Sei sulla pagina 1di 14

Please Do Not Cite

Hefer, A.W., D.N. Little, and B.E. Herbert. 2005. Bitumen Surface Energy
Characterization by Inverse Gas Chromatography. ASTM. Journal of ASTM
International (JAI) or Cement Concrete and Aggregates or Geotechnical Testing Journal
(In review).
Arno W. Hefer,1 Dallas N. Little,2 and Bruce E. Herbert,3

Bitumen Surface Energy Characterization by Inverse Gas


Chromatography
ABSTRACT : Modern surface energy theory has been identified in recent years as an attractive
tool by which to select compatible bitumen-aggregate combinations. In addition, this approach
offers the potential to quantitatively assess moisture susceptibility of these material
combinations. The success of implementing this technology depends on the availability of
techniques that allow efficient and reliable surface energy characterization of the materials under
consideration. This paper focuses on bitumen surface energy characterization employing inverse
gas chromatography (IGC).
The authors report detail column preparation and test
methodologies, and provide a precision statement for this technique. Analysis procedures and
results presented demonstrate successful application of this technique with modern
thermodynamic theory to acquire bitumen surface energy components. The technique allows
testing at different temperatures and results are in agreement with conceptual and theoretical
expectations. Surface energies derived from IGC compare reasonably well with mechanical
surface tension values from the literature. Results suggest that surface energies do not vary
considerably between different bitumen types, indicative of the controlling role of aggregate
type in bitumen-aggregate adhesion.
KEYWORDS: Bitumen, adhesion, bitumen-aggregate adhesion, surface energy, surface
tension, inverse gas chromatography

Introduction
Adhesion between bitumen and aggregate is fundamentally responsible for the
integrity of asphalt mixes under dynamic wheel loads and environmental influences such
as moisture fluctuation. Researchers in the first halve of the 20th century recognized the
importance of surface tension of the materials involved, but quantitative application of
these numbers remained a challenge. In 1964, Fowkes [1] introduced concepts that
stimulate a different approach to the application of physical chemistry to adhesion
science. The theory by van Oss, Chaudhury, and Good [2] is widely applied and
1

PhD Graduate , Dept. of Civil Engineering, Texas A&M University, College Station. Currently
with AFRICON Engineering International, South Africa. E-mail: arnoh@africon.co.za
2
Professor, Dept. of Civil Engineering, Texas A&M University, College Station.
E-mail: d-little@tamu.edu
3
Professor, Dept. of Geology and Geophysics, Texas A&M University, College Station.
E-mail: Herbert@geo.tamu.edu

expresses the relationship between Gibbs free energy of adhesion (Ga), work of
adhesion (Wa), and surface energy () components in units mJ/m2, as follows:
! #G ija = Wija = 2 " iLW " LW
+ 2 " i! " +j + 2 " !j " i!
j

(1)

Where i and j represent two materials that form an adhesive bond. According to Fowkes
[1] total surface energy (equal to surface tension for liquids) is comprised of components
that represent the important forces responsible for the formation of adhesive bonds at an
interface. In the theory presented above, LW is the Lifshitz-van der Waals component, +
the acid (or electron acceptor) component, and - the base (or electron donor) component.
The Lifshitz-van der Waals component represents non-polar species (such as alkanes,
paraffins, or aliphatic compounds in bitumen) that interact through non-specific van der
Waals interactions, while the acid and base components represent polar compounds
associated with specific interactions. Hefer et al. [3] describe theories and mechanisms
responsible for bitumen-aggregate adhesion in more detail.
Equation (1) implies that any technique capable of measuring the free energy of
adhesion (Ga) can be used to resolve the surface energy components of the unknown
material under consideration. Elphingstone [4] and Cheng et al. [5] used the Wilhlemy
plate technique to calculate Ga from contact angles of different liquids in contact with
bitumen. Li [6] and Cheng et al. [5] used a static gravemetric gas sorption technique to
calculate the Ga from equilibrium spreading pressures of different probe gases in
contact with aggregate. In this paper, inverse gas chromatography (IGC) is proposed as a
candidate technique for surface energy characterization of bitumen. Simplicity,
convenience of experimentation, and control of test conditions has inspired the use of this
technique in many industries. This approach has proven to be effective in characterizing
surface energies of various materials such as polymers, fibers, minerals, and pigments
[7].
This paper introduces inverse gas chromatography and provides a detailed
methodology for column preparation and testing. An analysis procedure used to calculate
surface energy components from the measured parameters in accordance with the theory
developed by van Oss and his colleagues is presented. Finally, bitumen surface energies
are compared to mechanical surface tension data obtained from the literature.
Inverse Gas Chromatography
Analytical chromatography is a simple technique used to separate a mixture into its
constituents. When an inert gas (carrier gas) carries a gas mixture through a
chromatographic column, each compound interacts differently with a known material in
the column (stationary phase) that results in different travel times (retention times) for
different constituent compounds (mobile phase) through the column. Inverse gas
chromatography differs from this, more conventional, analytical chromatography in that
the stationary phase is the unknown material under investigation; while different probe
gases with know characteristics move through the column. The net retention time (tN) is
the measured parameter that allows one to calculate the net retention volume (VN), which
in turn, is related to the free energy of adhesion. VN is linearly related to tN through the
flow rate (F) of the carrier gas [8]:

(2)

VN = F ! t N

When infinitely diluted, or zero coverage conditions, the probe gas can be considered to
behave as an ideal gas and the free energy of adsorption (as it is normally referred to in
this context) relates to VN through the following equation, where R is the universal gas
constant and T is the temperature.
" !G a = RT ln(VN )

(3)

When a polar probe molecule interacts with the surface, specific interactions (acidbase interactions) and non-specific interactions (van der Waals forces, mainly London or
dispersive forces), take place simultaneously. It is therefore assumed that the dispersive
contributions (GD) and specific contributions (GS) are additive to the total free energy
[9].
!G a = !G aD + !G Sa

(4)

- G = RT.Ln(VN)

As illustrated in Figure 1, the free energies of adsorption of different non-polar


probes, or normal alkanes, vary linearly with their associate molecular descriptors, e.g.,
number of carbons or molecular weight. It follows that the free energy of a polar probe
with the same molecular descriptor value as a, hypothetical, n-alkane should be located
above the alkane reference line. The specific contribution of free energy of adsorption is
obtained by subtracting the dispersive contribution represented by the portion below the
alkane line [9].

S
aP

Ga

Polar
probe

Comparison scale

FIG. 1 Principle of determining the dispersion and non-specific components of Ga


[9]
Methodology
Inverse-Liquid-Gas Chromatography (IGLC) has been applied to characterize
oxidative aging of bitumen since the 1960s [10,11,12]. In these methodologies, bitumen

was coated from a solution onto Teflon particles and then packed into a column. More
recently, Barbour and co-workers [13,14] adapted the original method to a more userfriendly and time efficient approach where a fused silica capillary column (30m long with
an inner diameter of 0.53mm) was coated with bitumen from a toluene solution. The test
methodology followed in the research that supports this paper was adapted from the
capillary IGLC technique developed at the Western Research Institute under a FHWA
contract [14].
Five bitumen types with different chemical composition were tested in this study.
Core bitumens used in the Strategic Highway Research Program (SHRP) were selected
due to the vast amount of research that has been conducted with these materials. The
material reference library (MRL) codes are AAB-1, AAD-1, AAF-1, AAM-1, and ABD.
Column Preparation Method
Untreated fused silica columns (15m long with a 0.25mm inner diameter, SigmaAldrich, Suppelco), were used in these experiments. Approximately one gram of
bitumen is diluted in 10ml of toluene to produce a 10 percent solution. A rinsing kit
(Sigma-Aldrich, Suppelco) consisting of a 25ml plastic coated (pressure safe) glass
reservoir with a screw-on stainless steel inlet-outlet unit is the key apparatus in the
coating process. The solution is transferred to the 25ml vial, and one end of the column
is inserted into the solution. The solution is then pushed through the column with dry
nitrogen by applying a pressure of approximately 35 kPa. The column rinsing assembly
is illustrated in Figure 2. With about 6 to 7 coils of the transparent capillary column
filled, the inlet of the column is pulled out of the solution, and the plug of bitumen
solution is pushed through the column by the nitrogen pressure. This process is repeated
two more times to ensure adequate coating of the column walls. No significant darkening
of the column occurs due to the thin film produced. After the final plug is pushed
through, the pressure is increased to approximately 200 kPa and the water bath is turned
on. This initial drying process is allowed to continue for 60 minutes at 40C.

Waste outlet

Dry Nitrogen

Capillary column

Rinsing reservoir
With 10% bitumen solution

Water bath

FIG. 2 Column rinse assembly (Adapted from [14])

Test Method
A HP5890 Series II Gas Chromatograph (GC) equipped with Electronic Pressure
Control (EPC) and a flame ionization detector (FID) was used in this study. ChemStation
software was used for experimental set-up, control, and data acquisition. The inlet and
outlet temperatures were set to 175C and 250C, respectively. The bitumen column was
installed in the GC oven and conditioned for an hour at 130C before testing. A
computer controlled temperature program lowered the temperature to the test temperature
of 25C after conditioning. The column flow was set to 1.5 ml/min, corresponding to a
column head pressure in the order of 70 kPa, and the total flow (or injection-spit flow)
was set to 15 ml/min. Operating in split mode allows one to inject a minute amount with
a syringe by removing some of the injected sample before elution through the column,
thereby increasing peak resolution.
Injection of the probe molecules was carried out manually with gas tight glass micro
syringes (Sigma-Aldrich, Hamilton). All the chemicals were HPLC grade supplied by
Sigma-Aldrich. A mixture of normal alkanes, including normal pentane, hexane,
heptane, octane and nonane was made with methane as the reference probe.
Approximately 0.1l of the mixture was injected in liquid form. Methane is considered
an inert hydrocarbon and its retention time serves as a measure of the dead volume of the
column. While the same series of n-alkanes were used as in the original oxidation
experiments, polar probes with known surface energy components are required in these
experiments.
Two probes with acidic character, namely chloroform and
dichloromethane, and two probes with basic character, namely toluene and ethyl acetate,

were selected. One replicate consisted of a sequence of the n-alkane mixture followed by
individual injections of each of the polar probes. Five-minute of conditioning at 130C
was sufficient between sequences to remove all excess gas from the system.
Figure 3 shows a typical retention time output obtained for an n-alkane mixture,
including methane (0.731 min.) to nonane (8.156 min.). Retention time was defined as
the time corresponding to the maximum peak height.

0.912

counts
80000

0.731

40000
30000

8.156

50000

3.163

60000

1.519

0.985

70000

20000
10000
0
0

FIG. 3 Typical IGC output for an n-alkane mixture (Captured from ChemStation
Software)

Analysis
The basic relationship between retention time, tR, and net retention volume, VN, was
previously discussed. A generalized relationship is presented by Skoog and Leary [15].
VN = j / m ! F ! (t R " t M )!

T
273.15

(5)

Where, tM is the dead time, accounting for column characteristics, or essentially any
volume other than that of the sample. The parameter tM is usually determined by passing
an inert hydrocarbon, usually methane, through the column. Some researchers use air
and record an air peak. The difference tR tM is therefore the net retention time, tN. The
flow-rate of the carrier gas (F), usually helium, the test temperature (T), and sample mass
(m) must be known. The variable, j, is the dimensionless James-Martin compressibility
factor, which corrects for the pressure drop in the column. Due to the nature of the data
analysis, most of these test parameters are not directly required in the calculations as
discussed below.
A typical analysis plot compiled from data obtained for bitumen ABD is presented as
Figure 4. It is essential that a good fit be obtained for the alkane line. Decane was
abandoned because a good fit could not be established with this high molecular weight nalkane. The slope of this line is associated with the Lifshitz-van der Waals (LW)
component of surface energy (LW). This relative approach makes it possible to simplify
Equation (5) and therefore Equation (4) so that only net retention time is required as an

input, which is done by plotting RTln(tN) against the molecular descriptor. Employing
the theory adopted in this research, Equation (1), the LW component of the free energy
(GLW in mJ/m2) is expressed by the Berthelot geometric mean.
1
2

1
LW 2
S

( ) (! )

"G LW = 2 ! LW
L

(6)

Where L and S traditionally represent liquid (vapor in this case) and solid, respectively.
If expressed in kJ/mol,
1
2

1
2

( ) (! )

"G LW = aN A 2 ! LLW

LW
S

(7)

where a is the cross-sectional area of the solute, and NA is Avogardos number (6.0221 x
1

1023 species per mol). In order to facilitate a straight-line plot Equation (7), a (! LLW )2
was used as the molecular descriptor in this research. The LW component of surface
energy in mJ/m2 is then given by,

& slope #
!!
' SLW = $$
% 2N A "

(8)

Elution of monopolar basic and monopolar acid compounds enables one to determine
the acid and the basic surface energies, respectively. Equation (4) can be rewritten in
terms of the descriptors associated with the theory adopted in this research.

!G = !G LW + !G AB

(9)

As previously discussed, the difference between free energy determined from polar
interactions and the corresponding position on the alkane line defines the free energy of
specific adsorption, i.e. GAB, either acidic or basic in this case due to the use of
monopolar probes. The simplified from applicable to surface energy analysis can
therefore be expressed as:
&tp
'G AB = RT ln$$ Nn
% tN

#
!!
"

(10)

Where t Nn is the net retention time for the hypothetical n-alkane, and t Np is the net
retention time for the polar molecule under consideration.

6E+06

RT ln(tN )

3E+06

0E+00
1E-18

1.6E-18

2.2E-18

-3E+06

2.8E-18

3.4E-18

y = 8E+24x - 2E+07
R2 = 0.9998

-6E+06

a(! LW )1/2
n-alkanes

Chloroform

Toluene

DCM

Ethyl acetate

FIG. 4 Typical IGC analysis plot for bitumen: ABD


Using the approach by van Oss, Good and Chaudhury, the acid-base contribution to
free energy can be expressed in terms of the known surface energy components of the
polar probe compound and that of the surface under consideration.
&
'G AB = aN A 2 $ ) l+) s(
%

1
2

) + ()

1
( + 2
l s

) #!

(11)

"

If the acid (+ , electron acceptor) and base (- , electron donor) properties of the two
monopolar compounds are known, and either + or - equals zero, then the unknown
surface energy components can be calculated alternately. With the three surface energy
components known, the total surface energy of the bitumen under consideration can be
calculated.

" Total = " LW + 2 " + " !

(12)

Relevant characteristics of the probe molecules are presented in Table 1. The crosssectional areas ( a ) of the n-alkanes were calculated by assuming that each CH2 group
occupies 0.06 nm2, and each CH3 group occupies 0.08 nm2. This was first proposed by
Dorris and Grey, and applied by Sun and Berg [16].

TABLE 1 Characteristics of probe molecules used in IGC experiments


Probe Molecule
n-Pentane (C5H12)
n-Hexane (C6 H14)
n-Heptane (C7 H16)
n-Octane (C8 H18)
n-Nonane (C9 H20)
Toluene (C7 H8)
EA (CH3 CO2 C2 H5)
DCM (CH2Cl2)
Chloroform (CHCl3)

Surface Energy Characteristics, mJ/m2


(20C)
Total
LW
+
-

Crosssectional area
(m2)

16.1
18.4
20.1
21.6
22.8
28.5
23.9
26.5
27.2

3.399 10-19
3.999 10-19
4.598 10-19
5.198 10-19
5.799 10-19
4.200 10-19
3.293 10-19
2.986 10-19
3.505 10-19

16.1
18.4
20.1
21.6
22.8
28.5
23.9
26.5
27.2

0
0
0
0
0
0
0
5.2
3.8

0
0
0
0
0
2.3
19.2
0
0

Legend:
: Surface energy; LW: Lifshitz van der Waals; +: acid (electron acceptor); -: base (electron donor)

A liquid density model assuming a spherical molecular shape and hexagonal packing
was used to approximate the cross-sectional areas of the polar probes:
1

(13)

3
amol = 1.33 N 3! mol

Where amol is the molar area, !mol the liquid molar volume, and N is Avogadros
number.
Results and Discussion
The parameters presented in Table 2 were derived only from the measured retention
times and are essentially free from any theoretical manipulations. These parameters may
therefore be used to establish a precision statement for the inverse gas chromatography
technique to determine bitumen surface energies. Reproducibility of the technique for
the conditions of one operator and one laboratory are presented in Table 3.
TABLE 2 Parameters calculated from IGC retention times (25C)
Bitumen
MRL
Code

LW
n-Alkanes
Avg.
s
2
mJ/m

G+
Toluene
Avg.
s
kJ/mol

G+
Ethyl
Acetate
Avg.
s
kJ/mol

GChloroform
Avg.
s
kJ/mol

GDCM
Avg.
s
kJ/mol

AAB-1

48.3

0.9

1.3

0.04

1.6

0.02

0.6

0.04

0.5

0.1

AAD-1

45.4

0.6

1.5

0.03

1.9

0.12

0.9

0.03

0.5

0.1

AAF-1

47.5

0.5

1.4

0.02

1.8

0.05

0.5

0.03

0.7

0.1

AAM-1

44.9

0.4

1.1

0.01

1.3

0.01

0.2

0.03

0.1

0.1

ABD

44.8

0.2

1.3

0.01

2.0

0.03

0.7

0.02

0.4

0.04

Legend:
Avg.: average; s: standard deviation; LW : Lifshitz van der Waals surface energy components;
G+: specific free energy (acid); G- : specific free energy (basic)

TABLE 3 Pooled standard deviations (Sp) for specific free energy parameters
from IGC
LW (mJ/m2)
n-Alkanes
0.481

G+ (kJ/mol)
Toluene
0.029

Ethyl Acetate
0.074

G- (kJ/mol)
Chloroform
0.046

DCM
0.068

With the Lifshitz-van der Waals (LW) component presented in Table 2, surface
energy components derived from different polar probes and total surface energies are
presented in Tables 4 and 5, respectively. It is evident that the LW components are
dominant if compared to the acid and base components. This observation is in
accordance with previous research and can be rationalized by the fact that bitumen is
comprised primarily of non-polar organic hydrocarbons. According to Lytton [17],
predominance of LW forces lead to contraction of newly formed crack faces which
manifests in retarded fracture healing ability of the material. While the LW components
only vary between 44.8 and 48.3 mJ/m2, total surface energy values only vary between
45.1 and 50.4 mJ/m2 regardless of the combination of acidic or basic probes used in its
calculation. The polar components (acid and base) are generally small, and although they
differ statistically between different bitumen types, these differences do not appear to be
of any practical significance. Although small, the polar acid-base components ought to
contribute to long term healing if cracks are allowed to close by overcoming the effect of
van der Waals forces through other mechanisms, such as relaxation.
It is interesting that the acidic components derived from toluene and ethyl acetate are
generally larger than the base components derived from both chloroform and
dichloromethane. While this may seem to coincide with the generally acidic character of
bitumen, these values simply imply that the ratio between the surface energies of water
and bitumen (AAB-1, for example) is 25.5/2.6 for the acidic component and 25.5/0.8 for
the basic component (if ethyl acetate is selected). This is attributed to the fact that polar
surface energy components of all these probe compounds are referenced to water with an
assumed scale, or ratio between acid and base component of water, of 1:1 as originally
proposed by van Oss and his colleagues [18,19]. Other scales have been proposed but
are generally only available for liquids used in contact angle experiments. In addition,
these scales remain relative scales since the true scale is still unknown. Although this
problem is not expected to be resolved soon [20], the scale problem does not adversely
impact free energy of adhesion calculation [21].
TABLE 4 Surface energy components determined by IGC at 25C, mJ/m2
Bitumen
MRL Code

+
(Toluene)
Avg.
s

+
(Ethyl Acetate)
Avg.
s

(Chloroform)
Avg.
s

(DCM)
Avg.
s

AAB-1

2.6

0.1

0.8

0.02

0.5

0.1

0.4

0.2

AAD-1

3.1

0.1

1.1

0.15

1.2

0.1

0.4

0.1

AAF-1

2.9

0.1

1.1

0.06

0.4

0.05

0.6

0.2

AAM-1

1.7

0.01

0.6

0.01

0.1

0.02

0.01

0.01

ABD

2.3

0.03

1.3

0.04

0.8

0.03

0.3

0.1

Legend:
Avg.: average; s: standard deviation; : Surface energy; LW: Lifshitz van der Waals; +: acid (electron
acceptor); -: base (electron donor)

TABLE 5 Total surface energies (Total) determined by IGC at 25C, mJ/m2

MRL Code

Toluene/
Chloroform
Avg.
S

Avg.

AAB-1

50.4

0.9

50.2

1.2

49.5

0.9

50.4

0.9

AAD-1

49.2

0.5

47.7

0.7

47.7

0.7

48.0

0.8

AAF-1

49.6

0.5

50.2

0.8

48.8

0.5

50.0

0.4

AAM-1

45.7

0.4

45.1

0.6

45.4

0.4

46.5

0.4

ABD

47.4

0.3

46.4

0.3

46.8

0.3

47.4

0.2

Bitumen

Toluene/ DCM

Ethyl Acetate/
Chloroform
Avg.
s

Ethyl Acetate/
DCM
Avg.
s

Legend:
Avg.: average; s: standard deviation; L: liquid anti-strip agent (Akzo-Nobel C-450)

While intrinsic values for surface energy of bitumen are not known, especially for
surface energy components, total surface energies could be compared with surface
tension values measured by classic mechanical means. Elphingstone [4] used the
pendant drop method to measure surface tension of bitumen at 80C, while WRI [22]
employed the du Ny ring to obtain surface tension values at 60C. These data were
available on four of the five bitumen types selected for this study. Due to the
convenience offered by IGC to perform tests at different temperatures, surface energies
were also determined at 60C (in addition to 25oC). Total surface energies (of surface
tensions) are presented in Figure 5. The fact that surface energy decreases with
temperature can conceptually be explained based on the definition of surface energy: the
work required to increase the surface by a unit area. Heating results in mobility of the
molecules, and hence less work to increase the surface area. Theoretically, this
phenomenon implies that increased temperature (T), and thus entropy (S), will result in
a decrease of the free energy of the system (G), i.e. G = H - TS, where H is the
total energy or enthalpy of the system.
Figure 5 shows that du Ny ring and pendant drop data compare reasonably well with
IGC data obtained at 60C with the understanding that these techniques are very different.
The trends followed by the data sets also compare favorably. An important observation
is that all data sets indicate that surface energies of different bitumen types do not vary
significantly. Numerous researchers agree that aggregate type plays a dominant role in
bitumen-aggregate adhesion and the fact that bitumen surface energies do not vary

Total Surface Tension/ Energy (mN/m or mJ/m2)

considerably, corroborates this observation. Future research should focus on surface


energy characterization of oxidized bitumen and bitumen treated with liquid additives. If
this approach is sufficiently sensitive, it is expected that these effects will be pronounced
in the polar components of bitumen surface energy.
55.0

50.0

45.0

40.0

35.0

30.0

25.0
AAB-1

AAD-1

AAF-1

AAM-1

Bitumen Type

Inverse Gas Chromatography 25C

Inverse Gas Chromatography 60C

Du Ny Ring 60C (WRI, 2001)

Pendant Drop 80C (Elphingstone, 1997)

FIG. 5 Comparison of total surface energy values


Conclusions
The physical chemistry theory proposed by van Oss, Chaudhury, and Good allows
one to calculate the free energy of adhesion between two materials, provided that the
surface energy characteristics of the materials are known. The significance of such an
application is that compatibility of bitumen-aggregate combinations can be assessed and
moisture susceptibility of mixes quantitatively evaluated. While this theory has been
applied to these materials in previous research, more reliable and efficient techniques for
surface energy characterization need to be explored to implement this technology
successfully.
Inverse gas chromatography was used in this research to demonstrate its application
to bitumen surface energy characterization. Results demonstrate that the Lifshitz-van der
Waals component of surface energy is the largest contributor to the total surface energy,
while polar components are relatively small. This result coincides with previous research
and implies that apart from the importance of this component to bitumen-aggregate

adhesion, fracture healing may initially be retarded by contraction of fractured surfaces


due to predominance of van der Waals forces acting in a cohesive fashion. Surface
tension values from the literature corroborate total surface energy trends established with
IGC and compare reasonably well with magnitudes of these numbers. Results reveal a
general decrease of total surface energy with increased temperature, which is consistent
with the fundamental concept of Gibbs free energy. Both results from mechanical
measurements and IGC suggest that surface energies of different bitumen types fall
within a close practical range. The significance of this observation is that aggregate type
dictates bitumen-aggregate adhesion. Given the fact that this technique has been used to
assess oxidative aging, future research should further investigate the sensitivity of
bitumen surface energies determined from IGC for the effects of oxidative aging and
liquid additives.
Acknowledgements
The authors acknowledge the support of the Aggregates Foundation for Technology,
Research, and Education (AFTRE) and the National Cooperative Highway Research
Program (NCHRP), project 9-37 for funding support.
References
[1]
[2]
[3]

[4]
[5]
[6]
[7]

[8]

Fowkes, F.M., Attractive forces at interfaces, Industrial Engineering &


Chemistry, Vol. 56, No. 12, 1964, pp. 40-52.
Van Oss, C.J., Chaudhury, M.K., and Good, R.J. Interfacial Lifshitz-van der Waals
and polar interactions in macroscopic systems, Chemical Review, Vol. 88, 1988, p.
927.
Hefer, A.W., Little, D.N., and Lytton, R.L., A synthesis of bitumen-aggregate
adhesion including recent advances in quantifying the effects of water, Submitted
to the Association of Asphalt Paving Technologists for publication in Volume 74 of
the AAPT Journal, 2005.
Elphingstone, G.M., Adhesion and cohesion in asphalt-aggregate systems, Ph.D.
Dissertation, Chemical Engineering Dept., Texas A&M University, College Station,
Texas, 1997.
Cheng, D., Little, D.N., and Holste, J.C. Use of surface free energy of asphaltaggregate systems to predict moisture damage potential, Proceedings of the
Association of Asphalt Paving Technologists, Vol. 71, 2002, pp. 59-84.
Li, W., Evaluation of the Surface Energy of Aggregate Using the Chan Balance,
Unpublished Manuscript, Texas A&M University, Chemical Engineering
Department, College Station, Texas, 1997.
Charmas, B., and Leboda, R. Effect of heterogeneity on adsorption of solid
surfaces: Application of inverse gas chromatography in the studies of energetic
heterogeneity of adsobents, Journal of Chromatography A, No. 886, 2000, pp.
133-152.
Brendl, E., and Papirer, E. A new topological index for molecular probes used in
inverse gas chromatography, Journal of Colloid and Interface Science, No. 194,
1997, pp. 217-224.

[9]

Brendl, E., and Papirer, E. A new topological index for molecular probes used in
inverse gas chromatography for the surface nanorugosity evaluation, Journal of
Colloid and Interface Science, No. 194, 1997, pp. 207-216.
[10] Davis, T.C., and Petersen, J.C. An adaptation of inverse-liquid chromatography to
asphalt oxidation studies, Analytical Chemistry, Vol. 38, 1966, pp. 138-1940.
[11] Barbour, A.F., Barbour, R.V., and Petersen, C.J., A study of asphalt-aggregate
interactions using inverse-liquid chromatograph, Journal of Applied Chemistry and
Biotechnology, Vol. 24, No.11, 1974, pp. 645-654.
[12] Kim, S-S., Gardner, G.W., McKay, J.F., Robertson, R.E., and Branthaver, J.F,
Detection of strongly acidic compounds in extensively aged asphalt, In A.M.
Usmani (Ed.), Asphalt Science and Technology, Marcel Dekker, Inc., New York,
1997, pp. 103-117.
[13] WRI., Fundamental properties of asphalts and modified asphalts, Volume 1:
Interpretive Report, Draft Final Report (Contract DTFH61-99C-0022), Western
Research Institute, Laramie, Wyoming, 2003.
[14] WRI., Fundamental properties of asphalts and modified asphalts, Volume 2: New
test methods, Draft Final Report (Contract DTFH61-99C-0022), Western
Research Institute, Laramie, Wyoming, 2003.
[15] Skoog, D.A., and Leary, J.J., Principles of instrumental analysis, 4th Ed., Saunders
College Publishing, USA, 1992.
[16] Sun, C., and Berg, J.C. Effect of moisture on the surface free energy and acid-base
properties of mineral oxides, Journal of Chromatography A, No. 969, 2002, pp.
59-72.
[17] Lytton, R.L., Characterizing asphalt pavements for performance, Transportation
Research Record, No. 1723, 2000, pp. 5-16.
[18] Della Volpe, C. and Siboni, S. Some reflections on acid-base solid surface free
energy theories, Journal of Colloid and Interface Science, No. 195, 1997, pp. 121136.
[19] Della Volpe, C. and Siboni, S. Acid-base surface free energies of solids and the
definition of scales in the Good-van Oss- Chaudhury theory. Journal of Adhesion
Science and Technology, Vol. 14, No. 2, 2000, pp. 235 -272.
[20] Holtz, T. Personal communication. Digital Scientific, Germany, 2003.
[21] Van Oss, C.J., Interfacial Forces in Aqueous Media, Marcel Dekker, Inc., New
York, 1994.
[22] WRI., Fundamental properties of asphalts and modified asphalts, Volume 1:
Interpretive Report, Research Report FHWA-RD-99-212, Western Research
Institute, Laramie, Wyoming, 2001.

Potrebbero piacerti anche