Download as pdf or txt
Download as pdf or txt
You are on page 1of 127

This PDF is available at http://nap.nationalacademies.

org/24959

Long-Term Aging of Asphalt Mixtures for


Performance Testing and Prediction
(2017)

DETAILS
124 pages | 8.5 x 11 | PAPERBACK
ISBN 978-0-309-46863-3 | DOI 10.17226/24959

CONTRIBUTORS
Y. Richard Kim, Cassie Castorena, Michael Elwardany, Farhad Yousefi Rad, Shane
Underwood, Akshay Gundha, Padmini Gudipudi, Mike J. Farrar, and Ronald R.
BUY THIS BOOK Glaser; National Cooperative Highway Research Program; Transportation Research
Board; National Academies of Sciences, Engineering, and Medicine

FIND RELATED TITLES SUGGESTED CITATION


National Academies of Sciences, Engineering, and Medicine. 2017. Long-Term
Aging of Asphalt Mixtures for Performance Testing and Prediction. Washington,
DC: The National Academies Press. https://doi.org/10.17226/24959.

Visit the National Academies Press at nap.edu and login or register to get:
– Access to free PDF downloads of thousands of publications
– 10% off the price of print publications
– Email or social media notifications of new titles related to your interests
– Special offers and discounts

All downloadable National Academies titles are free to be used for personal and/or non-commercial
academic use. Users may also freely post links to our titles on this website; non-commercial academic
users are encouraged to link to the version on this website rather than distribute a downloaded PDF
to ensure that all users are accessing the latest authoritative version of the work. All other uses require
written permission. (Request Permission)

This PDF is protected by copyright and owned by the National Academy of Sciences; unless otherwise
indicated, the National Academy of Sciences retains copyright to all materials in this PDF with all rights
reserved.
Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

N AT I O N A L C O O P E R AT I V E H I G H W AY R E S E A R C H P R O G R A M

NCHRP RESEARCH REPORT 871


Long-Term Aging of Asphalt
Mixtures for Performance
Testing and Prediction

Y. Richard Kim
Cassie Castorena
Michael Elwardany
Farhad Yousefi Rad
North Carolina State University
Raleigh, NC

Shane Underwood
Akshay Gundla
Padmini Gudipudi
Arizona State University
Tempe, AZ

Mike J. Farrar
Ronald R. Glaser
Western Research Institute
Laramie, WY

Subscriber Categories
Design • Materials • Pavements

Research sponsored by the American Association of State Highway and Transportation Officials
in cooperation with the Federal Highway Administration

2018

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

NATIONAL COOPERATIVE HIGHWAY NCHRP RESEARCH REPORT 871


RESEARCH PROGRAM
Systematic, well-designed research is the most effective way to solve Project 09-54
many problems facing highway administrators and engineers. Often, ISSN 2572-3766 (Print)
highway problems are of local interest and can best be studied by ISSN 2572-3774 (Online)
highway departments individually or in cooperation with their state ISBN 978-0-309-44683-9
universities and others. However, the accelerating growth of highway Library of Congress Control Number 2018935615
transportation results in increasingly complex problems of wide inter-
© 2018 National Academy of Sciences. All rights reserved.
est to highway authorities. These problems are best studied through a
coordinated program of cooperative research.
Recognizing this need, the leadership of the American Association
of State Highway and Transportation Officials (AASHTO) in 1962 ini- COPYRIGHT INFORMATION
tiated an objective national highway research program using modern Authors herein are responsible for the authenticity of their materials and for obtaining
scientific techniques—the National Cooperative Highway Research written permissions from publishers or persons who own the copyright to any previously
Program (NCHRP). NCHRP is supported on a continuing basis by published or copyrighted material used herein.
funds from participating member states of AASHTO and receives the Cooperative Research Programs (CRP) grants permission to reproduce material in this
full cooperation and support of the Federal Highway Administration, publication for classroom and not-for-profit purposes. Permission is given with the
United States Department of Transportation. understanding that none of the material will be used to imply TRB, AASHTO, FAA, FHWA,
FMCSA, FRA, FTA, Office of the Assistant Secretary for Research and Technology, PHMSA,
The Transportation Research Board (TRB) of the National Academies
or TDC endorsement of a particular product, method, or practice. It is expected that those
of Sciences, Engineering, and Medicine was requested by AASHTO to reproducing the material in this document for educational and not-for-profit uses will give
administer the research program because of TRB’s recognized objectivity appropriate acknowledgment of the source of any reprinted or reproduced material. For
and understanding of modern research practices. TRB is uniquely suited other uses of the material, request permission from CRP.
for this purpose for many reasons: TRB maintains an extensive com-
mittee structure from which authorities on any highway transportation
subject may be drawn; TRB possesses avenues of communications and NOTICE
cooperation with federal, state, and local governmental agencies, univer-
The research report was reviewed by the technical panel and accepted for publication
sities, and industry; TRB’s relationship to the National Academies is an
according to procedures established and overseen by the Transportation Research Board
insurance of objectivity; and TRB maintains a full-time staff of special- and approved by the National Academies of Sciences, Engineering, and Medicine.
ists in highway transportation matters to bring the findings of research
The opinions and conclusions expressed or implied in this report are those of the
directly to those in a position to use them. researchers who performed the research and are not necessarily those of the Transportation
The program is developed on the basis of research needs identified by Research Board; the National Academies of Sciences, Engineering, and Medicine; or the
chief administrators and other staff of the highway and transportation program sponsors.
departments, by committees of AASHTO, and by the Federal Highway The Transportation Research Board; the National Academies of Sciences, Engineering,
Administration. Topics of the highest merit are selected by the AASHTO and Medicine; and the sponsors of the National Cooperative Highway Research Program
Special Committee on Research and Innovation (R&I), and each year do not endorse products or manufacturers. Trade or manufacturers’ names appear herein
R&I’s recommendations are proposed to the AASHTO Board of Direc- solely because they are considered essential to the object of the report.
tors and the National Academies. Research projects to address these
topics are defined by NCHRP, and qualified research agencies are
selected from submitted proposals. Administration and surveillance of
research contracts are the responsibilities of the National Academies
and TRB.
The needs for highway research are many, and NCHRP can make
significant contributions to solving highway transportation problems
of mutual concern to many responsible groups. The program, however,
is intended to complement, rather than to substitute for or duplicate,
other highway research programs.

Published research reports of the

NATIONAL COOPERATIVE HIGHWAY RESEARCH PROGRAM


are available from

Transportation Research Board


Business Office
500 Fifth Street, NW
Washington, DC 20001

and can be ordered through the Internet by going to


http://www.national-academies.org
and then searching for TRB
Printed in the United States of America

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

The National Academy of Sciences was established in 1863 by an Act of Congress, signed by President Lincoln, as a private, non-
governmental institution to advise the nation on issues related to science and technology. Members are elected by their peers for
outstanding contributions to research. Dr. Marcia McNutt is president.

The National Academy of Engineering was established in 1964 under the charter of the National Academy of Sciences to bring the
practices of engineering to advising the nation. Members are elected by their peers for extraordinary contributions to engineering.
Dr. C. D. Mote, Jr., is president.

The National Academy of Medicine (formerly the Institute of Medicine) was established in 1970 under the charter of the National
Academy of Sciences to advise the nation on medical and health issues. Members are elected by their peers for distinguished contributions
to medicine and health. Dr. Victor J. Dzau is president.

The three Academies work together as the National Academies of Sciences, Engineering, and Medicine to provide independent,
objective analysis and advice to the nation and conduct other activities to solve complex problems and inform public policy decisions.
The National Academies also encourage education and research, recognize outstanding contributions to knowledge, and increase
public understanding in matters of science, engineering, and medicine.

Learn more about the National Academies of Sciences, Engineering, and Medicine at www.national-academies.org.

The Transportation Research Board is one of seven major programs of the National Academies of Sciences, Engineering, and Medicine.
The mission of the Transportation Research Board is to increase the benefits that transportation contributes to society by providing
leadership in transportation innovation and progress through research and information exchange, conducted within a setting that
is objective, interdisciplinary, and multimodal. The Board’s varied committees, task forces, and panels annually engage about 7,000
engineers, scientists, and other transportation researchers and practitioners from the public and private sectors and academia, all
of whom contribute their expertise in the public interest. The program is supported by state transportation departments, federal
agencies including the component administrations of the U.S. Department of Transportation, and other organizations and individuals
interested in the development of transportation.

Learn more about the Transportation Research Board at www.TRB.org.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

COOPERATIVE RESEARCH PROGRAMS

CRP STAFF FOR NCHRP RESEARCH REPORT 871


Christopher J. Hedges, Director, Cooperative Research Programs
Lori L. Sundstrom, Deputy Director, Cooperative Research Programs
Edward T. Harrigan, Senior Program Officer
Anthony P. Avery, Senior Program Assistant
Eileen P. Delaney, Director of Publications
Natalie Barnes, Associate Director of Publications
Scott E. Hitchcock, Senior Editor

NCHRP PROJECT 09-54 PANEL


Field of Materials and Construction—Area of Bituminous Materials
Dean A. Maurer, Practical Asphalt Solutions Technology, Lewisberry, PA (Chair)
Bouzid Choubane, Florida DOT, Gainesville, FL
Shongtao Dai, Minnesota DOT, Maplewood, MN
Dale S. Decker, Dale S. Decker, LLC, Eagle, CO
Changlin Pan, Nevada DOT, Carson City, NV
Murari M. Pradhan, Arizona DOT, Phoenix, AZ
Maria C. Rodezno, Auburn University, Auburn, AL
Mansour Solaimanian, Pennsylvania State University, University Park, PA
Jack Youtcheff, FHWA Liaison
Nelson H. Gibson, TRB Liaison

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

FOREWORD

By Edward T. Harrigan
Staff Officer
Transportation Research Board

This report presents a proposed standard method for long-term laboratory aging of
asphalt mixtures for performance testing; the method is intended for consideration as a
replacement for the method in AASHTO R 30, “Mixture Conditioning of Hot Mix Asphalt
(HMA).” Thus, the report will be of immediate interest to materials and design engineers in
state highway agencies and in the construction industry with responsibility for design and
production of hot and warm-mix asphalt.

Accurately characterizing the in situ aging of asphalt pavement materials over their long-
term service life is of utmost importance to the implementation of mechanistic–empirical
pavement design and analysis methods. Current pavement performance prediction models
vary in their ability and level of sophistication for numerically simulating the increased
stiffness of asphalt materials from oxidative aging and the competing reduction in modulus
caused by accumulated pavement damage and deterioration.
For the past 25 years, the most commonly used method for aging asphalt materials for
performance testing for input to prediction models has been the long-term procedure in
AASHTO R 30. R 30 prescribes aging compacted asphalt mixture specimens at 85°C for
5 days, a time and temperature combination that the original Strategic Highway Research
Program (SHRP) research estimated—based on limited field calibration—to reflect a criti-
cal duration of field exposure from 5 to 10 years. However, an accumulation of laboratory
and field data since the end of SHRP has demonstrated that a single time–temperature com-
bination cannot reasonably simulate the effects of the range of climates found throughout
the United States. Further, these data suggest that this single combination even under­
estimates the level of the asphalt pavement aging that occurs in as little as 5 years in many
climates.
The objective of this research was to develop a long-term aging procedure for asphalt
mixtures appropriate to fabricate performance test specimens that provide input data for
use with AASHTOWare Pavement ME Design software and other mechanistic design and
analysis systems. The research was performed by North Carolina State University, Raleigh,
North Carolina; in conjunction with Arizona State University, Tempe, Arizona, and Western
Research Institute, Laramie, Wyoming.
The long-term aging procedure for asphalt mixtures was developed, calibrated, and vali-
dated through a series of laboratory experiments using field cores and original asphalt binders
and aggregates from 18 asphalt mixtures, including both HMA and WMA. The mixtures
were obtained from nine field projects representing the wide range of climates found in in
the United States and Canada. The method improves on R 30 in that the laboratory aging
time is specifically determined by the climate at the project location. All aging is conducted

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

on loose mix at 95°C. The use of loose mix provides fast, even aging compared to compacted
specimens. Aging at 95°C also substantially reduces the necessary aging time compared to
the 85°C specified in R 30; however, the research found that using an aging temperature of
100°C or greater can introduce chemical changes in the asphalt binder not present in field-
aged binders. Besides climate, the procedure also accounts for the asphalt mixture’s depth
in the pavement. Finally, to simplify the selection of laboratory aging time, the proposed
procedure presents a series of laboratory aging duration maps to match 4, 8, and 16 years
of field aging at depths of 6 mm, 20 mm, and 50 mm below the pavement surface.
The practical outcome of the project is a proposed AASHTO standard method for long-
term aging that can replace the procedure in AASHTO R 30.
This report fully documents the research and includes the proposed standard method. In
addition, the following appendices, not printed herein, are available for download from the
TRB website (trb.org) by searching for “NCHRP Research Report 871.”
Appendix A: Previous Studies of Laboratory Aging of Compacted Specimens and Loose
Mixtures
Appendix B: Previous Studies of Modeling Asphalt Binder Aging in Pavements
Appendix C: Evaluation of the Sensitivity of the Mechanical Properties of Asphalt Concrete
to Asphalt Binder Oxidation
Appendix D: Evaluation of Different Chemical and Rheological Aging Index Properties
Appendix E: Factors Affecting Oxidation Reaction Mechanisms in Asphalt Concrete
Appendix F: Evaluation of Asphalt Mixture Laboratory Long-Term Aging
Appendix G: Investigation of Proper Long-Term Aging Temperature
Appendix H: Climatic Aging Index
Appendix I: Performance Testing of Field Cores

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

CONTENTS

1 Summary
3 Chapter 1 Background
4 Project Objectives and Scope
5 Previous Research into Long-Term Aging of Asphalt Mixtures
5 Compacted Specimen Aging Versus Loose Mixture Aging
6 Oven Aging Versus Pressure Aging
7 Laboratory Aging Temperature
7 Aging Index Properties (AIPs)
8 Modeling of Oxidative Aging
12 Chapter 2 Research Approach
12 Overview of Research Approach
12 AIP Selection
12 Sensitivity Study
14 Selection of the Proposed Aging Method
15 Determination of Project-Specific Aging Durations
15 Climate-Based Determination of Predefined Aging Durations
15 Development of Pavement Aging Model
16 Test Materials and Field Projects
16 Group A Materials
17 Group B Materials/Projects
17 Sample Preparation Methods
17 Asphalt Mastic Preparation
18 FAM Preparation
19 Asphalt Mixture Aging
21 Asphalt Binder Aging
22 Field Core Preparation
22 Micro-Extraction and Recovery
23 Test Methods
23 Asphalt Binders
23 Asphalt Mixtures
24 Chapter 3 Findings and Applications
24 Findings
24 Sensitivity Study
26 Selection of the Chemical and Rheological Aging Index Properties
29 Selection of Long-Term Aging Method
58 Climate-Based Determination of Predefined Aging Durations
78 Aging Model to Predict Field Aging Throughout Pavement Depth
90 Applications
90 Integration of Pavement Aging Model in Mechanistic–Empirical Design

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

94 Conclusions
94 Sensitivity Study
94 Selection of the Chemical and Rheological Aging Index Properties
95 Selection of the Long-Term Aging Method
96 Climate-Based Determination of Predefined Aging Durations
97 Aging Model to Predict Field Aging
97 Integration of the Pavement Aging Model in Mechanistic–Empirical
Design
98 Chapter 4 Suggested Future Research
100 Chapter 5 Proposed Standard Method of Test for Long-Term
Conditioning of Hot Mix Asphalt (HMA)
for Performance Testing
112 References
116 Appendices A–I

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

SUMMARY

Long-Term Aging of Asphalt


Mixtures for Performance
Testing and Prediction
This report presents the experimental, analytical, and computational research conducted
under NCHRP Project 09-54. This project’s goals were to (1) develop a long-term aging
procedure for asphalt mixtures appropriate to the fabrication of performance test specimens
and (2) develop an asphalt pavement aging model for mechanistic–empirical (ME) pave-
ment design and analysis. Original component materials and field cores from in-service and
test track pavements from eight states and the Province of Manitoba, Canada, were used in
this study. These pavement projects yielded a total of 18 different asphalt mixtures, includ-
ing warm-mix asphalt (WMA).
In order to track the aging level of the laboratory-aged mixtures and field cores, a com-
prehensive study was first performed to identify efficient and accurate binder aging index
properties (AIPs). Both chemistry-based and rheology-based parameters were included in
the study. The logarithm of the binder shear modulus, log G*, was selected as the rheologi-
cal AIP and the total absorbance under the carbonyl and sulfoxide infrared (IR) peaks was
selected as the chemical AIP. Also, a sensitivity study was performed to evaluate the sensitiv-
ity of the mixture dynamic modulus to changes in the binder AIP that are caused by oxida-
tive aging. The sensitivity study provided thresholds by which to evaluate the significance
of observed differences in asphalt binder AIPs in terms of asphalt mixture performance. As
a rule of thumb, it was concluded that a 15% change in the binder dynamic shear modulus
value would lead to about a 10% change in the mixture dynamic modulus value.
Factors that were investigated as part of the selection of the aging procedure included:
(1) loose mixture aging versus compacted specimen aging, (2) oven aging versus pressure
aging, and (3) a 95°C aging temperature versus 135°C. The selection process was based on
considerations of practicality, efficiency, and versatility. Loose mixture aging reduced the
aging time significantly compared to compacted specimen aging. Also, significant aging gra-
dients were found in the aged compacted specimens, which violated the representative vol-
ume element requirement for performance test specimens. Pressure aging expedited the aging
process; however, a much larger pressure aging vessel (PAV) than the binder PAV would be
needed in order to age a sufficient quantity of loose mixture for the fabrication of performance
test specimens. Aging at 135°C caused changes in the chemistry of the binder; such changes
do not occur in the field. These chemical changes led to significantly different cracking perfor-
mance results compared to the results obtained after aging at 95°C. Based on these findings,
loose mixture aging in the oven at 95°C is proposed as the long-term aging procedure for the
fabrication of asphalt mixture performance test specimens. Based on a limited study, the same
recommendation is tentatively made for the laboratory long-term aging of WMA mixtures.
A comprehensive aging model for asphalt mixtures in pavement systems should account
for both binder oxidation kinetics and diffusion. In this study, a rheology-based kinetics

1  

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

2   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

model was developed for loose mixture aging based on an existing chemistry-based kinetics
model that was initially developed for binder aging. The rheology-based kinetics model was
verified using isothermal and non-isothermal laboratory aging conditions. This kinetics
model can predict the binder dynamic shear modulus as a function of aging temperature
and duration. The kinetics model and laboratory experiments demonstrate that the labora-
tory aging duration that is required to match a given field condition is independent of the
material-specific kinetics.
A climatic aging index (CAI), based on a simplification of the developed kinetics model,
was developed to prescribe the required laboratory aging durations to reflect different pave-
ment temperature histories at different locations and depths. The field aging levels obtained
from field cores were compared against loose mix aging rates at 95°C in the laboratory to
determine the aging duration that is required to match field aging. This CAI study resulted
in a procedure that calculates the laboratory aging durations to match the field aging at any
pavement depths for any locations. This procedure was used to develop a series of laboratory
aging duration maps to match 4, 8, and 16 years of field aging at depths of 6 mm, 20 mm,
and 50 mm below the pavement surface as examples.
Pavement temperature as a function of pavement depth was predicted using Enhanced
Integrated Climatic Model (EICM) data. The predicted pavement temperature history was
coupled with the kinetics model to predict field aging throughout the pavement depth. The
kinetics model was calibrated successfully (R 2 = 0.75) against field core measurements
of log G* at a depth of 20 mm, which represents a reasonable depth for the evaluation of
the bulk behavior of surface layers and avoids the effect of ultraviolet (UV) oxidation at the
pavement surface. These results suggest that the kinetics model can be extended in the future
to replace the Global Aging System (GAS) model to improve the prediction of changes in
asphalt binder properties with oxidative aging in pavement performance prediction software
(e.g., Pavement ME Design). The calibrated kinetics model tends to under-predict aging near
the pavement surface and over predict aging at depths below 20 mm, which matches intuition
in the absence of a diffusion model and consideration of UV oxidation for the pavement
near-surface. These results highlight the need for the development of a diffusion model that
considers the morphological properties of asphalt mixtures in order to predict field aging
more accurately.
The new model accounts for mixture-specific kinetics parameters that can be determined
directly from loose mix aging; however, it still requires cumbersome mixture preparation
and corresponding binder extraction and recovery. An empirical model thus was developed
to determine loose mix aging kinetics parameters from Universal Simple Aging Test binder
aging. The developed empirical model facilitates the determination of mixture-specific
kinetics parameters and eliminates the need for binder extraction and recovery.
Within pavement performance prediction frameworks, the predicted changes in binder
properties as a result of oxidation must be translated to the corresponding changes in asphalt
mixture properties. Therefore, as an example application, three common asphalt mixture
dynamic modulus predictive models [i.e., Witczak, Hirsch, and North Carolina State Uni-
versity (NCSU) Artificial Neural Networks (ANN)] were used to predict the dynamic mod-
ulus values at different aging levels, and the prediction results were compared against the
measured dynamic modulus values. Significant errors were found, which suggests the need
to adjust the existing models so that they can predict aged mixture properties based on
inputs from binder AIPs.
Finally, suggestions are made for future research.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

CHAPTER 1

Background

Aging has long been recognized as a major driver of distress for asphalt concrete and, by
extension, asphalt pavements. Aging causes the material to stiffen and embrittle, which leads to
a high potential for cracking. The term aging with regard to asphalt concrete can have multiple
interpretations. The term has been applied to mean the overall deterioration of an asphalt pave-
ment from exposure to both climatological and load factors. In other cases, aging refers only to
the effects of environment, which include oxidative aging, ultraviolet radiation, thermal effect,
steric hardening, and moisture-related damage (Wright 1965). However, the most common
usage of the term aging, and the one used in this report, is to describe the process of asphalt
binder oxidation.

The issue of oxidative aging of asphalt binder has been recognized and studied for over a cen-
tury. Hubbard and Reeve (1913) published the results of a study that examined the effects of a
year of outdoor weathering on the physical (weight, hardness, etc.) and chemical (solubility)
properties of paving-grade asphalt cements. Subsequent studies have confirmed the basic findings
that oxidation, and not volatilization alone, is responsible for the changes in asphalt properties
that occur due to exposure (Thurston and Knowles 1936, Van Oort 1956, Corbett and Merz
1975), although volatilization may play a greater role in some of the binders produced today. Sig-
nificant research also has been conducted to investigate the chemical aspects of the aging process,
and excellent reviews of these studies can be found in the literature (Wright 1965, Lee and Huang
1973, Jemisson et al. 1992, Petersen 2009). In some of the cited studies, researchers have used
sophisticated experimental studies to propose conceptual, empirical, and/or analytical models
for the aging of asphalt binder (Lunsford 1994, Liu et al. 1996, Petersen and Harnsberger 1998,
Glover et al. 2008).

A review of the pertinent literature showed that although significant research has been devoted
to understanding and modeling the aging of asphalt binder, relatively little has been devoted to
the aging of asphalt mixtures (Bell 1989, Brown 2000, Airey 2003, Houston et al. 2005). The lack
of significant research in this area, particularly in the years since the conclusion of the Strategic
Highway Research Program (SHRP), reflects the complexities involved in studying mixture aging.
For example, when physicochemical interactions between asphalt binder and aggregate occur,
and when the material becomes structured and contains air voids, the aging process becomes
more complex than for binder alone. To date, no comprehensive study has been undertaken that
links the known behavior of an asphalt binder to that of an asphalt mixture, which is affected by
these and other interacting factors. Without such a study, and given the current state of under-
standing of the relationship between the properties of an asphalt binder and an asphalt mixture
that contains that asphalt binder, it is argued that the only way to determine the impact of aging
on mixture properties is direct experimentation of the mixtures (Bell and Sosnovske 1994, Bell
et al. 1994, Anderson et al. 1994).

3  

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

4   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

The accurate characterization of asphalt mixture properties in terms of the service life of a
pavement is becoming more important as more powerful pavement design and performance
prediction methods are implemented. One example of such characterization is shown in Fig-
ure 1. The results shown in this figure were obtained from the viscoelastic continuum damage
finite element program, FlexPAVE™, a software program that was developed under the FHWA-
sponsored Asphalt Mixture Performance-Related Specifications project (Kim et al. 2018).
The figure presents the damage contours within a 10-cm thick asphalt layer in a pavement
under California climatic conditions. This analysis employed the material properties obtained
from aging a mixture composed of SHRP AAD-1 asphalt binder and Federal Highway Adminis-
tration Accelerated Loading Facility­– (FHWA ALF–) graded aggregate in a forced-draft oven at
95°C for about 9 days. More details of the material properties can be found elsewhere (Qi et al.
2004). The effects of aging on the propensity of the pavement to exhibit top-down cracking can
be seen clearly by comparing Figure 1 (a) and (b); i.e., the damage intensity at the top of the
asphalt layer becomes much greater when the aged material properties are used in the analysis.
The accuracy of the aging model to represent the long-term aging of asphalt mixtures at different
depths of in-service pavements is critical to a mechanistic pavement analysis model’s ability to
predict top-down cracking.

Project Objectives and Scope


The objective of this project was to develop a calibrated and validated procedure to simulate
the long-term aging of asphalt mixtures for performance testing and prediction. The key prod-
ucts are a laboratory aging procedure and associated procedures that prescribe a set of laboratory
aging conditions to represent the long-term-aged state of asphalt mixtures in a pavement as a
function of climate and depth. The results of this project provide a basis for the future develop-
ment of a methodology that integrates the effects of long-term aging in Pavement ME Design.

0 1 0 1

1 0.9 1 0.9

2 0.8 2 0.8

3 0.7 3 0.7

4 0.6 4 0.6
Z (cm)
Z (cm)

5 0.5 5 0.5

6 0.4 6 0.4

7 0.3 7 0.3

8 0.2 8 0.2

9 0.1 9 0.1

10 10
-1.5 -1 -0.5 0 0.5 1 1.5 -1.5 -1 -0.5 0 0.5 1 1.5
X (m) X (m)

Short- Term Aged Loose Mix, 95°C, 8.9 Days

(a) (b)

Figure 1.   Damage contours from 10-cm thick asphalt pavements in California after 20 years of service:
(a) short-term aging only and (b) long-term aging.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Background  5  

In addition, this procedure would allow long-term aging to be considered in terms of full-scale
and accelerated pavement test results.
Future research is suggested to establish a diffusion model framework that accounts for differ-
ences in asphalt mixture morphology (e.g., air voids, film thickness) within the aging modeling
framework and that can be incorporated into mechanistic design and analysis methods. In addi-
tion, future work is suggested to calibrate the aging procedure for warm-mix asphalt (WMA)
and mixtures containing reclaimed asphalt pavement (RAP).

Previous Research into Long-Term Aging


of Asphalt Mixtures
Several asphalt mixture laboratory aging procedures have been tried. These procedures can
be broadly classified based on the (a) state of the material during aging (compacted specimen
versus loose mix), (b) pressure level (oven aging versus pressurized aging), and (c) aging tem-
perature. Thus, the discussion of the laboratory aging of asphalt mixtures in this section focuses
on these factors. In addition, a review of relevant aging models is presented. Key elements
required for developing an aging model include binder aging index properties (AIPs), oxida-
tion kinetics modeling, and diffusion modeling.

Compacted Specimen Aging Versus Loose Mixture Aging


The standard method used to assess the long-term aging of asphalt mixtures in the United
States is American Association of State Highway and Transportation Officials (AASHTO) R 30
(2002), which is meant to represent 5 to 10 years of aging in the field. However, two speci-
men integrity problems have been found using compacted specimen aging as recommended
in AASHTO R 30:
1. Distortion: Changes in air void content and geometry due to specimen softening and slump-
ing have been reported when using AASHTO R 30 (Reed 2010). To overcome this problem,
the NCHRP Project 9-23 protocol proposes wrapping specimens in metal wire mesh secured
with three clamps to prevent the samples from undergoing geometric distortion (Houston
et al. 2005). However, this approach has been reported only to reduce, but not eliminate,
specimen distortion during aging (Reed 2010).
2. Oxidation gradient: The NCHRP 9-23 project (Houston et al. 2005) demonstrated that the
long-term oven aging of compacted specimens leads to both radial and vertical oxidation
gradients in mixture specimens, which is a concern for the use of long-term oven aging of
compacted specimens in performance testing because properties differ throughout a speci-
men (Houston et al. 2005).
It is important to note that the objective of this study was to develop a long-term aging pro-
cedure that could be used in the fabrication of performance test specimens. A performance
test specimen should satisfy representative volume element requirements; i.e., it should have
uniform material properties throughout the specimen so that the performance measured from
the specimen can be related to a specific state of the material. An aging gradient would lead to
multiple dynamic modulus (|E*|) values within the specimen as well as to variable fatigue per-
formance throughout the specimen.
In addition, the aging gradient in a laboratory-aged specimen does not replicate field aging
gradients. In the field, aging is greatest at the surface and reduces with depth. However, in
laboratory-aged specimens, aging is greatest at the periphery and reduces towards the center of
the specimen. Hence, the direction of aging gradients differs between field cores and compacted
specimens prepared in the laboratory (Houston et al. 2005).

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

6   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Although the laboratory aging of loose (uncompacted) asphalt mixtures has received less
attention than compacted specimens, several studies recommend aging loose mixtures in the
laboratory to simulate the aging of asphalt pavements instead of aging compacted specimens
(Mollenhauer and Mouillet 2011, Van den Bergh 2011). However, most of these studies were
intended to prepare RAP materials rather than to investigate compaction and subsequent per-
formance testing (Partl et al. 2013, Mollenhauer and Mouillet 2011, Van den Bergh 2011).
Geometry distortion is not a concern with loose mix aging because the specimens are compacted
following aging. In addition, the aging gradient is not a problem, because the loose mix is aged
as a single layer of coated aggregate particles and, thus, oxygen and heat can circulate easily and
evenly throughout the mix. Also, the increased surface area of the binder film that is exposed to
oxygen is believed to accelerate aging in loose mixtures more so than in compacted specimens.
However, the compaction of aged loose mix for performance testing can lead to a potential spec-
imen integrity concern because aged binder is stiffer than unaged binder and thus is expected to
be less compactable than unaged material. Table 1 provides a summary of the advantages and
disadvantages of using loose and compacted asphalt mixture aging.

Oven Aging Versus Pressure Aging


Long-term oven aging is the most common method used to simulate the oxidative aging of
asphalt mixtures in the laboratory. As discussed, the current standard procedure, AASHTO
R 30, consists of conditioning compacted asphalt concrete specimens in an oven at 85°C for
5 days. However, other oven aging procedures have been tried and are cited in the literature.
Although these procedures are somewhat similar in terms of methodology, they differ in terms
of temperature and duration (e.g., Bell et al. 1994, Houston et al. 2005, Reed 2010).
The summaries of compacted and loose mixture aging trials (refer to Appendix A) dem­
onstrate that long-term oven aging requires considerable time to produce oxidation levels that
correspond to field conditions near the end of a pavement’s service life. Hence, researchers
have attempted to conduct long-term aging trials using both loose mix and compacted speci-
mens under air pressure to expedite oxidative aging (e.g., Kumar and Goetz 1977, Von Quintus
et al. 1992, Bell et al. 1994). Different pressure and temperature combinations were tried in these
studies. Generally, these earlier studies suggest that air/oxygen pressure expedites oxidative
aging. However, in the case of aging compacted specimens, the loss of integrity during aging is a
concern unless low temperatures are used, which will adversely affect the aging rate.
The standardized pressure aging equipment that is widely used in the asphalt industry is the
PAV. The PAV has been used to age both loose mix and compacted specimens. One of the

Table 1.   Comparison between loose mix and compacted specimens


in the aging procedure.

Homogenous aging in the mixture


Pros Higher oxidation rate than compacted mix
Maintaining specimen integrity a non-issue
Loose Mix Difficulties associated with compaction of aged loose mix, which limits
its use for producing specimens for performance testing
Cons
Limited amount of materials can be aged in a standard pressure aging
vessel (PAV) chamber
Can produce aged sample for performance tests if slumping is
Pros
minimized through use of wire mesh
Slower oxidation rate than loose mix
Compacted
Specimen Integrity of the specimens is compromised at high temperatures and
Cons pressures due to slump, cracking upon pressure release, and differences
in the coefficient of thermal expansion between binder and aggregate
Oxidation gradients exist radially and throughout height of the specimen

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Background  7  

main shortcomings of PAV aging compared to oven aging is the amount of material that can
be aged. Based on previous studies of loose mixture aging, in order to obtain uniform aging,
a uniform thin layer of loose mix should be placed in the PAV, which reduces the capacity of
the equipment to a very low level (around 1 kg) (Partl et al. 2013). Despite this capacity prob-
lem, however, it is believed that the air/oxygen pressure in the PAV can expedite aging and
reduce the time required to achieve any desired level of aging in an asphalt mixture specimen.
Furthermore, if a smaller geometry for asphalt mixture cores is used, more specimens can be
aged together. However, concern remains with respect to maintaining compacted specimen
integrity during pressure aging and upon pressure release, which needs to be considered care-
fully when assessing PAV aging of compacted mix specimens (Bell et al. 1994). Table 2 presents
a brief summary of the pros and cons of conducting oven versus pressurized aging.

Laboratory Aging Temperature


The oxidation of asphalt binder involves several independent and concurrent reactions. Tem-
perature can affect the rate of oxidation, the binder species that are oxidized, and the nature of
the oxidized species that are formed. Thus, temperature is a critical factor in laboratory simula-
tions of long-term aging. Increasing the aging temperature increases the rate of oxidation, which
is a desirable attribute. However, it can also disrupt polar molecular associations, which leads to
the oxidation of molecules that are inaccessible at lower temperatures. In addition, at ambient
temperatures, the oxidation of sulfides in asphalt binders leads to the formation of sulfoxides.
Elevated temperatures can deplete these sulfoxides through secondary oxidation reactions that
convert sulfoxides to sulfones. Thus, accelerated aging of asphalt binder at significantly high
temperatures may lead to a fundamentally different aged asphalt binder than asphalt aged in the
field (at a lower temperature) (Branthaver et al. 1993). The literature indicates that the disrup-
tion of polar molecular associations and sulfoxide decomposition become critical at tempera-
tures that exceed 100°C (Petersen 2009). Furthermore, aging at temperatures above 100°C can
lead to asphalt mastic drain-down from the loose mix because the low viscosity of the asphalt
binder at elevated temperatures.

Aging Index Properties (AIPs)


Properly quantifying the extent of aging for both laboratory- and field-aged binders is essen-
tial to this study. Asphalt binder is the asphalt mixture constituent that undergoes oxidative
aging. Hence, the oxidation of a pavement is best tracked using binder properties, because the
mixture is subjected to other factors, including mechanical degradation that is caused by traffic
loading, thermal-induced stress, and moisture, all of which could confound test results. Herein,
the binder properties that are used to measure the extent of oxidation are referred to as AIPs.

Table 2.   Comparison between oven and PAV aging methods.

Available and easy to perform and control


Pros
Large amount of material can be aged
Oven High variability among ovens, especially in terms of air drafting
Aging More time needed to age materials in the oven than in the PAV
Cons
Maintaining compacted specimen integrity is required, especially at high
temperatures
Pressure can expedite the aging process
Pros More reliable than oven aging due to less equipment variability between
laboratories
Pressure
Aging Due to limited capacity of the vessel, less material can be aged in each
aging cycle unless new device is developed
Cons
Integrity of compacted samples during and after testing is a major
concern

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

8   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

AIPs can be divided into two main categories: chemical functional groups and rheological
parameters. Chemical AIPs provide the most direct indicators of oxygen uptake and can be mea-
sured using Fourier transform infrared (FTIR) spectroscopy techniques. Also, different rheo-
logical properties have been introduced as aging indices (e.g., the crossover modulus, low shear
viscosity, etc.) (Petersen 2009, Underwood et al. 2010, Farrar et al. 2013). Rheological AIPs are
advantageous over chemical AIPs because they can be related directly to the mechanical proper-
ties of an asphalt mixture and are readily relatable to pavement distresses (Glaser et al. 2013a).
However, a comprehensive study to elucidate the rheological properties that most directly relate
to the chemical changes that result from oxidation is lacking in the literature.
Carbonyl and sulfoxide are the two major chemical functional groups that are produced
upon oxidation of an asphalt binder. The carbonyl chemical functional group has long been
used to indicate the level of oxidation within asphalt binder and is directly related to changes
in asphalt binder viscosity (Petersen 2009). The ketone functional group is the major compo-
nent of the carbonyl infrared absorption region. Ketone formation changes the polarity of the
associated aromatic ring components within an asphalt binder, which leads to an increase in
the asphaltene content, thereby increasing viscosity (Petersen 2009, Petersen and Glaser 2011).
The change in the sulfoxide fraction during aging has received less consideration in the past.
Petersen and Glaser (2011) conducted a thorough study to understand the role of sulfoxide
formation on physical properties during oxidative age hardening in asphalt binders. By evaluat-
ing the relationship between the sum of the absorbances of the ketones and sulfoxides and the
logarithm of viscosity, Peterson and Glaser (2011) showed that the alcohols that are produced
during oxidative aging have a significant impact on viscosity, especially for asphalt with a high
sulfur content.
FTIR-based AIPs have been represented by both absorbance peaks (e.g., Petersen 2009,
Petersen and Glaser 2011) and areas (e.g., Han, 2011, Jin et al. 2011, Lamontage et al. 2001,
Zhang et al. 2011). Peaks are determined directly by evaluating absorbance at specific wave
numbers. In this study, changes in the carbonyl and sulfoxide peaks were tracked at wave num-
bers 1702 cm–1 and 1032 cm–1, respectively. Areas are calculated by determining the area under
the FTIR spectrum that lies between specific wave numbers. In this study, the carbonyl area,
carbonyl + sulfoxide (C + S) area, and C + S peaks are used in comparisons. Although numerous
AIPs have been proposed in the literature, a comprehensive study to elucidate the chemical and
rheological AIPs that are the best for tracking oxidation is not available.

Modeling of Oxidative Aging


To prepare an aged compacted specimen in the lab that represents specific time in service,
climate, depth within the pavement, and level of air voids, a solid understanding of binder
oxidation kinetics and diffusion is required. The Global Aging System (GAS) model (Mirza
and Witczak 1995) is an empirical model that allows for the prediction of the change in binder
viscosity as a function of age, given the mean annual air temperature (MAAT); the model also
considers the gradient of aging with pavement depth. The GAS model assumes a hyperbolic
aging function that predicts a decreasing rate of change of viscosity with an increase in age, under
the assumption that most age hardening occurs within the first 10 years of a pavement’s service
life. The simplicity of the GAS model makes it an attractive option. However, the GAS model has
been criticized because it does not account for aging deeper than 1.5 in. below the pavement
surface (Mirza and Witczak 1995, Prapaitrakul 2009). In addition, the model does not directly
account for differences in asphalt binder kinetics among binder types. Also, the method used to
account for air void content is based on a relatively small set of conditions, which is the reason
this adjustment factor is considered optional (Mirza and Witczak 1995).

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Background  9  

A preliminary fundamentals-based, one-dimensional combined asphalt oxidation kinetics


and diffusion model was developed at Texas A&M University. This model is referred to as the
transport model (Lunsford 1994, Prapaitrakul 2009, Han 2011). The transport model includes
several sub-models to predict pavement aging, including 1) oxygen diffusion/flow from the atmo-
sphere into the interconnected air voids in a pavement, 2) oxygen diffusion through the asphalt-
aggregate matrix, 3) heat transfer, and 4) oxidation kinetics. Implementation of the transport
model currently requires detailed information about air void content and distribution as well as
the temperature gradient throughout the pavement depth at a higher level of accuracy than can
be provided by available databases, including the Enhanced Integrated Climatic Model (EICM).
Although the oxygen percolation and diffusion components of the transport model are still
under development, oxidation kinetics modeling has received significant attention in the past.
All asphalt materials exhibit relatively similar kinetics, which can be described as trends in AIPs
that are indicated by an initial fast reaction period, also known as the spurt, followed by a slower
reaction period that has an approximately constant rate (Petersen et al. 1996, Petersen 1998,
Petersen et al. 2011, Prapaitrakul 2009, Han 2011). Asphalt binder oxidation kinetics has been
modeled successfully using the Arrhenius expression of temperature and pressure dependency
(Petersen 2009, Prapaitrakul 2009, Glaser et al. 2013b).
The Arrhenius equation, presented as Equation 1 in this report, describes the temperature
dependency of the oxidation rate. Herrington et al. (1994) proposed Equation 2 through Equa-
tion 4 for modeling the oxidation kinetics of asphalt binder using viscosity as an AIP. Herrington
et al.’s (1994) kinetics model considers both the initial fast reaction period (i.e., the spurt) and
the constant rate period. Researchers at Texas A&M University also conducted a series of proj-
ects that led to a kinetics model, as expressed in Equation 3 through Equation 8, using similar
principles and the same Arrhenius equation as Herrington et al. (1994) (Lau et al. 1992, Davison
et al. 1994, Liu et al. 1996, Domke et al. 2000, Glover et al. 2014). In the Texas A&M University
kinetics model, the carbonyl area is used as the AIP. In addition, Glover et al. (2014) presented
an interrelationship among kinetics parameters to describe the fast and constant rate reactions.
Glover et al. (2005) developed interrelationships between Arrhenius parameters of fast and con-
stant rates, as shown in Equation 6 through Equation 8. This simplification reduced the number
of unknowns to only two parameters in addition to the short-term-aged binder, CA.
In addition, Glaser et al. (2013b) used stoichiometry to derive the kinetics model given in
Equation 9 using C + S absorbance peaks as an AIP. Their kinetics model has only one adjust-
able parameter besides short-term-aged binder aging level, which implies that the kinetics model
can be calibrated using isothermal aging. They validated their model using 12 asphalt binders
from a wide variety of sources. Equation 3 and Equation 4 are applied along with Glaser et al.’s
(2013b) model to describe the temperature dependency of the oxidation reaction. Glaser et al.
(2013b) proposed that the kinetics of asphalt binder can be modeled using a single material-
dependent parameter that corresponds to the quantity of reactive material within an asphalt, M,
and proposed that the Arrhenius parameters, kf and kc, are the same for all binders. Equation 10
shows the form of Glaser et al.’s (2013b) model, which incorporates the pressure dependency of
the reaction rate. If the Arrhenius parameters can be applied universally, as Glaser et al.’s results
suggest, the measurement of the AIPs that correspond to various durations of aging at a single
temperature is all that is needed to characterize the kinetics model for a given binder (Glaser
et al. 2013b, 2015).

 − Ea 
k = AP α exp  (1)
 RT 

log η = log ηo + M (1 − exp ( −k f t )) + kc t (2)

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

10   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

 − Eaf 
k f = A f exp  (3)
 RT 

 − Eac 
kc = Ac exp  (4)
 RT 

CA = CAo + M (1 − exp ( −k f t )) + kc t (5)

Eaf = 0.85Eac − 10.4 (6)

A f = 0.52exp ( 0.3328 Eaf ) (7)

Ac = 0.0266exp ( 0.3347 Eac ) (8)

 kc 
(C + S ) = (C + S )o + M  1 − (1 − exp ( −k f t )) + kc Mt (9)
 k f 

 kc . P n 
(C + S ) = (C + S )o + M  1 − (1 − exp ( −k f P mt )) + kc MP nt (10)
 k f . P m 

where
k= rate of reaction,
kf = rate of fast reaction,
kc = rate of constant reaction,
A= frequency (pre-exponential) factor,
Af = fast reaction frequency factor (s–1),
Ac = constant reaction frequency factor (s–1),
Eaf = fast reaction activation energy (kJ/mol),
Ea = activation energy,
Eac = constant reaction activation energy (kJ/mol),
P= absolute oxygen pressure,
α= reaction order with respect to oxygen pressure,
m= reaction order of fast reaction,
n= reaction order of constant reaction,
R= universal gas constant, or ideal gas constant (kJ/mol K), •

t= reaction time (s),


T= reaction temperature (Kelvin),
M= fitting parameter related to fast reaction reactive material,
h= long-term aged binder viscosity,
ho = short-term aged viscosity,
CA = long-term aged binder carbonyl area,
CAo = short-term aged binder carbonyl area,
(C + S) = long-term aged binder C+S absorbance peaks, and
(C + S)o = short-term aged binder C+S absorbance peaks.
Although significant research efforts have been dedicated to modeling the kinetics of asphalt
binder aging, relatively little attention has been devoted to modeling the kinetics of asphalt
binders within asphalt mixtures. The lack of major studies in this area reflects the complexities
involved in studying mixture aging. Oxidation reaction rates and mechanisms are affected by

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Background  11  

the complex physicochemical interactions between the asphalt binder and the aggregate
(Moraes and Bahia 2015, Wu et al. 2014, Petersen 2009, Little et al. 2006, Recasens et al. 2005,
Huang et al. 2002, Jones 1997, Petersen et al. 1987).
The kinetics modeling used in this study applied Glaser et al.’s (2013b) work using a rheology-
based AIP of loose mix aging data to account for the effect of mineral filler on aging rates. The
kinetics model outputs can be calibrated based on concepts related to diffusion from the trans-
port model in an effort to incorporate more fundamental behaviors than the GAS model and
still be less cumbersome than the transport model.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

CHAPTER 2

Research Approach

Overview of Research Approach


This section briefly explains the research approach. Figure 2 presents a flow chart to describe
the research plan. The subsequent sections explain the components within the flow chart. The
suggested future research topics shown in the figure are described in Chapter 4.
The individual tasks included in the overall research approach are as follows:
1. AIP selection to track the oxidation levels of laboratory-aged mixtures and field cores.
2. Sensitivity study to estimate the sensitivity of the mechanical properties of asphalt concrete
to asphalt binder oxidation.
3. Selection of long-term aging method.
4. Determination of project-specific aging durations by matching the AIPs measured from
laboratory-aged loose mixtures with those from field cores.
5. Climate-based determination of predefined aging durations to determine the required labo-
ratory aging duration to match the field aging at any United States location of interest and
depth of interest using EICM hourly pavement temperature data.
6. Development of pavement aging model by calibrating the rheology-based kinetics model
against the AIPs measured from different depths of field cores using temperature profiles
obtained from the EICM.

AIP Selection
The development of the long-term aging procedure and kinetics model hinged on the com-
parison of key AIPs of laboratory-aged binders and binders extracted and recovered from field
cores. Therefore, an evaluation of candidate chemical and rheological AIPs was conducted in
order to select AIPs to track the oxidation levels of field- and laboratory-aged materials when
developing the long-term aging procedure and associated kinetics model. A wide range of
laboratory- and field-aged materials were used to evaluate and subsequently select the AIPs.
The chemical AIPs evaluated include the carbonyl absorbance area, C + S absorbance area, and
C + S absorbance peaks determined using attenuated total reflectance (ATR) FTIR. The chemi-
cal AIPs were evaluated based on their correlation to the duration of the laboratory aging. The
rheological AIPs evaluated included the dynamic shear modulus, zero shear viscosity, Glover-
Rowe (G-R) parameter, and crossover modulus. The rheological AIPs were evaluated based on
the strength of their relationship to the chemical changes that were induced by oxidation.

Sensitivity Study
The goal of the sensitivity study was to estimate the sensitivity of the mechanical properties of
asphalt concrete to asphalt binder oxidation. The sensitivity study provided thresholds by which

12

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Research Approach   13  

Figure 2.   Research approach flow chart.

to evaluate the significance of observed differences in asphalt binder AIPs in terms of asphalt
mixture performance. To accomplish this goal, experimental characterization was performed
at multiple length scales, i.e., asphalt binder, asphalt mastic, and asphalt fine aggregate matrix
(FAM). Here, asphalt mastic refers only to the portion of the asphalt mixture that includes
the asphalt binder and the filler (i.e., particles finer than 75 mm). FAM includes portions of
asphalt mastic, fine aggregate (< 0.6 mm), and some of the air voids. Testing was performed
using (a) asphalt binder to establish baseline properties and evaluate the degree of oxidation,
(b) asphalt mastic to consider physicochemical aspects, and (c) FAM to consider air voids and
aggregate inter­action effects.
For all the tests, the asphalt binder was first oxidized using standard laboratory aging proce-
dures [rolling thin film oven (RTFO) and PAV]. The aging process was undertaken to create
test materials that replicated, to the degree possible, asphalt binder that exists in asphalt pave-
ments of relatively young, medium, and old ages. For this purpose, the relationships among
PAV aging time, MAAT, and in-service aging time that was developed during NCHRP Proj-
ect 9-23 were utilized (Houston et al. 2005). Correlations between the MAAT and asphalt
binder grade were established to utilize this function, and the fact that these aging conditions
affected the rheology of the asphalt was verified by determining the performance grade (PG) of
the aged asphalt.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

14   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

After oxidizing the asphalt, it was either tested directly, blended with filler particles to create
mastic, which was then tested, or blended with filler and fine aggregate to create FAM, which
was then tested. The testing consisted of temperature and frequency sweep tests to establish the
dynamic modulus values of the materials and time sweep tests to establish the fatigue properties
of the materials. Subsequently, these experiments were analyzed to determine the sensitivity of
the material responses to binder oxidation.

Selection of the Proposed Aging Method


An experimental program was executed to select the proposed aging method, in which the
following factors were evaluated: (a) state of the material during aging (compacted speci-
men versus loose mix), (b) pressure level (oven aging versus pressurized aging), and (c) aging
temperature (95°C versus 135°C). The integrity of the specimens following aging, the rate of
oxidation quantified using the AIPs of the extracted binder, versatility, ability to mimic field
oxidation reactions, and the cost of the various procedures were compared in order to select
the most promising aging procedure. The aforementioned analysis was conducted using hot
mix asphalt (HMA) mixtures. However, a complementary analysis of the long-term aging of
WMA mixtures also was conducted. To assess the aging level achieved during the aging trials,
comparisons were made between the AIPs of the binder extracted and recovered from long-
term laboratory-aged mixtures, binder aged using the standard RTFO and PAV, and binder
extracted and recovered from field cores acquired from in-service pavements.

Loose Mix Versus Compacted Specimen Aging


Both compacted and loose mixture aging trials were conducted to determine the optimal
state of the material to use for long-term aging. The specimen integrity following laboratory
aging and the efficiency of oxidation were used to evaluate the state of the material.
For the compacted specimens, two geometries were considered: standard 100-mm diameter
specimens and small-specimen geometry 38-mm diameter specimens. The motivation behind
the use of small specimens was to reduce the diffusion path. The primary concerns associated
with the laboratory aging of compacted specimens are slump, changes in air void content, and
the existence of an oxidation gradient. Therefore, the dimensions and air void contents of
compacted specimens were compared before and after aging to determine if the specimens
were damaged during the laboratory aging process. In addition, differences in the rheology
and chemical compositions, along with the distance from the specimen periphery, were used
to detect the presence (if any) of an aging gradient.
For the loose mixtures, the primary specimen integrity concern was compactability. There-
fore, the number of gyrations required to meet the target air void content was compared between
the short- and long-term aged mixtures in order to assess compactability. In addition, air void
content measurements were used to verify that the desired compaction level had been met.
Dynamic modulus and cyclic fatigue tests were conducted using short- and long-term aged
specimens to further assess if specimen integrity had been compromised as a result of the long-
term aging procedures applied to both the loose mixtures and compacted specimens. The AIPs
of the binder extracted and recovered from the long-term aged materials were used to assess the
relative efficiency of the loose and compacted specimen aging procedures.

Oven Versus Pressure Aging


The application of pressure to expedite the long-term aging of the loose mixtures and com-
pacted specimens was evaluated by conducting aging trials in both an oven and a binder PAV.
The ability of the PAV to improve the efficiency of laboratory aging was assessed through

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Research Approach   15  

comparisons of extracted and recovered asphalt binder AIPs following long-term aging.
Damage induced by the application and release of pressure to the compacted specimens also
was assessed.

Aging Temperature, 95çC Versus 135çC Aging


Several researchers have proposed the oven aging of loose (uncompacted) asphalt mix-
tures at 135°C for efficient laboratory long-term aging (e.g., Braham et al. 2009, Dukatz 2015,
Blankenship 2005). However, the literature indicates that the oxidation reaction mechanism
can change when the temperature exceeds 100°C, thus suggesting that accelerated aging at
135°C may lead to a fundamentally different aged asphalt binder than asphalt aged in the field
(at a lower temperature). Therefore, the performance implications of aging laboratory loose
mixtures at 135°C were evaluated by comparing the dynamic modulus values and the cyclic
fatigue performance of mixtures subjected to long-term aging at 95°C and 135°C to yield the
same rheology. In this study, 95°C was selected instead of 100°C in order to avoid the aging
temperature reaching close to 100°C due to possible temperature fluctuations in the oven.
Although the rheology of the mixtures aged at 135°C and 95°C matched, their chemistry dif-
fered, and thus, the experiments allowed for the assessment of the significance of the chemical
differences that are caused by aging at 135°C. The results then were used to inform the selection
of the laboratory aging temperature.

Determination of Project-Specific Aging Durations


After selecting the most promising aging method, the aging procedure was applied to some
selected component materials for a prolonged duration. Samples were removed periodically
and subjected to extraction and recovery, after which the binder AIPs were measured and used
to derive the oxidation kinetics. In addition, binders were extracted and recovered from vary-
ing depths of a selected group of field cores obtained from in-service pavements. The AIPs of
the field-aged binders were measured and compared to the laboratory-aged oxidation rates to
determine the laboratory aging duration that is required to match the target field AIPs for spe-
cific projects.

Climate-Based Determination of Predefined Aging Durations


The project-specific aging durations were used to calibrate a kinetics-derived climatic aging
index (CAI) that can be used to determine the required laboratory aging duration to match
the field aging at any location and depth of interest using EICM hourly pavement temperature
data. The CAI, which can be used to relate the pavement temperature history to the required
laboratory aging duration, was derived using a simplification of a rigorous oxidation kinetics
model that retains the exponential relationship between the oxidation rate and temperature.
A diffusion correction factor is included within the CAI to allow the required laboratory aging
duration to match different pavement depths of interest. The CAI analysis was used to gener-
ate maps of the United States that allow the visual determination of the required laboratory
aging durations that match 4, 8, and 16 years of field aging at depths of 6 mm and 20 mm
from the surface.

Development of Pavement Aging Model


Kinetics Modeling of Field Aging Using Mix-Specific Kinetics Parameters
An existing kinetics model was applied successfully to predict loose mix aging rates at dif-
ferent temperatures using rheology-based AIPs. The kinetics model was then validated using

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

16   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

a separate set of mixtures. The validated kinetics model provides a basis for the future devel-
opment of a methodology that integrates the effects of long-term aging on performance in
Pavement ME Design and other mechanistic design and analysis methods. Field aging levels
were measured at different depths from field cores obtained at different service lives. Pavement
temperatures obtained from the EICM as a function of depth were used along with the devel-
oped kinetics model to predict field aging levels as a function of pavement depth. The predicted
aging levels were compared against measured field aging in order to calibrate the predictive
model. The calibrated kinetics model can be coupled with a diffusion model to enhance the
prediction capabilities of that model to account for the effects of mixture morpho­logical proper-
ties on field aging.

Determination of Mix-Specific Kinetics Parameter from Universal


Simple Aging Test (USAT) Binder Aging
Aging loose mixture in the oven allows the physicochemical effects of filler on asphalt binder
oxidation rates to be captured. Thus, this aging method is assumed to provide a good representa-
tion of field aging. However, loose mix aging requires the cumbersome extraction and recovery
of the binder from the aged mixture prior to the AIP evaluation. Development of a model that
can predict the effect of filler on asphalt binder oxidation would negate the need to perform
loose mixture aging in order to relate laboratory aging to field aging levels. In order to evalu-
ate the idea of obtaining the binder aging rate from USAT (Farrar et al. 2015) and relating it
to the loose mix aging rate at a given temperature, binder and loose mix samples were short-
term aged at 135°C for 4 hours followed by long-term aging in the same oven at 95°C. Then,
both aged binders from the USAT as well as extracted and recovered binders from loose mix
aging were tested using a dynamic shear rheometer (DSR) and FTIR spectroscopy. The aging
rates obtained from this binder aging and loose mix aging were compared in order to find a
relationship between the obtained aging rates.

Test Materials and Field Projects


The experimental plan involves two groups of materials. The Group A materials are a set of
laboratory-prepared materials for which relatively large quantities were available. The main
purpose of investigating these materials was to evaluate different long-term aging scenarios.
Also, the Group A materials allowed for a systematic sensitivity study to investigate the effects
of changes in the AIPs on the performance of asphalt concrete. The Group B materials are
the original component materials (i.e., binder and aggregate) and field cores extracted from
in-service pavements. The Group B materials were used to develop the interim long-term
aging procedure.

Group A Materials
Table 3 presents a summary of the Group A materials. Several binders are included to cover
a wide range of aging characteristics. The limestone aggregate and PG 58-28 and PG 76-16
binders used in the Asphalt Research Consortium (ARC) study were used in the sensitivity

Table 3.   Summary of Group A materials.


Material
Material Source
Aggregate Asphalt Binder
ARC Limestone PG 58-28, PG 76-16
North Carolina Granite PG 64-22
SHRP Granite PG 58-28 (AAD-1), PG 58-10 (AAG-1)

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Research Approach   17  

Table 4.   Selected sections for development of long-term aging


protocol and field calibration under Group B materials.

Binder/ Date Date Core


Site ID Location
Modification Built Extracted
Control, slices,
and aged binders
FHWA ALF Virginia 2001 2013
styrene butadiene
styrene (SBS)-LG
Control, Foam
Manitoba Manitoba, Canada WMA, Evotherm 2010 2014
WMA
National Center of
Asphalt Control, Foamed
Alabama 2009 2013
Technology WMA
(NCAT)
New Mexico, Grant County, Asphalt Cement
1996 2006, 2014
I-10 Frontage Rd., MP 51 AC-20
South Dakota, Campbell Asphalt Cement
1993 2006, 2014
County, 101st St., MP 400.1 120-150 pen
LTPP a Texas, Brazos County, Old
Asphalt Cement
Cameron Ranch Rd., MP 1996 2007, 2014
AC-20
404.2
Wisconsin, Marathon
unknown 1997 2005, 2014
County, Apple Ln.
WesTrack Fine
1995, 1997,
Sections (1, 2, 3, Dayton, Nevada PG 64-22 1995
1999, 2014
4, 14, 17, 18)
WesTrack Coarse
Dayton, Nevada PG 64-22 1997 1999, 2014
Sections (36, 39)
a
LTPP stands for long-term pavement performance.

study. The PG 64-22 binder and granite aggregate were acquired from North Carolina and used
in the sensitivity study and for the evaluation of candidate aging methods. The SHRP binders
AAD-1 and AAG-1 were used to study the effects of aging temperature due to their known dif-
ferences in chemistry.

Group B Materials/Projects
Table 4 summarizes the Group B materials that are composed of the original binders and
aggregate and field cores from the selected pavement sections. These sections cover a wide range
of pavement design, climatic conditions, ages, binder and aggregate characteristics, air void
contents, asphalt contents, and gradations. The Group B materials were used to develop the
long-term aging procedure and for validation and calibration.
Figure 3 summarizes the geographic coverage of the selected materials/projects. The figure
indicates that a broad range of geographic locations across the United States and Canada were
used to develop the long-term aging procedure.

Sample Preparation Methods


Asphalt Mastic Preparation
The composition of the mastics was determined based on the asphalt mixture design for each
source. Also, the specific gravity of each aggregate source in the gradation was measured. Trials
were conducted to determine the optimum binder content for the mastics. Asphalt binders aged
in the PAV for different durations were mixed with the prepared gradations and mixed to form
a uniform mix. Samples were taken from the mastic to test in the DSR.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

18   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Figure 3.   Locations of materials/projects included in the study


(MB = Manitoba).

FAM Preparation
The compositional design was developed for FAM based on the mixture design for each
source. The maximum specific gravity, Gmm, was calculated at the aging level of 0 year. The
aged binder for the FAM samples was pre-aged in the PAV and then blended with aggregate in
a conventional bucket mixer. The blended FAM mix was poured into a compaction mold and
kept inside the oven until it reached the compaction temperature. After reaching the correct
temperature, FAM plugs were compacted in the Superpave® gyratory compactor and then left
at room temperature to cool. The Ø20-mm × 50-mm test specimens were extracted by coring
the Superpave gyratory plugs. The target FAM air void contents correspond to mixture air
void contents of 4.5% and 8.5% were 6.5% and 11%, respectively. Figure 4(a) and 4(b) show a
FAM plug with a cored sample and the test set-up, respectively.

(a) (b)

Figure 4.   (a) FAM plug with cored and cut sample and (b) instrumented FAM sample.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Research Approach   19  

Asphalt Mixture Aging


All of the asphalt mixtures aged in the laboratory were prepared using component materials
that were used in constructing the pavements from which field cores were obtained. All of the
component materials were prepared based on the original mix design. All HMA mixtures were
subjected to short-term aging at 135°C for 4 hours in accordance with AASHTO R 30 prior to
long-term aging. The WMA mixtures were subjected to short-term aging at 117°C for 2 hours
prior to long-term aging.

Aging of Compacted Specimens


For all the compacted specimen aging trials, the short-term aged mixtures were compacted
in a Superpave gyratory compacter to fabricate Ø150-mm × 178-mm specimens. Subsequently,
large specimens for aging were prepared through coring to obtain Ø100-mm × 178-mm speci-
mens. To fabricate small specimens, Ø38-mm cylindrical specimens were horizontally cored
from initial gyratory specimens (Ø150-mm × 178-mm). To fabricate the specimens for the
WMA aging trials, the specimens were cored to obtain Ø100-mm × 178-mm specimens. Then,
the inner 150 mm of the Ø100-mm specimens were sliced to obtain 25-mm thick disks.
For the oven aging of the large compacted specimens, the ends of the cores were not sawn
before aging because of the high probability of an aging gradient that could affect the per-
formance test results. Also, wire mesh supports were utilized to minimize distortion under
self-weight. A single aging temperature of 85°C, as specified by AASHTO R 30, was used for
the compacted specimen aging trials. For the small specimens, the ends were sawn to obtain
Ø38-mm × 100-mm specimens. Figure 5 shows large and small specimens aging in the oven.
For the WMA compacted specimen aging trials, thin disk specimens were placed on end in the
oven, as shown in Figure 6, to maximize oxygen exposure.
Pressure aging trials using the compacted specimens were conducted in a standard asphalt
binder PAV. The pressure level of 300 kPa was selected for the PAV aging trials because SHRP
researchers (Bell et al. 1994a) had found that higher pressure levels damage specimens upon
pressure release. To reduce the stress on the large specimens under self-weight in the PAV, a wire
mesh hammock-like support was developed to allow the specimen’s weight to be distributed
over a larger area on its sides than if it were positioned vertically, as shown in Figure 7.

Pressure Aging of Loose Mix


The oven aging of the loose mix was accomplished by separating the mix into several pans
such that each pan had a relatively thin layer of loose mix that was approximately equal to the

Figure 5.   Long-term aging of large and small


specimens in the oven.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

20   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Figure 6.   WMA compacted specimens aging


in an oven.

nominal maximum aggregate size (NMAS) of the aged mix, as shown in Figure 8. The loose
mix was agitated several times during oven aging, and the pans in the oven were rotated sys-
tematically to minimize any effects of an oven temperature gradient and/or draft on the degree
of aging. After long-term aging, the materials were taken out of the oven and mixed together in
order to obtain a uniform mixture, and then the mixture was left to cool to room temperature.
The loose mixture was then reheated to the compaction temperature for 75 minutes. Speci-
mens were compacted following aging for performance testing. Because the compactability of
the aged loose mix was a potential concern, the effect of increasing the compaction tempera-
ture on the compactability of the aged loose mix was investigated.
The pressure aging trials for the loose mix utilized the standard binder PAV pressure of
2.1 MPa, because pressure was not anticipated to induce any integrity concerns regarding the
loose mixture samples, as compaction follows the aging process. Figure 9 shows the specimen

Figure 7.   Simple set-up for holding large specimens during aging in the PAV.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Research Approach   21  

Figure 8.   Long-term aging of loose mix in thin layers


for long-term aging in oven.

set-up. The loose mix was dispersed in thin layers, consistent with the process for oven aging.
The size of the binder PAV prohibited aging a large quantity of mix efficiently, and thus, any
gain in the oxidation rate had to be balanced with the amount of material that could be aged
at one time or, conversely, the associated costs of developing a mixture-specific pressure aging
device. Due to the capacity constraints of the PAV, the long-term aging trials of the loose mix
in the PAV were limited to simply assessing how much pressure would expedite the oxida-
tion of the loose mix. Insufficient quantities of material were aged to produce compacted
specimens.

Asphalt Binder Aging


RTFO Aging
RTFO aging was conducted using selected original asphalt binder samples according to
AASHTO T 240 to simulate short-term aging.

Figure 9.   Aging rack developed for long-term aging of loose mix in PAV.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

22   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

PAV
Asphalt binder residue obtained from the RTFO aging was subjected to PAV aging based on
AASHTO R 28 at 100°C for 20 hours.

USAT
The USAT, developed by Farrar et al. (2015) at the Western Research Institute (WRI), was
applied in this study to derive the oxidation kinetics of asphalt binders as part of the aging model
development. In the USAT, the binder is placed in grooved plates to achieve a film thickness of
300 micrometers. The USAT plates were placed in an oven at 135°C for 4 hours to simulate the
short-term aging of loose mixtures. After this binder short-term aging process, the USAT plates
were placed in an oven at 95°C to simulate long-term aging.

Field Core Preparation


Full depth cores were acquired from in-service pavements and from the FHWA Materials
Reference Library. The acquired field cores were wrapped with plastic wrap and placed in a
temperature-controlled room to minimize further aging during storage. Following storage, the
upper 2 inches of each field core was sliced into 0.5-inch thick sections. The remainder of the
field core was sliced into 1-inch thick slices. Special attention was given to avoid the tack coat
and prime coat layers when slicing the field cores. Figure 10 provides a depiction of the field
core slicing pattern used.

Micro-Extraction and Recovery


The micro-extraction and recovery of the asphalt binder from the asphalt mixtures was per-
formed following the procedure proposed by Farrar et al. (2015) at the WRI. This procedure
uses a solvent mixture of toluene and ethanol (85:15) for extraction and recovery. The mixture
sample size is limited to 200 g to produce approximately 10 g of asphalt binder per extraction,
which is adequate for both FTIR spectrometry testing and DSR testing. In order to prevent
further aging of the binder samples during the extraction and recovery procedure, a distillation
flask was subjected to vacuum pressure of 80.0 ± 0.7 KPa (600 ± 5 mm Hg) under nitrogen gas.
FTIR spectrometry testing was conducted following extraction and recovery to ensure that no
detectable solvent was present.

Figure 10.   Depiction of field core slices used to determine


oxidation gradient with depth.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Research Approach   23  

Test Methods
Asphalt Binders
FTIR Test Procedure
ATR FTIR spectroscopy collects absorbance data within a wide spectral range (400 cm-1 to
4000 cm-1). The ATR spectra were collected using 64 scans at a resolution of 4 cm-1 using a
minimum of two replicates. For each binder, all replicates under different conditions were
normalized to the same absorbance value at wave number 1375 cm-1. This wave number was
selected for normalization because absorbance at this wave number is not affected by the level
of oxidation. Changes in the C + S peaks were tracked at wave numbers 1702 cm-1 and 1032 cm-1,
respectively. Additionally, C + S areas were measured as the areas under the FTIR absorbance
curves at wave number ranges 1650–1820 cm-1 and 1000–1050 cm-1, respectively. The trap-
ezoidal rule was used to numerically determine the area under the band between the specified
ranges of wave number.

DSR Test Procedure


Frequency sweep testing was conducted at frequencies ranging from 0.1 Hz to 30 Hz and
multiple temperatures (5°C, 20°C, 35°C, 50°C, and/or 64°C) using asphalt binders in the DSR
with 8-mm parallel plate geometry. A strain amplitude of 1% was applied at all frequencies
and temperatures of testing. The rheological properties analyzed included the dynamic shear
modulus (G*) at 64°C and 10 rad/s, the crossover modulus (G c*) [Farrar et al. 2013, defined as
the G* value that corresponds to the reduced frequency where the storage modulus (G′) and
loss modulus (G″) master curves cross (i.e., where the phase angle equals 45°)], the zero shear
viscosity (ZSV), defined as the viscosity when the shear rate approaches zero (Binard et al.
2004, Brio et al. 2009), and the Glover-Rowe (G-R) parameter (Rowe et al. 2014), which has
been proposed as an indicator of ductility and is equal to G*cos2d/sind evaluated at 15°C and
0.005 rad/s.

Asphalt Mixtures
Dynamic Modulus Test Procedure
Frequency sweep tests were conducted at multiple temperatures in accordance with AASHTO
PP 342 to build dynamic modulus master curves. The initial test temperatures used to build
these master curves were -10°C, 5°C, 20°C, 40°C, and 54°C. However, the dynamic modulus
test results from the first set of specimens revealed insufficient overlap between the dynamic
modulus values at the different test temperatures for highly aged materials. This lack of overlap
in the dynamic modulus values precluded an accurate application of time–temperature super-
position in order to construct the master curves. Therefore, to enable the successful construction
of dynamic modulus master curves, most of the long-term aged mixture dynamic modulus tests
were conducted at -10°C, 5°C, 15°C, 27°C, 40°C, and 54°C. At least two test replicates were
conducted for each mixture and condition evaluated.

Asphalt Mixture Performance Tester (AMPT) Cyclic Fatigue Test Procedure


Cyclic fatigue testing was conducted in accordance with AASHTO TP 107. The test tem-
perature was determined using either the binder PG or designated regional PG estimated from
the LTPPBind program as per AASHTO TP 107. The testing frequency was 10 Hz. Three tests
were conducted for each mixture at three different actuator displacement amplitudes (low,
medium, and high).

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

CHAPTER 3

Findings and Applications

Findings
Sensitivity Study
Temperature–frequency sweep tests were first performed to determine the modulus values
of the materials and subsequently the aging ratio (AR). As represented in Equation 11, AR
is the mathematical ratio of |G*| or dynamic modulus, |E*|, after aging to the value before
aging. It is calculated at multiple temperatures (Tj) and frequencies (wi) to gain a com-
plete picture of the impacts of oxidation. For comparison across different study materials, a
singular frequency was required and 10 rad/s was chosen for this purpose. The ARs were
used to assess the sensitivity of both the mastics and FAM to their corresponding binder
oxidation levels.

 G * ( ω i , Tj )after aging 
 mastic and binder
 G * ( ω i , Tj )before aging 
AR =  (11)
 E * ( ω i , Tj )after aging 
 FAM
 E * ( ω i , Tj )before aging 

The aging parameter (AP), described as the change in the parameter of interest with time, was
also used to represent the oxidative aging of the mastic and FAM samples. Equation 12 shows an
example of the AP calculated based on G*c from the unaged case, e.g., at t = 0, G*c0, and crossover
modulus at different oxidation times, G*ct. The kt represents the aging rate of the material.

1 1
AP = kt = − (12)
G ct* G c*0

The approach adopted here to evaluate sensitivity is based on the concept of the crossover
modulus and the principles of second-order rate kinetics of binder oxidation. Because a direct
analysis of the AR and binder AP did not provide a clear understanding of the quantification
of the differences, a more involved methodology was followed. In the first application of this
approach, the binder AP accuracy required to match varying levels of accuracy (1% to 20%) for
the binder, mastic, and FAM ARs was estimated. In the second application, the errors in the AR
when the AP was matched at 1% to 20% were evaluated.
The primary conclusions drawn from these two analyses are summarized in Table 5 and
Figure 11, which indicate the accuracy levels that are required for the binder AP to match the
binder, mastic, and FAM AR at 1%, 10%, and 20% levels of accuracy. It should be noted that,
based on recent advances reported in studies that compare the mechanical properties of FAM

24

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   25  

Table 5.   Summary of sensitivity study findings.


Range of Required AP Accuracy
Error in AR
Binder Mastic FAM
1% 1.1% – 1.9% 1.1% – 1.9% 1.5% – 3.6%
10% 11.0% – 18.9% 11.1% – 19.5% 14.8% – 35.6%
20% 22.0% – 37.7% 22.2% – 39.1% 29.6% – 71.2%
Range of Required AP Accuracy
Error in AR
Binder Mastic FAM
1% 0.5% – 0.9% 1.1% – 1.9% 1.5% – 3.6%
10% 6.3% – 9.1% 11.1% – 19.5% 14.8% – 35.6%
20% 12.3% – 18.3% 22.2% – 39% 29.6% – 71.2%

75 75
Binder Mastic
Percent Accuracy in Binder AP

Percent Accuracy in Binder AP


Change in Binder AR Change in Mastic AR
is 1.1 to 1.9 times less is 1.1 to 1.9 times less
than the change in AP than the change in AP
50 50

25 25

Line of Equality Line of Equality


0 0
0 5 10 15 20 25 0 5 10 15 20 25
Percent Error in AR Percent Error in AR
(a) (b)

75
FAM

Change in FAM AR is
Percent Accuracy in Binder AP

1.5 to 3.6 times less


than the change in AP
50

25

Line of Equality
0
0 5 10 15 20 25
Percent Error in AR
(c)

Figure 11.   Summary of results from the sensitivity study.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

26   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

and asphalt concrete, the sensitivity level of a full asphalt concrete mixture would be expected
to be lower than for FAM, but by only a relatively small amount. The findings presented in this
section are based on the AP-based sensitivity assessment; however, the same conclusions are
reached when the assessment is based on G*.
The significance of these findings is two-fold. First, they demonstrate that the properties of
asphalt concrete are not proportionally sensitive to changes in the asphalt binder modulus. That
is, increases in the binder modulus by a given factor do not necessarily result in equivalent factor
changes for the asphalt concrete modulus. Second, as demonstrated most clearly in Figure 11,
these findings show that for a laboratory aging procedure to replicate the impacts of binder oxi-
dation on the modulus of asphalt concrete with a given level of accuracy (say 10%), the desired
binder oxidation can be replicated at less accuracy by a factor of 1.5 to 3.6 (15%–36%). Con-
versely, if a laboratory aging procedure is found to match the in-service level of binder oxidation
with a certain percentage of error (say 10%), then the expected percentage of error in the resulting
modulus value of an asphalt mixture that is tested after being subjected to that laboratory process
would be 1.5 to 3.6 times higher (6.7%–2.8%).
For the assessment of fatigue sensitivity, the parameter used to quantify the sensitivity was
the strain ratio (SR). This SR parameter is based on the ratio of the strain levels (SL) that
are required to achieve given number of failure cycles after and before aging as shown in the
Equation 13.

SLafter aging
SR = (13)
SLbefore aging

Assessments were made using a polynomial relationship between the AP and SR parameters.
The general trend observed in fatigue sensitivity analysis is that if a laboratory aging procedure
is found to match the in-service level of binder oxidation with a certain percentage of error
(say 10%), then the expected percentage of error in the resulting fatigue properties (the SR in
this case) of an asphalt mixture that is tested after being subjected to that laboratory process
will be lower than or equal to 10%. Both sensitivity parameters, i.e., the AR and SR, and their
relationship to the AP, show that the sensitivity of the mechanical properties to the oxidation in
asphalt binders decreases when going from mastic to FAM. Presumably, the sensitivity would
decrease even further in the asphalt mixture. Although general conclusions can be drawn for
fatigue behavior using the relationship between the SR and AP, the actual quantification of the
accuracy of the SR with respect to the error in the AP of the binder cannot be generalized for
different binders and material types. A detailed discussion of the sensitivity study is included in
Appendix C.

Selection of the Chemical and Rheological Aging Index Properties


Six material sources, detailed in Table 6, were used to select the chemical and rheological AIPs
to track the oxidation levels in this project. The materials evaluated encompass a wide range of
binder types. All of the mixtures were subjected to laboratory aging. For three of the mixtures,
both original component materials and field cores from in-service pavements were available
and evaluated. In the case of the WesTrack section, field cores were available at three different
times after placement whereas only one level of aging was available for the FHWA ALF-styrene
butadiene styrene (SBS) and control field cores.
To evaluate the changes in the chemical and rheological AIPs in terms of oxidation, laboratory
loose mixture aging was conducted at multiple temperatures ranging from 70°C to 95°C with
durations ranging from 8 days to 35 days. Samples were collected at different time intervals for
binder extraction and recovery and subsequent AIP testing.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   27  

Table 6.   Asphalt mixtures used for AIP evaluation.

Aggregate Binder Field Material


Mixture Additive
Type/Source PG/Modification Availability
NC Granite/NC 64-22/None N/A N/A
FHWA ALF-
Granite/VA 70-22/None Available Hydrated Lime
Control
FHWA ALF-
Granite/VA 70-22/SBS Available Hydrated Lime
SBS-LG
The same source
SHRP AAD 58-28/None N/A Hydrated Lime
as FHWA ALF
The same source
SHRP AAG 58-10/None N/A Hydrated Lime
as FHWA ALF
WesTrack Granite/NV 64-22/None Available Hydrated Lime

Note: NC = North Carolina, VA = Virginia, NV = Nevada.

Results
Selection of Chemical AIP.   The chemical AIPs were evaluated based on their correlation to
the aging duration. Asphalt materials exhibit relatively similar kinetics, with an initial fast reaction
period, also known as the spurt, followed by a slower reaction period that has an approximately
constant rate (Glaser et al. 2013a, Han 2011, Jin et al. 2011, Petersen 1998, Petersen et al. 1996).
The focus herein is on long-term aging. Therefore, the evaluation of chemical AIPs was conducted
within the constant rate period because it corresponds to higher age levels. Figure 12 shows the
comparisons between the three chemical AIPs evaluated (i.e., carbonyl area, C + S area, and C + S
peaks) with respect to aging duration. The rate of oxidation is temperature-dependent. Therefore,
the chemical AIPs measured from long-term aging at different temperatures for a given mixture
are plotted separately. All of the data included in the analysis of the chemical AIPs correspond to
aging trials where extracted and recovered binder analysis was conducted for at least three aging
durations. As demonstrated in Figure 12, generally all three chemical AIPs strongly correlate
with the aging duration. However, it can be noted that the effect of the sulfoxide functional
group on the chemical AIP-based aging rates is of great importance. Although the C + S area and
C + S peaks show very similar aging rates and rankings, the carbonyl area versus aging duration
shows a different ranking of the materials. That is, in Figure 12(a), the carbonyl area is greater
for the AAG-1 mixture than for the AAD-1 mixture, although their slopes are similar. However,
in Figure 12(b), the C + S area is greater for the AAD-1 mixture than for the AAG-1 mixture,
and the slope of the C + S area is higher for the AAD-1 mixture than for the AAG-1 mixture. The
literature indicates that sulfoxides have a significant effect on rheology, and thus, this observa-
tion suggests that it is important to consider the sulfoxide functional group when tracking oxi-
dative aging. The C + S peaks versus aging duration graph exhibits the overall highest R2 values.
Furthermore, the C + S peaks can be calculated using the direct output of the ATR–FTIR data,
whereas the calculation of the C + S area requires numerical integration under the infrared (IR)
spectrum. Therefore, the C + S peaks is an easy parameter to calculate and is not sensitive to
the method chosen for the calculation. Thus, the C + S peaks was selected as the most promising
chemical AIP identified and was used in the subsequent analysis of the rheological AIPs.

Selection of Rheological AIP.   The correlation between rheological AIPs and the C + S peaks
for each mixture was used to evaluate the rheological AIPs. Data that correspond to different
aging temperatures ranging from 70°C to 95°C were included in the evaluation. Past studies
have demonstrated that the relationship between chemistry and rheology is not affected by
aging temperature if the temperature is below 100°C (Petersen 2009, Elwardany et al. 2017a,
Yousefi Rad et al. 2017). Therefore, the data for multiple aging temperatures are included for
each evaluated mixture.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

28   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

5.0 9.0
NC - 95°C FHWA ALF-SBS - 85°C NC - 95°C FHWA ALF-SBS - 85°C
FHWA ALF-SBS - 95°C SHRP AAD-1 - 70°C FHWA ALF-SBS - 95°C SHRP AAD-1 - 70°C
SHRP AAD-1 - 95°C SHRP AAG-1 - 95°C

Carbonyl + Sulfoxide Area


SHRP AAD-1 - 95°C SHRP AAG-1 - 95°C
WesTrack - 95°C 8.0 WesTrack - 95°C
Carbonyl Area (AU)

R² = 0.989 R² = 0.976
4.0
R² = 1 R² = 0.9895
R² = 0.938

(AU)
R² = 0.9704 7.0
R² = 0.9965
3.0 R² = 0.9645 R² = 0.973
R² = 0.9367
6.0 R² = 0.9438
R² = 0.8867

R² = 0.1406 R² = 0.4373

2.0 5.0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
Oven Aging Duration (Day) Oven Aging Duration (Day)
(a) (b)

0.15
NC - 95°C FHWA ALF-SBS - 85°C
FHWA ALF-SBS - 95°C SHRP AAD-1 - 70°C
SHRP AAD-1 - 95°C SHRP AAG-1 - 95°C
Carbonyl + Sulfoxide Peak

0.13 WesTrack - 95°C


R² = 0.9786
R² = 0.9931
R² = 0.9614
(AU)

0.11
R² = 0.9549 R² = 0.9971
R² = 0.9892
0.09
R² = 0.8378

0.07
0 5 10 15 20 25 30 35
Oven Aging Duration (Day)
(c)

Figure 12.   Sensitivity of different chemical AIPs to aging duration: (a) carbonyl area, (b) C + S area,
and (c) C + S peaks.

Log G* at 64°C and 10 rad/s was selected as the rheological AIP to evaluate oxidation levels
within the project. It was found that oxidative age hardening affects G* most significantly at high
temperatures and/or low frequencies. Therefore, the AIPs that were evaluated at low reduced
frequencies were found to be the most effective. Figure 13 presents the correlation between
log G* at 64°C and 10 rad/s and the C + S peaks for all the mixtures evaluated. Each data point
corresponds to a different aging duration and/or aging temperature. A log scale of the rheo-
logical parameters was used to evaluate the data in a linear trend for easier interpretation of
the test results. The results include those of the binder extracted and recovered from loose
mixture aging that was conducted at various temperatures and depths. The results demon-
strate that the G* at 64°C and 10 rad/s is highly correlated to the chemical changes induced by
oxidation for all of the mixtures evaluated.
The G-R parameter also demonstrated a high correlation with the C + S peaks. Detailed data
are included in Appendix D. In addition, it was found that the G-R is highly correlated with
log G* at 64°C and 10 rad/s. The G-R parameter was evaluated at a very low frequency (0.005 rad/s)
at which the direct measurement of the G* and phase angle is not possible. Therefore, the deter-
mination of the G-R parameter required frequency sweep testing at multiple temperatures and
the fitting of a master curve and time-temperature shift models to the rheological data because

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   29  

3.0 3.0
NC FHWA ALF-Control
log G*, at 64°C, 10 rad/s

log G*, at 64°C, 10 rad/s


2.5 WesTrack Fine 2.5 FHWA ALF-SBS

2.0 R² = 0.9789 2.0 R² = 0.9907


(kPa)

(kPa)
1.5 1.5
R² = 0.9879

1.0 1.0 R² = 0.9152

0.5 0.5
0.07 0.09 0.11 0.13 0.07 0.08 0.09 0.1 0.11 0.12
Carbonyl + Sulfoxide Peak (AU) Carbonyl + Sulfoxide Peak (AU)

3.0
log G*, at 64°C, 10 rad/s

SHRP AAD-1
2.5
SHRP AAG-1
2.0
R² = 0.978
(kPa)

1.5

1.0 R² = 0.9783
0.5

0.0
0.07 0.09 0.11 0.13
Carbonyl + Sulfoxide Peak (AU)

Figure 13.   Correlation between G* at 64°C and 10 rad/s with C + S peaks for six mixtures.

testing at 0.005 rad/s was not possible. In contrast, the G* at 64°C and 10 rad/s is a single point
measurement that can be acquired directly from testing at a single temperature and frequency.
The problems associated with the ZSV and crossover modulus are related to the interpolation
and extrapolation of the data. In addition, the crossover modulus is less sensitive to the chemical
changes caused by oxidation compared to the other rheological parameters evaluated. Therefore,
the G* at 64°C and 10 rad/s constitutes the simplest and most efficient and effective rheological
AIP evaluated. The complete results and discussion that led to the selection of the chemical and
rheological AIPs are provided in Appendices D and E.

Selection of Long-Term Aging Method


To evaluate candidate aging procedures, preliminary aging trials were conducted using a typi-
cal North Carolina mix with 9.5 mm NMAS and PG 64-22 binder, hereinafter referred to as the
NC mix. The integrity of the specimens following aging, the rate of oxidation quantified using the
AIPs of the extracted binder, versatility, and the cost of the various procedures were compared in
order to select the most promising aging procedure. The selected procedure was then applied to
a FHWA ALF-SBS–modified mixture, which is known to be highly susceptible to hardening with
oxidation and difficult to compact, in order to verify that the procedure would not degrade the
specimen integrity. Aging trials using the SBS mix included laboratory aging to match the oxida-
tion level of the surface of an 8-year-old field core obtained from McLean, Virginia (the location
of the FHWA ALF).
Based on the literature review, two candidate aging methods were identified: oven aging and
pressurized aging. For the preliminary evaluation of these two aging procedures, a standard
binder PAV was utilized for the pressurized aging trials. Oven and pressurized aging methods

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

30   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

were applied to both loose mix and compacted specimens and the potential integrity problems
associated with the two specimen types were evaluated.

Study of Aging of Compacted Specimens


Trials for aging the compacted specimens were conducted using both the oven and PAV. The
current standard procedure for asphalt mixture laboratory aging, AASHTO R 30, consists of
aging compacted specimens in the oven at 85°C for 5 days. However, previous research identified
two specimen integrity problems using this procedure:
1. Distortion: Changes in the air void content and geometry due to slumping under self-
weight have been reported when using AASHTO R 30 (Reed 2010). To overcome this issue,
the NCHRP Project 9-23 protocol recommends wrapping specimens in metal wire mesh
secured with three clamps to prevent the samples from geometry distortion (Houston et al.
2005). However, this approach has been reported only to reduce, but not eliminate, specimen
distortion during aging (Reed 2010).
2. Oxidation gradient: NCHRP Project 9-23 demonstrated that the long-term oven aging of com-
pacted specimens leads to both radial and vertical oxidation gradients in mixture specimens,
which is a concern for its use in performance testing because properties differ throughout a
specimen (Houston et al. 2005).
To overcome these specimen integrity problems, two potential remedial approaches were
tried in this study:
1. The application of pressure to increase the diffusion of the oxygen and, hence, potentially
reduce the oxidation gradients.
2. The use of small specimens, 38 mm in diameter and 100 mm in height, to reduce the diffusion
path distances and reduce the slump under self-weight.
To evaluate compacted specimen aging as rigorously as possible, various procedures were
tried, including pressurized and oven aging of both the large and small specimens. All specimens
were fabricated using the NC mix according to the procedures explained in Chapter 2. Three
criteria were used to evaluate the integrity of the specimens that were subjected to compacted
specimen aging:
1. Initial integrity check: Specimen integrity was evaluated initially by visual inspection,
dimension measurements, and air void content comparisons before and after aging to
determine if the specimen had been damaged during the laboratory aging process. If an
aging procedure was found to have disturbed the specimen integrity, it was eliminated from
further consideration.
2. Performance testing: Performance test data were analyzed to detect any integrity issues that
were not related to measurable geometry changes (e.g., microcracking). Dynamic modulus
tests (AASHTO TP 79-12) and cyclic direct tension fatigue tests (AASHTO TP 107-14) were
conducted, and the results were compared to the short-term aged mixture properties to deter-
mine if the specimen integrity had been affected during aging (e.g., if the dynamic modulus
value had decreased upon applying the aging procedure).
3. Oxidation gradient: Oxidation gradients in the aged specimens were evaluated through FTIR
and DSR testing of binders extracted and recovered from locations in an aged specimen that
varied in terms of distance from the specimen periphery. Differences in the rheology and
chemical compositions, along with the distance from the specimen periphery, were used to
detect the presence (if any) of an aging gradient.

Study of Aging of Loose Mixtures


The aging of loose mixtures provided another potential solution to overcome the issues
associated with the AASHTO R 30 procedure. Geometry distortion was not a concern because

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   31  

the loose mix specimens are compacted following aging. In addition, the aging gradient was
not a problem, because the loose mix is aged as a single layer of coated aggregate particles and,
thus, oxygen and heat could circulate easily throughout the mix. Also, the increased surface
area of the binder film that is exposed to oxygen was expected to accelerate aging in loose mix-
tures compared to compacted specimens. However, the compaction of aged loose mix for per-
formance testing remained a potential specimen integrity concern because aged binder is very
stiff and, thus, was expected to be less compactable than unaged material. A past study on the
loose mix aging of an asphalt rubber friction course (ARFC) reported that significantly more
effort is required to compact long-term aged loose mix than short-term aged loose mix (Reed
2010). However, it is important to note that an ARFC represents an extreme case with rubber-
modified asphalt and relatively thick asphalt film. It also has been found that the increased
force/effort required to reach target air void contents when compacting aged loose mixtures
may cause degradation of the aggregate structure and alter the mixture properties (Gatchalian
2006). Also, the compactability of aged loose mix can potentially be improved by increasing
the compaction temperature or by adding a compaction aid such as zeolite.
For this study, preliminary loose mix aging trials consisted of both oven and PAV aging of the
NC mix at 85°C. Loose mix aging trials in the oven and PAV were conducted using the proce-
dures described in Chapter 2. Two criteria were used to evaluate the integrity of the specimens
compacted following loose mix aging:
1. Initial integrity check: The number of gyrations required to meet the target air void content
was compared with that required for the short-term aged mix in order to assess compactabil-
ity. Air void content measurements were used to verify that the desired compaction level
was met. In addition, digital imaging processing software was used to analyze the internal
coarse aggregate structure of the compacted short-term and long-term aged loose mixes to
determine if the aggregate structure had been degraded by compacting the mixes following
long-term aging.
2. Performance testing: Performance tests, including dynamic modulus tests (AASHTO PP 342)
and cyclic fatigue tests (AASHTO TP 107-14), were conducted using specimens that were
compacted following loose mix aging to assess any further potential integrity problems (e.g.,
dynamic modulus decrease upon aging).

Evaluation of the Aging Procedures


To select the most promising aging procedure, compacted and loose mix aging procedures
were compared based on the following criteria:
1. Specimen integrity: Specimen integrity, as related to compacted and loose mix aging, as previ-
ously discussed, is important for reliable performance evaluation.
2. Efficiency: The relative rate of oxidation achieved in each procedure, as evaluated through
comparisons of chemical and rheological AIPs, can be quantified using ATR FTIR and DSR
temperature–frequency sweep testing, respectively.
3. Practicality and versatility: The relative cost and availability of the required equipment were
considered in selecting the most promising aging procedure. Furthermore, the versatility of
the specimen geometries that could be produced for performance testing was an important
consideration.
It was necessary to verify the selected procedure for an additional mixture, particularly to
evaluate potential concerns regarding specimen integrity. For this purpose, the FHWA ALF-
SBS–modified mixture was selected as it is known to be both difficult to compact and highly
susceptible to hardening with oxidation. Asphalt binder was extracted and recovered from a
field core obtained after 8 years in service at the FHWA ALF in McLean, Virginia, for com-
parison to the loose mix aging trial results. Aging for 21 days at 95°C was found to match the

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

32   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

level of oxidation of the binder extracted from the surface of the field core. This condition is
considered to represent an extreme level of oxidation, as oxidation levels are drastically reduced
with the depth of the pavement. To evaluate the integrity of the aged loose mix, the aged mix
was compacted with no adjustment to the short-term aging compaction temperature and then
subjected to dynamic modulus and cyclic fatigue performance testing. The number of gyrations
required for compaction and the performance test results were compared to the results of the
short-term aged specimens to assess the integrity of the specimens that were compacted after
loose mix aging.

Initial Integrity Check Results


Compacted Specimen Aging.   The initial integrity checks of the compacted specimens
included visual inspection, air void content, and dimensions of the specimens. In cases where
air void content and dimensional integrity problems were encountered, remedial strategies were
developed to eliminate them. These strategies are detailed in Table 7.
The comparison of the specimen air void contents before and after aging indicated that very
minor air void changes occurred during the aging of both the small and large specimens with and
without pressure. Note that the initial integrity checks for the small specimens aged in the oven
at 85°C for 8 days indicated that no wire mesh was needed to avoid geometry distortion due to
the relatively low weight of the small specimens.

Loose Mix Aging.   The number of gyrations required to reach the target air void content and
analysis of the coarse aggregate structure were used as initial integrity checks for loose mix aging
because the primary integrity concern was the ability to compact aged material for performance
testing. Two compaction temperatures were tried: 144°C and 157°C. The results showed no
significant difference in the compaction effort required for the short-term and long-term aged
materials. The results thus indicate that it is possible to compact aged loose mix with no adjust-
ment to the compaction temperature.

Performance Testing
Following the initial integrity checks, the specimen integrity of the aged mixtures was assessed
using performance testing. Only a limited number of aging conditions were selected for perfor-
mance testing due to the constraints of time and resources. For specimens that were aged using
compacted specimen aging, the strategies detailed in Table 7 were utilized to minimize integrity-
related problems.

Dynamic Modulus Testing Results.   Figure 14 presents the dynamic modulus master curves;
the results shown correspond to the average values of two replicates. No meaningful difference is

Table 7.   Summary of findings from Level 1 integrity check


of compacted specimen long-term aging.
Recommendation to Avoid Integrity
Material State Temperature Pressure
Problems
Compacted Large Wire mesh support should be used for
85°C -
Specimen specimens.
Compacted Large Specimen should be placed in a
85°C 0.30 MPa
Specimen hammock-like support on its side.
Compacted Small Specimen should be placed in the oven
85°C -
Specimen on its side on a flat surface.
Controlled air pressure application
Compacted Small
85°C 0.30 MPa should be considered and specimen
Specimen
placed on its side on a flat surface.
Note: “-” equals no pressure.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   33  

1E+05 3E+04
S-NC S-NC
L-O-85-8D-NC Compacted at 144°C L-O-85-8D-NC Compacted at 144°C
L-O-85-8D-NC Compacted at 157°C 2E+04 L-O-85-8D-NC Compacted at 157°C
C-O-85-8D-NC C-O-85-8D-NC
1E+04 C-P3-85-1D-NC C-P3-85-1D-NC
2E+04

|E*| (MPa)
|E*| (MPa)

1E+04
1E+03

5E+03

1E+02 1E+02
1E-08 1E-06 1E-04 1E-02 1E+00 1E+02 1E-08 1E-06 1E-04 1E-02 1E+00 1E+02
Reduced Frequency (Hz) Reduced Frequency (Hz)
(a) (b)

Figure 14.   Dynamic modulus results: (a) log-log scale and (b) semi-log scale.

evident between the dynamic modulus values of the short-term aged and PAV-aged compacted
specimens, indicating that either (a) no significant aging occurred or (b) the application of
pressure damaged the specimens. The specimens aged using other methods show a significant
increase in their dynamic modulus values. Loose mix aging appears to lead to slightly higher
dynamic modulus values than compacted specimen aging in the oven for the same duration,
thereby indicating that aged loose mix can be compacted for performance testing. In addition,
these results suggest that higher levels of oxidation are achieved using loose mix aging com-
pared to compacted specimen aging, given the same temperature and duration of conditioning
in an oven. The compaction temperature utilized for the aged loose mixes had little effect on
the dynamic modulus values.

Cyclic Fatigue Performance Testing Results.   Figure 15 presents the damage characteristic
curves obtained from analysis of the cyclic direct tension test results for both the short-term
and long-term aged materials. Typically, damage characteristic curves are used to describe the

1.0
Short-term Aged - Compacted at 144°C
Oven, Loose Mix, 85°C, 8 days - Compacted at 144°C
0.8 Oven, Loose Mix, 85°C, 8 days - Compacted at 157°C
PAV, Compacted Spec., 85°C, 1 day, 300 kPa
Oven, Compacted Spec., 85°C, 8 days
0.6
Failure Point
C

0.4

0.2

0.0
0.0E+00 2.0E+05 4.0E+05 6.0E+05 8.0E+05 1.0E+06
S

Figure 15.   Comparison of damage characteristic curves for the NC


mixes subjected to different aging conditions (C = pseudo stiffness
and S = damage parameter).

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

34   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

relationship between material integrity and damage and are path-independent (i.e., independent
of loading and thermal history). The curves presented in Figure 15 represent the averages obtained
from three different cross-head displacement amplitudes used in the tests. The damage character-
istic curves of the stiffer materials usually are higher than the damage characteristic curves of the
softer mixtures, as demonstrated also in Hou et al. (2010). Because aging is expected to increase
the stiffness of asphalt mixtures, it was expected that the damage characteristic curves of the long-
term aged mixtures would be higher than those of the short-term aged mixtures, unless an integ-
rity problem existed in the specimen. The results demonstrate that all the oven-aged specimens
(both aged loose mix and aged compacted specimens) have higher damage characteristic curves
than the short-term aged specimens, indicating no integrity issues. In other words, the compac-
tion of aged loose mix does not appear to lead to integrity problems. Furthermore, the results
for the long-term aged loose mixture trials show similar damage characteristic curves regard-
less of the compaction temperature, indicating that specimen integrity can be achieved without
elevating the compaction temperature. In addition, the results indicate that the oven-aged loose
mix specimens have slightly higher damage characteristic curves than the oven-aged compacted
specimens, which is consistent with the dynamic modulus test results.
The research team compared two sets of compacted specimens that were subjected to oven
aging: aged 178-mm tall specimens, whereby the specimen ends were cut following aging to
reduce the height to 130 mm for testing, and 130-mm tall specimens aged in the oven after cut-
ting the ends of the specimens. The damage characteristic curves were similar for the two sets
of samples. However, the specimens that were aged at a height of 130 mm (i.e., ends cut before
aging) demonstrated a high propensity for end failure, indicating a higher level of oxidation at
the ends of the specimen and, hence, a testing concern. The damage characteristic curves of the
PAV-aged compacted specimens lie below those of the short-term aged specimen, indicating
that the specimens were damaged by the application and/or release of pressure during aging.
Thus, the application of pressure when aging compacted specimens should be avoided.
A new fatigue failure criterion, termed DR (Wang and Kim 2017), was also used to evaluate
mixtures with different aging levels. The DR criterion uses the average reduction in pseudo stiff-
ness (i.e., C) up to failure. The DR value is calculated as the summation of (1 – C), illustrated in
Figure 16, divided by the fatigue life (number of cycles to failure) for individual test replicates.
DR is a material constant that is independent of mode of loading, temperature, and stress/strain
amplitude. Note that the calculation of DR is in arithmetic scale rather than in log-log scale. Con-
sequently, DR results are not as affected by test variability as another pseudo energy-based failure
criterion, termed GR, that is calculated in log-log scale (Sabouri and Kim 2014).
Pseudo Stiffness (C)

Number of Cycles (N)

Figure 16.   Illustration of summation of (1–C).

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   35  

7E+04
Short-Term Aged, Compacted at 144°C
6E+04 Oven-Aged Loose Mix, 85°C, 8 Days, Compacted at 144°C
Oven-Aged Loose Mix, 85°C, 8 Days, Compacted at 157°C
5E+04 Oven-Aged Compacted Spec., 85°C, 8 Days
Cumulative (1-C)

PAV-Aged Compacted Spec., 300 kPa, 85°C, 1 Day


4E+04

3E+04

2E+04

1E+04

0E+00
0E+00 2E+04 4E+04 6E+04 8E+04 1E+05
Nf (Cycle)

Figure 17.   DR failure criterion lines of NC mix subjected to different


aging conditions (Nf = number of load applications to failure).

Figure 17 depicts the DR energy-based failure criterion plots. Average DR values for different
aging conditions are presented in Table 8. For severely aged materials, it is anticipated that the
DR failure criterion lines will fall below those of short-term aged materials due to the loss of
fatigue resistance associated with embrittlement imposed by oxidation. Any integrity problem
in the specimens is also expected to result in lower DR values. Thus, the failure criterion results
for the PAV-aged compacted specimens indicate a potential integrity problem, as the dynamic
moduli values for the PAV-aged compacted specimens were lower than those from the short-
term aged mixtures. All the other failure criterion lines fall close to that of the short-term aged
materials, which indicates no integrity problems.

Evaluation of Aging Gradient in Compacted Specimen Aging


The chemical and rheological data for extracted and recovered binder that were obtained
from the different aging trials were compared using AIPs. Based on the sensitivity study results
(see Table 5 and Figure 11), a 2% change in the C + S absorbance and a 15% change in the binder
complex modulus both correspond to an approximately 10% change in the dynamic modulus
values of the mixture samples. A 10% change in the dynamic modulus value is considered a
reasonable threshold of significance, and hence, a 2% change in the C + S absorbance and 15%
change in the G* value at 64°C and 10 rad/s were used as thresholds for detecting significant
differences when interpreting the AIP results.
In compacted specimen aging, the oxygen diffusion from the periphery to the specimen center
is impeded by the binder film and aggregate, thus leading to the high possibility of an oxidation
gradient within specimens aged in a compacted state. The extraction and recovery of binder
from different distances from the periphery of compacted specimens following long-term aging

Table 8.  Average DR values for NC mix subjected


to different aging conditions.
Mix ID Average DR
Short-Term Aged, Compacted at 144°C 0.574
Oven-Aged Loose Mix, 85°C, 8 Days, Compacted at 144°C 0.541
Oven-Aged Loose Mix, 85°C, 8 Days, Compacted at 157°C 0.535
Oven-Aged Compacted Spec., 85°C, 8 Days 0.494
PAV-Aged Compacted Spec., 300 kPa, 85°C, 1 Day 0.454

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

36   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Figure 18.   Schematic of components of small


specimen (38 mm ë 100 mm) used to evaluate
oxidation gradients.

allows for the assessment of aging gradients. In this study, this assessment was accomplished
by first coring and cutting the specimens into slices from which the binder was recovered. For
the small specimens, the outer segments were obtained by slicing the specimens vertically, as
depicted in Figure 18. The large specimens were cored using a 38-mm core bit and 75-mm core
bit to obtain radial slices, as shown in Figure 19. Binder extraction and recovery was carried out
only for the outer layer, middle layer, and core of the oven-aged large specimens.
Figure 20 presents the results of the AIP tests for the various aging trials. The data shown
in Figure 20 were further processed to calculate percent changes in C + S absorbance and G*
between different aging conditions. These data are summarized in Table 9. Cells highlighted in
green indicate conditions that show the percent change in G* values is less than 15%, meaning
that the conditions used in calculating the percent change do not result in significant differences
in material properties. The 15% criterion used in the comparison resulted from the sensitivity
study, where 15% change in binder G* resulted in approximately 10% change in mixture
|E *|. The results first demonstrate that the level of oxidation in the short-term aged mixture
surpasses the level of oxidation in the RTFO-aged binder. The results shown in Table 9 indi-
cate that the C + S absorbance and G* values of the long-term aged compacted specimens are

Figure 19.   Schematic of


components of large specimen
(100 mm ë 150 mm) used to
evaluate oxidation gradients.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   37  

0.14 6.0
Carbonyl+Sulfoxide Peak
log G*
0.12 PAV-Aged Oven-Aged Oven-Aged PAV and Oven-
5.5
Aged Binders STA Small Specimen Small Specimen Large Specimen Aged Loose Mix
85°C, 3 days 85°C, 8 days 85°C, 8 days 85°C
Carbonyl + Sulfoxide Peak (AU)

0.10 0.0918 0.0926 0.0972


0.0905

log G* at 64°C, 10 Hz (Pa)


0.0885 0.0871 0.0868 5.00 4.94
4.93 0.0850 5.0
4.84 4.84
4.75
4.80 0.0823 0.0801 4.82
0.0746 4.71
0.08 0.0635 4.63
4.59
4.45 4.5

0.06

4.0
0.04

3.5
0.02

0.00 3.0

0.14 5.5
Carbonyl+Sulfoxide Peak
log G*
0.12
PAV-Aged Oven-Aged Small Oven-Aged Large PAV and Oven-
Aged Binders 5.0
STA Small Specimen Specimen Specimen Aged Loose Mix
Carbonyl + Sulfoxide Peak (AU)

0.10 0.0918 0.0972


0.0926

log G*, 64°C, 10 rad/s (Pa)


0.0885 0.0905
0.0871 0.0868 0.0850
0.0823 0.0801 4.43 4.34 4.5
0.08 4.34 0.0746
4.24 4.23 4.21
0.0635 4.15 4.19
4.09
0.06 4.00
3.96 4.0
3.79

0.04

3.5
0.02

0.00 3.0

Figure 20.   Comparison between C + S absorbance peaks and log G* at 64°C and 10 rad/s for extracted
and recovered binders from different compacted specimen aging processes, loose mix aging trials, and
aged binders.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

38   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Table 9.   Percentage change in chemical and rheological AIPs according


to the different aging methods.

% Change
Between STA &
Between Between Between Outer
RTFO+PAV Binder for Aged Binder,
RTFO & Outer Layer Layer & Mid
Outer Layer for Compacted Specimens,
PAV & Core Layer
and Loose Mix
C+S G* C+S G* C + S G* C + S G*
Aged Binder 44.6 254.8 23.1 139.9
PAV-Aged Small Specimen 21.3 90.5 2.2 18.7
Oven-Aged Small Specimen 16.4 86.2 0.3 8.8
Oven-Aged Large Specimen 13.9 77.8 3.2 38.3 5.8 24.1
Oven-Aged Loose Mix 24.1 195.1
PAV-Aged Loose Mix 30.3 139.9

Note: STA = short-term aged.

significantly higher than those of the short-term aged loose mixtures based on the previously
defined thresholds. However, the AIP values of all the long-term aged compacted specimens
are lower than those for the RTFO+PAV-aged binder.
The AIP results of binders that were extracted and recovered from the large compacted speci-
mens aged for 8 days at 85°C indicate greater levels of oxidation at the outer portion of the
specimen than the core; this indicates the presence of an oxidation gradient. However, the AIP
values corresponding to the outer segments and the cores of the long-term aged small specimens
are similar, thus confirming that shorter diffusion paths can mitigate oxidation gradient con-
cerns. The AIP values of binder extracted from the small specimens that were aged in the oven
are relatively similar to the outer portion of large specimens aged under the same conditions.
Based on the AIP results, the small specimens aged in the PAV experienced a higher level of
oxidation than the oven-aged specimens, indicating that pressure expedites the oxidation of
mixtures. However, for the small specimens aged in the PAV (3 days at 85°C and 300 kPa), a
significant aging gradient is observed based on the G* results, thereby indicating that pressure
does not alleviate the effects of an oxidation gradient.
The long-term aging of the loose mixture in an oven yields AIP values that exceeded those of
the compacted specimens aged for the same duration. Thus, aging loose mix appears to expedite
oxidation significantly compared to aging compacted specimens. The AIP values of the long-
term aged loose mixtures also exceeded the value of the RTFO+PAV binder. The long-term
aging of the loose mix in the PAV for 2 days at 85°C and 2.1 MPa led to a level of oxidation
relatively similar to that of the loose mix oven aging for 8 days at the same temperature. These
results suggest that the addition of pressure can expedite aging almost four times faster than a
conventional aging oven. However, it is important to note that the standard binder PAV does
not allow enough space for aging sufficient quantities of material for performance testing in a
single trial. Thus, if PAV aging of loose mix were to be adopted, a new and larger PAV would
have to be developed.

Summary of Compacted Specimen and Loose Mix Aging


Compacted Specimen Aging.   Although no integrity issues in terms of changes in air void
content or specimen dimensions were encountered with the AASHTO R 30 protocol, the aging of
large (100-mm diameter) compacted specimens with wire mesh support in an oven at 85°C led
to the development of an aging gradient within the specimens. This lack of uniform properties
throughout the specimen is of concern for performance testing and was observed directly as a

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   39  

high rate of failure at the end locations where oxidation was most significant in the cyclic direct
tension fatigue tests. However, the high rate of end failure could be overcome by aging specimens
178 mm in height and then trimming the ends to produce a 130-mm tall specimen for testing.
The aging gradient observed in the large compacted specimens subjected to oven aging was
eliminated by using small specimens (38-mm diameter with 100-mm height) due to the shorter
diffusion paths of the smaller specimens.
The application of pressure in compacted specimen aging was found to expedite aging. How-
ever, oxidation gradients were observed in the pressure-aged specimens. In addition, although
no changes in air void contents or specimen dimensions were induced by pressurized aging,
the performance test results indicate that the application and/or release of pressure can damage
specimens. Therefore, the results indicate that the most promising method for aging compacted
specimens is to age small specimens in an oven without pressure.

Loose Mix Aging.   The primary concern associated with loose mixture aging is the ability to
compact the material after long-term aging. However, in this study, the compaction of the NC
mix after 8 days of oven conditioning at 85°C was possible with no adjustment to the compac-
tion temperature. A similar number of compaction gyrations was required for both the short-
and long-term aged loose mixes. The image analysis of the aggregate structure also indicated
comparable compaction of both the short- and long-term aged loose materials. Furthermore,
the performance test results indicate a significant increase in the dynamic modulus values of
the long-term aged material compared to those of the short-term aged material, and the fatigue
performance test results indicate no integrity concerns. In addition, loose mix aging exposes a
large surface area of the binder to oxygen, and thus, a faster rate of oxidation was observed in the
loose mix oven aging compared to the compacted specimen aging based on the measured asphalt
binder chemical and rheological aging index values. The application of pressure also was found
to expedite the oxidation of the loose mixes. However, only 500 g of loose mix could be aged at
one time in the binder PAV. One Superpave gyratory-compacted specimen requires the prepara-
tion of 7000 g to 8000 g of loose mix. Thus, the binder PAV would need to be run approximately
15 times to generate enough loose mix to prepare a compacted specimen for performance test-
ing, which is inefficient and therefore was deemed impractical. Again, selection of pressure aging
of loose mix would necessitate development of a new, larger PAV.

Selection of Aging Procedure


Based on the findings for the compacted and loose mix aging trials with and without pressure,
the research team identified oven aging of loose mix as the optimal aging procedure. To avoid
specimen integrity issues, the only option for the oven aging of compacted specimens is to age
small specimens in an oven. This small specimen geometry allows for only dynamic modulus
and direct tension testing. Other tests (e.g., permanent deformation tests) would not be possible
under this scenario. Loose mix aging is more versatile than compacted specimen aging in that any
specimen geometry (e.g., slabs or beams) can be produced using aged loose mix. In addition,
loose mix aging leads to faster oxidation than compacted specimen aging and therefore offers
efficiency gains. As discussed, the PAV aging of loose mix would require the development of a
new, larger, mixture-specific PAV, which would be costly. Thus, given that aging loose mix can
be accomplished relatively quickly in an oven using multiple pans, this method is considered a
more practical approach at present. In this research, a conventional oven with inner chamber
dimensions of 36 in. × 24 in. × 19 in. (W × H × D) was used. Using only six shelves, 18 pans
(13 in. × 18 in. × 1 in.) can fit inside the oven. Loose mix spread in four pans is sufficient
for the preparation of one Superpave gyratory-compacted specimen that is 150 mm in diameter
and 178 mm in height. Also, the preliminary results indicate that the compaction of aged loose
mix can be accomplished using the same temperature that is required to compact short-term
aged material.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

40   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Evaluation of the Selected Aging Procedure


To verify the findings of the loose mix aging trials with the NC mix, the FHWA ALF SBS-
modified mixture was subjected to loose mix oven aging trials at 70°C, 85°C, and 95°C. Small
samples were removed periodically from the oven and subjected to extraction and recovery to
determine the binder AIPs. In order to compare the level of oxidation achieved in the laboratory to
that found under field conditions, the AIPs determined for binder extracted from the aged loose
mix were compared with the AIPs determined for the asphalt binder extracted from the top lift of
a field core extracted after 8 years in service. The results were used to approximate the oven condi-
tioning time needed for the loose mix to reach the oxidation level of the field core at various depths.
Two aging procedures were used to evaluate the integrity of the specimens that were compacted
after loose mix aging: 8 days of conditioning at 85°C (consistent with the NC mix aging trials) and
21 days of conditioning at 95°C, which was found to correspond to the same oxidation level as the
surface of the 8-year-old field core. The conditioning temperature of 95°C was selected because
it can expedite aging faster than 85°C but is not expected to lead to volatilization or degradation
of the polymers (Petersen 2009). Note that the asphalt binder oxidation level at the surface of the
field core is thought to represent an “extreme” level of aging, as the surface of the field core was
found to be severely oxidized compared to samples extracted from deeper within the pavement.
Following compaction, the aged loose mix was evaluated using the same protocol utilized for the
NC mix, which included an assessment of compactability and performance testing.

Results.  The 8-year-old field core obtained from the FHWA ALF was cut (sawn) to obtain
3 half-inch-thick slices. The asphalt binder was extracted and recovered from the field core slices,
and then the AIPs were determined. The results were used to evaluate the oxidation level of
the field core in terms of depth for comparison to the oxidation levels of the laboratory-aged
samples. A significant oxidation gradient was found within the field core, with the surface of the
field core being more severely oxidized than from deeper within the pavement.
Figure 21 provides a graphic representation of the estimated oxidation levels for loose mix aging
at 95°C that were needed to match the level of oxidation of the field core at different depths. Even
at 95°C, 21 days of oven conditioning were needed to match the level of oxidation of the surface of
the field core. However, it is worth noting that the smallest sample utilized in the performance tests
of asphalt concrete is 38 mm in diameter or thickness. If a 38-mm specimen is obtained from the

Carbonyl + Sulfoxides Absorbance Peaks (AU)


0.104 0.106 0.108 0.11 0.112
0
21 Days of Aging at 95°C
-5 First Slice of
Top Layer
-10

-15
Depth (mm)

Second Slice
-20 15 Days of Aging at 95°C
of Top Layer
-25

-30 Third Slice


13 Days of Aging at 95°C of Top Layer
-35

-40
Tack Coat Layer
-45

Figure 21.   Comparison between field core aging gradient


with respect to depth and long-term aging of loose mix in
oven at 95°C.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   41  

top surface of a field core, the material in the top 19 mm of the specimen has aged more than the
material at the 19 mm depth. Also, the material in the bottom 19 mm of the specimen has aged less
than the material at the 19 mm depth. Therefore, it is reasonable to assume that the representative
aging level for a depth of 38 mm would be close to the oxidation level at a depth of 19 mm. This
explanation would indicate that 15 days are required for aging loose mix to match the age level of
the top 38 mm of an 8-year-old field core extracted in Virginia. Additionally, for structural modeling
and analysis purposes (e.g., using the FlexPAVE™ program), typically the averaged properties of each
layer are used as inputs. Therefore, for this study, laboratory-prepared samples were aged in order to
meet the average level of aging of each layer (in this case, 15 days of aging at 95°C). Based on these
results, the oxidation level of the surface is considered to be a severe condition, because oxidation
greatly dissipates with depth. Thus, loose mix aged at 95°C for 21 days is considered an “extreme”
condition for evaluating the compactability and integrity of aged loose mix.
Figure 22 shows comparisons of the degree of aging, using the C + S absorbance peaks and
*
G AIPs, among the aged loose mixes used for the compacted specimen integrity assessment,
the short-term aged mix, and binders aged in the RTFO and PAV. The results demonstrate
that 21 days of loose mix conditioning at 95°C greatly exceeds the oxidation level of the binder
PAV, which is comparable to 8 days of loose mix aging at 85°C.
The compactability of the long-term aged FHWA ALF-SBS–modified mix was evaluated as
an initial specimen integrity check by comparing the number of gyrations needed to reach the
target air void content with the number needed for short-term aged materials. The number of
gyrations needed to reach the target air void contents was similar for both the short-term aged
loose mixture and the two levels of long-term aged loose mixture (8 days at 85°C and 21 days
at 95°C) with no adjustment of the compaction temperature. Thus, compacting the long-term
aged loose mix was not problematic.
Performance testing was utilized as an additional means to evaluate the integrity of the
FHWA ALF-SBS–modified specimens that were compacted following long-term loose mix aging.

0.16 8.0
Carbonyl + Sulfoxide Peak
log G*
Short-term Field Core
Aged Binder Oven Aged Loose Mix
Aged Mix Top 6 mm
Carbonyl + Sulfoxide Peak (AU)

0.12 0.1188 6.0


0.1111
log G* at 64°C, 10 rad/s (Pa)

5.18
4.98
0.0923
0.0875 4.37 4.28
3.85 3.80
0.08 4.0
0.0705

0.0577

0.04 2.0

0.00 0.0

Figure 22.   Comparisons between C + S absorbance peaks and log G* at 64°C and 10 rad/s
for extracted and recovered binders from different loose mix aging trials, field core top
layer surface slices, and aged binders (SBS-modified mixture).

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

42   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

2E+05
Short-term Aged Short-term Aged
Oven-Aged Loose Mix, 85°C, 8 days 3E+04 Oven-Aged Loose Mix, 85°C, 8 days
Oven-Aged Loose Mix, 95°C, 21 days Oven-Aged Loose Mix, 95°C, 21 days

2E+04

|E*| (MPa)

|E*| (MPa)
2E+04

2E+03
1E+04

2E+02 1E+02
1E-08 1E-05 1E-02 1E+01 1E-08 1E-05 1E-02 1E+01
Reduced Frequency (Hz) Reduced Frequency (Hz)
(a) (b)

Figure 23.   Dynamic modulus test results: (a) log-log scale and (b) semi-log scale.

Comparisons between the dynamic modulus and cyclic fatigue damage characteristic curves of
the short-term aged material and the long-term aged material were used to evaluate the specimen
integrity of the compacted long-term aged material based on expected trends of increased aging.
Figure 23 presents the dynamic modulus master curves. The master curves represent the aver-
aged values of two replicates. The results indicate that the oven-aged loose mix specimens have
higher dynamic modulus values than the short-term aged specimens. Furthermore, the results
suggest that the specimens that were compacted after 21 days of oven aging at 95°C have signifi-
cantly higher dynamic modulus values than the specimens compacted after 8 days of oven aging
at 85°C, as was expected based on the AIPs. If severe integrity problems had been present in the
long-term aged specimens, then the dynamic modulus values would not be significantly higher
than those of the short-term aged specimens, regardless of the aging level (because damage in the
specimen reduces the dynamic modulus value), and thus, no integrity problems were detected
in the dynamic modulus test results.
Figure 24 presents the cyclic fatigue characteristic curves for the short- and long-term
aged materials. These curves define how damage grows in a material and represent the averaged

1.0
Short-term Aged
Oven, Loose Mix, 85°C, 8 days
0.8
Oven, Loose Mix, 95°C, 21 days
Failure Point
0.6
C

0.4

0.2

0.0
0.0E+00 5.0E+05 1.0E+06 1.5E+06 2.0E+06
S

Figure 24.   Comparison of damage characteristic curves


from FHWA ALF SBS mixtures subjected to different aging
conditions.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   43  

results of three tests that were conducted using various cross-head displacement amplitudes.
The long-term aged specimens have higher damage characteristic curves than the short-term
aged specimen, which follows expected trends. Thus, the performance test results indicate that
no integrity issues are associated with the compaction of the aged loose mix.
Figure 25 shows the DR energy-based failure criterion results for the SBS-modified mix. The
failure criterion line for the specimens that were compacted following long-term aging for 21 days
at 95°C falls significantly lower than the failure criterion lines for the short-term aged specimens
and the specimens aged for 8 days at 85°C. This outcome suggests that the severe level of aging
led to embrittlement and consequently degraded the resistance to fatigue and, hence, suggests no
integrity problems.
These results indicate that the compaction of long-term aged loose mix is possible with no
adjustment to the compaction temperature, based on both the number of compaction gyrations
required to reach the target air void contents and the performance test results. The long-term
aged loose mix that was aged at 95°C for 21 days had an oxidation level that was equivalent to
that of the surface of an 8-year-old field core obtained from the FHWA ALF in McLean, Virginia.
This level is assumed to represent an extreme oxidation level that nonetheless allows for com-
paction of the mix. Additional details on the selection of the aging procedure are included in
Appendix F and elsewhere (Elwardany et al. 2017b).

Selection of Laboratory Aging Temperature


Petersen (2009) suggested that aging temperatures that exceed 100°C can induce chemical
changes to the oxidation reaction in asphalt materials. To evaluate the performance impli-
cations of long-term aging temperatures below and above 100°C, comparative tests between
loose mixtures aged at 95°C and 135°C were conducted to evaluate the implications of loose
mix aging at 135°C in terms of asphalt mixture performance using three mixtures, all prepared
with the same FHWA ALF aggregate, but with different binders: ALF-SBS, SHRP AAD, and
SHRP AAG. The SHRP AAD and SHRP AAG binders were selected due to their known dif-
ferences in chemistry. The SHRP AAD binder has a high sulfur content (6.9%) and is highly
structured (i.e., its components exhibit a high degree of incompatibility). Thus, the SHRP AAD
binder was expected to be especially susceptible to changes in oxidation kinetics and mechanics
at 135°C. The SHRP AAG binder has a low sulfur content (1.3%) and is less structured (more
compatible) than SHRP AAD. Thus, the SHRP AAG binder was expected to be less susceptible to
changes in oxidation kinetics and mechanisms when the temperature for loose mixture aging was

5E+04
Short-Term Aged
Oven-Aged Loose Mix, 85°C, 8 days
4E+04
Oven-Aged Loose Mix, 95°C, 21 days
Cumulative (1-C)

3E+04

2E+04

1E+04

0E+00
0E+00 2E+04 4E+04 6E+04 8E+04
Nf (Cycle)

Figure 25.   Comparison of DR failure criterion lines


for SBS-modified mixtures.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

44   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

increased from 95°C to 135°C. The FHWA ALF-SBS–modified mixture was selected in order to
include a common asphalt-modified asphalt binder in the study and because field core data were
available for it. Figure 26 presents a summary of the experimental plan that was implemented for
each mixture to evaluate the implications of loose mixture aging at 95°C and 135°C. Note that the
rheological AIP used in the selection of the laboratory aging temperature is G* at 64°C and 10 Hz,
which differs from the selected rheological AIP of G* at 64°C and 10 rad/s. However, the use of G*
at 64°C and 10 Hz versus G* at 64°C and 10 rad/s is not anticipated to affect the findings related to
the selection of the laboratory aging temperature.
First, two batches of loose mix were short-term aged at 135°C for 4 hours and then subjected
to long-term aging at 95°C and 135°C using the proposed aging method. Small samples of the
loose mixture were taken from the pans at periodic intervals for binder extraction and recovery.
The changes in the asphalt binder oxidation level versus the aging duration were assessed for the
two aging temperatures by means of chemical and rheological AIPs.
Figure 27 shows the relationship between the G* value at 64°C and 10 Hz frequency and
the C + S absorbance peaks for the binder samples extracted and recovered from the SHRP
AAD mix aged at 95°C and 135°C. Figure 27 indicates that the binders aged at the two dif-
ferent temperatures have different C + S absorbance peaks for the same G* value, indicating
that a change in the oxidation reaction mechanism occurred when the aging temperature
was increased from 95°C to 135°C. An analogous trend was observed for the FHWA ALF-
SBS mixture and, to a lesser extent, the SHRP AAG mixture. The G* value was selected as
the aging index value that matched the degree of aging between the two aging temperatures
because binder rheology is speculated to be related to asphalt mixture performance more
directly than binder chemistry. Also, the selection of G* as the primary aging index allowed
the effects of different C + S absorbance peaks on mixture performance to be evaluated,

Oven Aging of Loose Mix at Oven Aging of Loose Mix at


95 °C for Different Durations 135 °C for Different Durations

Binder Extraction Binder Extraction


and Recovery and Recovery

FTIR-ATR Testing DSR Testing DSR Testing FTIR-ATR Testing

Determining the Required


Aging Durations at 95 °C
and 135 °C to Match the
Specified G* Values at 64 °C
and 10 Hz

Oven Aging of Loose Mix at Oven Aging of Loose Mix at


95 °C to Match Target G* 135 °C to Match Target G*

Performance Testing Comparing Performance Testing


(Dynamic Modulus and the Test (Dynamic Modulus and
Cyclic Fatigue) Results Cyclic Fatigue)

Figure 26.   Summary of experimental plan.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   45  

3.5 Short-Term Aged

log G* at 64°C, 10 Hz (kPa)


Loose Mix, 95°C, UC
3
Loose Mix, 135°C, UC
2.5

1.5

0.5

0
0.07 0.08 0.09 0.1 0.11 0.12
Carbonyl + Sulfoxide Absorbance Peaks (AU)

Figure 27.   SHRP AAD loose mix prepared for


long-term aging.

because the same G* value for mixtures aged at two different temperatures will result in two
different C + S absorbance peaks.
The next step in the experimental plan was to determine the aging durations for the samples
tested at 95°C and 135°C that would yield the same G* value. For the 95°C aging, 21 days of con-
ditioning was determined for the FHWA ALF-SBS–modified mixture. Note that 21 days of con-
ditioning at 95°C led to an equivalent oxidation level for the top 6 mm of an 8-year-old field core
obtained from McLean, Virginia. The G* value for 21 days of conditioning at 95°C was deter-
mined to be approximately four times the G* value for PAV-aged asphalt binder. Field cores that
corresponded to the SHRP AAD and SHRP AAG mixtures were not available. Therefore, based
on the finding for the FHWA ALF-SBS–modified mixture, four times the G* value of the PAV-
aged binder was used as the target G* value that would reflect reasonable field aging levels of the
pavement surface after a prolonged in-service period. Thus, the required aging durations at 95°C
and 135°C to achieve G* values equal to four times that of PAV-aged binder were determined and
used to evaluate the effects of aging at 135°C for both the SHRP AAD and SHRP AAG mixtures.
Figure 28, Figure 29, and Figure 30 show the procedures that were used to match the aging
levels between the loose mixtures aged at 95°C and at 135°C for the FHWA ALF-SBS, SHRP AAD,

1E+04
Short-term Aged
Loose Mix, 95°C, UC
Loose Mix, 135°C, UC
G*, 64°C, 10 Hz (kPa)

G* = 43.1e0.0431 (H)
1E+03
Target G*

1E+02
G* = 57.068e0.0034 (H)

52 H 21 D

1E+01
0 200 400 600 800
Aging Duration (Hours)

Figure 28.   Determination of FHWA ALF-SBS aging


durations.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

46   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

1E+05
Short-term Aged
Loose Mix, 95°C, UC
Loose Mix, 135°C, UC
1E+04

G*, 64°C, 10 Hz (kPa)


G* = 65.4010e0.0700 (H)
1E+03
Target G*

1E+02
G* = 52.818e0.0065 (H)
16.8 H 8.9 D

1E+01
0 200 400 600
Aging Duration (Hours)

Figure 29.   Determination of SHRP AAG aging


durations.

and SHRP AAG materials, respectively. To match the aging levels between the loose mixtures aged
at 95°C and 135°C, first the relationship between the log G* and aging duration was obtained
for each aging temperature. Then, in the case of the FHWA ALF-SBS material, the G* value that
corresponded to 95°C before compaction was determined. Based on the relationship between
log G* and the time that corresponded to the loose mix aged at 135°C, the required duration
of aging at 135°C to match this G* value was then determined via interpolation. For the SHRP
AAD and SHRP AAG materials, the aging durations required to match the predetermined
G* values, which corresponded to four times the value of G* for the binder after PAV aging
(the target G*), at both 95°C and 135°C were determined. Note that the results presented in Fig-
ure 29 and Figure 30 demonstrate that the SHRP AAD mix required shorter aging times at both
95°C and 135°C to achieve the requisite G* values than the SHRP AAG mix. This finding matches
expectations in terms of the binders’ microstructures. That is, the SHRP AAD mix is incompatible
and therefore was expected to have a high level of hardening susceptibility with the oxidation level
compared to the SHRP AAG mix, which is compatible and therefore less structured.
Comparisons of the performance of the mixtures prepared by aging at 95°C and 135°C (with
equivalent binder G* values) were used to assess the performance implications of the aging
temperatures. Note that to build the dynamic modulus master curves, frequency sweep tests

1E+05
Short-term Aged
Loose Mix, 95°C, UC
1E+04 Loose Mix, 135°C, UC
G*, 64°C, 10 Hz (kPa)

G* = 13.9933e0.0644 (H)
1E+03
Target G* G* = 13.176e0.0055 (H)

1E+02

37.6 H 19 D
1E+01
0 200 400 600
Aging Duration (Hours)

Figure 30.   Determination of SHRP AAD aging


durations.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   47  

were conducted at multiple temperatures. In addition to material testing, the effect of the aging
temperature on fatigue performance was evaluated at the pavement level. The pavement perfor-
mance was predicted using FlexPAVE™.

FHWA ALF-SBS.   Figure 31 presents the mixture performance test results for the FHWA
ALF-SBS mixture. Figure 31 (a) presents the FHWA ALF-SBS mixture dynamic modulus master
curves that correspond to the specimens fabricated after aging at 95°C for 21 days and 135°C for
52 hours. The dynamic modulus test results for the specimens fabricated after short-term aging
only also are provided for reference. The results indicate a significant increase in dynamic mod-
ulus values with long-term aging. The results also indicate very little difference in the dynamic
modulus master curves that correspond to mixtures aged at 95°C for 21 days and 135°C for
52 hours, suggesting that the chemical changes, represented by the C + S absorbance peaks,
induced by aging at 135°C do not have a significant effect on the mixture dynamic modulus
of the FHWA ALF-SBS mixture. Differences in the time–temperature shift factors between the
short-term and long-term aged materials led to the differences in the reduced frequency range
observed in the corresponding master curves. It can be seen that the data obtained from testing
long-term aged specimens cover a larger reduced frequency domain than the short-term aged
data, whereas the specimens that correspond to both conditions were tested over the same range
of temperatures.
Direct tension cyclic tests were performed on the FHWA ALF-SBS mixture specimens that are
aged at 95°C for 21 days and at 135°C for 52 hours to assess the implications of aging temperature
with regard to fatigue cracking performance. The fatigue tests were conducted using three cross-
head displacement levels (low, intermediate, and high) that were selected based on the displace-
ment of the test machine’s actuator, thus resulting in different on-specimen strain levels. Figure 31

1E+5

1E+4
|E*| (MPa)

Short-term Aged
Oven, Loose Mix, 95°C, 21 days
1E+3
Oven, Loose Mix, 135°C, 52 hours

1E+2
1E-8 1E-4 1E+0 1E+4
Reduced Frequency (Hz)
(a)

1.0 5E+04

0.8 4E+04
Cumulative (1-C)

0.6 3E+04
C

0.4 2E+04

0.2 1E+04

0.0 0E+00
0E+0 3E+5 6E+5 9E+5 1E+6 2E+6 0E+00 2E+04 4E+04 6E+04 8E+04
S Nf (Cycle)
(b) (c)

Figure 31.   FHWA ALF-SBS mixture performance test results: (a) dynamic modulus
curves, (b) C versus S curves, and (c) DR failure criterion lines.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

48   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

(b) presents the C versus S curves. The results show similar C versus S curves for the two long-
term aging conditions, whereas the short-term aging C versus S curves are considerably lower,
which is attributable to the short-term aged mixtures’ lower stiffness values (age level) compared
to those of the long-term aged material. These results indicate no significant effect of aging
temperature on the fatigue performance of the FHWA ALF-SBS mixture.
In addition to the C versus S curves, the DR failure criterion was evaluated in this study. Fig-
ure 31 (c) presents the failure criterion results for the FHWA ALF-SBS mixture, showing that
the failure criterion lines for the long-term aged materials fall below the line for the short-term
aged material, which is indicative of the lower fatigue resistance of long-term aged materials as
a result of the brittleness caused by oxidative aging in the asphalt binder. However, the results
indicate very similar failure criterion lines for the specimens prepared with loose mixture aged
at 95°C for 21 days and at 135°C for 52 hours.
These results suggest that the chemical changes induced by aging at 135°C do not signifi-
cantly affect the performance of the FHWA ALF-SBS mixture. The FHWA ALF-SBS mixture
contains SBS modification, which may mask the effects of microstructural changes induced
by aging at 135°C. Hence, the evaluation of asphalt mixtures that contain unmodified asphalt
binders (i.e., SHRP AAD and SHRP AAG) is also important.
Table 10 presents the statistical t-test analysis outcomes for the dynamic modulus and fatigue
test results that correspond to the FHWA ALF-SBS materials with different aging conditions.
For each pair of compared samples, the data points that correspond to the selected reduced
frequencies or S values were used for the analysis. The selection of the reduced frequencies and
S values was based on the ranges of both the dynamic modulus master curves and C versus
S curves. Very high reduced frequencies were avoided due the fact that the aging effect on the
dynamic modulus is more pronounced at low frequencies or high temperatures. Also, very low
S values were not considered because the properties of asphalt mixture specimens are more
distinguishable at higher S values (or smaller C values) when different materials are compared
with each other.
The data points for two dynamic modulus test replicates and three fatigue test replicates were
used for the statistical analysis. As mentioned earlier, a confidence level of 95% was employed to
evaluate the difference between the pairs of aging treatments.

Table 10.   Statistical t-test analysis outcomes for dynamic


modulus and cyclic fatigue test results for FHWA ALF-SBS
materials.
Reduced Frequency
Sample 2.0E-06 1.0E-04 1.0E-03 1.0E-01
p-value
Dynamic Short-term aged
0.0263 0.0014 0.0037 0.0201
Modulus Loose mix, 95°C, 21 days
Short-term aged
0.0061 0.0028 0.0051 0.0150
Loose mix, 135°C, 52 hours
Loose mix, 95°C, 21 days
0.3529 0.1453 0.0591 0.3346
Loose mix, 135°C, 52 hours
S
Sample 1.0E+05 3.0E+05 5.0E+05 7.0E+05
p-value
Short-term aged
Cyclic 0.0005 0.0001 0.0006 0.0025
Loose mix, 95°C, 21 days
Fatigue
Short-term aged
0.0078 0.0004 0.0004 0.0029
Loose mix, 135°C, 52 hours
Loose mix, 95°C, 21 days
0.3599 0.0918 0.1230 0.4560
Loose mix, 135°C, 52 hours

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   49  

The pairs with significant differences (p < 0.05) are shaded in Table 10. The statistical anal-
ysis results suggest a significant difference between the dynamic modulus test results for the
short-term aged and long-term aged materials. However, as inferred from the comparison of the
master curves, no significant difference is seen between the ALF-SBS materials aged at 95°C and
at 135°C. The p-values for the C versus S curves indicate a similar conclusion.
The material properties shown in Figure 31 were input to FlexPAVE™ to predict the cracking
performance of the study pavement. Considering the different positions of the failure criterion
lines between the short-term aged and long-term aged conditions, very different field perfor-
mance predictions were anticipated for these two conditions.
A simple pavement structure was considered: a 10-cm asphalt concrete layer over a 20-cm
aggregate base and 380-mm subgrade. For each binder type, based on the reported PG, different
EICM data were selected for the performance predictions. EICM data for Washington, D.C.,
were used for the ALF SBS material, and climatic data for Ann Arbor, Michigan, and San Luis
Obispo, California, were used for the SHRP AAD and SHRP AAG materials, respectively. The
traffic input was 3,500 daily equivalent single-axle loads (ESALs). The analysis was performed
for a 20-year service life. FlexPAVE™ was used to predict the distribution of damage within a
cross-section of the asphalt pavement layer after 20 years of traffic loading. The wheel path was
directly above the region of damage localization. To compare the results of the fatigue crack-
ing predictions, the percentage of damage (referred to as “percent damage”) was computed
as a function of time for each case. Percent damage is defined as the ratio of the sum of the damage
factors within the reference cross-section area to the reference cross-section area itself, as shown
in Equation 14. (Note: the percent damage area is defined schematically in Figure 33, also.)

 N 
∑ i=1  N
M
 × Ai
f i
Percent Damage = (14)
∑ i=1 Ai
M

where
i = nodal point number in finite element mesh,
M = total number of nodal points in finite element mesh,
N/Nf = damage factor,
N = number of load applications,
Nf = number of load applications to failure,
Ai = area represented by nodal point i in finite element mesh, and
Σ Ai = reference area.
Figure 32 shows the distribution of damage within a cross-section of the asphalt concrete
layer of the pavement for different aging treatments predicted using the FlexPAVE™ program.
This figure suggests no significant differences between the distribution of damage predicted for
asphalt loose mixture aged at 95°C compared to that aged at 135°C. However, as expected, the
short-term aged material performed better than the long-term aged material. It is also noted that
the long-term aging of the SBS mixture increases the top-down cracking propensity of the study
pavement greatly. Although this observation is based on the unrealistic aging condition (i.e.,
constant aging through the thickness of asphalt layer), it signifies the importance of including
accurate aging condition in LTPP prediction.
FlexPAVE™ employs two overlapping triangles to form the reference cross-section area
within which the damage evolution can be considered (Kim et al. 2017), as shown in Figure 33.
The top inverted triangle has a 170-cm wide base that is located at the top of the surface layer
and a vertex that is located at the bottom of the bottom asphalt layer. The 12-cm wide base of the

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

50   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

0 1 0 1 0 1

1 0.9 –0.01 0.9 –0.01 0.9

2 0.8 –0.02 0.8 –0.02 0.8

3 0.7 –0.03 0.7 –0.03 0.7

4 0.6 –0.04 0.6 –0.04 0.6


Z (cm)

Z (m)

Z (m)
5 0.5 –0.05 0.5 –0.05 0.5

6 0.4 –0.06 0.4 –0.06 0.4


7 0.3 –0.07 0.3 –0.07 0.3
8 0.2 –0.08 0.2 –0.08 0.2
9 0.1
–0.09 0.1 –0.09 0.1
10
–1.5 –1 –0.5 0 0.5 1 1.5 –0.1 0 –0.1 0
–1.5 –1 –0.5 0 0.5 1 1.5 –1.5 –1 –0.5 0 0.5 1 1.5
X (m)
X (m) X (m)

Short-Term Aging Loose Mix, 95°C, 21 Days Loose Mix, 135°C, 52 Hr

Figure 32.   Damage contours for FHWA ALF-SBS mixture aged at different conditions.

second triangle is located at the bottom of the bottom asphalt layer and its vertex is positioned at
the surface layer. Figure 33 presents the final shape of these overlapping triangles, which defines
the reference area for percent damage.
FlexPAVE™ program output data were used to calculate the fatigue damage area as a function
of traffic load repetition. Figure 34 presents the fatigue damage area versus the service life for the
three aging treatments. A significant difference can be observed between the fatigue performance
of the short-term and long-term aged materials, whereas the results for the two long-term aging
treatments at 95°C and 135°C do not indicate any significant differences. These observations are
in agreement with the fatigue performance data.

SHRP AAD.   Figure 35 presents the mixture performance test results for the SHRP AAD mix-
ture. Figure 35 (a) presents the SHRP AAD mixture dynamic modulus master curves that cor-
respond to the specimens fabricated after aging at 95°C and 135°C for 8.9 days and 16.8 hours,
respectively. The dynamic modulus test results of the specimens fabricated after short-term aging
only are provided also for reference. The results indicate a significant increase in the dynamic

170 cm

Top Layer

Intermediate Layer

Bottom Layer

120 cm
Figure 33.   Area for percent damage definition.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   51  

70
Short-Term
60 Loose Mix, 95°C, 21 Days
Loose Mix, 135°C, 52 Hr

Percent Damage
50

40

30

20

10

0
0 50 100 150 200 250
Month

Figure 34.   Comparison of fatigue damage area


versus service life for FHWA ALF-SBS mixture
aged at different conditions.

modulus value with long-term aging. The results also indicate a significant difference in the
dynamic modulus master curves that correspond to the mixtures aged at 95°C for 8.9 days and
at 135°C for 16.8 hours, despite the mixtures’ equivalent binder rheology. A 10% threshold was
used to determine the significance of difference between the dynamic modulus test results at the
intermediate and high temperatures. The mixture aged at 135°C shows a reduction in modu-
lus value compared to the material aged at 95°C. These results suggest that the chemical changes
and/or other effects (e.g., absorption, drain-down) induced by aging at 135°C had a significant
effect on the performance of the SHRP AAD mixture. Note that the SHRP AAD binder has high

1E+5

1E+4
|E*| (MPa)

Short-term Aged
Oven, Loose Mix, 95°C, 8.9 days
1E+3
Oven, Loose Mix, 135°C, 16.8 hours

1E+2
1E-8 1E-6 1E-4 1E-2 1E+0 1E+2
Reduced Frequency (Hz)
(a)

1.0 5E+04

0.8 4E+04
Cumulative (1-C)

0.6 3E+04
C

0.4 2E+04

0.2 1E+04

0.0 0E+00
0.0E+0 4.0E+5 8.0E+5 1.2E+6 0E+00 5E+04 1E+05 2E+05
S Nf (Cycle)
(b) (c)

Figure 35.   SHRP AAD mixture performance test results: (a) dynamic modulus
curves, (b) C versus S curves, and (c) DR failure criterion lines.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

52   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

sulfur content and is highly structured, indicating high potential for changes between the oxidation
products of laboratory aging at 135°C versus 95°C.
Figure 35 (b) presents the damage characteristic curves for the SHRP AAD mixture. The
results show that the damage characteristic curves for the two long-term aging conditions are
both higher than the short-term aging damage characteristic curve, which is attributable to the
short-term aged mixture’s lower stiffness value (age level) compared to that of the long-term
aged materials. The same observation can be made between the two long-term aging tempera-
tures. The results indicate that the damage characteristic curves for the mixture aged at 135°C
are consistently lower than the curves for the mixture aged at 95°C, which is consistent with the
dynamic modulus test results that indicate that the mixture aged at 135°C is less stiff than the
mixture aged at 95°C. Moreover, the C value at failure, indicated by the end point of the C versus
S curve, is considerably higher for the mixture aged at 135°C than at 95°C. The average C value
at failure is 0.18 for the short-term aged material. The average C values at failure are 0.37 and
0.21 for the long-term aged material conditioned at 95°C and 135°C, respectively. The observed
difference in the C values at failure for the long-term aged materials indicates that the mixture
aged at 135°C is more brittle than the mixture aged at 95°C.
Figure 35 (c) presents the DR failure criterion results for the SHRP AAD mixture and shows
that the failure criterion line for the mixture that was long-term aged at 95°C is similar to that
for the short-term aged mixture. However, the failure criterion line for the mixture that was
long-term aged at 135°C is considerably lower than that for the short-term aged mixture, which
is indicative of less fatigue resistance. These results suggest that long-term aging at 135°C leads
to the degradation of fatigue resistance.
The results presented suggest that long-term aging at 135°C should be avoided due to negative
performance implications. In addition to the trends seen in the performance test results, visual
observations of fractured specimens also indicate changes between the SHRP AAD mixtures
aged at 95°C and 135°C. Table 11 presents the statistical analysis outcomes for the SHRP AAD
mixture performance test results. The pairs with significant differences (p < 0.05) are shaded in
the table. Both long-term aging conditions have a significant effect on the dynamic modulus

Table 11.   Statistical t-test analysis outcomes for dynamic


modulus and cyclic fatigue test results for SHRP AAD
materials.

Reduced Frequency
Sample 1.0E-05 1.0E-04 1.0E-03 1.0E-02
p-value
Short-term aged
Dynamic Loose mix, 95°C, 8.9 days 0.0055 0.0002 0.0058 0.0063
Modulus
Short-term aged
0.0264 0.0213 0.0188 0.0072
Loose mix, 135°C, 16.8 hours
Loose mix, 95°C, 8.9 days
0.0537 0.0338 0.0218 0.0043
Loose mix, 135°C, 16.8 hours
S
Sample 1.0E+05 2.0E+05 2.5E+05 3.0E+05
p-value
Short-term aged
Cyclic 0.0337 0.0062 0.0002 0.0006
Loose mix, 95°C, 8.9 days
Fatigue
Short-term aged
0.0093 0.0036 0.0037 0.0044
Loose mix, 135°C, 16.8 hours
Loose mix, 95°C, 8.9 days
0.4232 0.1304 0.0176 0.0134
Loose mix, 135°C, 16.8 hours

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   53  

0 1 0 1 0 1

1 0.9 1 0.9 1 0.9

2 0.8 2 0.8 2 0.8

3 0.7 3 0.7 3 0.7

4 0.6 4 0.6 4 0.6

Z (cm)
Z (cm)

Z (cm)
5 0.5 5 0.5 5 0.5

6 0.4 6 0.4 6 0.4

7 0.3 7 0.3 7 0.3

8 0.2 8 0.2 8 0.2

9 0.1 9 0.1 9 0.1

10 10 10 0
–1.5 –1 –0.5 0 0.5 1 1.5 –1.5 –1 –0.5 0 0.5 1 1.5 –1.5 –1 –0.5 0 0.5 1 1.5
X (m) X (m) X (m)

Short-Term Aged Loose Mix, 95°C, 8.9 Days Loose Mix, 135°C, 16.8 Hr

Figure 36.   Damage contours for SHRP AAD mix aged under different conditions.

values and fatigue performance for the SHRP AAD mixture. In spite of the matched binder
rheology at the 95°C and 135°C aging temperatures, the significant difference is observed in
the dynamic modulus and fatigue performance test data between the 95°C and 135°C aging
treatment. Note that, due to the very brittle nature of the SHRP AAD materials aged at 135°C,
the comparison is limited to low S values. As suggested by the t-test results, 95°C and 135°C have
different effects on the SHRP AAD materials.
Figure 36 presents the damage contours for the SHRP AAD mixture with different aging
conditions. The short-term aged material shows better performance than the long-term aged
asphalt materials. Notably, based on the damage contours, the asphalt mixture aged at 95°C
shows better performance than the material aged at 135°C; the same conclusion was drawn from
the material-level fatigue performance data as well.
Figure 37 presents the fatigue damage area versus ESALs for the SHRP AAD mixture with dif-
ferent aging conditions. A significant difference between the fatigue performance of the asphalt
mixture aged at 95°C and at 135°C is evident.

SHRP AAG.   Figure 38 shows the mixture performance test results for the SHRP AAG mix-
ture. Figure 38 (a) presents the SHRP AAG mixture dynamic modulus master curves that cor-
respond to the specimens fabricated after aging at 95°C and 135°C for 19 days and 37.6 hours,
respectively. The dynamic modulus test results of specimens fabricated after short-term aging

70
Short-Term
60 Loose Mix, 95°C, 8.9 Days
Loose Mix, 135°C, 16.8 Hr
Percent Damage

50

40

30

20

10

0
0 50 100 150 200 250
Month

Figure 37.   Comparison of fatigue damage area


versus service life for SHRP AAD mixture aged
under different conditions.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

54   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

1E+5

1E+4

|E*| (MPa)
Short-term Aged
Oven, Loose Mix, 95°C, 19 days
1E+3
Oven, Loose Mix, 135°C, 37.6 hours

1E+2
1E-8 1E-6 1E-4 1E-2 1E+0 1E+2
Reduced Frequency (Hz)
(a)

1.0 4E+04

0.8

Cumulative (1-C)
3E+04
0.6
2E+04
C

0.4
1E+04
0.2

0.0 0E+00
0.0E+0 4.0E+5 8.0E+5 1.2E+6 0E+00 5E+04 1E+05
S Nf (Cycle)
(b) (c)

Figure 38.   SHRP AAG mixture performance test results: (a) dynamic modulus
curves, (b) C versus S curves, and (c) DR failure criterion lines.

only also are provided for reference. The results indicate a significant increase in the dynamic
modulus value with long-term aging. The results also suggest similar dynamic modulus master
curves that correspond to mixtures aged at 95°C for 19 days and 135°C for 37.6 hours, with
slightly higher dynamic modulus values for the mixture aged at 135°C compared to 95°C. How-
ever, the binder rheology test results shown in Figure 38 suggest that the binder contained within
the mixture aged at 135°C is slightly stiffer than the binder contained within the mixture aged
at 95°C. Thus, it is difficult to ascertain whether or not the results suggest a change in mixture
performance as a result of laboratory aging at 135°C as opposed to 95°C. The results presented in
Figure 38 suggest that little chemical change is induced by aging SHRP AAG at 135°C. Hence,
any difference noted in performance that results from aging at 135°C as opposed to 95°C is
speculated to reflect changes other than chemistry (e.g., absorption, drain-down). Figure 38
(b) presents the C versus S fatigue damage characteristic curves that correspond to the SHRP
AAG mixture. The results show somewhat different trends in the C versus S curves for the two
long-term aging conditions, with the short-term aging C versus S curve falling considerably
lower than the long-term aged condition curves. Figure 38 (c) presents the failure criterion
results for the SHRP AAG mixture, showing that the failure criterion line for the mixture
long-term aged at 95°C is similar to that of the short-term aged mixture, but that the failure
criterion line for the mixture long-term aged at 135°C falls somewhat lower, which is indica-
tive of less fatigue resistance.
Table 12 presents the statistical analysis results for each pair of aging treatments. The p-values
indicate a significant difference between the performance of the short-term aged and long-term
aged SHRP AAG materials. [The pairs with significant differences (p < 0.05) are shaded in the
table.] However, although no significant difference can be observed between the dynamic mod-
ulus values of the material aged at 95°C and at 135°C, a significant difference between the fatigue

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   55  

Table 12.   Statistical t-test analysis outcomes for dynamic


modulus and cyclic fatigue test results for SHRP AAG materials.

Reduced Frequency
Sample 2.0E-06 1.0E-04 1.0E-03 1.0E-01
p-value
Short-term aged
Dynamic 0.0102 0.0084 0.0044 0.0109
Loose mix, 95°C, 8.9 days
Modulus
Short-term aged
0.0296 0.0120 0.0023 0.0049
Loose mix, 135°C, 16.8 hours
Loose mix, 95°C, 8.9 days
0.0928 0.1169 0.3271 0.2768
Loose mix, 135°C, 16.8 hours
S
Sample 1.0E+05 2.0E+05 3.0E+05 5.0E+05
p-value
Short-term aged
Cyclic 0.0082 0.0020 0.0001 0.0314
Loose mix, 95°C, 8.9 days
Fatigue
Short-term aged
0.0062 0.0001 0.0004 0.0235
Loose mix, 135°C, 16.8 hours
Loose mix, 95°C, 8.9 days
0.0161 0.0289 0.0415 0.0940
Loose mix, 135°C, 16.8 hours

test results for these two temperatures is evident. This observation suggests that the 135°C aging
temperature induces more brittle behavior than the 95°C aging temperature.
The FlexPAVE™ program results also suggest less fatigue resistance of material aged at
135°C compared to 95°C. The damage contours presented in Figure 39 indicate that more
severe fatigue cracking is associated with long-term aging at 135°C than at 95°C. Figure 40
also suggests a significant decrease in fatigue cracking resistance when the 135°C long-term
aging treatment is applied.

Proposed Long-Term Aging Procedure


Based on the findings from this study, oven aging of loose asphalt mixtures at 95°C is pro-
posed as the long-term aging procedure for fabrication of performance testing specimens.
Aging asphalt mixtures in a loose mix state expedites oxidation compared to compacted speci-
men aging under the same conditions. The performance test results indicate no problems with
loose mixtures compacted after long-term aging. Loose mixture aging at 95°C provides shorter

0 1 0 1 0 1

1 0.9 1 0.9 1 0.9

2 0.8 2 0.8 2 0.8

3 0.7 3 0.7 3 0.7

4 0.6 4 0.6 4 0.6


Z (cm)

Z (cm)
Z (cm)

5 0.5 5 0.5 5 0.5

6 0.4 6 0.4 6 0.4

7 0.3 7 0.3 7 0.3

8 0.2 8 0.2 8 0.2

9 0.1 9 0.1 9 0.1

10 10 0 10
–1.5 –1 –0.5 0 0.5 1 1.5 –1.5 –1 –0.5 0 0.5 1 1.5 –1.5 –1 –0.5 0 0.5 1 1.5
X (m) X (m) X (m)

Short-Term Aged Loose Mix, 95°C, 19 Days Loose Mix, 135°C, 37.6 Hr

Figure 39.   Damage contours for SHRP AAG mixture aged at different conditions.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

56   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

70
Short-Term
60 Loose Mix, 95°C,19 Days
Loose Mix, 135°C, 37.6 Hr
50

Percent Damage
40

30

20

10

0
0 50 100 150 200 250
Month

Figure 40.   Comparison of fatigue damage area versus


service life for SHRP AAG mixture aged at different
conditions.

durations of laboratory aging to reach the oxidation levels in the field than lower temperatures,
whereas aging temperatures higher than 100°C can result in changes in binder chemistry that
do not occur in the field.

Aging of WMA Mixtures


Loose mixture aging in the oven at 95°C is the proposed long-term aging method for HMA
mixtures. Although the compaction of long-term aged HMA loose mixtures typically is not prob-
lematic, some WMA technologies (e.g., foam) may not maintain their compaction-aiding prop-
erty if they are aged as loose mix. Consequently, a separate investigation of the feasibility of using
compacted specimen aging was conducted to select the optimal laboratory aging procedure for
WMA mixtures. Table 13 provides a summary of the Group B materials that were used in this
study to evaluate WMA aging. This evaluation of WMA aging includes data obtained from two
projects, Manitoba (MB) and NCAT. Both projects included a control HMA mixture. The results
for HMA compacted specimen aging indicate that the oxidation gradient within the specimen
is insignificant if the minimum sample dimension is equal to or less than 38 mm. Therefore,
compacted specimen aging trials were conducted on thin disk specimens prepared from gyratory-
compacted samples with a thickness of 25 mm. Compacted specimen aging was conducted at
85°C, which is the maximum temperature at which compacted specimen aging can be conducted
without causing damage or distortion of the sample. In order to evaluate the efficiency of com-
pacted specimen aging at 85°C compared to loose mixture aging at 95°C, both procedures also
were conducted using the NCAT-HMA mixture. The results were compared to those for binder
extracted and recovered from field cores obtained after 4 years in service to allow the determina-
tion of the time required to match 4 years of field aging.
Figure 41 (a) presents the G* values at 64°C and 10 rad/s of binder extracted and recovered
from the MB-HMA and MB-Evotherm (WMA) field cores. Figure 41 (b) presents the field core

Table 13.   Group B materials used for WMA aging evaluation.

Date Date Core


Site ID Location Binder / Modification
Built Extracted
Manitoba, Control HMA, Foam WMA,
MB 2010 2014
Canada Evotherm WMA
Alabama,
NCAT Control HMA, Foamed WMA 2009 2013
United States

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   57  

G*, 64°C, 10 rad/s (kPa) G*, 64°C, 10 rad/s (kPa)


1 2 3 4 5 0 10 20 30 40 50 60
0 0

10
10
20

Depth (mm)
Depth (mm)

30 20

40
MIT-HMA 30 NCAT-HMA
50
MIT-Evotherm NCAT-Foam
60 40
(a) (b)

Figure 41.   G* values at 64°C and 10 rad/s versus depth for (a) MB field cores and (b) NCAT
field cores (MIT = Manitoba Infrastructure and Transportation).

data obtained from the NCAT-HMA and NCAT-Foam (WMA) sections. These results show
that the HMA pavement section exhibits higher G* values than the WMA pavement section at
all depths where measurements were taken.
Figure 42 (a) shows the evolution of log G* at 64°C and 10 rad/s with laboratory aging dura-
tions for the MB compacted specimen aging trials conducted at 85°C. The results show that the
MB-HMA mixture has higher G* values than the MB-Foam and MB-Evotherm mixtures at a
given laboratory aging duration, suggesting that the effects of reduced short-term aging from
using WMA has long-term implications. However, the slopes of the WMA and HMA log G* versus
aging duration curves are similar. The laboratory aging finding of elevated age levels in HMA
compared to WMA matches the trends observed in the field core results presented in Figure 41.
The MB-Evotherm and MB-Foam mixtures exhibit a similar G* evolution, indicating that the
WMA technologies used did not significantly impact oxidative aging.
Figure 42 (b) shows the evolution of log G* at 64°C and 10 rad/s with laboratory aging dura-
tions for the NCAT compacted specimen aging trials conducted at 85°C. The NCAT-HMA loose
mixture aging results for tests conducted at 95°C are included for comparison. The results of
the NCAT mixtures match the findings of the MB mixtures; i.e., reduced short-term aging in
WMA compared to HMA leads to lower G* values at a given aging duration, but the slopes of the

0.8 2.5
MIT-HMA y = 0.0249x + 0.1151 2.3
log G*, 64°C, 10 rad/s (kPa)

log G*, 64°C, 10 rad/s (kPa)

MIT-Evotherm R² = 0.9825
2.1
0.6 MIT-Foam
1.9 y = 0.0675x + 1.1788
1.7 R² = 0.98

0.4 y = 0.0217x + 0.0935 1.5


y = 0.021x + 0.88
R² = 0.9935 1.3 R² = 0.94
1.1
0.2 y = 0.0181x + 0.1442
R² = 0.9995 0.9 Oven, Loose Mix, 95°C, NCAT-HMA
y = 0.0188x + 0.8397
Oven, Compacted, 85°C, NCAT-HMA
0.7 R² = 0.95
Oven, Compacted, 85°C, NCAT-Foam
0 0.5
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Duration (Days) Duration (Days)
(a) (b)

Figure 42.   Laboratory aging G* evolution results for (a) MB mixtures and (b) NCAT mixtures.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

58   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

compacted WMA and compacted HMA log G* versus the laboratory aging duration curves are
similar. The comparison of the loose and compacted specimen aging log G* versus aging dura-
tion curves for the NCAT-HMA mixture shown in Figure 42 (b) indicates a significant increase
in the efficiency of loose mixture aging at 95°C compared to compacted specimen aging at 85°C.
The results presented in Figure 42 combined with the log G* results for binder extracted
and recovered from field cores were used to determine the laboratory aging durations that
are required to match 4 years of field aging at a depth of 19 mm. The MB project did not include
any field sections that contained the Foam mixture, but the MB-Foam mixture was studied in the
laboratory to investigate the effect of foaming on the aging of MB and NCAT base binders. Table 14
presents the required laboratory aging durations. The results show that the required laboratory
aging durations for the WMA mixtures are slightly shorter than for the HMA for the same field
aging. Furthermore, the large difference in laboratory aging durations that are required to match
4 years of aging in the NCAT versus MB sections demonstrates that climate has a significant effect
on long-term aging. The required durations of compacted specimen laboratory aging at 85°C to
match just 4 years of field aging are long (i.e., roughly 2 to 5 weeks), making compacted speci-
men aging impractical. Compacted specimen aging at 85°C required 35.6 days to match 4 years
of field aging for the NCAT-HMA section. In contrast, only 6.7 days of loose mixture aging at
95°C were required to match the same condition. Therefore, it is proposed that loose mixture
aging at 95°C be used for laboratory long-term aging of both WMA and HMA mixtures. Future
research is needed to investigate the relationship between laboratory aging durations and field
aging for WMA materials and to resolve the compactability problems associated with WMA after
long-term aging. It is expected that the compaction of long-term aged WMA loose mixture will
require the use of HMA compaction temperatures. The compaction advantages of surfactant and
wax additives may be lost after long-term aging. Foaming will cease prior to the completion of the
long-term aging of foamed WMA loose mixtures.

Climate-Based Determination of Predefined Aging Durations


Kinetics Modeling of Loose Mix Aging
Determining the laboratory aging durations representing a given time in service, climate,
and depth within a pavement requires a solid understanding of binder oxidation kinetics and
diffusion. Kinetics modeling enables the prediction of the rate of change in asphalt properties
as a function of aging time and temperature. Therefore, kinetics models generally are calibrated
using experimental measurements of AIPs at various aging durations under isothermal aging
conducted at multiple temperatures. Diffusion models enable the prediction of oxygen partial
pressure within binder films. In this project, a rheology-based kinetics model was developed
based on existing chemistry-based kinetics models and AIP measurements obtained from the
laboratory oven-aged loose mixtures and field cores. However, the time and resources available
to this project did not allow the development of a fully validated diffusion model, which is a key
suggestion for future research.

Table 14.   Laboratory aging durations required to match 4 years


of in-service aging.
Required Compacted Required Loose Mixture
Depth
Field Section Specimen Aging Aging Duration at 95°C
(mm)
Duration at 85°C (days) (days)
NCAT-HMA 19 35.6 6.7
NCAT-Foam (WMA) 19 31.3 NA
MB-HMA 19 16.5 NA
MB-Evotherm (HMA) 19 16.1 NA

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   59  

All of the models presented herein were obtained when investigating the AIPs of asphalt
binders aged in thin films at elevated temperatures and/or under pressure. The oxidation of an
asphalt will be kinetics-controlled if the binder film thickness is thin enough to allow oxygen
to diffuse through the film faster than the rate of the reaction itself. It is critical that the film
thickness is sufficiently thin to ensure a kinetics-controlled reaction when calibrating a kinetics
model (Glaser et al. 2015). If the reaction is not kinetics-controlled, the diffusion effects will be
significant and, therefore, must be considered within the model, which adds complexity.
Glaser et al.’s (2013b) kinetics model framework discussed previously was applied to loose
mixture aging to facilitate linking laboratory aging durations to field aging; log G* at 64°C and
10 rad/s frequency was selected as the AIP because mixture performance relates more directly to
rheology than to chemistry, leading to Equation 15.

 kc 
log G * = log Go* + M  1 −  (1 − exp ( −k f t )) + kc Mt (15)
 kf 

where
G* = long-term aged binder shear modulus at 64°C and 10 rad/s (kPa) and
G*o = short-term aged binder shear modulus at 64°C and 10 rad/s (kPa).
Table 15 provides the detailed properties of the 10 mixtures that were used to calibrate and
validate the kinetics model presented in Equation 15. Because the binder source, aggregate
source, and filler type may affect loose mixture aging rates, the experimental plan was designed
to include a wide range of materials.
Initially, an experiment was designed to verify that loose mixture oven aging leads to a kinetics-
controlled reaction mechanism, which is a requirement for applying kinetics models without
considering diffusion. Three loose mixtures were prepared using component materials from the
WesTrack Fine sections with various binder contents (4.7%, 5.4%, and 6.1%). If loose mixture
oven aging leads to a kinetics-controlled oxidation reaction, then the binders in these three mix-
tures will oxidize at the same rate despite their various film thicknesses. However, if loose mixture
oven aging leads to a diffusion-controlled oxidation reaction, then the rate of oxidation will be
inversely proportional to the binder film thickness and thus to the binder content. After verifying
that loose mixture aging results in a kinetics-controlled reaction, isothermal loose mixture aging was
conducted using a broad range of mixtures at three temperatures: 95°C, 85°C, and 70°C. Asphalt
binder was extracted and recovered from the loose mixtures at various aging durations to measure
the rheological and chemical AIPs. The relationship between laboratory aging duration and log G*
at 64°C and 10 rad/s was derived from loose mixture aging conducted at 95°C and used to deter-
mine the parameter M in Equation 15 for all the study mixtures. Universal values of the reaction
parameters included in kf and kc were obtained from the least mean square error optimization of
Equation 15 to match the values of log G* at 64°C and 10 rad/s obtained from binders extracted
and recovered from five calibration mixtures (FHWA ALF-Control, FHWA ALF-SBS, SHRP AAD,
WesTrack Fine, and WesTrack Coarse) after various durations of laboratory aging at 95°C, 85°C,
and 70°C. The remaining mixtures in Table 15 were used to validate that the kf and kc parameters are
mixture-independent. Equation 15 was used to predict the values of log G* at 64°C and 10 rad/s cor-
responding to the various durations of loose mixture aging at 85°C and 70°C. The predicted log G*
values were compared against the measured data to evaluate the accuracy of the proposed model.
To validate the kinetics model further, two loose mixtures, FHWA ALF-Control and WesTrack
Fine with optimum binder content (5.4%), were subjected to non-isothermal aging. Then, the
ability of the kinetics model to predict the oxidation of the loose mixtures under non-isothermal
conditions was evaluated. Asphalt binder was extracted and recovered at several times during

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

60   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Table 15.   Mixtures used to calibrate and validate the kinetics model for loose
mix oven aging.

Binder
Modifier / Hydrated
Binder Binder Binder Aggregate
Project Mix MIX ID Anti- Lime
Source Grade Content Type
Stripping (Content)
Agent

Citgo- Limestone
ALF PG
ALF CTRL Venezuelan None 5.3% / Diabase 1%
Control 70-22
Bachaquero (traprock)
FHWA
ALF
Limestone
Mid PG
ALF SBS ALF SBS SBS-LG 5.3% / Diabase 1%
Continent 70-28
(traprock)
AAD-1
Binder + Limestone
SHRP California PG
FHWA SHRP AAD None 5.3% / Diabase 1%
Binder Coast 58-28
ALF (traprock)
Aggregate
4.7% Andesite /
Fine WesTrack - PG
West Coast None 5.4% Granite / 1.5%
Section Fine 64-22
6.1% Sand
WesTrack
Crushed
Coarse WesTrack - Idaho PG
None 5.7% Andesite / 1.5%
Section Coarse Asphalt 64-22
Sand

Citgo-
PG
NC DOT NCS9.5B NC Wilmington, None 6.6% Granite None
64-22
NC

South 120-150
LSD N/A None 5.9% N/A None
Dakota Pen

New
LNM N/A None AC-20 7.6% N/A None
Mexico
LTPP

Texas LTX N/A None AC-20 5.4% N/A None

Wisconsin LWI N/A None N/A 5.9% N/A None

the non-isothermal aging process and subjected to AIP testing. The measured AIPs were then
compared against those predicted by the kinetics model.

Evaluation of the Loose Mixture Aging Reaction Mechanism.   Loose mixture oven aging
was hypothesized to lead to a kinetics-controlled reaction mechanism because the film thick-
ness of asphalt binders in loose mixtures is typically less than 15 microns (Radovskiy 2003).
Glaser et al.’s (2015) evaluation of binder aging suggests that a film thickness of 15 microns will
lead to a kinetics-controlled reaction. The hypothesis that loose mixture aging is a kinetics-
controlled reaction is verified by the results presented in Figure 43, which shows the relationship
between laboratory duration and both chemical and rheological AIPs for the three WesTrack
Fine mixtures with various binder contents and, thus, film thicknesses. The results show equiva-
lent AIPs at various aging durations irrespective of the mixture film thickness, thereby verifying
the hypothesis that oven aging loose mixtures results in a kinetics-controlled oxidation reaction.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   61  

0.13 3.5
C+S = 0.0013(D) + 0.0884 log G* = 0.0552(D) + 1.6508
0.12 3
R² = 0.96 R² = 0.95
Absorbance Peaks (AU)

0.11
Carbonyl+Salfoxides

log G*, 64 °C, 10 rad/s


2.5
0.1
2

(kPa)
0.09
WT-Fine-%ac=6.1, Short-Term Aged 1.5 WT-Fine-%ac=6.1, Short-Term Aged
0.08 WT-Fine-%ac=6.1, Oven, Loose Mix, 95°C WT-Fine-%ac=6.1, Oven, Loose Mix, 95°C
WT-Fine-%ac=5.4, Short-Term Aged 1 WT-Fine-%ac=5.4, Short-Term Aged
0.07
WT-Fine-%ac=5.4, Oven, Loose Mix, 95°C WT-Fine-%ac=5.4, Oven, Loose Mix, 95°C
0.06 WT-Fine-%ac=4.7, Short-Term Aged 0.5 WT-Fine-%ac=4.7, Short-Term Aged
WT-Fine-%ac=4.7, Oven, Loose Mix, 95°C WT-Fine-%ac=4.7, Oven, Loose Mix, 95°C
0.05 0
0 10 20 30 0 10 20 30
Aging Duration (Days) Aging Duration (Days)

Figure 43.   Loose mix aging rates obtained from WesTrack Fine mix prepared with high (6.1%),
optimum (5.4%), and low (4.7%) binder contents: (a) aging rates in terms of C + S absorbance peaks and
(b) aging rates in terms of logarithm of binder shear modulus log G* at 64°C and 10 rad/s.

These results indicate that kinetics models can be derived from loose mixture aging results with-
out the need to consider diffusion.

Kinetics Modeling of Loose Mixtures under Isothermal Aging.   Table 16 presents the values
of the kinetics model parameters. The universal values of the reaction parameters included in
kf and kc (Eaf, Eac, Af, and Ac) were obtained from a least mean square error optimization of the
data obtained at 95°C from five mixtures (FHWA ALF-Control, FHWA ALF-SBS, SHRP AAD,
WesTrack Fine, and WesTrack Coarse), assuming M is equal to one. After optimizing the kf
and kc parameters, the material-dependent M values were determined by conducting least mean
square error optimizations using individual mixtures. It should be noted that M accounts for the
mix-specific aging kinetics. The table shows that the material-dependent parameter, M, varies
significantly (from 0.56 to 1.10) among the different materials evaluated. Higher M values mean
higher oxidation susceptibility.
Figure 44 presents comparisons between the measured and predicted log G* at 64°C and
10 rad/s values for the loose mixture aging temperatures of 70°C, 85°C, and 95°C for the mixtures

Table 16.   Kinetics model parameters using a single fitting parameter


and universal reaction rates.

Project Mix ID Af Eaf Ac Eac M


FHWA ALF ALF-CTRL 1.25×1013 95.04 3.68×107 62.21 0.743
ALF-SBS 1.25×1013 95.04 3.68×107 62.21 0.623
SHRP SHRP AAD 1.25×1013 95.04 3.68×107 62.21 1.104
Binder
WT-Fine 1.25×1013 95.04 3.68×107 62.21 0.871
WesTrack
WT-Coarse 1.25×1013 95.04 3.68×107 62.21 0.725
Validation Set
NCDOT NC 1.25×1013 95.04 3.68×107 62.21 0.937
LTPP-SD 1.25×1013 95.04 3.68×107 62.21 0.747
LTPP-NM 1.25×1013 95.04 3.68×107 62.21 0.546
LTPP
LTPP-TX 1.25×1013 95.04 3.68×107 62.21 0.88
LTPP-WI 1.25×1013 95.04 3.68×107 62.21 1.016

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

62   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

4
3.5 M = 0.743

log G*, 64°C, 10


3

rad/s (kPa)
2.5
2
1.5
1
0.5
(a) ALF Control
0
0 10 20 30 40
Aging Duration (Days)

4 4
3.5 M = 0.623 3.5 M = 1.104
log G*, 64°C, 10

log G*, 64°C, 10


3 3
rad/s (kPa)

rad/s (kPa)
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
(b) ALF SBS (c) SHRP AAD
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

4 4
3.5 M = 0.871 3.5 M = 0.725
3 log G*, 64°C, 10 3
log G*, 64°C, 10

rad/s (kPa) 2.5


rad/s (kPa)

2.5
2 2
1.5 1.5
1 1
0.5 0.5
(d) WesTrack Fine (e) WesTrack Coarse
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

Figure 44.   Comparisons between measured and predicted log G* values


for the calibration mixtures: (a) ALF-Control, (b) ALF-SBS, (c) SHRP AAD,
(d) WesTrack Fine, and (e) WesTrack Coarse.

used to calibrate the kf and kc parameters. The results demonstrate good agreement between the
measured and predicted values of log G*, thereby verifying the applicability of Equation 15 to
the kinetics modeling of loose mixture aging.
Figure 45 presents the comparisons between the measured and predicted log G* at 64°C and
10 rad/s values for the loose mixture aging temperatures of 70°C, 85°C, and 95°C for the validation
mixtures. To apply the kinetics model to the validation mixtures, Eaf, Eac, Af, and Ac derived from
the evaluation of the calibration mixtures were adopted. The material-dependent variable, M, was
optimized against the AIP measurements obtained from the various laboratory aging durations at
95°C. The results demonstrate very good agreement between the measured and predicted values,
thereby validating the use of Equation 15 coupled with the universal Eaf, Eac, Af, and Ac parameters.

Kinetics Modeling of Non-Isothermal Aging.   Field aging involves a variable temperature


history. Therefore, the ability of the kinetics model to predict the oxidative aging of loose mix-
tures under non-isothermal conditions was evaluated using two mixtures. Figure 46 (a) shows

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   63  

4
3.5 M = 0.937
3

log G*, 64°C, 10


rad/s (kPa)
2.5
2
1.5
1
0.5 (a) NCS9.5B
0
0 10 20 30 40
Aging Duration (Days)

4 4
3.5 M = 0.747 (b) LTPP-SD 3.5 M = 0.546 (c) LTPP-NM
log G*, 64°C, 10

3 3

log G*, 64°C, 10


rad/s (kPa)

rad/s (kPa)
2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

4 4
3.5 M = 0.88 3.5 M = 1.016
log G*, 64°C, 10

log G*, 64°C, 10

3 3
rad/s (kPa)

rad/s (kPa)

2.5 2.5
2 2
1.5 1.5
1 1
0.5 0.5
(d) LTPP-TX (e) LTPP-WI
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

Figure 45.   Comparisons between measured and predicted log G* values for
the validation mixtures: (a) NCS9.5B, (b) LTPP-SD, (c) LTPP-NM, (d) LTPP-TX, and
(e) LTPP-WI.

the non-isothermal aging history applied to the loose mixture specimens. The temperatures
were varied between 40°C and 70°C over the course of 43 days. Equation 15, with the mixture-
specific M value derived from laboratory aging at 95°C, was used to predict the AIP evolution
under the non-isothermal conditions that were applied to the loose mixtures. The predicted AIP
values were compared with those measured from the binders that were extracted and recovered
at the times (in days) indicated in Figure 46 (a). Figure 46 (b) through Figure 46 (d) present
comparisons between the measured and predicted AIP values. In general, the predicted and mea-
sured values for log G* are in good agreement. These results suggest that the presented kinetics
framework can be coupled with the field pavement temperature history and a diffusion model
to predict the evolution of asphalt binder AIPs within a pavement over its service life.

Kinetics Derivation of Laboratory Aging Duration


Based on the successful prediction of non-isothermal aging, the kinetics modeling was
extended to predict the required duration of loose mixture aging at 95°C in the laboratory to

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

64   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

80 40
6D Loose Mix Sampling 35 WesTrack, Fine
70
30

G*, 64°C, 10 rad/s


Predicted
Temperature (°C)

10 D
60 25 Measured
1D 13 D 20
50 15
6D 2% Error
10
40
10% Error 23% Error
2D 5D 5
30 0
0 5 10 15 20 25 30 35 40 45 16D 33D 43D
Aging Time (Days) Aging Duration
(a) (b)

60 60
ALF Control
50 y = 1.0035x - 1.9924

Measured log G*, 64°C,


50
R² = 0.99
G*, 64°C, 10 rad/s

Predicted
40 40
Measured

10 rad/s
30 30

20 lity
20 6% Error q ua
fE WesTrack, Fine
10 eo
10 8% Error 2% Error Lin ALF Control
0
0 0 20 40 60
16D 33D 43D
Predicted log G*, 64°C, 10 rad/s
Aging Duration
(d)
(c)
Figure 46.   Predictions of non-isothermal loose mixture oven aging: (a) non-isothermal laboratory aging
history, (b) WesTrack Fine mix prediction, (c) ALF-Control mix prediction, and (d) overall prediction quality.

match field aging as a function of pavement temperature history and depth. It is important
to note that loose mixture oven aging is kinetics-controlled whereas field aging is diffusion-
controlled. Thus, to link laboratory aging to field aging in a rigorous way, the kinetics model
would need to be coupled with a diffusion model. However, a suitable diffusion model is not
available. Furthermore, current diffusion model frameworks require detailed mixture mor-
phology information that may make the use of a diffusion model to determine the laboratory
aging duration required to match the field aging impractical. Therefore, in lieu of a diffusion
model, the kinetics modeling framework was simplified and calibrated against field data that
corresponded to a wide range of climatic conditions and depths.

Derivation of Laboratory Aging Duration Using the Kinetics Model.   To determine the
required laboratory aging duration at 95°C that is required to match a given field age level using
kinetics modeling, Equation 15 must be written in terms of laboratory and field thermal histories,
which are expressed by Equation 16 and Equation 17, respectively. Within these kinetics expres-
sions, the temperature history and corresponding temperature dependence of the reaction rate are
predicted by kf and kc , whose parameters are universal. In addition, Equation 16 and Equation 17
include the log G* values that correspond to short-term aging. To determine the required labora-
tory aging duration at 95°C to match a given field temperature history, G*lab and G*field are equated,
as shown in Equation 18. The time-dependent pavement temperature history is input into the field
side of Equation 18 and 95°C is input as the laboratory aging temperature that allows the labora-
tory aging time (tlab) to be solved. Equation 18 demonstrates that the material-dependent param-
eter, M, appears on both the lab and field sides of Equation 18 and, therefore, cancels itself out.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   65  

Furthermore, if the short-term aging of the field and laboratory mixtures is assumed to be equal
(i.e., G*o,lab = G*o,field), then Equation 18 reduces to Equation 19. Equation 19 shows that the labora-
tory oven aging duration that is required to match field aging for a given mixture is independent of
the G* of the short-term aged material and M. In other words, the laboratory aging duration that
is required to match a given field condition is independent of the material-specific kinetics. This
finding is significant because it indicates that mixture-specific kinetics model parameters are not
required for the determination of laboratory aging durations.

 k 
* = log Go*,lab + M  1 − c  (1 − exp ( −k f t )) + kc t 
log Glab (16)
  kf  lab

 kc  
log G *field = log Go*,field + M  1 −  (1 − exp ( −k f t )) + kc t  (17)
 kf   field

 kc  
log Go*,field + M  1 −  (1 − exp ( −k f t )) + kc t 
  k f   field

 kc  
= log Go*,lab + M  1 −  (1 − exp ( −k f t )) + kc t  (18)
 kf  lab

 kc    kc  
 1 − k  (1 − exp ( −k f t )) + kc t  =  1 − k  (1 − exp ( −k f t )) + kc t  (19)
 f  field  f lab

Laboratory and Field Validation.   Laboratory and field data were used to validate that
the laboratory aging duration required to match a given field condition is independent of the
mixture-specific kinetics. The laboratory non-isothermal aging experiment presented in Fig-
ure 46 was used first for validation. Isothermal aging was conducted at 95°C and binder was
extracted and recovered at different times to determine the G* value at 64°C and 10 rad/s for
both the FHWA ALF-Control and WesTrack Fine mixtures. The results were used to determine
the laboratory aging duration at 95°C that is required to match the G* value at 64°C and 10 rad/s
of binder extracted and recovered during the non-isothermal aging experiments after 16, 34,
and 44 days of aging. If the kinetics model derivation is valid, then the laboratory aging duration
at 95°C that is required to match the age level in the non-isothermal experiment should be the
same for the FHWA ALF-Control and WesTrack Fine mixtures, despite their differing oxidation
susceptibilities, which are based on their material-dependent M values (FHWA ALF-Control,
M = 0.743 versus WesTrack Fine, M = 0.871). Figure 47 presents comparisons between the

3 3
ALF - Control
ALF-Control (Days)
Duration at 95°C,

2.5 2.5
Duration at 95°C
Measured Aging

Measured Aging

WesTrack - Fine1995 y = 0.8665x + 0.2657


2 2 R² = 0.96
(Days)

1.5 1.5
1 1
0.5 0.5
0 0
0 20 40 60 0 0.5 1 1.5 2 2.5 3
Non-Isothermal Oven Aging Duration Measured Aging Duration at 95°C,
(Days) WesTrack-Fine (Days)
(a) (b)

Figure 47.   Laboratory non-isothermal aging validation: (a) measured aging


duration for non-isothermal aging of ALF-Control and WesTrack Fine mixtures and
(b) measured field aging for ALF-Control and WesTrack Fine mixtures.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

66   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

laboratory aging durations at 95°C that are required to match the oxidation level at several times
during the non-isothermal aging experiment. The results indicate that the aging duration at
95°C that is required to match the level of oxidation induced by the non-isothermal experi-
ment is the same for the FHWA ALF-Control and WesTrack Fine mixtures, thereby validating
the use of Equation 19 for equating laboratory aging to field aging.
The FHWA ALF field data also were used to validate the kinetics derivation that was imple­
mented to determine the laboratory aging durations. Asphalt binder was extracted and recovered
from various depths of field cores extracted after 8 years of service from FHWA ALF-Control
and SBS pavement sections in McLean, Virginia. The G* values at 64°C and 10 rad/s of the
field-aged binders were measured. Isothermal laboratory loose mixture aging was conducted
at 95°C and binder was extracted and recovered at various times to determine the G* value at
64°C and 10 rad/s for both the FHWA ALF-Control and SBS-modified mixtures. These data
were then used to determine the duration of laboratory aging that is required to match the
aging level of the field core. The measured laboratory aging durations at 95°C that are needed
to match the field aging level at various depths were then determined.
Figure 48 presents comparisons of the laboratory aging durations that are required to
match the field core age level based on the G* values at 64°C and 10 rad/s values for the
FHWA ALF-Control and SBS-modified mixtures. The FHWA ALF-Control (M = 0.743) and
SBS-modified (M = 0.623) mixtures exhibited significantly different kinetics model M val-
ues. However, the results demonstrate that the laboratory aging durations that are required
to match the field age levels are similar for the two mixtures, despite their differing kinetics
behavior.
It should be noted that the data presented in Figure 47 were obtained solely from
loose mixture aging, which is a kinetics-controlled reaction, whereas field oxidation is

(a)

Laboratory Aging Duration to 30


Match Field Aging (Days) 6 mm depth
ALF Control Measured

0 5 10 15 20 25 25 19 mm depth
Duration (Days)

0 54 mm depth
20
Pavement Depth (mm)

10
15
20
10
30
5 y = 1.0277x + 0.8943
40 R² = 0.99
ALF Control 0
50 0 5 10 15 20 25 30
ALF SBS
60 ALF SBS Measured Duration (Days)
(b) (c)

Figure 48.   FHWA ALF field case study validation: (a) FHWA ALF location in McLean,
VA (b) measured field aging at different depths, and (c) measured field aging for
ALF-Control and ALF-SBS–modified mixtures.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   67  

diffusion-controlled. Figure 48 demonstrates that the duration of laboratory aging that is


required to match field aging is the same for the FHWA ALF-Control and SBS-modified
mixtures. However, these mixtures were constructed using the same aggregate structure,
aggregate type, target air void content, and binder content. Thus, the diffusion proper-
ties of the mixtures would be expected to be similar, and hence, this finding might not
extend to all scenarios. It is not practical to consider diffusion in prescribing laboratory
aging durations due to unknown morphological properties, such as the in situ density of
asphalt layer.

Development of Climatic Aging Index (CAI)


The use of Equation 19 to determine laboratory aging durations as a function of pavement
temperature history is computationally expensive. Furthermore, the kinetics model expressed
in Equation 19 does not consider the diffusion that affects oxidation levels within pavements.
Therefore, the application of Equation 19 to determine laboratory loose mixture aging dura-
tions would likely be inaccurate without either the inclusion of a diffusion model or calibra-
tion against the field data. To overcome these challenges, the kinetics model was simplified
and calibrated against field data that correspond to a wide range of climatic conditions and
depths in order to develop a CAI to determine laboratory aging durations using pavement
temperature history.
Equation 20 shows the kinetics model for the prediction of log G* as a function of pavement
temperature and pressure history.

 kc . P n  
log G *field = log G *o,field + M   1 − m
(1 − exp ( −k f P mt )) + kc P nt  (20)
 kf .P   field

The required inputs of this model are pavement temperature history, the mixture-specific
kinetics parameter M, and the log G* that corresponds to the short-term aged condition.
Although the effect of the oxidation spurt is clearly evident in laboratory aging that is con-
ducted at an elevated temperature, as shown in Figure 44 and Figure 45, its effect on pave-
ment temperature is relatively small because pavement temperatures do not go as high as
85° and 95°C as shown in Figure 44 and Figure 45. It is noted that the log G * and aging
duration relationship is almost linear at 70°C even though that at 85°C and 95°C show the
nonlinear relationship with the initial spurt. Therefore, when Equation 20 is used to predict
log G* as a function of time using pavement temperature history data, the effect of the fast
rate oxidation spurt is negligible, which allows Equation 20 to be simplified to the form
given in Equation 21.

log G *field = log G0,* field + M × kc × P n × t (21)

To further support the simplification of Equation 20 to Equation 21, kc and kf are shown as
functions of temperature in Figure 49. It can be seen that, at temperatures below 15°C, kf is very
small, which gives the exponential term in Equation 20 a value that is close to one. Consequently,
at temperatures below 15°C, the red portion of Equation 20 tends towards zero. At temperatures
between 15°C and 60°C, kc /kf is relatively close to one, which causes the red term in Equation 20
to be close to zero. Hence, at typical pavement temperatures, the red term in Equation 20 tends
to be zero and Equation 20 can be simplified to Equation 21.
Equation 19 demonstrates that the determination of laboratory aging durations at 95°C
is independent of the mixture-specific parameters, M and G*o. Thus, with regard to the duration
of laboratory aging, the pertinent parameters in Equation 21 are k (rate of reaction) and P n

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

68   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

1E+00
(kc /kf) 1

Fast and Slow Reaction


1E-01

Rates (log G*/day)


1E-02
kf 0
1E-03

1E-04

1E-05 Fast Reaction Rate (kf)


Slow (Constant) Reaction Rate (kc)
1E-06
0 20 40 60 80 100
Temperature (°C)

Figure 49.   Fast and slow reaction rates at


different temperatures.

(the oxygen pressure term). The P n parameter is affected by altitude and pavement depth. Because
most of the pavement sections included in Group B that were used in the climatic index study
are not in locations with high altitudes, the P n parameter is affected largely by pavement depth.
Therefore, the P n parameter is replaced by the D term in Equation 22 to define the CAI, where D is
an empirical depth correction factor to account for the differences in oxygen partial pressure, P n,
with pavement depth. The CAI in Equation 22 is now defined by the rate of reaction (k), reaction
time (t), and pavement depth (represented by D).

CAI = D × k × t (22)

The slow reaction rate term, k in Equation 22, can be represented using Arrhenius equation
parameters, as shown in Equation 23, where the units of time are days. In order to consider
hourly temperature data, a summation of the CAI values is applied based on the hourly tem-
perature history, as shown in Equation 24.

CAI = D × A × exp − ( ) Ea
RT
×t (23)

t oven = CAI = ∑ i =1 D × A × exp  −


N Ea 
24 (24)
 RTi 

where
toven = required laboratory aging duration at 95°C to match field aging (day),
CAI = climatic aging index,
D = depth correction factor,
A = frequency (pre-exponential) factor (unit-less),
Ea = activation energy (kJ/mol),
R = universal gas constant (kJ/mol K), and •

T = hourly pavement temperature obtained from EICM at the depth of interest (Kelvin).
The CAI is an empirical index that is employed to represent the extent of oxidation. To apply the
CAI to determine the loose mixture aging duration at 95°C to match a given pavement tempera-
ture history and depth of interest (toven), the laboratory aging duration needed to match the level
of field aging was first determined experimentally using Group B materials. Loose mixture aging
was conducted using the Group B materials at 95°C, and binders were extracted and recovered

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   69  

at periodic intervals and from corresponding field cores at various depths. The laboratory aging
duration that is required to match the level of aging of the field cores was determined subsequently
based on matching the G* values at 64°C and 10 rad/s. The pavement temperature history acquired
from EICM data for the Group B sections was then input into Equation 24. For each section, CAI
was calculated for pavement temperatures greater than 20°C. Determination of 20°C as the mini-
mum temperature for aging in the field was made based on the results shown in Figure 50. This
figure shows the prediction of log G* using the kinetics model for two binders with vastly different
aging susceptibilities (SHRP AAD with M = 1.104 and ALF-SBS with M = 0.623). The predictions
show that the two binders do not experience any aging when the temperature is 20°C for prolonged
periods of time despite their different aging susceptibilities. Therefore, it is safe to assume that no
aging occurs when pavement temperatures in the field fall below 20°C. The parameters Ea and A
were regressed to provide the best fit between CAI values and measured laboratory aging durations
without depth correction.
Figure 51 (a) shows the relationship between the CAI values and measured laboratory dura-
tions at 95°C (toven) without applying the depth correction factor D. The figure shows significant
scatter in the data and an overall low R2 value. Figure 51 (b) presents the CAI calibration sepa-
rately for three depths: the near-surface layer (6 mm from the surface), 20 mm from the surface,
and deeper layers (below 20 mm). Figure 51 (b) also shows that separating the data according to
depth improves the relationship between the CAI values and laboratory aging durations signifi-
cantly, thereby highlighting the need for the depth correction. Thus, the Ea and A parameters
were first calibrated using the data that correspond to a depth of 6 mm to provide a CAI value
that is equivalent to the required laboratory aging duration at 95°C to match the pavement
temperature history. Then, the D values were calibrated using data that correspond to depths
of 20 mm and deeper. It is worth mentioning that the depth correction factor (D) is not only
affected by diffusion mechanisms but also by photo-oxidation effects at the pavement surface.
Figure 51 (c) shows the relationship between the laboratory aging durations and CAI values
after applying the depth correction factors shown in Table 17. As shown in Figure 51 (c), the
CAI values and measured durations that are needed to match field aging correlate linearly in a
one-to-one relationship. Thus, the CAI values represent the required duration at 95°C that is
needed to match the field aging for a given pavement temperature history and depth (toven), as
shown in Equation 24.

25 25
SHRP AAD ALF SBS
20 T=95°C 20 T=95°C
T=70°C T=70°C
T=60°C T=60°C
15 T=50°C 15 T=50°C
logG*
logG*

T=40°C T=40°C
T=30°C T=30°C
10 T=20°C 10 T=20°C

5 5

0 0
0 100 200 300 400 0 100 200 300 400
Duration (days) Duration (days)
(a) (b)

Figure 50.   Predictions of log G* under different isothermal loadings for prolonged periods of time for:
(a) SHRP AAD binder and (b) ALF-SBS–modified binder.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

70   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

35
WesTrack-Fine WesTrack-Coarse
30 ALF-Control ALF-SBS

Measured Duration at
LTPP-NM LTPP-SD
25
LTPP-TX LTPP-WI
95°C (Days) 20 y = 0.4944x
R² = 0.28
15

10

0
0 5 10 15 20 25 30 35
Climatic Aging Index, CAI
(a)

35
Surface Layer (6 mm)
30 20 mm depth
Measured Duration at

Deeper Layers (below 20 mm)


25
95°C (Days)

20 y = 0.4565x
y = 1x R² = 0.67
15
R² = 0.56
10
y = 0.2967x
5
R² = 0.70
0
0 5 10 15 20 25 30 35
Climatic Aging Index, CAI
(b)

35
Surface Layer (6 mm)
30 20 mm depth
Measured Duration at

Deeper Layers (below 20 mm)


25
95°C (Days)

20

15 y = 1x
R² = 0.7142
10

0
0 5 10 15 20 25 30 35
Climatic Aging Index, CAI
(c)

Figure 51.   CAI predictions: (a) overall CAI fitting without


depth correction factor D, (b) CAI fitting based on layer
depth without depth correction factor D, and (c) overall CAI
fitting after applying depth correction factor (D).

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   71  

Table 17.   CAI fitting coefficients.

Arrhenius Equation,
Depth Correction Arrhenius Equation,
MIX ID Pre-exponential
Factor (D) Activation Energy Ea
Factor (A)
Surface Layer (6 mm) 1.0000 1.40962 13.3121
20-mm depth 0.4565 1.40962 13.3121
Deeper Layers
0.2967 1.40962 13.3121
(below 20 mm)

It should be noted that climatic indices have been used in other applications to quantify
the effect of temperature history on aging. The GAS model uses the MAAT to predict G* as
a function of time and depth in Pavement ME Design. The cumulative degree days (CDD)
concept, defined as the sum of the daily high temperatures above freezing, was proposed as
a potential means to link laboratory aging durations to an equivalent pavement temperature
history in NCHRP Project 09-49 and NCHRP Project 9-52 (Newcomb et al. 2015, Martin
et al. 2014). The use of MAAT and CDD as aging indices was tried in this research, as well as
prescribing laboratory aging durations as a function of the high temperature PG. The analysis
of alternative aging indices to relate laboratory aging durations to field conditions are pre-
sented in Appendix H. In summary, the following advantages of the CAI over the alternative
aging indices were found:
1. The CAI is based on pavement temperature rather than air temperature.
2. The CAI can be applied at different depths and uses a depth correction factor (D) for variations
in pavement temperature with depth.
3. The CAI considers the exponential relationship between aging rate and temperature.
4. The CAI captures hourly pavement temperatures.

Laboratory Aging Duration Maps


Using the coefficient values given in Table 17, Equation 24 was used to calculate the CAI val-
ues for various locations in the United States using hourly pavement temperature history data
obtained from the EICM to provide an overview of the proposed laboratory aging durations for
various climates, field aging durations, and depths. Laboratory aging durations were calculated
for three field ages: 4 years, 8 years, and 16 years. For each field age, the laboratory aging durations
were determined at three depths (6 mm, 20 mm, and 50 mm) and rounded to the nearest day:
1. Determine the field aging duration to be simulated in the laboratory.
2. Determine the location of the pavement section of interest.
3. Obtain the air temperature history for the location of interest.
4. Run the EICM to obtain the hourly pavement temperature. Note that running the EICM
requires additional inputs, such as the material and structural properties of the pavement
section of interest.
5. Determine the depth in the pavement where aging is to be simulated; then, find the value of
D according to Table 17. Note: The values of A and Ea also are given in Table 17.
6. Calculate the CAI value by adding together the CAI values calculated for each hour of pave-
ment service life for temperatures greater than 20°C, in accordance with Equation 24. Note
that for simplicity, the CAI value can be calculated using hourly data for one year and then
multiplied by the number of years of interest.
7. For practicality, round toven to the nearest day because the CAI value represents the required
oven aging duration at 95°C that is needed to match the field aging for a given pavement
temperature history and depth (toven).
Figure 52 shows the CAI-determined loose mixture aging durations at 95°C that are required
to match 4, 8, and 16 years of field aging at a depth of 6 mm. Figure 51 demonstrates that

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

(a)

(b)

Figure 52.   Required oven aging duration at 95°C to match level of field aging 6 mm below pavement surface
for (a) 4 years of field aging and (b) 8 years of field aging.
(continued)

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   73  

(c)

Figure 52.   (Continued) Required oven aging duration at 95°C to match level of field aging 6 mm below
pavement surface for (c) 16 years of field aging.

climate has a significant effect on the required laboratory aging duration that is required to
match a given field age. For example, as shown in Figure 52 (c), replicating 16 years of aging in
Hawaii requires 40 days of oven conditioning, whereas only 4 to 5 days of oven conditioning are
required to replicate the same duration of aging in Alaska. The evaluation of asphalt mixtures
that have been conditioned to replicate aging at a depth of 6 mm may be useful for evaluating
and predicting top-down fatigue and thermal cracking in pavements.
Figure 53 shows the CAI-determined loose mixture aging durations at 95°C that are required
to match 4, 8, and 16 years of field aging at a depth of 20 mm. A comparison between the labo-
ratory aging durations presented in Figure 52 and Figure 53 demonstrates that significantly
shorter laboratory aging durations are required to match the field aging at a depth of 20 mm
compared to 6 mm, indicating that the temperature gradient and diffusion in pavements sig-
nificantly affect oxidation levels. On average, the required laboratory aging durations that are
needed to match field aging at a depth of 20 mm are 45% shorter than the durations required
to match field aging at a depth of 6 mm. Also, significantly longer laboratory aging durations
are required to match field age levels at a depth of 20 mm for warmer climates compared to
colder climates.
A depth of 20 mm represents a reasonable depth for the evaluation of surface layer asphalt mix-
tures because it better reflects bulk behavior within a pavement structure than nearer the surface and
avoids the effect of ultraviolet (UV) oxidation. Furthermore, the laboratory aging durations that are
required to reflect field aging are much shorter and, therefore, more practical for the 20-mm depth

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

(a)

(b)

Figure 53.   Required oven aging duration at 95°C to match level of field aging 20 mm below pavement surface
for (a) 4 years of field aging and (b) 8 years of field aging.
(continued)

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   75  

(c)

Figure 53.   (Continued) Required oven aging duration at 95°C to match level of field aging 20 mm below
pavement surface for (c) 16 years of field aging.

compared to the 6-mm depth. Also, it should be noted that the recently developed small speci-
men geometry for asphalt mixture performance testing consists of Ø38-mm diameter specimens.
Therefore, if field cores are to be evaluated to complement laboratory investigations, the center of
horizontal cores extracted from the pavement surface would be at a depth of approximately 20 mm.
Figure 54 shows the CAI-determined loose mixture aging durations at 95°C that are required
to match 4, 8, and 16 years of field aging at a depth of 50 mm. These results demonstrate that
considerably shorter aging durations are required to simulate aging at a depth of 50 mm com-
pared to depths of 20 mm and 6 mm, thus indicating the presence of a significant oxidation gra-
dient with depth near the surface of the pavements. In Figure 54, the required aging duration of
zero days in a few cold northern states means that long-term aging at 50 mm below the pavement
surface in these cold regions are not significant enough to require long-term oven conditioning
to mimic field conditions.
Figure 55 and Figure 56 present the gradients in terms of age level with depth measured in field
cores acquired from in-service pavements. Figure 55 shows that field aging is more or less constant
below a depth of 50 mm. It is also noted that the measured log G* values below 50 mm depth are
considerably larger than those from the laboratory short-term aged materials, which are shown in
Figure 55 at the pavement surface. Different from the GAS model’s assumption that long-term aging
below 50 mm depth is negligible, it can be concluded from this study that long-term aging does
take place below 50 mm depth, but does not change appreciably. Consequently, the evaluation of
asphalt mixtures that have been prepared to match field aging at a depth of 50 mm could be useful
for evaluating intermediate and base asphalt layers that play a critical role in bottom-up cracking.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

(a)

(b)

Figure 54.   Required oven aging duration at 95°C to match level of field aging 50 mm below pavement surface
for (a) 4 years of field aging and (b) 8 years of field aging.
(continued)

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   77  

(c)

Figure 54.   (Continued) Required oven aging duration at 95°C to match level of field aging 50 mm below
pavement surface for (c) 16 years of field aging.

log G*, 64°C, 10 rad/s (kPa)


0.0 0.5 1.0 1.5 2.0 2.5 3.0
0

50
Depth (mm)

100

150
WesTrack Fine - Short-Term Aged
WesTrack Fine - 19 Years Old
200
FHWA ALF Control - Short-Term Aged
FHWA ALF Control - 8 Years Old
250

Figure 55.  Log G* values for measured field gradient


throughout pavement depth compared to laboratory
short-term aged materials for FHWA ALF-Control and
WesTrack Fine section with optimum asphalt content (%ac)
and high air void content (%Va).

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

78   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

log G*, 64°C, 10 rad/s (kPa)


0.0 0.5 1.0 1.5 2.0 2.5 3.0
0

50

Depth (mm)
100

150 FHWA ALF Control - 8 Years Old


FHWA ALF SBS - 8 Years Old
LTPP South Dakota - 21 Years Old
200 LTPP New Mexico - 18 Years Old
LTPP Wisconsin - 17 Years Old
WesTrack Fine - 19 Years Old
250
(a)

log G*, 64°C, 10 rad/s (kPa)


0.5 1.0 1.5 2.0 2.5 3.0 3.5
0

20
Depth (mm)

40

Low %ac and High %Va - 19 Years Old


60 Low %ac and Medium %Va - 19 Years Old
Optimum %ac and High %Va - 19 Years Old
Optimum %ac and Medium %Va - 19 Years Old
80 Optimum %ac and Low %Va - 19 Years Old
High %ac and Medium %Va - 19 Years Old
High %ac and Low %Va - 19 Years Old
100
(b)

Figure 56.   Measured log G* values as a function of pavement


depth for (a) field cores obtained from various locations in the
United States and (b) field cores obtained from 19-year-old
WesTrack Fine sections constructed with different binder
contents and air voids.

Aging Model to Predict Field Aging Throughout Pavement Depth


Prediction of Field Aging Using Mix-Specific Kinetics Parameters
The kinetics model developed herein could potentially be used as the basis for improving
the prediction of changes in asphalt binder properties with oxidative aging within pavement
performance prediction frameworks, including Pavement ME Design. To evaluate the ability
of the kinetics model to predict the evolution of oxidative aging in pavements, Equation 25
was applied to predict the log G* at 64°C and 10 rad/s in the field using hourly pavement
temperature history data at varying depths obtained from the EICM for the eight mixtures
detailed in Table 18.

* _ Predicted = log Go*, field + M  1 − kc  (1 − exp ( −k f t )) + kc t 


log G field (25)
 kf  
  field

Figure 57 and Figure 58 present the results of the kinetics model predictions at various depths
of the pavements. The predicted log G* values were compared to corresponding measurements

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   79  

Table 18.   Mixtures used to calibrate field aging kinetics model predictions.

Modifier /
MIX Section Binder Anti- Binder Binder
Project Mix Field Age
ID Location Source Stripping Grade Content
Agent
Anti-
Citgo-
ALF- ALF- McLean, Stripping PG 70-
8 Years Venezuelan 5.3%
Control CTRL VA Agent 22
Bachaquero
FHWA (0.001)
ALF
ALF- McLean, Mid PG 70-
ALF-SBS 8 Years None 5.3%
SBS VA Continent 28

South LTPP- 120-150


SD 21 Years N/A None 5.9%
Dakota SD pen

New LTPP-
NM 11, 18 Years N/A None AC-20 7.6%
Mexico NM
LTPP
18 Years
LTPP-
Texas TX (15 Years before N/A None AC-20 5.4%
TX
microsurfacing)*
LTPP-
Wisconsin WI 8, 17 Years N/A None N/A 5.9%
WI

Fine WT- Dayton, West Coast PG 64- 5.4%


0, 4, 19 Years None
Section Fine NV Refinery 22 (Optimum)
WesTrack
Coarse WT- Dayton, Idaho PG 64- 5.7%
2 Years None
Section Coarse NV Asphalt 22 (Optimum)

*It was assumed that no aging occurred within the asphalt concrete layers after microsurfacing placement.

of log G* obtained for binder extracted and recovered from field cores. Note that, although the
determination of the required laboratory aging duration that is needed to match field aging for a
given mixture could be derived using the kinetics model without the mixture-specific parameters,
i.e., G*o and M, the prediction of log G* at pavement depths requires these mixture-specific kinet-
ics parameters. The mixture-specific parameter M is determined from the isothermal aging of
loose mix at 95°C. It is assumed that the short-term aging in the field and laboratory is the same
(i.e., G*o,lab = G*o,field).
Because the kinetics model accounts for only the temperature dependence of the oxida-
tion reaction, Equation 25 is not expected to yield accurate predictions of field aging. In the
field, the partial pressure of the oxygen that is available to the binder will vary with depth,
and thus, diffusion is expected to affect the field reaction rates significantly. In addition, UV
aging is expected to contribute to oxidative aging near the pavement surface, which is not
considered within the kinetics model that is calibrated using thermal oxidation laboratory
aging. Therefore, the kinetics model was calibrated against field core measurements of log
G* at a depth of 20 mm, which represents a reasonable depth for the evaluation of the bulk
behavior of surface layers and avoids the effect of UV oxidation at the pavement surface. The
calibration was accomplished using Equation 26. The C1 and C2 parameters were determined
by linear regression between the kinetics model predictions and field core measurements of
log G* at 20 mm depth.

* _ calibrated = C1 + C2 × log G field_predicted


log G field * (26)

where
C1, C2 = calibration factors.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

80   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

log G*, 64°C, 10 rad/s (kPa) log G*, 64°C, 10 rad/s (kPa)
0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.0 0.5 1.0 1.5 2.0 2.5
0 0
20 10
40 20
60 30
Depth (mm)

Depth (mm)
80 40
100 50
120 60
140 70
160 80 Measured for 8 Years
Measured for 8 Years
180 90
Predicted for 8 Years Predicted for 8 Years
200 100
(a) (b)

log G*, 64°C, 10 rad/s (kPa) log G*, 64°C, 10 rad/s (kPa)
0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0
0 0

10 10

20 20
Depth (mm)

Depth (mm)

30 30

40 40

50 50

60 60
Measured for 21 Years Measured for 18 Years
70 70
Predicted for 21 Years Predicted for 18 Years
80 80
(c) (d)

log G*, 64°C, 10 rad/s (kPa)


0.0 0.5 1.0 1.5 2.0
0

10

20
Depth (mm)

30

40

50

60
Measured for 17 Years
70
Prediction for 17 Years
80
(e)

Figure 57.   Predicted versus measured log G* values for (a) FHWA ALF-Control, (b) FHWA
ALF-SBS, (c) LTPP South Dakota, (d) LTPP New Mexico, and (e) LTPP Wisconsin sections.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   81  

log G*, 64°C, 10 rad/s (kPa)


0.5 1.0 1.5 2.0 2.5 3.0 3.5
0

10

20

Depth (mm)
30

40

50

60
(a)

log G*, 64°C, 10 rad/s (kPa) log G*, 64°C, 10 rad/s (kPa)
0.5 1.0 1.5 2.0 2.5 3.0 3.5 0.5 1.0 1.5 2.0 2.5 3.0 3.5
0 0

10 10

20 20
Depth (mm)

Depth (mm)

30 30

40 40

50 50

60 60
(b) (c)

log G*, 64°C, 10 rad/s (kPa) log G*, 64°C, 10 rad/s (kPa)
0.5 1.0 1.5 2.0 2.5 3.0 3.5 0.5 1.0 1.5 2.0 2.5 3.0 3.5
0 0

10 10

20 20
Depth (mm)

Depth (mm)

30 30

40 40

50 50

60 60
(d) (e)

Figure 58.   Predicted versus measured log G* values after 19 years of aging for
WesTrack sections with fine gradation: (a) all sections, (b) extreme sections,
(c) sections with optimum asphalt content, (d) sections with high asphalt
content, and (e) sections with low asphalt content.

Figure 59 presents the relationship between the kinetics model predictions of log G* at a
depth of 20 mm and those measured from field cores. The results demonstrate a high correlation
between the measured and predicted values of log G* (R2 = 0.75). The kinetics model tends to
over predict field aging, which is expected due to diffusion. That is, the partial pressure of oxygen
within the binder at a pavement depth of 20 mm will be lower than the oxygen partial pressure
at the surface of loose mixture from which mixture-specific M and G*o values were determined.
To enable accurate predictions of field aging, Equation 26 can be used with the experimentally
determined coefficients given in Figure 59 (i.e., C1 = 0.4867 and C2 = 0.615).

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

82   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

3.5
WT-Fine - Opt. %ac, Low %Va
y = 0.615x + 0.4867
3 WT-Fine - Opt. %ac, Medium %Va

log G*_measured (kPa)


R² = 0.75
2.5 WT-Fine - Opt. %ac, High %Va
WT-Coarse
2
LTPP-WI
1.5 LTPP-NM
LTPP-SD
1
LTPP-TX
0.5 ALF-CTRL
0 ALF-SBS
0 0.5 1 1.5 2 2.5 3 3.5
log G*_predicted (kPa)

Figure 59.   Calibration of kinetics prediction model of field aging by


comparing predicted versus measured log G* values at a depth of
20 mm (%ac = asphalt content, %Va = air void content).

Three WesTrack sections constructed with fine gradation were included in the calibration step
shown in Figure 59. These sections were constructed with optimum binder content (5.4%) and
various air void contents (i.e., high 12%, medium 8%, and low 4%). The calibration shown in
Figure 59 includes three data points for 19-year-old field cores obtained from the three WesTrack
Fine sections. Because the current kinetics prediction model does not account for the effect of
mixture morphological properties on field aging, all three sections have the same predicted
log G* value for 19 years of field aging (i.e., log G*_predicted = 2.325). However, the measured log G*
values obtained from 19-year-old field cores show different values (i.e., log G*_measured values range
between 1.765 and 2.077), indicating the need for the future consideration of the diffusion and
mixture morphological effects on field aging. The importance of diffusion modeling is discussed
further in the following paragraphs.
The calibrated kinetics model, i.e., Equations 25 and 26, was used to predict log G* as a func-
tion of depth using EICM hourly pavement temperature data and the log G* values at different
pavement depths were compared to field core measurements. Although the calibrated kinetics
model cannot account for diffusion differences with depth, the temperature gradients predicted
from the EICM data were used within the kinetics model predictions. Figure 57 shows the com-
parisons between the calibrated kinetics model predictions and the field core measurements for
the FHWA ALF and LTPP projects. It is important to note that the binder was extracted and
recovered only at a depth of 20 mm for the LTPP Texas field core, which is the reason those
results are not presented. With the exception of the LTPP Wisconsin project, the calibrated
kinetics model tends to under-predict aging near the pavement surface and over predict aging
at depths below 20 mm, which matches intuition in the absence of a diffusion model for the
depths below 20 mm and consideration of UV oxidation for the pavement near-surface. UV
oxidation will lead to more aging than is taken into account by the thermal oxidation kinet-
ics model herein. The partial pressure of oxygen within the binder will decrease as a function
of pavement depth, thereby leading to reduced aging, which is not considered in the current
model framework. Therefore, the results presented highlight the need for the development of
a diffusion model to complement the kinetics model to improve the prediction of field aging
in future work.
The WesTrack project offers a unique opportunity to study the effects of asphalt mixture
morphological properties on diffusion. The WesTrack Fine sections, detailed in Table 19, were
used in a systematic study of the changes in binder and air void contents and were used also to
conduct a preliminary evaluation of the effects of morphological properties on field aging.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   83  

Table 19.   WesTrack Fine sections used to evaluate the effects


of mixture morphological properties on field aging.

Binder Air Void


Location Field Age Mix Mix ID
Content Content
Low %aca, High %Vab 4.7% 12%
Low %ac, Medium %Va 4.7% 8%
WesTrack Optimum %ac, High %Va 5.4% 12%
Dayton, Fine
19 Years Optimum %ac, Medium %Va 5.4% 8%
NV (constructed
in 1995) Optimum %ac, Low %Va 5.4% 4%
High %ac, Medium %Va 6.1% 8%
High %ac, Low %Va 6.1% 4%
a
%ac stands for asphalt content.
b
%Va stands for air void content.

The calibrated kinetics model was used to predict log G* values after 19 years of field aging
in the WesTrack sections and these values were compared against field core measurements. Air
void content and film thickness do not affect loose mixture aging kinetics and, therefore, are
not taken into account by the calibrated kinetics model. Consequently, the predictions of the
log G* values were the same for all the WesTrack Fine sections. However, air void content and
film thickness affect the diffusion of oxygen in the field, and therefore, the field core results differ
among the different WesTrack sections.
Figure 58 shows the comparisons between the calibrated kinetics model predictions and the
field core measurements of log G* for 19-year-old WesTrack Fine sections. Figure 58 (a) includes
all of the WesTrack sections and shows a significant spread in the data. The general trends in oxi-
dation levels with changes in air void content and binder content suggest that mixture morphol-
ogy affects field aging; however, the wide spread of the data also may reflect field production and
placement variability. Figure 58 (b) shows the field core results of the most extreme cases (i.e.,
most severe aging versus least aging). The figure shows that the WesTrack section with the high-
est air void content and lowest binder content experienced the most aging whereas the section
with the lowest air void content and highest binder content experienced the least aging based on
the log G* values, which matches expectations. A higher air void content allows oxygen to perco-
late through the mix more easily than a lower air void content, and a lower film thickness creates
a shorter diffusion path. Figure 58 (c), (d), and (e) present the results of the WesTrack field sec-
tions with optimum binder content, high binder content, and low binder content, respectively.
Generally, a higher air void content at a given asphalt content leads to higher field aging.
These results highlight the need for the development of a diffusion model that considers the
morphological properties of asphalt mixtures in order to predict field aging more accurately. In
addition, the calibrated kinetics modeling framework presented here allows only for the predic-
tion of the G* at 64°C and 10 rad/s. Therefore, future research is suggested to establish a means
to predict the effects of aging over the range of temperatures and loading rates experienced by
pavements and thus enable the prediction of the changes in asphalt mixture properties with time
in pavement performance prediction frameworks.

Determination of Loose Mix Aging Kinetics Parameter


from USAT Binder Aging
The mixture-specific parameter values within the kinetics model expressed in Equation 25
are determined using loose mixture aging. Aging loose mixture in the oven allows the physico­
chemical effects of filler on asphalt binder oxidation rates to be captured, which is important
when predicting field aging (Moraes and Bahia 2015, Wu et al. 2014, Petersen 2009, Little
et al. 2006, Recasens et al. 2005, Huang et al. 2002, Jones 1997, Petersen et al. 1987). However,

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

84   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

determining the kinetics model parameter values using loose mixture aging is cumbersome.
The loose mixture must be prepared and conditioned in an oven at 95°C. The binder must
be extracted and recovered at a minimum of three durations of laboratory conditioning. The
extracted and recovered binder then must be subjected to DSR testing to determine the G* value
at 64°C and 10 rad/s as a function of laboratory aging time. Therefore, an alternative means
to calibrate the kinetics model using binder aging tests rather than mixture aging tests could
improve the feasibility of implementation. Correspondingly, an experiment was conducted to
relate the binder-derived aging rates to the corresponding loose mixture aging rates. The results
of this effort allow two options for calibrating the kinetics model:
1. Loose mixture aging coupled with extraction and recovery (as described).
2. USAT binder aging coupled with an empirical model to predict loose mixture aging based on
filler content and type.
As previously discussed, the calibration of the kinetics model parameters requires the use
of a kinetics-controlled oxidation experiment. Farrar et al. (2014) proposed the USAT for the
efficient simulation of asphalt binder aging in the laboratory. The USAT uses thin binder films
(0.3-mm thick) to induce a kinetics-controlled reaction (Farrar et al. 2014). Binder samples
conditioned for different durations can be tested in the DSR to derive a kinetics model. There-
fore, the USAT was selected for binder aging herein. To best mimic loose mixture aging, USAT
short-term aging was conducted in an oven at 135°C for 4 hours followed by long-term aging in
the same oven at 95°C. The evolution of log G* at 64°C for the USAT-aged binder was compared
to that of binder extracted and recovered from equivalent loose mixture aging and was used
to establish an empirical model to relate USAT binder aging to loose mixture aging. Figure 60
shows a summary of the experimental plan.
Because the properties of the binder, filler, and aggregate may affect the relationship between
the binder aging rates and the loose mixture aging rates, a broad range of materials, detailed in
Table 20, was used to develop an empirical model relating the USAT binder aging rates to the
loose mixture aging rates. A subset of the materials detailed in Table 20 was used to validate the
model’s extendibility to other mixtures rather than to calibrate the empirical model.

Oven Aging at
Binder USAT Aging 95°C for Various Loose Mix Aging
Durations

Binder Extraction
and Recovery

DSR Testing DSR Testing

Compare
Age
Hardening
Rates

Empirical Relationship between


Loose Mix Aging Rates and
Binder USAT Aging Rates

Figure 60.   Summary of experimental plan.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   85  

Table 20.   Mixtures used to compare binder aging rates obtained from USAT binder
aging and loose mix aging.

Modifier / %Pass
Hydrated
Binder Anti- Binder Binder Aggregate Sieve
Project Mix MIX ID Lime
Source Stripping Grade Content Type #200
Content
Agent (P200)
Anti-
Citgo- Limestone /
FHWA ALF- Stripping PG 70-
Venezuelan 5.3% Diabase 1.0% 6.3%
ALF CTRL Agent 22
Bachaquero (traprock)
(0.001)
120-
South
LTPP-SD N/A None 150 5.9% N/A None 4.6%
Dakota
pen
New LTPP-
N/A None AC-20 7.6% N/A None 6.1%
Mexico NM
LTPP
Texas LTPP-TX N/A None AC-20 5.4% N/A None 5.7%

Wisconsin LTPP-WI N/A None N/A 5.9% N/A None 4.1%

Andesite /
Fine WesTrack West Coast PG 64-
None 5.4% Granite / 1.5% 5.4%
Section Fine Refinery 22
Sand
WesTrack
Crushed
Coarse WesTrack Idaho PG 64-
None 5.7% Andesite / 1.5% 6.5%
Section Coarse Asphalt 22
Sand
Validation Sections
Limestone
SHRP California PG 58-
AAD None 5.3% / Diabase 1.0% 6.3%
AAD Coast 28
(traprock)
SHRP
Limestone
SHRP California PG 58-
AAG None 5.3% / Diabase 1.0% 6.3%
AAG Valley 10
(traprock)
Citgo-
PG 64-
NC DOT NCS9.5B NC Wilmington, None 6.6% Granite None 5.7%
22
NC

Figure 61 shows the log G* at 64°C and 10 rad/s results for both the USAT and loose mix aging
trials for the five mixtures evaluated that contain hydrated lime. Figure 62 shows the results of the
five different mixtures evaluated that do not contain hydrated lime. Hydrated lime is known to delay
oxidation significantly (Petersen 2009, Little et al. 2006, Recasens et al. 2005, Huang et al. 2002, Jones
1997, Petersen et al. 1987). It is clear that the binder that was conditioned using the USAT aged
significantly faster than when it was aged within a loose mixture, indicating that the filler inhib-
its oxidation. Similarly, the binder that was conditioned to simulate short-term aging using USAT
aged more quickly than binder extracted from short-term aged loose mixture. The inclusion of hydrated
lime appears to lead to a greater difference in the log G* evolution between loose mixture aging and
USAT binder aging, especially the slope of the secondary region that exhibits a linear relationship.
Figure 63 presents the log G* evolution results for all ten mixtures evaluated using (a) loose
mixture aging and (b) USAT binder aging. It can be seen that the relative log G* results among
the different mixtures are remarkably similar for USAT and loose mixture aging, which indicates
that the effect of the binder on the relative aging rate of different loose mixtures is more signifi-
cant than the aggregate or filler type.
An empirical model was developed to relate the binder and loose mixture aging rates as shown
in Equations 27 through Equation 35 to obtain loose mixture aging rates using USAT testing with
least squares regression. Equation 27 represents the shape of the log G* evolution, including the
spurt and constant rate reaction periods. Separate model parameter values are included for the

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

86   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

4
(a) ALF Control

log G*, 64°C, 10 rad/s (kPa)


3.5
3
2.5
2 log G* = 0.0019 (D) + 0.696
1.5 R² = 0.99
1 log G* = 0.0445 (D) + 1.7861
0.5 R² = 0.98

0
0 10 20 30 40
Aging Duration (Days)

4 4
log G* = 0.0625 (D) + 1.6739 log G* = 0.0285 (D) + 2.5199
log G*, 64°C, 10 rad/s (kPa)
log G*, 64°C, 10 rad/s (kPa)

3.5 3.5
R² = 0.97 R² = 1
3 3
2.5 2.5
2 2
log G* = 0.0161 (D) + 2.1611
1.5 1.5 R² = 0.64
log G* = 0.0437 (D) + 1.4079
1 R² = 0.99 1

0.5 0.5 (c) WesTrack Fine


(b) WesTrack Coarse
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

4 4
(d) SHRP AAD (e) SHRP AAG
log G*, 64°C, 10 rad/s (kPa)

log G*, 64°C, 10 rad/s (kPa)

3.5 3.5
3 3
2.5 2.5 log G* = 0.0436 (D) + 0.976
R² = 1
2 2
1.5 log G* = 0.0708 (D) + 1.7775 1.5
R² = 0.99
1 1
log G* = 0.0519 (D) + 1.4637 log G* = 0.0409 (D) + 0.567
0.5 0.5
R² = 0.94 R² = 1
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

Figure 61.   Measured USAT binder aging rates and loose mix aging rates at 95°C for
mixtures with hydrated lime: (a) ALF-Control, (b) WesTrack Coarse, (c) WesTrack Fine,
(d) SHRP AAD, and (e) SHRP AAG.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   87  

log G*, 64°C, 10 rad/s (kPa)


3.5 log G* = 0.0473 (D) + 0.6663
3 R² = 0.99

2.5 log G* = 0.0476 (D) + 0.2006


R² = 0.98
2
1.5
1
0.5
(a) LTPP-NM
0
0 10 20 30 40
Aging Duration (Days)

4 4
(b) LTPP-SD log G*, 64°C, 10 rad/s (kPa) log G* = 0.0024 (D) + 1.6942
log G*, 64°C, 10 rad/s (kPa)

3.5 3.5
R² = 0.98
3 log G* = 0.0663 (D) + 0.9795 3
R² = 0.98
2.5 2.5
2 2
1.5 1.5
log G* = 0.0023 (D) + 1.1155
1 1 R² = 0.99
log G* = 0.058 (D) + 0.389
0.5 0.5
R² = 0.94 (c) LTPP-TX
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

4 4
log G* = 0.0592 (D) + 1.466 (e) NCS9.5B
log G*, 64°C, 10 rad/s (kPa)

log G*, 64°C, 10 rad/s (kPa)

3.5 3.5
R² = 0.99
3 3
2.5 2.5
2 2
log G* = 0.0719 (D) + 1.7854
1.5 1.5
R² = 0.99
1 log G* = 0.0497 (D) + 1.1089
1
R² = 0.96 log G* = 0.064 (D) + 1.434
0.5 0.5 R² = 0.98
(d) LTPP-WI
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

Figure 62.   Measured USAT binder aging rates and loose mix aging rates at 95°C
for mixtures without hydrated lime: (a) LTPP New Mexico, (b) LTPP South Dakota,
(c) LTPP Texas, (d) LTPP Wisconsin, and (e) NCS9.5B.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

88   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

6 6
WesTrack Fine WesTrack Coarse WesTrack Fine WesTrack Coarse
ALF Control NCS9.5B ALF Control NCS9.5B
5 5
log G*, 64°C, 10 rad/s (kPa)

SHRP AAD SHRP AAG SHRP AAD SHRP AAG

log G*, 64°C, 10 rad/s (kPa)


LTPP Wisconsin LTPP New Mexico LTPP Wisconsin LTPP New Mexico
4 LTPP South Dakota LTPP Texas 4 LTPP South Dakota LTPP Texas

3 3

2 2

1 1
Loose Mix Aging at 95°C USAT Aging at 95°C
0 0
0 10 20 30 0 10 20 30
Aging Duration (Days) Aging Duration (Days)
(a) (b)

Figure 63.   Ranking comparisons: (a) loose mix aging rates and (b) USAT binder aging rates.

mixtures with and without hydrated lime. The model also considers the amount of filler that is
present in the mixture using the mass percentage of aggregate that passes the No. 200 sieve, as
presented in Equations 28 through Equation 35. In order to account for the effect of filler on the
reaction rate of loose mix aging during the slow (constant) reaction rate stage, Equation 28 is used
to convert the slope of the USAT binder log G* curve to that of the corresponding loose mixture.
Figure 64 shows the relationship between the slope correction factors (s) and percentage passing
the No. 200 sieve (P200) and the regression equations used to develop Equation 28. Equation 29 is
used to determine the intercept adjustment from USAT binder aging to loose mixture aging. The
parameter F is another shape factor that corrects the first-degree term constant during the fast
reaction state. Figure 65 summarizes the developed empirical model.

a ( D )2 + b ( D ) + c , D≤8
log G * =  (27)
a ( 8 )2 + b ( 8 ) + c + d ( D − 8 ) + e , D>8

3.51 − 0.32 ( P200 ) , hydrated lime


s=  (28)
1.60 − 0.1( P200 ). other fillers

2
Hydrated Lime
Other Mineral Fillers
Slope Correction

1.6
y = -0.32x + 3.51
R² = 0.99
1.2

y = -0.10x + 1.61
R² = 1.00
0.8
3 4 5 6 7
% Passing No. 200 Sieve (P200)

Figure 64.   Relationship between slope correction


factor and percentage passing No. 200 sieve.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   89  

Figure 65.   Summary of developed empirical


model to relate binder and loose mix aging
rates.

034, hydrated lime


I= (29)
0.5, other fillers

al = au (30)

bl = bu − F (31)

F=
(cl + bu × 8 + al × (8)2 ) − (dl × 8 + el ) (32)
8

c l = cu − I (33)

dl = du S (34)

e l = eu − I (35)

where
D = aging duration (days),
al , bl , cl , dl , el = loose mix dual-mechanism model parameters,
au, bu, cu, du, eu = USAT dual-mechanism model parameters,
S = slope correction factor to account for filler effects,
I = intercept correction factor to account for filler effects,
F = correction factor between fast and slow (constant) rates of reaction mechanism,
and
P200 = percentage passing No. 200 sieve.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

90   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

log G*, 64°C, 10 rad/s (kPa)


3.5
3
2.5
2
1.5
1
0.5
(a) ALF Control
0
0 10 20 30 40
Aging Duration (Days)

4 4

log G*, 64°C, 10 rad/s (kPa)


log G*, 64°C, 10 rad/s (kPa)

3.5 3.5
3 3
2.5 2.5
2 2
1.5 1.5

1 1

0.5 0.5 (c) WesTrack Fine


(b) WesTrack Coarse
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

Figure 66.   Predicted age hardening rates for loose mix aging at 95°C from USAT
binder aging for mixtures with hydrated lime: (a) ALF-Control, (b) WesTrack Coarse,
and (c) WesTrack Fine.

The proposed model was applied to the binder log G* results obtained from the USAT binder
aging tests to predict the corresponding loose mixture aging evolution. Figure 66 and Figure 67
present comparisons between the measured and predicted loose mixture aging rates for the five
mixtures that were used to calibrate the empirical model using mixtures with and without hydrated
lime, respectively. The results demonstrate excellent agreement between the measured and pre-
dicted log G* values.
Figure 68 shows comparisons between the measured and predicted loose mixture aging rates
for the independent validation mixtures. Note that two of the validation mixtures do not contain
hydrated lime whereas the third does. Generally, the prediction accuracy is very good, thereby
validating the use of the empirical equation to obtain the evolution of log G* for the loose mix
using USAT binder test results. These results suggest that the calibrated mixture-specific param-
eters of the kinetics model can be determined using USAT binder aging, thereby negating the
need for loose mixture aging and corresponding extraction and recovery.

Applications
Integration of Pavement Aging Model
in Mechanistic–Empirical Design
In Pavement ME Design, the material properties, pavement structure, and traffic and climatic
conditions are used as inputs to evaluate a trial pavement structural design. The trial design is
then examined using the response models that are incorporated in the mechanistic–empirical

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   91  

4 4
(a) LTPP-NM (b) LTPP-SD

log G*, 64°C, 10 rad/s (kPa)


log G*, 64°C, 10 rad/s (kPa)

3.5 3.5
3 3
2.5 2.5
2 2
1.5 1.5

1 1

0.5 0.5

0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

4 4
log G*, 64°C, 10 rad/s (kPa)

log G*, 64°C, 10 rad/s (kPa)

3.5 3.5
3 3
2.5 2.5
2 2
1.5 1.5
1 1
0.5 (c) LTPP-TX 0.5 (d) LTPP-WI
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

Figure 67.   Predicted age hardening rates for loose mix aging at 95°C from USAT
binder aging for mixtures without hydrated lime: (a) LTPP New Mexico, (b) LTPP
South Dakota, (c) LTPP Texas, and (d) LTPP Wisconsin.

design to determine stresses, strains, and deformations using layered elastic analysis. The
response model output is used to predict the corresponding pavement distress development
over the design period. The dynamic modulus of the asphalt mixtures in the individual layers
within the pavement is a key property for predicting the pavement responses and correspond-
ing distress development over the designated service life of the pavement. Based on the pre-
dicted pavement performance, the trial pavement design is adjusted to obtain the thicknesses
and material properties that are required to achieve adequate performance.
The GAS model, developed by Mirza and Witczak (1995), is applied within Pavement ME
Design to predict the changes in asphalt binder viscosity that are due to oxidative aging over
the course of a pavement’s service life. The predicted asphalt binder viscosity values are then
incorporated into the Witczak equation to predict the corresponding dynamic modulus ( |E *|)
evolution with oxidative aging. The GAS model assumes that aging is limited to the top 1.5 in. of
the pavement. Therefore, within Pavement ME Design, the material properties below 1.5 in.
are not considered. Consequently, bottom-up fatigue cracking predictions are not sensitive
to pavement age. Based on the findings from this study, oxidative aging occurs throughout
the entire thickness of the asphalt layers within a pavement. Hence, the current Pavement

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

92   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

log G*, 64°C, 10 rad/s (kPa)


3.5
3
2.5
2
1.5
1
0.5 (a) SHRP AAD
0
0 10 20 30 40
Aging Duration (Days)

4 4
log G*, 64°C, 10 rad/s (kPa)

log G*, 64°C, 10 rad/s (kPa)


3.5 3.5
3 3
2.5 2.5
2 2
1.5 1.5
1 1
0.5 (b) SHRP AAG 0.5 (c) NCS9.5B
0 0
0 10 20 30 40 0 10 20 30 40
Aging Duration (Days) Aging Duration (Days)

Figure 68.   Validation of predictive model: (a) SHRP AAD, (b) SHRP AAG,
and (c) NCS9.5B mixtures.

ME Design methodology needs to be modified and improved to consider the effects of


aging on cracking performance. A detailed explanation of the GAS model is presented in
Appendix B.
The kinetics modeling framework developed herein could potentially be used to replace the
GAS model in Pavement ME Design for the prediction of the binder G* evolution throughout
a pavement’s service life. However, the predicted G * values must be input into a model to
predict the corresponding dynamic modulus evolution. Therefore, in this study, the Witczak
equation (Bari and Witczak 2006), Hirsch model (Christensen et al. 2003), and the North
Carolina State University Artificial Neural Network (NCSU ANN) model (Sakhaei Far 2011)
were used to evaluate the accuracy of these existing dynamic modulus predictive models when
they are applied to highly aged materials. These models were evaluated using the laboratory-
aged mixture dynamic modulus test results of the LTPP South Dakota asphalt mixture.
Laboratory-fabricated loose mixture was short-term aged and then subjected to long-term
aging at 95°C in the oven for 4, 8, and 16 days prior to producing specimens for dynamic
modulus testing. The aged loose mixture samples were compacted in a Superpave gyratory
compactor to achieve 4.5% air voids in the test specimens. Four small Ø38-mm diam-
eter, 110-mm tall cores were extracted vertically from the inner 100-mm diameter of the
gyratory-compacted specimens. The small specimens were subjected to dynamic modulus
testing at 4°C, 20°C, and 40°C with loading frequencies ranging from 0.1 Hz to 25 Hz. Fig-
ure 69 presents the resulting dynamic modulus master curves that demonstrate the rate of
increase in the dynamic modulus values that is caused by oxidation decay with the duration
of laboratory aging.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   93  

1.0E+08

1.0E+07
|E*| (MPa)

1.0E+06
Short-Term
Oven, 95°C, 4 Days
Oven, 95°C, 8 Days
Oven, 95°C, 16 Days
1.0E+05
1.0E-04 1.0E-02 1.0E+00 1.0E+02 1.0E+04
Reduced Frequency (Hz)

Figure 69.   Comparison of dynamic modulus master curves at


different aging levels for LTPP South Dakota mix.

Asphalt binder samples were extracted and recovered from the aged loose mixtures and sub-
jected to DSR testing to determine the binder G* input for the dynamic modulus predictive
models. Table 21 shows the G* results at 20°C and 10 Hz for the asphalt binders extracted and
recovered from loose mixtures aged for 4, 8, and 16 days. The G* results demonstrate a con-
tinuous increase in the asphalt binder G* value even after 8 days of aging. However, the asphalt
mixture dynamic modulus test results show a comparably small change in dynamic modulus
values between 8 and 16 days of aging.
An intermediate temperature of 20°C and loading frequency of 10 Hz were selected for the
dynamic modulus predictions. Figure 70 shows the comparisons between the model predic-
tions and the measured dynamic modulus values at 20°C and 10 Hz frequency. The results
demonstrate that the measured dynamic modulus values of the long-term aged material are
significantly lower than the model predictions. The Witczak model over-predicted the dynamic
modulus values most significantly. The Hirsch and ANN model predictions have a maximum
error of approximately 50%.
The preliminary evaluation of the three dynamic modulus prediction equations suggests
that the Witczak model, currently used within Pavement ME Design, lacks the ability to predict
dynamic modulus values of aged mixtures accurately. Therefore, the replacement of the Witczak
equation in Pavement ME Design with a more accurate dynamic modulus prediction model
merits consideration in future work. Alternative dynamic modulus predictive equations
(i.e., NCSU ANN and Hirsch), although more accurate than the Witczak model, can still lead to
significant error in the prediction of asphalt mixture dynamic modulus values, suggesting the
need for recalibration using a broad set of materials with various age levels in future work.

Table 21.   G* results for asphalt binders extracted and


recovered from loose mix aged at different durations.

Aging Aging G* at 20°C and Dynamic Modulus


Sample
Temperature Duration 10 Hz at 20°C and 10 Hz
STA 135°C - 7,390 kPa 4,523 MPa
4D 95°C 4 days 13,646 kPa 6,364 MPa
8D 95°C 8 days 16,621 kPa 7,744 MPa
16 D 95°C 16 days 21,567 kPa 8,473 MPa

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

94   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

2.4E+04
|E*| _Measured |E*| _ Witczak
|E*| _Hirsch |E*| _NCSU ANN
1.9E+04
1.8E+04
1.6E+04
1.4E+04 1.2E+04

|E*| (MPa)
1.0E+04 1.2E+04
1.2E+04
1.2E+04 9.1E+03
8.7E+03
7.2E+03 7.6E+03 8.5E+03
6.3E+03 7.7E+03
6.4E+03
6.0E+03 4.5E+03

0.0E+00
Short-Term Oven, 95°C, Oven, 95°C, Oven, 95°C,
Aged 4Days 8Days 16Days

Figure 70.   Comparison between dynamic modulus values measured


at 20°C and 10 Hz and predicted values based on extracted and
recovered binder data measured at 20°C and 10 Hz.

Conclusions
The accurate characterization of asphalt mixture properties in terms of the service life of a
pavement is becoming more important as more powerful pavement design and performance
prediction methods are implemented. This project sought to develop a procedure to simulate
the long-term aging of asphalt mixtures as a function of climate and depth for performance test-
ing and prediction. The results of this project provide a basis for the future development of a
methodology that integrates the effects of long-term aging in mechanistic–empirical pavement
analysis programs such as Pavement ME Design.

Sensitivity Study
The goal of the sensitivity study was to estimate the sensitivity of the mechanical properties of
asphalt concrete to asphalt binder oxidation. The sensitivity study provided thresholds by which
to evaluate the significance of observed differences in asphalt binder AIPs in terms of asphalt
mixture performance. The following conclusions pertain to the sensitivity study.
• The properties of asphalt concrete are not proportionally sensitive to changes in the asphalt
binder modulus.
• To replicate the effects of binder oxidation on the modulus of asphalt concrete with a given
level of accuracy (say 10%), the desired binder oxidation can be replicated at less accuracy by
a factor of 1.5 to 3.6 (15% to 36 %).
• If the binder AIPs are replicated with a certain percentage of error (say 10%), then the expected
percentage of error in the resulting fatigue properties of an asphalt mixture that is tested will
be lower than or equal to 10%.

Selection of the Chemical and Rheological Aging Index Properties


Candidate chemical and rheological AIPs were evaluated in order to select the AIPs needed
to track the oxidation levels of field-aged and laboratory-aged materials when developing the

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   95  

long-term aging procedure and associated kinetics model. The following conclusions pertain to
the selection of the chemical and rheological AIPs.
• All three chemical AIPs evaluated [i.e., carbonyl plus sulfoxide (C + S) peaks, carbonyl area,
and C + S area] showed good sensitivity to aging duration; however, the C + S peaks were the
most reliable among the three chemical AIPs.
• Including the effect of the sulfoxide functional group in the oxidative aging evaluation
was crucial, as different aging rates were observed from carbonyl alone and from the C + S
index values.
• The dynamic shear modulus (G*) as a direct output of the DSR can be used to develop a
relationship between the chemistry and rheology of asphalt binder. It is simple and more
consistent with chemical indices than other rheology-based AIPs.

Selection of the Long-Term Aging Method


An experimental program was conducted to select the proposed long-term aging method. The
following factors were evaluated to select the proposed laboratory aging procedure: (a) state of
the material during aging (compacted specimen versus loose mix), (b) pressure level (oven aging
versus pressurized aging), and (c) aging temperature (95°C versus 135°C). The following con-
clusions pertain to the selection of the proposed long-term aging method.
• The current standard procedure for the long-term aging of asphalt mixtures (AASHTO
R 30), which consists of conditioning large 100-mm diameter compacted specimens in an
oven at 85°C, leads to the development of an oxidation gradient from the specimen’s center
to its periphery. The lack of uniform properties throughout the specimen is of concern for
performance testing and was observed directly in cyclic direct tension fatigue tests through
a high rate of end failure at specimen locations where oxidation was most significant.
• The application of pressure in the compacted specimen aging process expedites oxidation.
However, the performance test results indicate that the application and/or release of pressure
damages specimens.
• The aging gradient observed in large compacted specimens that were subjected to oven aging
was eliminated by the use of small specimens (38-mm diameter with 100-mm height) due to
their shortened lateral diffusion paths. Therefore, the oven aging of small compacted speci-
mens is the most promising compacted specimen aging procedure, as no integrity issues were
observed.
• Aging asphalt mixtures in a loose mix state expedites oxidation compared to compacted speci-
men aging under the same conditions.
• The compactive effort required to compact long-term aged loose mixes is comparable to
that required for short-term aged mixes, with no adjustment to the compaction temperature
needed based on the results for two mixtures, PG 64-22 and PG 70-28 SBS-modified, the latter
of which is known to be difficult to compact.
• The performance test results indicate no problems with loose mixtures compacted after
long-term aging.
• Pressure expedites the aging of loose mix. However, the size of the standard binder PAV pro-
hibits the generation of enough aged material for performance testing. If the pressure aging
of loose mix were to be selected as the best aging procedure, then a larger PAV would need
to be developed.
• Loose mix aging in an oven is the most promising aging procedure to produce mixture speci-
mens for performance testing in terms of efficiency and integrity, without the need to develop
costly new equipment. In addition, any specimen geometry (e.g., beams) can be produced
using aged loose mix, also making this procedure the most versatile option.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

96   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

• The results indicate that loose mix aging at 95°C is the optimal procedure for the long-term
aging of asphalt concrete for performance testing. The literature indicates that the disruption
of polar molecular associations and sulfoxide decomposition become critical at temperatures
that exceed 100°C (Petersen 2009). In this study, 95°C was selected as the aging temperature
instead of 100°C in order to avoid the aging temperature reaching close to 100°C due to pos-
sible temperature fluctuations in the oven. Aging at lower temperatures precludes reaching
field levels of oxidation within a reasonable time.
• A significant change in the relationship between binder rheology and chemistry can occur for
certain asphalts when the aging temperature is increased from 95°C to 135°C. This change
implies corresponding changes in the kinetics and mechanisms of the oxidation reactions that
are associated with an increase in temperature from 95°C to 135°C, which is consistent with
findings described in the literature.
• Asphalt mixture performance can be negatively impacted by long-term aging at 135°C.
Despite having matching rheological characteristics, two of the three mixtures evaluated
in this study exhibited decreases in both dynamic modulus values and fatigue resistance.
These results suggest that aging at 135°C for performance assessment and prediction should
be avoided.
• When the loose mix laboratory aging temperature is at or below 95°C, the relationship
between binder chemistry and rheology is unaffected by the aging temperature, based on all
three mixtures evaluated, indicating that the aging temperature does not affect the oxidation
reaction mechanism.
• The rate of oxidation increases with an increase in temperature, and thus, the results suggest
that the optimal loose mixture laboratory aging temperature is 95°C.
• It is proposed that loose mixture aging at 95°C should be used for the laboratory long-term
aging of both WMA and HMA mixtures. Future research is necessary to investigate the advan-
tages and disadvantages of compacting WMA after long-term aging.

Climate-Based Determination of Predefined Aging Durations


After selecting the proposed long-term aging method, a means to determine the laboratory
aging durations that are required to represent a given time, climate, and depth within a pave-
ment was developed. Project-specific laboratory aging durations that are required to match
the AIPs of field cores at varying depths were determined for a broad set of materials. The
project-specific aging durations were used to calibrate a kinetics-derived CAI that can be used
to determine the required laboratory aging duration to match the field aging at any location
of interest. The following conclusions pertain to the selection of the proposed long-term aging
method.
• Loose mixture oven aging leads to a kinetics-controlled oxidation reaction, indicating that
kinetics models can be derived and applied to loose mixture aging without the need to con-
sider diffusion.
• A kinetics model for loose mixture aging was developed and then validated using log G* at
64°C and 10 rad/s frequency as the AIP within Glaser et al.’s (2013a) kinetics model frame-
work that was developed initially for binder aging. The kinetics model can be calibrated using
AIP measurements obtained from isothermal aging at a single temperature.
• The laboratory aging duration required to match a given field condition is independent of the
material-specific kinetics.
• The CAI, developed by simplifying the kinetics model and calibrating against field data, can
prescribe laboratory aging durations to match a given field condition using hourly pavement
temperature histories at depths of 6 mm, 20 mm, and 50 mm or below.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Findings and Applications   97  

Aging Model to Predict Field Aging


This study evaluated the ability of the kinetics model developed herein to predict the evolu-
tion of oxidative aging in pavements. The prediction of the changes in asphalt binder properties
with oxidative aging is important for the simulation of changes in asphalt mixture properties
that are induced by aging within pavement performance prediction frameworks, including
Pavement ME Design. The following conclusions pertain to the kinetics aging model.
• The kinetics model was calibrated successfully (R2 = 0.75) against field core measurements
of log G* at a depth of 20 mm, which represents a reasonable depth for the evaluation of the
bulk behavior of surface layers and avoids the effect of UV oxidation at the pavement surface.
• The calibrated kinetics model tends to under-predict aging near the pavement surface and
over predict aging at depths below 20 mm, which matches intuition in the absence of a dif-
fusion model and consideration of UV oxidation at the pavement near-surface. These results
highlight the need for the development of a diffusion model that considers the morphological
properties of asphalt mixtures in order to predict field aging more accurately.
• An empirical model was developed and validated to relate the USAT asphalt binder aging
rates and the loose mixture aging rates. The model allows USAT binder testing to be used to
determine the mixture-specific coefficients that are included within the calibrated kinetics
model, thereby negating the need for cumbersome loose mixture aging and corresponding
extraction and recovery.

Integration of the Pavement Aging Model


in Mechanical–Empirical Design
Within pavement performance prediction frameworks, including Pavement ME Design, the
predicted changes in binder properties as a result of oxidation must be input into a model
to predict the corresponding changes in asphalt mixture properties. Therefore, a preliminary
investigation into the accuracy of existing dynamic modulus predictive models with regard to
highly aged materials was conducted using the Witczak equation (Bari et al. 2006), Hirsch model
(Christensen et al. 2003), and the NCSU ANN model (Sakhaei Far 2011). The following conclu-
sions pertain to the evaluation of dynamic modulus predictive models.
The three evaluated dynamic modulus prediction equations lack the ability to accurately pre-
dict dynamic modulus values from binder properties. The Witczak model, used in Pavement
ME Design, over-predicts the dynamic modulus values most significantly. The Hirsch and ANN
model predictions have a maximum error of approximately 50%.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

CHAPTER 4

Suggested Future Research

Suggested future research is presented in Figure 2. In order to arrive at a final aging protocol,
the interim aging protocol should be applied to RAP mixtures, WMA mixtures, and polymer-
modified mixtures. The AIPs of the binders extracted and recovered from the laboratory-aged
loose mixtures would be compared to those of the binders extracted and recovered from cor-
responding field cores. In addition, the compaction of loose mixtures after long-term aging for
the various WMA technologies should be investigated.
A diffusion model that considers the morphological properties of asphalt mixtures (e.g., binder
film thickness, air voids) should be developed and coupled with the developed kinetics model
to arrive at a pavement aging model. Such a model would improve the prediction of changes
in asphalt binder properties with oxidative aging within pavement performance prediction
frameworks, including Pavement ME Design. The diffusion model could be developed largely
by evaluating the WesTrack mixtures, which include systematic changes in asphalt mixture
morphology. The combined kinetics–diffusion model would then be calibrated against field
core AIP measurements at varying depths using materials and field cores from various pave-
ments in different climatic conditions, including the Group B materials and pavements used
in this study.
The predicted changes in the binder properties with oxidative aging must be related to the
changes in the asphalt mixture properties to facilitate the integration of the pavement aging
model in pavement performance prediction frameworks. Therefore, a systematic study of
laboratory-aged loose mixtures should be conducted, whereby loose mixtures are prepared at
three levels of long-term aging and subjected to dynamic modulus and cyclic fatigue perfor-
mance testing. Asphalt binders would be extracted and recovered from the long-term aged loose
mixtures and evaluated. The changes in the asphalt binder properties with oxidative aging would
be related to the corresponding changes in the asphalt mixture dynamic modulus values and
cyclic fatigue performance. Note that the current kinetics model allows only for the prediction
of G* at 64°C and 10 rad/s. Thus, linking changes in binder properties with oxidative aging to
corresponding changes in mixture performance may require establishing a means to predict
the effects of aging over the range of temperatures and loading rates actually experienced by
pavements. The performance test results of the systematic aging study will be integrated into
FlexPAVE™ to predict pavement performance, with consideration given to the changes in
mixture properties as a function of depth and time due to aging. The results will be analyzed
to evaluate the effect of aging on cracking performance and to identify the depth within the
pavement that is most critical to pavement performance.
There is also the possibility to extend the kinetics–diffusion modeling framework to address
the impact of oxidative aging in accelerated pavement tests (APTs). The kinetics-diffusion

98

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Suggested Future Research   99  

modeling framework would be used to derive a tool for the prediction of the required APT
conditions (i.e., time and temperature) to represent a given level of aging. In addition, a means
to determine the appropriate laboratory aging conditions to produce performance test speci-
mens with the same level of aging as a given APT and that has the capability to consider the
effects of environmental exposure both prior to and/or after the APT and of thermal treatment
during the APT should be developed.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

CHAPTER 5

Proposed Standard Method of


Test for Long-Term Conditioning
of Hot Mix Asphalt (HMA)
for Performance Testing

100

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Proposed Standard Method of Test for Long-Term Conditioning of Hot Mix Asphalt (HMA) for Performance Testing   101  

Standard Method of Test for

Long-Term Conditioning of Hot Mix Asphalt


(HMA) for Performance Testing
AASHTO Designation: TP xxx-xx

1 SCOPE
1.1 This standard practice describes a procedure for the long-term conditioning
of uncompacted hot mix asphalt (HMA) for performance testing to simulate
the aging that occurs over the service life of a pavement. The procedure for
long-term conditioning in performance testing is preceded by a procedure
for short-term conditioning in mixture mechanical property testing in
AASHTO R 30.

1.2 This standard may involve hazardous material, operations, and equipment.
This standard does not purport to address all safety problems associated
with its use. It is the responsibility of the user of this procedure to establish
appropriate safety and health practices and to determine the applicability of
regulatory limitations prior to use.

2 REFERENCED DOCUMENTS
2.1 AASHTO Standards:
PP 3, Preparing Hot Mix Asphalt (HMA) Specimens by Means of the Rolling
Wheel Compactor
PP 60, Preparation of Cylindrical Performance Test Specimens Using the
Superpave Gyratory Compactor (SGC)
R 30, Mixture Conditioning of Hot Mix Asphalt (HMA)
T 312, Preparing and Determining the Density of Hot Mix Asphalt (HMA)
Specimens by Means of the Superpave Gyratory Compactor
T 316, Viscosity Determination of Asphalt Binder using Rotational
Viscometer

2.2 Other Document:


Draft Final Report for NCHRP Report 9-54, Long-Term Aging of Asphalt
Mixtures for Performance Testing and Prediction, October 1, 2017.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

102   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

3 TERMINOLOGY
3.1 Definions:

3.1.1 Enhanced Integrated Climac Model (EICM) – A one-dimensional coupled heat


and moisture flow model that can be used to determine pavement temperature
at any depth of interest.

4 SUMMARY OF Prac‚ce
4.1 A mixture of binder and aggregate is condi‚oned in a forced-dra† oven at
95°C a†er prior short-term condi‚oning for mixture mechanical property
testing according to R 30. The dura‚on of condi‚oning at 95°C is selected
to reflect the ‚me, climate, and pavement depth for a given pavement
loca‚on in the United States using the clima‚c aging index (CAI). The long-
term condi‚oned loose mixture samples are prepared for maximum
specific gravity (Gmm) tes‚ng and then compacted for subsequent
performance tes‚ng.

5 SIGNIFICANCE AND USE


5.1 The long-term performance of HMA can be predicted more accurately by
using condi‚oned test samples rather than uncondi‚oned samples. The
long-term mixture condi‚oning in the performance tes‚ng procedure is
designed to simulate the aging that the mixture will undergo in service.

6 APPARATUS
6.1 Oven: A forced-dra† oven, thermostatically-controlled with horizontal air
flow, and capable of maintaining any desired temperature se›ng from
room temperature to 176°C within ± 3°C.

6.2 Thermometers: Thermometers having a range from 50°C to 260°C and


readable to 1°C.

6.3 Miscellaneous: A metal pan for heating aggregate, a shallow metal pan for
the short-term conditioning of uncompacted HMA, a 1-inch tall metal pan
for the long-term conditioning of uncompacted HMA, a metal spatula or
spoon, ‚mer, and gloves for handling hot equipment.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Proposed Standard Method of Test for Long-Term Conditioning of Hot Mix Asphalt (HMA) for Performance Testing   103  

7 HAZARDS
7.1 This standard involves the handling of HMA that can cause severe burns if it
contacts skin. Follow safety precauons to avoid burns.

8 LONG-TERM MIXTURE CONDITIONING PROCEDURE


8.1 The long-term conditioning for performance tesng applies to laboratory-
prepared loose mixtures that have been subjected to short-term
condioning as part of the mixture mechanical property tesng procedure
described in R 30.

8.2 Determine the required long-term conditioning duraon that reflects the
desired field aging in terms of age, climate, and depth using Equaon 1.
N
toven DA exp( Ea / RTi ) / 24 (1)
i 1

where
toven = required oven aging duration at 95°C to reflect field aging (days);
CAI = climatic aging index;
D = depth correction factor;
A = frequency (pre-exponential) factor;
Ea = activation energy;
R = universal gas constant, or ideal gas constant;
Ti = pavement temperature obtained from the EICM at the depth of interest
at the hour of interest, i (Kelvin);
Table 1 lists the values of D, A, and Ea.

Table 1 — Climatic Aging Index Fitting Coefficients


Depth Correction Pre-exponential Activation
Pavement Layer
Factor (D) Factor (A) Energy (Ea)
Surface Layer (6 mm) 1.0000 1.40962 13.3121
20-mm Depth 0.4565 1.40962 13.3121
Deeper Layers (below 20 mm) 0.2967 1.40962 13.3121

8.2.1 Break down any large chunks of asphalt mixtures in the short-term aged loose
mixture sample, taking care to avoid fracturing the aggregate, so that the clusters
of the fine aggregate poron are not larger than the nominal maximum aggregate
size (NMAS). This step needs to be performed shortly ašer short-term aging to
ensure that the asphalt mixture is sufficiently soš to be separated into pans for
oven aging. If an HMA sample is not sufficiently soš to be separated manually,

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

104   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

then place it in a pan and warm it in an oven unl it can be separated as


described.

8.2.2 Apporon the short-term aged loose mixture into several pans such that each pan
has a relavely thin layer of loose mix, approximately equal to the mixture NMAS.

8.2.3 Place the pans that contain mixture in a forced-dra oven at 95°C ± 3°C for the
duraon determined by Equaon (1). Place the pans on different shelves so that
the pans are arranged vercally within the oven as much as possible. Ensure that
adjacent pans do not overlap.

8.2.4 Rotate the pans to different shelves at four evenly-spaced me intervals during
the long-term condioning period so that each pan has similar exposure to heat
and air flow at different locaons within the oven.

8.2.5 Aer long-term condioning, remove the condioned mixtures from the oven
and mix all of the mixtures together in order to obtain a uniform mixture.

8.2.6 Allow the mixture to cool to room temperature. The long-term condioned loose
mixture sample is now ready for compacon or subsequent tesng as required.

8.3 Preparing Specimens from Loose HMA

8.3.1 Specimens Compacted Using the Superpave Gyratory Compactor

8.3.1.1 Compact the specimens in accordance with PP 60. Cool the test specimens at
room temperature for 16 ± 1 hour.
Note 4. Extrude the specimen from the compaction mold after cooling for 2 to 3
hours.
Note 5. Specimen cooling usually is scheduled as an overnight step. Cooling may
be accelerated by placing the specimen in front of a fan.

8.3.2 Specimens Compacted Using the Rolling Wheel Compactor

8.3.2.1 Compact the specimens in accordance with PP 3.

8.3.2.2 Cool the test specimens at room temperature for 16 ± 1 hour.

8.3.2.3 Remove the slab from the mold, and saw or core the required specimens from the
slab.

9 REPORT
9.1 Report the binder grade, binder content (to nearest 0.1 percent), and the
aggregate type and gradaon, if applicable.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Proposed Standard Method of Test for Long-Term Conditioning of Hot Mix Asphalt (HMA) for Performance Testing   105  

9.2 Report the following long-term condi oning information for the
performance tes ng condi ons, if applicable:

9.2.1 Long-term mixture condi oning temperature in laboratory (nearest 1°C);

9.2.2 Long-term mixture condi oning dura on in laboratory (nearest 5 min); and

9.2.3 Laboratory compac on temperature (nearest 1°C).

10 Keywords
10.1 Condi oning; hot mix asphalt; long-term conditioning.

APPENDICES
(Nonmandatory Information)

X1. LABORATORY AGING DURATION MAPS

X1.1. The CAI values for various locations in the United States were calculated, as
described in Section 9, using hourly pavement temperature history data obtained
from the EICM to provide an overview of the proposed laboratory aging durations
for various climates, field aging durations, and depths.

X1.2. The CAI values and measured durations that are needed to match field aging
correlate linearly in a one-to-one relationship. Thus, the CAI values represent the
required duration at 95°C that is needed to match the field aging for a given
pavement temperature history and depth.

X1.3. Laboratory aging durations were calculated for three field ages: 4 years, 8 years,
and 16 years. For each field age, the laboratory aging durations were determined
at three depths: 6 mm, 20 mm, and 50 mm, and rounded to the nearest day.

X1.4. Figure X1.1 shows the CAI-determined loose mixture aging durations at 95°C
that are required to match 4, 8, and 16 years of field aging at a depth of 6 mm.
Figure X1.1 demonstrates that climate has a significant effect on the required
laboratory aging duration that is required to match a given field age.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

106   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

(a)

(b)

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Proposed Standard Method of Test for Long-Term Conditioning of Hot Mix Asphalt (HMA) for Performance Testing   107  

(c)

Figure X1.1-Required oven aging duration at 95°C to match level of field aging 6 mm below
pavement surface for (a) 4 years of field aging, (b) 8 years of field aging, and (c) 16 years of
field aging.

X1.5. Figure X1.2 shows the CAI-determined loose mixture aging durations at 95°C
that are required to match 4, 8, and 16 years of field aging at a depth of 20 mm. A
comparison between the laboratory aging durations presented in Figure X1.1 and
Figure X1.2 demonstrates that significantly shorter laboratory aging durations are
required to match the field aging at a depth of 20 mm compared to 6 mm,
indicating that the temperature gradient and diffusion in pavements significantly
affect oxidation levels.

X1.6. A depth of 20 mm represents a reasonable depth for the evaluation of surface


layer asphalt mixtures because it better reflects bulk behavior within a pavement
structure than nearer the surface and avoids the effect of ultraviolet oxidation.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

108   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

(a)

(b)

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Proposed Standard Method of Test for Long-Term Conditioning of Hot Mix Asphalt (HMA) for Performance Testing   109  

(c)

Figure X1.2-Required oven aging duration at 95°C to match level of field aging 20 mm
below pavement surface for (a) 4 years of field aging, (b) 8 years of field aging, and (c) 16
years of field aging.

X1.7. Figure X1.3 shows the CAI-determined loose mixture aging durations at 95°C
that are required to match 4, 8, and 16 years of field aging at a depth of 50 mm.
The results demonstrate that considerably shorter aging durations are required to
simulate aging at a depth of 50 mm compared to depths of 20 mm and 6 mm, thus
indicating the presence of a significant oxidation gradient with depth near the
surface of the pavements. In Figure X1.3, the required aging duration of zero days
in a few cold northern states indicates that long-term aging at 50 mm below the
pavement surface in these cold regions are not significant enough to require long-
term oven conditioning to mimic field conditions.

X1.8. Based on the results presented in the draft NCHRP Project 9-54 report, the
researchers concluded that long-term aging does take place below a depth of 50
mm, but does not change appreciably below that depth. Consequently, the
evaluation of asphalt mixtures that have been prepared to match field aging at a
depth of 50 mm could be useful for evaluating intermediate and base asphalt
layers.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

110   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

(a)

(b)

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Proposed Standard Method of Test for Long-Term Conditioning of Hot Mix Asphalt (HMA) for Performance Testing   111  

(c)

Figure X1.3-Required oven aging duration at 95°C to match level of field aging 50 mm
below pavement surface for (a) 4 years of field aging, (b) 8 years of field aging, and (c) 16
years of field aging.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

References

Airey, G. D. (2003). State of the Art Report on Ageing Test Methods for Bituminous Pavement Materials.
The International Journal of Pavement Engineering, 4(3), 165–176. https://doi.org/10.1080/102984304
2000198568.
American Association of State Highway and Transportation Officials (AASHTO). (2002). Standard Practice for
Mixture Conditioning of Hot Mix Asphalt, AASHTO R 30, Washington, D.C.
American Association of State Highway and Transportation Officials (AASHTO). (2008). Standard Method of
Test for Effect of Heat and Air on a Moving Film of Asphalt (Rolling Thin-Film Oven Test), AASHTO T 240,
Washington, D.C.
American Association of State Highway and Transportation Officials (AASHTO). (2011). Standard Method of
Test for Determining Dynamic Modulus of Hot Mix Asphalt (HMA), AASHTO 342, Washington, D.C.
American Association of State Highway and Transportation Officials (AASHTO). (2012). Standard Practice for
Accelerated Aging of Asphalt Binder Using a Pressurized Aging Vessel (PAV), AASHTO R 28, Washington, D.C.
American Association of State Highway and Transportation Officials (AASHTO). (2014). Determining the
Damage Characteristic Curve of Asphalt Concrete from Direct Tension Cyclic Fatigue Tests, AASHTO TP 107,
Washington, D.C.
Anderson, D. A., D. W. Christensen, H. U. Bahia, R. Dongre, M. G. Sharma, C. F. Antle, and J. Button. (1994).
Binder Characterization and Evaluation Volume 3: Physical Characterization, Strategic Highway Research
Program, SHRP A-369, National Research Council, Washington, D.C.
Bari, J., and M. W. Witczak. (2006). Development of a New Revised Version of the Witczak E* Predictive Model
for Hot Mix Asphalt. Electronic Journal of the Association of Asphalt Paving Technologists, 75, 381–423.
Bell, C. A. (1989). SHRP-A-305: Summary Report on Aging of Asphalt Aggregate Systems. National Research
Council, Washington, D.C.
Bell, C. A., and D. Sosnovske. (1994). SHRP-A-384: Aging: Binder Validation, Strategic Highway Research Program.
National Research Council, Washington, D.C.
Bell, C. A., Y. AbWahab, R. E. Cristi, and D. Sognovske. (1994). SHRP-A-384: Selection of Laboratory Aging,
Strategic Highway Research Program,. National Research Council, Washington, D.C.
Binard, C., D. Anderson, L. Lapalu, and J. P. Planche. (2004). Zero Shear Viscosity of Modified and Unmodified
Binders Proceedings of the 3rd Eurasphalt & Eurobitume congress, Vienna, Vol. 2
Biro, S., T. Gandhi, and S. Amirkhanian. (2009). Determination of Zero Shear Viscosity of Warm Asphalt
Binders. Construction & Building Materials, 23(5), 2080–2086. https://doi.org/10.1016/j.conbuildmat.
2008.08.015.
Blankenship, P. (April 2005). Evaluation of Laboratory Performance Tests for Fatigue Cracking of Asphalt Pave-
ments. Presentation, Federal Highway Administration (FHWA) Asphalt Mixture Expert Task Group (ETG)
Meeting, Fall River, MA.
Braham, A. F., W. G. Buttlar, T. R. Clyne, M. O. Marasteanu, and M. I. Turos. (2009). The Effect of Long-term
Laboratory Aging on Asphalt Concrete Fracture Energy. Journal of the Association of Asphalt Paving Tech-
nologies, Vol. 78.
Branthaver, J. F., J. C. Petersen, R. E. Robertson, J. J. Duvall, S. S. Kim, P. M. Harnsberger, T. Mill, E. K. Ensley,
F. A. Barbour, and J. F. Scharbron. (1993). SHRP-A-368: Binder Characterization and Evaluation–Vol. 2:
Chemistry. National Research Council, Washington, D.C.
Brown, S. F. and T. V. Scholz. (2000). Development of Laboratory Protocols for the Ageing of Asphalt Mixtures.
2nd Eurasphalt and Eurobitume Congress, Barcelona, Spain.
Christensen, D. W., Jr., T. K. Pellinen, and R. F. Bonaquist. (2003). Hirsch Model for Estimating the Modulus of
Asphalt Concrete. Electronic Journal of the Association of Asphalt Paving Technologists, 72, 97–121.

112

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

References  113  

Corbett, L. W., and R. E. Merz. (1975). Asphalt Binder Hardening in the Michigan Test Road after 18 Years of
Service. Transportation Research Record, No. 544, 27–34.
Davison, R. R., J. A. Bullin, C. J. Glover, J. M. Chaffin, G. D. Peterson, K. M. Lunsford, M. S. Lin, M. Liu, and
M. A. Ferry. (1994). Verification of an Asphalt Aging Test and Development of Superior Recycling Agent and
Asphalts, FINAL REPORT, No. FHWA/TX-94/1314-1F, Washington, D.C.
Domke, C. H., R. R. Davison, and C. J. Glover. (2000). Effect of Oxygen Pressure on Asphalt Oxidation Kinetics.
Industrial & Engineering Chemistry Research, 39(3), 592–598. https://doi.org/10.1021/ie9906215.
Dukatz, E. (2015). STH 77 Project Objectives. Presentation, Federal Highway Administration (FHWA) Asphalt
Mixture Expert Task Group (ETG) Meeting, April 2015, Fall River, MA.
Elwardany, M. D., F. Yousefi Rad, C. Castorena, and Y. R. Kim. (2017b). Evaluation of Asphalt Mixture Labora-
tory Long-Term Ageing Methods for Performance Testing and Prediction. Road Materials and Pavement
Design, 18(1), 28–61. https://doi.org/10.1080/14680629.2016.1266740.
Elwardany, M. D., F. Yousefi Rad, C. Castorena, and Y. R. Kim. (2017a). Factors Affecting Oxidation Reaction
Mechanisms in Asphalt Concrete Transportation Research Board 96th Annual Meeting Compendium of
Papers.
Farrar, M. J., R. W. Grimes, S. Wiseman, and J. P. Planche. (2015). Asphalt Pavement–Micro-sampling and Micro-
Extraction Methods, Fundamental Properties of Asphalts and Modified Asphalts III, Quarterly Technical
Report, Federal Highway Administration (FHWA), Contract No. DTFH61-07-D-00005.
Farrar, M. J., J. P. Planche, R. W. Grimes, and Q. Qin. (2014). The Universal Simple Aging Test (USAT): Simulating
Short-and Long Term Hot and Warm Mix Oxidative Aging in the Laboratory. In Y. R. Kim (Ed.), Asphalt
Pavements, 79–87. London: CRC Press, Taylor & Francis Group. https://doi.org/10.1201/b17219-17.
Farrar, M. J., T. F. Turner, J. P. Planche, J. F. Schabron, and P. M. Harnsberger. (2013). Evolution of the Cross­
over Modulus with Oxidative Aging: Method to Estimate Change in Viscoelastic Properties of Asphalt
Binder with Time and Depth on the Road. Transportation Research Record: Journal of the Transportation
Research Board, No. 2370, 76–83. http://dx.doi.org/10.3141/2370-10.
Gatchalian, D., E. Masad, A. Chowdhury, and D. Little. (2006). Characterization of Aggregate Resistance to Deg-
radation in Stone Matrix Asphalt Mixtures. Transportation Research Record: Journal of the Transportation
Research Board, No. 1962, 55–63. https://doi.org/10.3141/1962-07.
Glaser, R., J. Schabron, T. Turner, J. P. Planche, S. Salmans, and J. Loveridge. (2013a). Low-Temperature Oxi-
dation Kinetics of Asphalt Binders. Transportation Research Record: Journal of the Transportation Research
Road, No. 2370, 63–68. https://doi.org/10.3141/2370-08.
Glaser, R., T. F. Turner, J. L. Loveridge, S. L. Salmans, and J. P. Planche. (2013b). Fundamental Properties of
Asphalts and Modified Asphalts III, Quarterly Technical Report, Federal Highway Administration (FHWA),
Contract No. DTFH61-07-D-00005.
Glaser, R., T. F. Turner, J. L. Loveridge, S. L. Salmans, and J. P. Planche. (2015). Aging Master Curve (FP 10) and
Aging Rate Model (FP 11), Fundamental Properties of Asphalts and Modified Asphalt, Volume III, Technical
White Paper, Western Research Institute, Laramie, WY.
Glover, C. J., A. E. Martin, A. Chowdhury, R. Han, N. Prapaitrakul, X. Jin, and J. Lawrence. (2008). Evaluation
of Binder Aging and its Influence in Aging of Hot Mix Asphalt Concrete: Literature Review and Experimental
Design, Report 0-6009-1. Texas Transportation Institute, College Station, TX.
Glover, C. J., R. R. Davison, C. H. Domke, Y. Ruan, P. Juristyarini, D. B. Knorr, and S. H. Jung. (2005). Development
of a New Method for Assessing Asphalt Binder Durability with Field Validation, Federal Highway Administra-
tion (FHWA) Report, No. FHWA/TX-05/1872-2.
Glover, C. J., R. Han, X. Jin, N. Prapaitrakul, Y. Cui, A. Rose, and A. E. Martin. (2014). Evaluation of Binder
Aging and Its Influence in Aging of Hot Mix Asphalt Concrete, Publication FHWA/TX-14/0-6009-2. Federal
Highway Administration (FHWA). U.S. Department of Transportation.
Han, R. (2011). Improvements to a Transport Model of Asphalt Binder Oxidation in Pavements: Pavement Tempera-
ture Modeling, Oxygen Diffusivity in Asphalt Binders and Mastics, and Pavement Air Void Characterization,
Doctoral dissertation, College Station, TX: Texas A&M University.
Herrington, P. R., J. E. Patrick, and G. F. Ball. (1994). Oxidation of Roading Asphalts. Industrial & Engineering
Chemistry Research, 33(11), 2801–2809. https://doi.org/10.1021/ie00035a033.
Hou, T., B. S. Underwood, and Y. R. Kim. (2010). Fatigue Performance Prediction of North Carolina Mix-
tures Using the Simplified Viscoelastic Continuum Damage Model. Electronic Journal of the Association of
Asphalt Paving Technologists, 79, 35–80.
Houston, W. N., M. W. Mirza, C. E. Zapata, and S. Raghavendra. (2005). NCHRP Web-Only Document 113:
Environmental Effects in Pavement Mix and Structural Design Systems. Transportation Research Board
of the National Academies of Sciences, Engineering, and Medicine, Washington, D.C.
Huang, S. C., J. C. Petersen, R. Robertson, and J. Branthaver. (2002). Effect of Hydrated Lime on Long-Term
Oxidative Aging Characteristics of Asphalt. Transportation Research Record: Journal of the Transportation
Research Board, No. 1810, 17–24. https://doi.org/10.3141/1810-03.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

114   Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Hubbard, P., and C. S. Reeve. (1913). The Effect of Exposure on Bitumens. Journal of Industrial and Engineering
Chemistry, 5(1), 15–18. https://doi.org/10.1021/ie50049a009.
Jemison, H. B., B. L. Burr, R. R. Davison, J. A. Bullin, and C. J. Glover. (1992). Application and Use of the ATR,
FT-IR Method to Asphalt Aging Studies. Petroleum Science and Technology Journal, 10(4), 795–808.
Jin, X., R. Han, Y. Cui, and C. J. Glover. (2011). Fast-Rate–Constant-Rate Oxidation Kinetics Model for Asphalt
Binders. Industrial & Engineering Chemistry Research, 50(23), 13373–13379. https://doi.org/10.1021/
ie201275q.
Jones, G. M. (1997). The Effect of Hydrated Lime on Asphalt in Bituminous Pavements. NLA Meeting, Utah
DOT, UT.
Kim, Y.R., M.N. Guddati, Y.T. Choi, D. Kim, A. Norouzi, Y. Wang, B. Keshavarzi, M. Ashouri, A. Ghanbari, A.D.
Wargo, and B.S. Underwood. (2018). Development of Asphalt Mixture Performance-Related Specifications,
Final Report, FHWA Project DTFH61-08-H-00005.
Kumar, A., and W. Goetz. (1977). Asphalt Hardening as Affected by Film thickness, Voids, and Permeability in
Asphaltic Mixtures. Proceedings, Association of Asphalt Paving Technologists, 46, 571–605.
Lamontagne, J., P. Dumas, V. Mouillet, and J. Kister. (2001). Comparison by Fourier Transform Infrared (FTIR)
Spectroscopy of Different Ageing Techniques: Application to Road Bitumens. Fuel, 80(4), 483–488. https://
doi.org/10.1016/S0016-2361(00)00121-6.
Lau, C. K., K. M. Lunsford, C. J. Glover, R. R. Davison, and J. A. Bullin. (1992). Reaction Rates and Harden-
ing Susceptibilities as Determined from Pressure Oxygen Vessel Aging of Asphalts. Transportation Research
Record, No. 1342, 50–57.
Lee, D. Y., and R. J. Huang. (1973). Weathering of Asphalts as Characterized by Infrared Multiple Internal
Reflection Spectra. Journal of Applied Spectroscopy, 27(6), 419–490.
Little, D. N., J. A. Epps, and P. E. Sebaaly. (2006). Hydrated Lime in Hot Mix Asphalt. National Lime Association.
Liu, M., K. M. Lunsford, R. R. Davison, C. Glover, and J. A. Bullin. (1996). The Kinetics of Carbonyl Formation
in Asphalt. American Institute of Chemical Engineers Journal, 42(4), 1069–1076. https://doi.org/10.1002/
aic.690420417.
Lunsford, K. M. (1994). The Effect of Temperature and Pressure on Laboratory Oxidized Asphalt Films with Com-
parison to Field Aging, Ph.D. Dissertation, Texas A&M University, College Station, TX.
Martin, A. E., E. Arambula, F. Yin, L. G. Cucalon, A. Chowdhury, R. Lytton, J. Epps, C. Estakhri, and E. S. Park.
(2014). NCHRP Report 763: Evaluation of the Moisture Susceptibility of WMA Technologies, Transportation
Research Board of the National Academies, Washington, D.C. https://doi.org/10.17226/22429.
Mirza, M. W., and M. W. Witczak. (1995). Development of Global Aging System for Short and Long Term Aging
of Asphalt Cements. Electronic Journal of the Association of Asphalt Paving Technologists, 64, 393–430.
Mollenhauer, K., and V. Mouillet. (2011). Re-road—End of Life Strategies of Asphalt Pavements. European Com-
mission DG Research.
Moraes, R., and H. U. Bahia. (2015). Effect of Mineral Filler on Changes in Molecular Size Distribution of Asphalt
during Oxidative Aging. Road Materials and Pavement Design, 16(sup2), 55–72. https://doi.org/10.1080/
14680629.2015.1076998.
Newcomb, D., A. E. Martin, F. Yin, E. Arambula, E. S. Park, A. Chowdhury, R. Brown, C. Rodezno, N. Tran, E. Coleri,
and D. Jones. (2015). NCHRP Report 815: Short-Term Laboratory Conditioning of Asphalt Mixtures. Trans-
portation Research Board of the National of Academies. Washington, D.C. https://doi.org/10.17226/22077.
Partl, M. N., H. U. Bahia, F. Canestrari, C. De la Roche, H. Di Benedetto, H. Piber, and D. Sybilski. (2013).
Advances in Interlaboratory Testing and Evaluation of Bituminous Materials. RILEM. https://doi.org/10.1007/
978-94-007-5104-0.
Petersen, J. C. (1998). A Dual, Sequential Mechanism for the Oxidation of Petroleum Asphalts. Petroleum Science
and Technology, 16(9-10), 1023–1059. https://doi.org/10.1080/10916469808949823.
Petersen, J. C. (2009). Transportation Research Circular E-C140: A Review of the Fundamentals of Asphalt Oxida-
tion: Chemical, Physicochemical, Physical Property, and Durability Relationships. Transportation Research
Board, Washington, D.C.
Petersen, J. C., and R. Glaser. (2011). Asphalt Oxidation Mechanisms and the Role of Oxidation Products on
Age Hardening Revisited. Road Materials and Pavement Design, 12(4), 795–819. https://doi.org/10.1080/
14680629.2011.9713895.
Petersen, J. C., and P. M. Harnsberger. (1998). Asphalt Aging: Dual Oxidation Mechanism and Its Inter­
relationships with Asphalt Composition and Oxidative Age Hardening. Transportation Research Record,
No. 1638, 47–55. https://doi.org/10.3141/1638-06.
Petersen, J. C., P. M. Harnsberger, and R. E. Robertson. (1996). Factors Affecting the Kinetics and Mecha-
nisms of Asphalt Oxidation and the Relative Effects of Oxidation Products on Age Hardening. Preprints of
Papers, American Chemical Society, Division of Fuel Chemistry, Vol. 41, No. CONF-960807.
Petersen, J. C., H. Plancher, and P. M. Harnsberger. (1987). Lime Treatment of Asphalt to Reduce Age Hardening
and Improve Flow Properties (Vol. 56). Association of Asphalt Paving Technologists.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

References  115  

Prapaitrakul, N. (2009). Toward and Improved Model of Asphalt binder Oxidation in Pavements, Collage Station.
TX: Texas A&M University.
Qi, X., G. Al-Khateeb, T. Mitchell, K. Stuart, and J. Youtcheff. (2004). Determining Modified Asphalt Binder
Properties for the Superpave Specification. Report on the Construction of Pavements with Modified Asphalt
Binders. Federal Highway Administration.
Radovskiy, B. (2003). Analytical Formulas for Film Thickness in Compacted Asphalt Mixture. Transportation
Research Record. Journal of the Transportation Research Board, No. 1829, 26–32. https://doi.org/10.3141/
1829-04.
Recasens, R., A. Martínez, F. Jiménez, and H. BianSchetto. (2005). Effect of Filler on the Aging Potential of
Asphalt Mixtures. Transportation Research Record: Journal of the Transportation Research Board, No. 1901,
10–17.
Reed, J. (2010). Evaluation of the Effects of Aging on Asphalt Rubber Pavements, M.S. Thesis, Arizona State Uni-
versity, Tempe, AZ.
Rowe, G. M., G. King, and M. Anderson. (2014). The Influence of Binder Rheology on the Cracking of
Asphalt Mixes in Airport and Highway Projects. Journal of Testing and Evaluation, 42(5), 1–10. https://
doi.org/10.1520/JTE20130245.
Sabouri, M., and Y. Kim. (2014). Development of a Failure Criterion for Asphalt Mixtures Under Different
Modes of Fatigue Loading. Transportation Research Record: Journal of the Transportation Research Board,
No. 2447, 117–125. http://dx.doi.org/10.3141/2447-13.
Sakhaei Far, M. S. (2011). Development of New Dynamic Modulus (|E*|) Predictive Models for Hot Mix Asphalt
Mixtures, Ph.D. Dissertation, North Carolina State University, Raleigh, NC.
Thurston, R. R., and E. C. Knowles. (1936). Oxygen Absorption Tests on Asphalt Constituents. Journal of Indus-
trial and Engineering Chemistry, 28(1), 88–91. https://doi.org/10.1021/ie50313a023.
Underwood, B. S., M. S. Sakhaei Far, and Y. R. Kim. (2010). Using Limited Purchase Specification Tests to
Perform Full Viscoelastic Characterization of Asphalt Binder. ASTM Journal of Testing and Evaluation.
Van den Bergh, W. (2011). The Effect of Aging on Fatigue and Healing Properties of Bituminous Mortars, Ph.D
Thesis, Delft University of Technology, Delft University, Netherlands.
Van Oort, W. P. (1956). Durability of Asphalt—It’s Aging in the Dark. Journal of Industrial and Engineering
Chemistry, 48(7), 1196–1201. https://doi.org/10.1021/ie50559a033.
Von Quintas, H., J. Scherocman, T. Kennedy, and C. Hughes. (1992). NCHRP Report 338: Asphalt Aggregate
Mixture Analysis System (AAMAS). Transportation Research Board of the National Academies, Washing-
ton, D.C.
Wang, Y., and Y. R. Kim. (2017). Development of a Pseudo Strain Energy-Based Fatigue Failure Criterion for
Asphalt Mixtures. The International Journal of Pavement Engineering, 1–11. https://doi.org/10.1080/
10298436.2017.1394100.
Wright, J. R. (1965). Weathering: Theoretical and Practical Aspects of Asphalt Durability. In A. J. Hoiberg (Ed.),
Bituminous Materials: Asphalts, Tars and Pitches (Vol. II). New York: Interscience Publishers.
Wu, J. T., W. P. Han, G. Airey, and N. I. M. Yusoff. (2014). The Influence of Mineral Aggregates on Bitumen
Ageing. International Journal of Pavement Research and Technology, 7(2), 115–123.
Yousefi Rad, F., M. D. Elwardany, C. Castorena, and Y. R. Kim. (2017). Investigation of Proper Long-Term
Laboratory Aging Temperature for Performance Testing of Asphalt Concrete. Construction & Building
Materials, 147, 616–629. https://doi.org/10.1016/j.conbuildmat.2017.04.197.
Zhang, F., J. Yu, and J. Han. (2011). Effects of Thermal Oxidative Ageing on Dynamic Viscosity, TG/DTG,
DTA and FTIR of SBS-and SBS/Sulfur-Modified Asphalts. Construction & Building Materials, 25(1),
129–137. https://doi.org/10.1016/j.conbuildmat.2010.06.048.

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Appendices A–I

Appendices A though I are not printed herein but are available for download from the TRB
website (trb.org) by searching for “NCHRP Research Report 871.” The appendices include the
following:
Appendix A: Previous Studies of Laboratory Aging of Compacted Specimens and Loose Mixtures
Appendix B: Previous Studies of Modeling Asphalt Binder Aging in Pavements
Appendix C: Evaluation of the Sensitivity of the Mechanical Properties of Asphalt Concrete to
Asphalt Binder Oxidation
Appendix D: Evaluation of Different Chemical and Rheological Aging Index Properties
Appendix E: Factors Affecting Oxidation Reaction Mechanisms in Asphalt Concrete
Appendix F: Evaluation of Asphalt Mixture Laboratory Long-Term Aging
Appendix G: Investigation of Proper Long-Term Aging Temperature
Appendix H: Climatic Aging Index
Appendix I: Performance Testing of Field Cores

116

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

Abbreviations and acronyms used without definitions in TRB publications:


A4A Airlines for America
AAAE American Association of Airport Executives
AASHO American Association of State Highway Officials
AASHTO American Association of State Highway and Transportation Officials
ACI–NA Airports Council International–North America
ACRP Airport Cooperative Research Program
ADA Americans with Disabilities Act
APTA American Public Transportation Association
ASCE American Society of Civil Engineers
ASME American Society of Mechanical Engineers
ASTM American Society for Testing and Materials
ATA American Trucking Associations
CTAA Community Transportation Association of America
CTBSSP Commercial Truck and Bus Safety Synthesis Program
DHS Department of Homeland Security
DOE Department of Energy
EPA Environmental Protection Agency
FAA Federal Aviation Administration
FAST Fixing America’s Surface Transportation Act (2015)
FHWA Federal Highway Administration
FMCSA Federal Motor Carrier Safety Administration
FRA Federal Railroad Administration
FTA Federal Transit Administration
HMCRP Hazardous Materials Cooperative Research Program
IEEE Institute of Electrical and Electronics Engineers
ISTEA Intermodal Surface Transportation Efficiency Act of 1991
ITE Institute of Transportation Engineers
MAP-21 Moving Ahead for Progress in the 21st Century Act (2012)
NASA National Aeronautics and Space Administration
NASAO National Association of State Aviation Officials
NCFRP National Cooperative Freight Research Program
NCHRP National Cooperative Highway Research Program
NHTSA National Highway Traffic Safety Administration
NTSB National Transportation Safety Board
PHMSA Pipeline and Hazardous Materials Safety Administration
RITA Research and Innovative Technology Administration
SAE Society of Automotive Engineers
SAFETEA-LU Safe, Accountable, Flexible, Efficient Transportation Equity Act:
A Legacy for Users (2005)
TCRP Transit Cooperative Research Program
TDC Transit Development Corporation
TEA-21 Transportation Equity Act for the 21st Century (1998)
TRB Transportation Research Board
TSA Transportation Security Administration
U.S.DOT United States Department of Transportation

Copyright National Academy of Sciences. All rights reserved.


Long-Term Aging of Asphalt Mixtures for Performance Testing and Prediction

ADDRESS SERVICE REQUESTED

Washington, DC 20001
500 Fifth Street, NW
TRANSPORTATION RESEARCH BOARD

ISBN 978-0-309-44683-9
NON-PROFIT ORG.
COLUMBIA, MD
PERMIT NO. 88

U.S. POSTAGE

90000
PAID

9 780309 446839

Copyright National Academy of Sciences. All rights reserved.

You might also like