Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Materials Chemistry and Physics 146 (2014) 261e268

Contents lists available at ScienceDirect

Materials Chemistry and Physics


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

Thermo-mechanical properties and microstructural considerations


of MDI isocyanate-based bituminous foams
M.A. Izquierdo, M. García-Morales, F.J. Martínez-Boza, F.J. Navarro*
Departamento de Ingeniería Química, Facultad de Ciencias Experimentales, Universidad de Huelva, Campus de “El Carmen”, 21071 Huelva, Spain

h i g h l i g h t s

 Bituminous foams were obtained by reaction of bitumen þ MDI-PPG þ water.


 The results are explained on the basis of the poly-urethane/urea chemistry.
 Bituminous foams obtained present improved in-service properties.
 Bitumen colloidal structure is altered after this chemical modification.
 An unexpected rheological behaviour is noticed at high temperatures.

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

Article history: Stable bituminous foams have been obtained by using a 35/50 penetration grade bitumen and a reactive
Received 6 March 2013 prepolymer (MDI-PPG) synthesized by the reaction of polymeric 4,40 -diphenylmethane diisocyanate
Received in revised form (MDI) with a low molecular weight polypropylene glycol (PPG). In a first step, MDI-PPG and bitumen
20 February 2014
were allowed to react for up to 7 days. Foams were then obtained by adding an excess of water. The
Accepted 7 March 2014
results obtained point out important changes in the material colloidal structure, as a consequence of the
reactions between the eNCO groups and bitumen most polar fractions. Modification led to bituminous
Keywords:
foams with low density and improved in-service properties (higher elasticity and resistance to defor-
Polymers
Solegel growth
mation at high in-service temperatures, reduced thermal susceptibility and better flexibility at low
Thermo-mechanical analysis (TMA) temperatures), demonstrating their adequacy to be used in building applications. Properties were
Thermal properties enhanced by subjecting the material to longer periods of curing prior to promote foaming.
Differential scanning calorimetry (DSC) Ó 2014 Elsevier B.V. All rights reserved.

1. Introduction passive agents. In the case of passive modification, the material


added does not produce any type of chemical interaction with
MDI-based prepolymers are versatile engineering materials bitumen and only physical phenomena happens (polymer swelling,
widely used in the preparation of a large and varied family of phase inversion, droplets dispersion, etc.) [3e6]. By contrast, active
products, namely elastomers, foams, adhesives, binders, sealants, modifiers are reactive compounds able to alter the bitumen
etc., and new uses emerge almost every day [1]. Lately, they are chemical composition. In this sense, MDI-based prepolymers pro-
being used as new additives in the modification of bitumen, a by- vide highly reactive functional groups (isocyanates) that can react
product of petroleum refineries [2]. Besides its main use as a under mild conditions with polyfunctional active hydrogen com-
binder of mineral aggregates in paving, bitumen is also used as a pounds of bitumen, giving rise to new linkages and so higher mo-
raw material to manufacture mixtures in building construction lecular weight molecules [7e9]. Previous papers point out the
with improved physical properties. In such products, bitumen is potential use of MDI-based prepolymers as chemical modifiers of
often modified by significant amounts of other compounds, mainly bitumen to produce road paving binders, bitumen emulsions, or
polymers, giving rise to the so-called modified bitumen. roofing membranes in building construction. It is also reported that
Attending to the type of interaction between the modifying the modification degree depends on bitumen composition and
agent and bitumen, we can classify them into two groups: active or microstructure [10e13].
Thus, in such reactive systems, bitumen chemical composition is
a key factor which determines its degree of reactivity with isocy-
* Corresponding author. anate groups. However, a detailed chemical characterization of
E-mail address: frando@uhu.es (F.J. Navarro).

http://dx.doi.org/10.1016/j.matchemphys.2014.03.018
0254-0584/Ó 2014 Elsevier B.V. All rights reserved.
262 M.A. Izquierdo et al. / Materials Chemistry and Physics 146 (2014) 261e268

bitumen is highly complex. Consequently, their compounds are Table 1


usually divided by chromatography techniques into four main Penetration, Ring & Ball softening temperature, molecular weight
and SARAs fractions for the neat bitumen.
families, referred to as SARAs fractions (Saturates, Aromatics, Resins
and Asphaltenes) with molecular weight and polarity increasing in Penetration [dmm] 45
this same order. In this sense, the presence of eOH and eNH groups R&B softening point [ C] 52
Saturates (wt. %) 6.30
in its most polar fractions determine its reactivity with the isocy- Aromatics (wt. %) 46.1
anate groups of the prepolymer [9]. Resins (wt. %) 27.6
Concerning its microstructure, bitumen is considered to be a Asphaltenes (wt. %) 20.0
colloidal system in which micelles of high molecular weight organic Mw (g mol1) 1679
Mw/Mn 2.55
molecules (asphaltenes), surrounded by a shell of resins, are
dispersed into an oily phase (maltenes), consisting of low molecular
weight saturated and aromatic hydrocarbons together with the (Mw/Mn) of 1.33 and isocyanate content of 31 wt.%. The prepolymer
remaining resins. Bitumen behaviour depends on the relative MDI-PPG was synthesized by reaction of PPG and polymeric MDI, in
concentration and the chemical features of asphaltenes and mal- N2 atmosphere, at 40  C, for 48 h and under agitation. It presents an
tenes. Thus, a variation in its composition strongly affects its me- average molecular weight of 915 g mol1, polydispersity (Mw/Mn)
chanical properties [2,13]. of 1.9, and isocyanate content of 22.9 wt.%. The molecular weight
On the other hand, MDI-based prepolymers are extensively used distributions corresponding to the raw materials are included in
to prepare polyurethane foam products, widely used in applica- Fig. 1A).
tions such as furniture industry, packing, coatings, decorating,
building construction, insulation, shoe industry and transportation. 2.2. Samples preparation
With regard to building applications, polyurethane rigid foams and
sprayed foams have extensively been used for thermal insulation as Two set of samples, modified bitumen and modified bituminous
the ultimate energy savers, since the air trapped within the foams, were prepared:
honeycomb-like structure confers quite low thermal conductivity. Firstly, modified bitumen was obtained by mixing bitumen and
Examples in building applications are layers of floor and ceiling 10 wt.% of MDI-PPG at 90  C for 1 h, in a cylindrical glass vessel
coverings, roofing boards, sheathing, perimeter insulation, sprayed (60 mm diameter, 140 mm height). Agitation of 1050 r.p.m. was
walls/ceilings, industrial tanks, curtain wall panels, etc. [1,14]. conducted with a four-blade turbine, by using an IKA RW-20 stir-
Polyurethane foams commonly result from the reaction of iso- ring device (Germany). Samples of the resulting modified bitumen
cyanate based prepolymers in the presence of water. The isocyanate were set, inside closed vessels, in an oven at 90  C, and allowed to
groups have three functions: a) first, they react with water to form cure for 24 h or 7 days.
CO2, which promotes foaming; b) second, they serve as a di- or On the other hand, bituminous foams were prepared in a second
polyfunctional reactant, joining polyol molecules together, from step. Thus, an excess of water (2 wt.%) was added to the previous
their reaction with hydroxyl groups. Thus, the entire system builds modified bitumens. Water reacted with the free eNCO groups
itself into a highly cross-linked polymer; c) third, the isocyanate
groups produce polyurea linkages. The polyurea linkages become
part of the polymer structure and provide rigidity and thermal
resistance to the foam [1].
With the aim to combine the waterproofing properties of
modified-bitumen membranes along with the thermal insulation
capacity of polyurethane foams, a new product is proposed here: a
bituminous foam. This novel material is expected to present
improved mechanical properties with respect to the base bitumen,
reduced density and be dimensionally stable [15,16].
In the present paper, we investigate the rheological and thermal
behaviour of stable foams produced by reaction of water with free
isocyanate groups of a mixture of a hard bitumen with an MDI-
derived prepolymer.

2. Experimental

2.1. Materials

A 35/50 penetration grade bitumen (EN 12591), provided by


Construcciones Morales S.A. (Spain), was used as base material for
polymer modification. Chemical composition and results of
selected technological tests are included in Table 1.
The polymer used was a polypropylene-glycol (PPG) function-
alized by polymeric MDI (4,40 -diphenylmethane diisocyanate),
henceforth MDI-PPG. The PPG, Alcupol D-0411, donated by Repsol
(Spain) presents an average molecular weight of 440 g mol1, a
polydispersity (Mw/Mn) of 1.1 and a hydroxyl index of 280 mg KOH/
g. The polymeric MDI, supplied by TH Tecnic (Spain), is a mixture of
reactive monomers (4,40 -MDI, 2,40 - and 2,20 - isomers) and oligo-
mers (three-ring and poly-ring-containing isocyanates oligomers) Fig. 1. Molecular weight distributions: A) PPG, polymeric MDI and MDI-PPG; B) Neat
with an average molecular weight of 363.5 g mol1, a polydispersity bitumen and modified bitumen cured at 90  C for 7 days.
M.A. Izquierdo et al. / Materials Chemistry and Physics 146 (2014) 261e268 263

(which did not react with bitumen) and generated CO2, which was mixtures of bitumen containing 10 wt.% of MDI-PPG were prepared
used as chemical foaming agent to produce stable bituminous for 1 h and further cured for 24 h or 7 days. Both preparation and
foams. curing were conducted at 90  C. Viscous flow curves of these
After processing, all the samples were covered with aluminium samples, presented in Fig. 2, show that curing gives rise to a sig-
foil and maintained at 4  C. nificant increase in viscosity, with the change observed after just
24 h of curing being very noticeable. Paradoxically, the non-cured
2.3. Samples testing sample presents lower viscosity than its parent base bitumen.
This leads to a negative value of its modification index (see M.I. in
Viscous flow measurements, at 60  C, and temperature sweep Table 2), defined as the increase in viscosity after bitumen modi-
tests in oscillatory shear between 30 and 240  C (heating rate: fication relative to its original value:
1  C min1; frequency: 10 rad s1; strain: 1%), were conducted in a
controlled-stress rheometer RS-150 from Haake (Germany), using a
h0;sample  h0;neat
M:I: ¼ (1)
plate-and-plate geometry (10, 20 and 35 mm diameter, 1 mm gap). h0;neat
At least two replicates of each test were done.
DMTA experiments were done with a Seiko DMS 6100 device where h0 represents the Newtonian viscosities, that is, at the lowest
from Seiko Instruments (Japan), using 50  10  3 mm samples in values of shear rate.
double cantilever bending mode. All the experiments were carried As the MDI-PPG prepolymer is a liquid with a viscosity of
out at constant frequency of 10 Hz and a strain value within the approximately 1.2 Pa$s at 60  C, which is nearly three orders of
linear viscoelastic region. The selected temperature ramp was magnitude below the viscosity of bitumen at the same tempera-
2  C min1. Temperature ranged from 40 to 50  C. Liquid nitrogen ture, its plasticizing effect prevails over the effect of its reaction in
was used for cooling. the non-cured sample [12].
Modulated differential scanning calorimetry (MDSC) tests were More relevant information, in a wider temperature range, is
developed with a calorimeter Q-100 (TA Instruments, USA). Sam- provided by temperature sweep tests in oscillatory shear, portrayed
ples of 5e10 mg were pressed into hermetic aluminium pans and in Figs. 3 and 4. Neat and modified bitumens are characterized by
allowed to anneal for 7 days at room temperature before being values of G0 and G00 which monotonously decrease with increasing
analysed. As bitumen is a complex material composed of highly temperature. However, their viscoelastic behaviour differs as a
time-dependent structures, the annealing time selected ensures a consequence of the reactive modification. Thus, neat bitumen
fully developed stable microstructure [2,12]. The test protocol was curves directly go from the glass-transition to the viscous flow re-
as follows: samples were cooled down to 80  C, equilibrated at gion. Instead, modification provokes a shoulder in G0 around 80  C
that temperature for 5 min, and subsequently subjected to a (specially marked for the fresh and 1 day-cured samples) which
modulated heating ramp up to 150  C. A heating rate of 5  C min1, shifts the appearance of the viscous flow region up to higher
amplitude of modulation of 0.5  C, and a period of 60 s were used temperatures. Moreover, modification increases, mainly after 7
for all the samples tested. Nitrogen was used as purge gas, with a days of curing, both the elastic and viscous moduli (although in a
flow rate of 50 mL min1. different extent). In relation with the above comments, the loss
FTIR spectra were obtained with a Digilab FTS3500ARX (Varian) tangent evolution in Fig. 4 reveals samples with G0 values higher
apparatus. Samples of modified bitumen and foams were dissolved than G00 values (tand < 1) up to temperatures in the range 35e45  C,
in dichloromethane in a concentration of 2.8 wt.% (0.7 g sample in depending on the sample nature. For higher temperatures, all the
25 mL dichloromethane). After 1 day, the resulting solutions were samples showed a prevailing viscous behaviour. However, the
filtered using glass microfiber filters with a pore size of 1.5 mm. above described increases in G0 , when bitumen was modified and
Then, a drop of solid-free solution was laid on a potassium bromide cured, result to be more significant than their corresponding in-
(KBr) thin plate and the solvent was then evaporated. The spectra creases in G00 , as indicated by lower values of tand. Consequently,
were obtained in a wavenumber range of 400e4000 cm1 at elasticity in the region of high in-service temperature is enhanced.
4 cm1 resolution in the absorbance mode. In addition, all modified bitumens display a well-defined maximum
The insoluble fraction in dichloromethane, retained by the filter, in the loss tangent function (Fig. 4), which derives from the
was also quantified.
Bitumen SARAs fractions were determined by thin layer chro-
matography coupled with a flame ionization detector (TLC/FID),
using an Iatroscan MK-6 analyzer (Iatron Corporation Inc., Japan).
Elutions were performed in hexane, toluene, and a mixture of
dichloromethane/methanol (95/5 vol./vol.), following the proce-
dure outlined elsewhere [17].
Foam density was measured according to ASTM 1622 [18].
Molecular weight and polydispersity were determined by GPC,
with a Waters 2414 Refractive Index Detector, a StyragelÒ HR 4E
column and THF as solvent. To convert the elution profiles to mo-
lecular weight distributions, raw data were transformed by using a
calibration curve obtained from the elution times of standard poliol
compounds through the column.

3. Results and discussions

3.1. Polymer modified bitumen

In order to promote reactions between eNCO groups (in pre- Fig. 2. Viscous flow curves, at 60  C, for 10 wt.% MDI-PPG modified bitumens (different
polymer) and eOH/eNH groups (in bitumen most polar fractions), curing times), and their corresponding foams.
264 M.A. Izquierdo et al. / Materials Chemistry and Physics 146 (2014) 261e268

Table 2 down to 2.8  C. This fact suggests the onset of the glassy region to
Some technological and rheological parameters of modified bitumens and their be shifted more than 6  C below the non-modified bitumen.
corresponding bituminous foams.
The low in-service properties can also be analysed by means of

Curing Tmax tand
h0 (Pa s) M.I. Slope tand4070 C
modulated differential scanning calorimetry, a technique that al-
time ( C) lows for the separation of glass transitions from order-disorder
neat e 6.9  102 0 2.7  102 transitions. In Fig. 6, the derivative heat capacity curves, dCp/dT,
bitumen help to identify the glass transition region and to obtain a “calori-
Modified 1h 77.4 6.3  102 0.1 2.5  102
metric” Tg. This figure clearly shows that all the samples present, at
bitumen 24 h 79.4 2.8  103 3 2.1  102 DSC DSC
7 days 92.9 3.5  103 3.9 1.7  102 least, two well-defined glass transitions, Tg1 and Tg2 (Table 3),
Bituminous 1h 1.8  106 2.6  103 derived from distinct amorphous bituminous phases, attributed to
foams 24 h 1.6  106 2.3  103 the overlap of glass transitions corresponding to several bitumen
7 days 2.3  107 3.4  104
fractions [23,24]. As Tg increases with the polarity, aromaticity and
molecular weight of the amorphous phase from which it originates,
the Tg at 33  C mainly arises from the maltene phase and the
shoulder in G0 observed in Fig. 3. Such a rheological transition or second one, centred at 2  C, arises from a malteneeasphaltene
relaxation process may be attributed to a polymer-rich dispersed interfacial region of mixed composition likely rich in resins [23e
DSC
phase arisen from the reaction of bitumen and prepolymer [19,20]. 25]. For the 10 wt.% MDI-PPG modified bitumen, both Tg1 and
DSC
In that sense, longer curing times yield the shift of this transition Tg2 undergo a significant decrease, which is in agreement with the
temperature towards higher values (Table 2). This fact reveals the evolution observed in the “mechanical” glass transition tempera-
development of more temperature-resistant and complex micro- tures (Table 3). The discrepancy between the Tg values obtained
structures in the resulting material, during the 90  C-curing stage. from different techniques is not surprising, as the macroscopic
Moreover, the slope of the loss tangent vs. temperature plot, be- character of the bending technique and also its frequency depen-
tween 40 and 70  C (see Fig. 4), may be taken as a parameter which dence make the position of the loss modulus maximum removed
quantifies the material thermal susceptibility over the from the genuine glass transition [26]. In general, the decrease
intermediate-to-high temperature range. As can be seen in Table 2, observed in the Tg values may be partially attributed to the plasti-
the values of the slope decrease after MDI-PPG addition, mainly for cizing effect of unreacted MDI-PPG. In fact, as can be seen in Table 4,
the longest curing time. Consequently, an improved thermal the residual eNCO content obtained from FTIR measurements is
resistance of the modified bitumen should be expected [21,11]. still large even after curing at high temperature. Consequently, an
With the aim to get a deeper insight into the material me- important quantity of liquid prepolymer remains dissolved in the
chanical response at low temperature, neat bitumen and 24 h- modified bitumen.
cured modified bitumen were analysed by using DMTA in bending Again, Tg analysis demonstrates a bituminous material from
mode at 10 Hz (Fig. 5). The glassy region, characterized by nearly which enhanced performance at low in-service temperatures may
constant values of the storage modulus (E0 ), around 109 Pa, and a be expectable [27].
maximum in the loss modulus (E00 ), is clearly observed for both As a conclusion, MDI-PPG addition to hard bitumen and further
00
systems. Interestingly, the temperature at that maximum (Tgmax E ) curing, at 90  C, of the resulting blend give rise to a new bitumen-
defines a “mechanical” glass transition temperature, which relates polymer material which presents improved in-service properties:
to the material resistance to thermal fracture [22]. In that sense, higher elasticity, overall resistance to deformation and reduced
even though this parameter is frequency-dependent, it defines the thermal susceptibility at high in-service temperatures, and better
transition to the brittle state of the material, from a practical point cracking resistance at low temperatures.
of view. According to the results presented in Table 3, the addition In general, these results can be explained on the basis of the
of MDI-PPG to bitumen and further curing for 24 h at 90  C is seen isocyanate chemistry. Thus, at 90  C, in closed vessels and in
to reduce the “mechanical” glass transition temperature from 3.3  C absence of water from air, eNCO groups in the prepolymer are

Fig. 3. Elastic and loss moduli obtained from temperature sweep tests in oscillatory shear, at 10 rad s1, for MDI-PPG modified bitumens (different curing times), and their cor-
responding foams.
M.A. Izquierdo et al. / Materials Chemistry and Physics 146 (2014) 261e268 265

3.2. Modified bituminous foams

Three bituminous foams were obtained from the previous


modified bitumens. They were produced by the standard procedure
followed for polyurethane foams [28]. Thus, the chemical foaming
process involves the reaction between water and free eNCO (R3),
producing an amine and releasing carbon dioxide, which promotes
bitumen expansion:

R1  NCO þ H2 O/R1  NH2 þ CO2 [ (R3)

Additionally, the reaction between the resulting amine groups,


from R3, and the remaining eNCO groups occurs during the
blowing process (see R2).
Consequently, carbon dioxide acts as a chemical blowing agent
which leads to the formation of gas bubbles. These achieve their
maximum size when the viscosity of the bituminous phase is high
Fig. 4. Loss tangent values obtained from temperature sweep tests in oscillatory shear,
at 10 rad s1, for MDI-PPG modified bitumens (different curing times), and their cor-
enough so that it stops bubbles growing. In fact, an increase in the
responding foams. material viscosity of several orders of magnitude (Fig. 2) prevents
foam collapsing, as bubbles are confined within a continuous
bituminous phase which hardens whilst Reaction R2 proceeds. This
expected to react with bitumen most polar compounds containing result becomes even more noticeable for the longest curing time (7
active hydrogen atoms, namely, hydroxyl and amine groups, lead- days).
ing to the formation of urethane and urea linkages, respectively (R1 It is important to underline that all the bituminous foams
and R2) [10]: studied have similar densities (350e400 kg m3). This allows a
comparative analysis of their thermo-mechanical properties to be
established with no need for previous density correction.
R1  NCO þ R2  OH/R1  NH  COO  R2 (R1)
Regarding their linear viscoelastic response, the foams present,
as shown in Fig. 3, a dramatic change with respect to their parent
modified bitumens, from a predominant viscous behaviour of the
R2  NCO þ R1  NH2 /R2  NH  CO  NH  R1 (R2)
latter, to the prevailing elastic properties of the former. Moreover,
Consequently, the molecular weight and the complexity of the loss tangent values in Fig. 4 clearly fall below one, with a minimum
resulting molecules and structures increase along the curing pro- at around 120  C which demonstrates the development of a
cess. In that sense, Fig. 1B) reveals an increase in the bitumen rubbery plateau [29] after the foaming process. Additionally, the
molecular weight after a curing period of 7 days. The appearance in temperature interval over which the plateau above extends in-
the modified bitumen of larger molecules than the corresponding creases as curing time does, which denotes a lower thermal sus-
to the polymer and neat bitumen necessarily implies chemical re- ceptibility of this material at in-service temperatures (Table 2).
actions. Moreover, as GPC actually measures the molecular size and At low temperatures, the linear viscoelastic behaviour,
shape of the species, the development of new larger and more measured by dynamic flexural tests, is also altered (Table 3 and
complex molecules during the high temperature curing is Fig. 5). On the one hand, foaming produces a very notable decrease
confirmed. In fact, they are thought to be the responsible for the in the material overall stiffness under small deformations, due to
previously mentioned relaxation process observed in the loss the high volume occupied by the bubbles, as revealed by lower
tangent function in Fig. 4. values of E0 and E00 . On the other hand, if compared to the material

Fig. 5. Elastic and loss moduli obtained from temperature sweep tests in the bending mode, at 10 Hz, for 24 h-cured MDI-PPG modified bitumen and its corresponding foam.
266 M.A. Izquierdo et al. / Materials Chemistry and Physics 146 (2014) 261e268

Table 3 Table 4
“Mechanical” and “calorimetric” glass transition temperature of selected samples. Some chemical parameters of the modified bitumens and their corresponding
E00
bituminous foams.
Curing time Tmax ( C) TDSC 
g1 ( C) TDSC 
g2 ( C)
Curing IcR Gel fraction NCO NCO
neat bitumen e 3.3 32.6 2.0
time (wt.%) content conversion
Modified bitumen 24 h 2.8 36.5 6.7
(wt.%) (%)
Bituminous foam 24 h 0.2 33.5 2.3
neat 0.358 e e e
bitumen
Modified 1h 0.495 e 10.6 53.6
before foaming, the “mechanical” and “calorimetric” glass transi- bitumen 24 h 0.529 e 9.4 58.7
tion temperatures are moved up to higher values, as a consequence 7 days 0.676 e 9.1 60.1
of the reactions occurred when the foam is promoted. Those in- Bituminous 1h 0.578 8.1 2.3 90.0
foams 24 h 0.649 14.6 2.4 89.6
crease the degree of polymer crosslinks corresponding to the
7 days 0.800 17.1 2.4 89.6
bitumen continuous phase (higher density of urethane/urea
bonds), providing a greater rigidity [30]. However it is important to
The results shown in Fig. 7A and B can also be explained on the
underline that Tgs of foam remain below the corresponding values
basis of the poly-urethane/urea chemistry. Hence, Reactions R1 and
of neat bitumen. This result is very significant since bituminous
R2 are reversible at high temperatures, leading to the partial
products are always used above their glass transition temperatures,
thermal decomposition of urethane or urea bonds into their orig-
otherwise they present brittle properties and tend to crack under
inal isocyanate/alcohol or isocyanate/amine groups, respectively.
stresses. Consequently, from a practical point of view, low tem-
Further crosslinking reactions at high temperature, involving these
perature properties are improved.
groups and original compounds, may occur [31]. This complex set
Finally, an unexpected behaviour is noticed in the very high
of reactions, whose extent increases as temperature does, produces
temperature region studied in Fig. 3, as both elastic and viscous
a very notable enhancement in the shear complex modulus [32].
moduli pass through a minimum and then experience a remarkable
increase with temperature. Interestingly, the onset temperature at
which this phenomenon occurs increases with curing time. 3.3. Chemistry and microstructure
In order to get a deeper insight into its origin, a selected bitu-
minous foam (cured for 7 days) was submitted to different tem- As previously mentioned, bitumen modification would result
perature profiles. Fig. 7A presents the evolution of the shear from different reactions (R1 to R3). During the blending and curing
complex modulus with time, at a fixed frequency of 10 rad s1 and processes at 90  C, free isocyanate groups of the prepolymer react
deformation within the linear viscoelasticity interval. Temperature with active hydrogen-containing compounds of bitumen, leading
was raised at a constant rate of 1  C min1, from 30  C up to 150  C to a progressive decrease in the eNCO groups, which becomes
(below the onset temperature, 197  C, in Fig. 3), and then it was larger for the longest curing time. This reduction was quantified by
maintained constant. As expected, only a slight increase in jG*j is means of normalised peak areas corresponding to the stretching
noticed during the isothermal step, probably due to hardening mode of the N]C]O group at 2275 cm1 from FTIR analysis (Fig. 8,
resulting from the oxidation of the bituminous phase and by curing Table 4).
of residual MDI-PPG and bitumen. By contrast, when the foam is
heated up to 210  C (above its onset temperature, 197  C, in Fig. 3),
the shear complex modulus undergoes a remarkable increase of
more than two orders of magnitude during an isothermal step
(Fig. 7B). If the sample is further cooled down to 70  C at the same
rate, jG*j continues to increase. Consequently, an irreversible
chemical process triggers when temperature achieves the value
corresponding to the minimum peaks observed in Fig. 3.

Fig. 6. First derivate of the Heat Capacity, obtained from MDSC tests, for neat bitumen, Fig. 7. Evolution of the complex shear modulus with time, at different temperature
24 h-cured MDI-PPG modified bitumen and its corresponding foam. profiles, for a selected foam derived from non-cured MDI-PPG modified bitumen.
M.A. Izquierdo et al. / Materials Chemistry and Physics 146 (2014) 261e268 267

After the curing, the presence of urethane/urea linkages can


readily be observed due to their characteristic peaks at 1728/
1650 cm1, related to the stretching vibrations of the carbonyl
groups in urethane/urea, a wide peak centred at 3400 cm1 due to
the eNH stretching vibrations and urethane/urea NeH bending
vibration absorptions at 1521/1504 cm1 (Fig. 8). In general, Fig. 8
shows an increase in the peaks described as longer curing time
are considered, with the consequent shift of the molecular weight
distribution up to higher values (Fig. 1B)).
Furthermore, Fig. 9 reveals changes in the relative proportions
of the bitumen SARAs fractions as a result of the chemical modifi-
cation by MDI-PPG prepolymers. Thus, as the curing time increases,
the reaction proceeds with the consequent increase in the eNCO
conversion. The aromatic fraction is seen to decrease in favour of an
increase in the asphaltene content, by following the sequence:

MDIPPG MDIPPG
Aromatics ƒƒƒ! Resins ƒƒƒ! Asphaltenes (2)
Thus, bitumen colloidal structure is altered after this chemical
modification, which may be better understood by using a “modi-
fied” version of the so-called Gaestel colloidal index, defined in
terms of the SARAs fractions determined by thin layer chroma-
tography [13]:

saturates þ asphaltenes þ ðNCO=polarsÞ


IcR ¼ (3)
aromatics þ resins

In the formula above, NCO/polars refers to those species resulted


from chemical reactions between the prepolymer and polar

Fig. 9. SARAs fractions for: A) Neat and modified bitumens, as a function of curing
time; B) Neat bitumen, non-cured modified bitumen and its corresponding foam.

bitumen compounds which are not eluted by any of the solvents


employed in the separation sequence. Those new species are
detected at the asphaltene peak position. For neat bitumen NCO/
polars content is zero.
IcR values in Table 4 reveal larger asphaltene clusters by reaction
of polar groups in the asphaltene micelles as the curing time in-
creases, which has shown to lead to materials with enhanced gel-
like characteristics (i.e. higher elasticity) [13]. In fact, previous re-
sults obtained by modification of softer bitumen with MDI-PPG
demonstrate that, after a step of curing, larger and more compact
asphaltenic regions are developed through the resulting material
[15].
With regard to the foams, around 90% of the eNCO groups is
converted into new urethane/urea linkages (see more developed
peaks in FTIR spectrum in Fig. 8). In addition to Reaction R1,
Reactions R2 and R3 extensively occur when water is added to the
modified bitumen. SARAs fractions are also modified in the same
sense, as indicated by sequence (2).
Finally, Table 4 presents the results of percentage of insoluble
matter in dichloromethane (retained by a 1.5 mm-filter), carried out
on foams samples. It has been referred to as “gel” fraction. It is
interesting to note that larger amounts of this fraction are obtained
if the modified bitumen has been cured for longer times prior to its
foaming. This fact would be related to the progressive formation of
multifunctional reactive molecules of high molecular weight dur-
Fig. 8. FTIR spectra for selected systems from: A) 2000 to 3700 cm1; B) 1500 to ing the curing process. These molecules may act as crosslink
1800 cm1. junctions in further foaming reactions, being more effective on
268 M.A. Izquierdo et al. / Materials Chemistry and Physics 146 (2014) 261e268

building up more complex structures. Consequently, it would References


provide an explanation to the improvement in the rheological
behaviour observed. However, further studies need to be conducted [1] M. Szycher, Szycher’s Handbook of Polyurethanes, CRC Press, Boca Ratón,
1999.
in order to clarify the role of both gel and sol phases on the thermo- [2] F.J. Navarro, P. Partal, F. Martínez-Boza, C. Gallegos, J.C.M. Bordado, A.C. Diogo
mechanical behaviour of bituminous foams. Mech, Time Depend. Mater. 10 (2006) 347e359.
[3] A.H. Fawcett, T. McNally, Colloid. Polym. Sci. 281 (2003) 203e213.
[4] A.H. Fawcett, T. McNally, Polymer 41 (2000) 5315e5326.
4. Conclusions [5] H. Zhang, J. Yua, H. Wangb, L. Xue, Mater. Chem. Phys. 129 (2011) 769e776.
[6] D. Lesueur, Adv. Colloid Interface Sci. 145 (2009) 42e82.
The results obtained support the use of the reactive polymer [7] G. Polacco, J. Stastna, D. Biondi, F. Antonelli, Z. Vlachovicova, L. Zanzotto,
J. Colloid. Interface Sci. 280 (2004) 366e373.
proposed, synthesized by reaction of polymeric 4,40 -dyphenyl- [8] B. Singh, H. Tarannum, M. Gupta, J. Appl. Polym. Sci. 90 (2003) 1365e1377.
methane diisocyanate (MDI) and a low molecular weight poly- [9] A. Topal, Fuel Process Technol. 91 (2010) 45e51.
propylene glycol (PPG), in the manufacture of modified bituminous [10] M.J. Martín-Alfonso, P. Partal, F.J. Navarro, M. García-Morales, C. Gallegos, Eur.
Polym. J. 44 (2008) 1451e1461.
foams. A preparation protocol consisting in blending hard bitumen
[11] F.J. Navarro, P. Partal, M. García-Morales, F. Martínez-Boza, C. Gallegos, Fuel 86
(35/50 pen.) with 10 wt.% MDI-PPG, followed by a step of curing, of (2007) 2291e2299.
up to 7 days, and further addition of water was seen to produce [12] M.J. Martín-Alfonso, P. Partal, F.J. Navarro, M. García-Morales, C. Gallegos, Ind.
Eng. Chem. Res. 47 (2008) 6933e6940.
foams with reduced density and enhanced rheological properties at
[13] V. Carrera, P. Partal, M. García-Morales, C. Gallegos, A. Páez, Ind. Eng. Chem.
both low and high in-service temperatures. Res. 48 (2009) 8464e8470.
In general, the thermo-rheological properties of the modified [14] A. Guo, I. Javni, Z. Petrovic, J. Appl. Polym. Sci. 77 (2000) 467e473.
bitumens and their resulting bituminous foams are improved if [15] M.A. Izquierdo, F.J. Navarro, F.J. Martínez-Boza, C. Gallegos, Fuel 90 (2011)
681e688.
longer curing times are chosen. This effect is attributed to the [16] M.A. Izquierdo, F.J. Navarro, F.J. Martínez-Boza, C. Gallegos, Constr. Build.
progressive formation of multifunctional reactive molecules of high Mater. 30 (2012) 706e713.
molecular weight during the curing step. These reactive molecules [17] A. Eckert, Petrol. Coal 43 (2001) 51e53.
[18] American Society for Testing and Materials. Standard Test Method for
still present a significant quantity of non-reacted isocyanate groups Apparent Density of Rigid Cellular Plastics. ASTM D-1622e1703.
(available to react with water). Hence, they have demonstrated to [19] M.A. Izquierdo, F.J. Navarro, F.J. Martínez-Boza, C. Gallegos, Rheol. Acta 53
be very efficient in enhancing the foams properties at the final step. (2014) 123e131.
[20] A.A. Cuadri, V. Carrera, M.A. Izquierdo, M. García-Morales, F.J. Navarro, Constr.
In fact, Reactions R1 to R3 lead to larger and more complex mole- Build. Mater. 51 (2014) 82e88.
cules, as seen by significant changes in the bitumen SARAs fractions [21] S. Chebil, A. Chaala, C. Roy, Fuel 79 (2000) 671e683.
and so, in its colloidal structure. [22] A.A. Cuadri, P. Partal, F.J. Navarro, M. García-Morales, C. Gallegos, Energy Fuels
25 (2011) 4055e4062.
As a result, the bituminous foams obtained present improved in-
[23] J.F. Masson, Energy Fuels 22 (2008) 2637e2640.
service properties with respect to their original modified bitumen: [24] J.F. Masson, G.M. Polomark, P. Collins, Energy Fuels 16 (2002) 470e476.
higher elasticity and resistance to deformation at high in-service [25] J.F. Masson, G.M. Polomark, P. Collins, Thermochim. Acta 43 (2005), 696e100.
[26] E. Bormashenko, R. Pogreb, S. Sutovsky, V. Lusternik, A. Voronel, Infrared Phys.
temperatures, reduced thermal susceptibility and better cracking
Technol. 43 (2002) 397e399.
resistance at low temperatures. [27] X. Lu, U. Isacsson, J. Ekblad, Fuel 36 (2003) 652e656.
[28] D. Eaves, Handbook of Polymer Foams, Shawbury Smithers Rapra, 2004.
Acknowledgements [29] J. D.Ferry, John Willey & Sons, Inc., New York, 1980.
[30] L. Xiaobin, C. Hongbin, Y. Zhang, Sci. China Ser. B 49 (2006) 363e370.
[31] A.W. Douglas, W Zeno Wicks Jr., Prog. Org. Coat. 36 (1999) 148e172.
This work is part of a research project sponsored by a MEC- [32] S. Subramani, Y.J. Park, Y.S. Lee, J.H. Kim, Prog. Org. Coat. 48 (2003) 71e79.
FEDER Programme (Research Project MAT2007-61460) and by a
Junta de Andalucía Programme (TEP6689). The authors gratefully
acknowledge its financial support.

You might also like