Next Article in Journal
Near-Infrared Artificial Optical Synapse Based on the P(VDF-TrFE)-Coated InAs Nanowire Field-Effect Transistor
Next Article in Special Issue
Pavement Performance Investigation of Asphalt Mixtures with Plastic and Basalt Fiber Composite (PB) Modifier and Their Applications in Urban Bus Lanes Using Statics Analysis
Previous Article in Journal
Performance Test and Thermal Insulation Effect Analysis of Basalt-Fiber Concrete
Previous Article in Special Issue
Analysis of Rejuvenating Fiber Asphalt Mixtures’ Performance and Economic Aspects in High-Temperature Moisture Susceptibility
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessment of Aging Impact on Wax Crystallization in Selected Asphalt Binders

1
School of Architecture and Civil Engineering, Xihua University, Chengdu 610039, China
2
School of Civil Engineering, Southwest Jiaotong University, Chengdu 610031, China
3
Highway Engineering Key Laboratory of Sichuan Province, Southwest Jiaotong University, Chengdu 610031, China
*
Author to whom correspondence should be addressed.
Submission received: 17 October 2022 / Revised: 11 November 2022 / Accepted: 18 November 2022 / Published: 21 November 2022
(This article belongs to the Special Issue Performance-Related Material Properties of Asphalt Mixture Components)

Abstract

:
For a better understanding of the changing trend in crystalline components of asphalt binders, asphalt binders originating from the SHRP Materials Reference Library with different oxidation degrees (unaged, 20 h PAV, and 60 h PAV) were prepared. The native asphalt binders and their oxidized residues were characterized by liquid-state nuclear magnetic resonance (NMR) spectroscopy and high-temperature gas chromatography (HTGC). The results showed that, compared with other carbon types, the content of internal methylene carbons of long paraffinic chains between different SHRP binders was quite different. The NMR average length of a long paraffinic internal methylene chain showed a good correlation with the wax content obtained at −20 °C using the methyl ethyl ketone (MEK) precipitation method and also the recently developed variable-temperature Fourier-transform infrared spectroscopy (VT-FTIR) method. In most cases, the average length of straight internal methylene carbons of a long paraffinic chain terminated by a methyl group increased with the oxidation of the asphalt binder. However, the difference caused by oxidation was significantly smaller than the difference caused by the source of the asphalt binder. In general, oxidation will make the n-alkanes distributed in asphalt binder fall within a narrower range. The carbon number of n-alkanes in the asphalt binder generally grew with oxidation.

1. Introduction

The quality of asphalt binder is one of the important factors affecting the durability of asphalt pavement. In the past, researchers tried to establish a direct link between the chemical composition of asphalt and its quality; however, they failed to do so. This was mainly due to the complexity of asphalt chemical composition and a lack of in-depth understanding of the asphalt chemical composition’s classification. Among the chemical components of asphalt, wax is undoubtedly an important substance affecting the low-temperature service performance of asphalt [1,2,3,4]. Consequently, wax content and its behavior in asphalt with thermal history have received considerable attention [5,6,7]. Wax is generally considered to be a substance that crystallizes in asphalt at 25 °C and is mainly composed of saturated long linear hydrocarbons [8]. Owing to wax self-polymorphism and the difference in the state of existence in asphalt, there is no consensus on the wax effect on asphalt performance. This is one of the reasons why countries imposed different limits on wax content in asphalt.
Separating wax is the most direct method to study wax structure in asphalt [9]. Several methods were proposed to physically separate the wax from the asphalt binder. Due to the interference of the polar components (such as asphaltenes and resins) in the asphalt, it is difficult to separate the wax directly from the asphalt binder. Therefore, distillation or solvent precipitation is usually utilized to remove asphaltenes from the asphalt binder. The remaining components (maltenes) continue to be separated using chromatographic techniques. According to Corbett and Swarbrik’s method [10], the maltenes fraction is adsorbed on a chromatographic column (alumina is used as the adsorbent phase) and sequentially desorbed with solvents of increasing polarity. Saturates, aromatics, and resins are obtained from the maltenes. Using methyl ethyl ketone-benzene as a dewaxing solvent, solid wax can be separated from saturated and aromatic fractions at low temperatures. In contrast, Rostler and White [11] separated the maltenes into nitrogen bases, first acidaffins, second acidaffins, and paraffin fractions based on their reactivity with sulfuric acid (decreasing degree of hydration). The last fraction is what is called “wax”. In some studies, asphalt’s neutral fraction obtained from ion exchange chromatography (IEC) is regarded as a “wax fraction” [12]. When preparative size exclusion chromatography (SEC) is utilized to separate asphalt binder, the wax component is in the SEC-II fractions (non-associating components) [13]. Waxes are mainly composed of normal alkanes and isomeric alkanes, and the former produces the most adverse effect on asphalt performance. Considering urea readily forms crystalline adducts with straight-chain hydrocarbons, Netzel et al. [14] applied this technique to determine the content of n-alkanes and regarded this as solid wax content in asphalt.
With the development of modern analytical instruments, various advanced physical and chemical characterization methods have been employed to study wax behavior in asphalt [15,16,17,18]. Differential scanning calorimetry (DSC) is a common analytical method to study the crystalline fractions in asphalt binder. The number of crystallizable fractions in asphalt binder can be quantified by the size of the endothermic peak during heating. Using a thermal analysis of asphalt components, Corbett et al. [19] concluded that the endothermic peak of asphalt was mainly caused by saturates. However, Harrison et al. [20] believed that the saturates fraction was not the main factor leading to the endothermic behavior of asphalt binder, and the average linear side chain length was the main factor that affected the size of the endothermic peak. The effects of oxidation on the thermal behavior of pure wax-doped asphalt systems were studied by Kovinich et al. [6]. They found that the size of the endothermic peaks in the heat flow curve decreased with the oxidation of the binder. Atomic force microscopy (AFM) is often applied to observe the morphology of asphalt and the “bee structures” found in asphalt are thought to be the wax. Lu et al. [21] showed that the bee structures were fewer but larger after prolonged oxidation. They attributed this phenomenon to the highly increased stiffness of aged binder and reduced compatibility between the crystalline fraction and the more polarized asphalt matrix. Recently, Ding et al. [5,22,23] developed a variable-temperature Fourier-transform infrared spectroscopy technique to study the wax in asphalt, including quantification of the wax content and the precipitation and melting process of wax in asphalt. Aging is an important factor leading to the deterioration of asphalt performance. There have been a lot of literature reports on the mechanism and chemical reaction path of asphalt aging [24,25,26,27,28]. However, they did not find a clear rule regarding the effects of oxidation on wax content in asphalt binder.
As can be seen from the above literature review, although the wax in asphalt has been studied to a certain degree, the effect mechanisms of oxidative aging on wax in asphalt are still unclear. In this study, various asphalt binders from different crude oil sources were oxidized to a different degree in the laboratory. High-temperature gas chromatography (HTGC) and liquid-state nuclear magnetic resonance (NMR) spectroscopy were employed to analyze the structure of asphalt binders under different aging conditions. It is expected that some new insights into oxidation’s effects on the crystallizable fractions in asphalt could be provided by a series of laboratory tests.

2. Materials and Methods

2.1. Asphalt Binders

In this study, eight asphalt binders from the SHRP (Strategic Highway Research Program) materials reference library were utilized. The SHRP materials reference library was built during the SHRP and used to store the asphalt binders adopted in the implementation of the project. The physical and chemical properties of these asphalt binders have been extensively studied to reveal the mechanisms of differences in field performance. Due to the fact that the SHRP binders have different colloidal structures and crude oil sources, they are representative and generally applied as base binders for modified asphalts to validate the performance of the modifier. The short-term aging of asphalt was simulated by Rolling Thin-Film Oven (RTFO) experiments, and the long-term aging behavior of asphalt was simulated by a Pressure Aging Vessel (PAV). In addition to the traditional 20-h PAV test, 60 h oxidative aging time was also carried out to better simulate the degree of field asphalt. In a previous study [5], although the effects of oxidation on the wax contents of these SHRP binders were studied using the VT-FTIR method, no precise rule was found. This study adopts the same SHRP binders and aging modes but with different characterization methods. In this way, the obtained results could be mutually confirmed and compared. The basic properties of the selected SHRP binders can be found in the reference [23,29].

2.2. Nuclear Magnetic Resonance (NMR)

Nuclear magnetic resonance spectroscopy can provide information about the relative number of different types of hydrogen, carbon, or other atoms in a molecule, and the chemical environment in which these different types of atoms are exposed. The abscissa of the NMR spectrum is the chemical shift, and the ordinate is the intensity of the resonance absorption peak, whose peak area is proportional to the number of such nuclei in the molecule. Since the absolute value of the chemical shift of the general atomic nucleus is very small, the peak of tetramethylsilane (TMS) is usually utilized as a reference for comparison, and the relative change value is expressed as ppm (parts per million). In the case of mixtures such as asphalt, NMR provides average structure information for the entire sample. This information is useful in characterizing and differentiating the chemical properties of different asphalts as a whole. The nuclear overhauser effect (NOE) occurs due to the influence of 1H adjacent to 13C, which can significantly enhance the sensitivity of 13C NMR. However, its influence on different types of carbon is inconsistent and could result in the spectrum’s intensity not being proportional to the carbon amount. For this reason, this study adopted the measures of inverse gated decoupling and relaxation reagent to eliminate the NOE. In this way, a quantitative determination based on the integrated intensity of the spectrum could be carried out.

2.3. Gas Chromatography

Gas chromatography is used to separate the components of a sample with different partitioning coefficients between the mobile phase and stationary phase of the column [30]. In other words, the mobile phase has different adsorption values or solubility over the stationary phase. The mixture about to be separated is vaporized into a gas at the injection port, and the gas is carried by the mobile phase into the chromatographic column. Owing to the continuous flushing of the mobile phase, the components move downstream, and the components with the weakest adsorption (or dissolution) ability move downstream faster than other components. In this way, each component in the sample could flow out from respective weak to strong adsorption according to its adsorption (or dissolution) ability, so that the components can be separated in this way. The detector response versus retention time is the raw data obtained from the gas chromatography test. The time from injection to the chromatographic peak reaching the local maximum value is called the retention time of the component. Under a specific stationary phase and experimental conditions, each component has a specific retention time, which could be used as a qualitative indicator. The peak height reflects the response of the detector.

3. Results and Discussion

3.1. 13C NMR Spectra of Asphalt Binders

The 13C NMR spectra of SHRP asphalt binders under different oxidation states (unaged, 20 h PAV, and 60 h PAV) are presented in Figure 1. According to previous studies, the peaks between 110–160 ppm were attributed to aromatic carbons, and the peaks between 10–40 ppm were attributed to carbons in aliphatic chains and naphthenic rings [31]. Considering the objectives of this study, only aliphatic carbons were analyzed. It can be seen from the 13C NMR spectrum of unaged SHRP asphalt binders that all tested asphalts have five obvious main peaks. It was reported that the peak at 14.1 ppm was assigned to methyl carbons of aliphatic side chains [32]. The peaks at 22.7, 31.9, and 29.1 ppm correspond to carbons in methylene units successively further from the methyl group. The peak at 29.7 ppm, which is the largest peak in the aliphatic carbon region, is caused by internal methylene carbons of long paraffinic chains.
It is difficult to distinguish the difference between 13C spectra of various SHRP asphalt binders with the naked eye. In addition, the influence of the degree of oxidative aging in SHRP binders on the 13C chemical shift is also difficult to be evaluated directly. Therefore, the ratio of methylene in different positions to the 14.1 ppm methyl peak is presented in Figure 2. The ratio is an indication of the average length of a straight methylene chain terminated by a methyl group [33]. Compared with internal methylene carbons of long paraffinic chains, oxidative aging does not produce a major effect on the contents of methylene carbons adjacent to the methyl functional group in the average molecular structure of the asphalt binder. Moreover, in comparison with other carbon types, the content of internal methylene carbons of long paraffinic chains between various SHRP binders is quite different.
In order to determine whether the 13C spectra signal can be utilized to determine the wax content in the asphalt binder, the NMR average length of a long paraffinic internal methylene chain was compared with the wax content obtained at −20 °C using the methyl ethyl ketone (MEK) precipitation method and also the recently developed variable-temperature Fourier-transform infrared spectroscopy (VT-FTIR) method, respectively. Corresponding results are illustrated in Figure 3. It can be seen that the NMR signal shows a close correlation with the other two selected methods, especially the VT-FTIR method. Compared with the commonly used chemical precipitation method to determine the wax content, the NMR method has many merits, including straightforward steps, less effect on results by the operating factors, and usage of almost no toxic chemical solvents. The most important fact is that this method can effectively distinguish asphalt binders with different wax contents from each other. However, compared with VT-FTIR, the results obtained by the NMR method are less discriminating. The difference between the highest and lowest wax content obtained using the VT-FTIR method is up to 9.07, while it is only 4.5 using the NMR method. This may be due to the fact that NMR is tested at room temperature and deuterated chloroform is used as the solvent. The presence of solvents may destroy the form of the molecular structure in the asphalt matrix. In the future, low-temperature solid-state nuclear magnetic resonance spectroscopy will be employed to further test asphalt binders and verify the accuracy of the conjecture. Similar to a previous publication by Ding et al. [5], the NMR method has not found a clear rule of the effect of oxidation on the crystallizable fractions in asphalt binder. In most cases, the average length of straight internal methylene carbons of a long paraffinic chain terminated by a methyl group increases with the oxidation of the asphalt binder. However, the difference caused by oxidation is much smaller than the source of asphalt binder.

3.2. DEPT 135 Spectra of Asphalt Binders

The main function of using the distortionless enhancement by polarization transfer (DEPT) test is to distinguish methyl (CH3), methylene (CH2), and methine (CH) groups and clarify the aliphatic region of 13C spectra. In general, methyl and methine carbons are displayed as upward signals in the DEPT 135 experiment, whereas the methylene carbons appear as downward signals. DEPT 135 spectra of sample AAV and AAC-1 under three oxidation states (unaged, 20 h PAV, and 60 h PAV) are shown in Figure 4. As expected, the peaks at 22.7, 31.9, 29.1, and 29.7 ppm correspond to carbons in methylene units successively further from the methyl group, and internal methylene carbons of long paraffinic chains are downward. The peaks at 14.1 ppm and 19.71 ppm correspond to methyl peaks, and methyl branches on a straight methylene backbone are upward. The aromatic/aliphatic carbons ratio in asphalt binder can also be determined using the DEPT 135 NMR experiment. Considering the main purpose of this study, the corresponding results are not shown here.

3.3. Effect of Oxidation on n-Alkane Distribution

Under the selected optimal chromatographic analysis conditions, the carbon disulfide solutions of asphalt with different aging states (unaged, 20 h PAV, and 60 h PAV) were analyzed and determined. The results of removing the solvent peaks are presented in Figure 5. Since the correction factors of hydrocarbon compounds on the flame ionization detector (FID) are similar, the area normalization method is directly used for quantitative analysis to obtain the content of each n-alkane. The determination results of carbon number distribution are illustrated in Figure 6. Although the effects of oxidation on the total n-alkane content in asphalt cannot be determined by gas chromatogram, some interesting observations could be made. For the unaged ABG asphalt sample, the carbon number of n-alkanes is mainly distributed in a wide range between 30 and 80. Moreover, there are very few n-alkanes with carbon numbers below 20 and above 90. With the oxidation of the binder (20 h PAV), the carbon number of n-alkanes in the asphalt is mainly concentrated in a narrow range between 35 and 50. The sample hardly contains n-alkanes with carbon numbers lower than 30 and higher than 80. By extending the binder’s oxidation time (60 h PAV), the carbon number of n-alkanes in asphalt generally increased. A similar trend was found in other asphalt binders. Furthermore, the carbon number distribution of n-alkanes tends to be uniform, especially for AAM-1 and AAC-1.

4. Summary and Conclusions

In this study, the effects of oxidation on crystallizable fractions in different asphalt binders were investigated by liquid-state nuclear magnetic resonance (NMR) spectroscopy and high-temperature gas chromatography (HTGC). The following conclusions can be drawn from a series of laboratory tests:
(1)
The results of the 13C NMR spectra of unaged SHRP asphalt binders revealed that all tested asphalts had five obvious main peaks. Compared with internal methylene carbons of long paraffinic chains, oxidative aging did not produce a major effect on the contents of methylene carbons adjacent to the methyl functional group in the average molecular structure of the asphalt binder.
(2)
Compared with other carbon types, the content of internal methylene carbons of long paraffinic chains between different SHRP binders was quite different. The NMR average length of a long paraffinic internal methylene chain showed a close correlation with wax content obtained at −20 °C using the methyl ethyl ketone (MEK) precipitation method and also the recently developed variable-temperature Fourier-transform infrared spectroscopy (VT-FTIR) method.
(3)
The application of the NMR method could not find a clear rule in terms of the effect of oxidation on the crystallizable fractions in asphalt binder. In most cases, the average length of straight internal methylene carbons of a long paraffinic chain terminated by a methyl group increased with the oxidation of the asphalt binder. However, the difference caused by oxidation was considerably smaller than the source of asphalt binder.
(4)
In general, oxidation will cause the n-alkanes distributed in asphalt binder to fall within a narrower range. The carbon number of n-alkanes in asphalt binder generally increases with oxidation.
(5)
Future studies will cover the evaluation of UV effects, the short-term RTFOT test, as well as the determination of rheological properties.

Author Contributions

Conceptualization, H.D. and Y.Q.; Methodology, A.R.; Data curation, W.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by 1. Talent Introduction Project of Xihua University (School Key Project) (Z202111); 2. Scientific research project of Sichuan Provincial Department of Education (16ZB0164); 3. Open Research Fund of Sichuan Provincial Key Laboratory of Road Engineering (15206569); 4. Open project of Sichuan Key Laboratory of Green Building and Energy Saving (szjj2015-074).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data supporting the figures in this paper, as well as other findings of this study, are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ma, J.; Hesp, S.A.M.; Chan, S.; Li, J.Z.; Lee, S. Lessons learned from 60 years of pavement trials in continental climate regions of Canada. Chem. Eng. J. 2022, 444, 136389. [Google Scholar] [CrossRef]
  2. Ding, H.; Zhang, H.; Liu, H.; Qiu, Y. Thermoreversible aging in model asphalt binders. Constr. Build. Mater. 2021, 303, 124355. [Google Scholar] [CrossRef]
  3. Edwards, Y.; Isacsson, U. Wax in bitumen. Road Mater. Pavement Des. 2005, 6, 281–309. [Google Scholar] [CrossRef]
  4. Ding, H.; Zhang, H.; Xie, Q.; Rahman, A.; Qiu, Y. Synthesis and characterization of nano-SiO2 hybrid poly(methyl methacrylate) nanocomposites as novel wax inhibitor of asphalt binder. Colloids Surf. A Physicochem. Eng. Asp. 2022, 653, 130023. [Google Scholar] [CrossRef]
  5. Ding, H.; Hesp, S. Variable-temperature Fourier-transform infrared spectroscopy study of asphalt binders from the SHRP Materials Reference Library. Fuel 2021, 298, 120819. [Google Scholar] [CrossRef]
  6. Kovinich, J.; Hesp, S.; Ding, H. Modulated differential scanning calorimetry study of wax-doped asphalt binders. Thermochim. Acta 2021, 699, 178894. [Google Scholar] [CrossRef]
  7. Qiu, Y.; Zhang, H.; Ding, H.; Rahman, A. Measurement and thermodynamic modeling of wax precipitation in asphalt binder. Mater. Des. 2022, 221, 110957. [Google Scholar] [CrossRef]
  8. Knotnerus, J.; Krom, C.J. The constitution of wax isolated from bitumen. Am. Chem. Soc. Div. Pet. Chem. Prepr. 1964, 9, B-39. [Google Scholar]
  9. Ding, H.; Zhang, H.; Zhang, H.; Liu, D.; Qiu, Y.; Hussain, A. Separation of wax fraction in asphalt binder by an improved method and determination of its molecular structure. Fuel 2022, 322, 124081. [Google Scholar] [CrossRef]
  10. Thenoux, G.; Bell, C.; Wilson, J.; Eakin, D.; Schroeder, M. Experiences with the Corbett-Swarbrick procedure for separation of asphalt into four generic fractions. Transp. Res. Rec. 1988, 1171, 66–70. [Google Scholar]
  11. Rostler, F.S.; White, R.M. Composition and changes in composition of highway asphalts, 85–100 penetration grade. J. Assoc. Asph. Paving Technol. 1962, 31, 35–89. [Google Scholar]
  12. Redelius, P.; Lu, X.; Isacsson, U. Non-classical wax in bitumen. Road Mater. Pavement Des. 2002, 3, 7–21. [Google Scholar] [CrossRef]
  13. Garrick, N.W. Use of gel-permeation chromatography in predicting properties of asphalt. J. Mater. Civ. Eng. 1994, 6, 376–389. [Google Scholar] [CrossRef]
  14. Netzel, D.A.; Rovani, J.F. Direct separation and quantitative determination of (n-, iso-) alkanes in neat asphalt using urea adduction and high-temperature gas chromatography (HTGC). Energy Fuels 2007, 21, 333–338. [Google Scholar] [CrossRef]
  15. Hasheminejad, N.; Pipintakos, G.; Vuye, C.; De Kerf, T.; Ghalandari, T.; Blom, J.; Van den bergh, W. Utilizing deep learning and advanced image processing techniques to investigate the microstructure of a waxy bitumen. Constr. Build. Mater. 2021, 313, 125481. [Google Scholar] [CrossRef]
  16. Blom, J.; Soenen, H.; Van den Brande, N.; Van den bergh, W. New evidence on the origin of ‘bee structures’ on bitumen and oils, by atomic force microscopy (AFM) and confocal laser scanning microscopy (CLSM). Fuel 2021, 303, 121265. [Google Scholar] [CrossRef]
  17. Ding, H.; Zhang, H.; Zhang, H.; Qiu, Y. Direct Observation of Crystalline Wax in Asphalt Binders by Variable-temperature Polarizing Microscope. J. Mater. Civ. Eng. 2022, 34, 04022244. [Google Scholar] [CrossRef]
  18. Ding, H.; Zhang, H.; Zheng, X.; Zhang, C. Characterisation of crystalline wax in asphalt binder by X-ray diffraction. Road Mater. Pavement Des. 2022, 1–7. [Google Scholar] [CrossRef]
  19. Noel, F.; Corbett, L.W. A study of the crystalline phases in asphalts. J. Inst. Pet. 1970, 56, 261–268. [Google Scholar]
  20. Harrison, I.R.; Wang, G.; Hsu, T.C. A Differential Scanning Calorimetry Study of Asphalt Binders; National Research Council: Washington, DC, USA, 1993. [Google Scholar]
  21. Lu, X.; Sjövall, P.; Soenen, H.; Blom, J.; Makowska, M. Oxidative aging of bitumen: A structural and chemical investigation. Road Mater. Pavement Des. 2021, 23, 1091–1106. [Google Scholar] [CrossRef]
  22. Ding, H.; Hesp, S. Variable-temperature Fourier-transform infrared spectroscopy study of wax precipitation and melting in Canadian and Venezuelan asphalt binders. Constr. Build. Mater. 2020, 264, 120212. [Google Scholar] [CrossRef]
  23. Ding, H.; Hesp, S. Quantification of crystalline wax in asphalt binders using variable-temperature Fourier-transform infrared spectroscopy. Fuel 2020, 277, 118220. [Google Scholar] [CrossRef]
  24. Tauste, R.; Moreno-Navarro, F.; Sol-Sánchez, M.; Rubio-Gámez, M.C. Understanding the bitumen ageing phenomenon: A review. Constr. Build. Mater. 2018, 192, 593–609. [Google Scholar] [CrossRef]
  25. Malinowski, S.; Bandura, L.; Woszuk, A. Influence of atmospheric oxygen on the structure and electronic properties of bitumen components—A DFT study. Fuel 2022, 325, 124551. [Google Scholar] [CrossRef]
  26. Hofko, B.; Maschauer, D.; Steiner, D.; Mirwald, J.; Grothe, H. Bitumen ageing—Impact of reactive oxygen species. Case Stud. Constr. Mater. 2020, 13, e00390. [Google Scholar] [CrossRef]
  27. Rathore, M.; Haritonovs, V.; Zaumanis, M. Performance Evaluation of Warm Asphalt Mixtures Containing Chemical Additive and Effect of Incorporating High Reclaimed Asphalt Content. Materials 2021, 14, 3793. [Google Scholar] [CrossRef]
  28. Ding, H.; Liu, H.; Qiu, Y.; Rahman, A. Effects of crystalline wax and asphaltene on thermoreversible aging of asphalt binder. Int. J. Pavement Eng. 2022, 23, 3997–4006. [Google Scholar] [CrossRef]
  29. Zhang, H.; Xie, Q.; Ding, H.; Rahman, A. Spectroscopic ellipsometry studies on optical constants of crystalline wax-doped asphalt binders. Int. J. Pavement Eng. 2022, 1–13. [Google Scholar] [CrossRef]
  30. Gupta, A.K.; Severin, D. Characterization of petroleum waxes by high temperature gas chromatography—Correlation with physical properties. Pet. Sci. Technol. 1997, 15, 943–957. [Google Scholar] [CrossRef]
  31. Siddiqui, M.N.; Ali, M.F. Investigation of chemical transformations by NMR and GPC during the laboratory aging of Arabian asphalt. Fuel 1999, 78, 1407–1416. [Google Scholar] [CrossRef]
  32. Jennings, P.W.; Desando, M.A.; Raub, M.F.; Moats, R.; Mendez, T.M.; Stewart, F.F.; Hoberg, J.O.; Pribanic, J.A.S.; Smith, J.A. NMR spectroscopy in the characterization of eight selected asphalts. Fuel Sci. Technol. Int. 1992, 10, 887–907. [Google Scholar] [CrossRef]
  33. Musser, B.J.; Kilpatrick, P.K. Molecular Characterization of Wax Isolated from a Variety of Crude Oils. Energy Fuels 1998, 12, 715–725. [Google Scholar] [CrossRef]
Figure 1. 13C NMR spectra of SHRP asphalt binders. (a) Unaged, (b) 20 h PAV, and (c) 60 h PAV.
Figure 1. 13C NMR spectra of SHRP asphalt binders. (a) Unaged, (b) 20 h PAV, and (c) 60 h PAV.
Materials 15 08248 g001aMaterials 15 08248 g001b
Figure 2. Percent carbons from 13C NMR for SHRP asphalt binders. (a) [email protected] ppm/[email protected] ppm (b) [email protected] ppm/[email protected] ppm (c) [email protected] ppm/[email protected] ppm (d) [email protected] ppm/[email protected] ppm.
Figure 2. Percent carbons from 13C NMR for SHRP asphalt binders. (a) [email protected] ppm/[email protected] ppm (b) [email protected] ppm/[email protected] ppm (c) [email protected] ppm/[email protected] ppm (d) [email protected] ppm/[email protected] ppm.
Materials 15 08248 g002aMaterials 15 08248 g002b
Figure 3. Comparison between the average length of straight internal methylene carbon of long paraffinic chain terminated by a methyl group and wax contents obtained using (a) precipitation and (b) VT-FTIR methods.
Figure 3. Comparison between the average length of straight internal methylene carbon of long paraffinic chain terminated by a methyl group and wax contents obtained using (a) precipitation and (b) VT-FTIR methods.
Materials 15 08248 g003
Figure 4. DEPT135 spectra of selected asphalt binders. (a) AAV. (b) AAC-1.
Figure 4. DEPT135 spectra of selected asphalt binders. (a) AAV. (b) AAC-1.
Materials 15 08248 g004
Figure 5. Gas chromatogram of asphalt binders. (a) ABG, (b) AAV, (c) AAM-1, and (d) AAC-1.
Figure 5. Gas chromatogram of asphalt binders. (a) ABG, (b) AAV, (c) AAM-1, and (d) AAC-1.
Materials 15 08248 g005
Figure 6. N-alkane distribution in asphalt binders. (a) ABG, (b) AAV, (c) AAM-1, and (d) AAC-1.
Figure 6. N-alkane distribution in asphalt binders. (a) ABG, (b) AAV, (c) AAM-1, and (d) AAC-1.
Materials 15 08248 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, W.; Rahman, A.; Ding, H.; Qiu, Y. Assessment of Aging Impact on Wax Crystallization in Selected Asphalt Binders. Materials 2022, 15, 8248. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15228248

AMA Style

Wang W, Rahman A, Ding H, Qiu Y. Assessment of Aging Impact on Wax Crystallization in Selected Asphalt Binders. Materials. 2022; 15(22):8248. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15228248

Chicago/Turabian Style

Wang, Wenqi, Ali Rahman, Haibo Ding, and Yanjun Qiu. 2022. "Assessment of Aging Impact on Wax Crystallization in Selected Asphalt Binders" Materials 15, no. 22: 8248. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15228248

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop