Sie sind auf Seite 1von 13

Fuel 113 (2013) 454466

Contents lists available at SciVerse ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Tongyan Pan

h i g h l i g h t s
 At 130 C, asphalt generates alkanes and sulfoxides more rapidly than ketones.
 Amine groups do not contribute much to asphalt aging.
 Chain breaking, sulfoxidation, and ketonization are vital to asphalt aging.
 Coniferyl-alcohol lignin shows radical-scavenging effect in bulk asphalt.

a r t i c l e

AC

 Coniferyl-alcohol lignin is most effective when temperature is below 130 C.

i n f o

a b s t r a c t

Article history:
Received 15 April 2013
Received in revised form 2 June 2013
Accepted 4 June 2013
Available online 19 June 2013

Petroleum asphalt is an important base material for many industrial applications, such as the binding and
waterproong component in road pavements and roof shingles. Being an organic end product of petroleum serving under the general open-to-air conditions, asphalt can lose the desired rheological properties
with time due to oxidative hardening or aging that frequently leads to increase in viscosity, separation of
components, and loss of cohesion and adhesion, and thereby becomes hardened. A common practice to
alleviate asphalt aging today is using different chemical additives or modiers as antioxidants. The
current state of knowledge in asphalt oxidation and antioxidant evaluation is focused on monitoring
the degradation in asphalts physical properties, mainly the viscosity and ductility, which although
satisfying direct engineering needs does not contribute to the fundamental understanding of the aging
and anti-aging mechanisms. Within this context, this study was initiated to study the anti-oxidation
mechanisms of bio-based additives, using the coniferyl-alcohol lignin as an example, by developing a
quantum chemistry based chemophysical environment in which the various chemical reactions among
asphalt components, anti-oxidative additive and oxygen, as well as the incurred physical changes can
be studied. The techniques of X-ray photoelectron spectroscopy (XPS) was used to prove the validity
of the modied and unmodied asphalt models, from which the XPS results showed high agreement
to the model predictions.
2013 Elsevier Ltd. All rights reserved.

ET

Keywords:
Atomistic modeling
Asphalt
Oxidative aging
Coniferyl alcohol
Lignin

TE

The Catholic University of America, 620 Michigan Avenue, N.E., Washington, DC 20064, United States

Coniferyl-alcohol lignin as a bio-antioxidant for petroleum asphalt:


A quantum chemistry based atomistic study

1. Introduction

Being an organic product from the remains of ancient organisms, petroleum asphalt is an important base material for many
industrial purposes, such as the primary binding and waterproofing component in road pavements and roof shingles [1]. Today
around 95% of the roads we drive on are covered with asphalt
mixtures. Asphalt in its general service conditions however is subject to chemical oxidation by reactions with atmospheric oxygen,
which can cause the hardening of asphalt and the sacrice of its
desirable physical properties. In asphalt pavements, for example,
oxidative hardening is responsible for mixture embrittlement that
Tel.: +1 (202) 319 5165; fax: +1 (202) 319 6677.
E-mail address: pan@cua.edu
0016-2361/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.fuel.2013.06.003

could speed up pavement cracking [24]. As a roong material,


asphalt embrittlement from oxidative hardening can promote the
loss of protective granules, substrate shrinkage, and cracking [1].
The primary factor that controls the rate of asphalt aging is
temperature. In engineering practices, most aging occurs in the
process of making asphalt-based end products, such as asphalt
concrete in the intense heat of an asphalt plant [2,3]. During the
short duration of heating, the temperature can reach up to
165 C. Long-term oxidative aging begins immediately after the
end product, such as an asphalt pavement or roof, is put in service,
which occurs at a much slower rate than the initial oxidation during mixing and production. However, long-term oxidative aging
can cause failures as serious as the initial oxidation, including a
large increase in stiffness and loss of ductility, and cracking of
the end products under thermal and/or load stresses [1,4].

455

T. Pan / Fuel 113 (2013) 454466

TE

structures of lignin are benzene rings with attached hydroxyl


groups. Benzene rings are six carbon structures with each carbon
sharing a single and double covalent bond to another carbon. In
a phenolic group, there can be one or more hydroxyl groups attached to the benzene ring. The ability of phenolic compounds to
be antioxidants is the functional groups ability to neutralize free
radicals [16]. A phenol can neutralize a free radical by donating
either a proton or an electron [15]. Because of its structure, a phenol is able to do both while remaining relatively stable. Lignin contains a large amount of phenolic groups, making it an effective
antioxidant [16]. Many factors affect the antioxidant ability of lignin, of which the source of biological origin is the most prominent
factor [15]. Each plant is biologically and chemically different;
therefore, the lignin obtained from them will be different. The
extraction method is also very important in determining a lignins
antioxidant ability. Lignin can be extracted from plant material by
chemicals such as ethanol, acetone, acetic acid, methanol, and propanol, which produce lignins of different antioxidant abilities [15].
Lignins however can also get oxidized when exposed to oxygen
or atmosphere, which is similar to the oxidation of asphalt binder
but at a slower pace. At raised temperature (e.g., P150 C),
oxidation of lignins can be signicant. Based on a comprehensive
study of oxidation of lignins from different sources (150 C in
oxygen for 4 h), the combined gasliquid chromatography and a
sequential methylation technique indicated the presence of a
larger number of products, including different acids of low molecular weights (e.g., formic acid, two-carbon-atom acids, and some
three-carbon-atom and four-carbon-atom acids), and various
phenolic compounds such as vanillin, syringaldehyde, and
p-hydroxybenzaldehyde as the methyl ethers and carboxycontaining compounds as the methyl esters or methyl ether esters
[17]. According to Bryan, oxygen in the early stages of oxidation is
necessary to break lignin polymer into smaller molecular fragments [18]. Notably, such a level of temperature, i.e., 150 C, is
close to those at which many asphalt products are produced (e.g.,
asphalt mixtures), and might be of signicance when lignin is
evaluated as an asphalt antioxidant.
Within the current state of knowledge, a research study was recently initiated to explore the molecular-scale mechanisms of asphalt oxidation and anti-oxidation by different antioxidants,
aiming at elucidating the chemical bases of asphalt oxidative hardening and identifying effective antioxidants for asphalts. This manuscript presents the results from the study, focusing on reporting
the development of a quantum chemistry (QC) based atomistic
model, validation of the model, and utilization of the model in
studying asphalt oxidation and the coniferyl-alcohol lignin as a potential antioxidant.

ET

AC

Petroleum asphalt is composed of carbon and hydrogen, and


nitrogen, sulfur and oxygen of lower percentages [2,4]. Trace heavy
metals such as vanadium and nickel often are also present. These
elements combine to form the main fractions of asphalt cement:
asphaltenes, saturates, naphthalene and polar aromatics (also
known as light and heavy resins, respectively) according to the
Corbetts method [5]. The three fractions, each providing different
properties of asphalt, chemically and physically interact with each
other and form the complex mixture system of asphalt. Asphaltenes and saturates are normally incompatible compounds, and
are brought together by aromatics. Asphaltenes are the main contributors of viscosity and therefore hardening effects; and an abundance of aromatics and saturates decrease the ductility and hence
elastic effects [410].
In general, oxidative hardening of asphalt is believed to be
caused by the generation of oxygen-containing polar chemical
functionalities on asphalt molecules, which in turn can cause
agglomeration among molecules due to increased chemophysical
associations such as hydrogen bonding, van der Waals force, and
Coulomb force [410]. In addition to oxygen-containing functionalities, oxidation also can cause aromatization of certain asphalt
molecules that facilitates further agglomeration of asphalt components in ambient conditions [510]. However, the systematic study
of asphalt oxidation from the chemical perspective has not been
seriously attempted due to one underlying challenge in the state
of knowledge: the complexity of chemical composition in asphalt,
which has naturally led to the lack of an effective tool for studying
the manifold chemical reactions involved in asphalt oxidation
[3,8].
Chemical agents have been studied widely as antioxidants in
asphalt. The presence of polymers in asphalt, among various
chemical antioxidants, can effectively decrease aging rate and
reduce asphalts temperature susceptibility, such as the use of
styrenebutadienestyrene (SBS) and styrene-b-butadiene (SBR)
in high-volume roadways to increase the high- and low-temperature properties of asphalt and reduce the pavements tendency to
form ruts or crack [11]. There are many polymers that have shown
anti-oxidation abilities for asphalt; however, in the state of practices none of them truly demonstrates satisfactory in situ anti-oxidative performance [12]. As such, the identication of an effective
antioxidant of asphalt binder is of great benet to the asphalt-related industries. More information in the regard of asphalt aging
can be found in the literature per the references of [4,6,8,1012].
Of the various polymers as antioxidants in asphalt, wood lignin
has also been researched and once shown promising performance.
At a seven-percentage usage by weight of modied asphalt, wood
lignin signicantly slowed down the oxidation of asphalt [13]. Lignin is one of the most abundant organic polymers on Earth, constituting 30% of non-fossil organic carbon and from a quarter to a
third of the dry mass of wood, and exceeded only by cellulose
[14]. As a biopolymer, lignin is a cross-linked three-dimensional
hydrophobic and aromatic molecule. The degree of polymerization
in nature is difcult to measure, since it is fragmented during
extraction and the molecule consists of various types of substructures that appear to repeat in a haphazard manner. Although different forms of lignin have been described depending on the
means of isolation, there are three monolignol monomers in the
form of a benzene ring with a tail of three carbons, methoxylated
to various degrees: p-coumaryl alcohol, coniferyl alcohol, and sinapyl alcohol [14]. Coniferyl alcohol is the main monomer unit that
makes up the lignin in softwood. These lignols are incorporated
into lignin in the form of the phenylpropanoids p-hydroxyphenyl,
guaiacyl, and syringal, respectively [14].
The capability of lignins as a type of asphalt antioxidant is believed to arise from the scavenging action of their phenolic structures on oxygen containing free radicals [15]. The phenolic

2. Quantum chemistry perspective of asphalt oxidation and


anti-oxidation
A fundamental understanding of asphalt oxidation and anti-oxidation by chemical additives is highly desired when dealing with
the problems related to asphalt aging, this task however cannot
be readily accomplished within the frame of traditional organic
or polymer chemistry that has centered on experimenting bulk asphalt using different devices. The past few decades has seen significant advances made in using spectroscopic techniques for
characterizing molecular and atomic phenomena that involve
essentially the transfer of electrons [1924]. Since the oxidation
of asphalt is mainly a chemical process, an approach capable of
simulating electron-transfer processes is desired. In this study,
the quantum chemistry (QC) is resorted to as such an approach.
In recent years, the efforts made in atomic-level description of
phenomena have been accelerated by the availability of powerful

T. Pan / Fuel 113 (2013) 454466

3. Reactive force eld for asphalt oxidation and anti-oxidation

ET

Esystem Ebond Ev al Etors EHbond Ev dWaals ECoulomb

In Eq. (1), the bond energy Ebond describes the chemical energy
between each pair of bonded atoms. Valence angle energy Eval accounts for the energy contribution from valence angle; torsion
rotation energy Etors ensures proper dependence of the energy of
torsion angle for bond order approaching trivial and bond order
greater than 1; van der Waals interactions energy EvdWaals accounts
for the van der Waals interactions; and Coulomb interactions energy ECouloms between all atom pairs adjust for orbital overlap between atoms at close distances.
The fundamental ReaxFF assumption is that the bond order BO0ij
between a pair of atoms is dependent on the interatomic distance
rij according to Eq. (2), in which the parameter r0 is the bond radius
and the series of parameters ps describe the bond order. In calculating bond orders, ReaxFF distinguishes between contributions
from r bonds, p bonds, and pp bonds. The bond orders BO0ij are updated in each time step. The energy of the system is nally determined by summing up all the energy contributions per Eq. (1). The
three pairs of parameters: pbo,1 and pbo,2, pbo,3 and pbo,4, and pbo,5
and pbo,6 in Eq. (2) correspond to the orders of the r bond, the rst
p bond, and the second pp bond, respectively, of which the values
are given in Table 1. The values of the exponential terms is unity
below a particular interatomic distance r0 and negligible at a longer distance. The bond energy is calculated from the bond order BO0ij .

The connectivity related terms in Eq. (1) such as the bond energy, valence angle and torsion angle energy terms are also bond
order dependent and will disappear upon bond dissociation. This
feature of ReaxFF ensures a smooth transition of the energy and
force from a bonded system to a non-bonded system. In addition
to the valence interactions which depend on overlap, there are
repulsive interactions at short interatomic distances due to Pauli
principle orthogonalization and attraction energies at long distances due to dispersion. These interactions, comprised of van
der Waals and Coulomb forces, are included for all atom pairs, thus
avoiding awkward alterations in the energy description during
bond dissociation. In this respect, ReaxFF is similar in spirit to
the central valence force elds used earlier in vibrational spectoscropy. The following sections introduced these energy contribution
terms.
3.1. Bond energy

Ebond of the CAHAOANAS system was determined according to


Eq. (3). De and pbe,1 and pbe,2 are bond parameters. Upon the dissociation of a bond, the bond order BO0ij approaches zero making the
bond energy term Ebond disappear (Eq. (3)). To simulate oxidation of
asphalt, the energy contributions developed by van Duin et al.
[22,26,27] are used to determine the overall system energyhence
the ReaxFF potential functionfor the atomic-level modeling of asphalt oxidation, with the parameters given in Table 1.

AC

The Reactive Force Field (ReaxFF) developed by van Duin et al.


[22] is one such rst-principle based force eld method, which
since advent has been parameterized and implemented to a variety
of materials and processes, including high-energy materials,
hydrocarbon reactions, and transition-metal-catalyzed nanotube
formation. Recently ReaxFF has been extended to more materials
including various polymers, metals, ceramics, and silicon, and is
now used as a general tool for chemical simulations. The reactive
force eld of a CAHAOANAS system (carbon, hydrogen, oxygen,
nitrogen, and sulfur) originally developed by van Duin et al.
[22,26,27] is re-evaluated in this study to simulate the oxidation
of asphalt in exposure to oxygen.
The overall energy of the CAHAOANAS ReaxFF contains a series of energy contributions per Eq. (1), all determined using QC. The
actual number of energy contributions of a ReaxFF system depends
on the type of chemical species and processes to be modeled. The
ReaxFF for simulating asphalt oxidation in an oxygen environment
involves complex chemical reactions between the ve element
species of asphalt and external oxygen.

BO0ij BOrij BOpij BOpp


ij

 pbo2 

 pbo4 
r ij
rij
exp pbo1  r
exp pbo3  p
ro
ro

 pbo6 
r ij
exp pbo5  pp
ro

computers and parallel computing methods, advances in the statistical mechanics and new experimental data. The QC-based methods today are readily implemented using different computerbased programs, of which the dominant one is the Density Functional Theory (DFT) [25].
Extensive use of QC computation today however is limited to
small atomic/molecular systems and therefore is not practical for
studying asphalt oxidation even though recent computing capacity
has signicantly improved the speed of QC computation that enables predicting accurately the geometries and vibrational energies. As such, it is necessary to have an accurate force-eld based
method that enables quick evaluation of inter-atomic bonding
and forces. Such a rst principle based force-eld method has the
advantage of being able to simulate chemical reactions while
obtaining fast computation speed as traditional force eld methods
do, and therefore is highly desired for simulating asphalt oxidation
that involves thousands of atoms for a realistic simulation.

TE

456

Ebond De  BO0ij  exppbe1 1  BO0ij

pbe2

3.2. Valence angle energy


Eval was determined according to Eq. (4). Just like bond energy,
it is important that the energy contribution from valence angle
term goes to zero as the bond orders in the valence angle goes to
zero. Eq. (4) determines the energy associated with deviations in
valence angle Hijk from its equilibrium value H0. The term f1(BO)
per Eq. (4a) ensures that the valence angle energy contribution disappears smoothly during bond dissociation. Eq. (4b) deals with the
effects of over/under coordination in central atom j on the valence
angle energy. Atom undercoordination/overcoordination parameter Dj is dened for atoms as the difference between the total bond
order around the atom and the number of its bonding electrons
valence. The equilibrium angle H0 for Hijk depends on the sum
of p-bond orders around the central atom j and changes from
around 109.47 for sp3 hybridization (p-bond = 0) to 120 for sp2
(p-bond = 1) to 180 for sp (p-bond = 2) based on the geometry
of the central atom j and its neighbors. Table 2 lists the parametric
values used to determine the valence angle energy Eval.

Ev al f1 BOij  f1 BOjk  f2 Dj  fpv al1  pv al1


 exppv al2 Ho BO  Hijk 2 g
p

f1 BOij 1  expBOijv al3


f2 Dj

2 expDj
1 expDj exppv al4  Dj

4a
4b
4c

457

T. Pan / Fuel 113 (2013) 454466


Table 1
ReaxFF parameters for determining bond order and Ebond of asphalt as a CAHAOANAS system.
De (kcal/mol)

pbe,1

pbe,2

pbo,1

pbo,2

pbo,3

pbo,4

pbo,5

pbo,6

CAC
CAH
HAH
CAO
OAO
CAN
OAN
NAN
HAO
HAN
CAS
HAS
OAS
NAS
SAS

145.4070
167.1752
188.1606
171.0470
90.2465
134.9992
127.7074
151.9142
216.6018
223.1853
128.9942
151.5159
100.0000
0.0000
96.1871

0.2176
0.4421
0.314
0.36
0.995
0.042
0.4561
0.428
0.4201
0.4661
0.1035
0.4721
0.5563
0.4438
0.0955

0.1940
1.0000
1.0000
0.2660
0.1850
0.3161
0.3555
0.1614
1.0000
1.0000
0.2398
1.0000
0.4577
0.3153
0.2373

5.9724
8.5445
5.7082
5.0637
6.2396
5.4980
7.0000
5.3056
5.9451
6.1506
5.6731
7.0050
7.1145
5.6864
6.4757

1.0000
0.0000
0.0000
0.0000
1.0000
1.0000
1.0000
1.0000
0.0000
0.0000
1.0000
1.0000
1.0000
1.0000
1.0000

8.6733
0.0000
0.0000
7.4396
7.5281
7.0000
7.0000
12.1345
0.0000
0.0000
8.1175
0.0000
12.7569
9.1227
9.7875

1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000

0.7816
0.0000
0.0000
0.1696
0.2435
0.1370
0.1481
0.1001
0.0000
0.0000
0.5211
0.0000
0.4038
0.2034
0.4781

0.3217
0.5969
0.6816
0.3796
0.9704
0.2415
0.2000
0.6229
0.9143
0.5178
0.6000
0.6000
0.6000
0.6000
0.6000

H0 (degree)

pval,1 (kcal/mol)

CACAC
CACAH
HACAH
CAHAH
CAHAC
HAHAH
CACAO
OACAO
CACAN
OACAN
NACAN
HACAO
HACAN
CAHAN
CAOAC
CAOAO
CAOAN
OAOAO
OAOAN
NAOAN
CAOAH
HAOAO
HAOAN
HAOAH
CANAC
CANAO
CANAN
OANAO
OANAN
NANAN
CANAH
HANAO
HANAN
HANAH
CAHAO
CAHAN
CAHAS
OAHAO
OAHAN
NAHAN
HAHAO
HAHAN
CACAS
CASAC
HACAS
CASAH
CASAS
HASAH
HASAS
HAHAS

70.0265
69.7786
74.6020
0.0000
0.0000
0.0000
72.9588
80.0708
61.5055
71.9345
51.3604
66.6150
68.9632
0.0000
79.1091
83.7151
79.5876
80.0108
81.5614
85.3564
78.1533
84.1057
79.4629
79.2954
66.1477
91.9273
92.6933
73.4749
73.9183
74.0572
72.7016
82.4368
82.6883
71.2183
0.0000
0.0000
0.0000
0.0000
0.0000
0.0000
0.0000
0.0000
74.9397
86.9521
74.9397
86.1791
85.3644
93.1959
84.3331
0.0000

13.6338
10.3544
11.8629
0.0000
3.4110
27.9213
16.7105
45.0000
45.0000
45.0000
45.0000
13.6403
16.3575
0.0019
45.0000
42.6867
45.0000
38.3716
19.8012
36.5858
44.7226
9.6413
44.0409
26.3838
22.9891
38.0207
9.9708
42.7640
44.8857
15.4709
33.4153
44.1900
39.9831
14.4528
0.0019
0.0019
0.0019
0.0019
0.0019
0.0019
0.0019
0.0019
25.0560
36.9951
25.0560
36.9951
36.9951
36.9951
36.9951
0.0019

ET

pval,2 (1/radian2)

pval,3

pval,4

2.1884
8.4326
2.9294
6.0000
7.7350
5.8635
3.5244
2.1487
1.2242
1.5052
0.6846
3.8212
3.1449
6.3000
0.7067
0.9699
1.1761
1.1572
3.9968
1.7504
1.3136
7.5000
2.2959
2.2044
1.5923
0.5387
1.6094
1.7325
1.1980
5.4220
1.0224
1.9273
1.1916
3.6870
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
1.8787
2.0903
1.8787
2.0903
2.0903
2.0903
2.0903
6.0000

0.1676
0.1153
0.1367
0.0000
0.0000
0.0000
1.1127
1.1127
1.1127
1.1127
1.1127
0.0755
0.0755
0.0000
0.6142
0.6142
0.6142
0.6142
0.6142
0.6142
0.1218
0.1218
0.1218
0.1218
1.6777
1.6777
1.6777
1.6777
1.6777
1.6777
0.0222
0.0222
0.0222
0.0222
0.0000
0.0000
0.0000
0.0000
0.0000
0.0000
0.0000
0.0000
0.0559
0.0559
0.0000
0.0000
0.0559
0.0000
0.0000
0.0000

1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.1880
1.1880
1.1880
1.1880
1.1880
1.0500
1.0500
1.0400
1.0783
1.0783
1.0783
1.0783
1.0783
1.0783
1.0500
1.0500
1.0500
1.0500
1.0500
1.0500
1.0500
1.0500
1.0500
1.0500
1.0500
1.0500
1.0500
1.0500
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400
1.0400

AC

Angle atoms

TE

Table 2
Parameters for determining valence angle energy in the developed CAHAOANAS system.

Bond

458

T. Pan / Fuel 113 (2013) 454466

3.3. Torsion rotation energy

Table 4
Parameters for hydrogen bond interactions energy.

f3 BOij ; BOjk; ; BOkl 1  expBOij   1  expBOjk 

f4 Dj ; Dk

 1  expBOkl 

5a

2 expDj  Dk
1 expDj  Dk expDj Dk

5b

phb2

phb3

OAHAO
OAHAN
NAHAO
NAHAN
OAHAS
NAHAS
SAHAO
SAHAN
SAHAS

2.0431
1.674
1.4889
1.8324
2.6644
4.0476
2.1126
2.2066
1.9461

6.6813
10.9581
9.6465
8.0074
3.9547
5.7038
4.5790
5.7038
4.0000

3.5000
3.5000
3.5000
3.5000
3.5000
3.5000
3.5000
3.5000
3.5000

1.7295
1.7295
1.7295
1.7295
1.7295
1.7295
1.7295
1.7295
1.7295

Table 5
Parameters for determining van der Waals interaction energy.
Atom units

Dij (kcal/mol)

rvdW ()

cw

aij

PvdW1

C
H
O
N
S

0.1818
0.0600
0.088
0.1376
0.2099

1.8857
1.603
1.9741
1.9324
2.0677

2.0784
4.4187
7.7719
7.8431
4.9055

9.5928
9.3951
10.219
10.067
9.9575

1.5591
1.5591
1.5591
1.5591
1.5591

EHbond phb1  1  expphb2  BOXH 



 o



r
r HZ
8 HXHZ
 exp phb3 hb o  2  sin
r HZ r hb
2








f5 r ij
1
f5 r ij
Ev dWaals Dij  exp aij  1 
 2  exp  aij  1 
r v dW
2
r v dW

AC

Eq. (6) described the bond-order dependent hydrogen bond for


an XAHZ system as incorporated in ReaxFF. Table 4 lists the values of parameters rhb, phb1, phb2, and phb3 used to determine the torsion rotation energy EHbond.

ET

EvdWaals accounts for the van der Waals interactions using a distance-corrected Morse-potential given in Eq. (7). By including a
shielded interaction (Eq. (7a)), excessively high repulsions between bonded atoms (12 interactions) and atoms sharing a valence angle (13 interactions) are avoided. Dij is the basic energy
term of an atomic pair; rvdW is the van der Waal radius. The param-

Table 3
Parameters for determining torsion rotation energy (X represents any of elements C,
H, O, N, and S).

phb1

eter aij is set to be 1 for the elements of this study. The other
parameters for EvdWaals are given in Table 5.

3.4. Hydrogen bond interactions

3.5. Van der Waals interactions energy

rhb

Etors f3 BOij ; BOjk ; BOkl  sin Hijk



1
2
 sin Hjkl V 1  expfptor  BOjk  1 f4 Di ; Dk g
2



1
 1  2 cos xijkl V 2 1 cos 3xijkl
2

Hydrogen bond

TE

Etors ensures that dependence of the energy of torsion angle xijk


accounts properly as bond order approaches trivial and bond order
greater than 1. This is given in Eqs. (5), (5a), and (5b). Hijk and Hjkl
are valence angles. D is the atom undercoordination/overcoordination parameter. Parameters ptor3 and ptor4 in Eq. (5b) equal zero for
the specic types of elements in this study. Table 3 lists the values
of parameters V1, V2 and ptor used to determine the torsion rotation
energy Etors.

Torsion angle

V2 (kcal/mol)

V3 (kcal/mol)

ptor

CACACAC
CACACAH
HACACAH
XACAHAX
XAHAHAX
XACAOAX
XAHAOAX
XAOAOAX
XACANAX
XAHANAX
XAOANAX
XANANAX
XACACAX
NACANAN
XACASAX
XASASAX
XAHASAX

23.2168
45.7984
44.6445
0.0000
0.0000
16.7344
0.1000
68.9706
66.2036
0.1000
14.8049
37.4200
0.9305
43.6430
30.3435
42.7738
0.0000

0.1811
0.3590
0.3486
0.0000
0.0000
0.5590
0.0200
0.8253
0.3855
0.0200
0.0231
0.0107
0.0000
0.0004
0.0365
0.1515
0.0000

4.6220
5.7106
5.1725
0.0000
0.0000
3.0181
2.5415
28.4693
4.4414
2.5415
10.7175
3.5209
24.2568
11.5507
2.7171
2.2056
0.0000


 pv dW1 p 1
v dW1
1
p
f5 r ij rijv dW1

7a

cw

3.6. Coulomb interactions energy ECouloms

ECouloms is considered between all atom pairs. To adjust for orbital overlap between atoms at close distances a shielded Coulomb
potential ECouloms is used (see Eq. (8)).

ECoulcomb C 

qi  qj
r 3ij

1=cij 3 

1=3

Atomic charges are calculated using the electron equilibration


method [28,29]. The initial values for the electron equilibration
method parameters (g, v, and c) listed in Table 6 were determined
by Njo et al. [30] The optimization parameter cij in Eq. (8) is included for orbital overlap correction in the electron equilibration
method.
4. Atomistic modeling of asphalt oxidation and anti-oxidation
by lignin
Petroleum asphalt is a molecular system of different chemical
elements: carbon (85%), hydrogen (11%), sulfur (15%), nitrogen

Table 6
Parameters for determining Coulomb interactions energy.
Atom type

g (eV)

v (eV)

c ()

C
H
O
N
S

6.9235
9.8832
7.8386
6.3404
8.2545

5.7254
3.8196
8.500
6.8418
6.500

0.8712
0.7625
1.0804
0.8596
1.0336

459

T. Pan / Fuel 113 (2013) 454466

TE

warm-mix asphalts. Notably, with only one asphalt species studied, this atomistic work is a deterministic instead of a stochastic
process. Notably, lignin might get oxidized at raised temperature,
the temperature of 150 C, which is typical in producing asphalt
products such as asphalt pavements, was selected in this study
to test the oxidation behavior of lignins at raised temperature.
The ball-stick schemes of the polyaromatic portion, chain alkane
portion, and the average model are shown in the second column
of Fig. 1a. It is noteworthy that as the average asphalt molecule
was separated into the polyaromatic portion and saturate portion,
hydrogen atoms were added to each portion to make the included
molecules chemically balanced. The ball-stick schemes of the two
lignin molecules are shown in the rst column of Fig. 1b.
Thermodynamics and kinetics are two distinct aspects of chemical reactions. While thermodynamics describes the possibility and
direction of a reaction in terms of the free energy, DG, released or
consumed during a chemical reaction; kinetics concerns how fast
the chemical reaction can reach equilibrium, i.e., the rate of the
reaction, as inuenced by factors of reaction condition such as
the temperature and concentrations of reactants. The thermodynamics aspect of asphalt oxidation and lignin anti-oxidation is
evaluated in this study by observing respectively the oxidation
behavior of the lignin molecules, and the average molecule and
its polyaromatic/saturate portions in a domain of oxygen molecules, all under the same simulation condition. The kinetics of asphalt oxidation and lignin anti-oxidation depends on factors such
as asphalt composition, products formed in oxidation, reaction
temperature, oxygen partial pressure, and other physicochemical
effects of the system. Accordingly, the kinetics of the different
molecular species was evaluated by studying the bulk asphalt oxidized under the same condition as used in studying the thermodynamics of the average molecule and its components, i.e., at one
standard atmospheric pressure (1 ATM) pressure (oxygen only)
and 130 C. A total of twenty such average asphalt molecules are
compacted to the density of typical asphalt materials under the
ambient condition, i.e., 0.98  103 kg/m3.
For the kinetics of asphalt oxidation, the compacted asphalt was
then used to build a chemophysical environment, in form of an
atomistic model shown in Fig. 2a, to study the oxidation behavior
of bulk asphalt under the condition of temperature equal to 130 C
and oxygen partial pressure of 1 ATM. In the average asphalt model, its component models and the bulk chemophysical model, in
addition to the asphalt components oxygen molecules are included
that are separated at an average distance of 1.2 nm (same as the
inter-molecule distance in air at 1 ATM). Such a simulation
condition was determined based on the experience of numerous
asphalt-oxidation studies conducted before. The chemical
functionality developed in asphalt when oxidized at 130 C was
once found to be similar to that developed during normal pavement aging at ambient temperatures [34]. Also, it was shown that
oxidation of asphalts in exposure to high oxygen particle pressure
(e.g., 100% oxygen) is equivalent to that in air of 300 psi pressure
(such as in a Pressure Aging Vessel (PAV) condition), which is believed to be the same oxidation level as typically found in asphalts
after ve or more years of pavement service [35]. The use of 1 ATM
pressure for modeling is important for studying the escape of existing small-molecule-weight components of asphalt or newly generated ones during asphalt oxidation from bulk asphalt, which might
be an important mechanism contributing to asphalt hardening.
As such, each ReaxFF model contains an individual or a number
of the selected lignin molecules, average asphalt molecule(s) or its
components, forming a domain embedded in oxygen molecules
(see Fig. 1a and b for the single-molecule and molecular component models, and Fig. 2a, c and e for the bulk asphalt model). Each
ReaxFF model was run at a Velocity Verlet plus Berendsen ensemble [36,37] at a time step of 0.25 femtoseconds (fs) using a parallel

ET

AC

(0.31.1%), oxygen (0.20.8%), and some trace species such as


vanadium (41400 ppm) and nickel (0.4110 ppm). Bulk asphalt
consists of molecules with different structures formed by atoms
of these chemical species. Like in general organic substances, structures of the asphalt molecules are more important than the
amount of each element present. Some heteroatoms, i.e., sulfur,
nitrogen, oxygen, are attached to carbon atoms in different congurations, forming different polar molecules or functional groups
due to imbalance of electrochemical forces which are weaker than
the primary chemical bondthe covalent bond that holds atoms
within asphalt molecules. Molecular interactions among the polar
molecules of asphalt are the primary mechanism of agglomerations that strongly inuence the physical properties and macroscopic performance of asphalt. Usually a small amount of polar
molecules can have a great effect on most engineering behavior
of asphalt, such as the resistance to binder-aggregate stripping that
in general is believed to be controlled by the adsorption of polar
molecules to the surface of mineral aggregates [7,10].
A proper balance among the different functional groups hence is
important for producing durable and resistant asphalts against
detrimental physical property changes during oxidative aging
[31]. A classic method for separating asphalt into different functional groups is per their solubilities in a series of organic solvents,
supplemented with chromatographic analyzes (namely, the Corbett method) [5,10]. To study the oxidation of asphalts in their service conditions, a typical petroleum asphalt can be divided in three
representative functional groups: asphaltenes (relatively large
molecules insoluble in straight-chain alkanes such as n-heptane
or n-pentane), resins (naphthene aromaticsalkane-soluble and
elute in aromatic solvents such as benzene, and polar aromatics
alkane-soluble and elute in more polar solvents, such as an aromatic/alcohol mixture), and oils or saturates (molecules that elute
immediately in n-heptane) [5,10]. The resin and oil portions together also are known as maltenes or petrolenes. Asphaltenes, with
higher molecular weight and polarity than resins and saturates, if
not properly dispersed by the resinous components of maltenes
can cause reduced asphalt compatibility and asphalt durability
thereby [5,10].
In this study, a model asphalt system representing the three
functional groups of general petroleum asphalt was developed
according to the composition determined by Carbognani [32]. Such
an average asphalt model consists of a polyaromatic portion as
included in the asphaltene and naphthene groups of asphalt, and
some straight or branched chain alkanes like those in saturate
group of asphalt. The overall mixture composition, i.e., the relative
amounts of the three ingredient functional groups used for making
the average asphalt model was in agreement to the NMR-based
measurements by Storm et al. [33]. The lignin used for this study
was a commercially available one, i.e., the coniferyl alcohol as
the main monomer unit of lignins in softwood. The coniferyl-alcohol
lignin contains large amounts of phenol structures, giving it
antioxidant ability. Two specic types of coniferyl-alcohol lignin
molecules were mixed with asphalt, with two lignin-1 molecules
and three lignin-2 molecules, reaching a 22 weight percentage in
modied asphalt.
The lignin materials was mixed with the modeled average asphalt binder at 130 C, and then subject to two different levels of
temperature, 130 C and 150 C, to evaluate its anti-oxidation effect on the asphalt binder. The rational for selecting 130 C and
150 C conditioning for 2 h are to test the survival ability of the lignin, based on existing literature indicating that most lignin will
lose activities at above 150 C. The 2 h is the typical time for simulating asphalt aging. The coniferyl-alcohol lignin studied in this
work showed different survival abilities at 130 C vs. 150 C, which
indicates that fast mixing and compacting procedure is necessary
for hot mixes and that the lignin studied is more suitable for

T. Pan / Fuel 113 (2013) 454466

ET

AC

TE

460

Fig. 1. Thermodynamics study of lignin molecules, and average asphalt molecule and its components, (a) average asphalt molecule and its components and (b) lignin
molecules.

computing algorithm developed based on the Message-Passing


Interface (MPI) Standard. The Virginia Techs suite of high performance computersSystem X was used for the computation. System X is a supercomputer comprising 1100 Apple PowerMac G5
cluster nodes and capable of running at 12.25 Teraops (meaning
1012 FLoating point OPerations per Second). Each ReaxFF model
was kept running until the equilibrium of oxidation is reached,
i.e., no more new species observed for 1 h. The modied and

unmodied bulk asphalt models after oxidation simulation is


shown in Fig. 2b, d and f.

5. Validation of the chemophysical environment


The technique X-ray photoelectron spectroscopy (XPS), a powerful surface method for accurately detecting the presence and

461

T. Pan / Fuel 113 (2013) 454466

(b)

(d)

ET

(c)

AC

TE

(a)

(e)

(f)

Fig. 2. Oxidation of unmodied and lignin-modied asphalts at 130 C and 150 C, (a) bulk asphalt without lignin before oxidation, (b) after oxidation at 130 C for 2 h, (c)
bulk asphalt with lignin before oxidation, (d) after oxidation at 130 C for 1 h, (e) bulk asphalt with lignin before oxidation and (f) after oxidation at 150 C for 2 h.

relative quantities of chemical elements (except hydrogen and helium), was used in this study to validate the modied and unmodied bulk asphalt models by determining and comparing the
amounts of generated functional groups before and after asphalt
oxidation. The grade of the asphalt studied was PG 64-16, which
was determined using the SuperPave method. It is noteworthy that
the focus of the study is the chemical oxidation (aging) behavior.

The rheological properties of the binder are close to that of PG


64-16. Three replicates were prepared and tested in the XPS studies. XPS analysis can obtain information of the state and environment of atoms in the sample surface (usually 110 nm in depth
and 10 lm in width depending on the input energy), which can
further be analyzed for information of the surface structure of
the surface. Such powerful ability of XPS is particularly useful for

T. Pan / Fuel 113 (2013) 454466

K E ht  Eb  u

ET

AC

A Physical Electronics 5600 XPS setup was used in this study,


which provides the relative frequencies of binding energies of electrons measured in 0.1 electron-volts (0.1 eV). The binding energies
were then used to identify the elements to which each peak corresponded, as elements of the same kind in different states and environments usually have different characteristic binding energies.
Furthermore, comparing the areas under the peaks gave relative
percentages of the elements detected in a sample. The oxygen concentration adopted in the XPS analysis was controlled by oxygen
pressure. The 1 ATM oxygen pressure (oxygen only) used was higher than the atmospheric oxygen proportion (0.21 ATM), which was
intended to accelerate the oxidation reaction of asphalt. To maintain the oxygen at this relatively high concentration, the asphalt
samples were rst vacuumed in an environmental chamber, and
oxygen was then regulated into the chamber until the 1 ATM pressure was achieved. During the vacuum process at the room temperature, small oil molecules and low-molecular-weight alkanes
from asphalt are released from asphalt surface. The lost portion
was counted as the short-chain alkanes, and leftovers were tested
for ketones and sulfoxides. An initial survey XPS scan was conducted to identify the elements present in the modied and
unmodied asphalt samples (Figs. 3a and 4a), followed with a subsequent high-resolution scan of the peaks of interest, i.e., the carbon and sulfur in this study, to identify ketones and sulfoxides as
observed in the atomistic simulation (Figs. 3b and 4b). As no obvious generation of new nitrogen compounds in the atomistic model
and XPS survey scan, high-resolution scan was not conducted at
the nitrogen peak. For the high-resolution scan, the computer software associated with the Physical Electronics 5600 XPS device detects the transform of sigma carboncarbon single bonds (before
oxidation) to carbonoxygen double-bonded ketones (after oxidation), and the transform of suldes (before oxidation) to sulfoxides
(after oxidation) at the carbon and sulfur peaks, which was also
predicted by the atomistic models as shown in Figs. 1 and 2.
Furthermore, the generations of ketones and sulfoxides during
oxidation, as quantied by XPS for the same modied and unmodied asphalt materials as modeled atomistically, are shown in
Fig. 5, which further validates the ReaxFF-based bulk asphalt
model by showing close results in the amounts of newly generated
ketones and sulfoxides. The slight difference between the
experimental results and numerical predictions might come from

(a)

studying the oxidative aging behavior of asphalt binders [3840],


for which a sufciently thin sample (e.g., a few lms) can be easily
prepared and oxidized with uniform oxidation throughout the entire depth of the sample in a relatively short period of time (e.g., 2 h
at 130 C) [41].
XPS analysis is based on the photoelectric effect stemming from
the ejection of electrons from the surface of a sample in exposure
to electromagnetic radiation of sufcient energy. Electrons emitted
have characteristic kinetic energies proportional to the energy of
the radiation, according to Eq. (9), in which KE is the kinetic energy
of the electron, h is Plancks constant, m is the frequency of the incident radiation, Eb is the ionization or binding energy, and u is the
work functiona constant dependent on the X-ray photoelectron
spectrometer used. In an XPS analysis, a level of energy radiation
is used to expel core electrons from a sample, and the kinetic energies of the resulting core electrons are measured. Using Eq. (9) with
the kinetic energy KE and known frequency m of radiation, the binding energy Eb of the ejected electron can be determined. By Koopmans theorem, which states that binding or ionization energy is
equivalent to the negative of the orbital energy, the energy of the
orbital from which the electron originated can be determined.
These orbital energies are characteristic of the element and its
state.

TE

462

(b)

Fig. 3. XPS spectra of unmodied asphalt before and after oxidation, (a) survey scan
and (b) high-energy scan of carbon and sulfur zones.

the machine or operation errors related to XPS. In preparing samples for XPS analysis, the modied/unmodied asphalt samples
were vacuumed at the room temperature following the procedures
established by Ruiz et al. [41]. The samples (in dry powders) were
then pressed onto a piece of thin indium foil (0.1 mm thick) as the
sample substrate. The graphite tape was not used as sample substrate for the carbon-based asphalt to avoid peaks from the graphite tape, which would otherwise add to the carbon peak and
potentially skewing or overlapping the XPS spectra.

6. Simulation results and discussions


Fig. 1a also shows the thermodynamic aspect of the oxidation of
the average asphalt and its components. Under the same oxidation
condition, different component molecules of the average asphalt
molecule, i.e., the polyaromatics and alkanes oxidize at different
rates and lead to different products. The carbon atoms in saturate
alkanes are quite resistant to oxidation; instead some chains show
the sign of breaking into shorter chains. This could constitute an
important mechanism for asphalt hardening as liquid portions of
asphalt vaporize into air relatively easily under general service
conditions. However, no sulfoxidation or ketonization was observed in the saturate alkanes. The carbon atoms in the polyaromatic part of the average asphalt molecule show a partial level of
oxidation, i.e., oxidized at the benzylic position, not the carbon
atoms forming benzene rings, resulting in a ketone. For the polyaromatic, sulfoxidation occurs to the sulfur atom forming a sulfoxide
that consists of a double bond with an oxygen atom (S@O), however no ketonization was observed. Different from the two component models, both sulfoxidation and ketonization were observed in
the average asphalt model and the ketonization occurs to some
benzylic carbon atoms. In the average asphalt model, sulfoxidation

463

T. Pan / Fuel 113 (2013) 454466

AC

TE

(a)

(b)

Fig. 4. XPS spectra of lignin-modied asphalt before and after complete oxidation (10 h), (a) survey scan and (b) high-energy scan of the carbon zone.

ET

seems to be easier than ketonization as the sulfoxide appears earlier than the ketone.
The snapshot of the oxidation of unmodied asphalt at 2 h,
shown in Fig. 2b, displays two obvious phenomena: (1) generation
of light-molecular-weight saturate molecules (with shorter chains
than the two original saturates), which tend to leave the bulk asphalt via diffusion; and (2) agglomeration of oxidized saturates
and aromatics as attracted by the oxygen-bearing functional
groups. Some unbroken long saturates entangled with such
agglomeration due to larger molecular weight and/or electrostatic
forces. These two observations are in agreement to the phenomena
of chain breaking of saturates, sulfoxidation, and ketonization observed in the preceding thermodynamics studies shown in
Fig. 1a. The two mechanisms probably contribute signicantly to
the oxidative hardening of asphalt. The generated new species
was summarized in Fig. 5a. In general, the generations of new species all tend to slow down and approach an upper bound number
of production, as more reactants were consumed. The generation
rate of sulfoxides seems to exceed that of ketones in the early
stages, and gradually lags behind. The generation of shorter chains
was not very dramatic; however the generation rate of ketones
seemed to dominate the overall reaction at a rate higher than both
sulfoxide generation and saturate-chain breaking. Ketones therefore may contribute more to the long-term agglomeration of aged
asphalt.
Ketones and sulfoxides are the major oxidation products
formed in oxidative aging of the modeled average asphalt, which
is in agreement to results of existing laboratory experiments such
as by Functional Group Analyzes [5,10]. Moreover, such oxidation
products formed are consistently observed among eld asphalts

from different sources [34,35], and are in good agreement with


the general oxidation chemistry of hydrocarbon and sulfur-containing molecules [3840]. Ketones are a group of organic compounds carrying a carbonyl group, C@O, in which the carbon
atoms is bonded to two adjacent carbon atoms. The amine groups
(nitrogen compounds) however do not show contribution to asphalt hardening. Ketones are weak acids due to the a-proton adjacent to the carbonyl group that are much more acidic compared to
simple hydrocarbons, and can be removed by common bases such
as HO and RO. Sulfoxides are another group of chemical compounds, containing a sulnyl functional group, S@O, attached to
two carbon atoms. The generation of ketones and sulfoxides are
thermodynamically and kinetically favorable in general asphalt
under the typical open-to-air conditions. Although the bond between sulfur and oxygen atoms differs from conventional double
bonds like that between carbon and oxygen in ketones, sulfoxides
and ketones are both polar, containing an electronegative oxygen
atom producing a dipole in interaction or association with other dipoles or induced dipoles, thus both contributing to agglomeration
of oxidized asphalt. The XPS spectra clearly show these two species, which was in good agreement to the ReaxFF simulation results. Also, these two species have demonstrated effects of
increasing viscosity and molecular aggregation in aged asphalt as
reported in numerous research studies conducted to link the physicochemical properties of the ketones and sulfoxides to viscosity
and distresses of asphalt and pavements. More information in this
regards can be found in the series of SHRP I reports.
Fig. 1b shows the thermodynamic aspect of the oxidation of the
two lignin molecules. At 130 C lignin molecules did show signs of
oxidation, as can be seen from the second column of Fig. 1b.

T. Pan / Fuel 113 (2013) 454466

464

AC

TE

(a)

(b)

ET

Fig. 5. Experimental vs. numerical results of oxidized unmodied/modied bulk asphalt (rst 2 h), (a) generation of new functional groups at 130 C and (b) generation of
new functional groups at 150 C.

However, lignins do get oxidized when temperature is raised to


150 C at the same oxygen partial pressure, generating vanillin
and glycolaldehyde for the two coniferyl-alcohol lignin molecules.
More meaningfully, the addition of coniferyl-alcohol lignin signicantly changes the oxidation behavior of the average asphalt modeled in this study, as presented in Figs. 2d and e, 4a and b, and 5a
and b. At 130 C the lignin does show effectiveness in anti-oxidation,
which can be seen from the levels of agglomeration in the oxidized
asphalt models without lignin (Fig. 2b) vs. with lignin (Fig. 2d).
This anti-oxidation effect is owing to the scavenging action of
lignins phenolic structures on oxygen containing free radicals,
i.e., the ketones and sulfoxides formed in asphalt oxidation [15].
Phenolic structures are benzene rings with attached hydroxyl
groups, which can form p-conjugation regions that promote the
formation of donoracceptor complex in modied asphalt. Such
complexes act as catalysts or activators for scavenging the ketones
and sulfoxides generated in asphalt oxidation [16,42]. However,
when the temperature is raised to150 C the lignin molecules
started to get oxidized and generated vanillin and glycolaldehyde,
and their scavenging action began to decrease. Father oxidation of
lignin at 150 C will lead to glycolic acid as observed in both
atomistic simulation and XPS analysis. The increased level of
agglomeration as lignin molecules started to get oxidized can be
clearly seen by comparing Fig. 2df.

The generated new species at 130 C and 150 C are summarized in Fig. 5a and b. In unmodied asphalt the generation rate
of sulfoxides seems to exceed that of ketones in the early stages,
and gradually slows down and lags behind. The generation rate
of shorter chains shows a similar trend as sulfoxides but at higher
rates; the generation rate of ketones however continues at a higher
rate than both sulfoxides and shorter-chain alkanes, which contributes more to the agglomeration of aged asphalt in the long
run. When modied with lignin, the asphalt demonstrated a significantly slower oxidation rate at 130 C than the unmodied asphalt. Lignins in general are more resistant to oxidation than
asphaltenes and resins in asphalt under the same conditions. At
150 C however, the generation rates of shorter-chain saturates
and ketones are higher than those at 130 C, however a lower rate
in sulfoxides can be seen. Therefore, the coniferyl-alcohol lignins,
when added in asphalt and not oxidized, are capable of reducing
the aging speed of asphalt. Also, added in solid powders, lignins
usually can increase asphalts stiffness.
Furthermore, lignins tend to get oxidized at temperature
P150 C and/or raised oxygen partial pressure, generating vanillin
and glycolaldehyde for coniferyl-alcohol lignin. Sustained oxidation of the new species can lead to the generation of vanillic acid
(from vanillin) and glycolic acid (from glycolaldehyde) that can
be further oxidized to oxalic acid. When oxidized to low-molecular

465

T. Pan / Fuel 113 (2013) 454466

TE

1. Two distinct stages of asphalt aging can be identied based on


the composition and structure of generated chemical species
and their generation speeds. Asphalt molecules exhibit a
chain-breaking trend and a fairly high reactivity with oxygen,
causing a rapid spurt of light-molecular-weight alkanes and
sulfoxides, and a relatively low production of ketones. This
spurt is followed by a slower rate of asphalt oxidation and
hardening.
2. Different oxidation mechanisms appear operative during these
two periods. During the initial spurt, sulfoxides are the major
oxidation product that controls viscosity increase. Following
the spurt, ketones become the major product. The ratio of
ketones to sulfoxides formed depends on sulfur content in
asphalt. The amine groups however do not show obvious contribution to asphalt hardening.
3. The oxidation of asphalt and lignin involves many stable radicals, and a number of intermediate chemicals that can only be
observed at the atomic scale due to the high chemical instabilities of such intermediate products. Signicant decomposition
of saturates and evaporation of small-molecular-weight hydrocarbon were observed in the simulation. The sulfur atoms and
benzylic carbon atoms account for most of the oxidative reactions in the modeled asphalt. Therefore for oxidative hardening
of asphalt in typical service conditions, chain breaking of saturates, sulfoxidation, and ketonization could be the major
mechanisms.
4. The coniferyl-alcohol lignin can be used as antioxidant for
petroleum asphalt, with the maximum radical-scavenging
effectiveness achieved in a non-oxidative condition of the lignin
(e.g., <130 C under the equivalent oxygen particle pressure of
1 ATM). This means that when added in asphalt the mixing
temperature must be controlled to maximize the anti-oxidation
effect of lignins.

ET

AC

acids, methyl ethers, methyl ether esters, and methyl esters that
have smaller sizes than oxidized asphaltenes and even naphthalenes, these new products help reduce the viscosity (as caused by
ketones and sulfoxides in oxidized asphaltenes and naphthalenes)
and recover the lost owability and ductility of asphalt. However
the benet owing to the smaller-size molecules is limited as can
be seen from Fig. 5b, in which signicant amounts of ketones
and sulfoxides were generated from both asphalt and lignin molecules when the temperature is raised. Therefore, the anti-oxidation
effect of lignins, if demonstrated, comes mainly from its scavenging actions at a non-oxidative temperature.
It is noteworthy that although asphalts from different sources
have different chemical compositions, they include approximately the same set of chemical elements, i.e., carbon, hydrogen,
oxygen, sulfur, and nitrogen, with some trace species. Quantumchemistry based ReaxFF is a forceeld-based method that models a molecular system composed of the chemical elements
incorporated in the force eld. Hence, the present model is capable of modeling any material systems composed of the same elements as asphalt, i.e., carbon, hydrogen, oxygen, sulfur, and trace
elements. Such material systems reasonably include organic antioxidants for asphalt and the more general organic materials.
Moreover, the overall mixture composition, i.e., the relative
amounts of the three ingredient functional groups used for making the average asphalt model was in agreement to the three
representative functional groups of asphalt: asphaltenes, resins,
and oils or saturates. Although there exist different species of asphalt, their major differences lie in the relative proportions of
these three representative functional groups. Since the oxidation
behavior of these representative functional groups each is rather
stable, the species of oxides from the XPS tests and the numerically simulated asphalt species are not supposed to changes
much in different asphalt types. The amounts of such oxide species may be different, which however does not signicantly affect the analysis and conclusion of this study. It is admitted
that this work was focused on the major representative ingredients of asphalt, i.e., asphaltenes, resins, and saturates. The method developed is applicable to different virgin asphalt species. As
to modied or recycled asphalts, with more impurity species included in the force eld, the applicability can also be reasonably
expected. Regarding the performance of asphalt pavements, a
eld aging model of asphalt can be developed based on the
atomistic model presented in this study, which can be further
incorporated into the mixture and pavement design guides. Such
a model will signicantly improve the accuracy for predicting
the service lives of asphalt binders and pavements. Admittedly,
this work presents the early results of a study on asphalt aging,
in which only an average asphalt model was studied. More complex systems such as recycled asphalt pavement will signicantly add to the difculty for model building, simulation, and
identifying functional species. Anyhow, with the major and representative ingredients of asphalt, i.e., asphaltenes, resins, and
saturates included, the aging behavior of general asphalt is captured. More desirably, more lignin species can be evaluated to
eventually form a database that can be used to direct the use
of bio-based materials as asphalt antioxidants.

7. Summary and conclusions


The phenomena of asphalt oxidation concerns petroleum and
many manufacture industries. Based on a QC-based reactive force
eld re-evaluated for the CAHAOANAS system, the thermodynamics and kinetics of asphalt oxidation and the anti-oxidation
mechanism of coniferyl-alcohol lignin are studied, aided with
XPS analyzes. Important conclusions are made as follows.

References

[1] McNichol D. Paving the way: asphalt in America. 1st ed. National Asphalt
Pavement Association; 2005.
[2] Petersen JC, Branthaver JF, Robertson RE, Harnsberger PM, Duvall JJ, Ensley EK.
Effects of physicochemical factors on asphalt oxidation kinetics. In:
Transportation Research Record 1391, TRB, National Research Council,
Washington, D.C; 1993. p. 110.
[3] Davis TC, Petersen JC. An inverse glc study of asphalts used in the zacawigmore experimental test road. Proc, Assoc Asphalt Pav Technol
1967;36:115.
[4] Plancher H, Green EL, Petersen JC. Reduction of oxidative hardening of asphalts
by treatment with hydrated lime: a mechanistic study. Proc Assoc Asphalt Pav
Technol 1976;45:124.
[5] Corbett LW. Composition of asphalt based on generic fractionation using
solvent
deasphalteneing,
elution-adsorption
chromatography
and
densiometric characterization. Anal Chem 1969;41:5769.
[6] Petersen JC. A dual sequential mechanism for the oxidation of asphalts. Petrol
Sci Technol 1998;16(910):102359.
[7] Branthaver JF, Petersen JC, Robertson RE, Duvall JJ, Kim SS, Harnsberger PM,
Mill T, Ensley EK, Barbour FA, Schabron JF. Binder characterization and
evaluation, Chemistry, vol. 2. In: Report No. SHRP-A-368, SHRP, National
Research Council, Washington, D.C; 1993.
[8] Petersen JC. Asphalt aging: a dual oxidation mechanism and its
interrelationships with asphalt composition and oxidative age hardening.
Transportation Research Record 1638, TRB, National Research Council,
Washington, D.C; 1998. p. 4755.
[9] Hveem FN, Zube E, Skog J. Progress report on the zaca-wigmore experimental
asphalt test project, vol. 277. American Society for Testing Materials; 1959. p.
145.
[10] Petersen JC. Chemical composition of asphalt as related to asphalt durability:
state of the art. Transportation Research Record 999, TRB, National Research
Council, Washington, D.C; 1984. p. 1330.
[11] Ruan Y, Davison RR, Glover CJ. Oxidation and viscosity hardening of polymermodied asphalts. Energy Fuels 2003;17:9918.
[12] Lucena MC, Soares SA, Soares JB. Characterization of thermal behavior of
polymer-modied asphalt. Mater Res 2004;7:52934.
[13] Bishara SW, Robertson RE, Mohoney D. Lignin as an antioxidant: a limited
study on asphalts frequency used on kansas roads. In: The 42nd Peterson
Asphalt research conference, Cheyenne, WY; 2005.

T. Pan / Fuel 113 (2013) 454466

[29] Janssens GOA, Baekelandt BG, Toufar H, Mortier WJ, Schoonheydt RA.
Comparison of cluster and innite crystal calculations on zeolites with the
electronegativity equalization method (EEM). J Phys Chem 1995;99:3251.
[30] Njo SL, Fan J, van de Graaf B. Extending and simplifying the electronegativity
equalization method. J Mol Catal A 1998;134:79.
[31] White RM, Mitten WR, Skog JB. Fractional components of asphalt:
compatibility and interchangeability of fractions produced from different
asphalts. Proc Assoc Asphalt Pav Technol 1970;39:498531.
[32] Carbognani L. Molecular structure of asphaltene proposed for 510c residue of
venezuelan crude. In: INTEVEP S.A. Tech. Report; 1992.
[33] Storm DA, Edwards JC, DeCanio SJ, Sheu EY. Molecular representations of
ratawi and alaska north slope asphaltenes based on liquid- and solid-state
NMR. Energy Fuels 1994;8:5616.
[34] Petersen JC, Plancher H, Miyake G. Chemical reactivity and ow properties of
asphalts modied by metal complex-induced reaction with atmospheric
oxygen. Proc Assoc Asphalt Pav Technol 1983;32:3260.
[35] Griffen RL, Simpson WC, Miles TK. Inuence of composition of paving asphalts
on viscosity, viscosity-temperature susceptibility, and durability. J Chem Eng
Data 1959;4:34954.
[36] Verlet L. Computer experiments on classical uids: I. Thermodynamical
properties of LennardJones molecules. Phys Rev 1967;159(1):98103.
[37] Berendsen HJC, Postma JPM, van Gunsteren W, DiNola A, Haak JR. Molecular
dynamics with coupling to an external bath. J Chem Phys 1984;81.
[38] Gelius U, Heden RF, Hedman J, Lindberg BJ, Manne R, Nordberg R, et al.
Molecular spectroscopy by means of ESCA III: carbon compounds. Phys Scripta
1970;2(12):7080.
[39] Kelemen SR, George GN, Gorbaty ML. Direct determination and quantication
of sulfur forms in heavy petroleum and coals: 1. the X-ray photoelectron
spectroscopy (XPS) approach. Fuel 1990;69:93944.
[40] Lindberg BJ, Hamrin K, Johansson G, Gellius U, Fahlman A, Nordllng C, et al.
Molecular spectroscopy by means of ESCA II: sulfur compounds, correlation of
electron binding energy with structure. Phys Scripta 1970;1(56):28898.
[41] Ruiz JM, Carden BM, Lena LJ, Vincent EJ, Escaller JC. Determination of sulfur in
asphalts by selective oxidation and photometric spectroscopy for chemical
analysis. Anal Chem 1982;54:68891.
[42] Kuzina SI, Demidov SV, Brezgunov AY, Poluektiv OG, Grinberg OY, Dubinski
AA, et al. Study of free-radical centers in lignin with 20 mm band ESR
spectroscopy. J Polym Sci 1993;A35(7):798802.

ET

AC

[14] Boerjan W, Ralph J, Baucher M. Lignin bios. Annu Rev Plant Biol
2003;54(1):51949.
[15] Dizhbite T, Telysheva G, Jurkjane V, Viesturs U. Characterization of the radical
scavenging activity of lignins-natural antioxidants. Bioresour Technol
2004;95:30917.
[16] Boeriu CG, Bravo D, Gosselink RJ, Van Dam JE. Characterization of structuredependent functional properties of lignin with infrared spectroscopy. Ind
Crops Prod 2004;20:20518.
[17] Brink DL, Bicho JG, Merriman MM. Oxidation degradation of wood III, lignin,
structure and reactions. Adv Chem Ser 1966;177(59).
[18] Bryan CC. Manufacture of vanillin from lignin. United States Patent 2692291;
October 1954.
[19] Pan T, van Duin ACT. Steel surface passivation at a typical ambient condition:
atomistic modeling and X-ray diffraction/reectivity analyses. Electrocatalysis
2011;2(4):30716.
[20] Pan T, van Duin ACT. Passivation of steel surface: an atomistic modeling
approach aided with X-ray analyses. Mater Lett 2011;65(2122):32236.
[21] Pan T. Quantum chemistry-based study of iron oxidation at the iron-water
interface: an X-ray analysis aided study. Chem Phys Lett 2011;511(4
6):31521.
[22] van Duin ACT, Dasgupta S, Lorant F, Goddard WA. ReaxFF: a reactive force eld
for hydrocarbons. J Phys Chem A 2001;105.
[23] Pan T, Xi Y. Physicochemical nature of iron oxidation in a damp atmospheric
condition. Acta Metall Sin 2011;24(6):41522.
[24] Pan T, Lu Y. Quantum-chemistry based studying of rebar passivation in
alkaline concrete environment. Int J Electrochem Sci 2011;2(8).
[25] Parr RG, Yang W. Density-functional theory of atoms and molecules. New
York: Oxford University Press; 1989. ISBN 0-19-504279-4.
[26] Chenoweth K, van Duin ACT, Goddard WA. ReaxFF reactive force eld for
molecular dynamics simulations of hydrocarbon oxidation. J Phys Chem A
2008;112:104053.
[27] Keten S, Chou C-C, van Duin ACT, Buehler MJ. Tunable nanomechanics of
protein disulphide bond begets weakening in reducing and stabilization in
oxidizing chemical microenvironments. J Mech Behav Biomed Mater
2012;5:3240.
[28] Mortier WJ, Ghosh SK, Shankar S. Electronegativity-equalization method for
the calculation of atomic charges in molecules. J Am Chem Soc 1986;108:
4315.

TE

466

Das könnte Ihnen auch gefallen