 0.5 J/g), which was the lowest
value we have ever observed in an asphalt binder. As more than 20 different binders
were subjected to the MDSC experiment in our laboratory, a purely amorphous
binder may be considered as a rarity.

4.3.1. Effect of Evaporation


In this section, the effect of evaporation on the phase stability and Tg is
discussed. Evaporation plays a minor role in the overall aging process of pavement
(Kemp et al., 1981). Our objective was to estimate the effect of evaporation and
separate it from simultaneous oxidation. The data presented here may also be related
to the refining process setup – how the low-temperature properties of the product are
dependent on the degree of distillation.
Asphalt A was subjected to a thermogravimetric analyzer (TGA) experiment
under nitrogen atmosphere. The two sets of samples were prepared in the TGA.
First, the same soaking time was maintained (24 hours), and the temperatures,
between 60 and 200°C, were held constant. Second, the same soaking temperature
was maintained (200°C), and the soaking times were changed. Tg was determined for
both sets of samples: the results indicate that there was no qualitative difference
between the samples. It can be concluded, therefore, that the time and temperature
can be interchanged, within reasonable limits, during such an experiment (Figure 6).
Typical straight run asphalt is a residuum after crude oil distillation, which is an
equilibrium separation process; and, complete separation of the individual
components, based on their normal boiling point, is impossible (Smith et al., 2005).
Therefore, asphalt contains molecules with lower boiling points than the equilibrium
temperature (Speight, 1999). The boiling point depends not only on the carbon
Glass Transition in Asphalt Binders 15

number, but also on the molecular structure (Wessel et al., 1995). The heavy oil
distilled from asphalt consists predominantly of saturates (~50%) and naphthene-
aromatics (~40%) (Noel et al., 1970). Single bonds in saturated molecules possess a
rotational degree of freedom; and, a whole chain is, therefore, more flexible than in
molecules with multiple bonds or rings. Higher molecular mobility of saturated
molecules results in a lower Tg (Ngai, 2004), and removal of such a component
essentially leads to an increase in overall Tg.

262
y (●) = 0.66x + 247.02
260 R2 = 0.99
258
y (▲) = 0.66x + 246.75
256 R2 = 0.98
Tg [K]

254

252 y (◊) = 0.50x + 246.93


R2 = 0.80
250
y (□) = 0.65x + 247.92
248 R2 = 0.90
246
0 5 10 15 20
Weight loss [%]

Figure 6. Effect of Evaporation and Oxidation on Tg in Asphalt A: evaporation at


different temperatures (60 to 200°C) under nitrogen for 24 hours (●); evaporation
at 200°C, time varied between 50 min and 24 h (▲); oxidation in air at 200°C (◊);
oxidation in air at 275°C (□).

Figure 7. Effect of Evaporation on Glass Transition in Asphalt A. Original asphalt


(---) and after 5 %wt. (·−·) and 18 %wt. (−) was evaporated.
16 Road Materials and Pavements Design. Volume X – No X/2007

Figure 8. Effect of Oxidation on Glass Transition in Asphalt A .Original asphalt


(---) and after 200 min at 200°C (·−·) and after 50 min at 275°C (−).

The results obtained for Asphalt A (Figures 6 and 7) were consistent with the
literature, where Tg was already found to increase linearly with the decrease in
penetration (du Bois, 1966) or the logarithm of penetration (Schmidt et al., 1965). It
has also been found by several researchers that the overall Tg linearly increases with
the amount of mass fraction of asphaltenes, the heaviest asphalt fraction (Wada et
al., 1960, Giavarini, 1984). The simple linear behavior of Tg upon evaporation
(Figure 6) suggests that the raise in overall Tg was directed predominantly by a loss
of the “low Tgi” component. No separate glass transition outside the main transition
was observed in the neat Asphalt A (Figure 5). The solubility factor plays an
important role during the glass transition. The heaviest and most polar molecules
vitrify at much lower temperature than they would normally do as a separate entity,
due to being mobilized and prohibited to pack into a glass by molecules of a much
lower Tg.
Figure 9 presents the reversible calorimetric signal of Asphalt A and its
asphaltenes and maltenes. Asphaltenes were prepared using the American Society
for Testing and Materials’ ASTM D-3279-97 method. The glass transition of
asphaltenes is extremely broad, due to higher molecular weight and variety in
molecular types (see section 4.3.4.). The signal change above 200°C was caused by
the deformation of the pan bottom as the internal pressure increased upon heating.
This was confirmed visually. The asphaltenes and maltenes had glass transition
temperatures of 68 and −30.8°C, respectively. The weight fraction of the asphaltenes
was 12.9 %wt. The glass transition temperature of Asphalt A was calculated from
the glass transition temperatures of the asphaltenes and maltenes, using a simple
mixing rule (Krevelen et al., 1976). The value of −18.1°C was obtained (original
Asphalt A has Tg of −25.7°C). The difference was expected, as this approach omits
molecular interactions and solubility effects that were present in the original asphalt.
Glass Transition in Asphalt Binders 17

Figure 9. Glass transition as observed in (a) Asphalt A, in separated (b) asphaltenes


(12.9 %wt.) and (c) maltenes (87.1 %wt.).

4.3.2.Effect of Oxidation
Asphalt A was oxidized in air at 200 and 275°C. The glass transition temperature
was determined and plotted versus the weight loss. The results are presented in
Figures 6 and 8. An interesting fact was observed: there was only a negligible
difference between the glass transition temperatures of the evaporated (under
nitrogen) and oxidized samples. The glass transition still increased linearly with the
weight loss and the slope, and a y-intercept of the linear fit was not affected by
oxidation. The curves were very similar with a single Gaussian peak in both the
oxidized and evaporated binders, as shown in Figures 7 and 8.
It was expected that oxidation would affect the phase stability of the system and
that a phase separation would set in. This was, however, not observed in samples
with no thermal history. Thermal history turned out to be critical in this phase
stability issue (see section 4.4.).

4.3.3. No Phase Separation in Evaporated/Oxidized Asphalt?


Is the glass transition of evaporated/oxidized asphalt affected only by the loss of
low Tg components or is there any other parameter involved? Further
association/flocculation of polar molecules, for instance, may be one of the
additional effects that may occur upon loss of the light components. Essentially,
such association would lead to significantly bigger aggregates of polar molecules, as
proposed in earlier studies (Saal et al., 1940, Eilers, 1949, Rogacheva et al., 1980).
Molecular motion inside such an aggregate would be restricted, and the aggregated
molecules cannot be mobilized/solvated by other components in the asphalt.
Association of polar molecules would result in incompatibility with the rest of the
amorphous phase. This newly developed amorphous domain should be identified by
one of the two following features on the calorimetric signal. First, the separate
18 Road Materials and Pavements Design. Volume X – No X/2007

amorphous domain would be observable only if the associated molecules that form a
new phase still possess some degree of the molecular mobility and, therefore, the
ability to exhibit glass transition. The new phase should be observed as a new peak
on the calorimetric signal. This was not observed (Figures 7 and 8). Second, the
molecules forming the new phase have lost molecular motion (associated molecules
may have significantly restricted mobility since they are closely packed and possess
a very low degree of freedom) and now behave like an “immobilized” solid phase.
This phase would have extremely low molecular mobility; and, its glass transition, if
it exists, would be extremely weak and perhaps undetectable in the calorimetric
experiment. In such a case, the removal of the high molecular weight molecules
from the original amorphous phase would lower the glass transition of the remaining
amorphous phase. This is in conflict with the peak broadening and increase in Tg
upon evaporation/oxidation (Figures 7 and 8).

4.3.4. Transition Broadening


The glass transition was broadened with a higher degree of
evaporation/oxidation (Figures 7 and 8). The broadening of the glass transition is
described in polymer literature. Wunderlich (Wunderlich, 1994) reported that the
solutions involving macromolecules always have a symmetrically broadened glass
transition region, which is caused by the inability of the molecules to randomize the
structure of their molecular backbone on a nanometer scale.
The concentration of polar molecules (with a high partial Tg) per volume
increases as the continuous phase is being removed. With the absence of smaller
mobile molecules, the packing of large polar molecules into the glass upon cooling
is complicated and slow. On heating, devitrification of the glass is directed by
molecules with higher Tg, which may postpone the devitrification of the low Tg
molecules through spatial interference. The spatial interference becomes more
pronounced and the glass transition broader the higher the average molecular weight
of the system.
Other effects may also contribute. Peak broadening accompanied with the
relatively small discontinuity in heat capacity has been observed in highly cross-
linked polymeric systems (Chang, 1992). It has also been shown that the breadth of
the glass transition increases as polymers are mixed into miscible poly-blends or
covalently linked together into homogeneous copolymers (Bates et al., 1984). It is
improbable that new covalent bonds (bonding energy in the range of 180-450
kJ/mol) in between molecules (cross-links) are being created during evaporation or
oxidation in asphalt. Rather, very weak non-covalent “cross-links”, most probably
based on the van der Waals forces (~1 kJ/mol) or H-bonds (5-10 kJ/mol)
(Israelachvili, 1985), exist between the molecules. These forces are not strong
enough to bond the aggregates/molecules together into bigger agglomerates
permanently, yet strong enough to spatially coordinate them. The number of such
interactions may increase as the amount of the continuous phase decreases
(evaporation) or polarity increases (oxidation). Weak interactions may contribute to
Glass Transition in Asphalt Binders 19

peak broadening because higher energy/temperature would be needed to overcome


these forces and set molecules or molecular chains in motion.

4.4. Effect of Time

Figures 6-8 indicate that there is no significant difference between oxidized and
evaporated samples in terms of glass transition and Tg. Even the most severely
oxidized sample (275°C in air, 300 min) showed neither phase separation nor the
glass transition temperature deviated from the trend set by the evaporated samples.
It has been shown (Kriz et al., 2007a, Kriz et al., 2007b, Kriz et al., 2008) that
isothermal conditioning at a temperature within the glass transition range affects the
phase distribution and glass transition temperature. The glass transition temperature
was found to increase linearly with the logarithm of conditioning time. The
mechanical properties have also been shown to develop with time. A certain analogy
to the physical aging of amorphous polymers was suggested. The studies indicated
that changes within the amorphous phase at low temperatures are extremely slow, as
the diffusion and molecular mobility are reduced.

Figure 10. Asphalt A conditioned at −20°C for 140 days. Percentage represents a
weight loss of a particular sample during the “evaporation” experiment (200°C in
nitrogen). 0% wt. represents the control sample (conditioned but not evaporated).
Curves vertically shifted.

All samples (original, evaporated and oxidized) were subjected to isothermal


conditioning at −20°C for 3,360 hours (140 days). Results are presented in Figures
10-12. The differences in glass transition among the original, oxidized and
evaporated samples became more apparent. For example, with the conditioning, Tg
20 Road Materials and Pavements Design. Volume X – No X/2007

increased by almost 2°C in the original binder; while in the most oxidized sample
(275°C, 300 min, 10.7 %wt. loss), Tg increased by 16.5°C. The change in Tg upon
conditioning was greater in oxidized samples than in evaporated samples, perhaps
due to the increased content of high Tgi components (carbonyl and sulfonyl
functional groups and higher aromatic content The peak split at −20°C (conditioning
temperature) was observed in severely evaporated and oxidized samples (Figures 10
and 11). Upon isothermal conditioning, phase separation occurred with two glassy
domains, stable above and below the conditioning temperature. This indicates that
the phase stability of the oxidized asphalt was affected, and phase separation
occurred. The origin of the small peak observable at around 4°C in Figures 10 and
11 is not clear. It may represent melting enthalpy of water as frost may develop in
the sample upon long term storage at sub-zero temperature.

Figure 11. Asphalt A conditioned at −20°C for 3,360 hours (140 days). Percentage
represents weight loss of a particular sample during the “oxidation” experiment
(200°C in air). 0% wt. represents the control sample (conditioned but not oxidized).
Curves vertically shifted.

4.5.Thermo-reversibility

It has been shown that the changes in phase distribution and glass transition upon
isothermal conditioning within the glass transition range are fully reversible, if the
thermal history is deleted. This can be achieved by heating the sample to a
sufficiently high temperature (150°C) where the solid structure is
devitrified/melted/disassociate, and asphalt becomes a simple fluid (Kriz et al.,
2007a, Kriz et al., 2008).
Glass Transition in Asphalt Binders 21

Weight loss [% wt.]

0.0 2.0 4.0 6.0 8.0 10.0 12.0


0

-5
Tg [°C] -10

-15

-20

-25

-30

Figure 12. Asphalt A with no thermal history (unfilled symbols) and conditioned
at −20°C for 140 days (filled symbols). Original asphalt (○), evaporated in nitrogen
at 200°C (◊), and oxidized in air at 200°C (□).

Figure 13. The effect of low-temperature isothermal conditioning on glass transition


was fully reversible in Asphalt A, and an identical result was obtained for the
original samples (no isothermal conditioning) and for samples isothermally
conditioned with the thermal history deleted (samples marked DTH). Curves are
vertically shifted.

The thermal history was deleted in the evaporated and oxidized samples, and
MDSC data were recorded. The comparison of the original MDSC run (prior to
isothermal conditioning) and the MDSC run of the conditioned samples with deleted
thermal history is presented in Figure 13. The results indicate that the effect of
isothermal conditioning was fully reversible, as the specimen was heated up to
150°C. Reversibility indicates that the processes observed in asphalt upon isothermal
conditioning are controlled by changes in molecular mobility, which can be affected
22 Road Materials and Pavements Design. Volume X – No X/2007

by spatial interferences and molecular interactions (possibly stronger in oxidized


samples). As the sample is heated, temporary spatial interferences and molecular
interactions disappear, and the sample returns to the initial state. Changes in phase
stability can be reversed by heating the asphalt to a sufficiently high temperature.
The lowest temperature at which the asphalt retains its original properties was not
determined. This should be the subject of future study.

4.6. Long-term Conditioning of Asphalt B

Asphalt B was subjected to long-term conditioning at two temperatures −20°C


and +20°C. Results are presented in Figure 14. Interestingly, peak separation was
observed in both cases, i.e. in samples conditioned at both −20°C and +20°C.

Figure 14. Effect of long-term isothermal conditioning on glass transition in


Asphalt B. (a) conditioned at −20°C for 5,400 hours, (b) control sample (no thermal
history), (c) conditioned at +20°C for 8,445 hours. Curves vertically shifted.

The sample conditioned at −20°C started to devitrify upon heating very rapidly.
This suggests that molecules that normally devitrify between −50 and −20°C, were
built into the glassy structure. The structure of the already glassy molecules perhaps
obstructed (spatially or by molecular interaction) the molecular motion of these
molecules, and they eventually “froze” in the structure and became glassy – note the
analogy to the rigid amorphous phase (Kriz et al., 2008). Therefore, upon heating,
the restricted molecules could not devitrify until the “constraining molecules”
devitrified too. The restricted molecules are released instantaneously. This explains
the rapid start of the transition at −20°C (Figure 14, curve (a)).
The sample conditioned at 20°C also showed separation into two phases – one
above and one below the conditioning temperature. In this case, the conditioning
Glass Transition in Asphalt Binders 23

temperature is close to transition end (upon heating) yet within the glass transition
range. Therefore, the glassy domain above the conditioning temperature is very
small.
The phase separation was observed after long-term conditioning at the
temperature within the glass transition range in neat Asphalt B and in the severely
evaporated and oxidized samples of Asphalt A. These asphalts are considered as
phase instable asphalts. Phase separation upon long-term conditioning may be an
indictor of the phase stability of the whole asphalt system. If the binder is phase
stable, separation does not occur at the temperature of conditioning. The stability
may be affected by several processes, as shown in Figure 15.
If any disturbance occurs within the sensitive region, as indicated in Figure 15, a
phase separation may occur. With evaporation, the sensitive region is affected only
if significant fraction of light-end components is evaporated. Polar molecules are
sensitive to oxidation, and their polarity is increased during oxidation. Less polar
molecules are not easily oxidized, and their polarity is not significantly affected.
Gradual transition is discontinued, and phase separation may occur. In thermal
cracking, severe treatment at high temperatures leads to a cleavage of covalent
bonds and results in a low boiling product (polarity decreased) and coke (carbonized
residuum). With solvent recovery, the action of non-polar solvent affects the
continuum within the sensitive region and allows the most polar molecules to
associate: phase separation occurs.

Figure 15. Asphalt as a molecular continuum and the effect of 4 processes on phase
stability. Shades of grey represent molecular continuity in polarity or molecular
weight (white - non-polar or lightest, black - most polar or heaviest). The region
between white dashed lines represents the sensitive region, where molecular
continuity is essential for system stability.
As previously discussed, the phase compatible asphalt shows a single Gaussian
peak on calorimetric signal as the glass transition smoothly proceeds from the most
24 Road Materials and Pavements Design. Volume X – No X/2007

mobile to least mobile molecules or molecular chains. In such asphalt, the molecular
continuity from the lightest to heaviest molecules must exist. If it is broken, sooner
or later the phase incompatibility will occur in the asphalt. This material will behave
as a composite material with properties directed by a “weaker” constituent.

5. Conclusions

The glass transition in asphalt binders is always dependent on experimental


conditions; therefore, relating asphalt properties or performance to a particular
parameter may be impractical without further specification of the experiment.
The effects of evaporation of light-end components in an inert atmosphere and of
oxidation in air on the glass transition of asphalt were studied. It was shown that the
glass transition temperature increases linearly with the weight loss in both the
evaporated and oxidized asphalt samples and there was virtually no difference
between these two sets of samples.
The difference in glass transition temperature and phase stability between
evaporated and oxidized asphalt specimens became more apparent after long-term
isothermal conditioning at a temperature within the glass transition region. The
increase of Tg was observed upon isothermal conditioning. The increase of Tg was
higher in oxidized samples. The oxidized samples of Asphalt A showed wider glass
transition on the temperature scale that the evaporated. Also phase separation at the
conditioning temperature is somewhat more apparent in oxidized specimens. Two
separate glassy domains, one above and one below the conditioning temperature,
were observed, if a sufficient length of time was allowed for the conditioning.
Phase incompatibility may affect the low-temperature properties of asphalt
binder, as the asphalt failure may be directed by a failure within a domain with
higher rigidity/glass transition temperature.

Acknowledgements
The authors would like to acknowledge the Natural Sciences and Engineering
Research Council of Canada and Husky Energy Inc. for financial support, and the
Bituminous Materials Chair at the University of Calgary for providing the
performance grades for Asphalts A and B.

6. Bibliography

Andersen S.I., “Effect of Precipitation Temperature on the Composition of N-


Heptane Asphaltenes 2”, Fuel Sci Techn Int, Vol. 13, No. 5, 1995, p. 579-604.
Glass Transition in Asphalt Binders 25

Andersen S.I., Birdi K.S., “Influence of temperature and solvent on the


precipitation of asphaltenes”, Fuel Sci Techn Int, Vol. 8, No. 6, 1990, p. 593-615.
Andersen S.I., Birdi K.S., “Aggregation of Asphaltenes As Determined by
Calorimetry”, J Colloid Interf Sci, Vol. 142, No. 2, 1991, p. 497-502.
Andersen S.I., Speight J.G., “Petroleum resins: Separation, character, and role in
petroleum”, Pet Sci Technol, Vol. 19, No. 1-2, 2001, p. 1-34.
Asphalt Institute, Superpave Asphalt Binder Specification, Asphalt Institute,
Lexington, KY, USA, 1994.
Bair H.E., “Glass Transition Measurement by DSC” in Seyler R.J. (Ed),
Assignment of the Glass Transition, ASTM, Philadelphia, PA, USA, 1994, p. 50-74.
Bates F.S., Bair H.E., Hartney M.A., “Block Copolymers Near the Microphase
Separation Transition. 1. Preparation and Physical Characterization of A Model
System”, Macromolecules, Vol. 17, No. 10, 1984, p. 1987-1993.
Baur H., Wunderlich B., “About Complex Heat Capacities and Temperature-
Modulated Calorimetry”, J Thermal Anal, Vol. 54, No. 2, 1998, p. 437-465.
Carlson C.S., Langer A.W., Stewart J., Hill R.M., “Thermal hydrogenation -
Transfer of hydrogen from tetralin to cracked residua”, Ind Eng Chem, Vol. 50, No.
7, 1958, p. 1067-1070.
Chang S.S., “Effect of Curing History on Ultimate Glass-Transition Temperature
and Network Structure of Cross-Linking Polymers”, Polymer, Vol. 33, No. 22,
1992, p. 4768-4778.
Clarson S.J., Dodgson K., Semlyen J.A., “Studies of Cyclic and Linear
Poly(Dimethylsiloxanes). 19. Glass-Transition Temperatures and Crystallization
Behavior”, Polymer, Vol. 26, No. 6, 1985, p. 930-934.
Claudy P., Letoffe J.M., King G.N., Planche J.P., “Characterization of Asphalt
Cements by Thermomicroscopy and Differential Scanning Calorimetry: Correlation
to Classic Physical Properties”, Fuel Sci Techn Int, Vol. 10, No. 4-6, 1992, p. 735-
765.
Corbett L.W., “Refinery Processing of Asphalt Cement”, Trans Res Rec, Vol.
999, 1984, p. 1-6.
du Bois P., “Die Temperatur des Asphaltbitumens beim Übergang in den
glasigen Zustand”, Bitumen-Teere-Asphalte-Peche, Vol. 17, No. 7, 1966, p. 254-
255.
Eilers H., “The colloidal structure os asphalt”, J Phys Chem-US, Vol. 53, No. 8,
1949, p. 1195-1211.
Farcasiu M., “Fractionation and Structural Characterization of Coal Liquids”,
Fuel, Vol. 56, No. 1, 1977, p. 9-14.
26 Road Materials and Pavements Design. Volume X – No X/2007

Ferry J.D., Viscoelastic Properties of Polymers, Wiley, New York, 1961.


Flocke H.A., “Ein Beitrag zum mechanischen Relaxationsverhalten von
Polyäthylen, Polypropylen, Gemischen aus diesen und Mischpolymerisaten aus
Propylen und Äthylen”, Kolloid ZZ Polym, Vol. 180, No. Heft 2, 1962, p. 118-126.
Fox T.G., Loshaek S., “Influence of Molecular Weight and Degree of
Crosslinking on the Specific Volume and Glass Temperature of Polymers”, J Polym
Sci, Vol. 15, 1955, p. 371-390.
Frolov A.F., Vasileva V.V., Frolova E.A., Ovchinnikova V.N., “Strength and
Structure of Asphalt Films”, Chem Tech Fuels Oil+, Vol. 19, No. 7-8, 1983, p. 415-
419.
Fryer D.S., Nealey P.F., de Pablo J.J., “Thermal probe measurements of the glass
transition temperature for ultrathin polymer films as a function of thickness”,
Macromolecules, Vol. 33, No. 17, 2000, p. 6439-6447.
Giavarini C., “Visbreaker and Straight-Run Bitumens”, Fuel, Vol. 63, No. 11,
1984, p. 1515-1517.
Goodrich J.L., Goodrich J.E., Kari W.J., “Asphalt Composition Tests: Their
Application and Relation to Field Performance”, Trans Res Rec, Vol. 1096, 1986, p.
146-167.
Groenzin H., Mullins O.C., “Asphaltene molecular size and structure”, J Phys
Chem A, Vol. 103, No. 50, 1999, p. 11237-11245.
Groenzin H., Mullins O.C., “Molecular size and structure of asphaltenes”, Pet
Sci Technol, Vol. 19, No. 1-2, 2001, p. 219-230.
Heithaus J.J., “Measurement and significance of asphaltene peptization”, J I
Petrol, Vol. 48, No. 458, 1962, p. 45-53.
Hiemenz P.C., Rajagopalan R., Principles of colloid and surface chemistry,
Marcel Dekker, New York, 1997.
Huynh H.K., Khong T.D., Malhotra S.L., Blanchard L.P., “Effect of Molecular-
Weight and Composition on Glass-Transition Temperatures of Asphalts”, Anal
Chem, Vol. 50, No. 7, 1978, p. 976-979.
Israelachvili J.N., Intermolecular and surface forces with applications to
colloidal and biological systems, Academic Press, London, 1985.
Jones D.R., SHRP Materials Reference Library: Asphalt Cements - A Concise
Data Compilation, Report No. SHRP-A-645, Strategic Highway Research Program,
National Research Council, Washington, D.C., 1993.
Judovits L., Menczel J.D., Leray A.G., “Molecular Weight Effects on the
Reorganization of Poly(vinylidene fluoride), Polyamide 12, and Poly(p-phenylene
sulfide)”, J Thermal Anal, Vol. 54, No. 2, 1998, p. 605-622.
Glass Transition in Asphalt Binders 27

Katz D.L., Beu K.E., “Nature of Asphaltic Substances”, Ind Eng Chem, Vol. 37,
No. 2, 1945, p. 195-200.
Kemp G.R., Predoehl N.H., “A Comparison of Field and Laboratory
Environments on Asphalt Durablity”, Asphalt Paving Technology, 1981, p. 492-537.
Koots J.A., Speight J.G., “Relation of Petroleum Resins to Asphaltenes”, Fuel,
Vol. 54, No. 3, 1975, p. 179-184.
Krevelen D.W., Hoftyzer P.H., Properties of polymers, their estimation and
correlation with chemical structure, Elsevier Scientific Pub. Co, Amsterdam, 1976.
Kriz P., Stastna J., Zanzotto L., “Physical Aging in Semi-Crystalline Asphalt
Binders”, accepted by the 2008 AAPT Annual Meeting and Technical Sessions, April
27-30, 2008, Philadelphia, PA, 2008.
Kriz P., Stastna J., Zanzotto L., “Effect of Low-Temperature Isothermal
Conditioning on Glass Transition in Asphalt Binders”, Proceedings of the 52nd
Annual Conference of the Canadian Technical Asphalt Association, Niagara Falls,
ON, Canada, 2007a.
Kriz P., Stastna J., Zanzotto L., “A Calorimetric Study of the Low-Temperature
Properties of Asphalt Binders”, Proceedings of the 43rd International Petroleum
Conference, September 24-26, 2007, Bratislava, Slovakia, 2007b.
Kriz P., Stastna J., Zanzotto L., “Time Dependence of Asphalt Properties at Low
Temperatures”, The 6th International Symposium on Asphalt Binder Rheology and
Pavement Performance, April 2-3, 2007, Tampa Bay, FL, USA, 2007c.
Mack C., “Colloid Chemistry of Asphalts”, J Phys Chem-US, Vol. 36, 1932, p.
2901-2914.
Masson J.F., Collins P., Polomark G., “Steric hardening and the ordering of
asphaltenes in bitumen”, Energ Fuel, Vol. 19, No. 1, 2005a, p. 120-122.
Masson J.F., Leblond V., Margeson J., “Bitumen morphologies by phase-
detection atomic force microscopy”, J Microsc-Oxford, Vol. 221, 2006a, p. 17-29.
Masson J.F., Polomark G., Collins P., “Glass Transitions and Amorphous Phases
in SBS-bitumen Blends”, Thermochim Acta, Vol. 436, No. 1-2, 2005b, p. 96-100.
Masson J.F., Polomark G.M., “Bitumen Microstructure by Modulated
Differential Scanning Calorimetry”, Thermochim Acta, Vol. 374, No. 2, 2001, p.
105-114.
Masson J.F., Polomark G.M., Bundalo-Perc S., Collins P., “Melting and glass
transitions in paraffinic and naphthenic oils”, Thermochim Acta, Vol. 440, No. 2,
2006b, p. 132-140.
28 Road Materials and Pavements Design. Volume X – No X/2007

Masson J.F., Polomark G.M., Collins P., “Time-dependent Microstructure of


Bitumen and its Fractions by Modulated Differential Scanning Calorimetry”, Energ
Fuel, Vol. 16, No. 2, 2002, p. 470-476.
McKay J.F., Amend P.J., Cogswell T.E., Harnsberger P.M., Erickson R.B.,
Latham D.R., “Petroleum asphaltenes: Chemistry and composition” in Uden P.C.,
Siggia S., Jensen H.B. (Eds), Analytical Chemistry of Liquid Fuel Sources: Tar
Sands, Oil Shale, Coal, and Petroleum, American Chemical Society, Washington,
1978, p. 128-142.
Memon G.M., Chollar B.H., “Glass transition measurements of asphalts by
DSC”, J Therm Anal, Vol. 49, No. 2, 1997, p. 601-607.
Merino-Garcia D., Andersen S.I., “Interaction of asphaltenes with nonylphenol
by microcalorimetry”, Langmuir, Vol. 20, No. 4, 2004a, p. 1473-1480.
Merino-Garcia D., Murgich J., Andersen S.I., “Asphaltene Self-association:
Modeling and Effect of Fractionation with a Polar Solvent”, Pet Sci Technol, Vol.
22, No. 7-8, 2004b, p. 735-758.
Moynihan C.T., “Phenomenology of the Structural Relaxation Process and the
Glass Transition” in Seyler R.J. (Ed), Assignment of the Glass Transition, ASTM,
Philadelphia, U.S.A., 1994, p. 32-49.
Murgich J., “Intermolecular forces in aggregates of asphaltenes and resins”, Pet
Sci Technol, Vol. 20, No. 9-10, 2002, p. 983-997.
Murgich J., Strausz O.P., “Molecular mechanics of aggregates of asphaltenes and
resins of the Athabasca oil”, Pet Sci Technol, Vol. 19, No. 1-2, 2001, p. 231-243.
Nellensteyn F.J., “The colloidal structure of bitumens” in Dunstan A.E., Nash
A.W., Brooks B.T., Tizard H. (Eds), The Science of Petroleum, Oxford University
Press, London, 1938, p. 2760-2763.
Ngai K.L., “The Glass Transition and the Glassy State” in Mark J.E. (Ed),
Physical Properties of Polymers, Cambridge University Press, Cambridge, UK,
2004, p. 72-152.
Noel F., Corbett L.W., “A Study of the Crystalline Phases in Asphalts”, J I
Petrol, Vol. 56, 1970, p. 261-268.
Plazek D.J., Seoul C., Bero C.A., “Diluent effects on viscoelastic behavior”, J
Non-Cryst Solids, Vol. 131-133, No. 2, 1991, p. 570-578.
Porte G., Zhou H.G., Lazzeri V., “Reversible description of asphaltene colloidal
association and precipitation”, Langmuir, Vol. 19, No. 1, 2003, p. 40-47.
Rogacheva O.V., Rimaev R.N., Gubaidullin V.Z., Khakimov D.K.,
“Investigation of the Surface-Activity of the Asphaltenes of Petroleum Residues”,
Colloid J USSR, Vol. 42, No. 3, 1980, p. 490-493.
Glass Transition in Asphalt Binders 29

Saal R.N.J., Labout J.W.A., “Rheological properties of asphaltic bitumens”, J


Phys Chem-US, Vol. 44, No. 2, 1940, p. 149-165.
Santangelo P.G., Roland C.M., Ngai K.L., Rizos A.K., Katerinopoulos H.,
“Dielectric and mechanical relaxation in PMPS, BMC and their mixtures”, J Non-
Cryst Solids, Vol. 172-174, No. 2, 1994, p. 1084-1093.
Schmidt R.J., Barral E.M., “Asphalt Transitions”, J I Petrol, Vol. 51, No. 497,
1965, p. 162-168.
Selucky M.L., Kim S.S., Skinner F., Strausz O.P., “Structure-related properties
of athabasca asphaltenes and resins as indicated by chromatographic separation” in
Bunger J.W., Li N.C. (Eds), Chemistry of asphaltenes, American Chemical Society,
Washington DC, 1981, p. 83-118.
Sheu E.Y., Liang K.S., Sinha S.K., Overfield R.E., “Polydispersity Analysis of
Asphaltene Solutions in Toluene”, J Colloid Interf Sci, Vol. 153, No. 2, 1992, p.
399-410.
Smith J.M., Van Ness H.C., Abbott M.M., Introduction to chemical engineering
thermodynamics, McGraw-Hill, Boston, 2005.
Song M., Hammiche A., Pollock H.M., Hourston D.J., Reading M., “Modulated
Differential Scanning Calorimetry. I. A Study of the Glass-Transition Behavior of
Blends of Poly(Methyl Methacrylate) and Poly(Styrene-Co-Acrylonitrile)”,
Polymer, Vol. 36, No. 17, 1995, p. 3313-3316.
Song M., Hourston D.J., “Temperature-Modulated Differential Scanning
Calorimetry - IX. Some Comments on the Rigid Amorphous Fraction in Semi-
Crystalline Poly(ethylene terephthalate)”, J Thermal Anal, Vol. 54, No. 2, 1998a, p.
651-657.
Song M., Hourston D.J., Schafer F.U., Pollock H.M., Hammiche A., “Modulated
Differential Scanning Calorimetry: XVI. Degree of Mixing in Interpenetrating
Polymer Networks”, Thermochim Acta, Vol. 315, No. 1, 1998b, p. 25-32.
Speight J.G., The Chemistry and Technology of Petroleum, Marcel Dekker, Inc.,
New York, 1999.
Storm D.A., Sheu E.Y., “Characterization of Colloidal Asphaltenic Particles in
Heavy Oil”, Fuel, Vol. 74, No. 8, 1995, p. 1140-1145.
Struik L.C.E., “Some Problems in the Non-linear Viscoelasticity of Amorphous
Glassy Polymers”, J Non-Cryst Solids, Vol. 131-133, 1991, p. 395-407.
UOP Inc., Paraffin Wax Content of Petroleum Oils and Asphalts, UOP Method
46-85, Des Plaines, IL, 1985.
van der Hart D.L., Manders W.F., Campbell G.C., “Investigation of Structural
Inhomogeneity and Physical Aging in Asphalts by Solid State NMR”, Preprints
Division of Petroleum Chemistry, American Chemical Society, 1990, p. 445-452.
30 Road Materials and Pavements Design. Volume X – No X/2007

Vinter J.G., “Extended electron distributions applied to the molecular mechanics


of some intermolecular interactions. 2. Organic complexes”, J Comput Aid Mol Des,
Vol. 10, No. 5, 1996, p. 417-426.
Wada Y., Hirose H., “Glass Transition Phenomena and Reological Properties of
Petroleum Asphalt”, J Phys Soc Jpn, Vol. 15, No. 10, 1960, p. 1885-1894.
Wada Y., Hirose H., Asano T., Fukutomi S., “On the Dynamic Mechanical
Properties of Polymers at Ultrasonic Frequencies in Relation to Their Glass
Transition Phenomena”, J Phys Soc Jpn, Vol. 14, No. 8, 1959, p. 1064-1072.
Wessel M.D., Jurs P.C., “Prediction of Normal Boiling Points of Hydrocarbons
from Molecular-Structure”, J Chem Inf Comput Sci, Vol. 35, No. 1, 1995, p. 68-76.
Weyland H.G., Hoftyzer P.J., van Krevelen D.W., “Prediction of the glass
transition temperature of polymers”, Polymer, Vol. 11, No. 2, 1970, p. 79-87.
Williams M.L., Landel L.F., Ferry J.D., “The Temperature Dependence of
Relaxation Mechanisms in Amorphous Polymers and Other Glass-forming Liquids”,
J Am Chem Soc, Vol. 77, No. 14, 1955, p. 3701-3707.
Wunderlich B., “The Nature of the Glass Transition and Its Determination by
Thermal Analysis” in Seyler R.J. (Ed), Assignment of the Glass Transition, ASTM,
Philadelphia, PA, USA, 1994, p. 17-31.

View publication stats

You might also like