Anda di halaman 1dari 122

THE NATIONAL ACADEMIES PRESS

This PDF is available at http://nap.edu/24850 SHARE


Relationship Between Chemical Makeup of Binders and


Engineering Performance

DETAILS

118 pages | 8.5 x 11 | PAPERBACK


ISBN 978-0-309-39002-6 | DOI 10.17226/24850

CONTRIBUTORS

GET THIS BOOK William H. Daly; National Cooperative Highway Research Program; National
Cooperative Highway Research Program Synthesis Program; Synthesis Program;
Transportation Research Board; National Academies of Sciences, Engineering, and
FIND RELATED TITLES Medicine


Visit the National Academies Press at NAP.edu and login or register to get:

Access to free PDF downloads of thousands of scientic reports


10% off the price of print titles
Email or social media notications of new titles related to your interests
Special offers and discounts

Distribution, posting, or copying of this PDF is strictly prohibited without written permission of the National Academies Press.
(Request Permission) Unless otherwise indicated, all materials in this PDF are copyrighted by the National Academy of Sciences.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

NCHRP
NATIONAL
COOPERATIVE
HIGHWAY
RESEARCH
PROGRAM
SYNTHESIS 511

Relationship Between Chemical


Makeup of Binders and
Engineering Performance

A Synthesis of Highway Practice

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

TRANSPORTATION RESEARCH BOARD 2017 EXECUTIVE COMMITTEE*

OFFICERS
Chair: MALCOLM DOUGHERTY, Director, California Department of Transportation, Sacramento
Vice Chair: KATHERINE F. TURNBULL, Executive Associate Director and Research Scientist, Texas A&M Transportation Institute,
College Station
Executive Director: NEIL J. PEDERSEN, Transportation Research Board

MEMBERS
VICTORIA A. ARROYO, Executive Director, Georgetown Climate Center; Assistant Dean, Centers and Institutes; and Professor
and Director, Environmental Law Program, Georgetown University Law Center, Washington, DC
SCOTT E. BENNETT, Director, Arkansas State Highway and Transportation Department, Little Rock
JENNIFER COHAN, Secretary, Delaware DOT, Dover
JAMES M. CRITES, Executive Vice President of Operations, DallasFort Worth International Airport, TX
NATHANIEL P. FORD, SR., Executive DirectorCEO, Jacksonville Transportation Authority, Jacksonville, FL
A. STEWART FOTHERINGHAM, Professor, School of Geographical Sciences and Urban Planning, Arizona State University, Tempe
JOHN S. HALIKOWSKI, Director, Arizona DOT, Phoenix
SUSAN HANSON, Distinguished University Professor Emerita, Graduate School of Geography, Clark University, Worcester, MA
STEVE HEMINGER, Executive Director, Metropolitan Transportation Commission, Oakland, CA
CHRIS T. HENDRICKSON, Hamerschlag Professor of Engineering, Carnegie Mellon University, Pittsburgh, PA
JEFFREY D. HOLT, Managing Director, Power, Energy, and Infrastructure Group, BMO Capital Markets Corporation, New York
S. JACK HU, Vice President for Research and J. Reid and Polly Anderson Professor of Manufacturing, University of Michigan, Ann Arbor
ROGER B. HUFF, President, HGLC, LLC, Farmington Hills, MI
GERALDINE KNATZ, Professor, Sol Price School of Public Policy, Viterbi School of Engineering, University of Southern California,
Los Angeles
MELINDA McGRATH, Executive Director, Mississippi DOT, Jackson
PATRICK K. McKENNA, Director, Missouri DOT, Jefferson City
JAMES P. REDEKER, Commissioner, Connecticut DOT, Newington
MARK L. ROSENBERG, Executive Director, The Task Force for Global Health, Inc., Decatur, GA
DANIEL SPERLING, Professor of Civil Engineering and Environmental Science and Policy; Director, Institute of Transportation Studies,
University of California, Davis
GARY C. THOMAS, President and Executive Director, Dallas Area Rapid Transit, Dallas, TX
PAT THOMAS, Senior Vice President of State Government Affairs, United Parcel Service, Washington, DC
JAMES M. TIEN, Distinguished Professor and Dean Emeritus, College of Engineering, University of Miami, Coral Gables, FL
DEAN H.WISE, Vice President of Network Strategy, Burlington Northern Santa Fe Railway, Fort Worth, TX
CHARLES A. ZELLE, Commissioner, Minnesota DOT, Saint Paul

EX OFFICIO MEMBERS
ALBERTO AYALA, Deputy Executive Officer, California Air Resources Board, Sacramento
MARY R. BROOKS, Professor Emerita, Dalhousie University, Halifax, Nova Scotia, Canada, and Chair, TRB Marine Board
JACK DANIELSON, Executive Director, National Highway Traffic Safety Administration, U.S. DOT
AUDREY FARLEY, Executive Director, Office of the Assistant Secretary for Research and Technology, U.S. DOT
LeROY GISHI, Chief, Division of Transportation, Bureau of Indian Affairs, U.S. Department of the Interior, Washington, DC
JOHN T. GRAY II, Senior Vice President, Policy and Economics, Association of American Railroads, Washington, DC
MICHAEL P. HUERTA, Administrator, Federal Aviation Administration, U.S. DOT
DAPHNE Y. JEFFERSON, Deputy Administrator, Federal Motor Carrier Safety Administration, U.S. DOT
BEVAN B. KIRLEY, Research Associate, University of North Carolina Highway Safety Research Center, Chapel Hill, and Chair,
TRB Young Members Council
HOWARD McMILLAN, Acting Administrator, Pipeline and Hazardous Materials Safety Administration, U.S. DOT
WAYNE NASTRI, Acting Executive Officer, South Coast Air Quality Management District, Diamond Bar, CA
CRAIG A. RUTLAND, U.S. Air Force Pavement Engineer, U.S. Air Force Civil Engineer Center, Tyndall Air Force Base, FL
REUBEN SARKAR, Deputy Assistant Secretary for Transportation, U.S. Department of Energy
TODD T. SEMONITE (Lieutenant General, U.S. Army), Chief of Engineers and Commanding General, U.S. Army Corps of Engineers,
Washington, DC
KARL SIMON, Director, Transportation and Climate Division, U.S. Environmental Protection Agency
JOEL SZABAT, Executive Director, Maritime Administration, U.S. DOT
WALTER C. WAIDELICH, JR., Acting Deputy Administrator, Federal Highway Administration, U.S. DOT
PATRICK T. WARREN, Executive Director, Federal Railroad Administration, U.S. DOT
MATTHEW WELBES, Executive Director, Federal Transit Administration, U.S. DOT
RICHARD A. WHITE, Acting President and CEO, American Public Transportation Association, Washington, DC
FREDERICK G. (BUD) WRIGHT, Executive Director, American Association of State Highway and Transportation Officials, Washington, DC
PAUL F. ZUKUNFT (Admiral, U.S. Coast Guard), Commandant, U.S. Coast Guard, U.S. Department of Homeland Security
* Membership as of April 2017.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

NATIONAL COOPERATIVE HIGHWAY RESEARCH PROGRAM

NCHRP SYNTHESIS 511


Relationship Between Chemical Makeup of
Binders and Engineering Performance

A Synthesis of Highway Practice

Consultant
William H. Daly
Baton Rouge, Louisiana

S ubscriber C ategories
Construction Highways Materials Pavement

Research Sponsored by the American Association of State Highway and Transportation Officials
in Cooperation with the Federal Highway Administration

2017

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

NATIONAL COOPERATIVE HIGHWAY RESEARCH PROGRAM NCHRP SYNTHESIS 511


Systematic, well-designed research provides the most effective Project 20-05 (Topic 47-13)
approach to the solution of many problems facing highway administra- ISSN 0547-5570
tors and engineers. Often, highway problems are of local interest and ISBN 978-0-309-39002-6
can best be studied by highway departments individually or in coop- Library of Congress Control No. 2017936212
eration with their state universities and others. However, the accelerat-
ing growth of highway transportation develops increasingly complex
problems of wide interest to highway authorities. These problems are 2017 National Academy of Sciences. All rights reserved.
best studied through a coordinated program of cooperative research.
In recognition of these needs, the highway administrators of the
American Association of State Highway and Transportation Offi- COPYRIGHT INFORMATION
cials initiated in 1962 an objective national highway research pro-
Authors herein are responsible for the authenticity of their manuscripts
gram employing modern scientific techniques. This program is sup-
and for obtaining written permissions from publishers or persons who
ported on a continuing basis by funds from participating member
own the copyright to any previously published or copyrighted material
states of the Association and it receives the full cooperation and sup-
used herein.
port of the Federal Highway Administration, United States Depart-
Cooperative Research Programs (CRP) grants permission to repro-
ment of Transportation.
duce material in this publication for classroom and not-for-profit pur-
The Transportation Research Board of the National Research Coun-
poses. Permission is given with the understanding that none of the
cil was requested by the Association to administer the research pro-
material will be used to imply TRB, AASHTO, FAA, FHWA, FMSCA,
gram because of the Boards recognized objectivity and understanding
FTA, or Transit development Corporation endorsement of a particular
of modern research practices. The Board is uniquely suited for this
product, method, or practice. It is expected that those reproducing the
purpose as it maintains an extensive committee structure from which
material in this document for educational and not-for-profit uses will
authorities on any highway transportation subject may be drawn; it
give appropriate acknowledgment of the source of any development or
possesses avenues of communication and cooperation with federal,
reproduced material. For other uses of the material, request permission
state, and local governmental agencies, universities, and industry; its
from CRP.
relationship to the National Research Council is an insurance of objec-
tivity; it maintains a full-time research correlation staff of specialists
in highway transportation matters to bring the findings of research
NOTICE
directly to those who are in a position to use them.
The program is developed on the basis of research needs identified The report was reviewed by the technical panel and accepted for
by chief administrators of the highway and transportation departments publication according to procedures established and overseen by the
and by committees of AASHTO. Each year, specific areas of research Transportation Research Board and approved by the National Acad-
needs to be included in the program are proposed to the National emies of Sciences, Engineering, and Medicine.
Research Council and the Board by the American Association of State The opinions and conclusions expressed or implied in this report are
Highway and Transportation Officials. Research projects to fulfill those of the researchers who performed the research and are not neces-
these needs are defined by the Board, and qualified research agencies sarily those of the Transportation Research Board; the National Acade-
are selected from those that have submitted proposals. Administration mies of Sciences, Engineering, and Medicine; or the program sponsors.
and surveillance of research contracts are the responsibilities of the The Transportation Research Board; the National Academies of
National Research Council and the Transportation Research Board. Sciences, Engineering, and Medicine; and the sponsors of the National
The needs for highway research are many, and the National Coop- Cooperative Highway Research Program do not endorse products or
erative Highway Research Program can make significant contributions manufacturers. Trade or manufacturers names appear herein solely
to the solution of highway transportation problems of mutual concern because they are considered essential to the object of the report.
to many responsible groups. The program, however, is intended to
complement rather than to substitute for or duplicate other highway
research programs.

Published 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 at:


http://www.national-academies.org/trb/bookstore

Printed in the United States of America

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

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 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 Boards 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.


Relationship Between Chemical Makeup of Binders and Engineering Performance

TOPIC PANEL 47-13


WILLIAM E. AHEARN, Barre, VT
GAYLON L. BAUMGARDNER, Paragon Technical Services, Inc., Jackson, MS
AMIT BHASIN, University of Texas at Austin
LYNDI D. BLACKBURN, Alabama Department of Transportation, Montgomery
NICHOLAS I. BURMAS, California Department of Transportation, Sacramento
ALLEN J. GALLISTEL, Minnesota Department of Transportation, Maplewood
DARREN G. HAZLETT, Texas Department of Transportation, Austin
FREDERICK HEJL, Transportation Research Board
DEREK NENER-PLANTE, Mane Department of Transportation Cabinet, Augusta
JACK YOUTCHEFF, Federal Highway Administration (Liaison)

SYNTHESIS STUDIES STAFF


STEPHEN R. GODWIN, Director for Studies and Special Programs
JON M. WILLIAMS, Program Director, IDEA and Synthesis Studies
MARIELA GARCIA-COLBERG, Senior Program Officer
JO ALLEN GAUSE, Senior Program Officer
THOMAS HELMS, Consultant
GAIL R. STABA, Senior Program Officer
TANYA M. ZWAHLEN, Consultant
DON TIPPMAN, Senior Editor
CHERYL KEITH, Senior Program Assistant
DEMISHA WILLIAMS, Senior Program Assistant
DEBBIE IRVIN, Program Associate

COOPERATIVE RESEARCH PROGRAMS STAFF


CHRISTOPHER J. HEDGES, Director, Cooperative Research Programs
LORI L. SUNDSTROM, Deputy Director, Cooperative Research Programs
EILEEN P. DELANEY, Director of Publications

NCHRP COMMITTEE FOR PROJECT 20-05

CHAIR
BRIAN A. BLANCHARD, Florida Department of Transprtation

MEMBERS
STUART D. ANDERSON, Texas A&M University
SOCORRO COCO BRISENO, California Department of Transportation
DAVID M. JARED, Georgia Department of Transportation
CYNTHIA L. JONES, Ohio Department of Transportation
MALCOLM T. KERLEY, NXL, Richmond, VA
JOHN M. MASON, JR., Auburn University
ROGER C. OLSON, Minnesota Department of Transportation (retired)
BENJAMIN T. ORSBON, South Dakota Department of Transportation
RANDALL R. PARK, Utah Department of Transportation
ROBERT L. SACK, New York State Department of Transportation
FRANCINE SHAW WHITSON, Federal Highway Administration
JOYCE N. TAYLOR, Maine Department of Transportation

FHWA LIAISON
JACK JERNIGAN

TRB LIAISON
STEPHEN F. MAHER

Cover figure: AFM phase images of 25 m by 25 m of Asphalt ARC BI0002 derivatives taken at room temperature
(25C): (a) control; (b) asphaltene-doped blend; (c) naphthenic aromatic-doped blend; (d) polar (resin)-doped blend; and
(e) saturate-doped blend (enlarged locations are 10 m by 10 m for each asphalt blend). (Source: Allen et al. 2014.)

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

FOREWORD Highway administrators, engineers, and researchers often face problems for which
information already exists, either in documented form or as undocumented experience
and practice. This information may be fragmented, scattered, and unevaluated. As a con-
sequence, full knowledge of what has been learned about a problem may not be brought to
bear on its solution. Costly research findings may go unused, valuable experience may be
overlooked, and due consideration may not be given to recommended practices for solving
or alleviating the problem.
There is information on nearly every subject of concern to highway administrators and
engineers. Much of it derives from research or from the work of practitioners faced with
problems in their day-to-day work. To provide a systematic means for assembling and
evaluating such useful information and to make it available to the entire highway commu-
nity, the American Association of State Highway and Transportation Officialsthrough
the mechanism of the National Cooperative Highway Research Programauthorized the
Transportation Research Board to undertake a continuing study. This study, NCHRP Proj-
ect 20-5, Synthesis of Information Related to Highway Problems, searches out and syn-
thesizes useful knowledge from all available sources and prepares concise, documented
reports on specific topics. Reports from this endeavor constitute an NCHRP report series,
Synthesis of Highway Practice.
This synthesis series reports on current knowledge and practice, in a compact format,
without the detailed directions usually found in handbooks or design manuals. Each report
in the series provides a compendium of the best knowledge available on those measures
found to be the most successful in resolving specific problems.

PREFACE This synthesis documents the current practices of departments of transportation (DOTs)
in the selection of the chemical composition of a binder used in pavement applications. The
Mariela Garcia-
study will provide information to practice engineers about the selection of binders and post-
Colberg production additives and modifiers, as well as corresponding engineering performance.
Senior Program Officer A literature review and detailed survey responses from 45 out of 51 DOTs (88.2%
response rate) are provided. Detailed case examples from four different states are also
included in the report and provide additional insights on the state-of-the-practice, including
lessons learned, challenges, and gaps in information.
The synthesis will be useful for transportation agencies because it will provide their
engineers with the chemical composition knowledge they need. It will help these practitio-
ners make more informed choices and avoid potential costly mistakes.
William Daly collected and synthesized the information and wrote the report. The
members of the topic panel are acknowledged on page iv. This synthesis is an immediately
useful document that records the practices that were acceptable within the limitations of
the knowledge available at the time of its preparation. As progress in research and practice
continues, new knowledge will be added to that now at hand.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

CONTENTS

1 SUMMARY

7 CHAPTER ONE INTRODUCTION


Background, 7
Superior Performing Asphalt Pavements, 7
Chemical Analysis of Asphalt, 8
Synthesis Approach, 9

10 CHAPTER TWO LITERATURE REVIEW: CHEMICAL AND PHYSICAL CHARACTERIZATION OF ASPHALT BINDERS
Asphalt Production, 10
Asphalt Composition, 12
Asphaltene Structures, 14
Asphalt Morphology, 16

19 CHAPTER THREE TECHNIQUES FOR CHEMICAL CHARACTERIZATION OF ASPHALT


Asphalt Fractionation, 19
Analytical Instrumentation for Asphalt Analysis, 24
Fourier Transform Infrared Spectroscopy, 26
Gel Permeation Chromatography, 29
Nuclear Magnetic Resonance Analysis, 33
X-ray Fluorescence Spectroscopy, 36
Atomic Force Microscopy, 37

41 CHAPTER FOUR ASPHALT ADDITIVES AND MODIFIERS


Introduction, 41
Polymer Additives, 42
Biobinders, 45
Other Nonbituminous Modifiers, 46

53 CHAPTER FIVE CASE EXAMPLES OF BINDER CHARACTERIZATION PRACTICES


Introduction, 53
Louisiana, 53
Virginia, 55
Missouri, 57
Ontario, 58
Summary, 59

61 CHAPTER SIX SURVEY RESULTS: CURRENT U.S. AND CANADIAN EXPERIENCE

66 CHAPTER SEVEN CONCLUSIONS


Applications of Analytical Instrumentation to Binder Characterization, 66
Survey Findings, 67
Suggested Future Research, 68

Note: Many of the photographs, figures, and tables in this report have been converted from color to grayscale for printing.
The electronic version of the report (posted on the web at www.trb.org) retains the color versions.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

viii

70 ACRONYMS AND ABBREVIATIONS

73 REFERENCES

83 APPENDIX A SURVEY QUESTIONS AND RESULTS

108 APPENDIX B SURVEY RESPONDENTS

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

RELATIONSHIP BETWEEN CHEMICAL MAKEUP OF


BINDERS AND ENGINEERING PERFORMANCE

SUMMARY A number of factors and variables influence the chemical composition and, ultimately, the
engineering properties of a binder used in pavement applications. It is important that the
users most closely working with binders have a strong working knowledge of the relation-
ship between binder chemistry and engineering properties. In addition, the relationship
among standard process control actions and final chemical composition needs to be under-
stood so that correlation to physical performance can be made.

This synthesis can provide the chemical composition knowledge to help departments of
transportation (DOTs) and practicing engineers make more informed choices with regard
to selection of binders and post-production additives and modifiers and avoid potentially
expensive mistakes. Also, researchers might develop more informed methods to design
binder modifiers and additives that will result in low-cost and high-performance binders
for specific applications.

The information in this synthesis was gathered by a thorough review of the available
U.S. and international literature. A comprehensive literature search was performed using
the TRID database, SciFinder, Bing, and Google. In addition, a survey of U.S. state DOTs
and Canadian ministries of transportation was conducted to determine the current status
of asphalt chemical analysis. The U.S. state and District of Columbia response rate to
the survey was 88.2% (45 of 51). Only eight of the 14 Canadian provinces and territories
responded, but their comments were enlightening.

This synthesis includes an updated literature review on binder chemistry and the rela-
tionship between binder chemistry (including modifiers and additives) and engineering
properties. Analytical methods that can be used to qualitatively identify the chemical
makeup of asphalt binders based on modern analytical instrumentation are comprehen-
sively reviewed. Composition information is useful for understanding asphaltwhat
makes it behave as it does and what makes one asphalt behave differently from another.
With the asphalt sources available, composition information can be used to improve the
product through modification with additives, by blending, and so on, or to alter use design
procedures to accommodate specific properties. Composition information can be used to
match asphalt and aggregate, provide clues as to what modifications are necessary to make
an asphalt-aggregate system more serviceable under a given environment, diagnose fail-
ures, and provide information needed for corrective measures. This synthesis reviews our
current understanding of asphalt composition.

Only a limited number of crude oils yield quality asphalt. The price of liquid asphalt
has increased dramatically within the past decade mainly as a result of the advent of cok-
ing technologies, which enabled refineries to decrease their production of asphalt (residua
from crude oil refining processes) by converting the residua into synthetic fuel. This devel-
opment has led to a shortage of asphalt binders and an increase in its price independent
of the crude oil price. Asphalt usage, as reported by the Asphalt Institute, is summarized

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

2

according Petroleum Administration for Defense Districts. Trends in net production of


asphalt and road oil by refineries and blenders as reported by the U.S. Energy Information
Administration show a decline in production over the past 5 years based on 2010 production
figures. Demand for asphalt in paving applications is forecast to advance 3.1% annually to
19.6 million tons based on improving economic conditions and a pressing need to repair and
expand the nations infrastructure. Rising use of recycled asphalt pavement and increas-
ing interest in rehabilitating and repairing older or worn surfaces will serve as a check on
asphalt demand advances.

Asphalt (or asphalt cement) is the carefully refined product derived from selected crude
oils. Outside the United States, the product is often called bitumen. Asphalt is no longer
just the residua from crude oil refining. It is now known as an asphalt binder and is part of
an engineered system. All molecules in asphalt are hydrocarbons with small amounts of
sulfur, nitrogen, and oxygen and traces of metals such as vanadium and nickel. The hydro-
carbons may consist of polyaromatic structures containing different numbers of fused rings,
saturated polycyclic structures also with different numbers of rings, and combinations of
these. All these core structures contain saturated hydrocarbon side chains of different chain
lengths and different substitution patterns. Many of these side chains are lost during the
refining process, so the properties of the asphalt produced today are different from those of
the asphalt that was simply the residue of a crude oil distillation.

One school of thought is that asphalt could be considered a colloidal or micellar system.
The hydrocarbon insoluble components, asphaltenes and resins, are dispersed in a hydro-
carbon blend. The overall behavior of asphalt cement is controlled by the compatibility and
the relationships of the different components in this macroscopically homogeneous mixture
rather than by the quantitative amount of any single component. Historically, the study of
asphalt chemical composition has been facilitated by the separation of asphalt into compo-
nent fractions based on the polarity of the molecular components present or their adsorp-
tion characteristics, or both. The component fractions, sometimes called generic fractions,
although useful in classifying and characterizing asphalts and in providing simplified mix-
tures for further study, are still complex mixtures, and their composition is a function of
asphalt source. The component fractions are, however, sufficiently unique to identify their
particular contribution to the complex flow properties of asphalt. A proper balance of com-
ponent types is necessary for a durable asphalt. The method for fractionating asphalts into
the generic fractionssaturates, aromatics, resins, and asphalteneshas evolved recently,
and the new developments are discussed in this synthesis.

Asphaltenes are the insoluble fraction precipitated from a toluene solution of asphalt
binder by a nonpolar solvent such as pentane or heptane. The precipitating solvent affects the
quantity of asphaltenes precipitated. The nature of the asphaltene molecules in the precipi-
tate depends on the precipitating solvent; greater amounts of a precipitant are produced using
pentane. In general, the asphaltenes are defined as the mixture of materials precipitated by
heptane. The asphaltene molecules are very complex and exhibit a very high tendency to
associate into molecular clusters. The amount and characteristics of asphaltenes vary from
one asphalt source to another. They play a significant role in the rheology of asphalt binders.
The current status of our understanding of asphaltene structure is discussed in chapter two.

Recently, Mullins (2011) published a comprehensive review on asphaltenes. Virtually


all asphaltene chemical properties, except elemental composition, had been the subject of
debate; their molecular weight had been estimated at values spanning six orders of mag-
nitude. Molecular weight determination has been, and still is, a challenging problem in
asphaltene chemistry. The source of complications is attributable to three basic properties
of asphaltene, namely, compositional variance, size polydispersity, and, most important, the
high propensity of the covalent asphaltene molecules to form molecular aggregates.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 3

In contrast, a second school of thought is that different molecular species in the asphalt
binder exist in a colloidal dispersion of asphaltene micelles in the maltenes. Lesueur has
published a detailed review of the current asphalt structure concepts that defines links
between chemistry, structure, and mechanical properties in a framework of an updated col-
loidal picture of asphalt. The resinsthat is, the polar components of the malteneswere
thought to stabilize the asphaltene micelles. Extensive analytical procedures are support-
ing this model.

Traditional analytical techniques like ultraviolet spectroscopy, Fourier transform infra-


red spectroscopy (FTIR), nuclear magnetic resonance (NMR), or mass spectroscopy have
provided substantial information about the average chemical composition of asphalt. Most
asphalts give more or less identical spectra, and there are no known general correlations
between physical properties and any particular functional group as identified by the tech-
niques listed. The only technique that is frequently used is FTIR, which permits identifica-
tion of carbonyls and sulfoxides formed as a result of aging. Modern analytical techniques
allow more complete definition of the typical asphalt molecules. These techniques include
thermogrammetric analysis (TGA), differential thermal analysis, gel permeation chro-
matography (GPC), scanning electron microscopy, and atomic force microscopy (AFM).
These modern techniques provide insight into the molecular interactions that govern the
physical properties of the asphalt matrix. Changes in the physical properties produced by
specific additives can be ascertained to estimate their effectiveness. These techniques are
described in chapter three.

Renewed interest in asphalt morphology has been promoted by thermal analysis


research. TGA and differential scanning calorimetry are used to characterize petroleum
bitumens and their chromatographic fractions, including the glass transition temperature
and the percentage of crystalline fractions. The influence of the different constituents on
the thermal stability of bitumen can be studied by TGA. TGA measurements provided a
simple means to determine the thermal stability of bitumen and the presence of volatile
species in binders. The advent of modulated differential scanning calorimetry provides
new insight on asphalt microstructure. The development of bitumen microstructure and the
calculations of the entropy and enthalpy of transitions suggest that bitumen is a structured
amorphous phase containing small crystalline phase.

FTIR is one of the more important methods for fingerprinting asphalt materials and
quantifying the distribution of asphalt components. By determining the various chemical
functional groups in the binder, an understanding of its origin and history can be obtained.
The FTIR method is an efficient technique for identifying polymer additives in a binder.
Determination of polymer content is essential for quality control and quality assurance
during the processing and application of polymer-modified asphalts.

Differences in the molecular structures of asphalt components have prompted efforts to


separate these components using size exclusion chromatography, more commonly known
as gel permeation chromatography. GPC is a method of separating molecules based on
their size and shape in solution. GPCs ability to separate mixtures by molecular size rather
than by some complex property such as solubility or absorptivity is one of the great advan-
tages of the technique. This feature has made GPC a useful alternate technique for frac-
tionating complicated mixtures such as crude oil residua, asphalts, and asphaltenes and
a vital contributor to understanding asphalt mixes. The differences in molecular weights
and molecular weight distributions identified by the GPC data are used to predict aging,
viscosity aging index, viscosity number, and penetration. Because the polymer and asphalt
components of polymer-modified asphalt cements can be separated completely, GPC is an
excellent tool for measuring polymer content in modified binders.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

4

Proton nuclear magnetic resonance (1H NMR) spectroscopy has emerged as a very pow-
erful and versatile tool for bitumen characterization. Using 1H and 13C NMR can yield infor-
mation on average structural parameters of asphalt and asphaltenes, such as percentages of
aromatic carbons, aliphatic carbons, bridged carbons, methyl carbons, ring carbons, naph-
thenic carbons, paraffinic chain lengths, and other parameters. The specific environment of
the different types of hydrogens and carbons can be defined. Thus, NMR spectroscopy is a
powerful tool for predicting the structure of complex organic molecules. When information
from NMR and GPC is combined, possible structures for asphalt and mechanisms of aging
can be suggested.

Recently, researchers combined atomic resolution imaging using AFM with molecular
orbital imaging using scanning tunneling microscopy to show the actual atomic arrange-
ments in an asphaltene molecule. Identifying molecular structures provides a foundation to
understand all aspects of petroleum science, from colloidal structure and interfacial interac-
tions to petroleum thermodynamics, thus enabling a first-principles approach to optimiz-
ing resource utilization. The findings contribute to a long-standing debate about asphaltene
molecular architecture. The impact of AFM imaging on understanding the microstructure
and performance characteristics of an asphalt binder is immense. Certain asphalt chemical
parameters have a consistent and measurable effect on the asphalt microstructure that can
be observed with AFM.

As new efficiency allows refiners to extract more gasoline and other petroleum products
from crude oil and as the source of crudes that yield quality asphalt residua decreases, the
need for additives to upgrade straight-run asphalts increases. The most common additives
to asphalt binders are polymers (styrene-butadiene-styrene and styrene-butadiene rubber,
antistripping agents, softening agents, and polyphosphoric acid). The properties and contri-
butions of each of these nonbituminous additives are reviewed in chapter four.

Polymers are used to enable a wider spread of high and low performance grading (PG)
values for binder blends. The benefit of the polymer modifiers will depend on the concentra-
tion, morphology, molecular weight, chemical composition, and molecular structure of the
material. The crude source, refining process, and grade of the neat asphalt binder are equally
important. The residual reactivity of the polymer is useful for improving compatibility of the
additive with the binder.

Bio-based alternatives, which are being developed across the industry in various coun-
tries, could be a solution to reduce the asphalt industrys dependence on petroleum resources.
In addition to efforts to create alternative binders, bio-based additives and extenders are
being developed. Recent developments in this field are described in chapter four.

Low molecular weight modifiers are specifically used to improve the asphalt binders
ability to perform in a given environment. These materials are often proprietary, but they
can be classified by their intended application; that is, antioxidants, hydrocarbon supple-
ments, antistripping agents, and stiffening agents. The hot-mix recycling operation for bitu-
minous mixes commonly uses a modifier to restore the aged asphalt cement to a condition
that resembles virgin asphalt cement. The contribution of each of these modifiers to asphalt
binder chemistry is discussed in chapter four. The advantages and disadvantages of their
employment are elaborated.

The survey results in chapter six cover numerous aspects of asphalt chemistry. State
DOTs do not monitor the sources of their asphalt binders. Only 13.3% of the respondents
could identify their sources. Similarly, most DOTs are not familiar with the process used to
produce their binders. Significant differences in asphalt binder properties over the past 10
years were reported by 37.8% of the respondents. Binders appeared to be stiffer and exhibit
poor low-temperature properties, and some pavements aged prematurely. Although binders

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 5

met specifications, experienced engineers could detect quarterly changes as crude feeds
changed. Changes associated with seasonal or market fluctuations were reported by 35.5%
of the respondents.

The dominant analytical procedure used for binder testing is rheological; that is, the PG
protocol (100% of respondents). FTIR is employed by 24.4% of the DOTs, but this appears
to be primarily used for research. The third most common procedure was X-ray fluores-
cence. A number of modifiers/additives are currently employed, including polyphosphoric
acid (42.2% of respondents), antistripping agents (26.7%), and softening agents (22.2%).
Comments on specific DOT procedures can be found in Appendix A. More than 95% of the
40 respondents replying to the questionnaire do not measure the compatibility of modifiers
in the final binder.

It is hoped that this review will bring about a better understanding of the chemical
compositional factors that control the properties of asphalt and will assist in providing
direction to both research and application, and lead to improved asphalt products with bet-
ter performance and durability.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 7

CHAPTER ONE

INTRODUCTION

This chapter introduces background information and highlights the objectives, organization, and key definitions used in the
report. This synthesis includes an updated literature review on binder chemistry and the relationship between binder chem-
istry (including modifiers and additives) and engineering properties. Analytical methods that can be used to qualitatively
identify the chemical makeup of asphalt binders based on modern analytical instrumentation are comprehensively reviewed.
Compositional information is useful in helping to understand asphaltwhat makes it behave as it does and what makes
one asphalt behave differently from another. The efficiency of the refiners and producers to extract crude oil for reservoirs
increases the source of crudes, but the number of crudes that yield quality asphalt residua is limited. As refiners extract more
gasoline and other petroleum products from crude oil, the need for additives to upgrade straight-run asphalts increases. The
properties and contributions of nonbituminous additives are reviewed.

BACKGROUND

Asphalt pavements were first constructed in the 1870s using asphalt mined from a lake of native asphalt in Trinidad. The
asphalt was diluted with petroleum residuum to make an asphalt cement of the desired consistency (Welborn 1984). As early
as 1892, scientists realized that the durability of asphalt pavements depended on an asphalt cement capable of adhering to the
aggregate while remaining elastic. In 1903, an ASTM committee on Road and Paving Materials Procedures was formed to
develop test methods and specifications for highway materials. Test methods, which included volatilization, penetration, and
bitumen quality, were adopted by ASTM in 1911. As asphalt characterization techniques continued to evolve, three national
specificationsfederal, AASHTO, and ASTMfor asphalt cements to be used in hot-mix asphalt (HMA) were published.
With minor exceptions, the requirements for physical and chemical properties were essentially the same for the three national
specifications. The specification tests included penetration, flash point, ductility, and loss on heating (Welborn 1984).

The specifications proved to be inadequate for predicting the field hardening of the HMAs. Laboratory accelerated tests
were developed to assess age and oxidative hardening. Relating these tests to field performance became possible when Abson
developed a procedure for recovering asphalt liquid from asphalt-aggregate mixture in 1933. This procedure has evolved
to the standard used today (AASHTO 2011). Various thin-film oven (TFO) and rolling thin-film oven (RTFO) tests were
developed to assess the extent of hardening, since hardening was judged to be the one property most closely related to asphalt
performance (AASHTO 2013). These tests predict the properties of asphalt at the time of construction, but they do not provide
sufficient information on the change in properties of the asphalt during service in a pavement.

SUPERIOR PERFORMING ASPHALT PAVEMENTS

Strategic Highway Research Program (SHRP) research had a significant impact on characterization and specification of
asphalt binders through development of more fundamental rheology-based test methods. Advances in instrumentation and
personal computers permitted routine application of rheological methods in conducting and determining specification compli-
ance. In 1987, the SHRP began developing a new system for specifying asphalt materials; that is, Superpave, which stands for
Superior Performing Asphalt Pavements. The goal was to develop an improved system for specifying component materials,
asphalt mixture design and analysis, and pavement performance prediction (SHRP 1993a, b). As a result of SHRP research,
the 1990s saw the introduction of a new binder purchase specification now known as the Superpave binder specification. The
Superpave binder specification is based on rheological properties of the asphalt binder measured over a wide range of tempera-
tures and aging conditions. Measuring binders rheological properties over a wide range of temperatures, loading conditions,
and aging conditions allows performance relationships to be established between the test results and the pavement. The details
of this asphalt binder testing are described in an AASHTO Specification (AASHTO 2016).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

8

The new system for specifying binders is performance-based. It specifies binders on the basis of climate and attendant
pavement temperatures in which the binder is designed to serve. Physical criteria remain the same, but the temperature at
which the binder must attain the properties changes. Performance graded (PG) binders are designated by high and low tem-
perature limits. For example, in the designation PG 64-22, the first number, 64, is the high temperature grade, which desig-
nates the highest temperature in centigrade at which the binder exhibits adequate physical properties. This would correspond
to the average high temperature for the climate where the binder is expected to serve. The second number (-22) is the low
temperature grade; that is, the binder will meet specified physical properties in pavements down to at least 22oC.

A key feature of binder evaluation in the Superpave system is that physical properties are measured on binders. After test-
ing the unaged binder, the physical properties are also measured on binders exposed to rolling thin-film oven aging to simulate
oxidative hardening that occurs during hot mixing and installation. A pressure aging vessel (PAV) is used to expose the binder
to severe aging expected after it has served many years in the field. Various pieces of equipment are used to measure stress
strain relationships in the binder at the specified test temperatures. The equipment includes the dynamic shear rheometer
(DSR) and bending beam rheometer (BBR). These devices and others that may be used to characterize a binder, along with
the purpose of each test, are summarized in Table 1. The procedures focus on the physical properties of asphalt binders and
their respective asphalt concretes (HMAs).

TABLE 1
SUPERPAVE ASPHALT BINDER TESTING EQUIPMENT AND PURPOSE
Device Purpose Performance Parameter
Rolling Thin-Film Oven Simulate binder aging during hot-mix asphalt (HMA) production Resistance to aging and adequate stiffness during
and construction construction
Pressure Aging Vessel Simulate binder aging during HMA service life Resistance to aging (durability) in the field
Rotational Viscometer Measure binder properties at high and intermediate construction Facility in handling and pumping
temperatures
Dynamic Shear Viscometer Measure binder properties at high and intermediate service Resistance to permanent deformation (rutting) and
temperatures fatigue cracking
Bending Beam Rheometer Measure binder properties at low service temperatures Resistance to thermal cracking
Direct Tension Tester Measure binder properties at low service temperatures Resistance to thermal cracking

Understanding the chemistry of asphalt binders is complicated because asphalt composition is complex. Asphalt is the
residue from fractional distillation of crude oil, so its composition varies with the crude oil source (Corbett 1984). The distil-
lation process may be followed by further separation of the vacuum residue by solvent deasphalting in order to extract high
boiling fractions for lube oils or as a feedstock for catalytic cracking. The insoluble fraction (precipitated asphalt) may be used
as a blending component for asphalt cements. Residuum oil supercritical extraction uses low molecular weight hydrocarbon
solvents (propane or pentane) for production of asphaltenes, resins, and oils by extraction of the nonvolatile constituents of
petroleum crudes. Solvent selection depends on the nature of the feedstock and product usage (Gearhart and Garwin 1976).
If the viscosity of the vacuum residuum must be increased or the temperature sensitivity improved, a limited air-blowing
process may be used. The process involves continuous pumping of an asphalt resid (flux) through an oxidation tower while air
is passed through the flux stream (Corbett 1984).

CHEMICAL ANALYSIS OF ASPHALT

Efforts to use chemical analysis to predict the performance of asphalt as a construction material have been thwarted by the
complexity of the material. The atomic composition of asphalt is predominately carbon (83%87%) and hydrogen (10%11%)
(Peterson 1984). Asphalt is a mixture of a wide variety of chemical compounds that include aliphatic hydrocarbons and
aromatic ring (polypericyclic) systems. The hydrocarbon component consists of straight and branched chains referred to
as aliphatic or paraffinic molecules, simple or complex hydrogen saturated rings called naphthenic molecules, and simple
or complex unsaturated ring structures classified as aromatic molecules. The aromatic molecules could include benzene or
naphthalene, but more commonly consist of more complex multiring structures (polycyclic molecules). The smallest size of
the hydrocarbons is defined by the processing conditions of the crude oils (cut point in the vacuum distillation tower) and the
largest size is defined by the crude oil source. The hydrocarbons are considered to be nonpolar components.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 9

Polycyclic molecules may contain one or more heteroatoms; that is, sulfur (1%5%), nitrogen (0.3%1.1%), and oxygen
(0.2%0.8%). The heteroatoms are attached to the asphalt molecules in various forms, but each of the functionalized mol-
ecules is considered a polar component of an asphalt cement. These polar components vary depending on the source of the
asphalt. The nonpolar components of the asphalt cements act as solvents or dispersing agents of polar components and control
the compatibility of the asphalt cement.

Depending on the crude source, trace metals such as vanadium (41,400 ppm), nickel (0.41110 ppm), and iron (6250
ppm) occur in porphyrins present in the polar components. The molecular interactions among these polar molecules strongly
influence the physical properties and the performance of the asphalt.

SYNTHESIS APPROACH

The information in this synthesis was gathered through a thorough review of the available U.S. and international literature. A
comprehensive literature search was performed using the TRID database, SciFinder, Bing, and Google. The TRID database
and SciFinder complement each other in that TRID covers engineering subjects and SciFinder provides access to the chemical
literature. Journals not covered by either of these search engines can be accessed directly using Bing or Google. In addition, a
survey of U.S. departments of transportation (DOTs) and Canadian ministries of transportation was conducted to determine
the current status of asphalt chemical analysis. The U.S. state and District of Columbia response rate to the survey was 88.2%
(45 of 51). Only eight of the 14 Canadian provinces and territories responded, but their comments were enlightening.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

10

CHAPTER TWO

LITERATURE REVIEW: CHEMICAL AND PHYSICAL CHARACTERIZATION


OF ASPHALT BINDERS

ASPHALT PRODUCTION

According to the National Asphalt Pavement Association, of the 2.6 million miles of paved roads and highways in the United
States, approximately 93% are paved with petroleum-based asphalt. Asphalt binder supplies are shrinking, as only a limited
number of crude oils yield quality asphalt. The price of liquid asphalt has fluctuated dramatically within the past decade,
ranging from $200 per ton in 2005 to $560 per ton in 2012. This is mainly because of the increasing cost of crude oil and the
advent of coking technologies, which enabled refineries to decrease their production of asphalt (residua from crude oil refin-
ing processes) by converting it into synthetic fuel. This development has led to a shortage of high-quality asphalt binders and
an increase in their price independent of the crude oil price.

U.S. refiners are running at maximum capacity (>90%) and no new refineries are planned in the near term. Existing refin-
ery expansions must fill any gap but the expansion will focus on handling increased supplies of heavy crude oil and their con-
version capabilities to meet light product demand. Early refineries were simply distillation units that processed the material
boiling above 550C and left approximately 40% of heavy crudes as residue (residuum), which was primarily sold as asphalt.
The addition of various hydrocracking units allowed further processing of the residuum to produce more light products. These
additions reduced the typical asphalt yield to 12%15%. Currently, refineries are adding coker units that will convert the
residuum to fuel oil and coke, and no asphalt is produced as a by-product. A refiner can decide whether to produce asphalt
based on the relative price differential between light products and asphalt prices, the crack spread. As the price of gasoline
increases the resultant crack spread makes the production of asphalt uneconomical unless there is a corresponding increase in
asphalt prices. Thus, asphalt can no longer be considered a by-product, but now it is a product that can be produced based on
the market demand. The price of asphalt must keep pace with conversion feed values to encourage its production. In addition
to the absolute price of crude oil, the production of asphalt is influenced by the light/heavy crude price differential, the light
product crack spread, coking economics, and heavy crude availability. More heavy crude is becoming available and not all
refineries have installed coker units, so the asphalt supply is ensured at the right price.

A survey of asphalt usage prepared by the Asphalt Institute, based on responses from its 54 members, is summarized in
Table 2. The total amount of asphalt used in 2013 was 19.1 million tons, of which 16.3 million tons went to paving applications.
The categories of asphalt usages cited in the table are as follows:

Asphalt cement (AC): A solid or semisolid asphalt that has not been modified by the addition of a low or intermediate
boiling range solvent, emulsification, or the addition of inorganic fillers. It has sufficient quality and consistency for direct
use in the manufacture of bituminous pavements, roofing materials, or other industrial products. Outside the United States,
the product is often called bitumen.

Modified asphalt cement: An AC to which a performance-improving modifier and/or additive has been blended. Com-
monly used modifiers include elastomers and plastics (polymers). Other additives include fillers, extenders, fibers, oxidants,
antioxidants, hydrocarbons, polyphosphoric acid, and combinations. Normally, modifiers will improve the base ACs grade
by at least one level. Asphalts that are changed by the refining process, such as oxidation, are not considered modified for the
purposes of this survey.

Cutback asphalt (slow, medium, or rapid curing): Asphalt products produced by blending an asphalt with solvents such
as naphtha, kerosene, No. 2 fuel oil, diesel oil, or other volatile solvents. Upon exposure to the atmosphere, the volatile product
evaporates, leaving the asphalt. Also included under this heading are road oils, including residual asphalt oils, used as a dust
palliative or other surface treatments of pavements.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 11

TABLE 2
2013 U.S. ASPHALT USAGE IN SHORT TONS BY PADD DISTRICT
PADD Asphalt % of Modified % of Cutback Emulsified % of Total Nonpaving Total
Districts Cement Paving Asphalt Paving Asphalt Asphalt Paving Paving Asphalt Asphalt
Asphalts Cement Asphalts Asphalts Asphalt Usage
PADD I 4,541,850 88 377,833 7 5,085 222,109 4 5,146,877 897,320 6,044,197
East Coast
PADD II 4,415,858 72 872,041 14 93,860 753,877 12 6,135,636 694,320 6,829,956
Midwest
PADD III 1,599,298 69 355,128 15 63,906 306,437 13 2,324,769 863,677 3,188,446
Gulf Coast
PADD IV 239,969 44 153,026 28 6,402 140,210 26 539,607 5,578 545,185
Rocky
Mountain
PADD V 1,771,847 83 175,720 8 37,589 152,917 7 2,138,073 387,285 2,525,358
West Coast
Total U.S. 12,568,822 77 1,933,748 12 206,842 1,575,550 10 16,284,962 2,828,180 19,113,142
tonnage
Source: This work compiled from www.eia.gov data.

Emulsified asphalt: Emulsified asphalts consist of asphalt and water processed with emulsifying agents to produce a
stable suspension of minute globules of asphalt in water; or alternatively, a suspension of minute globules of water in liquid
asphalt. The listed figures are an estimate of the asphalt content and should not include water and other liquid additives that
typically range between 30% and 45% by weight of mixture. Emulsified asphalts may be either anionic or cationic emulsions.

Asphalt materials for nonpaving applications, primarily asphalt cements, emulsions, and fluxes used in the manufacture
of roofing asphalts.

Changes in geographic usage can be assessed when the usage is reported by Petroleum Administration for Defense Districts
(PADDs) as defined by the U.S. Department of Energy. The PADD concept is used today for data collection purposes and to
aid in understanding of supply and demand of domestic petroleum products. Figure 1 shows the current PADD delineation.

FIGURE 1 Petroleum Administration for Defense Districts (PADDs) (Source: U.S. Energy Information Administration:
http://www.eia.gov/petroleum/supply/monthly/pdf/append.pdf).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

12

It is well understood that crude oil sourcing for U.S. refineries varies over time, but in general it can be assumed that the
sources for a given PADD are consistent. PADD 1 refineries process crude oil shipped from all over the world, including Saudi
Arabia and Venezuela. PADD 2 and PADD 4 depend primarily on crude oil produced and moved by pipeline from Canada
supplemented by crude from PADD 3 as well as production from Rocky Mountain state sources. PADD 3 is the largest refin-
ing region and obtains crude oil from the Gulf Coast outer continental shelf, Mexico, Venezuela, and the rest of the world.
Currently, permitting issues stalling construction of the Keystone pipeline are forcing rail transport of Canadian syncrude
(from oil sands) to PADD 3 refineries. PADD 5 obtains crude oil primarily from Alaska (by tanker) and California (Kern
River Valley), and through imports.

Data compiled in 2013 are summarized in Table 2. Asphalt cement constitutes more than 80% of the paving asphalt
installed in the East and West Coasts districts. Modified asphalts and emulsified asphalts are used more extensively in the
Midwest, Gulf Coast, and Rocky Mountain districts. Overall, 77% of the total tonnage applied is asphalt cement, 12% is modi-
fied asphalts, and 10% is emulsified asphalts.

Data from the U.S. Energy Information Administration show the trends in net production of asphalt and road oil by refiner-
ies and blenders. The data are reported in annual thousand barrels of asphalt, so it is difficult to compare production directly
with the asphalt usage reported in tons above. However, the trends in production can be observed in Figure 2. In the past 6
years, maximum production occurred in 2010. Asphalt production decreased steadily through 2014, but the trend was reversed
in 2015. The largest increase in production occurred in the Louisiana Gulf Coast region of PADD 3.

FIGURE 2 Refinery and blender net production of asphalt and road oil (Source: U.S. Energy Information
Administration. This work was compiled from data at http://www.eia.gov/petroleum/supply).

According to a study published by the Freedonia Group, demand for asphalt in paving applications is forecast to advance
3.1% annually through 2019, to 19.6 million tons, based on improving economic conditions and a pressing need to repair and
expand the nations infrastructure. Among asphalt products, asphalt emulsions will see the fastest growth, boosted by interest
in on-site recycling since asphalt emulsions can be used to recycle older pavements while minimizing asphalt consumption.
Asphalt cement will remain the leading paving material used in the United States owing to the prevalence of hot-mix and
warm-mix asphalts in paving jobs. These products are favored because of their moderate cost and solid performance proper-
ties, such as durability and good drainage (Freedonia Group 2015).

ASPHALT COMPOSITION

The terms asphalt and bitumen are often used interchangeably to mean both natural and manufactured forms of the substance.
Asphalt (or asphalt cement) is the carefully refined product derived from selected crude oils. Outside the United States, the prod-
uct is often called bitumen. Asphalt is no longer just the residua from crude oil refining. It is now known as an asphalt binder and
is part of an engineered system. For asphalt that is less suitable for paving applications, modifiers, such as polymers, adhesion

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 13

promoters, and crosslinking agents, work together to improve the performance of the asphalt binder, which ultimately extends
the life of the pavement. Asphalt is no longer specified by its purely physical properties, but on how it is expected to perform.
Asphalt is considered an economical, higher-performing construction material.

It has sometimes been assumed that once the chemistry of bitumen is known we will be able to predict its performance as a
construction material, as well as specify the properties of good bitumen. However, knowledge of the chemical composition
is only a limited help for understanding bitumen, and advanced modern analytical techniques always provide average results
that are not easily translated to physical properties or performance properties.

All molecules in asphalt are hydrocarbons with small amounts of sulfur, nitrogen, and oxygen and traces of metals such as
vanadium and nickel. The hydrocarbons consist of polyaromatic structures containing different numbers of fused rings, satu-
rated polycyclic structures also with different numbers of rings, and combinations of these. All these core structures contain
saturated hydrocarbon side chains of different chain lengths and different substitution patterns. These side chains can be lost
during the refining process, so the properties of the asphalt produced today are different from those of the asphalt that was
simply the residue of a crude oil distillation as in the past. Asphalt consists of millions of different molecules, almost none
of them in sufficiently large quantities, which makes them impractical to isolate and characterize. So even if the structure is
known in principle, the exact structure is unknown.

Based on average analytical data, one can suggest an average bitumen molecule as shown in Figure 3. The polycyclic ring
structure is depicted as being composed of both aromatic and saturated rings, and the colored atoms depict heteroatoms. The
molecular size is large enough to give a boiling point above the cut point for the heaviest distillate. Sulfur is largely present
in thiophene structures, which is the most common structure for sulfur-based heterocyclic molecules. The average bitumen
molecule shown contains aromatic and saturate components in approximately the average amount known for bitumen. It is
possible to estimate that the smallest size of the molecules is about 20 carbons, and the number of carbons goes up to the larg-
est size in the residue.

FIGURE 3 Typical bitumen molecule (Source: Redelius and Soenen 2015).

Traditional analytical techniques, such as ultraviolet spectroscopy (UV), Fourier transform infrared spectroscopy
(FTIR), nuclear magnetic resonance (NMR), or mass spectroscopy (MS), have provided substantial information about the
average chemical composition of asphalt. Most asphalts give more or less identical spectra and there are no known general
correlations between physical properties and any particular functional group as identified by the techniques listed earlier.
The only technique that is frequently used is FTIR, which permits identification of carbonyls and sulfoxides formed as
a result of aging. Modern analytical techniques allow a more complete definition of the typical asphalt molecules. These
techniques include thermogravimetric analysis (TGA), differential thermal analysis (DTA), gel permeation chromatog-
raphy (GPC), scanning electron microscopy, and atomic force microscopy (AFM). These techniques will be discussed in
detail later in this synthesis. The modern techniques provide insight into the molecular interactions that govern the physi-
cal properties of the asphalt matrix. Changes in the physical properties produced by specific additives can be ascertained
to estimate their effectiveness.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

14

FTIR is one of the more important methods for fingerprinting asphalt materials and quantifying the distribution of asphalt
components. By determining the various chemical functional groups in the binder, an understanding of its origin and history
can be obtained. The FTIR method is an efficient technique for identifying polymer additives in a binder. Determination
of polymer content is essential for quality control and quality assurance during the processing and application of polymer-
modified asphalts (PMAs).

TGA and differential scanning calorimetry (DSC) are used to characterize petroleum bitumens and their chromatographic
fractions, including the glass transition temperature and the percentage of crystalline phases. The advent of modulated DSC
provides new insight on asphalt microstructure. The development of bitumen microstructure and the calculations of the
entropy and enthalpy of transitions suggest that bitumen is a structured amorphous phase containing a small crystalline phase.

Proton nuclear magnetic resonance (1H NMR) spectroscopy has emerged as a very powerful and versatile tool for bitumen
characterization. Using 1H and 13C NMR can yield information on average structural parameters of asphalt and asphaltenes,
such as percentages of aromatic carbons, aliphatic carbons, bridged carbons, methyl carbons, ring carbons, naphthenic car-
bons, paraffinic chain lengths, and other parameters. The specific environment of the different types of hydrogens and carbons
can be defined (Betancourt Cardozo et al. 2016).

GPC is a separation method that takes advantage of differences in the size of molecular structures of asphalt components.
The procedure separates these components based on their molecular size (hydrodynamic volume). The polymer and asphalt
components of polymer-modified asphalt cements can be separated completely. GPCs ability to separate mixtures by molecu-
lar size rather than by some complex property such as solubility or absorptivity is one of the great advantages of the technique.
This feature has made GPC a useful alternate technique for fractionating complicated mixtures such as crude oil residua,
asphalts, and asphaltenes and a vital contributor to understanding asphalt mixes (Altgelt 1965; Dickie and Yen 1967; Snyder
1969; Yapp et al. 1991). The chromatograms are used to predict aging viscosity, aging index, viscosity number, and penetra-
tion. GPC is an excellent tool for measuring polymer content in modified binders on a routine basis. When NMR and GPC
information is combined, possible structures for asphalt and mechanisms of aging can be suggested.

One school of thought is that asphalt binder may be considered as a colloidal or micellar system. The hydrocarbon insoluble
components, asphaltenes and resins, are dispersed in a hydrocarbon blend. Many different types of molecular interactions
contribute to the stability of the dispersion. Because asphalt consists of nonpolar hydrocarbons, the dominating interactions
are London dispersive interactions resulting from temporary dipoles caused by variations in the electrons surrounding the
molecules. Also contributing are polar interactions and hydrogen bonding interactions that are due to the content of the more
electronegative elements nitrogen and oxygen. There are also pipi interactions, which for nonpolar molecules involve aro-
matic components of the molecules. Quantification of these interactions is difficult since the molecules exact structure is not
known. Each type of interaction occurs in concert with the other types of interactions described previously. The presence of a
solvent may exert a strong effect on the relative strength of the interactions (Redelius and Soenen 2015). The overall behavior
of asphalt cement is controlled by the compatibility and the relationships of the different components in this macroscopically
homogeneous mixture rather than by the quantitative amount of any single component (Petersen 1984).

ASPHALTENE STRUCTURES

Asphaltenes are the insoluble fraction precipitated from a toluene solution of asphalt binder by a nonpolar solvent such as
pentane or heptane. The precipitating solvent affects the quantity of asphaltenes precipitated. The nature of the asphaltene
molecules in the precipitate depends on the precipitating solvent; greater amounts of a precipitant are produced by pentane.
In general, the asphaltenes are defined as the mixture of materials precipitated by heptane. The asphaltene molecules are very
complex and exhibit a very high tendency to associate into molecular clusters. The amount and characteristics of asphaltenes
vary from one asphalt source to another. They play a significant role as viscosity builders in the rheology of asphalt binders.

The composition of an asphaltene fraction has been the subject of debate for the past 30 years (Petersen 1984; Goodrich
et al. 1986). The complexity of the fraction and the tendency for the asphaltene molecules to associate has stimulated an
extensive debate on their structures. Virtually all asphaltene chemical properties, except elemental composition, had been the
subject of debate; their molecular weight (MW) had been estimated at values spanning six orders of magnitude (Mullins et
al. 2008; Yarranton et al. 2013). MW determination has been, and still is, a challenging problem in asphaltene chemistry. The
source of complications is attributable to three basic properties of asphaltene, namely, compositional variance, size polydis-
persity, and, most important, the high propensity of the covalent asphaltene molecules to form molecular aggregates (Strausz

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 15

et al. 2008). A whole gamut of methods has been applied for the MW determination of asphaltene over the decades, including
chemical, physical, absolute, relative, equilibrium, and nonequilibrium methods, but a consensus of the appropriate MW value
has not been reached. Nevertheless, an understanding of asphaltenes is essential because they have a significant impact on
many physical and chemical properties of crude oils and asphalts.

Many diverse studies have converged on a simple, hierarchical model of the molecular and colloidal structure of asphaltenes:
the modified Yen model (also known as the YenMullins model). The structural hierarchy in Figure 4 shows the extent of
aggregation of the polycyclic aromatic hydrocarbon (PAH), which is the asphaltene monomer. The predominant asphaltene
molecular architecture contains a single, moderately large PAH with peripheral alkanes (~1.5 nm). Asphaltene molecules
form nanoaggregates with aggregation numbers of approximately six, with a single disordered PAH stack (~2 nm). These
asphaltene nanoaggregates can form clusters with aggregation numbers of approximately eight (5 nm).

FIGURE 4 The modified Yen model (or the YenMullins model): polycyclic aromatic hydrocarbon monomer (left),
nanoaggregate (center), and clusters (right) (Source: Mullins 2011).

It was unknown whether asphaltene molecules contain predominantly one PAH (the island architecture) or many cross-
linked PAHs (the archipelago architecture) (Strausz et al. 1992). Molecular diffusion measurements, especially from time-
resolved fluorescence depolarization, helped to resolve these molecular properties of asphaltenes and provide evidence for the
island molecular architecture.

Petroleum asphaltenes consist of approximately 40% to 45% aromatic carbon; the remainder is aliphatic, as shown by 13C
NMR. Alkane chains are an average of four to five carbons long, as has been determined by integrated infrared (IR) spectroscopy
and NMR studies. Observations of active hydrogen in asphaltene (Gould and Wiehe 2007) show that naphthenic rings are fused to
aromatic rings. The aromatic ring system is also where almost all the nitrogen (~1% by mass) is located, as shown by X-ray absorp-
tion near-edge structure; both basic pyridinic nitrogen and acidic pyrrolic nitrogen are present (Mitra-Kirtley et al. 1997). Sulfur
X-ray absorption near-edge structure shows that most of the several percent sulfur consists of thiophene- and sulfide-type groups
and generally a small fraction of sulfoxide (George and Gorbaty 1989). The proposed PAH asphaltene monomers shown in Figure
5 have an MW of 750 Da, and a single fused aromatic ring system per molecule (island molecular architecture) (Mullins 2011).

FIGURE 5 Typical asphaltene molecular architecture (Source:


Mullins 2011).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

16

A variety of experimental techniques were applied to a single source asphaltene sample at the same experimental condi-
tions to reveal the possible size distributions of asphaltene monomers and aggregates (Yarranton et al. 2013). The asphaltene
sample was divided into solubility cuts by selective precipitation in a solution of toluene diluted with various concentrations
of heptane. Asphaltene self-association was assessed through a combination of density, vapor pressure osmometry (VPO),
elemental analysis, Fourier transform-ion cyclotron resonance mass spectrometry, and time-resolved fluorescence emission
spectra measurements performed on each cut. The physical dimensions of the asphaltenes were assessed using small angle
X-ray scattering (SAXS), dynamic light scattering (DLS), membrane diffusion, Rayleigh scattering, and nanofiltration mea-
surements. Molecular and nanoaggregate dimensions were also investigated through a combination of interfacial tension,
interfacial adsorption, and surface force measurements.

All of the measurements indicated that approximately 90 weight percent of the asphaltenes are self-associated. Ultrahigh
resolution spectrometry suggests that the nonassociated asphaltenes are smaller and more aromatic than bulk asphaltenes,
indicating that the associating species are larger and less aromatic. On the basis of VPO, the average monomer MW was
approximately 850 g/mole, while the MW of the nanoaggregates spanned a range of at least 30,000 g/mole with an average
on the order of 10,000 to 20,000 g/mol. SAXS and DLS gave MWs 10 times larger. However, these techniques are known to
be unreliable for asphaltene MW measurement (Mullins et al. 2008). The discrepancies demonstrate that the final judgment
on asphaltene structure remains in doubt.

The physical dimensions of the nanoaggregates were less than 20 nm based on nanofiltration and with average diameters of
5 to 9 nm based on diffusion and Rayleigh scattering. SAXS and DLS showed average diameters of 14 nm and indicated that
the nanoaggregates had loose structures. Film studies were consistent with the lower MWs and dimensions. The asphaltene
monolayers swell by a factor of 4 in the presence of a solvent. The most consistent interpretation of the data is that asphaltenes
form a highly polydisperse distribution of loosely structured (porous or low fractal dimension) nanoaggregates (Yarranton
et al. 2013).

ASPHALT MORPHOLOGY

Lesueur has published a detailed review of the current asphalt structure concepts that defines links between chemistry,
structure, and mechanical properties in a framework of an updated colloidal picture of asphalt (Lesueur 2009). Asphalt is
considered a colloidal dispersion of asphaltene micelles in the maltenes. The resinsthat is, the polar components of the
malteneswere thought to stabilize the asphaltene micelles. Extensive analytical procedures support this model. A more pre-
cise description of this process is starting to arise. SAXS and small angle neutron scattering (SANS) confirm that asphaltenes
form micelles in organic solvents in asphalt (Yen 1992). The diffusion pattern observed in SAXS or SANS experiments disap-
peared once the asphaltenes were removed from the asphalt.

The colloidal model is also consistent with results obtained by thermal analysis. Asphalt undergoes a glass transition at a
temperature very close to that of its aromatics moieties. This strongly suggests that the asphaltenes exist as dispersed solid
particles and do not directly participate in the glass transition. The asphaltenes extend the span of the glass transition range,
implying that at least some molecules of this family might contribute to the glass transition when mixed with maltenes (Claudy
et al. 1992a).

Recently researchers combined atomic resolution imaging using atomic force microscopy and molecular orbital imaging
using scanning tunneling microscopy to study more than 100 asphaltene molecules derived from coal and petroleum (Schuler
et al. 2015). The complexity and range of asphaltene polycyclic aromatic hydrocarbons are established in detail. Aromatic
hydrocarbon (PAH) moieties of asphaltene comprise the primary site of intermolecular interaction in the asphalt colloids.
The petroleum asphalt shown in Figure 6 is from a single crude oil. Sample PA1 has a large PAH core with two side chains
attached. PA2 shows a PAH that has a side chain of about 20 in length, The PAH in PA3 contains three five-membered rings,
which are identified in the image. The asphaltene molecules consist of a central aromatic core with peripheral alkane chains.
In some cases, this central core is divided into several distinct PAHs connected by a single bond, which proves the presence of
archipelago-type molecules. Nevertheless, a single aromatic core with peripheral alkanes is the dominant asphaltene molecu-
lar architecture, proving the main aspects proposed by the YenMullins model (Mullins 2011).

All this evidence makes it difficult to deny the colloidal nature of bitumen. The most convincing evidence is that which
comes from diffusion experiments. Scattering techniques highlighted an elementary structure consisting of diffusive
particles with a of radius 28 nm. Such particle size is reminiscent of the crystal size of pure asphaltenes observed by

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 17

AFM in a pure petroleum fraction and concurs with the earlier description of asphaltene micelles as being made of a few
individual molecules.

FIGURE 6 Atomic force microscopy (AFM) Laplace filtered images of petroleum


asphaltenes (Source: Schuler et al. 2015).

Identifying molecular structures provides a foundation for understanding all aspects of petroleum science from colloi-
dal structure and interfacial interactions to petroleum thermodynamics, enabling a first-principles approach to optimizing
resource utilization. Particularly, the findings contribute to a long-standing debate about asphaltene molecular. architecture.
AFM can be a powerful complementary tool to rheology and spectroscopy for characterizing asphalts. It makes it possible to
compare the mechanical properties of pure and modified bitumen. The contributions of the AFM studies to the understanding
of asphalt morphologies will be detailed in the section on instrumental analysis.

When polymers are added to binders, the equilibrium situation is a macroscopic phase separation of the two phases
(Lesueur 2009). The polymer is swollen by the light aromatic components from the parent bitumen, and the polymer-rich
phase (PRP) occupies between 4 to 10 times the volume of added polymer, especially for styrene-butadiene diblock copolymer
(SB) and ethylene-vinyl acetate copolymer (EVA). The continuous phase is an asphaltene-rich phase. Polymer/binder compat-
ibility is indeed a dynamic concept; compatible systems are those with a slow phase separation (creaming rate). The creaming
rate is a function of the relative densities of the microphases; the larger the density difference and the larger the phase sizes, the
faster the creaming rate. To prevent phase separation, a few stabilization mechanisms are developed. Adding a crosslinking
agent to the PMA under agitation allows the PMA to slightly crosslink, preventing PRP droplets from coalescing. The freez-
ing of the equilibrium droplet size under agitation yields a very low particle size and favors stability. High crosslink density
is desired because it would result in less swelling. Because each asphalt has its own particular chemical composition, ways to
predict whether a particular polymer will be compatible with a given asphalt are not well defined, so the formulator usually
relies on laboratory experiments rather than on theoretical predictions. In all cases, and even if the polymer has a potentially
compatible chemistry, the formulation of PMAs requires knowledge of the chemical properties of the initial bitumen. Critical
binder properties can be identified; that is, high asphaltene content decreases polymer/asphalt compatibility and the aromatic-
ity of the maltenes needs to fall between certain values to reach a good level of compatibility.

The importance of polyaromaticity on the elastic properties of bituminous binders was studied (Soenen and Redelius
2014). The size of the polyaromatic structures appears to play a crucial role. Observations indicate that larger conjugated
aromatic structures provide stronger pipi interactions between asphaltene stacks, which are known to be the main mecha-
nism responsible for the formation of clusters at the nanoscale. The larger extent of these interactions relates to the elastic
behavior at longer loading times or up to higher temperatures, while smaller aromatic structures determine more the elastic

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

18

behavior at short loading times or at low temperatures. In addition to aromatic interactions, natural wax can, upon crystalliza-
tion, also induce increased elastic effects, especially at low frequencies. A large variety of binders was investigated: bitumen
from straight distillation, visbreaking, solvent deasphalting, and oxidation were included in the sample set. Average levels of
aromaticity were determined by FTIR and by refractive index measurements. Chromatography combined with UV-visible
absorption spectroscopy was used as an indicator of the average size of the aromatic structures. Rheological properties were
determined using a dynamic shear rheometer, in a temperature range from 0 up to 90C. Good relationships were observed
between the phase angle measurements, at specific test conditions of frequency and temperature, and ultraviolet-visible
absorption levels at specific wavelengths (Soenen and Redelius 2014).

Changes in the stacking behavior of asphaltene units are investigated in the presence of hexadecanamide, a representative
amide-type additive. Molecular dynamics simulations and experiments using high-resolution transmission electron micros-
copy and X-ray powder diffraction support results obtained from rigorous quantum mechanical calculations through a high
quantum level of density functional-dispersion correction approach. Based on this multiscale bottom-up study, interaction of
the amide-type binder with asphaltenes disturbs the uniformity of the density throughout the aromatic region and creates
some polarization in this region. This alteration of the system over the aromatic zone disturbs the eventual pipi interactions
between asphaltene stacks. Disturbing these interactions alters the stacking distance, the corresponding binding energy, and
ultimately the extent of clustering of asphaltene units. Any change in the clustering of asphaltene affects the rheology and
morphological properties of asphalt, which in turn alters the asphalts performance, including but not limited to its resistance
to fatigue and low-temperature cracking (Mousavi et al. 2016).

A number of factors and variables influence the chemical composition and ultimately the engineering properties of a binder
used in pavement applications. It is important that the users most closely working with the binders have a strong working
knowledge of the relationship between binder chemistry and engineering properties. In addition, the relationship between
standard process control actions and final chemical composition needs to be understood so that correlation to physical per-
formance can be made.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 19

CHAPTER THREE

TECHNIQUES FOR CHEMICAL CHARACTERIZATION OF ASPHALT

ASPHALT FRACTIONATION

Historically, the study of asphalt chemical composition has been facilitated by the separation of asphalt into component
fractions based on the polarity or adsorption characteristics, or both, of the molecular components present. The component
fractions, sometimes called generic fractions, are useful in classifying and characterizing asphalts and in providing simpli-
fied mixtures for further study. They are still complex mixtures whose composition is a function of asphalt source. Many
techniques using different properties of the molecules for separations have been used. The fractions are generally very hetero-
geneous and are defined only by the method of separation. A common misunderstanding is to consider the fractions as the
components in bitumen and claim that bitumen consists of a mixture of three, four, or five types of compounds instead of a
continuum of molecules. The continuum consists of relatively large hydrocarbons with different sizes, polarity, and aromatic-
ity. The component fractions are, however, sufficiently unique to identify their particular contribution to the complex flow
properties of asphalt. A proper balance of component types is necessary for a durable asphalt.

Many techniques, using different properties of the molecules for separations, have been used as shown in Table 3. These
fractions have then been correlated to physical properties, sometimes rather successfully but more commonly without find-
ing any general correlation. Historically, the two most common methods to fractionate asphalt are the chemical precipitation
method (Rostler and White 1959) and the selective adsorptiondesorption (chromatographic) method (Corbett 1970). In both
methods, the asphaltenes are identified as the fraction insoluble in pentane or heptane. The pentane or heptane soluble frac-
tion is classified as maltenes, which can be further separated into mixtures with differing solubility properties. The chemical
precipitation method uses various sulfuric acid concentrations to identify and characterize four classes of molecular mixes
comprising the maltenes: nitrogen bases, first acidiffines, second acidiffins, and parafins. The fractional components of four
representative asphalts have been summarized by White et al. (1970). In the absence of asphaltene, the fractional maltene

TABLE 3
METHODS FOR FRACTIONATING ASPHALTS AND CRUDE OILS
Principle Products Precipitant References
Solvent extraction Parafinnics, cyclics, asphaltics n-butanol, acetone Traxler and Schweyer (1953)
Acid precipitation Paraffins, nitrogenbases, acidaffins no. 1, n-pentane Rostler and White (1970); White at al. (1970)
acidaffins no. 2, asphaltenes
Adsorption on alumina Saturates, naphtene aromatics, n-heptane Corbett and Swarbrick (1966)
polar aromatics, asphaltenes
Adsorption on clay Saturates, aromatics, polars, asphaltenes n-pentane Corbett and Swarbrick (1966)
Gel permeation chromatography Molecules of different sizes N/A Davison et al. (1995)
Gas chromatography Volatile hydrocarbons N/A Tang and Isacsson (2005);
Fernandes et al. (2009)
High-pressure liquid Molecules of different polarities N/A Such et al. (1979)
chromatography
Thin-layer chromatography Saturates, aromatics, resins A, resins B N/A Bharati et al. (1994)
(Iatroscan) (SARA)
Ion exchange chromatography Neutrals, bases, acids N/A Branthaver et al. (1992)
Flash chromatography Saturates, aromatics, resins A, resins B n-heptane Raki et al. (2000)
Asphaltene determinator Aromatics, saturates, resins, asphaltenes N/A Boysen and Schabron (2013)
Automatic saturates, aromatics, Aromatics, saturates, resins, asphaltenes N/A Schabron and Rovani (2013)
resins and asphaltene-determinator
Source: Adapted from Masson et al. (2001).
N/A = not available.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

20

components from different asphalts are compatible. Variations in the asphalts are the result of changes in the asphaltene com-
ponents. The research demonstrates that the durability of the asphalts depends on the proportions of the components and that
the asphalt quality could be improved by blending to change the proportions of maltene components to reach a desirable com-
position ratio. Rostler and White (1970) demonstrated that the primary differences in asphaltene fractions was their molecular
weights. Asphaltene molecular weight variations among the asphalts studied was not very significant (2,0006,000), but the
addition of compatible rubbers enhanced the properties of the corresponding HMAs.

TABLE 4
FRACTIONS OBTAINED USING CORBETT ANALYSIS
Rings/Mole
Group Weight Average Fraction Naphthene Aromatic Description
Percent Range Molecular Weight Aromatic
Saturates 515 650 0 3 2.6 Pure paraffins + pure naphthenes
mixed paraffin naphthenes
Naphthene aromatics 3045 725 0.25 3.5 7.4 Mixed paraffin naphthene aromatics +
sulfur-containing compounds
Polar aromatics 3045 1,150 0.42 3.6 Not Mixed paraffin naphthene aromatics in
determined multiring structures and sulfur, oxygen.
nitrogen-containing compounds
Asphaltenes 520 3,500 0.5 Not Not Mixed paraffin naphthenc aromatics in
determined determined polycyclic structures + sulfur, oxygen,
nitrogen-containing compounds
Source: Corbett (1969).

Corbett Fractionation

This Rostler method has been essentially superseded by the chromatographic method as developed by Corbett (Corbett and
Swarbrick 1966; ASTM 2008). Corbett used a densometric procedure coupled with molecular weight determination by VPO
at 37oC to determine the structure of his fractions (Corbett 1964). Asphaltenes could not be characterized completely because
of the difficulties in molecular weight determination as a result of asphaltene molecular association. However, the asphaltenes
precipitated by heptane can be further fractionated by the relative solubility of the asphaltenes in toluene and carbon disulfide
(Figure 7) (Speight 2004). After precipitating the asphaltene fraction by addition of n-heptane, the maltenes in the heptane
solution are coated unto an alumina column. Sequential elution of the column with heptane and benzene yields the least polar
saturated hydrocarbons followed by the naphthenes and aromatic hydrocarbons. Increasing the polarity of the eluting solvents by
adding methanol to the benzene and changing to a trichloroethylene eluent yields the polar aromatics and resins. Alternatively
the deasphaltened oil can be fractionated by sequential extraction with heptane, toluene, and pyridine. The fractions are denoted
as saturates, aromatics, resins, and asphaltenes (SARA). The properties of the fractions are compiled in Table 4. Bulk flash chro-
matography of the maltenes can be used to obtain gram quantities of saturates, aromatics, and resins fractions (Masson 2000).

FIGURE 7 Separation of bitumen into its various fractions, highlighting the SARA (saturates, aromatics,
resins, and asphaltenes) fractions. (Source: Speight 2004.)

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 21

Table 5 shows additional structural data estimated for the fractions (Corbett 1970). These results are all dependent on the
composition of the crude oil source, particularly heteroatom content and metals. Both nickel and vanadium are found primarily
in the heptane-precipitated asphaltenes and are evenly distributed in the resins and asphaltenes. They appear to be interchange-
able in structurein fractions of a given asphalt the ratio of vanadium to nickel is constant over wide ranges of composition.

TABLE 5
ELEMENTAL CHARACTERIZATION OF CORBETT FRACTIONS
Average Number of Atoms per Molecule in:
Element Saturates Naphthene Aromatics Polar Aromatics Asphaltenes
Carbon
Paraffin chain 31 21 24 85
Naphthene ring 14 17 18 29
Aromatic ring 0 13 25 115
Hydrogen 85 94 105 350
Sulfur 0 0.5 1 4
Nitrogen 0 0 1 3
Oxygen 0 0 1 2.5
Avg. molecular weight 625 730 970 3,400
Source: Corbett (1969, 1970).

Heteroatoms are important because of their inordinate contribution to resin properties. The presence of heteroatoms
enhances the activity of asphalt molecules toward oxidation. Large increases in asphalt hardening occur with the uptake of
only 1 weight percent oxygen. Petersen has carried out extensive work on functional group analysis. A typical analysis is
shown in Table 6 (Petersen 1986). When asphalt oxidizes, the principal increases in oxygenated species are in ketones and
sulfoxides. Carboxylic acids and anhydrides tend to concentrate at the aggregate surface in asphalt concrete and may produce
sensitivity to water damage.

TABLE 6
DISTRIBUTION OF FUNCTIONAL GROUPS IN FRACTIONS FROM CORBETT SEPARATION
Concentration in Fractiona (mole/liter)
Group Whole Asphalt Saturates Naphthene Aromatics Polar Aromatics Asphaltenes
Ketone 0 0 0 0.11 Trace
0.027 0 0 0 0.034
Anhydrides 0 0 0 Trace Trace
2-Quinoline types 0.021 0 0 0.023 0.046
Sulfoxides 0.019 0 0 0.12 0.09
Pyrrolics 0.17 0 0 0.21 0.23
Phenolics 0.035 0 0 0.055 0.075
aYieldof fractions based on whole asphalt were saturates, 9.9%; naphthene aromatics, 25.3%; polar aromatics, 38.1%; asphaltenes, 21.6%; loss on the column
(which should be added to polar aromatics) 5.1%.
Source: Petersen (1986).

Studies have shown that increases in asphalt viscosity with oxidation can be correlated with increases in carbonyl forma-
tion, which has been shown to be proportional to oxygen uptake (Liu et al. 1998). Almost certainly this hardening results
from hydrogen bonding between heteroatom groups in asphaltene molecules and also between polar aromatics, which then
may become asphaltenes (Barbour and Petersen 1974; Herrington and Wu 1999). This association strongly impacts attempts
to measure unimolecular size by GPC or colligative properties.

Iatroscan Analysis

The Corbett analysis became more routine when a procedure was developed that allowed rapid sample analysis. Thin
layer chromatography (TLC) on a thin layer of silica powder that is fused onto a quartz rod (a Chromarod) coupled with

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

22

a flame ionization detector has been incorporated into a commercial instrument, an Iatroscan. The instrument requires
milligram quantities of a sample, and the asphalt is separated into four fractions with better resolution and more rapidly
than the column chromatography procedure. The four fractions are pyrolyzed on the developed Chromarod and the
pyrolysates pass through the FID. The signal output of the FID detector is plotted versus the Chromarod peak position to
produce an Iatrogramme (Masson 2001). The signals are attributed to SARA; the procedure is called a SARA analysis
and it is used extensively in the petroleum industry to analyze crude oils as well as HMAs (Bharati et al. 1994). Cor-
relation of the asphaltene fraction with alternate separation methods is improved by precipitating the asphaltenes with
heptane and only depositing the maltenes on the Chromarod. It is important to realize that the SARA analysis provides
general information about the composition, depending on the solubility, the adsorption, and the partition coefficient of
the compounds in each solvent, all of them affected by mutual interactions between the sample components. Therefore,
the SARA fractionation has limitations for providing an accurate description of the particular structural and chemical
composition of the sample.

Automated SAR-AD

A new on-column precipitation and redissolution separation technique uses a continuous flow system to precipitate and
redissolve various chemical species from oil strictly on their relative solubility (Boysen and Schabron 2013). This saturates,
aromatics, resins, and asphaltene-determinator method (SAR-AD) involves precipitation of asphaltene components from
residua on a ground PTFE or activated silica column using a heptane mobile phase. The precipitated material is redissolved
and eluted at 30C in three steps using solvents of increasing solubility parameter: cyclohexane, toluene, and methylene chlo-
ride/methanol (98:2, v/v). A series of automated switching valves is used to direct solvent flow into the various columns in
forward and reverse directions in a complex series of steps. The eluted fractions are detected using both an evaporative light
scattering detector (ELSD) and a 500 nm detector. An illustrative chromatogram is shown in Figure 8.

FIGURE 8 Automated SAR-AD separation profile for 2 mg Lloydminster vacuum


residuum (Source: Boysen and Schabron 2013).
Note: SAR-AD = saturates, aromatics, resins, and asphaltene-determinator method;
ELSD = evaporative light scattering detector.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 23

The SAR-AD results in Table 7 are different than those from the manual SARA separations performed during the original
SHRP program (Jones 1993) and other SARA separations performed on these binders using different methods. In particular,
the quantity of saturates fractions is greater in the automated SAR-AD method. Differing asphaltene separation protocols
and separation mediums between these two studies is likely responsible for these discrepancies. The many different SARA
methods utilize procedures with different solvents and filters for isolating asphaltenes and different solvents and sorbents for
the maltenes separations. Results from these different methods are not necessarily comparable. What is important is that the
new method can be applied to separate and quantify the asphaltene components. Further applications include the ability to
study how asphalt binders change with oxidation, how oils change with processing, and also the differences between samples
in a repeatable manner.

TABLE 7
AUTOMATED SAR-AD RESULTS FOR SHRP CORE ASPHALT BINDERS
Maltenes Asphaltenes
Sample Detector Saturates Aromatics Resins CyC6 Toluene CH2Cl2/MeOH Total ELSD
AAA-1 ELSD 19.0 19.9 47.4 5.2 8.3 0.2 13.6
500 nm 0.6 31.9 27.6 37.7 2.3
AAE-1 ELSD 16.9 10.4 50.4 4.7 17.4 0.2 22.3
500 nm 0.4 28.4 19.2 49.4 2.6
AAB-1 ELSD 18.0 15.5 54.6 3.3 8.5 0.1 11.9
500 nm 0.5 36.6 18.1 41.5 3.3
AAC-1 ELSD 29.2 13.1 51.6 2.2 3.8 0.1 6.1
500 nm 0.9 48.7 17.0 30.5 2.9
AAD-1 ELSD 11.8 15.5 54.7 5.8 12.2 0.1 18.1
500 nm 0.6 26.6 23.8 46.3 2.7
AAF-1 ELSD 17.4 13.2 62.1 2.5 4.7 0.0 7.2
500 nm 0.8 46.5 18.8 31.9 2.1
AAG-1 ELSD 18.0 12.4 67.3 0.5 1.8 0.0 2.3
500 nm 1.5 65.8 9.4 20.9 2.5
AAK-1 ELSD 11.3 18.6 54.9 4.3 10.9 0.1 15.3
500 nm 0.5 30.7 23.5 42.4 3.0
AAM-1 ELSD 18.3 13.7 65.9 0.8 1.2 0.1 2.1
500 nm 1.1 76.6 8.8 11.4 2.2
Note: SAR-AD = saturates, aromatics, and resins asphaltene-determinator method; ELSD = evaporative light scattering detector.
Source: Boysen and Schabron (2013).

The results presented in Table 7 show real differences between binders. They also show a change resulting from oxidation;
for example, binder AAE-1 is air-blown AAA-1. The ELSD toluene-soluble asphaltenes-to-aromatics ratio for the original
AAA-1 is 0.4, and this ratio increases to 1.7 with air-blowing oxidation treatment. These ratios are less than 1 for all of the
other original asphalt binders. The 500 nm aging index ratio is the ratio of the toluene soluble asphaltenes area to the area of
the resins at 500 nm. This ratio is 1.2 for the original AAA-1 binder, and it increases to 1.7 with the air-blowing treatment.
All of the other binders except AAD-1 have values less than 1.7. The automated SAR-AD separation is highly repeatable and
gives real content differences between asphalt binders and residua, which allows for an understanding of the relationship
between chemical composition and physical properties. The total pericondensed aromaticity, ratio of 500 nm toluene soluble
asphaltenes to resins, and ratio of ELSD toluene soluble asphaltenes to aromatics are being explored for possible use as aging
indices and for asphalt performance correlations (Boysen and Schabron 2013).

Chemical composition is important in determining the physical properties and performance characteristics of asphalts. The
interactions of polar or polarizable chemical functionality, either naturally present or formed on oxidative aging, play a major
role in determining asphalt viscosity and related complex flow properties. Two major factors affecting asphalt durability are
(1) the compatibility of the interacting components of asphalt and (2) the resistance to changes resulting from oxidative aging.
Both factors are a function of chemical composition, which can vary widely from one asphalt source to another because of
inherent differences in crude sources or from processing and blending.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

24

ANALYTICAL INSTRUMENTATION FOR ASPHALT ANALYSIS

Thermal Analysis

Thermogravimetric analysis and differential scanning calorimetry are used to characterize petroleum bitumens and their
chromatographic fractions. The influence of the different constituents on the thermal stability of bitumen was studied by
TGA. In this analysis, the change in mass of a material is measured as a function of temperature or time. TGA facilitates
acquisition of information on properties of a material and its composition. When a sample is heated it often loses mass. Loss
of mass may be caused by vaporization or chemical reactions that evolve volatile products from the sample. A decomposition
as a result of chemical reaction leads to changes in the mass of the sample in a stepwise manner as the onset temperature for
each phase of the decomposition is reached. The onset temperature at which decomposition occurs provides information on
the stability of the material in that atmosphere. For example, if a reactive gas atmosphere is used, reaction of the material with
the gas can result in mass change. Typically, this mass change is exhibited in the form of mass loss; however, in cases such as
oxidation there may be a gain in mass. Composition of a material can be determined by analyzing the temperatures and the
extent of the individual weight losses.

The derivative of the weight loss curve, DTA, facilitates identification of the thermal transitions. TGA can be used to mea-
sure the thermal stability of a polymer and the thermal degradation of polymer blends owing to the simplicity of the weight
loss method. The potential of TGA for quantitative analysis of vulcanizates based on binary elastomer blends of natural rub-
ber (NR) and styrene-butadiene rubber (SBR) has been previously reported (Lee et al. 2007). A typical thermogram from an
analysis of a natural rubber vulcanizate is shown in Figure 9.

FIGURE 9 Thermogravimetric analysis/differential thermal analysis thermogram of natural rubber


vulcanizate (Source: Baumgardner et al. 2014).

The sample is heated under a nitrogen atmosphere according to a set protocol of isothermal and temperature ramping
sequences. Weight loss steps on the TGA output appear as downward slopes of the solid thermogravimetric curve and peak in
the dashed DTG curve. The slope of the thermogravimetric curve corresponds to the rate of change of sample mass. In region
1, volatile compounds such as water, residual solvents, and oils are evolved from 0 to 25 minutes at relatively low temperatures
from 40C to 250C (104F to 482F). In region 2, pyrolytic decomposition occurs in an inert atmosphere (nitrogen) from 25
to 50 minutes at 250C to 550C (482F to 1022F) (Juma et al. 2006), allowing for analysis of the content (step height) and
material type of the rubber hydrocarbon component. In region 3, the carbon black is combusted upon switching to an oxidative
(air) atmosphere from 50 to 70 minutes at 550C to 750C (1022F to 1382F). Region 4 is residual ash remaining from the

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 25

entire TGA process. When materials of binary NR/SBR compounds were decomposed in nitrogen between 250C and 550C
(482F to 1022F), two distinct regions of decomposition were observed. Calibration of TGA data with known samples led to
a protocol that allowed the ratio of NR to SBR in crumb rubber samples to estimated (Baumgardner et al. 2014).

TGA measurements also provide a simple means to determine the thermal stability of bitumen (Jimenez-Mateos et al.
1996; Mothe et al. 2008). TGA is an excellent procedure for determining the volatile asphalt components and residual solvents
in binder extracts. Thermal analysis can be considered as one method for measuring variations in asphalt microstructures.

If the energy required to maintain a constant temperature between a reference sample and the analyte is plotted, one
obtains a differential scanning chromatogram. The calorimeter is sensitive to thermal transitions such as glass transitions and
melting points of crystalline fractions in asphalt binders, which impact their physical and rheological properties. Noel and
Corbett used differential scanning calorimetry to estimate both the crystalline fraction and glass transition temperature, Tg, of
binders and recognized that it is critical to establish a consistent thermal history prior to comparing different binder samples
(Noel and Corbett 1970; Claudy et al. 1992b).

Crystalline Components of Asphalt

The crystallization process of asphalt fractions depends on thermal history. Slow heating or annealing before analysis to
experimentally realize a near-equilibrium state is necessary to study the system with more thermodynamic rigor. Using an
annealing protocol, a number of low-temperature transitions characteristic of a given crude and/or refinery source can be
identified; thus, comparison of thermograms from an unknown source with reference thermograms from documented sources
allows the unknown source to be identified. The crystallization process of a selected AC-20 asphalt, ACB, was systematically
studied by doping with the following pure crystalline hydrocarbons: octadecene-1, eicosane, and octacosane. The impact of
doping on the crystalline fractions in asphalt varies from asphalt to asphalt. The crystalline components in asphalt exhibit
distinct endothermic patterns that depend on their chemical structure, interactions with the amorphous phase, and interac-
tions among themselves. The most significant endothermic effect is produced by cocrystallization of components with similar
crystalline chain lengths. For example, the crystalline components of asphalt ACB do not interact with octadecene-1, but do
cocrystallize with eicosane and octacosane (Daly et al. 1996).

Chambrion observed two glass transitions in bitumen after cooling at constant rate. The magnitude and temperature of these
transitions depended on the cooling rate. At low cooling rates (<1K/min), the glass transition at the higher temperature vanishes
and an endothermic peak is obtained. From these observations, a segregation mechanism is proposed to explain the behavior of
bitumen during cooling (Chambrion et al. 1996). Glass transition temperatures (Tg) of an asphalt and its fractions obtained by
preparative GPC decreased drastically from 17.5C to 63.36C as the apparent molecular weight of the fractions increased from
500 to 800 Daltons, then increased regularly to 53C as the molecular weight increased further to 3,000 Daltons. Composition
rather than molecular weights of the fractions is responsible for the control of the Tg values (Hon et al. 1978).

Bitumen was analyzed by modulated DSC (Masson et al. 2002). This method allows for the deconvolution of overlapping
reversing and nonreversing thermal events and it allows for the observation of transitions not visible on the thermal curve
obtained by standard DSC. The reversing thermal curve revealed two Tgs in bitumen, which had an 85/100 penetration grade
and respective saturates, aromatics, resins, and asphaltenes contents of 9%, 27%, 43%, and 20%, as measured with the Iat-
roscan by successive elution in heptane, toluene, and tetrahydrofuran (THF). One transition, at 20C, was assigned to the
maltenes, the other at 70C to the asphaltenes. The heat capacity of these transitions was found to depend on thermal history.

After cooling from the melt and annealing at 22C, bitumen microstructure was found to develop in four stages. Most rapid
is an ordering process that occurs when bitumen is quenched from the melt. It is postulated that this first stage arises from
the partial ordering of simple aromatic structures into micro- and nano-phases; a second stage when low-MW saturated seg-
ments crystallize. In the third stage, high-MW saturated segments crystallize. In the fourth stage, resins and asphaltenes order
into a mesophase. The third and fourth stages are responsible for the room-temperature (steric) hardening of bitumen. The
development of bitumen microstructure and the calculations of the entropy and enthalpy of transitions suggest that bitumen
is a structured amorphous phase with a small crystalline phase (Masson and Polomark 2001; Masson et al. 2002). From the
enthalpy of the endotherms, H, it is estimated that more than 50% of the total endotherm arises from isotropization of these
aromatics, the rest arising from the melting of low-MW saturated segments (Masson and Polomark 2004).

Wax crystallization and melting in bitumen is usually considered detrimental to bitumen quality and asphalt performance.
DSC is used to study the wax morphology in bitumen with respect to time, temperature, and thermal cycling. Eight waxy

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

26

bitumens from different sources and three lab blends prepared by adding a slack wax (a mixture of oil and wax, obtained
from lubricating oil) and two isolated bitumen waxes to a nonwaxy bitumen were characterized using DSC, polarized light
microscopy, confocal laser scanning microscopy, and freeze etching (fracture) in combination with transmission electron
microscopy. The DSC results indicated that the selected bitumen samples differ widely in wax content and initial wax crystal-
lizing and melting-out temperatures. Nonwaxy bitumen displayed no structure or crystals in the other three methods, while
waxy bitumens from different crude origins showed a large variation of structures. The morphology of wax crystals was
highly dependent on crystallization temperature as well as thermal history. The wax that has been isolated from waxy bitu-
men and mixed into nonwaxy bitumen displayed similar morphology as the wax in the original bitumen. It was also found
that bitumen wax usually melted at temperatures lower than 60C although in one case a temperature of 80C was required
to complete melting of the wax (Lu et al. 2005).

The nature and origin of bee-like microstructures (bees) in asphalt binders and their impact on asphalt oxidation have been
the subject of extensive discussions in recent years. Although several studies refer to the bees as solely surface features, some
others consider them to be bulk microcrystalline components that are formed as a result of the co-precipitation of wax and
asphaltene molecules. Pahlavan et al. (2016) use a rigorous theoretical and experimental approach to study the interplay of asphalt
components (mainly asphaltene and wax) and their impact on bee formation In the theoretical section of their study, quantum-
mechanical calculations using density functional theory are used to evaluate the strength of interactions between asphaltene unit
sheets in the presence and absence of a wax component, as well as the mutual interactions between asphaltene molecules (mono-
mers and dimers) and paraffin wax. The results reveal that paraffin waxes not only do not reinforce the interaction between the
asphaltene unit sheets, they destabilize asphaltene assembly and dimerization. This destabilization among interacting systems
(asphalteneasphaltene and waxasphaltene) does not support the hypothesis that interaction between paraffin waxes and nonwax
components, such as asphaltene, is responsible for their co-precipitation and bee formation. To further examine the effect of wax
component on asphalt microstructure experimentally, the authors used AFM to study the surface morphology of an asphalt sample
doped with 1% to 25% paraffin wax. The experiments indicate that paraffin wax tends to crystallize separately and form lamellar
paraffin wax crystal inclusions with 10 nm thickness. Also, the addition of 3% wax into asphalt results in a significant increase
in surface roughness from 0.5 nm to 4.1 nm and an increase in bee wavelength from 651 nm to 1038 nm (Pahlavan et al. 2016).

FOURIER TRANSFORM INFRARED SPECTROSCOPY

FTIR is one of the more important methods for fingerprinting asphalt materials (Lamontagne et al. 2001; Durrieu et al. 2003),
based on its sensitivity and exploiting ease. It is able to quickly offer reliable information regarding aliphaticity, aromatic-
ity, and oxygenation rate. This technique can give more accurate data such as the average distribution length of aliphatic
chains, oxygenation, and substitution mode of aromatics (Lu and Isacsson 1998; Lamontagne et al. 2001). Combining infrared
spectrometry with the specialized use of selective chemical reactions and differential spectra allowed quantification of the
analytical absorption bands of interest including the naturally occurring functional groups; carboxylic acids and their salts,
2-quinolone types, phenolics, and pyrrollitcs; and those formed on oxidation; ketones, anhydrides, small amounts of acids,
and sulfoxides (Petersen 1986).

Relatively low-cost ($20,000 to $40,000) portable devices have become available for FTIR. These can be employed in the
field to test the chemical composition of the delivered materials. These are point-and-shoot applications that could potentially
be used by field technicians with accuracy similar to that obtained by using traditional stationary laboratory equipment
(Zolfka et al. 2013).

Characteristic Absorption Bands

An infrared spectrum is a plot of the energy absorbed at specific wavelengths by the chemical functional groups in the asphalt.
Each type of chemical bond can be identified by characteristic bands in the absorption spectrum of the transmitted IR radia-
tion. The concentration of the functional groups associated with a given absorption can be deduced from the intensity of these
bands using the BeerLambert Law. The absorption maxima for typical groups found in AC are compiled in Table 8.

Although transmission IR spectroscopy is the most efficient technique for making quantitative measurements, an FTIR
spectrometer in the attenuated total reflectance (ATR) mode is the routine tool for studying asphalt samples (Jemison et al.
1990; Yut and Zofka 2011). In an ATR mode, the infrared beam is directed onto an optically dense crystal with a high refrac-
tive index at a certain angle (Figure 10) (Perkin Elmer 2015). The internal reflection creates an evanescent wave that extends
beyond the surface of the crystal, into the sample held in contact with the crystal. This evanescent wave extends only a few

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 27

microns (0.55) beyond the crystal surface and into the sample. If the sample is placed in good contact with the surface of
the crystal, the sample absorbs energy and the evanescent wave will be attenuated or altered. The attenuated energy from each
evanescent wave is passed back to the IR beam and then exits the opposite end of the crystal and is passed to the detector in
the IR spectrometer. The detector then generates the IR spectrum. A graphical representation of ATR phenomena (a) and an
image of the real crystal (b) are shown in Figure 10.

TABLE 8
FTIR ABSORPTION MAXIMA OF ASPHALT FUNCTIONAL GROUPS
Functional Group Wavenumber, cm-1 Species Source
-OH 3,300 Alcohol/phenol Natural
-C-H stretch 3,000 Aromatic Natural
-CH2 stretch 2,920, 2850 Aliphatic Natural
-C(=O)-O-C(=O) 1780 Anhydride Oxidative aging product
-C=O 1,700 Ketone Oxidative aging product
-C=O 1,700 Esters Oxidative aging product
-C=O(OH} 1,650 Carboxylic acid Oxidative aging product
-C=C- aromatic rings 1,600 Natural
-CH2 bending 1,460 Aliphatic Natural
-CH3 bending 1,375 Aliphatic Natural
-O- Ether Natural
-S- 690 Sulfide Natural
-S=O 1,030 Sulfoxide Oxidative aging product
-(CH2)n 745 Aliphatic chain Natural
Source: Petersen (1986).

FIGURE 10 (a) Graphical representation of attenuated total reflectance (ATR)


phenomena; (b) an image of the real crystal (Source: Perkins Elmer 2015).

An FTIR spectrum illustrating the characteristic absorption bands for bitumen is shown in Figure 11. The functional and
structural indices are calculated from the band areas measured from valley to valley. The use of the areas rather than the band
heights allows incorporation of several vibrations of the same type (for example, the C=O ester, acid, and ketone vibrations
between 1,753 and 1,635 cm-1).

One approach for estimating the concentration of several important functional groups using band area ratios is defined
here (Lamontagne et al. 2001).

aromatic structures (aromaticity index) A1600 /

aliphatic structures (aliphatic index) (A1460 +A1376)/A

branched index A1376/(A1460 +A1376)

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

28

long chains index A724/(A1460 +A1376)

oxygenated functions (carbonyl index) A1700 /

sulfoxide A1030 /

The sum of the area represents:

= A1700 + A1600 + A1460 +A1376 + A1030 + A864 + A814 + A743 + A724 + A (2953,2923,2862)

Using band area ratios ensures spectrum normalization so that the bitumen film does not have a constant thickness.

FIGURE 11 Fourier transform infrared spectra showing valley to valley area integration (Source:
Lamontagne et al. 2001).

Asphalt Aging

The extent of aging of a given asphalt binder sample can be followed using the area ratio technique to monitor the oxygenated
functions, carbonyl index, and sulfoxide index. Extensive studies have confirmed that the aging process is dependent on the
type of binder (crude oil origin, refining process) and the type of aggregate used in the mix. Direct FTIR index comparisons
should be limited to the same asphalt/aggregate combination. The potential benefits and limits of the FTIR methods for
reclaimed asphalt characterization based on an international round robin test are summarized by Marsac et al. (2014).

The identification and characterization of the chemical functional types normally present in asphalt or formed on oxidative
aging that influence molecular interactions afford a fundamental approach to the chemical compositional factors that determine
physical properties, which in turn governs the performance properties of both asphalts and asphalt-aggregate mixtures. Elemen-
tal analysis revealed differences in carbon and sulfur contents and an increase in oxygen content on aging. FTIR confirmed that
an increase of oxygen was caused mainly by carbonyl and sulfoxide groups. FTIR also showed a higher content of arylalkylk-
etones and a higher content of oxidizable sulfur compounds in asphalts derived from sour crudes (Michalica et al. 2008).

In addition to the oxidative formation of polar chemical functional groups, aged asphalt physical properties are signifi-
cantly altered by reversible molecular structuring (also called steric hardening). This latter phenomenon is a slow process that
appears to proceed concurrently and synergistically with oxidative aging during the lifetime of the pavement and may be a
major factor contributing to asphalt pavement embrittlement in the later stages of pavement service life. Limited data indicate
that the complex flow properties of asphalt and the tendency of asphalt to form microstructures are directly related.

Additive Identification

The FTIR method is an efficient technique for identifying additives in a binder. Quantitative determination of polymer content
is essential to quality control and quality assurance during the processing and application of PMAs. Research on the quantita-

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 29

tive determination of polymer content in modified asphalt has been conducted (Sun and Zhang 2013). A general description
of four methods for additive analysisnamely, performance-based methods, dissolving-separation methods, gel permeation
chromatography methods, and IR spectroscopy methodsis given. Of these methods, IR spectroscopy is probably the most
appropriate test method to determine several types of polymer [SBR, styrene-butadiene-styrene copolymer (SBS), and EVA
content in modified asphalt].

A typical ATR FTIR spectrum of EVA exhibits several distinctive absorption peaks (Table 9). The polyethylene chain is
characterized by sharp peaks at 2,920, 2,850, and 724 cm-1. The carbonyl peak at 1,735 cm-1 coupled with a double peak at
1,240 to 1,270 cm-1 represents the acetate component in EVA. Finally, two peaks at ~1,640 and ~1,560 cm-1 are associated with
vinyl vibrations. When added to asphalt binder, EVA has two distinctive peaks at ~1,240 and 1,735 cm-1, which can be used
for positive identification of this additive. The most distinctive chemical bonds in a typical styrene-butadiene-based polymer
are aromatic CH bonds in polystyrene and trans-alkene (vinyl) CH bonds in polybutadiene. The out-of-plane vibrations
yield prominent IR peaks at ~700 and ~965 cm-1 for polystyrene and polybutadiene, respectively, and confirm the presence of
SBS. It is important to note that polybutadiene peak at 967 cm-1 is likely to be obscured by a strong and wide band centered
on 1,000 cm-1 that is associated with the silicate component of the aggregates (Zofka et al. 2013).

TABLE 9
FTIR ABSORPTION MAXIMA OF DISTINCTIVE POLYMER FUNCTIONAL GROUPS
Polymer Additive Functional Group Wavenumber, cm 1 Species
Elvaloy EVA Methylene 2,920, 2,285, 724 Polyethylene block
Elvaloy EVA Carbonyl 1,735 Vinyl acetate
Elvaloy EVA Double bond to oxygen 1,240, 1,270 Vinyl acetate
Elvaloy EVA Vinyl 1,640, 1,560 Vinyl substituents
Styrene butadiene copolymers Aromatic C-Hl 700 Polystyrene block
Styrene butadiene copolymers Trans-alkene C-H 965 to 970 Butadiene block
Source: Zofka et al. (2013).

Aging of Polymer-Modified Binders

One of the most important issues in asphalt chemistry is to identify the processes occurring during PMA aging. These processes
involve bitumen aging, polymer aging, or both at the same time (Mouillet et al. 2008). Moreover, most PMAs feature a two-
phase structure made of polymer-rich areas along with polymer-poor regions, depending on the bitumen chemistry, the polymer
nature, and content. It is therefore important to take this binary phase behavior into consideration when trying to sort out the
respective effect on aging of the polymer and the bitumen. IR microscopy, which can focus on separate phases in the blend,
facilitates the characterization of different phases in heterogeneous products. PMAs can be studied in their original state and
after conventional aging tests that simulate the aging during the mixing process and several years of road service (RTFO + PAV
tests). IR microscopy was used to determine for each phase the polymer aging rate and functional indices characterizing the
bitumen, such as aromaticity, aliphaticity, and condensation, and also to map the polymer distribution in the PMA.

Characterization of PMAs in their original state shows that species of bitumen are involved in the polymer swelling and
the effect of the polymer interaction with the bitumen phase. Characterization of the same PMAs after RTFO + PAV aging
shows how the bitumen species responsible for the swelling evolve during aging. In addition, kinetic studies can be performed
using a heating cell fitted to the IR microscope. The chemical composition of the bitumen part of the bitumen phase and the
polymer phase before and after aging could be assessed by subtracting infrared spectra. These studies demonstrate an inter-
dependence of the aging of the different constitutive phases in a PMA and of chemical exchanges between them, which leads
to modification of the micro-morphological structure during aging in a PMA. The general trend for SBS-modified bitumen is
that the binder becomes more homogeneous upon aging. This is the result of both some polymer degradation (chain scission)
and a better compatibility of the smaller polymer chains with the oxidized bitumen molecules.

GEL PERMEATION CHROMATOGRAPHY

Differences in the molecular structures of maltenes and asphaltenes have prompted efforts to separate these components
using size exclusion chromatography, more commonly known as gel permeation chromatography. In theory GPC provides a

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

30

simple separation of molecules in a sample according to their sizes or, more specifically, their hydrodynamic volumes. GPC
is a method of separating molecules based on their size and shape in solution. The column used for separating the molecules
(stationary phase) is packed with a porous bead-like crosslinked polymer network of styrene-divinylbenzene copolymer with
closely controlled pores of variable sizes that can separate molecules in a particular molecular weight range. Depending on
the size and shape, solute molecules may be able to enter the pores of the stationary-phase particles. Molecules larger than
the pores will be totally excluded and will elute first. Very small molecules can enter every pore and permeate well into the
stationary-phase particles. These are retained most and hence appear last in the chromatogram. Intermediate-size molecules
elute at times depending on their comparative size. The size-separated molecules are detectedtypically by either a differen-
tial refractive index (RI) detector or an ultraviolet detectorand recorded according to their concentration. Through calibra-
tion with molecules of known molecular weight, hydrodynamic volumes are converted to MWs and various MW parameters
for the sample are calculated from the MW concentration data. GPCs ability to separate mixtures by molecular size rather
than by some complex property such as solubility or absorptivity is one of the great advantages of the technique. This feature
has made GPC a useful alternate technique for fractionating complicated mixtures, such as crude oil residua, asphalts, and
asphaltenes, for nearly 50 years (Altgelt 1965; Dickie and Yen 1967; Synder 1969; Yapp et al. 1991; Jennings et al. 1993).

GPC ascertains the quantitative distribution of all species present in a binder, including maltenes, asphaltenes, and poly-
mers. The detector signalthat is, the difference between the refractive indices of the eluting solution containing the asphalt
and that of the solvent (RI)is plotted versus the eluting volume (milliliters). The process allows the differentiation of
asphalt species over a molecular weight range of 106 102 Daltons. A set of gel permeation columns is selected to optimize the
separation of asphalt samples. Choice of the solvent system is of great importance, particularly with a complex material such
as asphalt. The fractionation parameters in a given solvent include the sample concentration, temperature, injection volume,
and flow rate. All these parameters affect the column performance and determine the separation efficiency of a given column
set. The solvents most widely used in asphalt analysis are THF and toluene; the associated asphalt components remain unal-
tered in toluene but THF breaks up some association (Jennings et al. 1992b).

The key factors impeding an efficient separation are the tendency of polar materials in asphalt to associate or to be adsorbed
on the column. To a lesser extent, but still important, is the fact that selective solvent interactions can affect the apparent
hydrodynamic volume. For instance, associating substances, such as asphaltenes, show much higher molecular size in a poor
solvent owing to the presence of large complexes. Adsorption of polar species leads to slow desorption from the column (tail-
ing) and distorts the column calibration.

Asphalt from a given source of crude oil has its own characteristic chromatogram that changes only slightly with grade. For
this reason, GPC is a very effective tool for detecting changes in asphalt as a result of processing changes, crude source, or con-
tamination. Each asphalt chromatogram exhibits its characteristic shape, but some of the asphalts show considerable seasonal
change, probably reflecting asphalt processing changes (Glover et al. 1987). It must be emphasized that characterizations of this
kind require that all GPC parameters be held constant. This is a major disadvantage, making comparisons difficult between
laboratories and even over time. An asphalt standard should be run periodically to confirm constant operating parameters.

Fractionation by Molecular Size (Hydrodynamic Volume)

A correlation of the eluting volume with the relative molecular weight of the eluting fraction is difficult to achieve. Because
GPC responds directly to apparent molecular size, it appears to be a simple method for obtaining the molecular weight dis-
tribution of asphalt. However, it turns out not to be a straightforward determination for a number of reasons. The first is the
tendency of some asphalt fractions to associate in solution. These same fractions also may tend to be adsorbed in the column.
A final factor is the chemical complexity of asphalt. It is well known that the order of elution of polar and nonpolar compounds
can be considerably altered by changing solvents, so it is difficult to choose calibrating compounds for such a complex mix-
ture. A detailed evaluation of this calibration problem has been summarized (Davison et al. 1995). The chief utility of GPC
in molecular size distribution measurements is not to obtain absolute values of molecular weight but to determine apparent
molecular sizes that measure the degree of association in asphalts of different properties and composition, particularly to note
the changes that occur during aging. It is likely that the effect of solvent power on the change in apparent molecular size car-
ries information about the internal stability of the asphalt. Thus, a practical approach is to assign apparent molecular weights
relative to commercially available polystyrene standards.

In most of the reported GPC studies, the chromatogram is divided into three or more equal slices defined as large molecular
size (LMS), medium molecular size (MMS), and small molecular size (SMS) (Brule et al. 1986; Price and Burati 1989; Yapp
et al. 1991; Lee et al. 2008). It is also reported that the LMS region can be correlated with the binders physical properties and

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 31

field performance (Glover et al. 1987). Many of these reports have not presented the range of apparent MW of these fractions
relative to polystyrene or other standards, which can reflect the separation efficiency of the column set. A detailed compilation
of conditions reported for GPC determination of asphalts has been reported (Davison et al. 1995), and a standard practice for
determining MW averages and molecule weight distribution of hydrocarbons and terpene resins by size exclusion chroma-
tography (ASTM D 6579) can be applied to asphalts.

A correlation of the eluting volume with the apparent molecular weight of the eluting fraction can be achieved using
narrow-MW polystyrene standards (Daly et al. 2013). The apparent molecular weights of the eluting fractions, the GPC
chromatograms, have been divided into three regions: LMS, MMS, and SMS based on elution volume. Researchers have
stated that the LMS and SMS regions are significant with respect to predicting pavement performance (Garrick and Wood
1986; Garrick 1994; Wahhab et al. 1996, 1999). The impact of asphalt aging on the GPC chromatograms has been reported
extensively (Kim and Burati 1993; Churchill et al. 1995; Siddiqui and Ali 1999; Negulescu et al. 2006; Shen et al. 2006; Lee
et al. 2008; Xiao et al. 2009). Although the division of the chromatograms into arbitrary regions based on elution volume is
used routinely, it is preferable to calibrate the GPC chromatograms and identify the maltenes, asphaltenes, and polymer com-
ponents on the basis of their apparent MW ranges (Daly et al. 2013). By using MW regions, it is possible to divide the LMS
fraction into ranges that change when the asphalt ages or is modified.

To define LMS, MMS, and SMS fractions, the chromatogram was divided into three slices based on the apparent MW of
the eluting species using a calibration curve from polystyrene standards (Daly et al. 2013). The three fractions are polymers
(MW greater than 19,000), asphaltenes (MW from 19,000 to 3,000), and maltenes (MW less than 3,000), as shown in Figure 12.
Quantitative data can be obtained by determining the area under the curve. As asphalt ages, the asphaltenes aggregate and begin
to contribute to the polymer fraction, so the increase in the concentration of materials in the polymer fraction reflects the degree
of aging. Therefore, the polymer region was subdivided into three fractions: very high molecular weight, with MW greater
than 300,000 Daltons; high molecular weight, with MW between 45,000 and 300,000 Daltons; and medium molecular weight,
with MW between 19,000 and 45,000 Daltons. The polymers added to modify the asphalt binders generally exhibit MWs above
45,000 Daltons so the medium molecular weight subfraction reflects aggregates formed during binder aging. Deconvolution of
the chromatogram facilitates quantitative analysis by integration of the area under the curves shown in Figure 13.

FIGURE 12 Molecular weight zones assigned in PMAC gel permeation


chromatography (GPC) chromatogram (Source: Daly et al. 2013).

Earlier determinations by osmometry indicated that the average MW of maltenes (as heptane soluble binder fraction) is
700900 Daltons and that of asphaltenes (as heptane insoluble binder fraction) ranges between 2,000 and 10,000 Daltons
(Dickie and Yen 1967). Morgan et al. (2010) used laser desorption mass spectrometry along with size exclusion chroma-
tography and planar chromatography to study the MW of maltenes and asphaltenes of Mayan crude oil. The fractions were
separated using extraction with pentane. The results revealed a small portion of asphaltenes extending to 10,000 Daltons, and

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

32

maltenes extending to 2,000 Daltons, but continuing to fall within the polymer ranges defining the maltene and asphaltene
regions of the chromatograph. These MW data have been confirmed by the GPC method (Daly et al. 2013). The polymer and
asphalt components of polymer-modified asphalt cements could be separated completely with accurate determination of the
molecular weight of species by calibration with standard narrow-MW polystyrenes (Figure 14).

FIGURE 13 Determination of maltenes and asphaltenes content


of PG 64-22 binder by deconvolution of the GPC curve (Source:
Daly et al. 2013).

FIGURE 14 GPC elution curve of PG 70-22M containing 1%


polymer extracted from a mixture aged for 5 days at 85C
(Source: Daly et al. 2013).

Correlating Physical Properties with GPC Results

Attempts to correlate asphalt physical properties with chemical properties have not been particularly successful. This no doubt
is primarily the result of the lack of uniqueness in the chemical properties. For instance, a Corbett fraction from one asphalt
may have very different physical properties from those of the same fractions from another asphalt. Also, two asphalts with
similar physical properties can have radically different GPC chromatograms (Davison et al. 1995).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 33

Elseifi et al. (2010) studied the relationship between asphalt binder deformation properties at intermediate and low tem-
peratures, its molecular composition, and mix performance. Nine straight-run binders obtained from two asphalt suppliers
were tested using ductility and direct tensile tests. To assess the results of these tests, selected asphalt binders were evaluated
using GPC, DSC, and dynamic mechanical analysis. Measurements showed that an inverse correlation exists between binder
ductility at intermediate temperatures and failure strain at low temperatures. In other words, a binder that provides high duc-
tility at intermediate temperatures would be characterized by poor elongation properties at low temperatures. This trend was
related to the binder molecular composition, as characterized by GPC. An increase in the binder content of low MW results
in an increase in binder ductility at intermediate temperatures.

If the aggregate is heated higher than the specified level, the asphalt binder in the mixture will be aged at a much higher
level during the short-term oven aging (STOA) period. The asphalt binder in that mixture will be oxidized (aged) more than
expected during STOA time because of the highly elevated aggregate temperature. If the binder in the mixture is severely
oxidized, the asphalt pavement will have a diminished service life. A gel permeation chromatography technique was used on
the mixture particles without binder extraction to estimate the significance of aging for each case of STOA (Kim et al. 2016).

Churchill et al. (1995) evaluated three different slice profiles (three, four, and 10 slices) to correlate aging times with the GPC
LMS region. Partitioning of chromatograms into 10 segments provided a better resolution for predicting aging viscosity, aging
index, viscosity number, and penetration from GPC data. It was also found that the aging level of binders in the warm-mix asphalt
mixture was significantly lower than that of binders in hot-mix asphalt (Kim et al. 2016). GPC chromatograms were divided into
13 slices to evaluate relative impact of an RTFO test versus STOA on the properties of nine binders and six field-aged samples
(Lee et al. 2008). The LMS (slices 15) ratio of aged to unaged samples consistently coincided with the aging times. The RTFO
method was found to induce less aging than the STOA method of aging asphalt mixtures in the laboratory. Estimation of binder
viscosity of recycled asphalt pavement (RAP) could be accomplished using GPC by direct sampling. The results exhibited a
positive correlation with binders from corresponding samples recovered by the Abson method (Kim et al. 2006).

A correlation of multiple stress creep recovery (MSCR) parameters with elastic recovery and molecular weight of differ-
ent polymer-modified binders was demonstrated (Batten et al. 2011). The MSCR test that was recently developed by FHWA
enables the correct grading of the field performance of polymer-modified asphalts. The polymer modification of asphalts is
one of the solutions to overcome factors leading to asphalts demise; that is, the corresponding asphalt concrete defects such
as rutting, thermal cracking, fatigue, and stripping. However, the different polymer-modified asphalts can behave differently
even though they have the same performance grade. The MSCR test measures the high temperature binder specification
parameter, called nonrecoverable creep compliance or Jnr, and percent recovery. In this study eight different polymer-modi-
fied binders are measured. The molecular weights were determined using GPC. The correlation of the MSCR results with the
molecular weight provides an insight into how modification affects mechanical response and the rutting potential.

Mathematic models have been generated to predict rheological properties of asphalt cement based on the molecular size of its
components. A set of polymer-modified asphalt samples with 5%, 10%, and 15% crumb rubber and 3%, 6%, and 9% styrene-buta-
diene-styrene was prepared to evaluate the effect of polymer modification on the molecular size distribution of the asphalt samples.
In general, the distribution of increasing molecular size could be correlated with changes in the asphalt binder physical properties.
Predictive models could be generated for the different rheological properties from GPC chromatograms (Wahhab et al. 1999).

NUCLEAR MAGNETIC RESONANCE ANALYSIS

Proton nuclear magnetic resonance (1H NMR) spectroscopy has emerged as a very powerful and versatile tool for bitumen
characterization (Jennings et al. 1992b; Borrego et al. 1996; Jain et al. 1998). 1H NMR allows investigations in solids as well
as in solution. As an alternative to the conventional analysis, the 1H NMR spectrometry method does not require sample pre-
treatment and thus considerably reduces manipulation time. The 1H NMR method is also capable of simultaneously detecting
and quantifying a number of constituents in a single spectrum. The direct 1H NMR spectrometry quantitative method presents
advantages over some routine methods: simplicity, rapidity, selective recognition, and quantitative determination of aliphatic
hydrogens and aromatic hydrogens in bitumen.

Asphalt Molecular Structure

Ramsey et al. (1967) published the first structural characterization of asphalts by 1H NMR. Hasan et al. (1983) published a
more structured analytical approach to characterize a petroleum vacuum distillation residue (boiling point >454C); later,

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

34

this methodology was slightly modified by Siddiqui and Ali (1999) and Siddiqui (2010). Since then other authors have used
the 1H NMR spectrum, sometimes the carbon-13 nuclear magnetic resonance (13C NMR) spectrum, and elemental analysis
for the structural characterization of asphalt fractions or similar materials to define indices that intend to correlate with the
properties and future behaviors of asphalts in roadworks (Huang 2010; Ma et al. 2011).

Betancourt Cardoza et al. (2016) improved the characterization of some heavy fractions of petroleum based on elemental
analysis, 1H NMR, and 13C NMR spectra of some heavy fractions of petroleum. They proposed a methodology useful to
characterize oil, coal, or its fractions using a method of data processing and spectra interpretation that benefited from recent
advances in NMR (in hardware and software).

Three unfractionated fresh commercial asphalts named P1, P2, and P3 produced in Colombia were examined. The viscos-
ity, the penetration (80/100), the softening point, and the colloidal instability index (Table 10) indicate that P1 and P2 have
a softer consistency than P3. The three asphalts show high thermal susceptibility, with P3 being the most sensitive (penetra-
tion index 1.4). This result agrees well with the common origin. The SARA fractionation data (Table 10) suggest that their
higher content of resins and their smaller fraction of aromatics justify the softer consistency of P1 and P2 compared with P3
(Betancourt Cardozo et al. 2016).

TABLE 10
PHYSICAL-CHEMICAL PROPERTIES OF FRESH AND UNFRACTIONATED ASPHALTS
Samples
Assay Units P1 P2 P3
Conventional assays viscosity at 60C Poises 1403 1263 2193
Viscosity at 135C Poises 3.10 2.89 4.23
Viscosity at 150C Poises 1.27 1.17 1.46
Ductility at 25C cm >150 >150 >150
Penetration at 25C 0.1 mm 84 93 64
Specific gravity 1.004 1.006 1.008
Softening point C 45.8 44.4 47.0
Colloidal instability index 0.36 0.36 0.38
SARA separation
Saturatesa % 14.18 13.86 13.54
Aromaticsb % 39.59 39.59 43.43
Resinsc % 33.69 33.06 29.03
Asphaltenesd % 12.53 13.48 13.99
Elemental analysis carbon % 85.69 85.73 85.84
Hydrogen % 10.50 10.45 10.42
Nitrogen % 0.94 0.99 1.02
Sulfur % 1.64 1.65 1.65
Oxygene % 1.24 1.19 1.07
Source: Betancourt Cardozo et al. (2016).
Uncertainty of a 0.50; b 0.83; c 0.54; d 0.67.
e Measured values, not deducted as the difference from 100%.

The elemental composition (Table 10) and the 1H NMR spectra (Table 11) show very similar quantities of hydrogen in
the different structural fragments (aliphatic, aromatic, olefinic, or phenolic) of P1, P2, and P3. Although all these data do not
discriminate clearly between the three asphalts, they evidence a predominance of alkyl substituents with a length of four or
more carbon atoms in their average structure. Most of the hydrogen in its aliphatic fraction belongs to CH n moieties located
at three or more bonds from the aromatic fragment. In total, more hydrogen was found bonded to carbons in positions , , ,
or farther (H, H, H, ...) than bonded to carbons in positions to aromatic fragments, H.

13C
NMR spectra (Table 12), viscosity, penetration, and the softening point (Table 10) confirm some degree of similitude
between P1 and P2, and show differences between them regarding P3. Table 10 data show that the content of carbon bonded
to hydrogen present as methyl, methylene, or methine in the structure of P1, P2, or P3 is quite similar. The quaternary carbon
content in alkyl or aryl fragments is practically equal in the structure of P1 or P2. P3 has the lowest content of quaternary

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 35

carbon in its aliphatic fragments and the highest content of quaternary carbon in polycondensed aromatic fragments (most of
these are catacondensed carbons bonded to a heteroatom or an alkyl group distinct to methyl).

TABLE 11
HYDROGEN PERCENTAGE DETERMINED BY 1H NMR SELECTED ASPHALTS
Samples
Type (assignments) Interval (ppm) P1 P2 P3
General aliphatic or saturatesa 0.54.5 94.48 94.33 94.35
Olefins 4.66.2 0.26 0.21 0.21
Aromatic 6.39.3 5.23 5.44 5.43
Phenolic OH 5.09.0 0.02 0.02 0.02
Particular undefined
CHn , or more to aromatic 0.52.0 46.04 46.08 45.92
CHn , , or more to aromatic 0.52.0 81.37 81.26 80.95
CHn to aromatic 0.52.0 35.32 35.19 35.04
CHn to aromatic 2.04.5 13.10 13.05 13.38
CHn to CC (it is CHnCC) 1.92.1 0.01 0.01 0.01
CHn in a C sp3 joined to oxygenb 3.13.3 0.01 0.01 0.01
CH3, CH2 or CH , or morec 0.51.0 16.08 16.34 16.32
CH3, CH2 or , CH or d 1.02.0 59.93 59.48 59.19
CH in monocyclic aromatice 6.37.3 1.85 2.07 1.94
CH in polycyclic aromatic 7.29.3 3.38 3.37 3.49
a Paraffins and naphthalenes
b CH n-C-O
cTo aromatic
d To aromatic and CH or CH
2
e Estimated by excess (6.37.3).

Source: Betancourt Cardozo et al. (2016).

TABLE 12
CARBON PERCENTAGE DETERMINED BY NMR IN SELECTED ASPHALTS
Type (assignment) Interval (ppm) Samples
P1 P2 P3
General aliphatic or saturates 1060 70.5 71.35 68.11
Olefinic 105153 1.38 1.05 1.14
Aromatic 102165 28.12 27.6 30.74
Carbon types
Total CH3 2428 2.69 2.66 2.66
Total CH2 2360 8.27 8.17 8.13
Quaternary sp3 1060 56.09 57.10 53.91
Total aromatic CH 102131 1.55 1.6 1.59
Quaternary sp2 123165 26.57 26 29.15
Source: Betancourt Cardozo et al. (2016).

Using 1H and 13C NMR can yield information on average structural parameters of asphalt and asphaltenes, such as percent-
ages of aromatic carbons, aliphatic carbons, bridged carbons, methyl carbons, ring carbons, naphthenic carbons, paraffinic chain
lengths, and other parameters. The specific environment of the different types of hydrogens and carbons can be defined (Betancourt
Cardozo et al. 2016). Thus, NMR spectroscopy is a powerful tool for predicting the structure of complex organic molecules.

Mechanisms of Asphalt Aging

Combining NMR and GPC information, Saddiqui suggested possible structures for asphalt and mechanisms of aging (Sid-
diqui and Ali 1999; Siddiqui 2010). This generic model incorporates aliphatic branch structures elaborating pericondensed
rings with low ratios of hydrogen to carbon. The presence of heteroatoms in the rings is shown in Figure 15.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

36

FIGURE 15 Possible representation of reaction types in a hypothetical asphalt


structure on laboratory aging (Source: Siddiqui and Ali 1999).

In this work (Siddiqui and Ali 1999), the chemical properties of a commercial grade Saudi Arabian asphalt procured
from Ras Tanura refinery were evaluated. The RTFO short-term aging and PAV long-term aging tests were used to simulate
the laboratory aging of this asphalt. PAV has more severe effects on the chemical properties of asphalt than does the RTFO
method. The Corbett fractionation procedure was used to separate fresh and aged asphalts into four generic fractions; namely
asphaltenes, polar aromatics, naphthene aromatics, and saturates. Various analytical techniques were applied to evaluate the
chemical changes that occurred during the aging processes. High pressure-gel permeation chromatography molecular weight
and size distributions suggested that molecular rearrangements occur predominantly on aging. Based on the molecular size
distribution of isolated asphaltenes, it is suggested that dissociation, isomerization, and fragmentation were predominant reac-
tion types that occurred in asphaltenes during extensive PAV laboratory aging. Further chemical reactions, such as molecular
association, polymerization, or polycyclic condensation within the complex asphaltenes, were inferred from NMR studies of
extensively aged asphaltene fractions (Siddiqui 2010).

X-RAY FLUORESCENCE SPECTROSCOPY

X-ray fluorescence spectroscopy (XRF) is a simple technique for quantitative analysis of elements ranging from sodium to
uranium in the periodic table. The XRF method is widely used to determine the elemental composition of materials. Because
this method is fast and nondestructive to the sample, it is often used in field and industrial applications for quality control.
Portable XRF devices facilitate measurements in the field. The heavy metal content of an asphalt sample can be used to iden-
tify some asphalts. XRF can detect other characteristic elements in binder additives such as phosphorus (P) in polyphosphoric
acid (PPA); calcium, zinc, and molybdum in recycled engine oil bottoms (REOB); and crumb rubber modifier (CRM). The

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 37

use of XRF for forensic analysis of HMAs is facilitated by the ability to determine the concentration of multiple elements
simultaneously.

XRF is the emission of characteristic secondary (or fluorescent) X-rays from a material that has been excited by a bom-
bardment of high-energy X-rays or gamma rays. When materials are exposed to short-wavelength X-rays or gamma rays, their
component atoms may ionize. As a result, the material emits radiation at the energy characteristic of the atoms present. The
term fluorescence is applied to phenomena in which the absorption of high-energy radiation results in the re-emission of
lower-energy radiation (Jenkins 2012). The wavelength of this fluorescent radiation can be calculated and analyzed either by
sorting the energies of the photons (energy-dispersive analysis) or by separating the wavelengths of the radiation (wavelength-
dispersive analysis). Once sorted, the intensity of each characteristic radiation is directly related to the amount of each element
in the material.

Asphalt contains high levels of sulfur and varying amounts of iron, vanadium, and nickel. Zinc compounds are added to
control hydrogen sulfide emissions in the refinery. The distribution of the heavy metal elements in unmodified binders is a
fingerprint that can be used to identify their corresponding crude oil sources.

AC does not naturally contain phosphorus. If it is assumed that all the phosphorus comes from PPA, measurement of the
phosphorus content using XRF can estimate the amount of PPA used to produce the binder. However, other phosphorus-con-
taining additives such as REOB can distort the results (Reinke and Glidden 2010). Waste engine oil contains copper and iron
(wear metals) as well as zinc and molybdenum from anti-wear and corrosion-inhibiting oil additives. The phosphorus source
in lube oil is zinc thiophosphate. Zinc and molybdenum are not found in natural asphalt, therefore XRF analysis of these ele-
ments can help identify the presence of REOB. Quantifying REOB content can be achieved by determining the intensity of
those elements and correlating that intensity to a known dosage of REOB with a calibration standard (Clifton et al. 2016).

Ground tire rubber (GTR) also contains some of the metals (zinc and iron) found in REOB, so analysis of REOB in crumb
rubber-modified binders is quite complicated since the levels of zinc in GTR are much higher than those in REOB (Arnold and
Shastry 2014). If no REOB is present there is a linear correlation between the GTR content and zinc (Reinke and Glidden 2010).

FIGURE 16 Schematic of an AFM scanning an asphalt surface (not


to scale).

ATOMIC FORCE MICROSCOPY

In simple terms, the atomic force microscope has a nanometer-sized tip at the end of a cantilever that probes the surface of
the material (typically over an area of 1050 micrometers x 1050 micrometers) recording the topography of the surface and
changes in the surfacetip interaction. Figure 16 shows a schematic of an atomic force microscope (not to scale) scanning the
surface of an asphalt binder. For imaging, the reaction of the probe to the forces that the sample imposes on it can be used to
form an image of the three-dimensional shape (topography) of a sample surface at a high resolution. This is achieved by raster
scanning the sample surface and recording the height of the probe that corresponds to a constant tipsample interaction. The
surface topography is commonly displayed as a pseudocolor plot. An AFM image is a simulated image based on the height of

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

38

each point of the surface, and in fact each point (x, y) of the surface has a height h (x, y). The color variations reflect the rela-
tive heights of the surface topography. The impact of asphalt chemical composition on the microstructure and performance
characteristics of asphalt binders was studied. The methods implemented included adsorptiondesorption chromatography
analysis and a range of AFM and chemical force microscopy techniques. AFM can be a powerful complementary tool to rhe-
ology and spectroscopy for characterizing asphalts. It is possible to compare the mechanical properties of pure and modified
bitumen using appropriate techniques.

Asphalt Microstructure

An intriguing AFM image of an asphalt surface is a peculiar bee structure that was initially found only for a gel bitumen.
The phase images, 25 m by 25 m, of one of the asphalts are shown in Figure 17, where the continuous phase is represented
with white color contrast. The impact that each SARA chemical fraction has on asphalt phase structuring, most notably the
well-known asphalt bee structures, is shown. Certain asphalt chemical parameters have a consistent and measurable effect
on the asphalt microstructure that is observed with AFM. Particular microstructures that emerged through chemical doping
were then discovered to have unique chemical polarity, which explicitly impacts the durability and performance of asphalt.
A surprising correlation was found between the saturates chemical parameter and the effects of oxidative aging on asphalt
behavior (Allen et al. 2014).

FIGURE 17 AFM phase images of 25 m by 25 m of Asphalt ARC BI0002 derivatives


taken at room temperature (25C): (a) control; (b) asphaltene-doped blend; (c)
naphthenic aromatic-doped blend; (d) polar (resin)-doped blend; and (e) saturate-
doped blend (enlarged locations are 10 m by 10 m for each asphalt blend)
(Source: Allen et al. 2014).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 39

The same bee structure (also called catana phase) was repeatedly observed in other works with an average height between
22 and 85 nm and a typical distance between strips on an order of 150 nm (Masson et al. 2006). A link between the extent of
the bee phase and the asphaltenes has been confirmed in one study (Masson et al. 2007). The scanning electron microscopy
observation of the same gel bitumen that gave the bee structure in AFM showed connecting aggregates of what was believed
to be asphaltene particles with a diameter of around 100 nm (Loeber et al. 1996).

Several studies using the AFM and mathematical modeling at a similar scale have revealed that the binder has a distinct
microstructure that is related to its chemical composition and thermal history (Masson et al. 2006; Schmets et al. 2010; Das et
al. 2013; Allen et al. 2014; Rebelo et al. 2014; Bhasin and Ganesan 2015; Zhao et al. 2015; Menapace et al. 2015). Studies have
also shown that such a structure can influence the nucleation of damage within the binder (Jahanigir et al. 2015). Although
most of the aforementioned studies have been limited to the surface of the binder, some studies have also demonstrated that
structures observed on the surface are correlated with similar, albeit smaller-sized, structures (referred to as ant-like struc-
tures in some literature) below the surface in the sample mass (Fischer and Dillingh 2014; Ramm et al. 2016). For example,
Ramm et al. (2016) used optical techniques to demonstrate that (1) these structures exist in the bulk of the binder; (2) for a
given chemical composition and temperature, the thermal history of the binder can be manipulated to alter the size and dis-
tribution of the structures; and (3) the structures influence the rheological properties of the binder, such as complex modulus
(e.g., in some cases the G* of the binder varied by 20% at the same temperature, composition, and rate of loading, simply by
manipulating the thermal history and concomitant internal structure).

Property Changes at a Microscopic Level

AFM results were presented for a base binder and the mastics prepared with granite, portland cement, and hydrated lime. A
clear identification of three micro domainsparaphase, periphase, and beeswas displayed. It was apparent that the rela-
tive proportion and size of the features changes with the addition of filler and appears to be filler-specific, providing evidence
of physico-chemical interactioninduced changes to the effective binder matrix. Image analysis was used to determine the
relative surface area each micro domain occupies for the materials studied. The size decreased and relative number of bee
structures increased with the addition of filler. The relative change in the size, relative quantity, and special distribution of bee
features is greatest for the mastic prepared with lime, indicating a higher degree of physico-chemical interaction intensity.
This was expected because the lime filler has a much higher specific surface area than the other fillers and hence, given the
same volume concentration, provided more opportunity for adsorption. A greater stiffening effect was observed for fillers
that induce the greatest microstructural changes as a result of physico-chemical interaction. Physico-chemical interaction is
anticipated to lead to a softening of the effective asphalt binder matrix since polar components adsorb to the surface of the
filler. Thus, results suggest the formation of an adsorbed interphase layer plays a more critical role in determining the macro-
scopic rheology of mastics than does softening of the binder matrix (Davis and Castorena 2015).

Conventional rheological and chemical tests provide a global view of asphalt property and composition changes upon
aging, but offer little details on changes at the microscopic level. Using AFM, Yuhong Wang et al. (2015) analyzed the micro-
mechanical properties of five asphalts with different aging conditions. Aging was found to significantly increase the spatial
variations of the sample properties. The asphaltene content and the size of microstructures both appear to affect the microme-
chanical properties of the binders. Using two different antioxidant additives, PPA and cashew nut oil, at concentrations of 1%
and 2%, respectively, it is possible to use the AFM technique to distinguish the effects of both additives in the morphological
and micromechanical properties of bitumen films (Rebelo et al. 2014).

Moisture damage in polymer-modified asphalts has been studied for decades, yet the effects of chemical functional groups
on moisture sensitivity are not known. A nanoscale experiment was conducted to measure these effects in terms of adhesive/
cohesive forces using AFM. A base asphalt binder and asphalts modified by two polymers (styrene-butadiene and styrene-
butadiene-styrene) were tested in dry and wet conditions. Using the AFM, these samples were probed by silicon nitrite
(Si3N4), carboxyl (-COOH), methyl (-CH3), and hydroxyl (-OH) functionalized AFM tips, and nanoscale pull-off or adhe-
sion/cohesion forces between asphalt and tip molecules were measured. Based on the ratio of wet to dry adhesion/cohesion
forces, it was shown that the polymer modification makes binders less susceptible to moisture damage. The moisture perme-
ability is impacted by the polymer molecular weight and compatibility with the binder. Among the four tips, the -COOH tip
showed almost no difference in adhesion forces between wet and dry samples. Using -OH tips shows that the cohesion in SBS-
modified wet asphalt samples is significantly higher than the cohesion in SB-modified wet asphalt samples. The Si3N4 tip
showed higher adhesion in SB-modified wet samples than in the SBS polymer-modified wet samples. Based on the adhesion/
cohesion force, 3% polymer is found to be optimum for minimizing moisture susceptibility with both SB and SBS polymers.
This study illustrates the potential for nanoscale AFM testing on asphalt binders (Tarefder and Zaman 2010).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

40

Worldwide, several research groups are reporting AFM results on bitumen, so it is becoming important to improve the
understanding of the reproducibility and objectivity of the technique for studying bituminous samples. When reproducibility
and stability are proven, AFM can be a tool for asphalt professionals to rapidly screen bituminous binders. In this context,
two independent laboratories have developed a standard method for preparing and conditioning bitumen for AFM imaging.
By means of an interlaboratory comparison of independently imaged specimens, the reproducibility of microstructure mea-
surements was investigated. A quantitative comparison on different microstructures was developed, and the consistency of
independently obtained results was confirmed. The results from both labs were comparable: the microstructural properties
were found to be randomly distributed within a 5% interval. Also, the influence of temperature on the microstructure was
demonstrated to be reproducible and consistent. With the increase of temperature, the microstructure gradually disappeared;
however, traces of the microstructure remained visible up to the highest measurement temperature of 60C. The conclusion is
that given well-defined sample preparation and measurement procedures, the microstructure of bitumen can be reproducibly
imaged by AFM from room temperature up to temperatures where bitumen becomes liquid (Nahar et al. 2013).

The current understanding of bitumens surface microstructures as characterized by AFM is evolving. Microstructures of
bitumen develop in different forms depending on crude oil source, thermal history, and sample preparation method. Although
some bitumens display surface microstructures with fine domains, flake-like domains, and dendrite structuring, bee struc-
tures with wavy patterns several micrometers in diameter and tens of nanometers in height are commonly seen in other bind-
ers. Controversy exists regarding the chemical origin of the bee structures, which has been related to the asphaltene fraction,
or the crystallizing waxes in bitumen. The rich chemistry of bitumen can result in complicated intermolecular associations,
such as co-precipitation of wax and metalloporphyrins in asphaltenes. Therefore, it is the molecular interactions among the
different chemical components in bitumen, rather than a single chemical fraction, that are responsible for the evolution of bitu-
mens diverse microstructures, including the bee structures. Mechanisms such as curvature elasticity and surface wrinkling
that explain the rippled structures observed in polymer crystals might be responsible for the formation of bee structures in
bitumen. Despite progress on morphological characterization of bitumen using AFM, the fundamental question of whether
the microstructures observed on bitumen surfaces represent its bulk structure remains to be addressed. Combining AFM
with other chemical analytical tools that can generate high resolution comparable to AFM would provide an avenue to linking
bitumens chemistry to its microscopic morphology and mechanical properties and consequently benefit the efforts focusing
on developing structure-related models for bituminous materials across the different-length scales (Yu et al. 2015).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 41

CHAPTER FOUR

ASPHALT ADDITIVES AND MODIFIERS

INTRODUCTION

As refiners efficiency allows them to extract more gasoline and other petroleum products from crude oil and as the source of
crudes that yield quality asphalt residua decreases, the need for additives to upgrade straight-run asphalts increases. Further, the
physical specifications for ACs defined by the PG system have prompted the use of more additives and modifiers to meet state
DOT requirements. Highway agencies have recognized the benefits of using modified asphalts to reduce the amount and severity
of pavement distresses and to increase service life. The primary benefit of using these high-performance asphalts is improved rut-
ting resistance, with less thermal (cold-temperature) cracking and overall improved mixture durability being secondary benefits.
Additionally, some modified binders provide improved stripping (moisture damage) resistance. Many agencies estimate that an
additional 4 to 6 years of pavement life from a pavement constructed using a modified asphalt binder is a reasonable expectation.

The state DOT or other agency responsible for managing construction may not know what type of modifier is used in its
binders. The goal of Superpave was to make the PG system blind to the modifier that is used. As long as the asphalt meets the
PG requirements it should not matter what modifier is used. The type of additive used may be proprietary and not necessarily
known by the purchaser. In practice the asphalt supplier will normally provide some information about the type of modifier
used but may not give specific details.

Modifiers and additives being used to boost AC performance include polymers, chemical modifiers, extenders, oxidants
and antioxidants, hydrocarbons, and antistripping additives. Polymers cover a broad range of modifiers, with elastomers
(rubbers or elastics) and plastomers (plastics) being the most commonly used types. Styrene-butadiene rubber, styrene-
butadiene-styrene, and crumb rubber are frequently used elastomers. SBS is the most often used modifier. These modifiers
are used to reduce rutting and to improve fatigue and thermal cracking resistance. Crumb rubber is an elastomer made from
ground tires. Several technologies are in place for using ground tire rubber. This material is used primarily to address rutting
and fatigue. Plastomers are used to improve the high-temperature (rutting) properties of modified materials. Low-density
polyethylene (LDPE) and EVA are examples of plastomers used in asphalt modification.

The most commonly used chemical modifier is polyphosphoric acid. This modifier may be used in combination with poly-
mers to increase the high-temperature stiffness. Other modifiers that may be used include asphalt binder extenders (primarily
sulfur) and hydrocarbon materials. Hydrocarbons can produce either hardening or softening effects. The modifiers also include
antioxidants, antistripping agents, softening agents, stiffening agents, wax-based additives, and recycling (rejuvenating) agents.

The increase in the amount of modifiers used in asphalt cements can primarily be attributed to the following factors (Rob-
erts et al. 2009):

1. An increased demand on HMA pavements. Traffic volume, traffic loads, and tire pressures have increased significantly
in recent years, requiring that HMA be more resistant to rutting.

2. Superpave binder specification requires the asphalt binder to meet the stiffness requirements at high as well as low
pavement service temperatures. Most neat (non-modified) asphalt cements do not meet these requirements in regions
with extreme climatic conditions.

3. Environmental and economic pressure to dispose of some waste materials and industrial by-products (such as tires,
roofing shingles, glass, sulfur, and ash) as additives in HMA.

4. Public agency willingness to pay a higher first cost for pavements with a longer service life or that will reduce the risk
of premature failure.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

42

Depending on the asphalt source and the average climatic conditions, the main reasons for including a given modifier are
the following (Roberts et al. 2009):

Increase the stiffness of the mixture to minimize rutting


Soften and increase the elasticity of the mixture to minimize cracking
Improve the fatigue resistance of the mixture
Improve asphalt-aggregate binding to reduce stripping or moisture sensitivity
Improve abrasion resistance to reduce raveling
Rejuvenate aged asphalt binders
Reduce flushing or bleeding
Improve resistance to aging or oxidation
Reduce structural thickness of pavement layers
Reduce life-cycle costs of HMA pavements
Improve overall performance of HMA pavements.

An ideal binder should be modified to achieve the properties illustrated in Figure 18.

FIGURE 18 Stiffness characteristics of conventional binder and ideal modified


binder (Source: Terrel and Epps 1989a, b).

The lower service temperature region prefers lower stiffness and faster relaxation properties to reduce thermal cracking.
Lower stiffness (or viscosity) is desired at high construction temperatures to facilitate pumping of the liquid asphalt binder,
and mixing and compaction of HMA, but a higher stiffness should be retained at high service temperatures to reduce rutting
and shoving. Adhesion between asphalt binder and aggregate should be strong in the presence of moisture to reduce stripping.
Most of these criteria can be met with polymer additives.

POLYMER ADDITIVES

Elastomers

Currently the most common polymer additives are styrene-butadiene-styrene copolymer, SBR, and EVA. Polymers are mac-
romolecules made by chemically reacting many (poly) smaller molecules (monomers) with one another to form long chains
or clusters. The sequence and chemical structure of the monomers determine the physical properties of the resulting polymer.
Copolymers consist of a combination of two different monomers that can be in a random or block arrangement. For example,

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 43

polystyrene is a hard, brittle plastic whereas polybutadiene is soft and rubbery. If these two distinctly different monomers
are randomly mixed and reacted together, a new polymer called a copolymer is formed, with varying properties depending
on the molar ratio of monomers incorporated into the chains. Polymers can be engineered to obtain a broad range of physical
properties. However, they can be divided into three general categories: fibers, plastics (plastomers), and rubbers (elastomers).
The divisions are defined by the polymer thermal properties and morphology. A fiber will be predominately crystalline with
a high melting point and a modest glass transition temperature. A plastic is predominately amorphous with a Tg higher than
its use temperature. A rubber is amorphous with a Tg well below its use temperature; to prevent the rubber molecules from
flowing at room temperature, the rubber is crosslinked.

The SBR is usually crosslinked with sulfur (vulcanization); the level of crosslinking controls the final properties ranging
from a soft flexible material suitable for tire treads to a hard material suitable for bowling balls. The SBR emulsions used as
modifiers are lightly crosslinked rubber that does not completely melt during processing. The SBR is normally introduced as
a latex emulsion and the water content is flashed from the asphalt cement.

SBS triblock copolymer is a self-crosslinking elastomeric material; the polystyrene blocks have a Tg above the use tem-
perature and serve as crosslinks. When the material is heated above the Tg of the polystyrene block, the material becomes a
free-flowing liquid that is easy to disperse at asphalt processing temperatures. The structure of the SBS may vary from a linear
polymer chain to a branched polymer chain that is designated as a radial copolymer. Radial SBS copolymers exhibit lower
melt viscosities than linear SBS copolymers and thus allow lower processing temperatures.

The polymers are used to increase the PG high temperature grade of the binder. The benefit of the polymer modifiers will
depend on the concentration, morphology, molecular weight, chemical composition, and molecular structure of the material.
The crude source, refining process, and grade of the neat asphalt binder are equally important. The residual reactivity of the
polymer is useful for improving compatibility of the additive with the binder.

Plastomers

The utilization of plastomers in asphalt modification is limited. Polyethylene, which can be found in three formsLDPE,
high-density polyethylene (HDPE), and linear low-density polyethylene (LLDPE)is the most common plastic. Other poly-
olefins employed include polypropylene and ethylene-propylene copolymer, and EVA copolymer. Although the modification
of bitumen with virgin polymers can improve the properties of asphalt mixtures, the use of recycled plastic may also show a
similar result with additional environmental advantages (Garcia-Norales et al. 2006). A recent study evaluated the possible
advantages of modifying the bitumen with different plastic wastes, namely polyethylene (HDPE and LDPE), EVA, acryloni-
trile-butadiene-styrene, and crumb rubber. The performance of modified binders with recycled polymers was compared with
that of conventional bitumen and of a commercial modified binder (Styrelf) (Costa et al. 2013).

Reclaimed polyethylene (PE) is recovered from low-density domestic waste PE carry bags. An 80/100-paving grade asphalt
was blended with different PE ratios (10%, 7.5%, 5.0%, and 2.5% by asphalt weight). The blends were tested using Hamburg
wheel track tests, resilient modulus tests, indirect tensile tests, and unconfined dynamic creep tests. Test result analysis showed
that the PE-modified asphalt mixture exhibited better performance characteristics than a conventional mixture. Including 5 weight
percent PE in the asphalt mixture can reduce temperature susceptibility and rutting potential (Punith and Veeraragavan 2007).

EVA is a plastomer, which is a copolymer obtained by copolymerization of ethylene and vinyl acetate. Though it is a
potential modifier, problems of phase separation have been encountered attributable to the presence of two separate phases
of bitumen and polymer that are incompatible with each other. In an effort to ascertain the optimum blending requirements
for EVA, an 85100 binder was modified with varying percentages of EVA from 1% to 7%. Modification was carried out at
different combinations of mixing temperature, blending time, and shear rate and a total of 80 combinations were obtained.
Further, the paper evaluated the optimum modifier content for obtaining a homogenous blend that could be stable at high tem-
peratures. Physical and rheological properties of the modified binder were also evaluated and compared with the base binder.
It was found that temperature is the most critical parameter for EVA modification. Shear rate had minimum influence over
obtaining a storage-stable blend. Fluorescence microscopy showed a change in morphology as the modifier content increased,
which could be used to assess the optimum modifier content for modification. The rheological response of the modified binder
significantly improved. EVA modification was found to be best suited at high temperatures (Saboo 2015).

Reactive elastomeric terpolymers (RETs) can be used to minimize phase separation by forming polymer networks in
blends. An RET is functionalized with glycidyl methacrylate and can crosslink and/or chemically bond with asphalt mol-

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

44

ecules and functional groups to improve the functional performance of the asphalt. The functional groups improve the rheo-
logical performance of the asphalt as demonstrated using Superpave tests. Low-temperature PG cracking resistance and mass
loss before and after aging of different types of modified asphalts were assessed. The results show that an RET modifier can
significantly improve high-temperature stability and low-temperature crack resistance of asphalt (Cao 2015). The viscosity
functions of several polymer-modified asphalts were studied at different temperatures in steady-state rate sweep tests. The
materials were obtained by mixing different base asphalts with either SBS, EVA, or RET. In the presence of SBS or EVA, at
certain temperatures, the viscosity curves exhibit a Newtonian behavior at low shear rates, followed by two distinct shear-
thinning phenomena. In some cases, the first shear-thinning is preceded by a small shear-thickening region. Similar phenom-
ena are not present in the viscosity curves of the RET-modified asphalt and can be related to a temporary nature of the physical
polymer network (Polacco et al. 2004).

A Turkish bitumen was modified with RET, EVA, and SBS polymers. Penetration, penetration index, softening point,
ductility, and percent elastic recovery tests were performed with the modified bitumen and raw bitumen. The samples of raw
bitumen and modified bitumen with 2% RET, 1% SBS, and 1% EVA were investigated by means of IR spectroscopy (FTIR)
and thermogravimetric analysis DTA (TGA/DTA). The penetration and ductility values of the modified bitumen decreased
while the penetration index, softening point, and percent elastic recovery increased (Keyf 2015).

The use of ground poly (ethylene terephthalate) (PET) particles in asphalt may provide an environmentally friendly solution
for the disposal of large quantities of PET waste. The performance of PET as a modifier for asphalt binders was evaluated using
rheological and viscosity properties. Tests were performed on the unaged and RTFO aged modified binders with recycled PET
particles at contents of 5%, 10%, and 15% by weight of the binders. The addition of recycled PET increased the high-temperature
performance. A higher-temperature performance grade was achieved by adding 10% PET. The viscosity and resulting work-
ability of the modified binders were not adversely affected for the amounts of PET studied (Shen et al. 2016a).

PET-modified HMA cements were prepared using either a dry or wet process; both wet and dry process mixtures contained
10% PET by weight of the base asphalt. Mixture performance tests were performed using an asphalt pavement analyzer (APA)
and a retrofitted APA Hamburg test to determine rutting resistance. Moisture susceptibility was estimated by an indirect
tensile strength test, and dynamic modulus (E*) was deteremined by asphalt mixture performance tester (AMPT). The wet
process mixture exhibited better rutting resistance and a higher tensile strength ratio (TSR) than the control. The dry process
mixture exhibited better resistance to permanent moisture damage in APA Hamburg testing and also exhibited a higher TSR
than the control. The modified mixtures exhibited lower E* and higher phase angles than the control (Shen et al. 2016a).

Crumb Rubber (GTR)

With scrap tire stockpiles continually growing, the federal government pushed state agencies and private businesses to develop
environmentally friendly ways to dispose of tire waste. Currently, only two states mandate the use of GTR. One way to solve
this disposal problem is by grinding or breaking tire rubber into small crumb-like particles to be used in HMA pavements.
This crumb rubber material, also known as crumb rubber modifier (CRM), can be blended with HMA mixtures by either a
wet process or a dry process (Lo Presti and Airey 2013). In the wet process, also known as the MacDonald process, the rubber
is melted and blended with the asphalt binder, whereas in the dry process the CRM is added as an aggregate to substitute
for a small portion of the fine aggregate. Though dry addition of GTR has had only limited success, recent efforts have been
employed to recycle GTR by dry addition in the HMA mixing process using additives and processing aids (Baumgardner et
al. 2012).

Chemical analysis of the binderCRM blends is complicated by the insolubility of the GTR in the binder. In the MacDonald
process, two factors are critical: (1) the development of performance-related properties and (2) binder compatibility or stor-
age stability; the performance-related properties develop early in the process but compatibility may require a few hours to
stabilize. Mixing binder with CRM at a high temperature results in a mixture of swollen rubber particles and binder matrix
containing the soluble components in the CRM. The swelling and degradation (devulcanization and depolymerization) of
the rubber particles leads to improvement in the properties of the binder matrix. However, the swelling and degradation is
highly dependent on CRM variables, the binder source, and the mixing conditions (Shen et al. 2009). The surface area and
the particle size of the GTR are critical variables in the blending process. CRM is produced by one of two processes, ambient
grinding or cryogenic fracture. Ambient grinding operates at room temperature and tears the tire carcasses apart to process
particles with a very porous structure. Cryogenic fracture is conducted at temperatures below the glass transition temperature
of the rubber (liquid nitrogen) so that the rubber shatters like breaking glass to process particles with smooth surfaces. The
difference in the microstructure influences the properties of the CRM binders; a higher PG at high temperature was observed

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 45

with ambient modified binder than that observed with cryogenic modified binders under the same condition (Shen et al. 2009).
A significant analysis of trends reported in the literature regarding GTR reaction time and attempts to dissolve it in asphalt is
available (Bahia 2011). A dissertation titled Characterization and Implementation of Ground Tire Rubber as Post-Consumer
Polymers for Asphalt Concrete gives a comprehensive study of all the variables involved in modifying asphalt with GTR
(Baumgardner 2015).

BIOBINDERS

Bio-based alternatives, which are being developed across the industry in various countries, could be a solution to reduce the
asphalt industrys dependence on petroleum resources. In addition to efforts in providing alternative binders, a trend toward
more sustainable pavements has led the pavement industry to place more emphasis on the application of technology to reduce
carbon footprints of pavements, including the use of warm-mix asphalt (WMA), half-warm-mix asphalt, and cold-mix asphalt
to reduce fuel consumption and CO2 production (Fini et al. 2016).

Bio-oil is derived from non-petroleum-based renewable resources such as woody biomass, waste oil, and animal manure;
in addition to bio-oils, efforts have been made to convert these waste materials to liquid fuel utilizing different methods such
as pyrolysis, fast pyrolysis, and gasification (Mohan et al. 2006). Among commonly used resources to produce bio-modifiers
are woody biomass that is also a source for biofuel. Oasmaa et al. (2010) used different wood-based feedstocks and agri-
cultural residues to produce biofuel; their study showed that although the liquid yield from the process was relatively high,
the accompanying gas component was low compared with that using other agricultural residues, making wood feedstock a
promising source for bio-oil production. According to the Energy Information Administration, nearly 9% of all energy con-
sumed in the United States is renewable, with 49% of it being derived from biomass. It can be noted that the main focus on
the conversion of biomass has been to produce fuel from various sources such as sawdust and cottonseed cake (Ozbay et al.
2006; Salehi et al. 2011; Klabunde and Shrestha 2014).

Seidel and Haddock (2012) derived bio-oil from soy fatty acids (SFA); they modified four Strategic Highway Research Program
binders and one recycled asphalt binder with 1% and 3% SFA to study the effect of introducing SFA on high-temperature proper-
ties of asphalt. Their study concluded that the inclusion of SFA could facilitate reduction of mixing and compaction temperatures.

In addition, several studies evaluated the merits of utilizing food crops to produce biofuel and bio-modifiers (bio-oils/
biobinders); however, in order to not create strains and competing demands, it is necessary to look into non-food-crop raw
materials. Raouf and Williams (2010) investigated the physical and chemical properties of bio-oils derived from oakwood,
corn stover, and switch grass in order to determine their applicability as a bio-modifier for pavement application. They con-
cluded that bio-modifiers could be used as a full or partial replacement for asphalt binder; their study further showed that
addition of their bio-oils to the asphalt binder increased the stiffness and high-temperature properties of base asphalt binders.

Chailleux et al. (2012) performed research on bio-oil derived from microalgae. Their chemical and rheological character-
ization showed that microalgae-based bio-oil has similar temperature dependence to asphalt binders and could be a promising
candidate for use in asphalt. With the need for advancing waste management practices, it would be much more advantageous
to determine the feasibility of producing bio-modifiers from certain waste products in order to further help environmental
practices as well as decrease the cost of raw materials to produce biobinders (Wen et al. 2013). Wen at al. evaluated the feasibil-
ity of using waste cooking oil to generate a bio-based asphalt. Binder studies utilizing 0%, 10%, 30%, and 60% waste cooking
oil showed an increased susceptibility to fatigue and rutting. Mixture tests also showed a reduction in dynamic modulus and
thermal cracking (Wen et al. 2013).

Recently, several researchers have shown that using a bio-modifier along with the petroleum-based asphalt could produce
a bio-modified binder with enhanced performance (Williams et al. 2009; Fini et al. 2012). Although there have been several
studies on the effect of introduction of various bio-modifiers to asphalt binders, their variational impact on the physicochemi-
cal characteristics of the base asphalt before and after oxidative aging has not been fully studied. It is important to evaluate
several bio-oils derived from different raw materials in order to determine the merits of their application while conducting
a comparative study. The effects of introduction of four different bio-modifiers (biobinders) on the rheological and chemi-
cal properties of a selected asphalt binder (PG 64-22) before and after oxidative aging have been published (Fini et al. 2011,
2016). Overall, bio-modifiers were found to be significantly different in terms of their aging characteristics. Accordingly, their
surface and rheological properties were found to be ranked differently before and after aging when compared with those of a
control asphalt binder. The results showed that the BB from swine manure is less susceptible to aging relative to plant-based

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

46

bio-oils. This can be further attributed to the chemical structure and the high lipid contents of the BB from swine manure,
making it less affected by oxidative aging (Fini et al. 2016).

Fini et al. (2011) synthesized a bio-oil from swine manure. They further studied the effect of introduction of the bio-oil to
a control asphalt at 2%, 5%, and 10% (by weight of base asphalt). Their rheological characterizations showed a clear trend in
improving asphalt workability and low-temperature properties resulting from the introduction of the BB from swine manure.
They further examined the merits of adding their BB (5%) in several asphalt mixtures along with various RAP and recycled
asphalt shingles (RAS) percentages; they showed the introduction of their BB was effective in compensating the stiffening
effect of RAP and RAS, resulting in more workable mixtures while improving the mixtures low-temperature cracking prop-
erties (Fini et al. 2011).

Yang et al. (2013) evaluated the use of asphalt binder modified by bio-oils generated from wood waste. Samples were
blended at 5% and 10% original bio-oil, dewatered bio-oil, and polymer-modified bio-oil, by weight of the base binder. The
results concluded that the addition of bio-oil can reduce the mixing temperature and improve high-temperature performance
while compromising low-temperature performance. Another major concern noted by the authors is the aging susceptibility
of the bio-oils.

Williams and McCready (2008) characterize the rheological properties and aging mechanism of asphalt binders blended
with high percentages of biobinders using FTIR. The petroleum asphalt was partially replaced by the biobinders at fractions
of 30% and 70% by weight. Rotational viscometer and dynamic shear rheometer tests were conducted for the rheological
properties. Loss of volatiles was obtained from the RTFO test, whereas the oxidation was investigated by an FTIR test. The
rheological results showed that the bioblended asphalt binders exhibit different rheological properties as compared with the
control asphalt binder before and after the RTFO aging. The mass loss test showed that biobinders had a much greater loss
of volatiles than the control asphalt binder. FTIR spectra analysis confirmed that additional C=C, C-O, C=O, and OH bonds
were generated during the aging. Further chemical analysis revealed that the aging of a biobinder can be attributed to three
processes: the loss of volatiles, dehydrogenation/condensation that forms higher-MW molecules such as asphaltenes, and the
oxidation to produce carboxylic acids, alcohols, and esters (Yang et al. 2015).

Morphological features and structural characteristics of asphalt binders are strongly affected by factors such as aging,
which can alter the performance of petroleum-based asphalt binders. Biobinder (BB), a newly produced amide-enriched
bio-adhesive obtained from biomass, has been found to be promising for reducing the negative effects that aging can have on
molecular conformation. Doping of BB into a commonly used petroleum-based asphalt binder creates a bio-modified binder
that has experimentally performed especially well at low temperatures. This improvement can be attributed to the effect
of modifier fragments, which contain high concentrations of amide functional groups, on the stacking of asphaltenes.
Disturbing the interactions alters the stacking distance, the corresponding binding energy, and ultimately the extent of
clustering of asphaltene units. Any change in the clustering of asphaltene affects the rheology and morphological proper-
ties of asphalt, which in turn alters the asphalts performance, including but not limited to its resistance to fatigue and low-
temperature cracking (Mousavi et al. 2016).

OTHER NONBITUMINOUS MODIFIERS

Low-MW modifiers are used for the specific purpose of improving the asphalt binders performance in a given environment.
These materials are often proprietary, but they can be classified by their intended application; that is, antioxidants, hydrocar-
bon supplements, antistripping agents, and stiffening agents.

Antioxidants

Oxidation is the primary cause of long-term aging in asphalt pavements. As a pavement oxidizes, it stiffens and can eventu-
ally crack. The use of an antioxidant as a performance enhancer in an asphalt binder could delay aging and thus increase the
life of an asphalt pavement. Experience with antioxidants to date is limited. Laboratory evaluations have been conducted but
there is a dearth of adequate field experience and validation. Some of the most recent efforts to use antioxidants are cited here.

Investigating the cracking potential in asphalt binders containing various antioxidant levels through a rheological test
series suggests that there can be a significant reduction in asphalt pavement cracking through antioxidant use. Laboratory-
prepared mixtures of asphalt containing antioxidants were subjected to short-term and long-term aging simulation. The

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 47

samples were evaluated using dynamic modulus tests and creep tests in the indirect tensile mode. Testing and analysis results
indicate a significant reduction in asphalt mixture cracking resistance through asphalt aging. Predicted cracking susceptibility
was substantially lower through antioxidant use (Apeagyeai et al. 2008).

Lignin is a readily available, well-studied antioxidant. Four lignin-containing coproducts were combined with four asphalt
binders in varying amounts to discover the optimum amount of coproduct that would provide the greatest benefit to the asphalt
binders. The asphalt-coproduct blends were evaluated according to Superpave specifications and performance graded on a
continuous scale. The data indicate a stiffening effect of the binder caused by the addition of coproduct; the more coproduct
was added, the greater the stiffening. Binder stiffening benefits high-temperature properties, while the low-temperature
binder properties are negatively affected. However, the low-temperature stiffening effects are small. Instead of acting purely
as a filler and shifting the use temperature range, the lignin has an overall effect of widening the temperature range of the
binders. Testing reveals that the lignin in the coproducts benefits the intermediate- and low-temperature properties of the
binders. Oxidative aging products were evaluated and some antioxidant effects were noticed (Williams and McCready 2008).

The oxidation mechanism in polymer-modified bitumen tends to alter physical and chemical properties. Dessouky et al.
(2015) evaluated five antioxidants to examine their potential to mitigate polymer-modified bitumen oxidation. Styrene copo-
lymers and hindered phenol additives were evaluated at low, intermediate, and high temperatures using aging indices. The
styrene copolymers were effective at high temperatures while the hindered phenols were effective at low and intermediate
temperatures. Two blends of both additives were introduced to cover a wide spectrum of temperatures. Frequency and tem-
perature sweep testing suggested that the blend improved the recoverable imposed energy at a wide range of temperatures/
frequencies. The blends have also improved resistance to rutting and moisture susceptibility for asphalt mixes.

A further study evaluated the effect of antioxidants on retarding of asphalt mixture aging using the GPC technique. An
antioxidant, hydrated lime, and linear low-density polyethylene were used when preparing the dense-graded asphalt and
the stone-mastic asphalt mixes, and the effect of aging retardation was evaluated using GPC. The asphalt mixes were aged
artificially in the laboratory in two stages: short-term aging for 1 hour at 160C (170C) and long-term aging for 164 hours
at 76C. It was found that the antioxidant and hydrated lime were effective on retarding aging of asphalt mixtures when the
large molecular size ratio was compared before and after aging treatment. The estimated viscosity levels of antioxidant- and
hydrated-lime-added asphalt mixtures were found be reduced by 27% of normal dense-graded asphalt mix after long-term
aging. The normal and LLDPE-modified stone-mastic asphalt were found to show even further age-retarding effects (by 42%)
without any antioxidant (Kwon et al. 2016).

Antistripping Agents

Antistripping agents are used to minimize or eliminate stripping of asphalt cement from the aggregate in HMA mixtures.
Both liquid antistripping additives and lime additives are used to resist stripping. Most liquid antistripping agents are surface
active agents that, when mixed with asphalt cement, reduce surface tension and, therefore, promote increased adhesion to the
aggregate. The chemical composition of most commercially produced antistripping agents is proprietary. However, the major-
ity of antistripping agents currently in use are chemical compounds that contain amines. Many antistripping agents have been
used in asphalt mixtures in the past, including amidoamines, imidazolines, polyamines, hydrated lime, organo-metallics, and
acids. These antistripping agents must be heat stable; that is, they will not lose their effectiveness when the modified asphalt
cement is stored at high temperatures for a prolonged period of time.

Of these products, the amines and hydrated lime have been used most commonly. In all cases, the purpose of these products
is to inoculate the mixture against moisture damage, often called stripping. Many liquid antistripping compounds have an
objectionable odor. New formulations are less objectionable, but are typically more expensive. Considerable research is ongo-
ing among various industry groups to develop products that promote adhesion between the asphalt binder and the aggregate
in the presence of moisture.

Since the late 1970s much research has been done to better understand the stripping phenomenon in asphalt mixtures. As a
result, there have been changes in both materials and technology over the past 30 years to improve asphalt mixtures resistance to
moisture damage and the ability to test for performance under adverse moisture conditions. Because of changes in materials and
technologies related to antistripping agents, a research study was conducted to evaluate the effectiveness of current antistripping
agents used in hot-mix asphalt pavements. The objectives were to construct a field test section that used three different antistrip-
ping agents in a conventional Superpave surface mixture and conduct a series of laboratory performance test comparisons using
different aging periods to make long-term comparisons of the effectiveness of hydrated lime, liquid additive, and warm-mix

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

48

asphalt antistripping additives. Sets of zero, one, five, and 10 freezethaw cycles were used for a portion of the research study.
The results showed that hydrated lime had the highest tensile strength and highest TSR values and was the only additive treat-
ment to meet the minimum of 80% TSR for all freezethaw cycle combinations. Both five and 10 freezethaw cycles were sig-
nificantly more discriminating for moisture susceptibility than one freezethaw cycle alone. Warm-mix asphalt treated mixtures
produced low initial tensile strengths, but the strength of these mixtures improved with time (Watson et al. 2013).

An important material property that influences the performance of an asphalt mixture is the surface free energy of the
asphalt binder and the aggregate. Surface free energy governs the adhesive bond strength between the asphalt binder and the
aggregate as well as the cohesive bond strength of the asphalt binder. These bond energies in turn influence the resistance
of the asphalt mixture to distresses such as fatigue cracking and moisture-induced damage. Asphalt binders undergo several
types of engineering and natural modifications that influence their chemical and mechanical properties. Three common exam-
ples of modifications are the addition of polymers, addition of additives (e.g., antistripping agents), and oxidative aging of the
asphalt binder. Bhasin et al. (2007) conducted a study examining the effect of different types of modifications on the surface
free energy components of the asphalt binder. The change in surface free energy was used to calculate parameters related to
the performance of the asphalt mixtures. Results from this study demonstrate that the magnitude and nature of change to the
surface free energy and concomitant performance-related parameters varied significantly among different asphalt binders.

Adding liquid antistripping agents typically reduced the surface free energy and consequently the work of cohesion of the
asphalt binders. This modification can indirectly improve fracture resistance by promoting better adhesion between the fine
aggregate particles and the binder during the mixing and compaction process. Use of liquid antistripping agents either improved
or did not significantly change the moisture resistance of the asphalt binder with the selected aggregates (gauged using the
parameter elastic recovery). The liquid antistripping agents from the two sources demonstrated different levels of changes in
moisture resistance when used with the same combination of asphalt binder and aggregate. In most cases, long-term aging
reduced the work of cohesion and thus indicated lower fracture resistance of the aged binder. In the case of one unmodified
binder and one modified binder, the work of cohesion increased after long-term aging. At that time, asphalt binders from one
source demonstrated a decrease in moisture sensitivity, while asphalt binders from the other source demonstrated an increase or
no change in moisture sensitivity with the two aggregates used in this study. The difference in the behavior of the two asphalt
binders is attributed to the influence of aging on the magnitudes of the polar functional groups (Bhasin et al. 2007).

Stiffening Agents

The modification of asphalt binders to improve performance properties has grown significantly since the implementation of
the SHRP binder specifications. The use of polymers and crumb rubber modifiers has increased. To compensate for the loss
of stiffness at high temperature ranges, stiffening agents such as polyphosphoric acid and gilsonite are added.

Polyphosphoric Acid

Several highway agencies have been concerned about the performance characteristics of PPA modification and possible nega-
tive interactions with other mix components, such as lime and liquid antistripping agents. A workshop on Polyphosphoric
Acid Modification of Asphalt Binders was held in April 2009 in an attempt to pull together the facts about PPA-modified
asphalt and performance.

The PPA workshop covered extensive laboratory and field evaluations on the use of PPA as a modifier for asphalt binders.
The following points were made by various workshop participants during the discussions:

1. The stiffening effect of PPA on the binder is crude source dependent with anywhere from 0.5% to more than 3% needed
to increase the binder grade.

2. PPA works as a stiffener and crosslinker when used with polymers such as SBS and ethylene terpolymers (e.g., Elvaloy).

3. PPA can significantly improve the delayed elastic response of the polymer-modified binder.

4. There is some indication that hydrated lime can somewhat reduce the stiffening effect of PPA but the increased stiffen-
ing from the lime outweighs any loss.

5. Limestone aggregate could not reverse or reduce the stiffening effect of PPA on the binder (DAngelo 2012).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 49

PPA has been used in 3.5% to 14% of the asphalt placed in the United States over the past 10 years. This represents up to
400 million tons of hot mix (Fee et al. 2010). Experienced industry practitioners have found that the addition of small amounts
(about 0.5%) of PPA to polymer-modified binders improved both their handling and performance. When used with styrene-
butadiene-styrene polymers it enables suppliers to achieve higher Superpave PG while improving mixing and compaction
characteristics. With ethylene terpolymers, PPA catalyzes the reactivity of the glycidyl methacrylate groups. Both types of
polymer-modified binders have shown the addition of PPA to increase rut resistance of the binder. More recently, the increas-
ing popularity of PPA has led to its use as a partial replacement for polymer modification (Arnold et al. 2012).

Used in conjunction with polymers, PPA enables suppliers to achieve performance grades that can be handled, mixed, and
compacted at reasonable temperatures. Both non-polymer-modified and polymer-modified plus PPA-modified binders have
shown improved rutting resistance with the addition of PPA. Though successful PPA modification of asphalt binders has a long
track record, its use is often debated, sometimes to the point that PPA-modified asphalt binders have been banned. Opinion
has it that such actions result from misinformation as well as lack of understanding of the benefits of PPA as an available tool
to improve the performance of asphalt binders. Although PPA is banned in some states, others continue its use without issue.
Baumgardner (2010) discusses common asphalt binder specifications, the relationship of asphalt binder chemical composition to
PG properties of asphalt binders, necessary use of PPA modification to meet specifications, the relationship of asphalt composi-
tion to PPA loadings necessary to achieve desired properties, expected PG enhancements of PPA-modified asphalt binders, and
effects of using PPA in non-polymer-modified and polymer-modified asphalt binders to meet current PG specifications.

The stiffening effect of phosphoric acid was found to be dependent on the particular asphalt being modified. Asphalt
binders from eight different sources were tested: AAD-1, AAK-1, AAM-1, ABM-1, two asphalts from Venezuela provided
by Citgo (a 60% Bachequero and a 94% Bachequero), an asphalt from BP Whiting Refinery, and an asphalt from Holly Cor-
poration. Of these binders, AAK-1 (Boscan) exhibited the greatest reactivity to phosphoric acid, whereas ABM-1 (California
Valley) was the least reactive and showed only a very slight increase in stiffness even at high dosage levels (Arnold et al. 2012).
The reactivity appears to be related to the effect of asphaltene dispersion. Asphalts with highly dispersed asphaltenes (sol
asphalts) are strongly impacted; asphalts with more condensed asphaltenes (asphalt gels) show little reaction.

As with all other components of the mix, testing is required to demonstrate the performance of PPA with each formula-
tion of asphalt and aggregate, together with polymer, antistripping agents, and other additives that may be used. Results of
the following tests are presented: dynamic shear rheometer, Hamburg, Lottman, and multiple stress creep recovery tests on a
matrix of a common asphalt with aggregate, three antistripping agents, two types of polymers, and PPA. In cases supported
by laboratory data for the materials tested, the performance of PPA-modified asphalt can be improved with the addition of
antistripping agents such as a phosphate esters, a particular polyamine compound, and hydrated lime. These findings hold true
for cases where modification includes the use of polymers: styrene-butadiene-styrene and Elvaloy (Fee et al. 2010).

Gilsonite

In tropical countries, roads built with asphalt layers must be made with bituminous mixtures containing asphalt that is rea-
sonably stiff, to increase resistance against permanent deformations; that is, rutting. Gilsonite (10 weight percent) modified
HMAs were prepared using either wet or dry processes. Gilsonite increases stiffness and improves the performance grade of
a virgin binder at high temperatures of service. However, at low temperatures, the mixture could experience embrittlement,
thus decreasing its resistance to low-temperature fatigue cracking (Quintana et al. 2016).

Sixty to 70 penetration-grade asphalt cement and Iranian natural bitumen (gilsonite) were subjected to physical and per-
formance tests. Physical and conventional tests were conducted on modified and unmodified bitumen (penetration, softening
point, ductility, and viscosity). Performance tests, including Marshall stability, indirect tensile strength, moisture suscepti-
bility, resilient modulus, and rutting resistance, were conducted on unmodified and modified stone-mastic asphalt mixtures.
Results of conducted tests on the modified asphalt binders with 5%, 10%, and 15% gilsonite show that using gilsonite
improves performance of stone-mastic asphalt mixtures (Babagoli et al. 2015).

Gilsonite was added to the biobinder, and rheological characterization of each bio-modified gilsonite (BMG) sample was
then conducted to investigate its performance. Results indicate that BMG samples show improved high-temperature perfor-
mance as well as comparable low-temperature performance and adhesion compared with PG 64-22. In some cases, BMG
samples show even better performance at low temperatures as evidenced by higher m-values and lower creep stiffness when
compared with PG 64-22. Specifically, 30% BMG had the most effective low-temperature properties (12C) among samples
studied. Furthermore, FTIR spectra were collected to compare the chemical functional groups of PG 64-22 with those of

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

50

BMG. Bio-modified gilsonite can be a promising candidate for use in asphalt mixtures because of its enhanced rheological
properties (Yaya et al. 2016).

A detailed study of the chemical composition of gilsonite has been reported. Gilsonite, naturally occurring asphaltic
bitumen, consists of a complex mixture of organic compounds. The gilsonite was characterized by elemental analysis to
determine the concentrations of carbon (C), hydrogen (H), nitrogen (N), sulfur (S), and oxygen (O), by FTIR for compara-
tive analysis of the chemical structures; by nuclear magnetic resonance spectroscopy of hydrogen (1H NMR) to ascertain the
aliphatic and aromatic hydrogen fractions; and by thin layer chromatography-flame ionization detection (Iatroscan TLC-FID)
to quantify saturated and aromatic hydrocarbons, and resin/asphaltene fractions (Nciri et al. 2014).

Asphalt Recycling

Asphalt recycling has become an important instrument used to minimize production costs of new pavements as well as to
mitigate their effects on the environment. Furthermore, stricter environmental regulations and depleting resources have led
to the exploitation of recycled materials by promoting the application of higher RAP and RAS in new mixtures.

Some of the benefits of utilizing recycled materials include the conservation of nonrenewable natural resources such as virgin
aggregates and asphalt binder, reduction in the amount of construction debris disposed of in landfills, decreased variability in
material expenditures, and potential reduction of the overall life-cycle cost. Recycling also helps to cut greenhouse gas emis-
sions by reducing the energy spent on extraction and processing of petroleum products and aggregates. Moreover, the increas-
ing price of asphalt binder along with more restrictive environmental legislation has forced highway agencies and contractors
to search for novel materials and construction techniques. Such efforts are aimed at fulfilling the current sustainability needs
without compromising the pavement quality and performance. There is, at this time, considerable emphasis on the use of RAP
as the preferred recycled material for highway construction as a result of its abundance and successful prior experiences. Collins
and Ciesielski (1994), in an NCHRP Synthesis of Highway Practice, noted that highway agencies have been proactive in the
recycling of reclaimed and by-product materials into construction materials, with RAP being the material most frequently used.
RAS, defined by AASHTO MP 23-14, Standard Specification for Use of Reclaimed Asphalt Shingles as an Additive in Hot-Mix
Asphalt (HMA), as any type of waste roofing asphalt shingles that have been processed into a recyclable product, have become
another promising recycling candidate. The utilization of RAP and RAS in modified asphalts has been reviewed extensively
(Stroup-Gardiner and Wattenberg-Komas 2013; Zhou et al. 2014; del Barco Carrin et al. 2015; Hoppe et al. 2015; Hossain et
al. 2015; Lee et al. 2015; Stroup-Gardiner 2016). A detailed discussion of these modifiers is beyond the scope of this synthesis.

Recycling Agents/Softening Agents

With increased interest in RAP and RAS, the use of recycling agents (RAs) is considered essential for softening and/or reju-
venating aged and stiff binders in RAS. Recycling agents are classified as two types: rejuvenating agents and softening agents.
Softening agents lower the viscosity of the aged binder, while rejuvenating agents are intended to restore the rheological and
chemical properties of the aged binder (Im et al. 2014). Most recycling agents are proprietary mixtures, but generic compounds
can be identified. Examples of softening agents include asphalt flux oil, lube stock, and slurry oil. Examples of rejuvenating
agents include lubricating and extender oils, which contain a high proportion of maltenes and low saturates contents that do
not react with asphaltenes (Im et al. 2014). The effectiveness of the rejuvenators depends on the mixture types and engineering
properties evaluated. Rheological properties may be improved, but reversing the chemical effects of extensive aging is unlikely.

The hot-mix recycling operation for bituminous mixes commonly uses a recycling agent to restore aged asphalt cement to
a rheological condition that resembles a virgin asphalt cement. In the laboratory this is done on the extracted asphalt cement,
thus ensuring thorough mixing. In the actual recycling operation, however, the RA is added to the material during the mixing
process and merely coats the salvaged asphalt concrete particles that are being recycled. A certain amount of time is required
for the RA to combine with the old asphalt and redistribute the older component throughout the modified binder. Carpenter
and Wolosick (1980) investigated the influence of this diffusion process on material behavior shows that the diffusion process
exerts a strong influence on the material properties required for long-term performance. Immediately after sample prepara-
tion, the mix may have high stiffness and excellent resistance to rutting. A week after preparation the stiffness had decreased
by a factor of two, and the resistance to rutting had decreased accordingly. This behavior is explained by a predicted rate of
the diffusion process. The redistribution of components is physically verified by extracting the outer and inner layers of the
modified asphalt concrete prepared in a simulated recycling operation and by comparing their consistency. Understanding this
phenomenon is critical for long-term performance predictions and in evaluating the effects of various laboratory conditioning
procedures (Carpenter and Wolosick 1980).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 51

Shen et al. (2007a) investigated the effects of an RA on properties of recycled asphalt binders and mixtures by adding
varying dosages of RAs. They found that the RAs significantly affected the properties of both recycled aged binders and their
mixtures. They also noted that the optimum percentages of the RAs could be obtained by satisfying SHRP specifications and
that they varied according to the neat binder properties. RAP rejuvenator-containing binder blending charts were developed.
Based on asphalt pavement analyzer evaluation of mixture rutting, as well as indirect tensile strength; mixtures incorporating
10% RAP; and RA mixtures were superior to HMA prepared with RAP blended with softer binders (Shen et al. 2007a, b).

Mogawer et al. (2013) queried if asphalt recycling agents can offset the stiffness attributed by the hardened binder from
RAP and RAS in mixtures that incorporate high RAP and RAS contents and if RAs can help the hardened binder from the
RAP/RAS comingle with the virgin binder. Overall, the results showed that asphalt RAs can mitigate the stiffness of the resul-
tant binder. The cracking characteristics of the mixture improved by the addition of the RAs; however, rutting and moisture
susceptibility were adversely impacted at the dosage and the testing conditions used.

The molecular composition of asphalt binders obtained from asphalt mixtures containing either RAP, RAS, or mixtures of the
two were characterized using gel permeation chromatography, the extent of aging using FTIR, and fracture resistance of laboratory-
produced mixtures using the Semi-Circular Bending test at intermediate temperature. Molecular fractionation through GPC of RAS
samples confirmed the presence of associated asphaltenes in greater concentrations than RAP pavement samples. High concentra-
tions of high-MW asphaltenes decrease the fracture resistance of asphalt mixtures. The use of recycling agentsCyclogen-L (a mix-
ture of naphthenes) and Hydrogreen (vegetable-derived oil)did not reduce the concentration of the highly associated asphaltenes,
and thus they failed to improve the cracking resistance of the asphalt mixtures evaluated in this study (Cooper et al. 2015).

REOB Asphalt Extenders/Softening Agents

Refined engine oil bottoms are obtained from the refining of recovered engine oil and have been used in the asphalt paving
industry since the 1980s. To obtain the low-temperature properties required in an asphalt binder, REOBs are typically used
from 3% to 10% by weight of the binder because REOB generally softens the base binder when used. When asphalt mixtures
are being designed with high RAP and RAS contents, resulting in stiffer binder blends because of the available recycled
asphalt binders, REOB may also be used as an RA in order to improve the cracking resistance properties of the asphalt mix-
ture given the stiffer composite binder blend (DeDene and You 2014).

FIGURE 19 Molecular weight (MW) distribution of molecular


species of extracted 70 PG binder containing 15% refined
engine oil bottoms and 5% recycled asphalt shingles
(70PG5P_B-RA1) (Source: Cooper et al. 2015).

The use of REOB in paving mixtures as a rejuvenating agent was examined. A series of mixtures were prepared containing
increasing REOB amounts: 5%, 10%, and 15%. The distribution of molecular species in REOB is concentrated in the domain
of asphalt maltenes. The higher molecular weight components of REOB contribute to the asphaltene domain, with a tail toward
the species of high-MW polymers (Figure 19). It appears that REOB acts only as a diluting agent, which is effective in extracting

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

52

the aged binder from RAS material. For example, the binder extracted from the mixture containing 15% REOB contained all
the asphalt binder present in RAS (100% RAS extraction; compared with only 36% availability for the mixture with no REOB).
Less RAS binder extraction was observed when the REOB addition to the mixtures dropped to 10% and 5%. However, the
REOB content did not improve the cracking resistance; Jc remained below the threshold limit of 0.5 kJ/m2. The extracting power
of REOB is reflected by the increase of the concentration of associated asphaltenes originating from RAS; that is, species with
apparent size of associated asphaltenes originating from RAS; that is, species exceeding apparent MW >40K Daltons (Figure
19). High-MW-associated asphaltenes are not significantly dissociated by adding rejuvenators. Use of rejuvenators negatively
impacted intermediate-temperature performance for the mixtures evaluated in this study (Cooper et al. 2015).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 53

CHAPTER FIVE

CASE EXAMPLES OF BINDER CHARACTERIZATION PRACTICES

INTRODUCTION

Based on the survey responses, a number of agencies indicated that they are in the process of implementing performance
specifications for asphalt mixtures or conducting research with performance tests to support their eventual implementation.

There were a number of reasons why these agencies were selected to serve as case examples. They represent a geographic
distribution that includes a wide range of climates and paving program sizes. The responses from the agency survey indicated
that Louisiana, Virginia, and Missouri in particular have made considerable advancements in binder evaluation.

This chapter discusses case examples from three agencies in the United States (Louisiana, Virginia, and Missouri DOTs) and
one Canadian Ministry of Transportation (Ontario). Their states of practice were captured through interviews with one or multiple
members of each agency to determine (1) how performance specifications for asphalt binders were developed and are used, (2) the
evaluation of performance testing, and (3) implementation efforts related to performance specifications for asphalt mixtures.

LOUISIANA

Lousiana Department of Transportation and Development (LDOTD) reported that asphalt mixture designs are approved in
the nine LDOTD district offices, and the responsibility to let pavement overlay and road contracts is also at the district level.
There are five major asphalt producers, and 15 to 20 contractors overall, that undertake state paving jobs. Four of the produc-
ers modify crude oil residuum and their products vary according to crude oil source. One of the producers blends asphalt
exclusively and ships a very consistent product that is more compatible with polymeric additives.

All contractor laboratories (in trailers or buildings) must be accredited through the Construction Materials Engineering
Council or AASHTO Materials Reference Laboratory. The state listing generally comprises 30 asphalt plants, and most contrac-
tors select a fully equipped central location to do their initial mix design testing, which does include some performance tests.
At this time, the state does not employ a consultant-based system and, as a result, LDOTD noted that it has been a challenge to
implement performance tests such as those conducted with AMPT equipment, owing to the cost and complexity of these tests.

Development and Use of Binder Specifications

The LDOTD Materials Laboratory has revised the binder specifications based on several years of experimentation with dif-
ferent testing techniques. The current protocol measures rotational viscosity, dynamic shear (DSR) G*/Sin Delta and phase
angle at 10 rd/s, flash point solubility, and polymer separation on the original binder. The rotational viscosity is measured to
determine product uniformity by the supplier. A binder having a rotational viscosity of 3.0 Pas or less will typically have
adequate mixing and pumping capabilities. The samples for the separation are prepared according to ASTM D7173. The
softening point is determined at the top and bottom of the tube per AASHTO T 53. The separation test is not required when
crumb rubber is used.

Tests on RTFO residue include mass change, DSR G*/Sin Delta, elastic recovery, ductility, and multiple stress creep recov-
ery. Further testing of PAV residue is completed using DSR and bending beam. Issues with the force ductility test have led to
the replacement of this test with an initial DSR phase angle maximum of 75@ 76C for polymer-modified asphalts. The force
ductility tests indicated the presence but not the quality of the polymer modifiers. All the base asphalts used in Louisiana meet
PG 67-22 specifications. A maximum of 10% crumb rubber is allowed in PF 82-22m and PG 76-22m mixes. Samples meeting the
specifications listed in Table 13 for polymer-modified grades are verified for acceptance; true grading of the sample is not done.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

54

TABLE 13
LOUISIANA PERFORANCE GRADED ASPHALT CEMENTS
Property AASHTO PG 82-22rm PG 76-22m PG 70-22m PG 67-22 PG 58-28
Test Method Spec. Spec. Spec. Spec. Spec.
Tests on original binder:
Rotational viscosity @ 135C, T 316 3.0 3.0 3.0 3.0 3.0
Pas
Dynamic shear, 10 rad/s, T 315 1.00+ 1.00+ 1.00+ 1.00+ 1.00+
G*sin delta, kPa @ 82C @ 76C @ 70C @ 67C @ 58C
Dynamic shear, 10 rad/s, T 315 75
phase Aangle, @ 76C
Flash point, C T 48 232+ 232+ 232+ 232+ 232+
Solubility, % T 44 N/A 99.0+ 99.0+ 99.0+ 99.0+
Separation of polymer, 163C, ASTM 2- 2-
48 hours, degree C difference D7173
in R & B from top to bottom
AASHTO
T 53
Tests on rolling thin-film oven T 240
residue:
Mass change, % T 240 1.00- 1.00- 1.00- 1.00- 1.00-
Dynamic shear, 10 rad/s, T 315 2.20+ 2.20+ 2.20+
G*/sin delta, kPa @ 82C @ 67C @ 58C
Elastic recovery, 25C, 10 cm T 301 60+
elongation, %
Multiple stress creep recovery T 350 0.5- 2.0-
(MSCR), 67C, Jnr(3.2 kPa)
Multiple stress creep recovery T 350 Meets curve Meets curve
(MSCR), 67C, % recovery
(3.2 kPa)
Ductility, 25C, 5 cm/min, cm T 51 90+
Tests on pressure aging vessel R 28
residue:
Dynamic shear @ 26.5C, 10 T 315 5,000- 6,000- 6,000- 5,000- 5000-
rad/s, G* sin delta, kPa @ 19C
Bending beam creep stiffness, T 313 300- 300- 300- 300- 300-
S, MPa @ 12C. @ 18C
Bending beam creep slope, m T 313 0.300+ 0.300+ 0.300+ 0.300+ 0.300+
value @ 12C @ 18C
Source: Louisiana DOTD (2016).

The MSCR curves and Jnr limits as defined by AASHTO M 332 are effective monitors of binder quality. Some of the plant-
blending operators can meet the Jnr, but they have had issues meeting the curve for PG 70-22m and PG 76-22m. They try to
balance stiffening the binder enough to meet the initial DSR phase angle maximum of 75C (an indicator for the presence of
polymer) but not stiffen it so much as to cause issues with the recovery. For the PG 70-22m, the problem appears to be similar
in that the blenders are trying to meet the recovery curve by adding more polymer (latex, generally), which stiffens the asphalt
and raises the recovery target. Some are now using additives that appear to help, and using a different base asphalt appears to
make a difference. MSCR percent recovery identifies elastomeric polymer, but MSCR recovery curves were developed using
SBS modifiers. Additional curves for SBR, crumb rubber, and so forth are needed to specify a percent recovery for a specific
Jnr. These curves are even more important if the actual Jnr is used rather than a maximum for a grade.

GPC Analysis

The Louisiana Transportation Research Center (LTRC) is the development arm for new techniques. LTRC developed a
method using deconvoluted GPC chromatograms to determine polymer percent in modified asphalt and examine asphal-
tene/maltene ratios. A robust state-of-the art GPC system was installed in an LDOTD Materials Laboratory room dedicated

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 55

solely to preparation and analysis of samples received from Louisiana suppliers of paving asphalt materials. An effective
asphalt binder extraction methodology without affecting the binder properties was developed. Both virgin asphalt cements
and asphalt concrete cores from field projects are analyzed. Asphalts supplied by each individual producer yield specific
GPC curves that confirm the specified quantities of the polymer added and serve as standards for forensic analysis. Asphalt
binders extracted from field cores yield the curves characteristic of the asphalt producer. The library of GPC curves has
already been used effectively to identify the source of asphalts used in the field. Plans to require each project to submit an
acceptance core from the first lot are being considered. Future research on the application GPC to asphalt binder changes
after field aging is planned.

LDOTD noted that there have not been many instances of rutting in existing Superpave asphalt pavements, except for some
extraneous reasons, such as issues at the plant during production, issues in the field during construction, or material problems.
Therefore, it indicated that the control of moisture damage and fatigue were performance-based tests the agency considers
most important for predicting the pavement performance typical for its roadway system. All mixtures are required to contain
at least 0.6% antistripping agent, which is added at the plant.

A maximum of 20% RAP is allowed in the mixtures. Recycling agents may be added to mixtures as long as all the speci-
fications for the mixtures are satisfied. There is a moratorium on the use of RAS in Louisiana. Future research reported by
LDOTD also includes its involvement in a Pooled Fund Study TPF-5(294) being conducted by the LTRC that will evaluate
various fatigue tests, including identifying differences in how the results from these tests differ for mixtures produced with
higher RAP or RAS contents. Specifically, one objective of the research project is to establish mechanistic test criteria for
achieving durable flexible pavements made from both WMA and HMA mixtures that contain high RAP and/or RAS con-
tents. A second objective is to develop preliminary asphalt mixture specifications that incorporate the resulting mechanistic
test criteria to be tested on plant-produced specimens and/or roadway cores, as based on the results of the study. The testing
of plant-produced mixtures and roadway cores will facilitate the evaluation of the impacts of higher RAP and/or RAS per-
centages on the durability of the asphalt mixtures investigated as part of the study.

VIRGINIA

Virginia has the third largest state-maintained highway system in the United States, with more than 57,000 miles of roadways,
the vast majority of which are constructed using asphalt. Virginia Department of Transportation (VDOT) reported that the
approval of asphalt mixture designs occurs at the district level in each of nine districts. The responsibility to let pavement
contracts also exists within the district offices. There are 27 asphalt producers in Virginia. Most contractors design their mix-
tures at a central laboratory. Contractor labs either receive AASHTO accreditation or are certified through VDOT inspection,
which uses a protocol similar to AASHTOs.

Development and Use of Binder Specifications

VDOT adopted Superpave binder performance grading in the mid-1990s and in 1997 began to using the Superpave mix design
system. Since adopting the PG system, no major changes have been made to VDOTs specifications.

In 2015, VDOT adopted the multiple stress creep recovery test. This test is more efficient than the Superpave elastic
recovery test, and makes identification of polymer modification in binders easier (Table 14). Virginia is identified as a PG
64 state, so all MSCR testing is conducted at 64C. The designations after the testing temperature (64C unless otherwise
specified) generally indicate the effective change in high temperature grade to accommodate expected traffic loading. These
letter designations are S for standard, H for heavy, V for very heavy, and E for extra heavy. The new MSCR binder
designations and their corresponding Superpave PG designations are listed in Table 1. Specific VDOT binder testing require-
ments are provided in Tables 15 and 16.

Asphalt pavements are subject to density requirements, operating within mix design tolerances for in-place density (as
measured by nuclear gauge), asphalt content, gradation, voids total in the mix, voids in the mineral aggregate, voids filled with
asphalt, and dust-to-asphalt-cement ratio. Asphalt content and gradation are checked for every 500 tons of mix. Volumetrics
are tested for every 1,000 tons of mix. Asphalt manufacturers typically take quality control (QC) samples of binder for every
batch, and VDOT does verification/independent assurance.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

56

TABLE 14
MULTIPLE STRESS CREEP AND RECOVERY VERSUS PERFORMANCE GRADE DESIGNATIONS
MSCR Designation Superpave PG Designation
PG 58S-28 PG 58-28
PG 64S-22 PG 64-22
PG 64H-22 PG 70-22
PG 64V-28 PG 70-28
PG 64E-22 PG 76-22
Source: Virginia DOT (2016).

TABLE 15
VDOT SPECIFICATIONS, REFERENCE AASHTO M-320
Property AASHTO PG 64E-22 PG6 4H-22 PG 64S-22 PG 58S-28
Test Method (PG 76-22) (PG 70-22) (PG 64-22) (PG 58-28)
Spec. Spec. Spec. Spec.
Tests on original binder:
Rotational viscosity @ 135C, Pas T 316 3.0 3.0 3.0 3.0
Dynamic shear, 10 rad/s, G*/sin delta, kPa T 315 1.00+ 1.00+ 1.00+ 1.00+
@ 76C @ 70C @ 64C @ 58C
Dynamic shear, 10 rad/s, phase angle, T 315
Flash Point, C T 48 230+ 230+ 230+ 230+
Solubility, % T 44 99.0+ 99.0+ 99.0+ 99.0+
Separation of polymer, 163C, 48 hours, degree C ASTM D7173 2- 2-
difference in R & B from top to bottom AASHTO T 53
Tests on rolling thin-film oven residue: T 240
Mass change, % T 240 1.00- 1.00- 1.00- 1.00-
Dynamic shear, 10 rad/s, T 315 2.20+ 2.20+ 2.20+ 2.20+
G*/sin delta, kPa @76 @70 @ 64C @ 58C
Elastic recovery, 25C, 10 cm elongation, % T 301
Ductility, 25C, 5 cm/min, cm T 51 90+
Tests on pressure aging vessel residue:
Dynamic shear, @ 26.5C, 10 rad/s, T 315 5000- 5000- 5000- 5000-
G* Sin Delta, kPa @ 31C @ 28C @ 25C @ 19C
Bending beam creep stiffness, S, MPa @ 12C. T 313 300- 300- 300- 300-
@ 18C
Bending beam creep slope, m value,@ 12C T 313 0.300+ 0.300+ 0.300+ 0.300+
@ 18C
Source: Virginia DOT (2016).

Binder Modification and/or Use of Reclaimed Material

VDOT permits up to 30% RAP in asphalt mixtures using a binder performance grade PG 64S-22 (PG 64-22) base. RAS
have a 5% allowance, but the binder must meet PG 64H-16 (PG 70-16) and must use PG 64S-22 (PG 64-22) as the base mix
binder grade. Rubber-modified binders are not prohibited by VDOT as long as they pass current binder specifications. There
is not a separate specification for rubber-modified binders, as they have received limited use. There are no specified polymer
modifiers; however, all modified binders are subject to the same binder testing (e.g., PG binder grading and MSCR testing).
In-line blending of polymers is very limited in Virginia and is subject to approval. Currently, most in-line blending is used
only for special projects.

VDOT does allow warm-mix asphalt to be used for paving projects. Although some projects use warm-mix additives to
produce WMA, many contractors also use additives or foaming technologies as a compaction aid for HMA or to place asphalt
at a production temperature between WMA and HMA. All additives are subject to product approval by VDOT.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 57

TABLE 16
VDOT SPECIFICATIONS, REFERENCE AASHTO M-332
Property AASHTO PG 64E-22 PG 64H-22 PG 64S-22 PG5 8S-28
Test Method (PG 76-22) (PG 70-22) (PG 64-22) (PG 58-28)
Spec. Spec. Spec. Spec.
Tests on original binder:
Rotational viscosity @ 135C, Pas T 316 3.0 3.0 3.0 3.0
Dynamic shear, 10 rad/s, G*/sin delta, kPa T 315 1.00+ 1.00+ 1.00+ 1.00+
@ 76C @ 70C @ 64C @ 58C
Dynamic shear, 10 rad/s, phase angle, T 315
Flash point, C T 48 230+ 230+ 230+ 230+
Solubility, % T 44 99.0+ 99.0+ 99.0+ 99.0+
Tests on rolling thin-film oven residue: T 240
Mass change, % T 240 1.00- 1.00- 1.00- 1.00-
Elastic recovery, 25C, 10 cm elongation, % T 301
Multiple stress creep recovery (MSCR), 64C, Jnr (3.2 kPa) T 350 @64C @ 64C @ 64C @ 58C
Multiple stress creep recovery (MSCR), 64C, T 350 > 29.371
% recovery (3.2 kPa)
Tests on pressure aging vessel residue: R 28
Dynamic shear, @ 26.5C, 10 rad/s, G* sin delta, kPa T 315 6,000 6,000 5,000- 5,000-
@ 19C
Bending beam creep stiffness, S, MPa @ 12C. T 313 300- 300- 300- 300-
@ 18C
Bending beam creep slope, m value @ 12C T 313 0.300+ 0.300+ 0.300+ 0.300+
@ 18C
Source: Virginia DOT (2016).

Specialty Testing

The Virginia Transportation Research Council is the research arm of VDOT. The council is often tasked with aiding the
VDOT Materials Office with evaluating potential changes to specifications and test procedures. It also leads research and
development projects that examine new test methodologies. Past and current research projects examined different nontradi-
tional techniques for testing asphalt binders. Among these are Fourier FTIR, GPC, and XRF.

In 2006, research was completed that used FTIR as a method to identify the presence of polymer modifiers in asphalt bind-
ers. This method was found to identify the presence of polymers with no false-positives and could be calibrated to quantify
the amount of modifier present based on polymer type and base binder source. Calibration, however, would require extensive
effort and was deemed to not be practical. Thus, FTIR remains more qualitative than quantitative for identifying polymer
modification. FTIR was also found to be useful in identifying whether asphalt binder solvents were completely removed from
the binder after recovery.

VDOTs use of GPC is primarily in the investigative stage. Its primary use, thus far, is to examine molecular changes
caused by binder oxidation. Potential future uses involve forensic analysis, investigation of polymer degradation with
age, and development or implementation of quantitative methods for determining polymer content. There is also interest
in utilizing XRF to determine binder composition, specifically to identify the presence of refined engine oil bottoms and
other additives.

MISSOURI

Missouri allows antistripping agents and wax-based additives (warm mix) in its mixture designs. Binder samples are taken
when they contain all of the additives, that is, just before the binder enters the mixing drum. Daily binder samples are required.
If the selected samples do not meet the standards, more extensive testing is employed before the material will be accepted.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

58

Missouri focuses on the end result. Randomly chosen field samples are checked using DSR to confirm the PG binder grade.
Acceptable tensile strength ratio data (quality control and quality assurance) are required on field samples for acceptance;
this is a check on the presence of antistripping agents. Occasionally the Missouri DOT observed that some antistripping agent
can affect the binder grade. The contractor may use whatever amount of antistripping agent is necessary to meet the TSR.

Missouri has one binder supplier that utilizes PPA. The supplier must indicate the use of PPA on the bill of lading and
advise the asphalt paving contractor of its presence so the PPA can be incorporated in the mix design. In a few jobs where the
asphalt paving contractor was not aware of the presence of PPA, the PPA neutralized the antistripping agent, which negatively
affected the field TSR. At least one antistripping agent that is compatible with PPA has been identified. The PPA also reacted
with the polymer and negatively affected the PG binder grade. To eliminate the possibility of further problems, the Missouri
DOT has developed a test method for determining the amount of PPA in PG-grade asphalt binders. PG-grade asphalt binder
is fused with sodium carbonate and dissolved in an acid solution. The solution is analyzed for phosphorous on an inductively
coupled plasma optical emission spectrophotometer and the phosphorus content is calculated as percent of PPA. This test is
used when PG binder grade is not confirmed through DSR testing.

Sustainability Efforts

Missouri has been incorporating RAP into asphalt cements since the 1990s. Beginning in 2013 95% of HMAs contained
recycled materials. An average of 23.8% RAP and 2.5% RAS is present in Superpave mixes. Stone-mastic asphalts have
been used on an experimental basis only. Contract-grade virgin binder may be used to blend up to 20 weight percent RAP
or up to 10 weight percent RAS. The contribution of the RAP to the binder content is estimated as an effective virgin binder
replacement percentage. If an RAP/RAS combination is present at less than 20 weight percent, the effective virgin binder
replacement estimated as [RAP + (2*RAS)] may be added to the contract grade virgin binder as long as the final binder blend
meets the designated PG specifications. If up to 40% RAP is blended, the PG of the virgin binder must be reduced one grade.
The standard PG 64-22 binder may be modified with recycling (softening) agents to produce a PG 58-28 binder grade before
the RAP addition. Specific blend charts are required for cements containing greater than 40 weight percent RAP. The incor-
poration of RAS began in 2009; initially up to 8% was added but the amount has now been reduced to 2%3%. If producers
use recycling agents, the dosage rates and the volumetrics must be designated in the mix design. Acceptance of the blended
binders depends on meeting the designated performance grade specifications.

Ground tire rubber may also be employed if the producers do terminal blending and provide an acceptable QC plan. Up
to 8% GTR may be added to PG 64-22 base binder to produce a PG 76-22 binder. All blends containing GTR include 4.5%
transpolyoctenamer rubber by weight of the GTR. The direct tension test is waived, but a separation test is performed in
accordance with ASTM D 5976. The number of Superpave mixes containing GTR was as high as 22% in 2011, but currently
approximately 10% of the designs include GTR.

ONTARIO

The Ontario Ministry of Transportation (MOT) is one of the largest in Canada, with more than 16,500 km (~10,300 m) of
roadways, the vast majority of which are constructed using asphalt. Ontario is Canadas most populous province by a large
margin, accounting for nearly 40% of all Canadians, and is the second-largest province in total area. The surrounding Great
Lakes greatly influence the climatic region of southern Ontario. Proximity to the Great Lakes gives some parts of southern
Ontario milder winters. The climate is similar to that of the inland Mid-Atlantic states and the Great Lakes portion of the
Midwestern United States. The region has warm to hot, humid summers and cold winters. The northernmost parts of Ontario,
primarily north of 50N, have a subarctic climate with long, severely cold winters and short, cool to warm summers with
dramatic temperature changes possible in all seasons. Temperatures of 40C (40F) are not uncommon; snowfall remains
on the ground for sometimes over half the year. Thus, MOT must cope with extreme environments in designing highways and
specifying the materials for their construction. Five MOT regional offices are responsible for the planning, design, construc-
tion, maintenance, operations, and management of the provincial highway network.

In the 1990s, the Strategic Highway Research Program Performance Graded Asphalt Cement specifications were imple-
mented with the expectation that identifying asphalt cements that meet the specifications would help to reduce pavement
cracking. Over the past 15 years, MOT has observed that some pavements experience more cracking within 2 to 3 years of
placement, whereas others experience little to no cracking in their early life. Through numerous trial contracts and exten-
sive research, MOT identified poor-quality asphalt cement as one of the primary causes for premature pavement cracking.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 59

Chemical analysis showed that the likely presence of waste engine oil residues, air-blown residues, and/or acids in some of
the asphalt cements may have contributed to excessive cracking.

The acceleration of the physical hardening phenomena using an isothermal conditioning is correlated to data from seven
trial sections and regular contracts (Evans et al. 2011). Tensile specimens are poured and conditioned for either 20 min or 72 h
at low temperatures before being subjected to a stress relaxation test at 10C. The extended conditioning period could more
than double the residual thermal stress at the end of the test. The one asphalt to show no physical hardening came from the
section that remained largely free of cracking. A second material that showed a moderate degree of physical hardening only
recently started to crack by an appreciable amount. In contrast, the remaining five materials significantly hardened during
the extended conditioning period and cracked prematurely and excessively in service. Materials with an unstable colloidal
structure were found most sensitive to physical hardening. The results of the study agree with earlier creep data obtained
according to an extended bending beam rheometer test method (Hesp et al. 2009). Hence, it is imperative that pavements be
designed with criteria that take physical hardening effects into account to limit premature and excessive performance failures.

Analysis of Binder Crack Resistance

In the past decade one lab noticed premature cracking (top-down mostly), starting in the wheel paths and then propagating
in a map cracking form (Tabib et al. 2015). The lab recovered the binder in several cases and noticed a great loss of low tem-
perature grade using an extended bending beam rheometer (ExBBR) test. This is an indication that the SHRP aging protocol
is not sufficient on the low temperature side. The research on binders extracted from poorly performing pavements led MOT
to focus on two of these test methods to better predict premature pavement cracking. Descriptions of these two tests follow.

ExBBR, which determines the low-temperature continuous performance grade of physically hardened asphalt cement
using different conditioning temperatures and times than those used in the AASHTO T 313 test method for bending beam
rheometer. The BBR method conditions samples for 1 h and tests them at a temperature 10C warmer than the low tempera-
ture grade (T). The ExBBR method conditions samples up to 72 h at 10C warmer (T +10C) and 20C warmer (T + 20C) to
determine, by extrapolation, the limiting temperature for each conditioning time and temperature.

Double-edge notched tension (DENT) is an indicator of asphalt cements resistance to ductile failure. The test is conducted
after intermediate thermal conditioning to determine the essential work of fracture, the plastic work of fracture, and an
approximate critical crack tip opening displacement at a specified temperature and rate of loading.

The ExBBR test is intended to predict reversible physical hardening that can cause premature low-temperature cracking.
The DENT test is used as a measure of asphalt cements elasticity, or ability to stretch and resist cracking at intermediate
temperatures. Because the tests look at different thermal states of the pavement, and strain tolerances change with tempera-
ture as a result of variations in hardening tendency, it is necessary to review both ExBBR and DENT results to accurately
predict the pavement cracking performance (Hesp et al. 2009; Shurvell et al. 2009). Hesp et al. (2009) found that factors such
as chemical and physical hardening cannot be captured by a single test at 15C and that the DENT test is intended to exclude
poor performers rather than to provide a perfect correlation with pavement cracking [5]. The criteria for passing these tests
has been incorporated into MOTs specification for performance graded asphalt cement acceptance (Table 17).

Ontario MOT requires the use of antistripping additives for northern Ontario and allows a limited quantity of PPA. Because
much PG XX-34 is used and some PG XX-40, MOT knows that softening agents (oils, etc.) are being used, but these are not
specified. The final product must comply with M320 specifications. Efforts to use XRF to detect PPA failed to give accurate
results owing to low dosage. MOT tried using FTIR to analyze for antistripping agents, but the low dosage is hard to detect.
Ongoing research includes the use of XRF and FTIR to quantify REOB, which currently is limited by the ash content speci-
fication. A modified PAV procedure is being evaluated, which would look at either reducing the asphalt film thickness or
increasing the oven aging time.

SUMMARY

The interviewees from DOTs of three states and one Canadian MOT consistently observed the following points:

The Strategic Highway Research Program Performance Graded Asphalt Cement specifications have deficiencies.
Estimation of low-temperature performance can be improved using combined ExBBR and DENT fatigue testing.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

60

The MSCR test is more efficient compared with the Superpave elastic recovery test, and makes identification of polymer
modification in binders easier.
Greater use of recycled materials (e.g., RAP, RAS, and crumb rubber) required adjusting the AC content to ensure suf-
ficient liquid in the mixes.
The interviewed agencies are moving toward incorporating performance testing for production acceptance; however,
the barriers to more widespread adoption of performance tests are the cost, manpower, and time required to run the
performance tests. Tests under extensive review include XRF, FTIR, and GPC.

TABLE 17
ONTARIO MINISTRY OF TRANSPORTATION SPECIFICATION FOR PERFORMANCE-GRADED ASPHALT CEMENT
ACCEPTANCE
Property and Attributes (Unit) Test Method Acceptable Major Rejectable PGAC Grade
Borderline
0.8 > 0.8 and 1.0 > 1.0 All PGAC except
PG 52-40 and PG
Ash content, % by mass of residue (%) LS-227 58-40
1.0 N/A > 1.0 PG 52-40 and PG
58-40
< -YY (low 3 to 6C warmer > 6C warmer
Low-temperature limiting grade (C) LS-308 PGAC grade) than low PGAC than low
grade PGAC grade
Grade loss (C) LS-308 06 68 >8 All PGAC
Nonrecoverable creep compliance at 3.2 kPa (Jnr3.2) (kPa-1 ) AASHTO T < 4.5 N/A 4.5 Grades except
PG 58-28 and PG
Average percent recovery at 3.2 kPa (R3.2) (%) 350 testing > the lesser of N/A the lesser of 52-34
conducted at [(29.371) [(29.371)
58C in (Jnr3.2)- (Jnr3.2)-
southern 0.2633] or 55 0.2633 -10] or
Ontario and 45
Percent difference in nonrecoverable creep compliance 52C in Testing carried out only for information purposes
between 0.1 kPa and 3.2 kPa, Jnrdiff (%) northern
Ontario
10.0 mm for < 6 and 4 mm < 4 mm
PGAC XX-28 for PGAC
XX-28
14.00 mm < 10 and 8 mm < 8 mm
CTODs, delta (mm) LS-299 for PGAC for PGAC
XX-34 XX-28
18.0 mm for < 14 and 12 < 12 mm
PGAC XX-40 mm for PGAC
XX-28
Source: Ontario MOT (2016).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 61

CHAPTER SIX

SURVEY RESULTS: CURRENT U.S. AND CANADIAN EXPERIENCE

This chapter presents the results of a survey of U.S. states and Canadian provinces and territories summarizing the criteria
that are used to establish relationships between binder chemistry and engineering properties of asphalt binders. See the appen-
dices for a copy of the questions in the survey, the tabulated survey responses, and a list of respondents.

An electronic survey was distributed in March 2016 to all 50 U.S. states and the District of Columbia and to 14 Canadian
provinces and territories. Responses were received from 45 states and the District of Columbia (a response rate of 88.2%).
Limited responses were received from eight Canadian provinces and territories. Because of the limited data on the Canadian
forms, these data were examined separately and published in the comments section of Appendix A only. The general survey
responses are summarized here; the percentages were calculated based on the U.S. responses. Since some respondents submit-
ted multiple answers to a given question, all percentages are based on the total of 46 state responses. The detailed summary
of the responses can be found in Appendix A.

The state departments of transportation do not monitor the sources of their asphalt binders. Only 13.3% of the respondents could
identify their sources. The questionnaire could also be directed to binder suppliers, since DOTs are not involved in production and
formulation of the binders they use. Similarly, most DOTs are not familiar with the process used to produce their binders.

Significant differences in asphalt binder properties over the past 10 years were reported by 42% of the respondents. Pos-
sible issues with changes in the testing requirements mean that most of the grades used today are different from those used 10
years ago, so direct comparison is difficult. Binders appear to be stiffer and exhibit poor low-temperature properties, and some
pavements aged prematurely. Although binders met specifications, experienced engineers could detect quarterly changes as
crude feeds change. Changes associated with seasonal or market fluctuations were reported by 38% of the respondents.

The physical properties measured to estimate changes in the crude stock or binder chemistry in the field are predominat-
edly rheological (Figure 20). The dominant analytical procedures used for binder testing are rheological; that is, the PG pro-
tocol (100% of respondents). FTIR is employed by 23.9% of the DOTs but this appears to be primarily used for research. The
third most common procedure was XRF (19.6%), which was employed by respondents incorporating PPA or REOB additives.
GPC is used routinely by only two DOTs (Figure 21).

FIGURE 20 Types of binder testing used to quantify binder properties (Source: Survey results).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

62

FIGURE 21 Rheological and analytical procedures employed (Source: Survey


results).

There was little interest in binder fractionation. Only 14 DOTs responded to this question and only three were working on new
separation procedures. Eleven DOTs stated that they do not fractionate. This issue is considered the purview of binder suppliers.

Binder aging characteristics are assessed routinely using the RTFO and PAV procedures defined by AASHTO M320 (100%
of respondents). Occasionally multiple PAVs (13.3%) may be employed. The use of Glover-Rowe parameters, delta Tc, fracture
energy, penetration index, and phase separation after oven aging is mentioned (Figure 22).

FIGURE 22 Procedures used routinely to estimate asphalt binder aging characteristics (Source:
Survey results).

A number of modifiers/additives are employed, including PPA (41.3% of respondents), antistripping agents (26.1%), and
softening agents (21.7%) (Figure 23). The use of the modifiers depends on suppliers reporting, so these data may not be
complete. The addition of compaction aids (waxes, wetting agents, antistripping agents, and rejuvenators) is paving related
and dictated by the job, aggregate, and plant. With the exemption of polymers and PPA, the additives in a binder are usually
not specified or quantified, because that is considered the responsibility of the supplier. Some concern over the excessive use
of PPA is reflected in the upper-limit specifications imposed by 12 DOTs. Comments on specific DOT procedures can be
found Appendix A. The compatibility of modifiers in the final mixes is not measured by more than 95% of the 40 respondents
replying to the questionnaire.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 63

The most common polymers used to prepare a polymer-modified binder are linear styrene-butadiene-styrene copolymers
(71.7%) and styrene-butadiene rubber (26.1%) (Figure 24). Crumb rubber or ground tire rubber is allowed by 31.4% of the
DOTs. The presence of polymers is controlled by an elastic recovery test. Frequently, the suppliers choose the type of polymer
used to meet the elastic recovery specification and the DOT imposed minimum concentration standard.

FIGURE 23 Modifiers/additives used in final binder preparation to achieve specific asphalt binder
grades (Source: Survey results).

FIGURE 24 Polymer modifiers employed (Source: Survey results).

Efforts to understand the impact of recycled materials by extracting binder from mixtures are focused on RAP (65.2%)
and RAS (47.8%). A limited number of DOTs (fewer than 10) are using recycling (rejuvenating) agents; the most common
materials employed are soft asphalt fluxes and abietic acid esters. The recycling agents appear to be used on an experimental
basis for specific projects, and nine DOTs expressly state that they do not use them.

A majority of the respondents (69%) did not consider the current PG criteria satisfactory in identifying compositional/
chemical properties of a given binder. PG criteria, which are physical measurements, currently appear to fail to capture some
poorly performing binders. To better identify compositional properties of polymer-modified asphalt, Illinois DOT specifies
PG+ tests (force ductility, elastic recovery, and separation of polymer). Louisiana and Virginia DOTs are employing GPC to

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

64

confirm the presence and concentration of polymers, but a quick, easy field identification of modifiers (polymer and rubber)
needs to be developed to determine what is contained in the mixes.

FIGURE 25 Recycled materials extracted from mixtures (Source: Survey results).

The respondents do not believe that current performance grade criteria satisfactorily identify compositional/chemical
properties. REOBs, PPA, and so forth can go undetected in the PG system because it is purely rheologically based. The jury is
very much out on how much any of these modifiers affect the long-term performance of a binder. Chemical analysis is required
to ensure that a binder has the specified composition. Virginia DOT considers this one potential benefit of running GPC as
well as employing XRF on binders. Several respondents expressed the opinion that the PG 64-22s have declined in quality,
based on the life of their asphalt pavements. The missing component of the PG criteria is long-term aging. A one-point aging
test such as PAV does not tell the whole story.

Procedures used to quantify the presence of modifiers/additives in final asphalt are surveyed. Polymer-modified binders
are checked using MSCR (30.4%), XRF (15.2%), viscosity (8.7%), and FTIR (6.5%). Softening agents are checked using
XRF. The blend compatibility with antistripping agents, polymers, and PPA is tested in a limited number of cases using mod-
ern variants of the cigar tube test. In one case evidence for incompatibility in blends containing PPA was reported. The PPA
neutralized the antistripping agent that negatively affected the field TSR. The PPA also reacted with the polymer, negatively
affecting the PG binder. A specification defining an upper limit on PPA is used by 12 DOTs.

Several laboratories are developing alternate techniques for estimating binder performance. Connecticut DOT currently
specifies the ductility test (AASHTO T-51), toughness and tenacity test (Colorado CP-L 2210), and elastic recovery test (AAS-
HTO T-301) as indicators for polymer modification on some grades of binder.

Ontario Ministry of Transportation has already implemented the DENT, MSCR, and ash content tests (to limit REOB). It
is phasing in the ExBBR test and looking at XRF and FTIR tests to quantify REOB. MOT has also developed a modified PAV
procedure that is currently being evaluated.

Quebecs Ministry of Transportation has developed a method to evaluate the adhesion between a binder and an aggregate.
For some binders, some criteria based on this method are added to the specifications to ensure a binder with superior adhesive
properties. These binders help to reduce stripping problems of the pavements.

The critical performance indicators for ensuring good high-temperature performance are either G*sin, RTFO (30.4%), or
the ASSHTO M 320 protocol (21.7%) (Figure 26). The corresponding indicators for intermediate temperature performance
are G*sin, PAV (28.3%), the ASSHTO M 320 protocol (15.2%), and MSCR (13%). Low-temperature performance is esti-
mated using primarily BBR (34.8%) (Figure 27).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 65

Although it is difficult to specify that field failures are the result of binder composition, at least eight DOTs cited concern.
In the past decade one lab noticed premature cracking (top-down mostly), starting in the wheel paths and then propagating in
a map cracking form. The lab recovered the binder in several cases and noticed a great loss of low temperature grade using an
ExBBR test. This is an indication that the SHRP aging protocol is not sufficient on the low temperature side.

FIGURE 26 Critical performance indicators for ensuring good high-temperature performance


(Source: Survey results).

FIGURE 27 Critical performance indicators for ensuring good low-temperature performance


(Source: Survey results).

Because there is no reliable field test, it is hard to determine if the failures can be attributed directly to the binder itself.
Stripping, bleeding, segregation, and other failures have been blamed on a bad binder, but nothing definitive could be repro-
duced in the laboratory. Failures can range from soft, tender mixes to prematurely cracked pavements. The problem is being
pursued by analyzing cores taken from failure locations to try to identify (1) the performance grade of asphalt used, (2) the
presence of any chemical contamination by gas chromatography/infrared spectroscopy, and (3) the presence of nonspecified
modifiers such as PPA, bio-oils, or REOB.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

66

CHAPTER SEVEN

CONCLUSIONS

An updated literature review on binder chemistry and the relationship between binder chemistry (including modifiers and
additives) and engineering properties reveals extensive progress in understanding asphalt materials. Composition informa-
tion is useful for understanding asphaltwhat makes it behave as it does and what makes one asphalt behave differently from
another. With the asphalt sources available, composition information can be used to improve the product through modification
with additives, by blending, and so on, or to alter use design procedures to accommodate specific properties. Composition
information can be used to match asphalt and aggregate, provide clues as to what modifications are necessary to make an
asphalt-aggregate system more serviceable under a given environment, diagnose failures, and provide information needed for
corrective measures. The survey reveals that DOTs current understanding of asphalt composition is very limited.

Asphalt binder supplies are shrinking because only a limited number of crude oils yield quality asphalt. Asphalt is no longer
a by-product of crude oil distillation, but rather it is a designed product produced at market demand rates and priced accordingly.
Currently, approximately 16.5 million tons are used in road construction each year. The demand for asphalt is increasing at approxi-
mately 3.1% per year based on improving economic conditions and a pressing need to repair and expand the nations infrastructure.

Asphalt is considered a colloidal or micellar system. The hydrocarbon insoluble components, asphaltenes and resins, are
dispersed in a hydrocarbon blend. The overall behavior of asphalt cement is controlled by the compatibility and the relation-
ships of the different components in this microscopically homogeneous mixture, rather than by the quantitative amount of any
single component. A proper balance of component types is necessary for a durable asphalt. Methods for fractionating asphalts
into the generic fractionssaturates, aromatics, resins, and asphaltenes (SARA)have evolved and new developments pro-
vide valuable information on changes in asphalt composition according to both source and date of production.

APPLICATIONS OF ANALYTICAL INSTRUMENTATION TO BINDER CHARACTERIZATION

The advent of modulated differential scanning calorimetry provides insight on asphalt microstructure. The development of
bitumen microstructure and calculations of the entropy and enthalpy of transitions confirm that bitumen is a structured amor-
phous phase containing a small crystalline phase.

Recently, researchers combined atomic resolution imaging using atomic force microscopy (AFM) to show the actual
atomic arrangement in an asphaltene molecule. Identifying molecular structures provides a foundation for understanding all
aspects of petroleum science, from colloidal structure and interfacial interactions to petroleum thermodynamics, enabling
a first-principles approach to optimizing resource utilization. Particularly, the findings contribute to a long-standing debate
about asphaltene molecular architecture. The impact of AFM imaging on understanding the microstructure and performance
characteristics of asphalt binders is immense. Each SARA chemical fraction influences asphalt phase structuring, most nota-
bly the well-known asphalt bee structures. Certain asphalt chemical parameters have a consistent and measurable effect
on the asphalt microstructure that is observed with AFM. Particular microstructures that emerged through chemical doping
were then discovered to have unique chemical polarity, which explicitly impacts the durability and performance of asphalt.

Gel permeation chromatography (GPC), a method of separating molecules based on their size and shape in solution, facili-
tates the analysis of polymer and asphalt components of polymer-modified asphalt cements. The ability of GPC to separate
mixtures by molecular size rather than by some complex property such as solubility or absorptivity is one of the great advan-
tages of the technique. This feature makes GPC a useful alternate technique for fractionating complicated mixtures, such as
crude oil residua, asphalts, and asphaltenes. Only one DOT appears to be using GPC on a routine basis, but the prevalence
of polymer-modified asphalts suggests that this technique could be employed more widely. The presence of recycled asphalt
pavement (RAP) and recycled asphalt shingles (RAS) in modified asphalt binders can be analyzed by carefully deconvoluting
the asphaltene/polymer region of the GPC chromatograms.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 67

Fourier transform infrared spectroscopy (FTIR) is one of the more important methods for fingerprinting asphalt materials
and quantifying the distribution of asphalt components. By determining the various chemical functional groups in a binder, an
understanding of its origin and history can be obtained. The FTIR method is employed to identify some antistripping agents
and to a more limited extent specialized additives in a binder. Amine-based antistripping agents are difficult to observe.
Quantitative redetermination of polymer content can be achieved with proper calibration techniques, but this is not a routine
procedure. FTIR is able to quickly assess aliphaticity, aromaticity, and extent of oxidative binder aging. Relatively low-cost
($20,000 to $40,000) portable devices have become available for FTIR. These can be employed in the field to test the chemical
composition of the delivered materials. These are point-and-shoot applications that could potentially be used by field techni-
cians with accuracy similar to that obtained by using traditional stationary laboratory equipment.

Proton nuclear magnetic resonance (1H NMR) spectroscopy has emerged as a very powerful and versatile tool for bitumen
characterization, but it is limited to the laboratory by the complexity of the instrumentation employed. Using 1H and 13C NMR
can yield information on average structural parameters of asphalt and asphaltenes. When information from NMR and GPC
is combined, possible structures for asphalt and mechanisms of aging are suggested.

As refiners efficiency allows them to extract more gasoline and other petroleum products from crude oil and as the source
of crudes that yield quality asphalt residua decreases, the need for additives to upgrade straight-run asphalts increases. The
most common additives to asphalt binders are polymers [styrene-butadiene-styrene (SBS) and styrene-butadiene rubber
(SBR), antistripping agents, softening agents, and polyphosphoric acid (PPA)]. Understanding the properties and contribu-
tions each of these nonbituminous additives to asphalt is critical for good binder design.

SURVEY FINDINGS

The survey results in chapter three cover numerous aspects of asphalt chemistry. State DOTs do not monitor the sources of
their asphalt binders. Only 13% of the respondents could identify their crude sources. Similarly, most DOTs are not familiar
with the process used to produce their binders. Significant differences in asphalt binder properties over the past 10 years were
reported by 42% of the respondents. Binders appear to be stiffer and exhibit poor low-temperature properties, and some pave-
ments aged prematurely. Although binders met specifications, experienced engineers could detect quarterly changes as crude
feeds change. Changes associated with seasonal or market fluctuations were reported by 38% of the respondents.

The dominant analytical procedures used for binder testing are rheological; that is, the performance grading (PG) proto-
col (100% of respondents). FTIR is employed by 23.9% of the DOTs, but this appears to be primarily used for research. The
third most common procedure was X-ray fluorescence spectroscopy (XRF). A number of modifiers/additives are currently
employed, including PPA (42.2% of respondents), antistripping agents (26.7%), and softening agents (22.2%). Documentation
of the use of modifiers depends on the binder suppliers; therefore, these data may not be complete. The addition of compaction
aids (waxes, wetting agents, and antistripping agents) is paving related and dictated by the job, aggregate, and plant. With the
exemption of polymers and PPA, the additives in a binder are usually not specified or quantified, because that is considered
the responsibility of the supplier. Some concern over the excessive use of PPA is reflected in the upper-limit specifications
imposed by 12 DOTs. Comments on specific DOT procedures can be found in Appendix A. More than 95% of the 40 respon-
dents replying to this question do not measure the compatibility of modifiers in the final binder blend. Failure to understand
this issue may lead to problems with long-term aging and pavement performance.

Prediction of asphalt cement aging is primarily accomplished using combined rolling thin-film oven and pressure again
vessel (PAV) procedures as defined in AASHTO M320. These procedures were developed for straight-run asphalts and they
may not apply to modified asphalts. The use of additives and chemical and polymer modifiers to enhance binder properties
has also greatly increased; in some cases, the oxidation kinetics of such modified binders are significantly different from those
of conventional binders. Stiffer binder grades may experience insufficient oxidation in the laboratory aging process. Studies
have confirmed that a better understanding of binder aging and oxidation can improve our ability to predict damage in asphalt
pavements. The most common suggestions to improve SHRP performance grading procedures focus on improving laboratory
aging procedures so that results relatable to field experience can be obtained. A one-point aging test such as PAV does not tell
the whole story; multiple PAVs provide some improvement, but an effective routine aging test is still lacking. Measurement of
aging rate would allow more flexibility in determining the long-term properties of binders and mixtures.

Efforts to understand the impact of recycled materials by extracting binder from mixtures are focused on RAP (65.2%)
and RAS (47.8%). A limited number of DOTs (fewer than 10) are using recycling (rejuvenating) agents; the most common

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

68

materials employed are soft asphalt fluxes and abietic acid esters. The recycling agents appear to be used on an experimental
basis for specific projects, and nine DOTs expressly state that they do not use them.

A majority of the respondents (69%) did not consider the current PG criteria satisfactory in identifying compositional/
chemical properties of a given binder. PG criteria, which are physical measurements, currently appear to fail to capture some
poorly performing binders. To better identify compositional properties of polymer-modified asphalt, Illinois DOT specifies
PG+ tests (force ductility, elastic recovery, and separation of polymer).

Procedures used to quantify the presence of modifiers/additives in final asphalt are surveyed. Polymer-modified binders
are checked using multiple stress creep recovery (MSCR) (30.4%), XRF (15.2%), viscosity (8.7%), and FTIR (6.5%). Soften-
ing agents such as refined engine oil bottoms (REOB) are checked using XRF. The blend compatibility with antistripping
agents, polymers, and PPA is tested in a limited number of cases using modern variants of the cigar tube test. In one case,
evidence for incompatibility in blends containing PPA was reported. A specification defining an upper limit on PPA is used
by 12 DOTs.

SUGGESTED FUTURE RESEARCH

Given the dearth of chemical information provided by DOTs and ministries of transportation in this study, a more diligent
effort to assess binder chemistry on a routine basis should be considered. More data are required to establish trends and the
proper analytical focus to prepare a practical guide for chemical analysis of binders. The relationship between the asphalt
chemistry of the various rheological properties and performance indicators of asphalt binders has yet to be defined. The
results of this synthesis indicate that the primary need for research is to identify more practical and cost-effective tests that
truly reflect field performance of asphalt mixtures. Better communication between asphalt producers and DOTs with regard
to asphalt cement composition would allow better understanding of the factors critical for long-term field performance.

Practical procedures to quantitatively confirm the presence of modifiers/additives in final asphalt cements are needed. In
addition to MSCR, viscosity, and FTIR, polymer-modified binders are readily characterized using GPC. Routine GPC analy-
sis of binder blends also provides useful information on asphaltene content and size distribution. Building a library of GPC
chromatograms will give DOTs a valuable resource for monitoring changes in the binder during field aging and identifying
the contribution of binder composition to field failures. Louisiana and Virginia DOTs are employing GPC to confirm the
presence and concentration of polymers, but a quick, easy field identification of modifiers (polymer and rubber) remains to
be developed.

It is important that protocols for ensuring the presence and effectiveness of antistripping agents be defined.

The blend capability of mixes with antistripping agents, polymers, and other additives needs to be evaluated more pre-
cisely. Several laboratories are developing alternate techniques for estimating binder performance. Connecticut DOT cur-
rently specifies the ductility test (AASHTO T-51), toughness and tenacity test (Colorado CP-L 2210), and elastic recovery
test (AASHTO T-301) as indicators for polymer modification on some grades of binder. Ontario Ministry of Transportation
(MOT) has already implemented the double-edge notch tension test, MSCR, and ash content test (to limit REOB). It is phasing
in the extended bending beam rheometer test and looking at XRF and FTIR tests to quantify REOB. MOT has also developed
a modified PAV procedure that is currently being evaluated. Quebec MOT has developed a method to evaluate the adhesion
between a binder and an aggregate. As these tests evolve, their utility could be shared with all DOTs and MOTs.

Asphalt pavement is Americas most recycled material. Despite the massive use of RAP in hot-mix asphalt production,
the chemico-physical phenomena that characterize the blending of these mixtures have not yet been completely explored.
The detection and understanding of these mechanisms, as well as the study of the heterogeneity that characterizes high-RAP
mix production, are fundamental to improving the approach to recycling, because they represent the source of the mixtures
characteristics and performance. It is important that the variability of binder rheology with different aging levels be studied
to evaluate RAP binder heterogeneity and its evolution with time. Protocols and blending charts defining the dosage of virgin
bitumen required in asphalt mixtures with or without RAP need further development. Investigating the specific contribu-
tions of all the components of the mixtures (virgin aggregates, fillers, recycling agents, and RAP) should be evaluated. The
overall goal of the study will be to provide guidance on the design and specification of RAP mixes and to reduce uncertainty
surrounding the performance of asphalt mixes designed and manufactured with RAP. Similar guidance is needed for mixes
containing RAP/RAS blends and ground tire rubber (GTR).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 69

Long-term aging of binder blends to ascertain the effect of modifiers and additives is needed. Continued research on
recycling agents and related additives to confirm their long-term efficacy could be pursued. The use of bio-based products
introduces new parameters to studies on compatibility and stability of asphalt mixtures. The issue is further complicated by
the addition of RAP, RAS, and GTR to mixtures.

Although it is desirable for a transportation agency to pursue the recycling of asphalt pavements, the economic assessment
to determine the best value among asphalt pavement recycling options remains a challenge. This issue has been long debated
at agency and industry levels. The problem is that there is no one size fits all answer to this issue. A number of considerations
can influence the outcome of an economic analysis of this issue. Several methods are available for recycling existing HMA
pavement materials into new pavements. Among these are inclusion as RAP in HMA mix designs, cold in-place recycling,
cold plant recycling, and hot in-place recycling. Continued research on life-cycle analysis of asphalt cements is needed.

It is important that studies of the morphology of asphalt binders using transmission electron microscopy be continued. The
various forms of microstructures in different types of asphaltenes and asphalt binder samples can be related to the impact of
additives as well as the progress of field aging. The prevalence, size, and shape of the microstructures in field-aged asphalt
binders may significantly affect binder engineering properties in pavement use as well as binder behavior in recycling or
rejuvenation. Hence, the microstructures are believed to be the key to understanding asphalt binder aging and aging control/
reversal methods.

Conventional rheological and chemical tests provide a global view of asphalt property and composition changes upon
aging, but offer few details on the changes at the microscopic level. Using atomic force microscopy (AFM), the micromechani-
cal properties of asphalts can be analyzed after different aging conditions. Aging was found to significantly increase the spa-
tial variations of the sample properties. It generally increases the ratio between the dissipated energy and total work to deform
the sample during the indentation process by AFM probe. It also appears to increase the adhesive and/or cohesive strength
of the sample. The asphaltene content and the size of microstructures both appear to affect the micromechanical properties
of the binders. It is also possible to use the AFM technique to distinguish the effects of additives on the morphological and
micromechanical properties of bitumen films.

Despite the progress made on morphological characterization of bitumen using AFM, the fundamental question of whether
the microstructures observed on bitumen surfaces represent its bulk structure remains to be addressed. Combining AFM
with other chemical analytical tools that can generate high resolution comparable to AFM would provide an avenue to linking
bitumens chemistry to its microscopic morphology and mechanical properties, and consequently benefit efforts to develop
structure-related models for bituminous materials across different-length scales.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

70

ACRONYMS AND ABBREVIATIONS

AC Asphalt cement

AFM Atomic force microscopy

AMPT Asphalt mixture performance tester

APA Asphalt pavement analyzer

ATR Attenuated total reflectance

BB Bio-oils/biobinders

BBR Bending beam rheometer

BMG Bio-modified gilsonite

CRM Crumb rubber modifier

Da Daltons, molecular weight in g/mole

DENT Double-edge notched tension

DLS Dynamic light scattering

DSC Differential scanning calorimetry

DSR Dynamic shear rheometer

DTA Differential thermal analysis

E* Dynamic modulus

ELSD Evaporative light scattering detector

EVA Ethylene-vinyl acetate copolymer

ExBBR Extended bending beam rheometer

FID Flame ionization detector

FTIR Fourier transform infrared spectroscopy

GPC Gel permeation chromatography

GTR Ground tire rubber

HDPE High-density polyethylene

HMA Hot-mix asphalt

IR Infrared

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 71

Jnr Nonrecoverable creep compliance

LDOTD Lousiana Department of Transportation and Development

LDPE Low-density polyethylene

LMS Large molecular size

LTRC Louisiana Transportation Research Center

MMS Medium molecular size

MOT Ministry of transportation

MS Mass spectroscopy

MSCR Multiple stress creep recovery

MW Molecular weight

NMR Nuclear magnetic resonance

NR Natural rubber

PADD Petroleum Administration for Defense Districts

PAH Polycyclic aromatic hydrocarbon

PAV Pressure aging vessel

PE Polyethylene

PET Poly (ethylene terephthalate)

PG Performance grading

PMA Polymer-modified asphalt

PPA Polyphosphoric acid

PRP Polymer-rich phase

QC Quality control

RA Recycling agent

RAP Recycled asphalt pavement

RAS Recycled asphalt shingles

REOB Refined engine oil bottoms

RET Reactive elastomeric terpolymer

RTFO Rolling thin-film oven

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

72

SANS Small angle neutron scattering

SARA Saturates, aromatics, resins, and asphaltenes

SAR-AD Saturates, aromatics, resins, and asphaltene-determinator method

SAXS Small angle X-ray scattering

SB Styrene-butadiene diblock copolymer

SBR Styrene-butadiene rubber

SBS Styrene-butadiene-styrene triblock copolymer

SEM Scanning electron microscopy

SFA Soy fatty acids

SMS Small molecular size

STOA Short-term oven aging

TFO Thin-film oven

Tg Glass transition temperature

TGA Thermogravimetric analysis

TLC Thin layer chromatography

TSR Tensile strength ratio

UV Ultraviolet spectroscopy

VDOT Virginia Department of Transportation

VPO Vapor pressure osmometry

WMA Warm-mix asphalt

XRF X-ray fluorescence spectroscopy

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 73

REFERENCES

Allen, R.G., D.N. Little, A. Bhasin, and C.J. Glover, The Effects of Chemical Composition on Asphalt Microstructure and
Their Association to Pavement Performance, International Journal of Pavement Engineering, Vol. 15, No. 1, 2014, pp.
922.
Altgelt, K.H., Fractionation of Asphaltenes by Gel Permeation Chromatography, Journal of Applied Polymer Science, Vol.
9, No. 10, 1965, pp. 33893393.
American Association of State Highways and Transportation Officials (AASHTO), Standard Practice for Recovery of Asphalt
Binder from Solution by Abson Method, R 59, AASHTO, Washington, D.C., 2011.
American Association of State Highway and Transportation Officials (AASHTO), Standard Specification for Superpave
Volumetric Mix Design, M323, AASHTO, Washington, D.C., 2013.
American Association of State Highway and Transportation Officials (AASHTO), Standard Specification for Performance
Graded Asphalt Binder, M320, AASHTO, Washington, D.C., 2016.
Apeagyei, A.K., W.G. Buttlar, and B.J. Dempsey, Investigation of Cracking Behavior of Antioxidant-Modified Asphalt Mix-
tures (with discussion), Association of Asphalt Paving Technologists, Lino Lakes, Minn., 2008 [Online]. Available: https://
trid.trb.org/view/890291.
Arnold, T.S., J. Youtcheff, and S.P. Needham, Transportation Research Board E Circular 160, Use of Phosphoric Acid as
Modifier for Hot-Mix Asphalt, Transportation Research Board of the National Academies, Washington, D.C., 2012, pp.
4051.
ASTM Book of Standards (ASTM), Standard Test Methods for Separation of Asphalt into Four Fractions, D47124-01, West
Conshohocken, Pa., 2008.
Babagoli, R., M. Hasaninia, and N.M. Namazi, Laboratory Evaluation of the Effect of Gilsonite on the Performance of Stone
Matrix Asphalt Mixtures, Road Materials and Pavement Design, Vol. 16, No. 4, 2015, pp. 889906.
Bahia, H.U., Synthesis of Use of Crumb Rubber in Hot Mix Asphalt, Transportation Research Board of the National Acade-
mies, Washington, D.C., 2011, 45 pp. [Online]. Available: http://rmrc.wisc.edu/wp-content/uploads/2012/11/Bahia-No.-
54-Synthesis-of-the-Use-of-Crumb-Rubber-in-Hot-Mix-Asphalt.pdf, https://trid.trb.org/view/1316214.
Barbour, R.V. and J.C. Petersen, Molecular Interactions of Asphalt. Infrared Study of the Hydrogen-Bonding Basicity of
Asphalt, Analytical Chemistry, Vol. 46, No. 2, 1974, pp. 273277.
Batten, E., Y. Mehta, A. Nolan, S. Zorn, and K. Dahm, Correlation Between PG Plus, Superpave PG Specifications and
Molecular Weight from GPC for Different Polymer Modified Binders. Geo-Frontiers Congress, Dallas, Tex., American
Society of Civil Engineers, Industrial Fabrics Association International, North American Geosynthetics Society, 2011
[Online]. Available: http://dx.doi.org/10.1061/41165(397)464, https://trid.trb.org/view/1109340.
Baumgardner, G.L., Why and How of Polyphosphoric Acid ModificationAn Industry Perspective, Journal of the Associa-
tion of Asphalt Paving Technology, Vol. 79, 2010, pp. 663678.
Baumgardner, G.L., Characterization and Implementation of Ground Tire Rubber as Post-Consumer Polymers for Asphalt
Concrete, PhD, Mississippi State University, ProQuest LLC, Ann Arbor, Mich., 2015.
Baumgardner, G.L., et al., Quantitative Analysis of Functional Polymer in Recycled Tyre Rubber Used in Modified Asphalt
Binders, Road Materials and Pavement Design, Vol. 15, Suppl. 1, 2014, pp. 263278.
Baumgardner, G.L., J.M. Hemsley, W. Jordan, III, and I.L. Howard, Laboratory Evaluation of Asphalt Mixtures Containing
Dry Added Ground Tire Rubber and a Processing Aid, Journal of the Association of Asphalt Paving Technology, Vol. 81,
2012, pp. 507539.
Betancourt Cardozo, F., E. Avella Moreno, and C.A. Trujillo, Structural Characterization of Unfractionated Asphalts by 1H
NMR and 13C NMR, Energy Fuels, Vol. 30, No. 4, 2016, pp. 27292740.
Bharati, S., G.A. Rostum, and R. Loberg, Calibration and Standardization of Iatroscan (TLC_FID) Using Standards Derived
from Crude Oils, Journal of Organic Geochemistry, Vol. 22, No. 3, 1994, pp. 835862.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

74

Bhasin, A. and V. Ganesan, Preliminary Investigation of Using a Multi-component Phase Field Model to Evaluate Micro-
structure of Asphalt Binders, International Journal of Pavement Engineering, Vol. 16, Suppl. 1, 2015, pp. 18.
Bhasin, A., J.E. Howson, E.A. Masad, D.N. Little, and R.L. Lytton, Effect of Modification Processes on Bond Energy of
Asphalt Binders, Transportation Research Record: Journal of the Transportation Research Board, No. 1998, Transporta-
tion Research Board of the National Academies, Washington, D.C., 2007, pp. 2937.
Borrego, A.G., C.G. Blanco, J.G. Prado, C. Diaz, and M.D. Guillen, 1H NMR and FTIR Spectroscopic Studies of Bitumen
and Shale Oil from Selected Spanish Oil Shales, Energy Fuels, Vol. 10, No. 1, 1996, pp. 7784.
Boysen, R.B. and J.F. Schabron, The Automated Asphaltene Determinator Coupled with Saturates, Aromatics, and Resins
Separation for Petroleum Residua Characterization, Energy & Fuels, Vol. 27, No. 8, 2013, pp. 46544661.
Branthaver, J.F., M.W. Catalfomo, and J.C. Petersen, Ion Exchange Chromatography Separation of SHRP Asphalts, Fuel
Science and Technology International, Vol. 10, No. 46, 1992, pp. 855885.
Brule, B., G. Ramond, and C. Such, Relationships Between Composition, Structure, and Properties of Road Asphalts: State
of Research at the French Public Works Central Laboratory, Transportation Research Record 1096, Transportation
Research Board, National Research Council, Washington, D.C., 1986, pp. 2234.
Cao, X.-j., Research on the Performance of New Reactive Asphalt Modifier, Dangdai Huagong, Vol. 44, No. 6, 2015, pp.
13501352.
Carpenter, S.H. and J.R. Wolosick, Modifier Influence in the Characterization of Hot-Mix Recycled Material, Transporta-
tion Research Record 777, Transportation Research Board, National Research Council, Washington, D.C., 1980, pp. 1522.
Chailleux, E., et al., Alternative Binder from Microalgae: Algoroute Project, Transportation Research E-Circular 165,
Transportation Research Board of the National Academies, Washington, D.C., 2012, pp. 714.
Chambrion, P., R. Bertau, and P. Ehrburger, Characterization of Bitumen by Differential Scanning Calorimetry, Fuel, Vol.
75, No. 2, 1996, pp. 144148.
Churchill, E., S. Amirkhanian, and J. Burati, Jr., HP-GPC Characterization of Asphalt Aging and Selected Properties,
Journal of Materials in Civil Engineering, Vol. 7, No. 1, 1995, pp. 4149.
Claudy, P., J.M. Letoffe, G.N. King, and J.P. Planche, Characterization of Asphalt Cements by Thermomicroscopy and Dif-
ferential Scanning Calorimetry, Fuel Science & Technology International, Vol. 10, No. 4, 1992a, pp. 735765.
Claudy, P., et al., A New Interpretation of Time-Dependent Physical Hardening in Asphalt Based on DSC and Optical Ther-
moanalysis, Division of Fuel Chemistry Preprints, Vol. 37, No. 3, 1992b, pp. 14081426.
Clifton, E.C., R.C. Barborak, and C. Coward, Detection and Estimation of Re-refined Engine Oil Bottoms in Asphalt Binders
Using Wavelength Dispersive X-Ray Fluorescence Spectroscopy, Texas Department of Transportations Approach 95th
Annual Meeting of the Transportation Research Board, Washington, D.C., Jan. 2016 [Online]. Available: ttps://trid.trb.
org/view/1394231.
Collins, R.J. and S.K. Ciesielski, NCHRP Synthesis 199: Recycling and Use of Waste Materials and By-products in Highway
Construction, Transportation Research Board, National Research Council, Washington, D.C., 1994, 92 pp.
Cooper, S.B., I. Negulescu, S.S. Balamurugan, L. Mohammad, and W.H. Daly, Binder Composition and Intermediate Tem-
perature Cracking Performance of Asphalt Mixtures Containing RAS, Road Materials and Pavement Design, 2015, pp.
275295.
Corbett, L.W., Densimetric Method for Characterizing Asphalt, Analytical Chemistry, Vol. 36, No. 10, 1964, pp.
19671971.
Corbett, L.W., Composition of Asphalt Based on Generic Fractionation, Using Solvent Deasphaltening, Elution-Adsorption
Chromatography, and Densiometric Characterization, Analytical Chemistry, Vol. 41, No. 4, 1969, pp. 576579.
Corbett, L.W., Relation Between Composition and Physical Properties of Asphalt, Proceedings of the Assocation of Asphalt
Paving Technology, Vol. 39, 1970, pp. 481491.
Corbett, L.W., Refinery Processing of Asphalt Cement, Transportation Research Record, 999, Transportation Research
Board, National Research Council, Washington, D.C., 1984, pp. 16.
Corbett, L.W. and R.E. Swarbrick, Structural Analysis of Asphalt Petrolene Fractions, Preprints, Division of Petroleum
Chemistray, American Chemical Society, Vol. 11, No. 2, 1966, pp. B-161B-166.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 75

Costa, L.M.B., H.M.R.D. Silva, J.R.M. Oliveira, and S.R.M. Fernandes, Incorporation of Waste Plastic in Asphalt Binders
to Improve Their Performance in the Pavement, International Journal of Pavement Research and Technology, Vol. 6, No.
4, 2013, pp. 457464.
DAngelo, J.A., Effect of Polyphosphoric Acid on Asphalt Binder Properties, Transportation Research Board of the National
Academies, Washington, D.C., 2012, pp. 2739 [Online]. Available: http://www.trb.org/Publications/Blurbs/166590.aspx.
Daly, W.H., I.I. Negulesci, and S.S. Balamurugan, Implementation of GPC Characterization of Asphalt Binders at Louisiana
Materials Laboratory, FHWA/LA, 13/505, 2013, 114 pp. [Online]. Available: http://www.ltrc.lsu.edu/pdf/2013/ts_505.
pdf.
Daly, W.H., Z.-Y. Qiu (Chiu), and I.I. Negulescu, Differential Scanning Calorimetry Study of Asphalt Crystallinity, Trans-
portation Research Record 1535, Transportation Research Board, National Research Council, Washington, D.C., 1996, pp.
5460.
Das, P., N. Kringos, V. Wallqvist, and B. Birgisson, Micromechanical Investigation of Phase Separation in Bitumen by Com-
bining Atomic Force Microscopy with Differential Scanning Calorimetry Results, Road Materials and Pavement Design,
Vol. 14, Suppl. 1, 2013, pp. 2537.
Davis, C. and C. Castorena, Implications of Physico-Chemical Interactions in Asphalt Mastics on Asphalt Microstructure,
Construction and Building Materials, Vol. 94, No. 2015, pp. 8389.
Davison, R.R., C.J. Glover, B.L. Burr, and J.A. Bullin, Size Exclusion Chromatography of Asphalts, In Handbook of Size
Exclusion Chromatography and Related Techniques, C.-S. Wu, Ed., Marcel Dekker, Inc., New York City, N.Y., 1995, pp.
211247.
DeDene, C.D. and Z. You, The Performance of Aged Asphalt Materials Rejuvenated with Waste Engine Oil, International
Journal of Pavement Research and Technology, Vol. 7, No. 2, 2014, pp. 145152.
del Barco Carrin, A.J., D. Lo Presti, and G.D. Airey, Binder Design of High RAP Content Hot and Warm Asphalt Mixture
Wearing Courses, Taylor & Francis, Abingdon, United Kingdom, 2015 [Online]. Available: http://dx.doi.org/10.1080/146
80629.2015.1029707.
Dessouky, S., D. Contreras, J. Sanchez, and D. Park, Anti-Oxidants Effect on Bitumen Rheology and Mixes Mechanical
Performance, 2015, http://dx.doi.org/10.1061/9780784479278.002.
Dickie, J.P. and T.F. Yen, Macrostructures of the Asphaltic Fractions by Various Instrumental Methods, Analytical Chem-
istry, Vol. 39, No. 14, 1967, pp. 18471852.
Durrieu, F., J.-P. Planche, J. Lamontagne, V. Mouillet, and J. Kister, Application of Infrared Micro-Spectroscopy in the Study
of Bitumen-Polymer Aging, Bulletin de Liaison des Laboratoires des Ponts et Chaussees, Vol. 240, 2003, pp. 1525.
Elseifi, M.A., et al., Relationship Between Molecular Compositions and Rheological Properties of Neat Asphalt Binder at
Low and Intermediate Temperatures, Journal of Materials in Civil Engineering, Vol. 22, No. 12, 2010, pp. 12881294.
Evans, M., R. Marchildon, and S.A.M. Hesp, Effects of Physical Hardening on Stress Relaxation in Asphalt Cements: Impli-
cations for Pavement Performance, Transportation Research Record: Journal of the Transportation Research Board,
2207, Transportation Research Board of the National Academies, Washington, D.C., 2011, pp. 3442.
Fee, D., R. Maldonado, G. Reinke, and H. Romagosa, Polyphosphoric Acid Modification of Asphalt, Transportation
Research Record: Journal of the Transportation Research Board, No. 2179, Transportation Research Board of the National
Academies, Washington, D.C., 2010, pp. 4957.
Fernandes, P.R., S.d.A. Soares, R.F. Nascimento, J.B. Soares, and R.M. Cavalcante, Evaluation of Polycyclic Aromatic
Hydrocarbons in Asphalt Binder Using Matrix Solid-Phase Dispersion and Gas Chromatography, Journal of Chromato-
graphic Science, Vol. 47, No. 9, 2009, pp. 789793.
Fini, E.H., I.L. Al-Qadi, Z. You, B. Zada, and J. Mills-Beale, Partial Replacement of Asphalt Binder with Bio-Binder: Char-
acterisation and Modification, International Journal of Pavement Engineering, Vol. 13, No. 6, 2012, pp. 515522.
Fini, E.H., S. Hosseinnezhad, D.J. Oldham, E. Chailleux, and V. Gaudefroy, Source dependency of rheological and surface
characteristics of bio-modified asphalts, Road Materials and Pavement Design, 2016, Ahead of Print.
Fini, E. H., et al., Chemical Characterization of Biobinder from Swine Manure: Sustainable Modifier for Asphalt Binder,
Journal of Materials in Civil Engineering, Vol. 23, No. 11, 2011, pp. 15061513.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

76

Fischer, H.R. and E.C. Dillingh, On the Investigation of the Bulk Microstructure of Bitumen - Introducing Two New Tech-
niques, Fuel, Vol. 118, 2014, pp. 365368.
Freedoniagroup, Asphalt, Freedonia Group Study, July 2015. [Online]. Available: http://www.freedoniagroup.com/industry-
study/asphalt-3304.htm.
Garcia-Morales, M., P. Partal, F.J. Navarro, and C. Gallegos, Effect of Waste Polymer Addition on the Rheology of Modified
Bitumen, Fuel, Vol. 85, No. 78, 2006, pp. 936943.
Garrick, N.W., Use Of Gel-Permeation Chromatography in Predicting Properties of Asphalt, Journal of Materials in Civil
Engineering, Vol. 6, No. 3, 1994, pp. 376389.
Garrick, N.W. and L.E. Wood, Relationship Between High-Pressure Gel Permeation Chromatography Data and the Rheo-
logical Properties of Asphalts, Transportation Research Record 1096, Transportation Research Board, National Research
Council, Washington, D.C., 1986, pp. 3541.
Gearhart, J.A. and L. Garwin, Resid-Extraction Process Offers Flexibility, Oil & Gas Journal, Vol. 74, No. 24, 1976, pp.
6366.
George, G.N. and M.L. Gorbaty, Sulfur K-Edge X-Ray Absorption Spectroscopy of Petroleum Asphaltenes and Model Com-
pounds, Journal of the American Chemical Society, Vol. 111, No. 9, 1989, pp. 31823186.
Glover, C.J., J.A. Bullin, J.W. Button, R.R. Davison, and G.R. Donaldson, Characterization of Asphalts Using Gel Permeation
Chromatography and Other Methods, Texas Transportation Institute Report, Texas A&M Transportation Institute, Col-
lege Station, 1987, 203 pp.
Goodrich, J.L., J.E. Goodrich, and W.J. Kari, Asphalt Composition Tests: Their Application and Relation to Field Perfor-
mance, Transportation Research Record 1096, Transportation Research Board, National Research Council, Washington,
D.C., 1986, pp. 146167.
Gould, K.A. and I.A. Wiehe, Natural Hydrogen Donors in Petroleum Resids, Energy Fuels, Vol. 21, No. 3, 2007, pp.
11991204.
Hasan, M.U., M.F. Ali, and A. Bukhari, Structural Characterization of Saudi Arabian Heavy Crude Oil by NMR Spectros-
copy, Fuel, Vol. 62, No. 5, 1983, pp. 518523.
Herrington, P.R. and Y. Wu, Effect of Inter-Molecular Association on Bitumen Oxidation, Petroleum Science and Technol-
ogy, Vol. 17, No. 3 and 4, 1999, pp. 291318.
Herrington, P.R. and Y. Wu, Effect of Inter-Molecular Association on Bitumen Oxidation, In Transportation Research
Circular E-C140, A Review of the Fundamentals of Asphalt Oxidation, Transportation Research Board of the National
Academies, Washington, D.C., 2015, pp. 1830.
Hesp, S.A.M., et al., Asphalt Pavement Cracking: Analysis of Extraordinary Life Cycle Variability in Eastern and Northeast-
ern Ontario, International Journal of Pavement Engineering, Vol. 10, No. 3, 2009, pp. 209227.
Hon, K.H., D.K. Trung, S.L. Malhotra, and L.P. Blanchard, Effect of Molecular Weight and Composition on the Glass Transi-
tion Temperatures of Asphalts, Analytical Chemistry, Vol. 50, No. 7, 1978, pp. 976979.
Hoppe, E.J., D.S. Lane, G.M. Fitch, and S. Shetty, Feasibility of Reclaimed Asphalt Pavement (Rap) Use as Road Base and
Subbase Material, Virginia Center for Transportation Innovations and Research, Richmond, 2015, 42 pp. [Online]. Avail-
able: http://www.virginiadot.org/vtrc/main/online_reports/pdf/15-r6.pdf.
Hossain, M.I., V. Veginati, and J. Krukow, Thermodynamics Between RAP/RAS and Virgin aggregates During Asphalt
Concrete ProductionA Literature Review, University of Illinois Center for Transportation, ChampaignUrbana, 2015,
79 pp. [Online]. Available: https://apps.ict.illinois.edu/projects/getfile.asp?id=3571.
Huang, J., Characterization of Asphalt Fractions by NMR Spectroscopy, Petroleum Science and Technology, Vol. 28, No.
6, 2010, pp. 618624.
Im, S., F. Zhou, R. Lee, and T. Scullion, Impacts of Rejuvenators on Performance and Engineering Properties of Asphalt
Mixtures Containing Recycled Materials, Construction and Building Materials, Vol. 53, 2014, pp. 596603.
Jain, P.K., O.S. Tyagi, and H. Singh, Physico-Chemical and Compositional Aspects of Bitumen Bearing Crudes and Their
Instrumental Characterization, Petroleum Science and Technology, Vol. 16, 1998, pp. 567582.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 77

Jahangir, R., D. Little, and A. Bhasin, Evolution of Asphalt Binder Microstructure Due to Tensile Loading Determined Using
AFM and Image Analysis Techniques, International Journal of Pavement Engineering, Vol. 16, No. 4, 2015, pp.
337349.
Jemison, H.B., B.L. Burr, R.R. Davison, J.A. Bullin, and C.J. Glover, Application and Use of the ATR, FT-IR Method to
Asphalt Aging Studies, Preprints, Division of Petroleum Chemistry, American Chemical Society, Vol. 35, No. 3, 1990, pp.
490495.
Jenkins, R., X-Ray Fluorescence Spectrometry, 2nd ed., New York: Wiley-Interscience, 2012.
Jennings, P.W., et al., NMR Spectroscopy in the Characterization of Eight Selected Asphalts, Fuel Science and Technology
Intl, Vol. 10, No. 46, 1992, pp. 887907.
Jennings, P.W., J.A. Pribanic, M.F. Raub, J.A. Smith, and T.M. Mendes, Advanced High Performance Gel Permeation Chro-
matography Methodology, SHRP-A-630, Transportation Research Board, National Research Council, Washington, D.C.,
1993, 122 pp. [Online]. Available: http://onlinepubs.trb.org/onlinepubs/shrp/SHRP-A-630.pdf.
Jennings, P.W., J.A.S. Pribanic, J.A. Smith, and T.M. Mendes, HP-GPC Analysis of Asphalt Fractions in the Study of Molecu-
lar Self-Assembly in Asphalt, Division of Fuel Chemistry Preprints, Vol. 37, No. 3, 1992, pp. 13121319.
Jimenez-Mateos, J.M., L.C. Quintero, and C. Rial, Characterization of Petroleum Bitumens and Their Fractions by Thermo-
gravimetric Analysis and Differential Scanning Calorimetry, Fuel, Vol. 75, No. 15, 1996, pp. 16911700.
Jones, D.R., SHRP Materials Reference Library: Asphalt Cements: a Concise Data Compilation, Transportation Research
Board, National Research Council, Washington, D.C., 1993, 28 pp. [Online]. Available: http://onlinepubs.trb.org/
onlinepubs/shrp/SHRP-A-645.pdf.
Keyf, S., The Modification of Bitumen with Reactive Ethylene Terpolymer, Styrene Butadiene Styrene and Variable Amounts
of Ethylene Vinyl Acetate, Research on Chemical Intermediates, Vol. 41, No. 3, 2015, pp. 14851497.
Kim, K.W. and J.L. Burati, Jr., Use of GPC Chromatograms to Characterize Aged Asphalt Cements, Journal of Materials
in Civil Engineering, Vol. 5, No. 1, 1993, pp. 4152.
Kim, K.W., K. Kim, Y.S. Doh, and S.N. Amirkhanian, Estimation of RAPs Binder Viscosity Using GPC Without Binder
Recovery, Journal of Materials in Civil Engineering, Vol. 18, No. 4, 2006, pp. 561567.
Kim, S., et al., Estimation of Service-Life Reduction of Asphalt Pavement Due to Short-Term Ageing Measured by GPC from
Asphalt Mixture, Road Materials and Pavement Design, Vol. 17, No. 1, 2016, pp. 153167.
Klabunde, K.J. and K. Shrestha, Sustainable Asphalt Pavements Using Bio-Binders from Bio-Fuel Waste, Report # MATC-
KSU: 164, Mid-America Transportation Center, University of NebraskaLincoln, 2014, 22 pp. [Online]. Available: http://
matc.unl.edu/assets/documents/matcfinal/Klabunde_SustainableAsphaltPavementsUsingBio-BindersfromBio-
FuelWaste.pdf, https://trid.trb.org/view/1344946.
Kwon, O.-s., et al. Evaluation of Age-Retardation of Asphalt Mixture Using Antioxidants Based on Large Molecular Size Ratio
of Asphalt Chromatogram, 2016. [Online]. Available: https://trid.trb.org/view/1393799.
Lamontagne, J., P. Dumas, V. Mouillet, and J. Kister, Comparison by Fourier Transform Infrared (FTIR) Spectroscopy of
Different Ageing Techniques: Application to Road Bitumens, Fuel, Vol. 80, No. 4, 2001, pp. 483488.
Lee, J., E. Denneman, and Y. Choi, Maximising the Re-Use of Reclaimed Asphalt Pavement: Outcomes of Year Two: RAP Mix
Design, AP-T286-15, Austroads, 2015, 67 pp. [Online]. Available: https://www.onlinepublications.austroads.com.au/
items/AP-T286-15.
Lee, S.-J., S.N. Amirkhanian, K. Shatanawi, and K.W. Kim, Short-Term Aging Characterization of Asphalt Binders Using
Gel Permeation Chromatography and Selected Superpave Binder Tests, Construction and Building Materials, Vol. 22,
No. 11, 2008, pp. 22202227.
Lee, Y.S., W.-K. Lee, S.-G. Cho, I. Kim, and C.-S. Ha, Quantitative Analysis of Unknown Compositions in Ternary Polymer
Blends: A Model Study on NR/SBR/BR System, Journal of Analytical and Applied Pyrolysis, Vol. 78, No. 1, 2007, pp.
8594.
Lesueur, D., The Colloidal Structure of Bitumen: Consequences on the Rheology and on the Mechanisms of Bitumen Modi-
fication, Advances in Colloid and Interface Science, Vol. 145, No. 12, 2009, pp. 4282.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

78

Liu, M., M.A. Ferry, R.R. Davison, C.J. Glover, and J.A. Bullin, Oxygen Uptake as Correlated to Carbonyl Growth in Aged
Asphalts and Asphalt Corbett Fractions, Industrial & Engineering Chemistry Research, Vol. 37, No. 12, 1998, pp.
46694674.
Lo Presti, D. and G. Airey, Tyre Rubber-Modified Bitumens Development: The Effect of Varying Processing Conditions,
Road Materials and Pavement Design, Vol. 14, No. 4, 2013, pp. 888900.
Loeber, L., O. Sutton, J. Morel, J.M. Valleton, and G. Muller, New Direct Observations of Asphalts and Asphalt Binders by
Scanning Electron Microscopy and Atomic Force Microscopy, Journal of Microscopy, Vol. 182, No. 1, 1996, pp. 3239.
Louisiana Department of Transportation and Development, Standard Specifications for Roads and Bridges, 2016.
Lu, X. and U. Isacsson, Chemical and Rheological Evaluation of Ageing Properties of SBS Polymer Modified Bitumens,
Fuel, Vol. 77, No. 9/10, 1998, pp. 961972.
Lu, X., M. Langton, P. Olofsson, and P. Redelius, Wax Morphology in Bitumen, Journal of Materials in Civil Engineering,
Vol. 40, No. 8, 2005, pp. 18931900.
Ma, L., Z. Li, and J. Huang, Investigation of Chemistry by FTIR and NMR During the Natural Exposure Aging of Asphalt,
Geotechnical Special Publications, Vol. 212, Pavements and Materials, Geo-Institute, 2011, pp. 150157.
Marsac, P., et al., Potential and Limits of FTIR Methods for Reclaimed Asphalt Characterisation, Materials and Structures
(Dordrecht, the Netherlands), Vol. 47, No. 8, 2014, pp. 12731286.
Masson, J.F., V. Leblond, and J. Margeson, Bitumen Morphologies by Phase-Detection Atomic Force Microscopy, Journal
of Microscopy (Oxford, U.K.), Vol. 221, No. 1, 2006, pp. 1729.
Masson, J.F., V. Leblond, J. Margeson, and S. Bundalo-Perc, Low-Temperature Bitumen Stiffness and Viscous Paraffinic
Nano- and Micro-Domains by Cryogenic AFM and PDM, Journal of Microscopy (Oxford, U.K.), Vol. 227, No. 3, 2007,
pp. 191202.
Masson, J.F. and G.M. Polomark, Bitumen Microstructure by Modulated Differential Scanning Calorimetry, Thermochi-
mica Acta, Vol. 374, No. 2, 2001, pp. 105114.
Masson, J.F. and G.M. Polomark, Bitumen Microstructure by Modulated Differential Scanning Calorimetry [erratum to
document cited in CA135:139515], Thermochimica Acta, Vol. 413, No. 12, 2004, p. 273.
Masson, J.F., G.M. Polomark, and P. Collins, Time-dependent Microstructure of Bitumen and Its Fractions by Modulated
Differential Scanning Calorimetry, Energy Fuels, Vol. 16, No. 2, 2002, pp. 470476.
Masson, J.F., T. Price, and P. Collins, Dynamics of Bitumen Fractions by Thin-Layer Chromatography/Flame Ionization
Detection, Energy & Fuels, Vol. 15, No. 4, 2001, pp. 955960.
Menapace, I., E. Masad, A. Bhasin, and D. Little, Microstructural Properties of Warm Mix Asphalt Before and after Labo-
ratory-Simulated Long-Term Ageing, Road Materials and Pavement Design, 2015, ahead of print.
Michalica, P., P. Daucik, and L. Zanzotto, Monitoring of Compositional Changes Occurring During the Oxidative Aging of
Two Selected Asphalts from Different Sources, Petroleum Coal, Vol. 50, No. 2, 2008, pp. 110.
Mitra-Kirtley, S., O. Mullins, C. Ralston, D. Sellis, and C. Pareis, Analysis of Sulfur X-Ray Absorption Near-Edge Spectros-
copy in Asphaltenes, Resins, and Maltenes of Two Different Crude Oils, Preprints of the Division of Fuel Chemistry, Vol.
42, 1997, pp. 419422.
Mogawer, W.S., A. Booshehrian, S. Vahidi, and A.J. Austerman, Evaluating the Effect of Rejuvenators on the Degree of
Blending and Performance of High RAP, RAS, and RAP/RAS Mixtures, Road Materials and Pavement Design, Vol. 14,
No. 2, 2013, pp. 193213.
Mohan, D., C.U. Pittman, Jr., and P.H. Steele, Pyrolysis of Wood/Biomass for Bio-oil: A Critical Review, Energy Fuels, Vol.
20, No. 3, 2006, pp. 848889.
Morgan, T.J., P. Alvarez-Rodriguez, A. George, A.A. Herod, and R. Kandiyoti, Characterization of Maya Crude Oil Maltenes
and Asphaltenes in Terms of Structural Parameters Calculated from Nuclear Magnetic Resonance (NMR) Spectroscopy
and Laser Desorption-Mass Spectroscopy (LD-MS), Energy Fuels, Vol. 24, No. 7, 2010, pp. 39773989.
Mothe, M.G., L.F.M. Leite, and C.G. Mothe, Thermal Characterization of Asphalt Mixtures by TG/DTG, DTA and FTIR,
Journal of Thermal Analysis and Calorimetry, Vol. 93, No. 1, 2008, pp. 105109.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 79

Mouillet, V., J. Lamontagne, F. Durrieu, J.-P. Planche, and L. Lapalu, Infrared Microscopy Investigation of Oxidation and
Phase Evolution in Bitumen Modified with Polymers, Fuel, Vol. 87, No. 7, 2008, pp. 12701280.
Mousavi, M., F. Pahlavan, D. Oldham, T. Abdollahi, and E.H. Fini, Alteration of Intermolecular Interactions Between Units
of Asphaltene Dimers Exposed to an Amide-Enriched Modifier, RSC Advances, Royal Society of Chemistry, Vol. 6, No.
58, 2016, pp. 5347753492.
Mullins, O.C., The Asphaltenes, Annual Review of Analytical Chemistry, Vol. 4, 2011, pp. 393418.
Mullins, O.C., B. Martinez-Haya, and A.G. Marshall, Contrasting Perspective on Asphaltene Molecular Weight, this com-
ment vs. the overview of A.A. Herod, K.D. Bartle, and R. Kandiyoti, Energy Fuels, Vol. 22, No. 3, 2008, pp. 17651773.
Nahar, S.N., et al., Is Atomic Force Microscopy Suited as Tool for Fast Screening of Bituminous Materials? Interlaboratory
Comparison Study, 2013 [Online]. Available: https://trid.trb.org/view/1242403.
Nciri, N., S. Song, N. Kim, and N. Cho, Chemical Characterization of Gilsonite Bitumen, Journal of Petroleum and Envi-
ronmental Biotechnology, Vol. 5, No. 5, 2014, pp. 193/191193/110.
Negulescu, I., et al., Chemical and Rheological Characterization of Wet and Dry Aging of SBS Copolymer Modified Asphalt
Cements: Laboratory and Field Evaluation, Journal of the Association of Asphalt Paving Technologists, Vol. 75, 2006, pp.
267296.
Noel, F. and L.W. Corbett, Crystalline Phases in Asphalts, Journal of the Institute of Petroleum, London, Vol. 56, No. 551,
1970, pp. 261268.
Oasmaa, A., Y. Solantausta, V. Arpiainen, E. Kuoppala, and K. Sipila, Fast Pyrolysis Bio-Oils from Wood and Agricultural
Residues, Energy Fuels, Vol. 24, No. 2, 2010, pp. 13801388.
Ontario Ministry of Transportaton, Special Provision No 111F09 June 2016, amended to Section 1101.02 of OPSS 110 Speci-
fications for Performance Graded Asphalt Cement Acceptance, 2016.
Origin Data Analysis and Graphing Software [Online]. Available: http://www.originlab.com/ [accessed April 30, 2015].
Ozbay, N., A.E. Putun, and E. Putun, Bio-Oil Production from Rapid Pyrolysis of Cottonseed Cake: Product Yields and
Compositions, International Journal of Energy Research, Vol. 30, No. 7, 2006, pp. 501510.
Pahlavan, F., M. Mousavi, A. Hung, and E.H. Fini, Investigating Molecular Interactions and Surface Morphology of Wax-
Doped Asphaltenes, Physical Chemistry Chemical Physics, Vol. 18, No. 13, 2016, pp. 88408854.
PerkinElmer Life and Analytical Sciences, FTIR Spectroscopy: Attenuated Total Reflectance (ATR) Technical Note
[Online]. Available: shop.perkinelmer.com/content/technicalinfo/tch_ftiratr.pdf [accessed April 30, 2015].
Petersen, J.C., Chemical Composition of Asphalt as Related to Asphalt Durability: State of the Art, Transportation Research
Record 999, Transportation Research Board, National Research Council, Washington, D.C., 1984, pp. 1330.
Petersen, J.C., Quantitative Functional Group Analysis of Asphalts Using Differential Infrared Spectrometry and Selective
Chemical ReactionsTheory and Application, Transportation Research Record 1096, Transportation Research Board,
National Research Council, Washington, D.C., 1986, pp. 111.
Polacco, G., J. Stastna, Z. Vlachovicova, D. Biondi, and L. Zanzotto, Temporary Networks in Polymer-Modified Asphalt,
Polymer Engineering and Science, Vol. 44, No. 12, 2004, pp. 21852193.
Price, R.P. and J.L. Burati, A Quantitative Method Using HPGPC to Predict Laboratory Results of Asphalt Cement Tests
(with discussion), 1989 [Online]. Available: https://trid.trb.org/view/486917.
Punith, V.S. and A. Veeraragavan, Behavior of Asphalt Concrete Mixtures with Reclaimed Polyethylene as Additive, Jour-
nal of Materials in Civil Engineering, Vol. 19, No. 6, 2007, pp. 500507.
Quintana, H.A.R., J.A.H. Noguera, and C.F.U. Bonells, Behavior of Gilsonite-Modified Hot Mix Asphalt by Wet And Dry
Processes, Journal of Materials in Civil Engineering, Vol. 28, No. 2, 2016, Content ID 04015114.
Raki, L., J.F. Masson, and P. Collins, Rapid Bulk Fractionation of Maltenes Into Saturates, Aromatics, and Resins by Flash
Chromatography, Energy & Fuels, Vol. 14, No. 1, 2000, pp. 160163.
Ramm, A., M.C. Downer, N. Sakib, and A. Bhasin, Optical Characterization of Temperature- and Composition-Dependent
Microstructure in Asphalt Binders, Journal of Microscopy, Vol. 262, No. 3, 2016, pp. 216225.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

80

Ramsey, J.W., F.R. McDonald, and J.C. Petersen, Structural Study of Asphalts by Nuclear Magnetic Resonance, Industrial
and Engineering Chemistry Product Research and Development, Vol. 6, No. 4, 1967, pp. 231236 [Online]. Available:
https://trid.trb.org/view/97922.
Raouf, M.A. and R.C. Williams, Temperature and Shear Susceptibility of a Nonpetroleum Binder as a Pavement Material,
Transportation Research Record: Journal of the Transportation Research Board, No. 2180, Transportation Research
Board of the National Academies, Washington, D.C., 2010, pp. 918.
Rebelo, L.M., et al., Micromorphology and Microrheology of Modified Bitumen by Atomic Force Microscopy, Road Mate-
rials and Pavement Design, Vol. 15, No. 2, 2014, pp. 300311.
Redelius, P. and H. Soenen, Relation Between Bitumen Chemistry and Performance, Fuel, Vol. 140, 2015, pp. 3443.
Roberts, F.L., P.S. Kandhal, E.R. Brown, D.Y. Lee, and T.W. Kennedy, Hot Mix Asphalt Materials, Mixture Design and Con-
struction, National Asphalt Pavement Association, Lanham, Md., 2nd ed., 2009, 585 pp.
Rostler, F. and R. White, American Society for Testing and Materials, Special Technical Publication, Vol. 277, 1959, pp.
6488.
Rostler, F. and R. White, Fractional Components of Asphalt-Modification of the Asphaltenes Fraction, Proceedings of the
Association of Asphalt Paving Technologists, Vol. 39, No. 1970, pp. 532574.
Saboo, N., Optimum Blending Requirements for EVA Modified Binder, International Journal of Pavement Research and
Technology, Vol. 8, No. 3, 2015, pp. 172178.
Salehi, E., J. Abedi, and T. Harding, Bio-oil from Sawdust: Effect of Operating Parameters on the Yield and Quality of
Pyrolysis Products, Energy Fuels, Vol. 25, No. 9, 2011, pp. 41454154.
Schabron, J.E. and J.E. Rovani, The University of Wyoming Research Corporation, U.S. 8,367,425 B1, 2013.
Schmets, A., N. Kringos, T. Pauli, P. Redelius, and T. Scarpas, On the Existence of Wax-induced Phase Separation in Bitu-
men, International Journal of Pavement Engineering, Vol. 11, No. 6, 2010, pp. 555563.
Schuler, B., G. Meyer, D. Pena, O.C. Mullins, and L. Gross, Unraveling the Molecular Structures of Asphaltenes by Atomic
Force Microscopy, Journal of the American Chemical Society, Vol. 137, No. 31, 2015, pp. 98709876.
Seidel, J.C. and J.E. Haddock, Soy Fatty Acids as Sustainable Modifier for Asphalt Binders, In Transportation Research
E-Circular E-C165, Alternative Binders, 2012, pp. 1522.
Shen, J., S. Amirkhanian, and J.A. Miller, Effects of Rejuvenating Agents on Superpave Mixtures Containing Reclaimed
Asphalt Pavement, Journal of Materials in Civil Engineering, Vol. 19, No. 5, 2007a, pp. 376384.
Shen, J., S.N. Amirkhanian, and S.-J. Lee, HP-GPC Characterization of Rejuvenated Aged Bcrm Binders, Journal of Mate-
rials in Civil Engineering, Vol. 19, No. 6, 2007b, pp. 515522.
Shen, J., S. Amirkhanian, and F. Xiao, High Pressure Gel Permeation Chromatography of Aging of Recycled Crumb Rubber-
Modified Binders with Rejuvenating Agents, Transportation Research Record, 2006.
Shen, J., S.N. Amirkhanian, F. Xiao, and B. Tang, Surface Area of Crumb Rubber Modifier and Its Influence on High-
Temperature Viscosity of CRM Binders, International Journal of Pavement Engineering, Vol. 10, No. 5, 2009, pp.
375381.
Shen, J., M. Earnest, and Z. Xie, High Temperature Properties of Recycled Polyethylene Terephthalate (PET) Modified
Asphalt Binders, 2016a, 14 pp. [Online]. Available: https://trid.trb.org/view/1393177.
Shen, J., M. Earnest, and Z. Xie, Selected Properties of Hot Mix Asphalt Modified with Recycled Polyethylene Terephthalate,
2016b [Online]. Available: https://trid.trb.org/view/1393152.
SHRP, Designations MPI and TP5, Provisional Standard Specification for Performance Graded Asphalt Binder Based on
Strategic Highway Research Program, SHRP Product 1001, AASHTO, 1993a.
SHRP, Background of SHRP Asphalt Binder Test Methods, National Asphalt Training Center, Demonstration Project 101,
Chapter VI, 1993.
Shurvell, H.F., S. Subramani, S.A.M. Hesp, S.N. Genin, and D. Scafe, Five Year Performance Review of a Northern Ontario
Pavement Trial: Validation of Ontarios Double-Edge-Notched Tension (DENT) and Extended Bending Beam Rheometer
(BRR) Test Methods, Polyscience Publications, Laval, QC, Canada, 2009 [Online]. Available: https://trid.trb.org/
view/911715.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 81

Siddiqui, M.N., NMR Fingerprinting of Chemical Changes in Asphalt Fractions on Oxidation, Petroleum Science and
Technology, Vol. 28, No. 4, 2010, pp. 401411.
Siddiqui, M.N. and M.F. Ali, Investigation of Chemical Transformations by NMR and GPC During the Laboratory Aging of
Arabian Asphalt, Fuel, Vol. 78, No. 12, 1999, pp. 14071416.
Snyder, L.R, Determination of Asphalt Molecular Weight Distributions by Gel Permeation Chromatography, Analytical
Chemistry, Vol. 41, No. 10, 1969, pp. 12231227.
Soenen, H. and P. Redelius, The Effect of Aromatic Interactions on the Elasticity of Bituminous Binders, Rheologica Acta,
Vol. 53, No. 9, 2014, pp. 741754.
Speight, J.G., Petroleum Asphaltenes. Part 1. Asphaltenes, Resins and the Structure of Petroleum, Oil & Gas Science and
Technology, Vol. 59, No. 5, 2004, pp. 467477.
Strausz, O.P., T.W. Mojelsky, and E.M. Lown, The Molecular Structure of Asphaltene: An Unfolding Study, Fuel, Vol. 71,
No. 12, 1992, pp. 13551363.
Strausz, O.P., I. Safarik, E.M. Lown, and A. Morales-Izquierdo, A Critique of Asphaltene Fluorescence Decay and Depolar-
ization-Based Claims About Molecular Weight and Molecular Architecture, Energy Fuels, Vol. 22, No. 2, 2008, pp.
11561166.
Stroup-Gardiner, M., NCHRP Synthesis 495: Use of Reclaimed Asphalt Pavement and Recycled Asphalt Shingles in Asphalt
Mixtures, Transportation Research Board of the National Academies, Washington, D.C., 2016, 133 pp.
Stroup-Gardiner, M. and T. Wattenberg-Komas, NCHRP Synthesis 435: Recycled Materials and Byproducts in Highway
Applications, Volume 6: Reclaimed Asphalt Pavement, Recycled Concrete Aggregate, and Construction Demolition Waste,
Transportation Research Board of the National Academies, Washington, D.C., 2013, 82 pp.
Such, C., B. Brule, and C. Baluja-Santos, Characterization of a Road Asphalt by Chromatographic Techniques (GPC and
HPLC), Journal of Liquid Chromatography, Vol. 2, No. 3, 1979, pp. 437453.
Sun, D.Q. and L.W. Zhang, A Quantitative Determination of Polymer Content in SBS Modified Asphalt. Part I: State of the
Art, Petroleum Science and Technology, Vol. 31, No. 24, 2013, pp. 26362642.
Tabib, S., F. Hoque, and P. Marks, Ontarios Strategy to Enhance Asphalt Cement Quality. Sixtieth Annual Conference of the
Canadian Technical Asphalt Association, Winnipeg, MB, Canada, 2015 [Online]. Available: https://trid.trb.org/
view/1399726.
Tang, B. and U. Isacsson, Determination of Aromatic Hydrocarbons in Asphalt Release Agents Using Headspace Solid-
Phase Microextraction and Gas Chromatography Mass Spectrometry, Journal of Chromatography A, Vol. 1069, No. 2,
2005, pp. 235244.
Tarefder, R.A. and A.M. Zaman, Nanoscale Evaluation of Moisture Damage in Polymer Modified Asphalts, Journal of
Materials in Civil Engineering, Vol. 22, No. 7, 2010, pp. 714725.
Terrel, R.L. and J.A. Epps, Using Additives and Modifiers in Hot Mix Asphalt, Section A, 1989a, v.p. [Online]. Available:
https://trid.trb.org/view/361507.
Terrel, R.L. and J.A. Epps, Using Additives and Modifiers in Hot Mix Asphalt, Sections B, C, D, and E, 1989b, v.p. [Online].
Available: https://trid.trb.org/view/363790.
Traxler, R.N. and H.E. Schweyer, Separating Asphalt Materials-Butanol-Acetone Method, Oil & Gas Journal, Vol. 52, No.
17, 1953, p. 133.
Virginia Department of Transportation, 2016 VDOT Road and Bridge Specications, 2016.
Wahhab, H.I.A.-A., I.M. Asi, and I.A. Al-Dubabi, Prediction of Asphalt Rheological Properties Using HP-GPC, Journal of
Materials in Civil Engineering, Vol. 11, No. 1, 1999, pp. 614.
Wahhab, H.I.A., M.F. Ali, I.M. Asi, and I.A. Dubabe, HP-GPC Characterization of Asphalt and Modified Asphalts from Gulf
Countries and Their Relation to Performance Based Properties, Preprints Division of Fuel Chemistry, Vol. 41, 1996, pp.
12961301.
Watson, D., J.R. Moore, A.J. Taylor, and P. Wu, Effectiveness of Antistrip Agents in Asphalt Mixtures, Transportation
Research Record: Journal of the Transportation Research Board, No. 2370, Transportation Research Board of the National
Academies, Washington, D.C., 2013, pp. 128136.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

82

Welborn, J.Y., Physical Properties as Related Asphalt Durability: State of the Art, Transportation Research Record 999,
Transportation Research Board, National Research Council, Washington, D.C., 1984, pp. 3136.
Wen, H., S. Bhusal, and B. Wen, Laboratory Evaluation of Waste Cooking Oil-Based Bioasphalt as an Alternative Binder for
Hot Mix Asphalt, Journal of Materials in Civil Engineering, Vol. 25, No. 10, 2013, pp. 14321437.
White, R.M., W.R. Mitten, and J.B. Skog, Fractional-Components of Asphalt Compatibility and Interchangeability of Frac-
tions Produced from Different Asphalts, Proceedings of the Association of Asphalt Paving Technologists, Vol. 39, No.
1970, pp. 492531.
Williams, R.C. and N.S. McCready, The Utilization of Agriculturally Derived Lignin as an Antioxidant in Asphalt Binder,
2008, 86 pp. [Online]. Available: http://publications.iowa.gov/id/eprint/19987 and http://www.ctre.iastate.edu/pubs/
midcon2007/McCreadyLignin.pdf.
Williams, R.C., J. Satrio, M. Rover, R.C. Brown, and S. Teng, Utilization of Fractionated Bio Oil in Asphalt, Iowa State Uni-
versity, 2009, 19 pp. [Online]. Available: https://trid.trb.org/view/882150.
Xiao, F., S.N. Amirkhanian, and J. Shen, Effects of Various Long-Term Aging Procedures on the Rheological Properties of
Laboratory Prepared Rubberized Asphalt, ASTM Journal of Testing and Evaluation, Vol. 37, 2009, pp. 129138.
Yang, X., Z. You, and Q. Dai, Performance Evaluation of Asphalt Binder Modified by Bio-Oil Generated from Waste Wood
Resources, International Journal of Pavement Research and Technology, Vol. 6, No. 4, 2013, pp. 431439.
Yang, X., Z. You, and J. Mills-Beale, Asphalt Binders Blended with a High Percentage of Biobinders: Aging Mechanism
Using FTIR and Rheology, Journal of Materials in Civil Engineering, Vol. 27, No. 4, 2015, pp.
04014157/0401415104014157/04014159.
Yapp, M.T., A.Z. Durrani, and F.N. Finn, HP-GPC and Asphalt Characterization Literature Review, Strategic Highway
Research Program SHRP-A/UIR-91-503, Transportation Research Board, National Research Council, Washington, D.C.,
1991.
Yarranton, H.W., et al., On the Size Distribution of Self-Associated Asphaltenes, Energy Fuels, Vol. 27, No. 9, 2013, pp.
50835106.
Yaya, A.S., D. Oldham, S. Hosseinnezhad, S. Aflaki, and E.H. Fini, Physiochemical and Rheological Characterization of
Biomodified Gilsonite, 95th Annual Meeting of the Transportation Research Board, Washington, D.C., Jan. 2016 [Online].
Available: https://trid.trb.org/view/1394191.
Yen, T.F., The Colloidal Aspect of a Macrostructure of Petroleum Asphalt, Fuel Science & Technology International, Vol.
10, No. 4, 1992, pp. 723733.
Yu, X., N.A. Burnham, and M. Tao, Surface Microstructure of Bitumen Characterized by Atomic Force Microscopy,
Advances in Colloid and Interface Science, Vol. 218, 2015, pp. 1733.
Yuhong Wang, P.E., et al., Effects of Aging on the Properties of Asphalt at the Nanoscale, Construction and Building Mate-
rials, Vol. 80, 2015, pp. 244254.
Yut, I. and A. Zofka, Attenuated Total Reflection (ATR) Fourier Transform Infrared (FT-IR) Spectroscopy of Oxidized
Polymer-Modified Bitumens, Applied Spectroscopy, Vol. 65, No. 7, 2011, pp. 765770.
Zhao, S., et al. Investigation on the Microstructure of Recycled Asphalt Shingle Binder and its Blending with Virgin Bitu-
men, Road Materials and Pavement Design, Vol. 16, Suppl. 1, 2015, pp. 2138.
Zhou, F., C. Estakhri, and T. Scullion, Literature Review, Performance of RAP/RAS Mixes and New Direction, 2014, 69 pp.
[Online]. Available: http://tti.tamu.edu/documents/0-6738-1.pdf and https://trid.trb.org/view/1308471.
Zofka, A., et al., Evaluating Applications of Field Spectroscopy Devices to Fingerprint Commonly Used Construction Materi-
als, SHRP 2 Report Transportation Research Board of the National Academies, Washington, D.C., 2013, 263 pp.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 83

APPENDIX A
Survey Questions and Results
Question 1: Do you know the crude source of oil used to produce the asphalt binder and refining process employed? If yes, please
identify the source.

FIGURE A1 Respondents Familiar with Crude Oil Source

Additional Responses to Question 1, Do You Know the Crude Source of Oil Used to Produce the Asphalt Binder and Refining
Process Employed?

Response
A majority of the crude refined in our area is Canadian
Bitumen, Axeon
This is not something typically reported to VDOT.
We do not manufacture PG Binder
We do not specify crude sources be identified.
We have a list of two pages of crude oil sources.
We have lots of different sources of asphalt with as many different sources.
Varies by supplier
Our binder suppliers may be willing to answer this question and many of the following survey questions.
CDOT is not involved with the production and formulation of the binders we use, and, with a few excep-
tions, simply test according to the PG specifications in AASHTO M320. All of our binder suppliers are
listed on CDOT's Approved Products List available online here: http://apps.coloradodot.info/apl/
SearchRpt.cfm?cid=Asphalt. I have tried to answer the survey questions to the best of my knowledge
from the agency side, but I would suggest contacting the suppliers directly for more information.
The refineries buy crude on the open market and sources change. Also the refineries will buy asphalt on
the open market and blend it in with what they produce.
In some cases, we do, but our intent is that the oil meet specifications testing, and so where it comes from
falls toward non-essential information.
It is mostly from Western Canada (heavy crude from Cold Lake and light crude from Alberta) but there
could also be blend of light oil from states, Venezuela, Saudi, etc.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

84

Question 2: What refining process was used to produce the asphalt?

FIGURE A2 Refining Processed Used to Produce Asphalt

TABLE A2 Survey Response to Question 2, What Refining Process Was Used to Produce the Asphalt?

Process Response Rate Responses


Distillation 32.6% 15: CA, CO, DE, FL, GA, IL, KY, MN,
MO, ND, OK, SC, SD, TN, TX
Residuum Oil Supercritical
Extraction (ROSE)
Continuous Air-Blowing 2.2% 1 ND
Blending 19.6% 9: CA, DE, GA, MO, NC, ND, OK, SC, SD
Other 2.2% 1 NE
Do not know 67.4% 31
Total 46

Additional Responses to Question 2, What Refining Process Was Used to Produce the Asphalt?

Response
Air-blowing is not allowed in our spec.
As an agency we do not know about the crude or refining process
Blending has been done in the past but it is rare.
Contact binder suppliers for more information.
Multiple manufacturers, likely multiple methods
Varies by supplier
We do not manufacture PG Binder
We have a performance specification that they meet.
degrees of modification. Also, air blowing is not allowed unless specifically ordered as such.
depends on suppliers
there are several asphalt suppliers in Ontario. Many of them do blending of different grades.
Basically it is how the refinery is set up and whether they are running gasoline or heating oil.
I am sure there are other processes employed, but from an agency perspective we do not ask or specify.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 85

Question 3: Have you observed significant differences in the asphalt binder PG properties derived from your oil feedstock in the
past 10 years?

FIGURE A3 Survey Responses to Question 3

Additional Responses to Question 3, Have You Observed Significant Differences in the Asphalt Binder PG Properties Derived
from Your Oil Feedstock in the Past 10 Years?

Response
Changes in testing requirements have caused changes in the asphalt binder
I've only been employed for 4 years and have not seen significant differences
Less sticky/M320 results are often close to the lower limit of the acceptable range (cold end)
Stiffer binders; varying penetration values, more sources than ever being used.
They seem to have less goodies than before.
We do not refine.
We dont think we see major differences for last 5-6 years.
higher loss on heating values
sulfur added to some binders
I haven't noted significant changes, as materials are meeting specifications. At least one refinery seems to
have issues at times with blending.
Testing is similarit meets the specs, but there are some pavements that seem to have prematurely aged.
There have been instances where a few projects have experienced premature oxidation (visual
observation).
They continue to meet the specification so it can't be that different. Our spec has not changed much dur-
ing that time.
we are the agency. We have noticed premature cracking on our highways and believe that PG quality is a
major player.
We recently completed a survey and have an Excel file with BBR data that might show cold temperature
property trends.
We test the product after it has been processed to meet our specifications, so from the agency's viewpoint,
the product does not change.
None significant, but most of the grades we use today are different from what we used 10 years ago, so it
is not an apples to apples comparison.
The physical properties change and one can see it sometimes quarterly as feeds change. Also in 15 years
the binders have gotten stiffer, poorer low temps.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

86

Question 4: Are the asphalt binder PG properties varying with seasonal or market fluctuations?

FIGURE A4 Survey Responses to Question 4

Question 5: What rheological and analytical procedures are you employing in your laboratory for asphalt binder testing? Please
mark all that apply.

TABLE A5 Rheological and Analytical Procedures Employed

Procedure Response Rate Respondents


Rheologyperformance grade (PG) test 100% 46
protocol
Iatroscan SARA analysis 0.0% 0
Fourier transform infrared (FTIR) 23.9% 11: AL, AR, FL, GA, IL, MN, MS, TN,
TX, VA, WY
Gel permeation chromatography (GPC) 4.4% 2: LA, VA
Functional group analysis 0.0% 0
X-ray fluorescence (XRF) 19.6% 9: CO, FL, IL, MA, MN, NC, SC, VT, VA
Other (Please specify) 2.2% 1: FL

Question 6: Are you conducting any of the following binder fractionation procedures? If yes, please describe what triggered the
decision to conduct this test.

TABLE A6 Binder Fractionation Procedures Employed

Procedure Response Rate Respondents


Selective precipitation 0.0% 0
Column fractionation 0.0% 0
Iatroscan SARA analysis 2.2% 1 NE
Gel permeation chromatography 4.4% 2: LA, VA
Compatibility/phase separation 0.0% 0
Do not fractionate 23.9% 11: AZ, DC, FL, GA, ME, MA, MI, NC, OR,
PA, TN
No answer 69.6% 32
Total 46

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 87

Additional Responses to Question 6, Are You Conducting Any of the Following Binder Fractionation Procedures?

Response
CDOT is not, but suppliers may be.
GPC, Others (FTIR) if needed (for research or forensic investigation). Not regularly.
No binder fractionation procedures are being conducted.
This is beyond the DOT function.
We are not doing any fractionation.
We do not perform any of the above.
We do not refine.
We have only done it for a research project to review the effects of rejuvenators on binders
We do not fractionate the binder since it is a moving target. Years ago I worked on TLC as a quick test.
Research facility (LTRC) developed a method using GPC to determine polymer percent in modified
asphalt and examine asphaltene/maltene ratios.
We have uncovered asphalt reaction with preservative treated woods resulting in liquefaction of bitumen.
Further chemical investigation is planned.

Question 7: Are you working on new procedures for separating various components of the asphalt binder? If yes, please describe
your new procedure

TABLE A7 Respondents Working on New Procedures for Separating Various Components of the Asphalt Binder

Response Response Rate Respondents


Yes 4.4% 2 LA, VA
No 95.7% 44
Total 46

Additional Responses to Question 7, Are You Working on New Procedures for Separating Various Components of the Asphalt
Binder?

Response
Contact binder suppliers for more information.
I do have a process for separating and collecting asphaltenes, but no other component.
Unknown
We have an x-ray procedure to identify REOB and PPA presence and quantity.
VTRC is utilizing GPC to study the effects of oxidation on binders as well as utilizing it as a potential
forensic tool for VDOT.

Question 8: What procedures are you using routinely to estimate asphalt binder aging characteristics?

TABLE A8 Procedures Used Routinely to Estimate Asphalt Binder Aging Characteristics

Procedure Response Rate Responses


TFOT, RTFO 100.0% 46
PAV 100.0% 46
Multiple PAVs 0.0% 0
Penetration index 4.4% 2: AL, AR
Oven aging/phase separation 2.2% 1: AL
Viscosity temperature coefficient 2.2% 1: AL
Other 0.0% 0
Total 46

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

88

Additional Responses to Question 8, What Procedures Are You Using Routinely to Estimate Asphalt Binder Aging Characteristics?

Response
CDOT uses the standard RTFO and PAV aging procedures in AASHTO M320.
Per Superpave Binder Specification
These are the conditioning requirements to follow M320.
We perform RTFO and PAV test for acceptance according to frequency guide.
We use Extended BBR test for evaluating effect of physical hardening
There are upper and lower limits on viscosity for the binder Abson recovered from RAP materials in our
spec.

Question 9: What procedures are you using occasionally to estimate asphalt binder aging characteristics? What conditions initiate
the additional investigation?

TABLE A9 Procedures Used Occasionally to Estimate Asphalt Binder Aging Characteristics

Procedure Response Rate Responses


TFOT, RTFO 40% 17
PAV 40% 17
Multiple PAVs 10.9% 6: FL, IL, IW, MA, MN, NE
Penetration idex 4.4% 2: AL, AR
Oven aging/phase separation 2.2% 1: AL
Viscosity temperature coefficient 2.2% 1 AL
Tc, fracture energy 2.2% 1: FL
Glover-Rowe 4.4% 2: FL, OK
Do not test 10.9% 6: CO, MI, MS, NV, PA, VA
No answer 39.1% 18
Total 46

Additional Responses to Question 9, What Procedures Are You Using Occasionally to Estimate Asphalt Binder Aging
Characteristics?

Response
A 3rd has looked at Glover-Rowe parameter when using rejuvenators in research projects.
CDOT only uses the routine tests to estimate asphalt binder aging.
FDOT is also looking some of new properties such as Delta Tc, Fracture Energy, Glover-Rowe, etc.
FTIR for research
No additional procedures are currently being used beyond what is used routinely.
No additional testing is performed to estimate binder aging characteristics.
Penetration index is part of our specification for some AC binders.
The same procedures as the ones using routinely.
We are starting to look into it on our rejuvenator project.
We perform the standard PG testing for quality assurance testing.
none. research done by a sub looking at multiple PAVs
If by multiple PAVs it is meant that extended PAV aging is done on an additional sample, this has been
considered/discussed in relation to asphalt extenders.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 89

Question 10: Are you employing any of the following modifiers/additives in final binder preparation to achieve specific asphalt
binder grades?

TABLE A10 Modifiers/Additives Used in Final Binder Preparation to Achieve Specific Asphalt Binder Grades

Additive Response Rate Responses


Softening agents 21.7% 10: FL, LA, ME, MA, MO, NH, NC,
OK, PA, WI
Antistripping agents 26.1% 12: AR, DE, DC, KS, LA, ME, MN,
OK, SC, TN, TX, WI
Rejuvenators 10.9% 6: FL, KS, LA, MO, NE, OK
Wax-based additives 15.2% 7: DE, FL, ME, PA, TN, TX, WI
Polyphosphoric acid (PPA) 41.3% 19
Antioxidants 0.0% 0
Other 17.4% 8: AZ, KY, MI, UT, VT, VA, WV, WI
No answer 32.6% 15

Additional Responses to Question 10, Are You Employing Any of the Following Modifiers/Additives in Final Binder Preparation
to Achieve Specific Asphalt Binder Grades?

Response
As a user, we are not blending asphalts. We know suppliers are using some of these.
Depends on suppliers
N/A
No air blown or PPA allowed unless approved by agency
No, the supplier controls these materials.
Softening agents and rejuvenators are only used for research projects. One supplier uses PPA.
Specification call for use of SBS only 64-22 to 76-22
Suppliers using polymers as well
This more suitable for terminals. We do testing for acceptance of project material.
To our knowledge these items are being used by suppliers in this area.
Unknownquestion for binder suppliers
We allow PPA up to 1.0%
We do not manufacture PG Binder
We do not refine.
We don't typically require modifiers but some suppliers use them to meet specifications.
We're not employing the additives, the suppliers are.
manufacturers are not required to disclose
maybe others, not sure
The refineries make the AC to meet the specified grade. They have used PPA and will back blend with
lighter stock. Compaction aids (waxes, wetting agents & antistrip) are paving related and dictated by the
job, aggregate and plant.
WMA Technologies (organic/wax based, chemical based (typically low water surfactant additive), water.
The DOT does not prepare binders for use on projects.
Missouri allows antistripping agents and wax-based additives (warm mix) but they are not used to
achieve the desired binder grade. We have observed that some antistripping agent can affect the binder
grade.
We are looking into rejuvenators to achieve specific grades, and do allow PPA (at a max of 0.5% by
weight of binder).
The producers may be adding other additives to the binder to meet grade, but we only have specifications
requiring or limiting some of them.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

90

Response
we have performed a pilot project using Evoflex. We have used Evoflex with a high recycle asphalt and it
seemed to have graded back to a PG64-22. This was only on a trial basis and is not approved for full time
use at this time.
We don't actually manufacture the binders but I'm sure most of these are used by our suppliers.
OR DOT does not monitor binder for any of the additives above, although suppliers may be utilizing one
or more of the additives above.
We require anti-strip for moisture-resistance properties, not grade modification. Companies have experi-
mented with other modifiers (rejuvenators, softening agents, etc.).
Antistripping agents are commonly used to meet TSR specification requirement. Rejuvenators have been
used on 5 trial projects.
We require antistripping additive for northern Ontario. We allow a limited quantity of PPA. We use a lot
of PGXX-34 and some PGXX-40 therefore we know that softening agents (oils, etc.) are being used
Contact binder suppliers for more information pertaining to specific formulations. PPA is currently not
allowed in any binder used by CDOT.
We allow SB, SBS, or SBR as modifiers. We also allow antistrip additives, but they are not commonly
used. We suspect PPA is being used, but it isn't allowed by our specifications.
We do not use liquid anti-stripping agents for on-system pavements. PPA/acid is restricted with less than
0.5% in our spec.

Question 11: If you extracted binder from mixtures containing recycled materials, which mixtures have you examined? If yes, com-
ment on your results.

TABLE A11 Recycled Materials Extracted from Mixtures

Recycled Material Response Rate Responses


RAP 65.2% 30
RAS 47.8% 22
REOB 10.9% 6: FL, MI, NH, NC, OH, PA
Rejuvenators 15.2% 7: NM, FL, KS, MO, NE, NY, VA
RAP/RAS 4.4% 2: MO, NC
No answer 26.1% 12: AL, AR, CO, DC, IL, ME, MA,
MS, NV, NJ, UT, VT
Total 46

Additional Responses to Question 11, If You Extracted Binder from Mixtures Containing Recycled Materials, Which Mixtures
Have You Examined?

Response
Binder extraction and grading has been performed on the rejuvenator trial projects.
Binders sent out for spot checking of REOB
CDOT does not test extracted binder.
Difficult to obtain consistent grading, stockpiles of recycled materials change daily.
Have also performed extractions on mixtures using both RAP and RAS.
I hate post-consumer RAS but we have to use it.
No RAS allowed
PG binder grade testing
PPA
Performed by HMA Producer
RAS has been and rejuvenators are planned to be studied on a trial basis.
We are currently doing a research project analyzing extracted RAP binder and rejuvenators.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 91

Response
crumb rubber
Polymer and Rubber have also been extracted with the understanding that Rubber particulates do get
trapped in the filter paper.
Typically done as a forensic investigation when issues during field production and construction occur.
This work for RAS was due to legislative interest in allowing the use of RAS. Industry has also been
showing interest in increasing the use of recycled materials.
We haven't performed in-depth examination of extracted binder for recycled materials. We have used
GPC to examine field core samples to verify polymer presence.
We have extracted the asphalt from high recycle mixes mainly consisting of 9% RAP and 4% RAS to
check the final grade of the asphalt. We will be checking more of these mixes this summer.
Depending on the degree of aging, these mixes can be hard to distinguish from typical binder properties.
See: http://l92018.eos-intl.net/L92018/OPAC/Common/Pages/GetDoc.aspx?ClientID=ML92018&Media
Code=3073725

Question 12: If you are employing polymer additives, which type are you using?

TABLE A12 Polymer Additives Employed

Response Rate Responses


SBS 71.7% 33
Radial (r-SBS) 15.2% 7: AR, LA, ME, NV, NJ, OH, SD
(SBR) 26.1% 12: AR, CO, GA, KY, LA, ME, MI, NC, OH,
SD, TN, TX
Crumb rubber 30.4% 14: AZ, CA, GA, FL, GA, IL, LA, MO, NH,
OH, OK, PA, SC, VA
Polyethylene copolymers 0.0%
Other plastomers 6.5% 3: NY, PA, SC
Hybrid (GTR + SBS) 4.4% 2: AZ, FL
No answer 17.4% 9: AL, KS, MS, ND, OR, UT, VT, WI, WY
Total responses 46

Additional Responses to Question 12, If You Are Employing Polymer Additives, Which Type Are You Using?

Response
Contact binder suppliers for more information pertaining to specific formulations.
Limited tests on plastomersHoneywell Products
Not sure if SBS is linear or radial.
Not sure if it's linear or radial SBS
Our TR+ specifications require both SBS and Ground Rubber
SBS & SBR. We test final grade se we do not necessarily know PG of base asphalt.
Suppliers modify AC.
The refineries are allowed to use SBS and SBR, and some of the others but not crumb rubber
We do not manufacture PG Binders
We do not specify type of polymers. It is up to asphalt suppliers
We do not test for polymer additive types, only for PG qualification.
crumb rubber is allowed, but has not been used
Meeting our specification normally requires the use of SBS polymer. They are free to use what they want
to meet the specification.
Crumb rubber/wet blend rubber is being investigated on a trial basis. Polyethylene is allowed but not typi-
cally used.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

92

Response
We are working on a county project that will use GTR. The base grade varied from PG 58-28 to PG 64-22
with a goal of PG 70-22 or better using 5, 10, 20% GTR by weight of binder.
Other polymers are permitted in combination with the crumb rubber to produce a PG 76-22. Currently we
are not told which polymer.
We do not require a specific polymer, but require elastic recovery. No supplier of PG binders is currently
using SBR.
SBS (don't know what type), SBR, and Crumb rubber are being used. We specify meeting grades not spe-
cific polymers. We specify a PG 64V-28 using any additive to make it. We specify SBR at 2% to result in
a PG 70-28, Crumb Rubber is specified using ASTM D 6690
We don't restrict the types of polymers that can be used. Most asphalt-rubber binders are not PG graded.
Other = approved use of Elvaloy, but actual use is very limited if any at all. We specify SBS block copo-
lymers, but we have not specifically specified linear or radial, so we are unsure which type (linear or
radial) asphalt binder suppliers are using.
OR DOT uses AASHTO T301 with an elastic recovery requirement of 50% on RTFO conditioned binder.
OR DOT does not monitor type of polymer utilized.
Basically everyone in the area uses SBS polymers. The % varies with the crude. One plant makes a con-
centrate and blends with 70-22 and the other uses 64-22 as the base.

Question 13: If you are employing rejuvenators to restore aged asphalt binder, which type are you using?

TABLE A13 Rejuvenators Employed to Restore Aged Asphalt Binder

Rejuvenator Response Rate Responses


Abietic acid esters (Hydrogreen) 10.8% 5: KS, KY, MO, NE, NC
Aromatic oils (Cyclogen-L) 4.4% 2: NE, NY
Soft asphalt binders 10.8% 5: AZ, FL, MN, NC, WY
REOB 2.2% 1: FL, Banned in ME
Evoflex 8.7% 4: KS, KY, MO, OK
Soybean oil 2.2% 1: NE
Lubricating/extender oils 2.2% 1: CO
None 19.6% 9: DE, DC, GA, IL, IN, IW,
MA, NV, PA
No answer 47.8% 22

Additional Responses to Question 13, If You Are Employing Rejuvenators to Restore Aged Asphalt Binder, Which Type Are You
Using?

Response
Ingevity characterizes EVOFLEX CA as a fatty acid derivative.
Also utilizing Evoflex sold by Ingevity.
Choice would be up to producer but we ban recycled engine oil based products.
DNA
FDOT does not limit or specify the type of rejuvenator the producers may use.
Hydrogreen and Evoflex have been the options on the rejuvenator trial projects.
N/AREOB banned
No
None
Not sure if rejuvenators are being used and which ones, if used.
These are for specific purposes and projects.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 93

Response
We currently do not specify rejuvenators.
We do not manufacture PG Binders
We do not specify rejuvenators
We do not specify. I know softer binder and REOB are likely to be used
We have experimented with Hydrogreen and Evoflex but have not accepted them for approved use.
None
Rejuvenators must meet ASTM 4552 Recycling Agent
Rejuvenator agent is not allowed, except in hot in-place recycling and CIR projects (special provision
only).
The manufacturer of our ARA and ARA-1P rejuvenating emulsions, Ergon Asphalt & Emulsions, Inc.,
will be able to better answer this question.
Have not started employing yet, but in research project we are looking at soybean oil, aromatic oil, and
abietic acid esters.

Question 14: How do you quantify the presence of modifiers/additives in final asphalt binder composition?

TABLE 14 Procedures Used to Quantify the Presence of Modifiers/Additives in Final Asphalt

Viscosity FTIR GPC MSCR XRF Do Not Measure


Softening agents MO IL IL, NC, VT
Antistripping agents SC DC, MO
Rejuvenators MO
Polymers MO, SC, GA, IL, MS LA AZ, AR, KT, AR, CO, IL,
TN, TX LA, ME NE, NC, VT,
NE, NV, NC, WY
ND, PA, SC,
TN, TX, VT
Polyphosphoric acid (PPA) SC SC
Recycled materials SC, TN, TX IL, ME PA, IN, TX
N/A 28

Additional Responses to Question 14, How Do You Quantify the Presence of Modifiers/Additives in Final Asphalt Binder
Composition?

Response
4C Ductility and Toughness and Tenacity
AASHTO T 301 for polymers
CDOT only uses the XRF to verify the lack of any PPA modification in our binders.
Crumb rubber through certification. Antistripping agents measured on site.
FTIR not internally but through outside entities occasionally
Moisture-resistance testing on mix determines if adequate anti-strip additive is used.
Polymers are usually most characterized by elastic recovery test.
RecycledNot commonly tested, restricted by aged binder in mix design
We are beginning the process to start XRF testing
We may start using XRF for PPA and REOB detection.
We use + specsphase angle, softening point, elastic recovery
We use ER for polymer and PPA.
XRF, GPC, and FTIR are being looked at for these types of applications.
elastic recovery test for polymers
we have specifications for polymers

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

94

Response
Recycled materials such as RAP or RAS when blended with virgin binder, still need to meet the specified
final grade of PG binder.
Two PG+ tests: recovery and separation. Mandate amine-based antistrip addition during PQL and verifi-
cation to assure binder won't change PG grade.
We use AASHTO M 320 plus specification including elastic recovery, an original DSR phase angle
requirement, and direct tension.
We occasionally use wet chemistry for PPA. Amine anti-strips are easily detected with the nose. The ter-
minal BOL is required to list all additives and modifiers.
we test the final product for rheology. We have tried using XRF to detect PPA. It wont give accurate results
due to low dosage. We tried FTIR for antistripping agents. Similarly, the low dosage is hard to detect
Field TSR (QC & QA) are required for acceptance; this is a check on the presence of antistripping agents.
Daily binder samples are required. Samples are taken just prior to the binder entering the mixing drum
and contain all of the additives, if used. DSR samples are performed on these samples to confirm the PG
binder grade. The Missouri DOT has developed a test method for determining the amount of PPA used in
the asphalt binder. Use this test when PG binder grade is not confirmed through DSR testing.
Most of the additives, if requested, are based on a certified statement from the producer that they meet the
specification requirements.
We have limits on REOB and PPA. Otherwise, our spec does not require specific modifiers, except that
we have an elastic recovery requirement.

Question 15: Do you measure the compatibility between modifiers and asphalt final binder after storage or aging? Please cite the
measurement technique employed in the comments box.

TABLE A15 How Compatibility Between Modifiers and Asphalt Final Binder Is Measured

After Initial Blending After Storage After Aging Do Not Measure


Softening agents SC
Antistripping agents SC MO, SC MO, SC SD
Rejuvenators SD
Polymers FL, IL, KS, SC GA, IL, LA, OH, SC SC SD
Polyphosphoric acid (PPA) SC OH, OK, SC SC
Rubber FL
N/A 29

Additional Responses to Question 15, Do You Measure the Compatibility Between Modifiers and Asphalt Final Binder After Stor-
age or Aging?

Response
ER only
All binder must pass some type of stability/settlement test, regardless of composition.
Contact binder suppliers for more information pertaining to specific formulations.
No
Supplier is relied upon to confirm compatibility of their formulations.
We don't measure compatibility.
We may run separation of polymer test after 48 hours of oven-aging.
We sample and test the asphalt binder directly from the contractors HMA plant.
We use the PG+ separation test.
it is up to asphalt suppliers. We test the final product for compliance to our specifications.
Compatibility as previous noted between wooden bridge decks and asphalt overlays may become an
issue.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 95

Response
Missouri has one binder supplier that utilizes PPA. They indicate the use of PPA on the bill of lading and
advise the asphalt paving contractor they are supplying so they can account for the use of PPA in the mix
design. We have had a few jobs where the asphalt paving contractor was not aware of the presence of
PPA. The PPA neutralized the antistripping agent which negatively affected the field TSR. The PPA also
reacted with the polymer, negatively affected the PG binder.
The PMA and PPA asphalts are stable in storage and never had a problem in the lab or the HMA plants.
Our PG+ spec includes a separation tests which is a difference in DSR results. This is similar to the old
cigar tube softening point difference method.
After initial blending at terminal (source samples taken by supplier and DOT), after storage: samples
taken from contractor storage tanks, after aging using PAV in Binder Lab.
We rely on the asphalt binder suppliers to perform any compatibility testing with the specific crude
sources they are using and any binder additives they are using to meet specified requirements.
Polymer separation test (softening point difference) is performed for SBR and crumb rubber modified PG
76-22.
Polymer elasticity is based on phase angle and MSCR %Recovery. A separation < 15F is required for the
rubber binder (cigar tube test).
No. PG binder need to meet the grade specified (with our ER requirement according to grade). Antistrip
agents need to work to allow mix to meet Hamburg requirements.
binder QA samples taken at HMA production facility on the day of mix production and need to meet
spec, so this accounts for some storage time

Question 16: If using an additive, what specifications are used for the amount of additive for a given grade?

TABLE A16 Specifications Are Used for Additive Amounts

Upper Limit Defined Minimal Required Specified Level Do Not Specify Amount
Softening agents MN
Antistripping agents PA, SC, VA FL, NC, ND, SD, AR, DC
TN, TX
Rejuvenators SD TX
Polymers AZ, MI, NV CT, IL, OH, SC, TN
Polyphosphoric acid (PPA) CO, FL, GA, IA, MA,
MS, NE, PA, SC, TN,
TX, WY
Other (rubber) AZ, FL
N/A 16

Additional Responses to Question 16, If Using an Additive, What Specifications Are Used for the Amount of Additive for a Given
Grade?

Response
Acid/PPA shall be less than 0.5% of the binder content.
Anti-strip: 0.5% PPA: 0.75% Maximum Ground tire rubber: Min. 7%
Antistrip: 1% Lime. PPA: 0.50 Max
Antistripping agents are allowed but rarely used
Binder grade and/or mix tests determine amount of modifier used.
Maximum allowable for REOB and PPA
No PPA is allowed in CDOT binder.
Other: asphalt rubber by state specification
Upper limit for REOBs
We do not manufacture PG Binder
We rely on ER test.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

96

Response
Levels of use for the most part are not specified but when additives are used I assume suppliers are dosing
at a level of minimal needed for desired change.
Anti-strip amounts come from the HMA plant. Polymer/PPA is very dependent on the crude source and
the refinery determines the amount to meet the appropriate PG grade. Softening agents/rejuvenators the
refineries probably use in blending to meet the PG grade. They are not used at the HMA plant.
No amount is currently specified, XRF testing has indicated low presence of additives in MD supplied
binders.
We took two approaches during the rejuvenator trials: (1) specified level and (2) level defined by target
grade for blended binder, Hamburg test and TSRST on mixture.
Liquid antistripping agents are typically specified at a minimum level of 0.25% by weight of total binder
in asphalt mixture or whatever the recommended minimum dosage is from antistripping agent manufac-
turer. Polymer amount is determined by the asphalt binder supplier based on their crude stock, type of
polymer, and production to meet the specified M335 grade (PG64E-22). PPA is typically specified at a
maximum % of binder and is only specified as a polymer cross-linking enhancer or polymer catalyst, but
this varies based on the polymer. (SBS vs. Elvaloy)
Contractors have used Evotherm at a dosage rate typically of 0.5%. Evotherm has been used as an anti-
strip agent and warm mix additive.
Missouri focuses on the end result. They use whatever amount of antistripping agent is need to meet TSR
requirements. Binder acceptance is based on meeting PG binder specifications. The supplier determines
the amount of additives needed to comply with these requirements.

Question 17: Are current PG grade criteria satisfactory in identifying compositional/chemical properties? If no, suggest criteria
that should be added or changed.

FIGURE A17 Current PG Grade Criteria Are Satisfactory in Identifying Compositional/Chemical Properties

Additional Responses to Question 17, Are Current PG Grade Criteria Satisfactory in Identifying Compositional/Chemical
Properties?

Response
Are we completely aware with how different components could affect binder performance?
Current criteria are performance-focused. XRF may be logical criteria that could be added.
Do not address durability.
Do not use other additives in our binder
Extended BBR, Extended PAV and DENT test
FTIR
Need existing equipment to tell us (XRF, FTIR, rheology)
No opinion
No suggestion
Something to address adhesive properties of the asphalt binder
Unknown.
Unsure.
We are in early stages of in-house researching XRF.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 97

Response
would likely need XRF testing or something similar to determine chemical composition
The missing part is long-term aging. A one-point aging test like PAV does not tell us the whole story.
In addition to M320 compliance, we require ash content not exceeding 0.8%. We also have acceptance
criteria for MSCR, DENT, and ExBBR tests.
We need to look at the chemical composition of our binders. It is my opinion that the PG64-22's are not
the quality they used to be. This is based on the life of our asphalt pavements.
MSCR % recovery identifies elastomeric polymer, but MSCR recovery curves were developed using SBS
modifiers. Need additional curves for SBR, crumb-rubber, etc. if we are going to specify a % recovery for
a specific Jnr (even more important if using the actual Jnr obtained rather than a maximum for a grade).
We dont believe current PG grade criteria are satisfactorily identifying compositional/chemical proper-
ties. REOBs, PPA, etc. can go undetected in the PG grading system because it is purely rheologically
based. The jury is very much out on how much any of these modifiers affect the long-term performance of
the binder. It takes an actual examination of the chemistry to determine whether or not they are present.
This one potential benefit of looking at utilizing Gel Permeation Chromatography (which were VTRC is
doing) as well as employing our XRF at VDOT Materials to look at our binders.
Delta Tc seems to be a quick method of using current test data to potentially identify a potential issue
with low temperature cracking in asphalt binders. Any test should be quick and ideally based on current
existing/common equipment.
PG criteria have nothing to do with composition/chemical properties. PG criteria are physical criteria.
A precise quantitative test procedure for type and % of REOB needs to be developed and standardized.
We haven't had a problem using them to identify use of polymer, but trying to determine which additional
tests to run, looking at SARA, AFM, FTIR, etc.to better analyze the binder.
PG criteria currently seem to fail to capture some poorly performing binders. However, composition/
chemical properties are not of interest to us; we need good performance properties.
CDOT uses the Ductility test (AASHTO T-51), Toughness and Tenacity test (Colorado CP-L 2210), and
Elastic Recovery test (AASHTO T-301) as indicators for polymer modification on some grades of binder.
A quick, easy field identification of modifiers (polymer & rubber) needs to be developed to determine
what we are actually receiving on the job.
The PG system or newer MSCR system does not identify chemistry. Basically fore either asphalt is some-
thing that is black and sticky and has certain physical properties.
For the most part current criteria measure physical properties which can indicate composition in some
instances.
It is satisfactory for the binder composition, the issues we see are more after the mix is put together and
RAP is added.

Question 18: Based upon internal research/knowledge, which physical properties are most sensitive to changes in the crude stock
or binder chemistry?

TABLE A18 Physical Properties Most Sensitive to Changes in the Crude Stock or Binder Chemistry

Physical Property Response Rate Responses


Initial rheology 26.1% 12
Aged rheology 34.8% 16
Ductility 13.4% 6
Forced ductility 10.9% 5
Elastic recovery 13.4% 6
Temperature susceptibility 4.4% 2
Stiffness 23.9% 11
Creep compliance 29.6% 8
Other 8.7% 4
No answer 43.5% 20

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

98

Additional Responses to Question 18, Which Physical Properties Are Most Sensitive to Changes in the Crude Stock or Binder
Chemistry?

Response
CDOT has not conducted any internal research on changes in crude or the chemistry of the binder.
Don't have this knowledge within DOT
Have not altered binder chemistry in the products we use
It is up to our suppliers to know this.
More appropriate to terminals or researchers.
No opinion
Toughness and tenacity
We are not aware of when crude or binder chemistry changes.
We do not deal with crude stock changes
We do not manufacture PG binder
We do not refine so no data vs crude stock
Binder suppliers could probably give best answer to this question; we see final product
Unknown
Penn DOT does not know the crude stock of its asphalt binder suppliers or when or how often they may
change this crude stock.
Unknown. As an agency, we only specify for the end result. We do not get involved in the crude selection
and changes.
Asphalts using certain crude stocks (i.e., Mayan) may have difficulty meeting elastomeric requirements.
Summer/winter blends may change properties as well when a refinery produces asphalt as a side
product.
As the crude stock varies and the main production product varies, so does the binder properties. Every-
thing will meet a PG grade. However, the PG system was based on all pavements are made with 100%
virgin binder. Using RAP/RAS means that not all PG binders are suitable.
Delta-T may be related to ductility, stiffness, and creep compliance

Question 19: Is your laboratory developing alternate techniques to estimate binder performance? If yes, please summarize details

FIGURE A19 Laboratory Is Developing Alternate Techniques to Estimate Binder Performance

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 99

Additional Responses to Question 19, Is Your Laboratory Developing Alternate Techniques to Estimate Binder Performance?

Response
Illinois DOT needs the ability to test for improved aging performance. Considering increasing the PAV
aging time to identify poor performing asphalt. IDOT recently developed the IFIT SCB test to help iden-
tify potentially poor asphalt mixes.
Vermont DOT Is considering extended BBR, FTIR, DENT and DELTA Tc.
West Virginia is Implementing MSCR Testing
Michigan DOT is looking into MSCR
New York DOT is starting to look more at mix performance tests
Virginia DOT On-going research is studying GPCs potential use as a performance indicator.
Washington DOT is working toward implementation of AASHTO T 350 and M 332 including percent
recovery.
Florida DOT We are looking into the newer provisional standards, such as the LAS, tack test for the resi-
due on the emulsions, binder fracture energy test recently developed by University of Florida
Louisiana DOT LTRC research facility is our development arm for new techniques. We may participate
in other ongoing research (i.e., NCHRP).
Colorado DOT currently specifies the Ductility test (AASHTO T-51), Toughness and Tenacity test (Colo-
rado CP-L 2210), and Elastic Recovery test (AASHTO T-301) as indicators for polymer modification on
some grades of binder.
Ontario MOT We have already implemented the DENT test (Double Edge Notch Tension Test-which is
an AASHTO test now), MSCR, and Ash content (to limit REOB). We are phasing-in the ExBBR test
(which is also an AASHTO Provisional test now). We are looking at XRF and FTIR tests to quantify
REOB. We have also developed modified PAV procedure which is being evaluated at this moment.

Question 20: What are your states critical performance indicators for assuring good high-temperature performance?

TABLE A20 Critical Performance Indicators for Assuring Good High-Temperature Performance

Indicator Response Rate Responses


G*sin , RTFO 30.4% 14
DSR, RV 8.7% 4: DC, FL, LA, NY
ASSHTO M320 21.7% 10
MSCR 13.0% 6: FL, LA, ME, NE, ND,
WY
APA, Hamburg 13.0% 5: GA, IL, KS, SD, WI
MSCR + PG+ 4.3% 2: TN, TX
No answer 13.05 6: AL, CT, IN, KY, NJ, NM

Additional Responses to Question 20, What Are Your States Critical Performance Indicators for Assuring Good High-Tempera-
ture Performance?

Response
PG high temp dynamic shear
Binder complies with PG binder requirements (AASHTO M 320).
Binders must pass high temperature PG grading.
DSR G*/sin value equal to or greater than 1.0 at the specified test temperature.
DSR on original and on RTFO binders (Phase angle, MSCR, etc.)
Dynamic Shear
G* and phase angle.
G*/sin delta APA and Hamburg wheel tester for mixture
G*/sin delta and the MSCR test

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

100

Response
Grade choice is traffic dependent and we use either the PG or MSCR system to determine the initial
binder grade.
High temperature DSR meeting some minimum threshold.
Initial DSR phase angle (replaced force-ductility) for presence of polymer, MSCR Jnr and % recovery
(replaced elastic recovery).
Jnr
MSCR, original DSR and RTFO DSR.
Meet PG and mix meets Hamburg
Moving to performance testing of mixes, Hamburg wheel. Binder onlyDSR/Jnr
Original & RTFO DSR and elastic recovery
Original DSR and MSCR
PG grade upper limit, ER, rut testing.
Penetration, ductility, viscosity and Marshall stability on asphalt concrete mix.
Per the QC/QA asphalt binder program for NC, true grading of the binder is required for the high end
DSRO, TRFO by the producer's AMRL lab. NCDOT also true grades the binder for samples received.
This has been identified as an area of research for northern climates.
Upper grading temperature on original and short-term aged binder (58, 64 or 70C depending of the area
and type of road).
We are an M320 state with Direct Tension requirements deleted.
We do not have rutting problem in Ontario. We rely on M320 results as well as MSCR compliance.
rutting: agg properties, mix volumetrics, binder grads (on high end) and Hamburg (on some projects)
we do not perform performance testing.

Question 21: What are your states critical performance indicators for assuring good intermediate-temperature performance?

FIGURE A21 Critical Performance Indicators for Assuring Good Intermediate-Temperature Performance

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 101

TABLE A21 Critical Performance Indicators for Assuring Good Intermediate-Temperature Performance

Indicator Response Rate Responses


G*sin, PAVt 28.3% 13
ER 8.7% 4: CO, KS, OH, UT
ASSHTO M320 15.2% 7: MO, NH, NY, NC, OK, OR, PA
MSCR 13.0% 6: FL, LA, ME, NE, ND, WY
MSCR + PG+ 4.3% 2: TN, TX
Beam fatigue, AMPT 4.3% 2: GA, KS
SCB 2.2% 1: WI
No answer 30.4% 14

Additional Responses to Question 21, What Are Your States Critical Performance Indicators for Assuring Good Intermediate-
Temperature Performance?

Response
Binders must pass intermediate temperature PG grading.
DSR, BBR
Elastic Recovery, Toughness and Tenacity, and PAV G*/sin(delta).
Elastic recovery and PAV DSR.
Fatigue cracking: binder elastic recovery and G*sin delta
G* sin delta No spec currently for this
G*sin delta Beam fatigue testing AMPT
G*sin delta at PAV
Intermediate DSR.
Intermediate temperature DSR maximum, but our spec ties it heavily to the low temperature grade.
M320 and DENT
Moving to performance testing of mixes, Hamburg wheel, SCB. Binder onlyDSR/Jnr
Penetration, ductility, viscosity and Marshall stability on asphalt concrete mix.
RTFO DSR > 2.2
There are no reliable tests out there to date. We are investigating some currently, but per PG Binder Spec-
ification, we perform PAV DSR.
We have an overlay test requirement on some mixtures (those that are designed to be more flexible).
unknown
We do not perform performance testing.

Question 22: What are your states critical performance indicators for assuring good low-temperature performance?

TABLE A22 Critical Performance Indicators for Assuring Good Low-Temperature Performance

Indicator Response Rate Responses


BBR 34.8% 26
DSR 8.7% 4: DC, ME, MA, MN
ASSHTO M320/M322 15.2% 7: MO, NH, NY, NC, OK, OR, PA
MSCR + PG 13.0% 6: FL, LA, ME, NE, ND, WY
No answer 17.4% 7: AL, CA, IL, IN, KY, NY, NJ

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

102

Additional Responses to Question 22, What Are Your States Critical Performance Indicators for Assuring Good Low-Temperature
Performance?

Response
BBR and Direct Tension. We are also looking at different mix tests.
BBR m-value (our binders seem to be creep-limitedstiffness is generally well within limits)
BBR maximum stiffness and minimum m-value. Almost all of our binders are m controlled.
BBR stiffness value less than 300 MPa at a test temperature of 12 degrees Celsius. BBR m-value greater
than 0.300 at a test temperature of 12 degrees Celsius.
BBR, DSC
Binders must pass low temperature PG grading.
Ductility at 4C, BBR stiffness and m-value.
Dynamic Shear and Creep Stiffness
Lower grading temperature on long-term aged binder (28C, 34C, or 40C depending on area).
Meet PG low on binder
Moving to performance testing of mixes, DCT. Binder onlyBBR
No cracking in pavement in cold climate areas like the NC mountains
PAV DSR < 5000 BBR: stiffness < 300, m value > 300
PAV TestingBBR
PAV-BBR
PG low temp dynamic shear and creep stiffness.
Penetration, ductility, viscosity and Marshall stability on asphalt concrete mix.
Transverse cracking: binder grading (low end)
evaluating DCT and SCB mixture cracking tests
we do not perform performance testing.

Question 23: Have you been able to identify field failures due to binder composition? If yes, describe failures.

FIGURE A23 Can Identify Field Failures Due to Binder Composition

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 103

Additional Responses to Question 23, Have You Been Able to Identify Field Failures Due to Binder Composition?

Response
Have experienced issues with mixes using high amounts of RAP and RAS.
I cannot recall anything specific to identify a failure.
Is it foundation, poor construction or binder.
Not by composition. Normally, identify issues related to the grade.
Premature aging is seen sometimes. Is it binder? Maybe.
Rapid embrittlement of mixes a couple of months after installation.
Suspected, but never able to determine a smoking gun.
Two research projects where cold-weather performance of PG 58-28 outperformed PG 52-34.
Since there is no reliable field test, it is hard to determine if the failures can be attributed directly to the
binder itself. We have had stripping, bleeding, segregation and other failures that have had binder
blamed, but nothing definitive.
We had a specification failure for DSR but not a field failure. PPA was used and with one ASA, it a true
grade high temperature drop of 5C.
It was a chip seal emulsion that performed fine after placement, but the snow plows pulled up the emul-
sion residue. The reason: excess diluent (diesel)
Several mixes flushed/bled and did not set-up; binder composition with specific aggregate suspected as
the cause; could not replicate issue in the lab
Incompatibility between wood deck preservative and asphalt binders. Initial stages of failure identified.
It is the opinion of our construction personnel that are binders are not of the quality they used to be 710
years ago. Our mix designs have basically not changed over the years, yet our asphalt mixes are only last-
ing 79 years. It is apparent from inspection, that the polymer quantity or quality has changed over the
last 7 years or so.
Yes, and no on this question. At times the wrong binder has been supplied to a project or PPA was used
with a limestone aggregate and was not compatible.
In the past decade we noticed premature cracking (top-down mostly), starting in the wheelpaths and then
propagating in a map cracking form. In many cases we recovered the binder and tested and noticed a great
loss of low temperature grade (especially using ExBBR test) whis is an indication that the SHRP aging
protocol is not sufficient on the low temperature side

Question 24: What procedures have you used to remedy any field failures described above?

Response
No Answer (9)
A Plan note was used on future project to allow no more than a 2C true grade drop.
As explained earlier, we have introduced new advanced binder testing
Change spec limits, change mix types etc.
Disallow the use of REOB or other recycled oils.
FDOT is looking into developing a field test for determining polymer/rubber presence and possibly quan-
tity. This is to try and prevent dilution at the plants and ensure that we are getting the high performing
binder that we have asked for by design.
Fog seal, remove and replace.
Have lowered the amount of RAP and RAS that can used in bituminous mixtures.
If it is something related to raveling or top down cracking, we tried fog seals with rejuvenator..
Proposed procedureremoval.
Remove and Replace (2)
Sent asphalt binder samples to TFHRC for REOB analysis.
Take good samples.
We have discussed the possibility of adding 0.2% virgin binder to our asphalt mixes and study them over
a few years to see if this aids in longer pavement life.
case by case

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

104

Question 25: Are you conducting ongoing research on binder characteristics that is not covered above and that you would be willing
to share? If yes, please provide contact information.

FIGURE A25 Conducting Ongoing Research on Binder Characteristics That You Would Be Willing to Share

Additional Responses to Question 25, Are You Conducting Ongoing Research on Binder Characteristics That Is Not Covered
Above and That You Would Be Willing to Share?

Response
Connecticut DOT is not currently conducting research on binder characteristics.
Virginia DOT for GPC. Contact Ben Bowers: ben.bowers@vdot.virginia.gov
Nebraska Department of Roads Please contact Robert Rea at 402-326-9934 to discuss further.
Rhode Island DOT Contact Professor Michael Greenfield URI greenfield@uri.edu
Iowa DOT Rejuvenator research with David Lee at University of Iowa
Oklahoma DOT Very early stages of in-house XRF testing for PPA and REOB.
Texas DOT Contact Tom Scullion, Texas A&M Transportation Institute; Amit Bhasin, Center for Trans-
portation Research at UT.
Ohio DOT Just started binder/RAS/RAP study with Ohio University, Dr Munir Nazzal. Just started mix
RAS/RAP cracking study with NCAT.
Wyoming DOT. I performed research on asphalt binders at WRI for roughly the last 11 years. Lead
author on about a dozen peer reviewed papers. Have a lot of ideas on binder research
michael.farrar@wyo.gov

Comments Provided by Canadian Respondents

Question 1: Do you know the crude source of oil used to produce the asphalt binder and refining process employed? If yes, please
identify the source. Comments

Ontario: It is mostly from Western Canada (heavy crude from Cold Lake and light crude from Alberta) but there could also be blend
of light oil from states, Venezuela, Saudi, etc.

SK Grade A crude typically from western Canada

Question 2: What refining process was used to produce the asphalt? Comments

there are several asphalt suppliers in Ontario. Many of them do blending of different grades.

Question 3: Have you observed significant differences in the asphalt binder PG properties derived from your oil feedstock in the
past 10 years? Comments

QC: We didnt observed differences in the PG properties, but we have observed differences in the MSCR results

Ontario: We are the agency. We have noticed premature cracking on our highways and believe that PG quality is a major player

Question 4: Are you conducting any of the following binder fractionation procedures? If yes, please describe what triggered the

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 105

decision to conduct this test? Comments

Question 5: What rheological and analytical procedures are you employing in your laboratory for asphalt binder testing? Please
mark all that apply.

GNB: Functional group analysis

Question 6: Are you conducting any of the following binder fractionation procedures? If yes, please describe what triggered the
decision to conduct this test.

NT: FTIR Compatibility/Phase separation

NS: No fractionation procedures are conducted as part of the Departments Quality Assurance program

BC: We have uncovered asphalt reaction with preservative treated woods resulting in liquefaction of bitumen. Further chemical
investigation is planned.

SK: we are not conducting any of the binder fractionation procedures mentioned

Question 7: Are you working on new procedures for separating various components of the asphalt binder? If yes, please describe
your new procedure. Comments

Question 8. What procedures are you using routinely to estimate asphalt binder aging characteristics? Comments

Ontario: We use Extended BBR test for evaluating effect of physical hardening

Question 9. What procedures are you using occasionally to estimate asphalt binder aging characteristics? What conditions initiate
the additional investigation? Comments

Question 10. Are you employing any of the following modifiers/additives in final binder preparation to achieve specific asphalt
binder grades? Comments

We require antistripping additive for northern Ontario. We allow a limited quantity of PPA. We use a lot of PG XX-34 and some PG
XX-40 therefore we know that softening agents (oils, etc.) are being used.

QC: Additives may be used by the asphalt binder suppliers to achieve our specifications. We do not request additives.

NL: Liquid antistripping agents are used in all binders but not to achieve PG grades.

Question 11: If you extracted binder from mixtures containing recycled materials, which mixtures have you examined? If yes, com-
ment on your results. Comments

QC: The use of 20% or less of RAP has a low impact on the asphalt binder properties

Question 12: If you are employing polymer additives, which type are you using? Comments

SK: SK does not use PG grade asphalt binder; typical asphalt used is 150200 and 200300. Rubber asphalt binder has been used
in past. the base grade is 200300 pen which is modified by addition of crumb rubber.

Ontario: We do not specify type of polymers. It is up to asphalt suppliers

QC: The majority of our asphalt binders are modified with polymers. The presence of polymers is controlled by an elastic recovery
test. The suppliers choose the type of polymer used to meet the elastic recovery specification.

SK: lime is used as anti-stripping agent for asphalt concrete mixes.

Question 13: If you are employing rejuvenators to restore aged asphalt binder, which type are you using? Comments

QC: No rejuvenators are used. A maximum of 20% of RAP and 3% of RAS are allowed in our mix.

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

106

NS: Rejuvenator used for cold in place recycling treatments

SK: cyclogen has been used on few occasions in past.

Question 14. How do you quantify the presence of modifiers/additives in final asphalt binder composition? Comments

Ontario: We test the final product for rheology. We have tried using XRF to detect PPA. It wont give accurate results due to low
dosage. We tried FTIR for antistripping agents. Similarly, the low dosage is hard to detect.

NS: Any additives need to be identified on weigh tickets

Question 15: Do you measure the compatibility between modifiers and asphalt final binder after storage or aging? Please cite the
measurement technique employed in the comments box. Comments

Ontario: It is up to asphalt suppliers. We test the final product for compliance to our specifications.

QC: It might be done by the suppliers. Not by the Ministry.

Question 16: If using an additive, what specifications are used for the amount of additive for a given grade? Comments

QC: The amount of additives used is determined by the suppliers. The final product shall meet all the specifications.

Question 17: Are current PG grade criteria satisfactory in identifying compositional/chemical properties? If no, suggest criteria
that should be added or changed. Comments

Ontario: In addition to M320 compliance, we require ash content not exceeding 0.8%. We also have acceptance criteria for MSCR,
DENT, and ExBBR tests.

QC: The PG grade is useful to evaluate the performance of binders at high and low temperatures, but its not a direct relationship
between the rheology of the binder and the chemical composition of the binder. Other properties such as the elasticity and the adhe-
sivity of the binder must be considered.

SKL: Not applicable as PG grades are not used in SK.

NL: For polymer grades MSCR or elastic recovery could be specified.

Question 18: Based upon internal research/knowledge, which physical properties are most sensitive to changes in the crude stock
or binder chemistry? Comments

Question 19: Is your laboratory developing alternate techniques to estimate binder performance? If yes, please summarize details.
Comments

Ontario: We have already implemented the DENT test (Double Edge Notch Tension Test-which is an AASHTO test now), MSCR,
and Ash content (to limit REOB). We are phasing-in the ExBBR test (which is also an AASHTO Provisional test now). We are
looking at XRF and FTIR tests to quantify REOB. We have also developed modified PAV procedure which is being evaluated at
this moment.

QC: We have a method to measure the ash content of the binders. We also have a method to evaluate the adhesivity between a binder
and an aggregate. For some binders, we added some criteria with this method in our specifications to be sure to get a binder with
superior adhesive properties. These binders help to reduce stripping problems of the pavements.

Question 20: What are your states critical performance indicators for assuring good high-temperature performance?

We do not have rutting problem in Ontario. We rely on M320 results as well as MSCR compliance.

BC: This has been identified as an area of research for northern climates.

QC: Upper grading temperature on original and short-term aged binder (58C, 64C, or 70C depending of the area and type
of road).

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 107

NS: Dynamic Shear

SK: Penetration, ductility, viscosity and Marshall stability on asphalt concrete mix.

NL: PG high temp dynamic shear

Question 21: What are your states critical performance indicators for assuring good intermediate-temperature performance?

NS: Dynamic Shear

SK: Penetration, ductility, viscosity and Marshall stability on asphalt concrete mix.

Question 22: What are your states critical performance indicators for assuring good low-temperature performance?

SK: Penetration, ductility, viscosity and Marshall stability on asphalt concrete mix.

QC: Lower grading temperature on long-term aged binder (28, 34 or -40C depending on area).

NS: Dynamic Shear and Creep Stiffness

NL: PG low temp dynamic shear and creep stiffness

Question 23: Have you been able to identify field failures due to binder composition? If yes, describe failures. Comments

Ontario: In the past decade we noticed premature cracking (top-down mostly), starting in the wheelpaths and then propagating in a
map cracking form. In many cases we recovered the binder and tested and noticed a great loss of low temperature grade (especially
using ExBBR test) this is an indication that the SHRP aging protocol is not sufficient on the low temperature side

Question 24: What procedures have you used to remedy any field failures described above?

Ontario: As explained earlier, we have introduced new advanced binder testing

Question 25: Are you conducting ongoing research on binder characteristics that is not covered above and that you would be willing
to share? If yes, please provide contact information. Comments

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

108

APPENDIX B
Survey Respondents
Title Agency/Organization
Asphalt Pavement Engineer MaineDOT
CEG ALDOT
Chemical Laboratory Director MnDOTOffice of Materials and Road Research
Deputy DirectorConstruction Division Texas Department of Transportation
Asphalt Binder Engineer Texas Department of Transportation
Office Chief Caltrans
Civil Engineering Manager 2 Tennessee Department of Transportation
Kentucky Transportation Cabinet/Division of
Materials
Bridge Area Manager Ministry of Transportation & Infrastructure
Bituminous Engineer South Dakota DOT
Assistant State Materials Engineer WSDOT
Manager Materials Engineering Transportation and Works
Asphalt Materials Engineer Ohio DOT
Director Materials Engineering Manitoba Infrastructure & Transportation
Bituminous Engineer AR State Highway and Transp Department
Asphalt Materials Manager SCDOT
Administrator VII Maryland SHA
MassDOT
Materials Engineer Montana Department of Transportation
Bituminous Engineering Specialist Florida Department of Transportation
M.Sc. Chemist Ministere des transports, de la mobilit durable et de
l'lectrification des transports
Transportation Engineering Specialist Arizona Department of Transportation
Civil Engineer 3 (Materials) NYSDOT
Transportation Engineer III CT DOT Materials
Field Operations Chief Illinois Department of Transportation
Principal Civil Engineer RIDOT
Chief Materials Engineer Wisconsin DOT
Bituminous Lab EIT Colorado Department of Transportation
Chief Chemist Delaware Dept of Transportation
Bituminous Engineer Montana Department of Transportation
New Brunswick Department of Transportation and
Infrastructure
Chief Materials Engineer DC Department of Transportation
Senior Bituminous Engineer Ontario Ministry of Transportation
Assistant State Materials Engineer Georgia DOT
HMA Materials Manager Vermont Agency of Transportation
ManagerHighway Construction Services Nova Scotia Transportation and Infrastructure Renewal
Bituminous Engineer Oklahoma DOT
Engineer of Tests PA Department of Transportation
Field Quality Assurance Administrator Louisiana DOTD

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

 109

Title Agency/Organization
State Materials Engineer NC Dept of Transportation Materials & Tests
Chief Chemist Mississippi Department of Transportation
Senior Materials and Surfacing Engineer Ministry of Highways and Infrastructure
Manager, Bureau of Materials NJDOT
VDOT
Iowa DOT
Materials Engineer Indiana Department of Transportation
Chief, Bureau of Research Kansas DOT
Pavement Quality and Materials Engineer Oregon DOT
Asphalt Group Supervisor West Virginia Division of Highways
State Asphalt Engineer UDOT
Materials Staff Engineer WYDOT
Trans Proj Manager NDDOT
Nebraska Department of Roads
Chief Materials Engineer Nevada DOT
Senior Geotechnical Engineer Department of Transportation, Northwest Territories,
Canada
Chief of Materials Technology NHDOT
HMA Operations Engineer Michigan DOT
Physical Laboratory Director Missouri DOT
State Asphalt Engineer NMDOT

Copyright National Academy of Sciences. All rights reserved.


Relationship Between Chemical Makeup of Binders and Engineering Performance

Abbreviations and acronyms used without denitions 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
ACINA Airports Council InternationalNorth 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 Americas 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.


Relationship Between Chemical Makeup of Binders and Engineering Performance

ADDRESS SERVICE REQUESTED

Washington, D.C. 20001


500 Fifth Street, N.W.
TRANSPORTATION RESEARCH BOARD

Copyright National Academy of Sciences. All rights reserved.

Anda mungkin juga menyukai