You are on page 1of 184

Project No. 9-58 Copy No.

__

THE EFFECTS OF RECYCLING AGENTS ON ASPHALT


MIXTURES WITH HIGH RAS AND RAP BINDER RATIOS

PHASE I
INTERIM REPORT

Prepared for
National Cooperative Highway Research Program
Transportation Research Board
of
The National Academies

TRANSPORTATION RESEARCH BOARD


OF THE NATIONAL ACADEMIES
PRIVILEGED DOCUMENT

This report, not released for publication, is furnished only for review to members
of or participants in the work of the CRP. This report is to be regarded as fully
privileged, and dissemination of the information included herein must be
approved by the CRP.

Amy Epps Martin


Fujie Zhou
Edith Arambula
Eun Sug Park
Arif Chowdhury
Fawaz Kaseer
Juan Carvajal
Elie Hajj
Jo Daniel
Charles Glover

Texas A&M Transportation Institute


The Texas A&M University System
College Station, Texas

March 2015
ACKNOWLEDGMENT OF SPONSORSHIP
This work was sponsored by one or more of the following as noted:
X American Association of State Highway and Transportation Officials, in cooperation with the
Federal Highway Administration, and was conducted in the National Cooperative Highway
Research Program,

Federal Transit Administration and was conducted in the Transit Cooperative Research
Program,

Federal Aviation Administration and was conducted in the Airport Cooperative Research
Program,

Research and Innovative Technology Administration and was conducted in the National
Cooperative Freight Research Program,

Pipeline and Hazardous Materials Safety Administration and was conducted in the
Hazardous Materials Cooperative Research Program,

which is administered by the Transportation Research Board of the National Academies.

DISCLAIMER
This is an uncorrected draft as submitted by the research agency. The opinions and
conclusions expressed or implied in the report are those of the research agency. They are not
necessarily those of the Transportation Research Board, the National Academies, or the program
sponsors.
Project No. 9-58 Copy No.__

THE EFFECTS OF RECYCLING AGENTS ON ASPHALT


MIXTURES WITH HIGH RAS AND RAP BINDER RATIOS

PHASE I
INTERIM REPORT

Prepared for
National Cooperative Highway Research Program
Transportation Research Board
of
The National Academies

Amy Epps Martin


Fujie Zhou
Edith Arambula
Eun Sug Park
Arif Chowdhury
Fawaz Kaseer
Juan Carvajal
Elie Hajj
Jo Daniel
Charles Glover

Texas A&M Transportation Institute


College Station, Texas

March 2015
TABLE OF CONTENTS

LIST OF FIGURES ........................................................................................................................ v 


LIST OF TABLES ....................................................................................................................... viii 
ACKNOWLEDGMENTS .............................................................................................................. x 
ABSTRACT ................................................................................................................................... xi 
CHAPTER 1.0 INTRODUCTION ................................................................................................. 1 
1.1 NCHRP Research on Recycled Materials in Asphalt Mixtures ............................... 2 
1.2 Scope of the Interim Report ...................................................................................... 3 
CHAPTER 2.0 BACKGROUND ................................................................................................... 5 
2.1  Aging and Recycling of Asphalt Materials ........................................................... 5 
2.1.1   Aging of Asphalt Materials ........................................................................... 5 
2.1.2  Recycling of Asphalt Materials ..................................................................... 15 
2.1.3  Degree of Blending between Virgin and Aged Binders ................................ 17 
2.1.4  Improving Binder Rheology in Recycled Materials...................................... 19 
2.2  Restoring and Characterizing Binder Rheology with RAs ................................. 21 
2.2.1  RAs: Definition, Types, and Properties ........................................................ 21 
2.2.2  Effect of RAs on Binders with High RBRs ................................................... 24 
2.2.3  Characterizing Binder Rheology with RAs ................................................... 31 
2.3  Improving and Characterizing High RBR Mixture Performance with RAs ....... 33 
2.3.1  Effect of RAs on Asphalt Mixtures with High RBRs ................................... 34 
2.3.2  Characterizing High RBR Mixture Performance with RAs .......................... 43 
CHAPTER 3.0 STATE DOT, CONTRACTOR, AND SUPPLIER WEB-BASED
SURVEYS .................................................................................................................................... 47 
3.1  State DOT Web-Based Survey ............................................................................ 47 
3.2  Contractor Web-Based Survey ............................................................................ 55 
3.3  RA Supplier Web-Based Survey ......................................................................... 61 
CHAPTER 4.0 PHASE II WORK PLAN Development .............................................................. 63 
4.1 Field Projects and Expanded Materials ................................................................ 64 
4.2 Field Activities ...................................................................................................... 69 
4.2.1  Material Sampling during Production and Construction............................... 69 
4.2.2  Coring Plan after Construction ...................................................................... 69 
4.2.3  Prior- and Post-Construction Assessment ..................................................... 70 
4.3  Laboratory Testing .............................................................................................. 70 
4.3.1  Binder Testing ............................................................................................... 72 
4.3.2  Mortar Testing ............................................................................................... 73 
4.3.3  Mixture Testing ............................................................................................. 74 
4.4  Mortar Experiment .............................................................................................. 80 
4.4.1  Preliminary Mortar Results ........................................................................... 80 
4.4.2  Expanded Mortar Experiment ....................................................................... 91 
4.5  Binder Experiments ............................................................................................. 94 
4.5.1  Preliminary Binder Results ........................................................................... 94 
4.5.2  Expanded Binder Rheology Experiments ................................................... 101 
4.5.3  Binder Compatibility Experiment ............................................................... 105 
4.5.4  Binder Aging Experiments .......................................................................... 107 
4.6  Mixture Experiments ......................................................................................... 110 
4.6.1  Preliminary Specimen Fabrication Results ................................................. 110 

iii
4.6.2  Aging Analysis ............................................................................................ 113 
4.6.3  Expanded Mixture Experiments .................................................................. 115 
4.7   Statistical Analysis, Technical Reports, and Phase III ...................................... 117 
4.7.1   Statistical Comparisons ............................................................................ 117 
4.7.2  Technical Reports ........................................................................................ 118 
4.7.3  Phase III....................................................................................................... 119 
REFERENCES ........................................................................................................................... 120 
LIST OF ACRONYMS .............................................................................................................. 131 
APPENDIX A. SURVEYS............................................................................................................. 1 
A.1  State DOT Web-Based Survey .............................................................................. 1 
A.2  Contractor Web-Based Survey .............................................................................. 8 
A.3  RAs Supplier Web-Based Survey ....................................................................... 15 
APPENDIX B. SPECIMEN FABRICATION ............................................................................... 1 
B.1  LMLC Specimens ................................................................................................. 1 
B.1.1  Materials Used and Mix Design ..................................................................... 1 
B.1.2  Recycled Material Handling .......................................................................... 3 
B.1.3  Specimen Fabrication .................................................................................... 4 
B.2  PMLC and RPMLC Specimens ............................................................................ 4 
B.3  Resilient Modulus Test.......................................................................................... 4 
B.4  Test Results ........................................................................................................... 6 
APPENDIX C. Texas FIELD PROJECT CONSTRUCTION REPORT ....................................... 1 
C.1  Mixtures and Materials.......................................................................................... 2 
C.2  Production of Mix and Paving .............................................................................. 2 
C.2.1  Section 1: Virgin Mix .................................................................................... 3 
C.2.2  Section 2: Control Mix with 10 Percent RAP and 5 Percent RAS ................ 7 
C.2.3  Section 3: Control Mix with Rejuvenator-Hydrogreen ................................. 7 
C.2.4  Section 4: Control Mix with Rejuvenator-Evoflex and Evotherm ................ 8 
C.2.5  Section 5: Control Mix with Rejuvenator-ERA-1 ......................................... 9 
C.3  Description of Asphalt Mix Plant........................................................................ 10 
C.4  Sample Collection ............................................................................................... 10 
C.5  Onsite Specimen Compaction ............................................................................. 11 
C.6  Stockpiles and Plant Details ................................................................................ 12 

iv
LIST OF FIGURES

Figure 2.1. Binder Rheology Due to Aging and Addition of RAs. ................................................ 6 
Figure 2.2. Overview of the Model Used to Estimate Time to Pavement Failure. ......................... 8 
Figure 2.3. Procedure to Determine Asphalt Reaction Kinetics Parameters. ................................. 9 
Figure 2.4. Example Binder Pavement Oxidation for Four Eac Values. 65 to 105 kJ/mol Span
the Observed Data; 65 and 75 kJ/mol Are Not Statistically Different. ........................................ 10 
Figure 2.5. Accelerated Method to Determine Asphalt Reaction Kinetics and Rheology. .......... 11 
Figure 2.6. Constant-Rate Reaction Kinetics Parameters for Binders at Both POV and PAV
Conditions. .................................................................................................................................... 12 
Figure 2.7. Constant-Rate Reaction Rates as a Function of Temperature for Binders at Both
POV and PAV Conditions (T in K, R = 8.314 J/mol/K). ............................................................. 13 
Figure 2.8. Example Plot Showing How Viscosity HS Is Determined. ....................................... 14 
Figure 2.9. Example Binder Pavement DSRfn Hardening for Four Eac Values. .......................... 15 
Figure 2.10. RAP–Virgin Binder Contact and Blending (Kriz et al. 2014).................................. 18 
Figure 2.11. Penetration of the RA into the Outer and Inner Layers of an Aged Binder by
Diffusion (Carpenter and Wolosick, 1980)................................................................................... 25 
Figure 3.1. Predominant Use of RAP and/or RAS in Surface Mixtures....................................... 48 
Figure 3.2. Predominant Type of Recycled Materials Used in Surface Mixtures. ....................... 48 
Figure 3.3. Current RAP Practice in Surface Mixtures. ............................................................... 49 
Figure 3.4. RAP Practice in Surface Mixtures. ............................................................................. 49 
Figure 3.5. Typical RAP Content in Surface Mixtures (no RAS). ............................................... 50 
Figure 3.6. Current RAS Practice in Surface Mixtures. ............................................................... 51 
Figure 3.7. RAS Practice in Surface Mixtures. ............................................................................. 51 
Figure 3.8. Typical RAS Content in Surface Mixtures................................................................. 52 
Figure 3.9. Performance Tests for Surface Mixtures with RAP and/or RAS. .............................. 53 
Figure 3.10. Common Distresses Observed on Surface Mixtures with RAP and/or RAS. .......... 53 
Figure 3.11. Current RA Practice in Surface Mixtures. ................................................................ 54 
Figure 3.12. Laboratory Tests to Characterize Properties of Binders with RAs. ......................... 55 
Figure 3.13. Upcoming Field Projects Using RAs and Availability............................................. 55 
Figure 3.14. Predominant Use of RAP and/or RAS in Surface Mixtures..................................... 56 
Figure 3.15. Current RAP Practice in Surface Mixtures. ............................................................. 56 
Figure 3.16. Typical RAP Content in Surface Mixtures (no RAS). ............................................. 57 
Figure 3.17. Current RAS Practice in Surface Mixtures. ............................................................. 57 

v
Figure 3.18. Typical RAS Content in Surface Mixtures............................................................... 58 
Figure 3.19. Performance Tests for Surface Mixtures with RAP and/or RAS. ............................ 58 
Figure 3.20. Common Distresses Observed on Surface Mixtures with RAP and/or RAS. .......... 59 
Figure 3.21. Current RA Practice in Surface Mixtures. ................................................................ 59 
Figure 3.22. Laboratory Tests to Characterize Properties of Binders with RAs. ......................... 60 
Figure 3.23. Upcoming Field Projects Using RAs and Availability............................................. 61 
Figure 3.24. Types of RAs Produced. ........................................................................................... 62 
Figure 3.25. Prescribed Blending Protocol for Plant and Laboratory Operations. ....................... 62 
Figure 4.1. SHRP-LTPP Environmental Zones. ........................................................................... 65 
Figure 4.2. Mixture Black Space Curves for Vermont Pooled Fund Phase I Materials. .............. 75 
Figure 4.3. Modified Mixture Glover-Rowe Values (0.005 rad/s and 15oC) against
Laboratory-Measured Critical Cracking Temperature for Pooled Fund Mixtures. ...................... 76 
Figure 4.4. Comparison of Measured and Predicted Modified Paris’ Law Coefficients (Luo
et al. 2013b). ................................................................................................................................. 78 
Figure 4.5. RAP and RAS Materials (#50 X #100) from the Texas Field Project. ...................... 81 
Figure 4.6. Continuous Grades for Virgin and Recycled Binders. ............................................... 82 
Figure 4.7. Effect of RAP and RAS Continuous PGH Grade with PG 70-22 Virgin Binder....... 85 
Figure 4.8. Effect of RAP and RAS on Continuous PGL Grade with PG 70-22 Virgin
Binder............................................................................................................................................ 86 
Figure 4.9. Effect of RAP and RAS on Continuous PGH Grade with PG 64-22 Virgin
Binder............................................................................................................................................ 86 
Figure 4.10. Effect of RAP and RAS on Continuous PGL Grade with PG 64-22 Virgin
Binder............................................................................................................................................ 87 
Figure 4.11. Effect of RAP and RAS on Continuous PGH Grade with PG 64-22 Virgin
Binder and RA (T1). ..................................................................................................................... 87 
Figure 4.12. Effect of RAP and RAS on Continuous PGL Grade with PG 64-22 Virgin
Binder and RA (T1). ..................................................................................................................... 88 
Figure 4.13. Effect of RAP and RAS on Continuous PGH Grade with PG 64-28 Virgin
Binder............................................................................................................................................ 88 
Figure 4.14. Effect of RAP and RAS on Continuous PGL Grade with PG 64-28 Virgin
Binder............................................................................................................................................ 89 
Figure 4.15. Estimated Continuous Grades for Blended Binders at 28.6 Percent RBR. .............. 90 
Figure 4.16. Estimating the Restoration of Blended Binders at 28.6 Percent RBR. .................... 91 
Figure 4.17. Validation of Regional Linear Blending Concept for Combined Blends of
PG 64-22, 10 percent TX RAP, 5 percent MWAS, and TX RAs Using M-Controlled Values
for PGL Grade............................................................................................................................... 95 

vi
Figure 4.18. Validation of Regional Linear Blending Concept for Combined Blends of
PG 64-22, 10 percent TX RAP, 5 percent MWAS, and TX RAs Using S-Controlled Values
for PGL Grade............................................................................................................................... 96 
Figure 4.19. Estimating the Restoration of PGH Grade for Binder Blends with MWAS. ........... 99 
Figure 4.20. Comparing Restoration of PGH Grade Based on Binder and Mortar Testing for
Blends with MWAS. ..................................................................................................................... 99 
Figure 4.21. Comparing Restoration of PGH Grade Based on Binder and Mortar Testing for
Blends with TOAS (Expected Results)....................................................................................... 100 
Figure 4.22. Comparing Restoration of PG Grade Based on Different RAS Sources................ 100 
Figure 4.23. Evaluating Restoration of PG Grade in Black Space (Zhou et al. 2015). .............. 101 
Figure 4.24. Comparison of Low-Temperature PG Grade from BBR Measurements to Those
Predicted from 8 mm DSR Frequency Sweeps Based on (a) S-Controlled PGL, and (b) m-
Controlled PGL. .......................................................................................................................... 105 
Figure 4.25. Example CA Growth over Time at 85°C. .............................................................. 110 
Figure 4.26. MR Results for the Mixtures Used in the Texas Field Project. ............................... 112 
Figure 4.27. Cumulative Degree Days for NCHRP 9-52 Field Projects. ................................... 113 
Figure 4.28. MR Ratio versus CDD for NCHRP 9-52 Post-Construction Cores and
Correlation of LTOA Protocols with Field Aging. ..................................................................... 114 
Figure 4.29. MR Ratio versus CDD for NCHRP 9-52 Post-Construction Cores for RAP
Mixtures. ..................................................................................................................................... 115 
Figure 4.30. Example Mix Design and Performance Evaluation Guidelines and Evaluation
Tools. .......................................................................................................................................... 119 

vii
LIST OF TABLES

Table 2.1. RA Categories and Types (NCAT 2014a). .................................................................. 22 


Table 2.2. Typical RA Properties. ................................................................................................ 23 
Table 2.3. Tests Used for the Proposed Specification by Kari et al. (1980). ................................ 24 
Table 2.4. Previous Research on the Effect of RAs on the Stiffness of Recycled Asphalt
Mixtures. ....................................................................................................................................... 29 
Table 2.5. Previous Research on the Effect of RAs on the Stiffness of Recycled Asphalt
Mixtures. ....................................................................................................................................... 35 
Table 2.6. Previous Research on the Effect of RAs on Rutting Resistance of Recycled
Asphalt Mixtures........................................................................................................................... 37 
Table 2.7. Previous Research on the Effect of RAs on Fatigue and Reflective Cracking
Resistance of Recycled Asphalt Mixtures. ................................................................................... 39 
Table 2.8. Previous Research on the Effect of RAs on Low-Temperature Cracking
Resistance of Recycled Asphalt Mixtures. ................................................................................... 41 
Table 2.9. Previous Research on the Effect of RAs on Moisture Susceptibility of Recycled
Asphalt Mixtures........................................................................................................................... 43 
Table 4.1. Proposed Field Projects and Associated Laboratory Testing Plan. ............................. 67 
Table 4.2. Texas Field Project. ..................................................................................................... 68 
Table 4.3. Selected Laboratory Tests. ........................................................................................... 72 
Table 4.4. Preliminary Mortar Experiment for Texas Field Project Materials and
Supplemental Virgin Binder. ........................................................................................................ 80 
Table 4.5. Effect of RAP on Continuous PG Grade. .................................................................... 84 
Table 4.6. Effect of RAS on Continuous PG Grade. .................................................................... 84 
Table 4.7. Verification of the Mortar Procedure for RAP and RAS Combination. ...................... 89 
Table 4.8. Expanded Mortar Experiment with RA T1.................................................................. 93 
Table 4.9. Expanded Mortar Experiment with RA A1. ................................................................ 94 
Table 4.10. Dosage Rates for Texas Field Project. ....................................................................... 96 
Table 4.11. Binder Blending Table for Restoring PGH Grade^ (Estimated by Binder |
Mortar Analysis with NA Results Not Available Yet). ................................................................ 98 
Table 4.12. Expanded Component Binders and RA Experiment ............................................... 102 
Table 4.13. Expanded Binder Blend Experiment for RA Type T1. ........................................... 103 
Table 4.14. Expanded Binder Blend Experiment for RA Type A1. ........................................... 104 
Table 4.15. Blends for Compatibility Experiment. ..................................................................... 106 
Table 4.16. Binder Aging Experiment. ....................................................................................... 108 
Table 4.17. Proposed Testing Protocol for Developing Binder DSR Master Curves................. 109 

viii
Table 4.18. Correlation of Field Aging in Terms of In-Service Time and Laboratory LTOA
Protocol of 5 days at 85°C (185°F) for NCHRP 9-52 Field Projects. ........................................ 114 
Table 4.19. Expanded Mixture Experiment for Texas Field Project. ......................................... 116 

ix
ACKNOWLEDGMENTS

The research reported herein was performed under NCHRP Project 9-58 by the Texas
A&M Transportation Institute with the Texas A&M Sponsored Research Services serving as
fiscal administrator. Dr. Amy Epps Martin, P.E., professor of civil engineering at Texas A&M
University, was the project director and principal investigator. Dr. Fujie Zhou, P.E., research
engineer with the Texas A&M Transportation Institute, was the co-principal investigator. Other
authors of this report are Dr. Edith Arambula, Dr. Eun Sug Park, Mr. Arif Chowdhury,
Mr. Fawaz Kaseer, Mr. Juan Carvajal, Dr. Elie Hajj, Dr. Jo Daniel, and Dr. Charles Glover.

x
ABSTRACT

This interim report documents the results of NCHRP Project 9-58 Phase I and describes
the Phase II work plan based on a literature review, survey, and preliminary laboratory results. A
comprehensive information search was completed and included a literature review to document
aging and recycling of asphalt materials, restoring and characterizing binder rheology with RAs,
and improving and characterizing mixture performance with high recycled binder ratios and
RAs. A survey of state departments of transportation, contractors, and RA suppliers was also
conducted to assess the current state of the practice on the use of RAs in hot-mix asphalt and
warm-mix asphalt mixtures containing high percentages of recycled materials. Preliminary
laboratory results included findings from testing of binders, mortars, and mixtures from a field
project in Texas and supplemental materials to expand an initial knowledge base. All of this
information was synthesized and used to develop the Phase II work plan, which is also described
in this report.

xi
CHAPTER 1.0 INTRODUCTION

More than 90 percent of highways and roads in the United States are built using hot-mix
asphalt (HMA) or warm-mix asphalt (WMA) mixtures. In the early 1990s, the Federal Highway
Administration (FHWA) estimated that more than 90 million tons of asphalt pavements are
milled off roads each year during resurfacing projects, and over 80 percent of reclaimed asphalt
pavement (RAP) is recycled in new asphalt mixtures (FHWA 1993). Following studies showed
that this trend is continuously increasing (Copeland 2011).
The use of RAP in HMA dates back to the early 1900s with renewed focus on research
and implementation in the 1970s and 1980s due to dramatic increases in the cost of oil and thus
asphalt binders and fuel needed to produce asphalt pavements. Newcomb and Epps (1981)
reviewed the technologies developed during this early period of recycling, which included drum
mix plants, cold milling machines, vibratory compaction rollers, cold and hot in-place recycling
techniques, and mix design methods toward high RAP contents to maximize economic and
environmental benefits. These symbiotic benefits are substantial and include: conservation of
natural resources (aggregate, binder, fuel, etc.), reduction in energy consumption, and reduction
in emissions (including greenhouse gases). For example for a relatively high 25 percent RAP
content HMA mixture, Robinett and Epps (2010) indicated 10 percent energy savings, 10 percent
emissions reductions, and 20–25 percent conservation of natural resources that translated into
reduced production and construction costs.
Interest in recycling waned during the 1990s and was not considered in the Strategic
Highway Research Program (SHRP), and these technologies remained largely unchanged until
2008, when the cost of petroleum products significantly increased again. So highway agencies
and the paving industry have developed a renewed interest in achieving higher recycled binder
ratios (RBRs) in asphalt mixtures through the use of larger percentages of RAP and/or the
addition of recycled asphalt shingles (RAS) from either manufacturer waste asphalt shingles
(MWAS) or tear-off asphalt shingles (TOAS) for the same economic and environmental benefits
described previously. To provide an overall indication of the possible binder contribution from
these recycled materials, RBR is defined as follows:

100
where PbRAP is the binder content of the RAP, PRAP is the percentage of RAP by weight of mix,
PbRAS is the binder content of the RAS, PRAS is the percentage of RAS by weight of mix, and
Pbtotal is the binder content of the combined mixture.
In spite of the symbiotic benefits, state departments of transportation (DOTs) limit the
use of RAP and/or RAS in asphalt mixtures for reasons that include variability of the recycled
materials and concerns about the long-term performance of the asphalt mixtures that contain
these materials. In addition, the mix design of these mixtures is more complicated and more time
consuming, particularly with high RBRs between 0.3 and 0.5. The potential for the following
construction and performance issues is also increased as RBRs increase and corresponding
mixtures become stiffer and more brittle:

 Compactibility/workability in cool weather.


 Low-temperature cracking with accumulation of thermally induced stresses.

1
 Fatigue cracking and microdamage accumulation leading to crack initiation and
propagation with repeated loading.
 Reflection cracking with repeated loading and daily/seasonal thermal stresses.
 Raveling with subsequent aging or moisture damage.

Thus, the environmental and economic benefits must be compared to the potential
increased risks associated with construction and performance to ensure engineering benefits can
also be realized. Mitigation of these construction and performance issues can be addressed
through mix design with the use of higher binder contents, material selection with the use of
softer binders that may be polymer modified, or additives such as RAs (RAs), as long as there
are not compatibility concerns and binder resistance to bleeding and mixture resistance to
rutting/permanent deformation is maintained. Mitigation by the use of RAs includes (Tran et al.,
2012; Mogawer, Booshehrian, et al. 2013; Im and Zhou, 2014):

 Alleviation of high stiffness caused the addition of recycled materials at high RBRs.
 Improvement in cracking resistance of mixtures with high RBRs without adversely
affecting their resistance to rutting/permanent deformation and moisture damage.
 Improvement in compactibility/workability (in some cases).

Utilizing lower production temperatures through the use of WMA technologies will also affect
these construction and performance issues and possibly offset benefits from reduced aging with
decreased blending of virgin and recycled binders and/or possible compatibility concerns with
WMA additives, RAs, and virgin and recycled binders.
RAs were utilized in HMA in the early period of widespread recycling in the 1970s and
1980s for the purpose of realizing all three types of benefits—environmental, economic, and
engineering. As RBRs continue to increase in the current period of widespread recycling, the use
of RAs holds promise once again with proper understanding of their effectiveness in restoring
binder rheology, its evolution with aging in HMA and WMA mixtures in both the laboratory and
the field, and performance of these binder blends and corresponding mixtures. Mix design
procedures, including component material characterization to ensure the recycled binders are
restored as much as possible rheologically, specimen fabrication protocols to simulate field
conditions, and production and construction best practices (including handling of recycled
materials—fractionation and drying, for example), are needed to ensure adequate performance
when using RAs. Guidelines developed in the 1970s and 1980s, such as those by Epps et al.
(1980) and, more recently, Copeland (2011), can be utilized as a starting point to produce high-
quality asphalt mixtures with high RBRs and adequate performance.

1.1 NCHRP Research on Recycled Materials in Asphalt Mixtures

In addition to this study, the National Cooperative Highway Research Program (NCHRP)
has funded the following four research projects to address the use of recycled materials in asphalt
mixtures (McDaniel and Anderson 2001; Advanced Asphalt Technologies 2011; West et al.
2013):

 NCHRP Project 9-12: Incorporation of Reclaimed Asphalt Pavement in the


Superpave System. Two main publications resulted from this project—the main report
(NCHRP Web Document 30) and the appendix (NCHRP Report 452). NCHRP Web

2
Document 30, Recommended Use of Reclaimed Asphalt Pavement in the Superpave
Mix Design Method, addressed the use of RAP in the Superpave mix design method.
This project included: first, an investigation of whether blending occurs between RAP
and virgin binders or RAP acts like a Black rock in the mix; second, an examination
of RAP binder tests, including RAP binder extraction and recovery procedures, and
the effect of RAP content and stiffness on the properties of blended binder; and third,
an examination of the effect of RAP addition on the properties of asphalt mixtures by
conducting shear, indirect tensile, and beam fatigue tests. NCHRP Report 452,
Recommended Use of Reclaimed Asphalt Pavement in the Superpave Mix Design
Method: Technician’s Manual, corresponds to the appendix of the NCHRP 9-12 final
report and provides a detailed description of the steps involved in designing and
testing HMA mixtures with RAP. The main points discussed include determining
properties of RAP binder, developing RAP mix design, and conducting field quality-
control testing for RAP mixtures.
 NCHRP Project 9-33 (Report 673): A Manual for Design of Hot Mix Asphalt with
Commentary. The main points regarding the use of RAP discussed in NCHRP Report
673 include RAP handling, RAP sampling, RAP blending and variability, RAP
properties, and design of HMA mixtures with RAP.
 NCHRP Project 9-46 (Report 752): Improved Mix Design, Evaluation, and Materials
Management Practices for Hot Mix Asphalt with High Reclaimed Asphalt Pavement
Content. The main objective of this project was to develop a mix design and
evaluation procedure for asphalt mixtures containing high RAP contents, up to
50 percent, to achieve acceptable long-term performance. Changes to existing
American Association of State Highway and Transportation Officials (AASHTO)
standards were also proposed to allow the design of asphalt mixtures with high RAP
content.
 NCHRP Project 9-55: Recycled Asphalt Shingles in Asphalt Mixtures with Warm Mix
Asphalt Technologies. This project is currently ongoing with an expected completion
date of fall 2016. The goal of this project is to develop a mix design and evaluation
procedure that provides satisfactory performance of WMA mixtures with RAS for
project-specific service conditions.

1.2 Scope of the Interim Report

This interim report completes Phase I of NCHRP Project 9-58. Phase I covers Tasks 1
through 3, which include gathering information, designing the laboratory experiment, and
documenting results.
Chapter 1 includes a brief history of the use of recycled materials in HMA and WMA
mixtures, construction and performance challenges associated with using high percentages of
these materials, the use of RAs to overcome these challenges, and the scope of this interim
report.
Chapter 2 provides the results of Task 1 (Subtask 1.1), where technical papers, reports,
and other literature documents (including laboratory and field data) were collected and assessed
to describe the current state of the knowledge on the use of RAs in HMA and WMA mixtures
containing high RBRs. Chapter 2 is divided into three main sections: (a) aging and recycling of
asphalt materials, (b) restoring and characterizing binder rheology with RAs, and (c) improving
and characterizing high RBR mixture performance with RAs.

3
Chapter 3 provides the results of Task 1 (Subtask 1.2), in which a web-based survey of
state DOTs, contractors, and RA suppliers was conducted to assess the current state of the
practice on the use of RAs in HMA and WMA mixtures containing high RBRs. In addition,
information on existing or upcoming field projects utilizing high RBRs, including laboratory and
field performance, was obtained.
Finally, Chapter 4 includes the results of Task 2 (design laboratory experiment). The
laboratory experiment test plan for Phase II has been prepared to characterize the effectiveness of
RAs, its evolution with aging, and the performance of binders and mixtures with high RBRs
through incorporation of RAP, RAS, or a combination of RAP and RAS materials from various
sources. The results from Task 1 were utilized to select the laboratory tests and material
combinations to evaluate the performance of the binders and mixtures with high RBRs.
With approval of the Phase II work plan proposed in this interim report (Task 3), Phase II
will begin with Task 4 (conduct laboratory experiment). Phase III will be conducted somewhat
concurrently with Phase II, as field projects will be identified and materials will be procured for
use in both Phase II and Phase III to tie the laboratory results to field performance.

4
CHAPTER 2.0 BACKGROUND

This chapter provides results of the literature review of technical papers and reports,
including laboratory and field data. Topics covered in this chapter include aging and recycling of
asphalt materials, restoring and characterizing binder rheology with RAs (RAs), and improving
and characterizing high RBR mixture performance with RAs. The information presented in this
chapter along with the information gathered from the surveys whose results are discussed in
Chapter 3 was considered when formulating the laboratory experiment presented in Chapter 4.

2.1 Aging and Recycling of Asphalt Materials

2.1.1 Aging of Asphalt Materials

In a presentation introducing a symposium on asphalt binder durability, Vallerga (1981)


emphasized that pavement cracking results when binders age and become brittle. These aged
binders are contained in recycled materials and provide a critical component in the composite
binder of mixtures with high RBRs. This concept of age embrittlement is more than age
hardening or stiffening as generally characterized through an index or ratio of binder stiffness
with and without laboratory aging. Binders are viscoelastic materials with two rheological
parameters that describe their behavior at any temperature and loading frequency: stiffness (shear
complex modulus [G*] at high and intermediate temperatures or asphalt binder stiffness [S] at
low temperatures) and binder phase angle (δ at high and intermediate temperatures or m-value at
low temperatures). Not only does the stiffness of a binder increase with age, but the phase angle
decreases as well. These two parameters can be utilized to capture age embrittlement by plotting
their response at a particular temperature and frequency on a Black space diagram (Figure 2.1)
with each point representing an aging state and further aging moving the binder rheologically
from the lower right to the upper left of the diagram by increasing G* and decreasing δ. Figure
2.1(King et al. 2012) shows this aging process for three performance grade (PG) 64 binders
(labeled WC, WTX, and GSE) aged in the rolling thin film oven (RTFO) and then for 0, 20, 40,
and 80 hr in the pressure aging vessel (PAV). While these binders are similar in terms of their
PG grade, the WC binder is controlled by S-value for its low-temperature PG grade
(S-controlled), while the WTX binder is controlled by m-value (m-controlled). After 80 hr in the
PAV, the WC binder might represent a typical RAP binder (PG 82), while the WTX binder may
represent a highly aged RAP binder (PG 88+). MWAS and TOAS binders may be shown as
more heavily aged with even higher G* values and lower δ values in Black space (Figure 2.1).

5
Figure 2.1. Binder Rheology Due to Aging and Addition of RAs.

Figure 2.1 also shows that different binders with similar PG grades start at different
locations in Black space and that each binder has a different rate of aging from the lower right of
the diagram to the upper left, as reported by others (Juristyarini et al. 2011; Ruan et al. 2003; Jin
et al. 2011). Although the WC binder has the highest aging index (ratio of G* with and without
laboratory aging), it starts out lower and further to the right in Black space and requires more
aging (longer time in the PAV) to move across the diagram. Thus, aging indexes alone cannot
capture the benefit of starting at a higher δ and lower G* prior to aging.
Figure 2.1 also shows a damage zone where cracking likely begins due to brittle
rheological behavior defined by an intermediate-temperature dynamic shear rheometer (DSR)
parameter called the Glover-Rowe (G-R) parameter between 180 and 450 kPa that correlates to
low ductility values of 5 cm to 3 cm, respectively. These limits were previously related to surface
raveling and cracking by Kandhal (1977). For the binders shown, laboratory aging in the RTFO
plus 15–45 hr in the PAV produced brittle binders with inadequate cracking resistance. This G-R
parameter was originally defined by Glover et al. (2005) as the DSR function (G’/(η’/G’)) and
reformulated for greater practical use by Rowe (2011) in a discussion of Anderson et al. (2011)
as G′/(η′/G′)/ = G* · (cos δ)2/(sin δ), where all rheological properties are referenced to
0.005 rad/s and 15°C.
Anderson et al. (2011) and Hanson et al. (2010) also recognized the importance of phase
angle for cracking resistance by using the bending beam rheometer (BBR) to determine the
difference between the low temperatures (ΔTc) where highly aged binders reach their respective
limits of 300 MPa stiffness (S) and 0.30 m-value. Although BBR test temperatures are almost
35°C below the temperature where ductility and G-R parameters are measured, Anderson
showed that ΔTc correlates well with both. Thus, to characterize the complete rheological
behavior of aged binders contained in recycled materials, DSR or BBR results are needed to
capture both stiffness (G* or S) and phase angle (δ or m-value) at intermediate to low
temperatures for adequate cracking resistance.
Cracking resistance decreases with aging; thus, mixtures with high RBRs that contain
aged binders from recycled materials are expected to have lower cracking resistance. Refinery

6
processing makes binders harder by removing smaller molecules through the physical process of
fractional distillation. Oxidation also makes binders in HMA or WMA harder during refining,
production, construction, and in service through an irreversible chemical reaction with oxygen
that changes the molecular structure of the binder. This reduces the binder’s ability to flow to
relieve stresses, resulting in a much lower phase angle as compared to that due to refinery
processing. The chemical changes during oxidative aging that cause changes in the rheological
parameters include one or more of the following three mechanisms:

 Oxygen removes hydrogen atoms from carbon atoms to increase the aromaticity of
the molecules, often converting polar aromatic molecules into asphaltenes. These
asphaltenes tend to stack in planes, forming large molecular clusters with polar
pi-bonding, little free volume, and few degrees of molecular motion to facilitate flow
to relieve stresses. This mechanism is favored at high temperatures during air
blowing, as is the case for the binder in RAS that has extremely low phase angle
values and thus a significant impact on the rheology of these materials.
 Oxygen atoms add to sulfur atoms in carbon chains to create sulfoxide and later
disulfoxide. This mechanism is rapid but does not have a large impact on binder
rheology.
 Oxygen atoms add to aliphatic carbon atoms attached to aromatic rings to form
functional groups called carbonyls and water through extraction of hydrogen atoms.
The ketones and organic acids that are produced are highly polar with strong
associations through Van der Waals forces with other active polar sites in the binder,
resulting in an increase in apparent molecular weight and associated increase in
stiffness. This mechanism is favored at ambient pavement temperatures and is the
predominant cause of age embrittlement.

2.1.1.1 Modeling of Asphalt Aging—Chemical Characterization


Members of the research team have worked over the past 15 years to develop a model to
predict the progression over time of asphalt pavement toward failure based on binder hardening
as a result of in-service oxidation. Figure 2.2 provides an overview of the model. As discussed,
when asphalt binder oxidizes, its ductility is reduced (Ruan et al. 2003). Asphalt oxidation can be
quantified in terms of asphalt carbonyl area (CA), which can be measured using Fourier
Transform Infrared (FTIR) spectroscopy. In pavements, reduced ductility leads to increased
cracking. Important factors influencing field asphalt oxidation rates include in-service pavement
temperature, asphalt-specific reaction kinetics parameters, asphalt-specific rheology and its
changes due to oxidation, and pavement-specific oxygen diffusion depth (a parameter that
characterizes the average path length for oxygen as it diffuses from air void pores into and
through the asphalt binder). Han et al. (2011) developed an improved model for in-service
pavement temperature as a function of time and depth as part of the oxidation model
development. The model uses hourly data from national weather databases to account for local
climatic conditions. Also, Jin et al. (2013) provided recently published work with field validation
and a further improved thermal and oxygen transport model that utilizes specific binder
properties (reaction kinetics parameters, viscosity hardening susceptibility, and DSR function
hardening susceptibility), weather data (hourly air temperature, wind speed, and solar radiation),
and diffusion depth (a parameter that represents an average binder diffusion distance from each
air void pore surface into the binder) to calculate binder oxidation and hardening (changes in
rheology) as a function of time in service and as a function of depth below the pavement surface.

7
Figure 2.2. Overview of the Model Used to Estimate Time to Pavement Failure.

Asphalt oxidation reaction kinetics may be described in terms of the kinetics model
described by the following equations (Jin et al. 2011):
CA CA M 1 e k t

k Ae

k A e
where CA is the carbonyl area of the binder sample; CAtank is the carbonyl area of the unaged
binder; M = CAtank − CA0 (with CA0 the intercept of the constant rate line); kf is the fast rate
period reaction constant; kc is the constant rate period reaction constant; Af and Ac are the pre-
exponential factors for fast rate and constant rate periods, respectively; Eaf and Eac are the
apparent activation energy for the fast rate and constant rate periods, respectively; and R is the
ideal gas constant. These kinetics parameters can be determined from experimental
measurements of oxidation over time at fixed temperatures and oxygen (or air) pressure. Once
the parameters have been determined, this kinetics model is used in the pavement oxidation
model to calculate changes in CA as the pavement temperature and oxygen pressure change in
the binder.
The asphalt reaction kinetics parameters may be determined by aging thin films of
asphalt at atmospheric pressure in a pressure oxidation vessel (POV) or in a forced draft oven.
Figure 2.3 illustrates the procedure. Three to five aging temperatures are typically used,
and samples are removed according to a specific time schedule, which encompasses
approximately 3 months. The extent of oxidation, in terms of CA, is measured using FTIR
spectroscopy. Numerical estimates of the reaction kinetics parameters are obtained by adjusting
the parameters to achieve the best possible fit of the kinetics model to the experimental results.

8
Figure 2.3. Procedure to Determine Asphalt Reaction Kinetics Parameters.

Of the five oxidation kinetics parameters, Eac can be used as a single parameter to
characterize a binder’s increase in oxidation (as represented by FTIR carbonyl area) over time in
pavements because of interrelations of the other parameters to it. Figure 2.4 shows a model-
calculated comparison between hypothetical asphalts that have different Eac values (and also
differences in other kinetics parameters in accordance with established correlations between the
parameters). A lower Eac results in a higher reaction rate, and vice versa. Thus, the material with
Eac = 65 kJ/mol has the highest rate (CA increases most quickly over time), while the material
with Eac = 105 kJ/mol has the lowest. Also, the lower Eac response declines in terms of rate over
time due to the increasingly stiffer binder (and consequently increasingly lower oxygen diffusion
coefficient) as oxidation proceeds and the resulting balance between oxidation rate and oxygen
diffusion rate (from the air voids into the binder film). Thus, for the higher Eac, the oxidation rate
limits the response, and for the lower Eac, the diffusion rate limits the response. Over a 10-year
period, the resulting oxidation varies significantly for these extremes in Eac values shown for a
single location (Lubbock, Texas) and single pavement depth of 0.5 inch. Other model parameters
are also shown (with fcf as the field calibration factor that accounts for changes in oxygen
transport to the binder due, e.g., to the development of microcracks).

9
Figure 2.4. Example Binder Pavement Oxidation for Four Eac Values. 65 to 105 kJ/mol
Span the Observed Data; 65 and 75 kJ/mol Are Not Statistically Different.

Recently, an accelerated method for determining oxidation kinetics parameters and


rheology was developed by Cui et al. (2014). Figure 2.5 provides an illustration of the
accelerated test method. The central concepts include accelerated asphalt aging using higher
pressure coupled with correlations between the reaction kinetics parameters at 1 atm and at
20 atm PAV conditions. It has been found that the high-pressure Eac may be correlated with the
atmospheric pressure Eac. All other atmospheric pressure reaction kinetics parameters (except M)
are also correlated with Eac. The necessary high-pressure aging can be completed in a standard
PAV in approximately 1 week. The reaction kinetics parameter, M, and the viscosity and DSR
function hardening susceptibilities can be determined using a relatively short POV test. Research
regarding the accuracy and precision of this newly developed method is ongoing.

10
Parallel Testing

Figure 2.5. Accelerated Method to Determine Asphalt Reaction Kinetics and Rheology.

Despite advances in trying to accelerate the procedure for determining oxidation reaction
kinetics parameters, issues remain. Figure 2.6 shows the kinetics parameters Ac and Eac for the
constant-rate kinetics portion of the oxidation reaction. Correlations for both POV (at
atmospheric pressure) and PAV conditions are shown; note the shift between the data for the two
conditions. Also of note is that at atmospheric air pressure conditions (POV), the range of Eac
values is from a little above 60 kJ/mol to above 100 kJ/mol, whereas for PAV conditions, the
range is from 14 kJ/mol to a little above 100 kJ/mol. It is also significant that both sets of data
have been obtained by multiple researchers (PAV conditions are especially notable because the
data have been obtained in three completely independent laboratories). Finally, note that for the
relationships shown, the units of Ac are CA/day for the POV conditions but CA/hr for PAV. The
increase in pressure to 20 atm simultaneously lowers Eac (and by a variable amount, depending
on its value at 1 atm) and increases the pre-exponential value (Ac).

11
1.00E+14
1.00E+13
1.00E+12
Jin 1 atm air
1.00E+11

Ac (POV: CA/day; PAV: CA/hr)


Domke 1 atm air
1.00E+10 Yuanchen 20 atm air
1.00E+09 Huh and Robertson 20 atm air
1.00E+08 Lu et al. 20 atm air

1.00E+07
1.00E+06
1.00E+05
1.00E+04
1.00E+03
1.00E+02
1.00E+01
1.00E+00
0 20 40 60 80 100 120
Eac (KJ/mol)
Figure 2.6. Constant-Rate Reaction Kinetics Parameters for Binders at Both POV and
PAV Conditions.

This type of correlation between Ac and Eac indicates an isokinetic temperature for this
constant-rate (at constant temperature) portion of the oxidation reaction. That is, for a perfect
correlation, there is a temperature at which all materials on this correlation have the same
reaction rate. This concept is shown in Figure 2.7, which shows reaction rate versus 1/RT for two
hypothetical asphalts that lie at the extremes of the Eac range used previously and for both the
POV and PAV conditions. All rates are shown as CA/day units. For POV conditions, the
isokinetic temperature is a little over 100°C; for PAV conditions, it is a little over 90°C. A direct
consequence of this isokinetic condition is that trying to measure kinetics parameters near this
isokinetic temperature is very difficult because measurement uncertainties will bracket the
differences between rates if the measurement temperature is too close to the isokinetic
temperature. Conversely, the farther away from this temperature, the better the determinations of
the kinetics. Thus, the standard PAV temperatures of 90 and 100°C are poor choices for
measuring reaction kinetics. The measurement is improved if multiple temperatures are used and,
in fact, the data in Figure 2.7 were all obtained at 90 and 100°C.

12
Figure 2.7. Constant-Rate Reaction Rates as a Function of Temperature for Binders at
Both POV and PAV Conditions (T in K, R = 8.314 J/mol/K).

Thus, measuring oxidation rates to predict oxidation in pavements is challenging. At


POV conditions, if oxidation and hardening rates are measured near the isokinetic temperature of
100°C, then differences between binders may not be shown. Furthermore, Figure 2.7 shows that
pavement aging in service is shifted considerably along the 1/RT scale so that extrapolating rates
at a higher temperature (near 90°C, say) can also result in considerable error. Rates measured in
the PAV suffer from these same complications plus the additional issue of shifting from PAV
conditions to POV. The work of Cui et al. (2014) addresses this complex issue and provides a
rational procedure for making the conversion. Also complicating this conversion is the fact that
hardening susceptibility at PAV conditions is different from that at POV conditions, and
unfortunately, it appears that the only way to determine values at pavement aging conditions is to
measure them at one atmosphere pressure.

2.1.1.2 Modeling of Asphalt Aging—Rheological Characterization


In addition to the asphalt reaction kinetics parameters, asphalt rheological properties may
be determined from POV (or oven) aged asphalt films. Specifically, asphalt low shear rate
limiting viscosity and the G-R parameter (or DSR function) can be measured using a DSR. For
each of these properties, a linear relationship exists between the natural logarithm of the
rheological property and the chemical characterization in terms of CA (Martin et al. 1990;
Juristyarini et al. 2011). The slope for the viscosity correlation is termed the viscosity hardening
susceptibility (HS). Likewise, the slope for the DSR function correlation is termed the DSR
function HS. Figure 2.8 provides an illustrative example for viscosity HS plotted on a semilog
scale.

13
Figure 2.8. Example Plot Showing How Viscosity HS Is Determined.

Both viscosity and the DSR function (or G-R parameter) are used in the oxidation model
to predict the time to asphalt pavement failure. Viscosity has been correlated with oxygen
diffusivity in asphalt, and higher diffusivity means faster transport of oxygen into the binder for
reaction. The DSR function (or G-R parameter) has been correlated with asphalt ductility, and as
mentioned previously, asphalt ductility reduction has been correlated with an increase in
pavement cracking.
To account for pavement structural effects on oxygen availability for reaction, the
concept of an oxygen diffusion depth is also considered (Jin et al. 2013). The oxygen diffusion
depth is analogous to the asphalt film thickness. To determine the asphalt film thickness, the
volume of asphalt is divided by the surface area of the aggregate; analogously, to determine the
diffusion depth, the volume of asphalt is divided by the surface area of the accessible air voids.
Determination of this surface area is made by analyzing X-Ray computed tomography (CT)
pavement core images. Research toward estimation of diffusion depth without X-Ray CT is
currently underway (Rose et al. 2014).
The core of the binder aging model is a partial differential equation that describes the
growth of CA as a function of time and position within the oxygen diffusion depth. The CA
value is averaged over the oxygen diffusion depth to provide CA as a function of time (at a
specific depth below the pavement surface). Using developed correlations, CA may be related to
the DSR function (or G-R parameter), and from the DSR function increase, progression toward
pavement failure may be predicted.
Figure 2.9 shows the effect that CA growth has on binder rheology, characterized by the
DSRfn (G′/(′/G′), referenced to 0.005 rad/s and 15°C. (This DSRfn value could also be
expressed as the G-R parameter by dividing by the frequency, 0.005 rad/s, and converting MPa
to kPa by multiplying by 1000.) CA represents the oxidation, and DSRfn represents changes in
the rheology (DSRfn) due to oxidation. The changes in rheology are more important, from a
pavement performance perspective, although increases in CA are more fundamental, from the
perspective of the cause of changes in durability. In terms of the DSRfn, a value of 0.0009 MPa/s
corresponds to a G-R parameter of 180 kPa. Thus, the difference between an Eac of 65 kJ/mol
versus 105 kJ/mol is very significant in terms of binder hardening after 10 years of service,
according to this pavement oxidation model, while differences between 65 and 75 kJ/mol are
minimal. (The model has been validated for a small number of field pavement oxidation
measurements; more validation is needed to improve confidence in conclusions drawn from the

14
model calculations.) If RAs change a binder’s oxidation kinetics and consequent hardening rate
enough, then appreciable differences in binder hardening (and thus pavement durability) will
occur over time.

Figure 2.9. Example Binder Pavement DSRfn Hardening for Four Eac Values.

In addition to understanding binder oxidation and resulting hardening, the effects of these
changes on mixture properties, such as stiffness and cracking resistance, are also important.
However, these effects are much less understood than the binder effects. Some of these data that
relate binder and mixture aging will be obtained in this study.

2.1.2 Recycling of Asphalt Materials

Many types of recycled materials are routinely used in HMA mixtures in the United
States. These materials include but are not limited to foundry sand, glass, slag, and tire rubber.
However, one of the most often used recycled materials in the asphalt industry is RAP. It is
expected that with appropriate RAP management and mix design and production considerations,
RAP can be utilized to produce asphalt mixtures that meet normal specification requirements.
Therefore, substantial testing is needed to characterize the recycled material when designing
asphalt mixtures containing RAP. The following properties of RAP are usually determined:

 Binder content (and sometimes grade).


 Aggregate gradation.
 Aggregate angularity (coarse and fine aggregate).
 Aggregate specific gravity.

Typically, binder contents of RAP range between 3 to 7 percent by weight depending on


the mix design of the original pavement. This binder is aged, with the degree of aging dependent

15
on many factors such as pavement age, climate type, mixture air void (AV) content, and binder
grade used in the original pavement. This aging results in a substantial increase in RAP binder
stiffness, which may improve the binder and corresponding mixture performance at high service
temperatures, but it may also reduce cracking resistance at medium and low service
temperatures. Therefore, RAP properties have a significant influence on the properties of
recycled asphalt mixtures. RAP contents in the range of 10–30 percent are commonly used in
recycled asphalt mixtures, and several studies showed that these mixtures perform as well as
virgin asphalt mixtures (Shah et al. 2007; Li et al. 2008; Hajj et al. 2009; West et al. 2009;
Hussain and Yanjun 2012). But with higher RAP contents (i.e., 30 percent and higher), many
studies have indicated an increase in rutting resistance of these mixes along with a decrease in
cracking resistance (Shah et al. 2007; Li et al. 2008; Hajj et al. 2009; West et al. 2009; Valdés et
al. 2011; Mogawer et al. 2012; Mogawer, Austerman, et al. 2013; West and Marasteanu 2013).
Recently, the use of RAS in HMA has received more attention in the asphalt paving
industry since this type of material contains larger amounts of binder, usually 20–30 percent
binder by total weight (Zhou et al. 2012). Two main types of RAS waste that can be recycled in
asphalt pavements exist: TOAS and MWAS. TOAS are shingles removed during re-roofing or
roof removal projects and thus represent the largest source of shingle waste. MWAS are
generated as waste during the manufacturing process, and therefore fewer and more localized
quantities exist.
When RAS is used in asphalt mixtures as a recycled material, it is important to
differentiate between MWAS and TOAS. For asphalt mixtures, MWAS have traditionally been
preferred over TOAS for several reasons. First, the composition of MWAS is better known, there
are usually fewer contaminants, and it is expected to be more consistent (Zhou et al. 2012).
Second, the presence of deleterious or harmful material in TOAS, such as wood, nails, and
asbestos, is expected. Third, the asphalt in TOAS is stiffer than in MWAS because of in service
exposure of the roof to the environment. Therefore, even though the MWAS binder is also stiff
after the production process, where the asphalt is air blown at elevated temperatures and the
combination of heat and oxygen causes the asphalt to age, the stiffness of the TOAS is even
higher. Zhou et al. (2013) found that the average high-temperature grade of TOAS taken from
different places in Texas was 175°C, with one TOAS binder registering a high-temperature grade
beyond 200°C, while the average high-temperature grade of MWAS binders was 131°C.
When using recycled materials, some factors such as mechanical mixing, diffusion, and
compatibility affect the final properties of the recycled mixture. Proper mechanical mixing is
important to ensure the virgin and aged binders are distributed uniformly throughout the asphalt
mixture, and mixing time and temperature play a crucial role in this process. Diffusion enables
the virgin and recycled binder to become a homogeneous blend distributed uniformly throughout
the asphalt mixture. Compatibility between virgin and aged binders is also a requirement for
creating homogeneous binder blends. Compatibility is mainly dependent on the nature and
distribution of the intermolecular associations of binders (Karlsson and Isacsson 2003).
Bennert et al. (2014) evaluated three different strategies that would allow for responsible,
increased use of recycled materials, particularly RAP, in asphalt pavements: (a) using softer
virgin binders, (b) limiting the RAP binder contribution in the mixture (i.e., increasing the virgin
binder contribution), and (c) using performance-based specifications. Using a softer virgin binder
grade was the easiest strategy to implement and resulted in slightly better low-temperature
cracking properties. However, in terms of crack initiation and propagation, the use of softer
virgin binders in high RAP mixtures helped in improving the resisting to the fatigue cracking

16
initiation but had minimal to no benefit in retarding the propagation of cracking. Limiting RAP
binder contribution, as expected, increased the effective asphalt content of the asphalt mixture
and thus increased the durability and fatigue resistance of the asphalt mixture. However, due to
the additional virgin asphalt binder needed, this strategy may not be cost effective. The
performance-based specifications strategy required the final mixture to meet a set of performance
tests that have established criteria. By specifying a minimum acceptable performance, some
assurance that the produced mixture will perform at a relatively high level will be guaranteed.
State DOTs and contractors are continually looking for better ways to incorporate greater
amounts of recycled materials into asphalt mixtures without adversely affecting pavement
performance. Many methods are available for compensating for the aged, stiff binder, and recent
studies focused on using softer virgin binders, WMA additives, or RAs, or increasing virgin
asphalt binder content, as possible solutions to produce higher RBR mixtures.

2.1.3 Degree of Blending between Virgin and Aged Binders

One of the challenges in increasing RBR in asphalt mixtures is the unknown degree of
blending between the virgin and aged binder. The degree of blending, or blending level, is
defined as the percentage of aged binder that is effectively active within the mixture (i.e.,
contributes with the virgin binder to bond the aggregate), and it depends on the aged binder
content, the difference in the stiffness of the virgin and aged binder, and the gradation of
aggregate in the recycled material (Coffey et al. 2013). The degree of blending has a
considerable impact on the volumetric and performance properties of the recycled asphalt
mixture. In mix design practice, if the degree of blending is assumed to be higher than it actually
is, a lower virgin binder content may be used, and thus will produce a stiff mixture that is more
susceptible to fatigue and thermal cracking. The opposite will occur and a soft mixture will be
produced if the degree of blending is assumed to be lower than it actually is, producing a mixture
that is more susceptible to rutting (Coffey et al. 2013). One of the major causes of
low-temperature pavement cracking in recycled mixtures is likely the lack of blending between
the virgin and aged binders (Copeland 2011).
The current standard specification for Superpave volumetric mix design, AASHTO
M323, assumes that complete blending occurs between the virgin and aged binders during HMA
production (i.e., forming a perfectly homogenous blend), and many state agencies assume this
full blending condition. However, as explained subsequently, in most cases, the degree of
blending is partial (Coffey et al. 2013; Kriz et al. 2014), and there is no standard method
available to determine the exact degree of blending. Moreover, current methods employed in the
mix design process in order to include high RBRs in asphalt mixtures are not representative of
plant operations (O’Sullivan 2011). Binder extraction is a good example; extraction is used to
determine the performance grade of the blended binder (virgin and aged binder) to evaluate to
what extent the aged binder will blend and alter the virgin binder grade. However, the extraction
procedure is well controlled in the laboratory and results in a practically full blended binder,
regardless of its actual state in the asphalt mixture produced at the plant. Therefore, it is essential
to establish a procedure that is capable of evaluating the degree of blending, and then
characterizing the effects of recycled materials on virgin asphalt binders without the use of
chemical extraction.
Blending between virgin and aged binders during HMA production is a complex process
and involves two major processes that do not happen instantaneously: first, direct contact
between the two binders is achieved by mechanical mixing; and second, blending of the binders

17
is achieved predominantly by diffusion (Karlsson and Isacsson 2003). Kriz et al. (2014)
explained four possible scenarios of virgin-aged binder contact (due to mechanical mixing) and
blending (due to diffusion) in mixtures, which are shown in Figure 2.10.

Figure 2.10. RAP–Virgin Binder Contact and Blending (Kriz et al. 2014).

Scenario A shows poor contact between the two binders. This mixture may be prone to
fatigue and/or low-temperature cracking because of the lack of uniformity in the rheological
properties of the blended binder throughout the mixture. The same performance is expected in
Scenario B, although the two binders partially blend by diffusion. Conversely, if good contact is
achieved but the two binders do not blend due to very slow diffusion between the two binders
(Scenario C), the properties of the mixture may be difficult to predict since mixture stiffness can
change with time due to continuing blending/diffusion of the two binders, as reported by
Carpenter and Wolosick (1980). The last scenario (D)—in which a good contact and blending is
achieved in the mixture, so the rheological properties of the blended binder will assume
properties of the homogenous blend—is desirable. However, there are regions where the virgin
binder is not in direct contact with the aged binder, and the aged binder in the final blend may
not be uniformly distributed throughout the asphalt mixture. Therefore, this virgin binder volume
basically cannot contribute to aged binder softening, and this may be why a perfectly
homogenous blend (i.e., 100 percent blending) does not occur practically.
Booshehrian et al. (2013) developed a step-by-step blending protocol method based on
developing binder master curves to evaluate the blending occurring between virgin and RAP
binders. The binder master curves were shifted using the Christensen-Anderson model and
further utilized to determine the degree of blending of different RAP plant-produced mixtures
containing 0, 20, 30, and 40 percent RAP. The results indicated an overall good degree of
blending for these mixtures. Nevertheless, a mixture fabricated with a softer binder (i.e.,
PG 58-28) did not show a satisfactory degree of blending compared to the other mixtures (with
PG 64-22), which could be the result of a lower plant discharge temperature and the effect of
temperature on diffusion.
Another case study on WMA evaluation by Copeland et al. (2010) included foaming
WMA technology with RAP. The study included WMA-RAP and HMA-RAP as controls to
compare the mechanical performance of the two types of mixtures, as well as the extent of
blending of virgin and aged binders. Comparisons between measured dynamic complex modulus
(|E*|) versus predicted |E*| were conducted to determine if assumed blending occurred in the
mixtures. The results suggest that WMA-RAP may not have blended completely. At the same

18
time, complete blending may not be necessary to achieve desired performance. Better blending
was observed in RAP-HMA mixtures.
Coffey et al. (2013) evaluated the impact of degree of blending in 25 percent RAP
mixtures. E* for different mixtures with different RAP sources were measured, and fatigue and
rutting performance for each mixture was predicted using mechanistic-empirical pavement
design guide (MEPDG) Level I analysis. Results indicated that the degree of blending has a
negligible effect on fatigue and rutting performance when the degree of blending exceeds 85
percent (calculated based on a previous study of similar mixtures). They concluded that when the
degree of blending is high or the percentage of RAP in a mixture is 25 percent or less, the full
blending assumption will be cost effective without compromising pavement performance.
However, that may not be the case for mixtures with high RBRs.

2.1.4 Improving Binder Rheology in Recycled Materials

In addition to showing the effects of aging on binders in recycled materials, Figure 2.1
also illustrates the possibility of restoring binder rheology in mixtures with high RBRs through
the addition of soft virgin binders and/or RAs to move back from the upper left to the lower right
of the Black space diagram. Selection of a soft virgin binder with good durability and a high
phase angle (S-controlled) would be better than one with a low phase angle (m-controlled) to
restore binder rheology as much as possible and move the starting point on the Black space
diagram lower and more to the right. The more restoration needed (as with RAS binders that are
further to the upper left in Black space), the more likely that a RA is needed in addition to a soft
virgin binder. However, these materials cannot fully compensate for the additional loss in phase
angle caused by structuring of oxidized molecules. Thus, the primary objective of these materials
must be to soften the aged asphalt from the recycled RAP and/or RAS materials (or restore its
stiffness), while at the same time adding molecular mobility to restore as much of the lost phase
angle as possible while maintaining chemical compatibility of the composite binder containing
RA, recycled binder, and virgin binder.
With aging, the composite binder will again rheologically shift from the lower right to the
upper left of the Black space diagram. If durability of the binder is restored to the extent
possible, mixtures with high RBRs should exhibit acceptable performance in terms of resistance
to cracking, although the components of the composite binder (recycled binder, virgin binder,
and RA) must be carefully selected and blended. As shown in Figure 2.1, since the binders that
might be represented by the aged WC binder (typical RAP) and the WTX binder (highly aged
RAP) start at different points on the Black space diagram, the materials needed to restore these
recycled binders would be different, and selection of an m-controlled virgin binder with the
WTX binder would likely age quickly into the damage zone indicated by an inadequate G-R
parameter.
An additive that breaks up the pi-bonded aromatic sheets in asphaltenes would be ideal,
but such chemistry has not been documented. The best alternative is to blend the recycled binder
with soft virgin binders that have very good phase angles (highly S-controlled binders) or RAs
compatible with the binder that have excellent molecular relaxation properties. In addition to
rheological considerations, the additive must meet the following prescreening criteria and not be
(a) hazardous to worker health and safety, (b) volatile per mass loss and flash point
specifications, (c) high in wax content such that it precipitates as a wax, (d) chemically
incompatible such that asphaltenes precipitate or phase separately over time, and (e) unavailable
in sufficient quantities or at a competitive cost.

19
2.1.4.1 Soft Virgin Binders as RAs
Although soft virgin binders with relatively lower viscosities have been used extensively
for compensating for the aged, stiff binder in recycled mixtures, the use of RAs (or rejuvenators
that restore both stiffness and phase angle) promises more benefits. The primary benefit is better
performance for asphalt mixtures since it helps in restoring the physical and chemical properties
of the aged binder, not just softening. Other benefits include (Yu et al. 2014; Zaumanis et al.
2014b):

 Ease of addition to the asphalt mixture in the plant using pump or any liquid additive
dosage system, since RAs do not require heating (in most cases).
 Ability to be used in higher RBRs mixtures, as compared to soft virgin binders.
 Ability to be added at the precise required dosage based on the recycled binder
properties and extent of aging.
 Ability to be added directly to RAP/RAS instead of the virgin binder.
 Often lower costs as a material, and lower storage cost since in most cases RAs do
not require heating.

2.1.4.2 Softening Agents as RAs


The concept of using additives in asphalt mixtures to improve the short and long-term
performance is not new. In particular, the use of additives to restore the rheology of aged binders
has been the subject of a large number of studies. The terms used to describe these additives
include softening agents, reclaiming agents, asphalt modifiers, fluxing oils, extender oils,
aromatic oils, RAs, and rejuvenators (Epps and Holmgreen 1980).
Some authors use the terms softening agents and RAs interchangeably; however,
according to the National Center for Asphalt Technology (NCAT 2009), a distinction between
softening agents and RAs is necessary due to the fact that the softening agents are aimed at
lowering the viscosity of aged binders, whereas RAs are used for restoring chemical and physical
properties of the aged binders. NACT (2009) compiled the following list showing examples of
softening agents and RAs:

 Softening Agents:
o Asphalt flux oils (generally blended with bitumen to reduce the viscosity).
o Lube stock (a fraction of crude oil that has a viscosity similar to lube oils).
o Lubricating or crackcase oil (usually highly aliphatic).
o Slurry oil (bottoms from the catalytic cracking process).
 RAs:
o Lube extracts (highly naphtenic or aromatic fractions removed from lube stock by
solvent extraction).
o Extender oils (aromatic oils from lube stock, mostly used for extending asphalt
rubber blends).

Tabatabaee (2015) compared softening agents and RAs and found that in terms of binder
chemistry, oxidative aging increases polarity and molecular weight, converting the non-polar or
solvent to polar or associated micelles. Softening agents supplement the solvent in the asphalt
colloidal structure by adding low polarity and/or lower molecular weight aromatic, naphthenic,
or paraffinic oils, while RAs break up micelle association and agglomerations to disrupt
structures formed by aging using hydrocarbons. As a result, RAs have an advantage over

20
softening agents in terms of better dosage efficiency for modulus reduction, better reversal of
aged asphalt embrittlement, better restoration of aged asphalt’s ability to heal, and better ability
to reverse shifting of aged asphalt binders’ viscoelastic response to elastic response.
Wang et al. (2014) emphasized that it is necessary for any RA to reverse the chemical
composition of aged binders and recover to that of the virgin asphalt, i.e., reverse the aging
process in which the content of aromatics decreases and the content of asphaltenes increases.
Furthermore, the RA added to the aged asphalt should be able to change the colloidal structure of
aged binders from a gel-type to a sol-type structure (or from a binder that exhibits highly non-
Newtonian behavior to one that exhibits Newtonian behavior). Typically, softening agents do not
have the potential to achieve these two objectives.
The term RA will be used henceforth in this report to describe the material/additive used
to restore the rheology of aged binders.

2.2 Restoring and Characterizing Binder Rheology with RAs

2.2.1 RAs: Definition, Types, and Properties

The Asphalt Institute (1986) defines RA as organic material with chemical and physical
characteristics selected to restore aged asphalt to desired specifications.
Much of the early work with RAs was done by Rostler and co-workers at Witco/Golden
Bear (Rostler and White 1959; Kari et al. 1980) and led to American Society for Testing and
Materials (ASTM) D 4552 for RAs, with six different RA grades covering a range of blending
proportions of a by-product oil from lube processing with a virgin asphalt to restore stiffness (in
terms of viscosity or penetration) while maintaining compatibility. The following purposes for
adding RAs to mixtures with recycled materials include (a) restore the aged binder by decreasing
the stiffness for construction purposes and mixture performance in the field; (b) restore the
recycled mixture in terms of durability or resistance to cracking by increasing the phase angle of
the binder; (c) provide sufficient additional binder to coat the recycled and virgin aggregates; and
(d) provide sufficient additional binder to satisfy mix design requirements (Kandhal and Mallick
1997; Epps et al. 1980; Newcomb et al. 1984; Newcomb and Epps 1981).
Peterson et al. (1994) studied the influence of the main RA components on the
performance of the recycled binder blend and concluded that (a) RAs should preferably be highly
aromatic to allow for improved performance of the pavement as measured by hardening and
temperature susceptibility; (b) saturates reduce the hardening susceptibility of the aromatics
when asphaltenes are absent; (c) if high amounts of asphaltenes are available, the saturates allow
for improved temperature susceptibility but diminish the hardening susceptibility; (d) waxes and
saturates have an adverse impact on the ductility, and the former is reported to affect low-
temperature properties; and (e) metals have little or insignificant effect on the oxidation or
hardening susceptibility, but their strong correlation with asphaltene content sometimes misled
previous research to indicate that they have a direct effect on these variables.
Common sources of RAs that satisfy the prescreening criteria for safety, compatibility,
and availability described previously include:

 Aliphatic, Naphthenic, and Paraffinic Rubber Processing Oils—These by-products of


lube oil production are good candidates, as they are not very volatile, are likely
compatible with binders at lower concentrations, and are likely low in wax content.

21
 Maltenes and Resins from Solvent De-Asphalting—These potential RAs are left after
butane or pentane precipitates the asphaltenes from refinery vacuum tower bottoms.
 Re-refined Waste Lube Oils—While lube oils are too expensive, recovered and
recycled lube oils from diesel train engines are good candidates in terms of
performance as long as compatibility is assessed, especially at higher concentrations
in highly aromatic binders.
 Derivatives of Lipid-Based Vegetable Oils—Bio-based oils from plants such as
soybeans, sunflowers, and palm are expensive but are potential RAs.

The various types of RAs that are currently commercially available, according to NCAT (2014a),
are listed in Table 2.1.

Table 2.1. RA Categories and Types (NCAT 2014a).


Category Types Description

Waste Engine Oil (WEO)


Waste Engine Oil Bottoms (WEOB)
Paraffinic Oils Refined used lubricating oils
Valero VP 165®
Storbit®

Hydrolene®
Reclamite® Refined crude oil products with polar
Aromatic Extracts
Cyclogen L® aromatic oil components
ValAro 130A®

SonneWarmix RJ™ Engineered hydrocarbons for asphalt


Napthenic Oils
Ergon HyPrene® modification

Waste Vegetable Oil


Triglycerides & Fatty Waste Vegetable Grease
Derived from vegetable oils
Acids Brown Grease
Oleic Acid

Paper industry by-products


Sylvaroad™ RP1000
Tall Oils Same chemical family as liquid
Hydrogreen®
antistrip agents and emulsifiers

RAs are generally characterized by laboratory procedures similar to those used for
binders, such as the ones listed in Table 2.2. A study by Shen et al. (2005) evaluated the effect of
conventional RAs on crumb rubber modified (CRM) binders and the feasibility of using these
products for restoring the CRM binders to a specific PG grade. This study analyzed the
properties of the RAs in terms of some of the variables presented in Table 2.2. In particular, the
properties measured by Shen and colleagues included (a) specific gravity, (b) viscosity, (c) flash
point, (d) RTFO (weight loss), (e) viscosity ratio before and after RTFO, (f) compatibility ratio,
(g) saturates, and (h) chemical compatibility.

22
Table 2.2. Typical RA Properties.
Properties Units Standard Authors
Viscosity at 135C Poises AASHTO T316-1 Yu et al. 2014; Shen et
al. 2005; Kari et al.
1980
Viscosity at 25C mPa. s AASHTO T316-1 Lin et al. 2014
Specific gravity NA ASTM D70 Yu et al. 2014; Shen et
al. 2005; Kari et al.
1980
Flash point C ASTM D92 Shen et al. 2005; Kari et
al. 1980
Volatility C ASTM D1160-13
RTFO at 163C Weight loss, % ASTM D2872 Shen et al. 2005; Kari et
al. 1980
Viscosity ratio (ratio of NA ASTM D2872 Shen et al. 2005; Kari et
60°C viscosity after RTFO al. 1980
to that before aging)
Saturates % by weight ASTM D2007-11 Shen et al. 2005; Kari et
al. 1980

In addition to the properties shown in Table 2.2, Yu et al. (2014) identified source
(refined or waste), dominant molecular structure, and relative polarity, and Lin et al. (2014)
identified solids content, boiling point, and density. Other properties of interest for RAs are low
free fatty acid content and moisture, impurities, and unsaponifiables (Yu et al. 2014; Asli et al.
2012). In particular, waste cooking oil filtered in accordance with ASTM D4124 has been used
as a RA by Asli et al. (2012). Details of the chemical properties of the oils are described by Asli
et al., and the main components that were evaluated include:

 Lauric acid.
 Myristic acid.
 Palmitic acid.
 Stearic acid.
 Oleic acid.
 Linoleic acid.
  linoleic acid.
 Cis-11-Eicosenoic acid.
 Heneicosanoic acid.

An early specification for RAs combining selected properties was published in 1980 by
Kari et al. The work was based on characterizing 33 RAs along with their blends with aged
binders. The specification was developed by establishing thresholds that would allow selecting
the RA and blend that best restored the aged asphalt, meaning the one providing the aged asphalt
with properties similar to those of the virgin asphalt used as a reference. Noncompliance with the
specification threshold does not necessarily indicate a poor blend, while compliance also does
not imply complete blending and full restoration of the aged binder. As a result, this specification
was intended to characterize RAs in terms of their physical and chemical properties as presented
in Table 2.3.

23
Table 2.3. Tests Used for the Proposed Specification by Kari et al. (1980).
Functional Reason Test Evaluated
Grade designation and Viscosity at 60°C, cSt
product consistency Viscosity at 135°C, cSt
Flash point
Handling and shipping
Fire point
Rolling thin film (RTF-C),
Oven weight change
Volatility
Smoke point
ASTM D 1160 distillation
Saturates, weight percent
Aniline and mixed-aniline point
Compatibility and solvency
refractive index
Compositional analysis
Durability Viscosity ratio
Accounting Specific gravity 60/60 F

2.2.2 Effect of RAs on Binders with High RBRs

Blending RAP and RAS binders with virgin binders in asphalt mixtures will alter the PG
grade of the blended binder by increasing the high- and low-temperature grades. Adding a RA
will help mitigate the stiffening effect of the aged binder by reducing the PG grade of the
blended binder (Shen and Ohne 2002; Shen, Amirkhanian, and Miller 2007; Tran et al. 2012;
Mogawer, Booshehrian, et al. 2013; Zaumanis et al. 2014a).
The process in which the RAs soften the aged binder has been investigated; however, the
working mechanisms are still not fully understood. In general, the working mechanism depends
on the following processes (Tran et al. 2012):

 Uniform dispersion of the RA within the mixture.


 Diffusion of the RA into the aged binder.

Dispersion is mixing caused by physical processes. The RA will uniformly dissipate


throughout the virgin binder and the mixture through mechanical mixing. Thus this imporant
factor influencing RA efficiency in asphalt mixtures is a function of mixing time. The
mechanical mixing at the plant is usually adequate to achieve uniform dispersion of the RA
within the recycled mixture, although in some cases the aged binder tends to quickly absorb any
hydrocarbon-type liquid before that liquid (i.e., the RA) is uniformly distributed throughout the
mixture (by the diffusion mechanism; Lee et al. 1983).
The second important process influencing RA efficiency is diffusion. Diffusion is the
process where a constituent moves from a higher concentration to a lower concentration. For
instance, when the RA is in direct contact with the recycled material, the aged binder tends to
quickly absorb any hydrocarbon-type liquid in the RA due to the diffusion mechanism. For the
diffusion mechanism, the RA spreads into the aged binder in the following four steps (Carpenter
and Wolosick 1980):

24
1. The RA forms a very low viscosity layer that surrounds the aged binder that is
coating the recycled material particles.
2. The RA begins to penetrate into the aged binder layer, softening the aged binder. The
amount of the RA surrounding the recycled material particles decreases as penetration
continues.
3. Penetration of the RA into the aged binder continues, decreasing the viscosity of the
inner layer and increasing the viscosity of the outer layer of the recycled material
particle.
4. Equilibrium is approached after a certain time, as shown in Figure 2.11.

Figure 2.11. Penetration of the RA into the Outer and Inner Layers of an Aged Binder by
Diffusion (Carpenter and Wolosick, 1980).

As shown in Figure 2.11, diffusion of the RA into the aged binder has a significant
influence on the penetration of the blended binder (virgin plus aged binder), and thus has a
significant effect on the performance of the resulting binder and asphalt mixture. Carpenter and
Wolosick (1980) conducted a staged extraction process to verify the diffusion mechanism
described. In this process, the inner and outer layers of the rejuvenated binder film were
separately extracted at certain time intervals, and the penetration values (at 25°C) for each layer
as a function of time after mixing were plotted. Results indicated that the RAs’ diffusion into the
aged binder occurred during mixing, construction, and a period after construction. Karlsson and
Isacsson (2003) pointed out that the rate of diffusion is governed by the viscosity of the maltene
phase and not the viscosity of the entire aged binder. The rate of diffusion can be accelerated
through the addition of diluent oil fractions or with increased mixing and compaction
temperatures (Oliver 1975). A dissolving-diffusing agent can also be used to produce a highly
diffusible RA (Defeng et al. 2014). The properties of mixtures containing such a RA were shown
to have better fatigue and low-temperature cracking resistance as compared to nondiffusible
RAs, although some negative effect on high-temperature rutting resistance was observed.
The method of adding the RA into the recycled materials may also have an influence on
its diffusion, and thus its effectiveness in the asphalt mixture. Better diffusion of the RA into the
aged binder is expected if the RA is mixed with the recycled materials before combining them
with the virgin binder and aggregate. However, this process is difficult to implement in an
asphalt plant where typically the RA is added to the virgin binder, and subsequently the blend is
added to the mix of virgin aggregate and the recycled materials (Tran et al. 2012). Nevertheless,

25
some asphalt plants are adding the RA to the recycled materials before mixing them with the
virgin binder and aggregate in order to have better diffusion of the RA into the aged binder,
despite the increase in the production cost.
Zaumanis et al. (2014b) pointed out that the RA’s ability to diffuse into the binder film,
and activate the aged binder, is one of the most cited concerns about RAs, and incomplete
diffusion can cause pavement distresses. If the RA does not fully diffuse into the aged binder,
part of the aged binder will remain as Black rock, which may effectively lower the active binder
content in the mixture and lead to a stiff mixture, and thus increased risk of cracking failures. At
the same time, even if the RA will fully diffuse into the aged binder but not before the pavement
is opened to traffic, the inactive dose of RA at the outer layer of the aged binder film may
provide a soft coating to the aggregate and cause early rutting failure in the pavement.
Thus the efficiency of the RA depends on both dispersion and diffusion processes. The
effectiveness of a RA to restore the rheological properties of an aged binder also depends upon a
number of factors, including but not limited to RA type, RA dosage, performance grade of virgin
and aged binders, and mixing time and temperature, as discussed in the following sections.

2.2.2.1 RA Type
Since one of the main purposes of using RAs is to increase the ratio of maltenes to
asphaltenes in the aged binder in order to restore its chemical composition, RA manufacturers
and suppliers produce and supply various types of RAs with different chemical bases and
compositions to achieve this goal. Various types of RAs have different impacts on the chemical
and physical properties of an aged binder, and thus restoring aged binder rheology is a function
of the RA type (Little et al. 1981; Lin et al. 2011; Mogawer, Booshehrian, et al. 2013; Zaumanis
et al. 2014b).
Zaumanis et al. (2013) evaluated the effectiveness of nine RAs in restoring the properties
of aged RAP binders in high RAP content mixtures by determining the penetration and the
kinematic viscosity. The RAs used were categorized by their origin (organic blend, refined
tallow, paraffinic base oil, aromatic extract, naphthenic flux oil, WEO, WEOB, WEO + Fischer-
Tropsch [FT] wax, and distilled tall oil). The RAs were added at two different dosage rates, 9
and 18 percent by weight binder, and thoroughly mixed with the RAP binder after 40 min of
heating at 135°C. Penetration results indicated that the effect of the RA on RAP binders varied
significantly among the different products. The use of refined tallow, for instance, allowed the
blend to reach the penetration level of the virgin binder at a dosage of 9.7 percent, while
naphthenic flux oil, WEO + FT wax, and WEO bottoms were found to be ineffective at reducing
the viscosity of the aged asphalt within a reasonable dosage rate. Kinematic viscosity test results
showed the same trend, where different types of RAs had dissimilar effects. Using the WEO
bottoms had the most notable effect, with significantly higher viscosity in the blend as compared
to other RAs. Other types of RAs significantly decreased the kinematic viscosity of the aged
binder, reaching a comparable level to that of the virgin PG 64-22 binder. In another study,
Zaumanis et al. (2014a) concluded that organic-based RAs require much lower dosage rates as
compared to the petroleum-based RAs to reduce the PG grade. Therefore, switching between
different RAs can play an important role in optimizing the RA dosage.
One of the most important factors with regard to RA type is compatibility. Compatibility
problems may occur when aged binders are blended with virgin binders and RAs because not all
RAs are compatible with the composition of the aged binders (Epps and Holmgreen, 1980;
Holmgreen et al. 1982). Therefore, high compatibility should be achieved when combining RAs
with virgin binder and recycled materials in order to ensure good diffusion and restoration

26
(O’Sullivan 2011). In fact, aged asphalts with more than 30 percent asphaltene content and/or
with relatively low N + A1 contents (less than 35 percent) require specific RAs with chemical
compositions, and Holmgreen et al. (1982) recommended ASTM D1754 for identifying
potential compatibility issues.
The chemical composition of the rejuvenated binder will change, so a blend of aged
binder and RA may be more temperature susceptible. The type of RA along with the nature of
the aged binder determines the temperature susceptibility of the blended binder (Holmgreen et al.
1982).
It is important to determine the best type and composition of the RA that will ensure the
best recovery of the specific aged binder properties. However, based on the previous discussion,
it is difficult to state which type of RA is best without first testing the impact of the RA on the
properties of the specific aged binder.

2.2.2.2 RA Dosage
When using RAs, it is important to determine the amount required to restore, as close as
possible, the aged binder rheology that meets the PG requirements of the virgin binder.
Typically, as the dosage of a RA increases, the impact on the aged binder increases too.
However, that does not imply that this impact will eventually be positive and beneficial. Adding
RAs above rational doses will be costly and even potentially detrimental to the performance of
the asphalt mixture. The most important factor to consider when determining the appropriate RA
dosage is its effect on the PG grade of the aged binder, and consequently on the performance of
the asphalt mixture. Using high doses of RAs will soften the aged binders to a large extent, and
that will be beneficial with respect to the cracking resistance of the asphalt mixtures. However,
that also may negatively impact the mixture’s resistance to rutting/permanent deformation.
Therefore, the dosage should be carefully optimized.
Normally, the dosage of a RA is recommended by the manufacturers based on their
experience, and small dosages are typically preferred. However, the dosage for a particular RA
cannot be fixed for asphalt mixtures with different types and amounts of recycled materials since
it may be affected by other factors such as the binder PG grade, the binder source, and the
aggregate type. Currently, there are many approaches used to determine the optimum RA dosage,
but no standard method is available. Some researchers determine the optimum dosage according
to asphalt blending charts based on the viscosity and/or penetration of the blends of aged binders
and the RA being used (Little et al. 1981; Zaumanis et al. 2013; Zaumanis et al. 2014b; Yan et
al. 2014). Other researchers use the PG system by evaluating the changes in the performance
grade of the aged binder, due to the addition of RAs, and then determining the optimum amount
of the RA needed to restore the performance properties of the aged binder (Shen and Ohne 2002;
Shen, Amirkhanian, and Miller 2007; Tran et al. 2012; Zaumanis et al. 2014a). Typically, the
minimum dosage is determined to ensure sufficient fatigue resistance and low-temperature
cracking resistance (intermediate- and low-temperature PG grade), while a high dosage is
determined to ensure adequate rutting resistance (high-temperature PG grade; Zaumanis et al.
2014a).
Shen and Ohne (2002) presented a comprehensive approach for determining the RA
dosage by considering the performance-related properties of the aged binder at the three
temperature requirements specified by the Superpave binder system. Blends of aged binder and
RA at different dosages (6, 9, and 14 percent by weight of the aged asphalt) were made, and
rejuvenated binders were tested using DSR and BBR testing at three aging conditions: original
(i.e., no lab aging), after RTFO aging, and after RTFO + PAV aging. For high-temperature

27
properties, results indicated that adding the RA caused a significant decrease in the parameter
G*/sin (δ). The reduction was nonlinear with respect to the increase in the RA dosage, which
was more rapid at lower contents as compared to higher contents. However, for intermediate
temperatures, the parameter G*sin (δ) of the blends decreased linearly with an increase in the RA
dosage. The same case was observed for low temperatures, where BBR results indicated that the
stiffness at low temperature decreased linearly, while the m-value increased linearly with an
increase in the RA dosage.
Tran et al. (2012) used a similar approach and determined the optimum RA dosage
needed to restore the performance properties of recycled binders by using the performance grade
system. Blends of aged binders with different dosages of RA (0, 10, 12, and 20 percent by
weight of RAP/RAS binder) were used. For each blended binder, the DSR and BBR were used to
determine the high- and low-temperature properties before and after PAV aging. The lowest
dosage that met the PG grade requirement was considered the optimum. They found that both the
critical low-temperature PG (PGL) and critical high-temperature PG (PGH) decreased linearly
with the increase in the RA dosage. This is different from the high-temperature PG grade
relationship found by Shen and Ohne (2002). Zaumanis et al.’s (2014a) results concur with Tran
et al.’s (2012) results, where the reduction rate for both low and high PG temperature was almost
linear with increasing RA dosage. However, the effect of RA dosage on the intermediate-
temperature PG (IPG) was not linear.
More research is needed to accurately define the relationship between RA dosage and its
effect on restored aged binder rheology, particularly at higher temperatures, because of the
concern of low rutting resistance associated with the use of higher dosages of RAs. Moreover, a
procedure for optimizing the dosage of RAs in asphalt mixtures with recycling materials should
be stablished in order to produce HMA and WMA mixtures with balanced performance.
Finally, it is important to understand how the RA dosage is added to the recycled
mixtures. Some manufacturers state that the RA should be added as a replacement for the binder
in the mixture. This means that if the RA dosage (by the weight of the binder or the mixture) is
0.2 percent as a replacement, for instance, then the same amount of virgin binder should be
removed from the mixture. Other manufacturers state that the RA should be added as an addition
to the mixture, which means that the RAs should be added without adjusting the virgin binder
content.

2.2.2.3 Degree of Binder Aging


Since recycled materials are highly variable in terms of aged binder PG grade, aged
binder content, and aggregate gradation, it is expected that the performance of asphalt mixtures
containing these materials will be variable as well. As explained previously, the type and dosage
of RA that needs to be added will also vary based on the recycled material being used. Therefore,
when using different sources of recycled materials with a specific type and dosage of RA, a
substantial difference in performance may be observed (Little et al. 1981).
The degree of aging, or stiffness, of the aged binders in the recycled materials is one of
the most influential factors that controls the effectiveness of the RA in restoring certain physical
and chemical properties of the aged binders. The higher the stiffness of the aged binder, the
higher the dosage and the lower the viscosity of the RA needed to restore the aged binder
properties. Shen and Ohne (2002) found that adding the same RA dosage to a highly aged binder
decreases the parameter G*/sin (δ) more quickly than for a less aged binder. In other words,
changes in the RA dosage will have a significant influence on the high PG temperature grade of
a highly aged binder, and maximum RA dosage should be controlled carefully. The same trend

28
has been observed for low-temperature properties, where the stiffness at low temperature
decreases and the m-value increases with an increase in the RA dosage, and the extent of the
stiffness decrease or m-value increase depends on the degree of recycled binder aging.
Table 2.4 summarizes selected research studies on the effect of the RAs of aged binders,
as well as the type of relationship between the RA dosage and the degree of improvement in
binder rheology.

Table 2.4. Previous Research on the Effect of RAs on the Stiffness of Recycled Asphalt
Mixtures.
Laboratory
Authors Experimental Plan Main Findings
Tests
Shen and  Virgin asphalt graded by PG and Properties at High Temperature
Ohne penetration as 60–80 was used Penetration  The incorporation of the RA dropped
(2002) and artificially aged by RTFO and Tests the high-temperature grade in a
PAV to target penetrations of nonlinear relationship, where the rate
20(Source A) and 30(Source A). of decline in G*/sinδ was more rapid at
Virgin asphalt graded by lower contents (0–6%) than higher
penetration as 80–100 was used contents (6–14%).
and artificially aged to a target  Although Pen 30(Source A) and Pen
penetration of 30(Source B). 30(Source B) have the same
 The aged asphalts were mixed penetration, the effect of adding the RA
with a RA at different dosages (0, differed a little for the parameter
6, and 11.6% by weight of the G*/sinδ, indicating that the source of
aged asphalt for Pen 20(A), and 0, the aged plays a role here.
6, 9, and 14% by weight of the Properties at Low Temperature
aged asphalt for Pen 30(Source A)  The incorporation of the RA decreased
and Pen 30(Source B). the low-temperature grade in a quite
linear relationship.
 The extent of the stiffness decrease or
m-value increase (in BBR test) was
greater in the harder-aged asphalt (Pen
20[Source A]) than the less one (Pen
30[Source A]).

Shen,  Extracted binder from RAP (PG PG Tests  The incorporation of RAs decreases the
Amirkhanian, 96-4) and an oil-type RA at 0, 5, high- and low-temperature parameters
and Miller and 10% by weight of the aged in a linear relationship.
(2007) binder were used to target asphalt  Using test results and extrapolating the
binder of PG 64-22. linear equation of RA dosage and PG
parameters, the optimum dosage was
determined.
 RA contents of maximum 12.7% was
required to satisfy the high-temperature
parameter, minimum 12.3% to satisfy
the intermediate-temperature parameter,
and minimum 2.9% to satisfy the low-
temperature parameters.
 An optimum dosage in the range of
12.3–12.7% was recommended, and a
dosage of 12.5% was selected.

29
Table 2.4. Previous Research on the Effect of RAs on the Stiffness of Recycled Asphalt
Mixtures (continued).
Laboratory
Authors Experimental Plan Main Findings
Tests
Tran et al.  A RA was added to RAP and RAS PG Tests  A linear correlation between the RA
(2012) binders at different percentages by dosages and the critical high and low
weight of recycled binder to temperatures of the RAP and RAS
determine the amount of the RA binders was observed.
required to restore PG of the  The continuous grade of RAP binder (at
recycled binders to meet the 0% RA) was 99.1–9.2. The addition of
requirements for a PG 67-22. the RA dropped this PG to 83.6–26.4
and 69.2–30.6 for 12% and 20%,
respectively.
 The continuous grade of RAS binder (at
0% RA) was 141.7–10.5. The addition
of the RA dropped this PG to 120.1–
22.8 and 96.9–32.5 for 10% and 20%,
respectively.
 RA dosage of 12% by the total weight
of recycled binders was selected as
optimum dosage based on the critical
low-temperature criteria.
Mogawer,  A virgin mixture (PG 58-28), a PG Tests  The incorporation of 35% RAP and 5%
Booshehrian, recycled mixture (35% RAP and RAS binders into the virgin binder
et al. 5% RAS), and a recycled mixture altered the continuous PG from 60.9–
(2013) (35% RAP and 5% RAS) with 31.1 to 74.2–25.2, and the addition of
three different types of RAs the RAs dropped this PG to 68.4–27.5,
(5.218% by weight of the virgin 68.9–26.5, and 69.1–26.2 for different
binder) were fabricated. RAs.
 The binders in these mixtures were
extracted and tested to determine
their continuous PG grade.
 The recovered binders were
assumed to be already short-term
aged during the recovery process.

Zaumanis et  Extracted binder from RAP (PG PG and  The incorporation of RAs decreases the
al. 94-12) and six different RAs (two Penetration high- and low-temperature parameters
(2014a) petroleum products and four Tests in an almost linear relationship for
organic products with doses from 6 most types of RAs.
to 18% from binder mass) were  The incorporation of RAs decreases the
used to target asphalt binder of PG intermediate-temperature parameter,
64-22. but the relationship is not linear with
the RA dosage.
 From penetration test results, RAs can
reduce the penetration of aged binder to
the target level of virgin binder’s
penetration (78 × 0.1 mm). However,
organic oils require lower dosage to
provide the same effect as petroleum
products.

30
2.2.2.4 Mixing Temperature
It is well established that mixing and compaction temperatures affect the viscosity of
asphalt binders, and low viscosity is important to ensure that the binder will be fluid enough to
cover and adhere to the aggregates. An aged binder, like that in RAP and RAS, needs a higher
mixing temperature, usually above 200°C (390°F), to behave like a fluid. A benefit of adding
RAs is that the mixing and compaction temperatures can be reduced significantly because the
blended binder will be fluid enough to coat the aggregate particles (Romera et al. 2006).
However, as mentioned previously, increasing the mixing and compaction temperatures will
improve the performance of the RA by accelerating its rate of diffusion into the aged binder,
which in turn has a significant influence on the performance of the asphalt mixture. Zaumanis et
al. (2013) found that for one type of RA, waste engine oil, mixing and compaction temperature
should be increased by approximately 22°C to ensure that the blended binder is fluid enough to
coat the aggregates.
The discussion in these subsections reveals how the RA type, RA dosage, properties of
the aged binder, and mixing operations have an effect on the performance properties of the
blended (virgin and aged) binders. All of these factors play a crucial role in restoring the aged
binder rheology in mixtures with recycled materials, and particularly those with high RBRs. If a
RA is added to an aged binder at the appropriate dosage and blended at a suitable temperature for
a sufficient period, the rejuvenated binder may meet the target PG grade and can be used to
produce asphalt mixtures with acceptable performance in terms of cracking and rutting resistance
(Tran et al. 2012). Conversely, if the dosage is not balanced and sufficient blending is not
achieved, the RA may not improve the performance of the aged binder or may even adversely
affect mixture performance.

2.2.3 Characterizing Binder Rheology with RAs

RAs can have varying effects on the chemical, physical, and mechanical properties of
binders and mixtures, and those effects require careful evaluation through laboratory tests. This
evaluation includes testing typically used for binders and mixtures, or other methods developed
through research. This section describes some of these test methods, presenting them in reverse
chronological order.
An evaluation of the correlation between rheology and microstructure for blends with
RAs was carried out by Nahar et al. (2014) using two RA types (i.e., emulsion type and liquid
type). For these materials, analyses were conducted with data gathered through atomic force
microscopy (AFM) and conventional DSR testing to construct master curves of the blends. A
mixing rule based upon mechanical properties of the component materials was considered for the
purpose of calculating the mixing ratios for the materials. Promising results existing between the
blends’ rheology and microstructure indicated that the RA induced not only changes in the
mechanical characteristics of the rejuvenated binder but also microstructural changes that
included the generation of new morphologies of the aged binder or modification of existing
microstructures. These changes were time dependent, which has to be considered in future
research to assess the chemophysical changes occurring at the micro scale. The conclusions
reported the rejuvenation process of the aged binder as concentration dependent, indicating that
the rejuvenation of aged binders can occur at different dosages depending on the RA type.
Another recent study by Wang et al. (2014) investigated the effectiveness of a newly
developed RA, named RA-A, produced through synthesis of furfural extract oil with lube base
oil and additives for improved low-temperature performance. The study evaluated these effects

31
by considering the initial penetration, softening point, and ductility of the binder before and after
aging. The main results suggest that RA-A positively influenced the properties of the aged
binders in comparison to other conventional RAs assessed. Moreover, the compatibility of the
RAs was found to be disparate, as supported by the aromatic carbon ratio and solubility
parameters. This compatibility analysis was mainly based upon previous research by Painter
(1993), in which the use of the atomic contribution method was proposed. This method is
typically used for calculating solubility parameters of asphalt and other complex mixtures.
Additional conclusions focused on the anti-aging properties of RA-A, especially when a
concentration equal to 8 percent was used, allowing for similar characteristics as those from the
original binder studied.
Zaumanis et al. (2013) characterized aged binder blends containing nine RA types and
developed a set of criteria for acceptance of these based on penetration measurements, creep
compliance, tensile strength, and fracture energy. Some novel calculations were introduced with
regard to the characterization of the blends in terms of the penetration index (PI) and the
penetration-viscosity number (PVN) as indicators of oxidative hardening and cracking according
to Burke and Hesp (2011) and Hesp and Shurvell (2010). The RAs studied belong to the
following NCAT defined categories: aromatic, paraffinic, naphtenic, and tall oils. Thresholds
were established for reducing the cracking potential of the RAP extracted binder and improving
the low-temperature performance. More specifically, PI was set as a maximum of 2.23, creep
compliance was set to more than 0.015 GPa, tensile strength was set to more than 2,469 kPa, and
fracture energy was set to more than 9.6 kPa.
Mogawer, Booshehrian, et al. (2013) conducted multiple stress creep recovery (MSCR)
and linear amplitude sweep (LAS) tests for characterizing fatigue of asphalt binders from
mixtures containing RAP, RAS, RAP/RAS, and RAs. In particular, the MSCR results indicated
that the addition of RAs had an effect on the nonrecoverable creep compliance of the virgin
binder, while the results from the LAS test showed an increased number of cycles to failure of
the virgin binders at two strain levels. Furthermore, analysis of the binder master curves
correlated with the data resulting from MSCR and LAS tests, leading to the conclusion that
hardness of the aged binders was reduced with the incorporation of RAs. The aforementioned
results are relevant indications of a softening process occurring due to the addition of the RAs.
As reported by Yu et al. (2014), characterization of RAs’ composition and their
interaction with aged binders can be achieved through conventional testing such as DSR or BBR,
whereas more advanced chemical techniques, such as saturate, aromatic, resin, and asphaltene
(SARA) fractionation and micro-scale assessment by using AFM, can also be used whenever
possible for a deeper understanding of the binder-RA interactions at the macro and micro scale.
Other studies have taken advantage of the separation methods such as high-pressure gel
permeation chromatography (HP-GPC) for characterizing blended binders containing crumb
rubber material (Shen, Amirkhanian, and Lee 2007). The results indicate that HP-GPC profiles
are susceptible to compositional changes when using varying amounts of either RAs or soft
binders, which is reflected by differences in molecular size of the blends (i.e., large molecular
size [LMS], medium molecular size [MMS], or small molecular size [SMS] from quantitative
measurements relative to the molecule size). More specifically, use of any type of aged binder or
rejuvenator led to increased SMS and decreased LMS of the blend, shifting the data to the right
of the HP-GPC profiles. This indicated that restoration of binder rheology occurred, since in
typical binder aging the opposite trend is observed. Furthermore, these test results allowed for
the development of empirical prediction models linking these variables with Superpave binder

32
properties (viscosity and high-temperature failure) that subsequently led to high correlation
between these variables.
Soleymani et al. (1999) developed a method for determining the time-temperature
dependency of blended and rejuvenated binders. According to their results, a linear relationship
can be used for the prediction of PG grade testing parameters (log(G*), δ, log(S), and m-value)
and performance criteria (log(G*/sin(δ)), log(G*sin(δ)), log(S), and m-value). Then, the RAs can
be selected through the blending charts that are based upon the PG grade parameters, in lieu of
the traditional approach involving traditional viscosity/penetration parameters. The authors
established an interaction parameter named G12 from the Irving (1997) model and concluded
that it is significant when low-viscosity RAs are blended with high viscosity binders. In a similar
study, Holmgreen et al. (1982) concluded that the type of aged binder and RA determines the
temperature susceptibility of the blend, while the time and temperature dependency is found to
be significant for the softening effect of the recycled agents. These conclusions have particular
implications with regard to the following: (a) addition of RAs during HMA operations, with
results indicating that most of the blending occurs during mixing, hauling, and compaction; and
(b) type of mixing technique affecting the early life properties of the recycled asphalt mixture.
Another study by Chaffin et al. (1995) developed a method for characterizing the
blending between virgin and aged binders and compared it with the Epps mixing rule (Epps et al.
1980). This method, known as the Grunberg model, can be used to describe the relationship
between viscosity and aged material mass fraction. Based on experimental work, using all the
aged-RA data, an interaction parameter was obtained and compared with Epps’ mixing rule and
the ASTM mixing rule. The results suggested that ASTM D4887 should be used when low-
viscosity asphalt is used as a softening agent, while for more accurate predictions on the amount
of RA needed to reach a desired viscosity, the dimensionless log viscosity (DLV) mixing rule is
superior to the other methods. The DLV was used for reducing differences between the RAs, and
its use proved to be good for normalizing the viscosity values. In particular, nonsignificant
differences were found between all the blends when taking into account the RAs’ chemical
composition and type.

2.3 Improving and Characterizing High RBR Mixture Performance with RAs

As discussed in the previous sections, the characterization of aged and blended binders
and RAs is an important component in understanding the behavior of mixtures with high RBRs.
However, in order to have a better understanding of the performance and properties of recycled
asphalt mixtures, several studies have conducted laboratory evaluations and correlated these
results to field performance. This section describes some of the efforts that have focused on
assessing the distresses that typically affect recycled asphalt mixtures.
The performance of asphalt mixtures containing recycled materials (with or without RAs)
in terms of rutting resistance, fatigue cracking resistance, and thermal cracking resistance does
not always follow the same trends. For example, incorporating high percentages of recycled
materials into asphalt mixtures could improve the resistance of the mixture to rutting, but it will
adversely impact its resistance to fatigue and thermal cracking. Conversely, the addition of RAs
could effectively improve the cracking resistance of the asphalt mixture but decrease its rutting
resistance. Therefore, designing recycled mixtures with balanced performance is important, and
properly designed asphalt mixtures with RAP, RAS, and RAs, even with high RBRs, can be
expected to perform as well as conventional virgin asphalt mixtures. Furthermore, similar to the
effect of RAs on binders with high RBRs, the effect of RAs on the performance of the

33
corresponding asphalt mixtures depends on a number of factors, most importantly, the type and
amount of the RA. Therefore, it is necessary to understand the impact of these factors on the
performance of asphalt mixtures in terms of rutting, fatigue cracking, thermal cracking, and
moisture susceptibility in order to develop a mix design procedure for high RBR mixtures and
achieve balanced rutting and cracking mixture performance.

2.3.1 Effect of RAs on Asphalt Mixtures with High RBRs

2.3.1.1 Stiffness
One of the main barriers to using high RBRs in asphalt mixtures is the increase in
stiffness that occurs after incorporating the recycled materials, which results in a mixture that is
often less workable and difficult to compact in the field (Mogawer et al., 2012). Incorporating
RAs in asphalt mixtures with high RBRs will decrease the stiffness of the recycled asphalt
mixtures to a level close to the stiffness of the virgin asphalt mixture. The decline in stiffness in
asphalt mixtures due to the addition of RAs is not necessarily similar for different types of
recycled materials. Mogawer, Booshehrian, et al. (2013) found that HMA mixtures containing
RAS, and HMA mixtures containing a combination of RAP and RAS, showed less stiffness
reduction after the incorporation of the RA as compared to HMA mixtures with only RAP. This
behavior might be attributed to the fact that RAS binders are usually stiffer than RAP binders.
One factor that should be considered for the change in stiffness of the rejuvenated asphalt
mixtures is the aging process. Tran et al. (2012) found that the effectiveness of a RA in reducing
the stiffness of RAP/RAS mixtures was less with long-term aging as compared to short-term
aging. Dynamic modulus master curve results indicated that the E* master curves for mixes with
RAs were closer to those of the control mixture with virgin binder after short-term aging.
However, after long-term aging, the E* master curves of the same mixtures were closer to those
of the RAP/RAS mixtures (without RA) than to those of the control mixture. This suggested that
RAP/RAS mixtures with RAs aged faster than the virgin binder mixtures, even though they both
had approximately the same level of stiffness before aging.
It should be noted that the impact of the RAs on the stiffness of asphalt mixtures is
different with various loading frequency levels and in different temperature ranges. Im and Zhou
(2014) found that the stiffness characteristic of mixtures containing recycled materials is affected
only over lower loading frequency levels or high temperature ranges.
Table 2.5 shows a number of studies on the impact of the RAs on the stiffness of recycled
asphalt mixtures.

34
Table 2.5. Previous Research on the Effect of RAs on the Stiffness of Recycled Asphalt
Mixtures.
Laboratory
Authors Experimental Plan Main Findings
Tests
Mallick et al.  100% RAP mixture, 100% RAP Dynamic  The addition of the RA dropped the
(2010) mixture with RA, and 100% RAP Modulus stiffness of the RAP mixture at high
mixture with RA and WMA loading frequencies (5 and 10 Hz).
additive.  The addition of the RA increased the
 The dosage of the RA was 0.9% of stiffness of the RAP mixture at lower
total mix. loading frequencies (1 and 0.1 Hz) at the
 The samples were kept at 60°C oven highest testing temperature (54.4°C).
for 5 weeks to facilitate the action of  The WMA additive increased the
the RA. stiffness at low temperatures and high
frequencies, and reduced the stiffness at
high temperature and low frequencies.
O’Sullivan  Virgin mixture, 100% RAP mixture, Dynamic  The addition of the RAs dropped the
(2011) 80% RAP mixture with 0.5% RA Modulus stiffness of the RAP mixture to be even
and 0.5% virgin binder, 90% RAP below the stiffness of virgin mixture.
mixture with 0.5% RA, and 80%  With time, the increase in stiffness was
RAP mixture with 1.0% RA. much lower for mixtures containing RAs
than RAP or virgin mixtures.
Tran et al.  Virgin mixture, control mixture with Dynamic • For the short-term aging, the addition of
(2012) 50% RAP, and control mixture with Modulus the RA dropped the stiffness of the RAS
20% RAP + 5% RAS. and RAP/RAS mixture closer to the
 For each recycled mixture, one type control mixture.
of RA added to the virgin binder  For the long-term aging, the RAS and
with dosage of 12% by the total RAP/RAS mixtures with RA appeared to
weight of recycled binder. age faster than the other mixtures
(control and recycled mixtures).
Mogawer,  Virgin mixture, control mixture with Dynamic  The addition of the RAs dropped the
Booshehrian, 40% RAP, control mixture with 5% Modulus stiffness of the RAP mixture closer to the
et al. (2013) RAS, and control mixture with 35% control mixture.
RAP + 5% RAS.  The drop in stiffness for the RAS and
 For each recycled mixture, three RAP/RAS mixtures was not as much as
different RAs were added to the it was for the RAP mixtures.
virgin asphalt binder with doses
ranging from 0.5 to 1.64% by
weight of the recycled material.
Im et al.  Three control recycled mixtures Dynamic  Regardless of the RA type, the addition
(2014) (19% RAP, 5% RAS, and 13% RAP Modulus of the RA did not affect the stiffness as
+ 5% RAS), and same mixtures with compared to the control mixture for the
three different RAs (R1, R2, and lower temperature range (4°C and 20°C
R3). zone).
 R1 and R2 directly added to virgin  The RAs’ addition decreased the
binder at a rate of 0.6 and 1.5% of stiffness of the recycled mixtures in
total asphalt binder. R3 directly high-temperature ranges (40°C zone), or
added to the recycled materials at a lower loading frequency levels.
rate of 2% by weight of the recycled
materials.

2.3.1.2 Rutting
HMA and WMA mixtures containing recycled materials typically have higher stiffness
compared to conventional mixes, and thus have better rutting resistance (West et al. 2009; Zhao
et al. 2012: Mogawer et al. 2012; Zhou et al. 2011). Due to the effect of RAs on reducing the

35
stiffness of asphalt mixtures, it would be expected that the rutting and moisture susceptibility of
the recycled asphalt mixtures containing RAs would be higher as compared to recycled asphalt
mixtures without RAs. Several researchers have observed this trend (Mogawer, Booshehrian, et
al. 2013; Shen, Amirkhanian, and Tang 2007). However, not all mixtures with RAs will fail to
meet the performance requirements in terms of rutting/permanent deformation as compared to
conventional mixtures without recycled materials (Tran et al. 2012). In fact, some studies have
shown that adding RAs to HMA and/or WMA mixtures containing high RBRs improves their
rutting resistance and moisture susceptibility as compared to virgin mixtures (Im and Zhou 2014;
Zaumanis et al. 2014b; Yan et al. 2014).
An important factor affecting the rutting resistance of asphalt mixtures with RAs is the
degree of aging of the recycled binder. Mogawer, Booshehrian, et al. (2013) found that adding
RAs to HMA mixtures with 35 percent RAP plus 5 percent RAS slightly increased the mixtures’
susceptibility to rutting and moisture damage, while adding the same RAs to HMA mixtures with
only 40 percent RAP or to mixtures with only 5 percent RAS significantly reduced the resistance
to rutting and moisture damage. This behavior was attributed to the fact that the stiffness of the
blended aged binder was different for each type of mixture.
Table 2.6 summarizes selected research studies on the effect of the RAs on rutting
resistance of recycled asphalt mixtures. Additional conclusions on the performance of recycled
mixtures with RAs or soft binders were drawn by Shen, Amirkhanian, and Miller (2007).
Moisture sensitivity and rutting measured through an indirect tensile strength (IDT) test and
asphalt pavement analyzer (APA), respectively, indicated that the mechanical performance of the
mixtures with added RAs through blending charts was superior to that with added soft binders.
Specifically, 40 to 48 percent of RAP could be included in the mixtures when using RA, whereas
10 percent less RAP was reported when using soft binder. Overall, the performance of the RAP
mixtures with RAs was comparable to that of virgin mixtures.

36
Table 2.6. Previous Research on the Effect of RAs on Rutting Resistance of Recycled
Asphalt Mixtures.
Laboratory
Authors Experimental Plan Main Findings
Tests
Shen et al.  Virgin mixture, artificially aged Wheel  The addition of RA significantly
(2004) binder mixture, and artificially Tracking decreased the DS of the mixtures.
aged binder mixtures with RA of Rut Test  The DS of the recycled mixture decreased
2.0 and 7.4% by weight of the (Dynamic by 20% due to the addition of 2.0% RA.
aged asphalt binder. Stability  The DS of the recycled mixture decreased
[DS]* at by 50% due to the addition of 7.4% RA.
60°C)
Shen,  Virgin mixture with PG 64-22, Asphalt  The use of softer binder or the addition of
Amirkhanian, control mixtures with two Pavement RA decreased the rut depths.
and Miller different sources of RAPs (each Analyzer  The rut depths of the mixtures containing
(2007) with 15 and 38% RAP) with PG (APA) RA were, in most cases, smaller than
52-28, and the same RAP those using softer binder.
mixtures with PG 64-22 and with
oil-type RA (12.5% by RAP
binder).
Lin et al.  35% RAP mixtures with three Marshall  Mixtures with RA-75 and RA-250 had
(2011) different types of RAs at four Stability much better stability value as compared to
different dosages (10, 20, 30, and mixtures with RA-25, regardless of the
40% by total weight of binder). amount of RA.
 The RAs used were RA-25, RA-  With increasing RA dosage from 10 to
75, and RA-250 with viscosity of 40%, the reduction in stability for RA-75
3620, 6340, and 31900 cSt at and RA-250 was about 25%, while the
60°C, respectively, based on reduction was more than 55% for RA-25.
ASTM classification (D4552).
Tran et al.  Virgin mixture, control mixture APA  The addition of the RAs to the recycled
(2012) with 50% RAP, and control mixtures increased the mixtures’
mixture with 20% RAP + 5% susceptibility to rutting. However, all
RAS. these mixtures exhibited APA manual rut
 For each recycled mixture, one depths less than 5.5 mm, which is
type of RA added to the virgin adequate to pass in terms of rutting.
binder with dosage of 12% by
the total weight of recycled
binder.
Mogawer,  Virgin mixture, control mixture Hamburg  The addition of the RAs to the 35% RAP
Booshehrian, with 40% RAP, control mixture Wheel + 5% RAS slightly reduced the mixtures’
et al. (2013) with 5% RAS, and control Tracking resistance to rutting and moisture damage.
mixture with 35% RAP + 5% Device  The addition of the RAs to the 40% RAP
RAS. (HWTD) and 5% RAS mixtures increased the
 For each recycled mixture, three mixtures susceptibility to rutting and
different RAs were added to the moisture damage.
virgin asphalt binder with doses  These two mixtures did not have enough
ranging from 0.5 to 1.64% by stiff binder to balance the impact of the
weight of the recycled material. RAs.

37
Table 2.6. Previous Research on the Effect of RAs on Rutting Resistance of Recycled
Asphalt Mixtures (continued).
Laboratory
Authors Experimental Plan Main Findings
Tests
Yan et al.  Virgin mixture and 30, 40, and Wheel  At a fixed 40% RAP content, the addition
(2014) 50% RAP mixtures. Tracking of RA-2 resulted in better rut resistance
 Three types of RAs used: RA-1, Rut Test compared to RA-1 and RA-3.
RA-2, with viscosity of 0.493 (DS*)  RAP mixtures containing RAs had slightly
and 0.254 Pa.s at 60°C, and RA- higher DS as compared to control mixture.
3, which consists of 40% RA-1
and 60% RA-2.
Im et al.  Three control recycled mixtures HWTD  For the 5% RAS mixtures, the use of R2
(2014) (19% RAP, 5% RAS, and 13% and R3 RAs significantly improved the
RAP + 5% RAS), and same rutting/moisture resistance of the
mixtures with three different mixtures, but the performance of the
RAs (R1, R2, and R3). mixture with R1 was similar to the control
 R1 and R2 directly added to mixture.
virgin binder with a dosage of  For the 19% RAP and 13% RAP + 5%
0.6 and 1.5% of total asphalt RAS mixtures, the addition of the RAs did
binder by weight. R3 directly not significantly affect the rutting
added to the recycled materials resistance.
with a dosage of 2% by dry
weight of the recycled materials.
* The number required for the passing wheel to produce vertical plastic deformation of 1 mm on the surface of a test sample.

2.3.1.3 Fatigue and Reflective Cracking


Asphalt mixtures with high RBRs typically exhibit lower resistance to cracking than
conventional mixtures, and cracking resistance reduction is proportional to the amount of
recycled materials in the mixture (West et al. 2009; Mogawer et al. 2012; Mogawer,
Booshehrian, et al. 2013). The addition of RAs improves the cracking performance of asphalt
mixtures containing recycled materials by mitigating the stiffness of the blended binder. The type
of the RA used is an influential factor on the degree of improvement. Mogawer, Booshehrian, et
al. (2013) found that the improvement in the cracking performance of asphalt mixtures
associated with the incorporation of one type of RA was relatively lower than the improvement
with different types of RAs, regardless of the type and percentages of recycled materials. Lin et
al. (2011) found that using RA RA-75 (per ASTM D4552 RA grades) increased the indirect
tensile strength better than when RAs RA-25 or RA-250 were used on asphalt mixtures
containing 35 percent RAP. They concluded that RA-75 offered the most significant
improvement in cracking performance. Similar results were observed by Im et al. (2014), where
some RAs exhibited better performance than others in terms of improving reflective cracking
(i.e., Overlay Test) resistance for HMA and WMA mixtures.
In the past few years, several studies were conducted to evaluate the effect of RAs on
fatigue and reflective cracking resistance of recycled mixtures. These studies are summarized in
Table 2.7.

38
Table 2.7. Previous Research on the Effect of RAs on Fatigue and Reflective Cracking
Resistance of Recycled Asphalt Mixtures.
Laboratory
Authors Experimental Plan Main Findings
Tests
Lin et al.  35% RAP mixtures with three Indirect  Mixtures with RA-75 and RA-250 had
(2011) different types of RAs (RA-25, RA- Tensile better cracking properties as compared
75, and RA-250) at three different Strength Test to the mixture with RA-25 under the
dosages (10, 18, and 36%, same conditions to achieve viscosity of
respectively). The doses were 2000 poises.
determined to achieve the viscosity
goal of 2000 poises for an aged
binder mixed with RA.
 The RAs used were RA-25, RA-75,
and RA-250 with viscosity of 3620,
6340, and 31900 cSt at 60°C,
respectively, based on ASTM
classification (D4552).
Tran et al.  Virgin mixture, control mixture Energy Ratio Top-Down Cracking
(2012) with 50% RAP, and control mixture Test  The addition of the RAs improved the
with 20% RAP + 5% RAS. fracture properties of the recycled
 For each recycled mixture, one type mixtures.
of RA added to the virgin binder
with dosage of 12% by the total
weight of recycled binder.

Texas Reflective Cracking


Overlay Test  The addition of the RAs increased the
(OT) average number of cycles to failure for
the 50% RAP and 20% RAP + 5% RAS
mixtures as compared to recycled
mixtures without RA. However, the
differences in the number of cycles to
failure among the recycled mixes were
not statistically significant
Mogawer,  Virgin mixture, control mixture OT  The addition of the RAs improved the
Booshehrian, with 40% RAP, control mixture cracking performance of the RAP, RAS,
et al. (2013) with 5% RAS, and control mixture and RAP/RAS mixtures.
with 35% RAP + 5% RAS.  The RA type affected the degree of
 For each recycled mixture, three improvement.
different RAs were added to the
virgin asphalt binder at doses
ranging from 0.5 to 1.64% by
weight of the recycled material.
Yan et al.  Virgin mixture and 30, 40, and 50% Beam  The addition of the RAs highly
(2014) RAP mixtures. Fatigue Test improved the fatigue properties of
 Three types of RAs used: RA-1, (Four-Point recycled mixes.
RA-2, with viscosity of 0.493 and Bending  At a fixed 40% RAP content, the
0.254 Pa.s at 60°C, and RA-3, Test) mixture with RA-2 had better fatigue
which consists of 40 % RA-1 and resistance compared to mixtures with
60 % RA-2. RA-1 and RA-3.
 With using RA-2, fatigue life first
increased and then decreased with
increases in RAP content.

39
Table 2.7. Previous Research on the Effect of RAs on Fatigue and Reflective Cracking
Resistance of Recycled Asphalt Mixtures (continued).
Laboratory
Authors Experimental Plan Main Findings
Tests
Im et al.  Three control recycled mixtures OT  The addition of the RAs yielded higher
(2014) (19% RAP, 5% RAS, and 13% OT cycles (from approximately 110% to
RAP + 5% RAS), and same 300% improvements) than the control
mixtures with three different RAs mixtures.
(R1, R2, and R3).
 R1 and R2 directly added to virgin
binder with a dosage of 0.6 and
1.5% of total asphalt binder by
weight. R3 directly added to the
recycled materials with a dosage of
2% by dry weight of the recycled
materials.

2.3.1.4 Low-Temperature Cracking


Asphalt mixtures with high RBRs typically appear to develop thermal stresses more
quickly than virgin asphalt mixtures, and therefore have less resistance to thermal cracking. The
incorporation of RAs is helpful in improving the low-temperature cracking resistance of the
mixtures (Tran et al. 2012; Mogawer, Booshehrian, et al. 2013; Yan et al. 2014).
As in the case of fatigue cracking, the type of RA is a factor that affects the degree of
improvement in low-temperature cracking resistance (Zaumanis et al. 2013; Yan et al. 2014).
Zaumanis et al. (2013) utilized an indirect tensile creep compliance test at −10°C, indirect tensile
strength, and fracture energy in characterizing the stiffness of asphalt mixtures containing 100
percent RAP, and thus the resistance to low-temperature cracking. The less stiff the material at
low temperature, the lower the possibility of it developing low-temperature cracking. They found
that creep compliance of the mixtures increased after the addition of a RA, and the increase
varied among different types of RAs. The paraffinic base oil, naphthenic flux oil, and refined
tallow showed the highest increase in creep compliance (and thus better reduction in low-
temperature cracking potential) as compared to the other six types of RAs. However, in the case
of the indirect tensile strength test, the addition of the organic blend, distilled tall oil, and
naphthenic flux oil yielded the highest increase in the tensile strength as compared to the other
six types of RAs. Some of the RAs caused a decrease in the tensile strength. The fracture energy,
in turn, showed the highest values when WEO, distilled tall oil, and aromatic extract were used.
Higher fracture energy means higher energy required to initiate the fracture of the mixture. It was
suggested by the authors that the RA that maintained or increased the creep compliance, without
reducing the tensile strength and fracture energy could be used in reducing the embrittlement of
the recycled asphalt mixture, and thus improving its low-temperature cracking resistance.
Table 2.8 shows a number of studies on the effect of the RAs on the low-temperature
cracking resistance of recycled asphalt mixtures.

40
Table 2.8. Previous Research on the Effect of RAs on Low-Temperature Cracking
Resistance of Recycled Asphalt Mixtures.
Laboratory
Authors Experimental Plan Main Findings
Tests
Shen et al.  Virgin mixture and artificially Thermal Stress  The addition of RA significantly
(2004) aged binder mixture with three Restrained improved the low-temperature fracture
different doses of RA (0, 2.0, and Specimen Test properties for the recycled asphalt
7.4%) by weight of the aged (TSRST) mixtures.
asphalt binder.  The fracture temperature of the recycled
mixture decreased from −22.0 to
−25.0°C due to the addition of 2.0%
RA.
 The fracture temperature of the recycled
mixture decreased from −22.0 to
−29.5°C due to the addition of 7.4%
RA.
Tran et al.  Virgin mixture, control mixture Indirect Tensile  The addition of the RAs reduced the
(2012) with 50% RAP, and control Test critical failure temperature.
mixture with 20% RAP + 5%  The control mixture exhibited the
RAS. lowest critical failure temperature,
 For each recycled mixture, one followed by the 50% RAP mixture with
type of RA added to the virgin RA, and then the 20% RAP plus 5%
binder with dosage of 12% by the RAS mix with RA.
total weight of recycled binder.
Mogawer,  Virgin mixture, control mixture TSRST  The addition of the RAs considerably
Booshehrian, with 40% RAP, control mixture helped to mitigate the loss in the low-
et al. (2013) with 5% RAS, and control temperature cracking due to the
mixture with 35% RAP + 5% recycling materials.
RAS.
 For each recycled mixture, three
different RAs were added to the
virgin asphalt binder with doses
ranging from 0.5 to 1.64% by
weight of the recycled material.
Zaumanis et  One virgin mixture and nine Indirect Tensile  The addition of five types of RAs
al. (2013) 100% RAP mixtures with RAs. Creep helped in maintaining or increasing the
 Nine different types of RA used Compliance low-temperature creep compliance and
with the RAP mixtures, and the Test at the same time increasing the indirect
dosage that returns the aged Indirect Tensile tensile strength and fracture energy.
binder penetration to reach virgin Strength Test  Improving low-temperature
binder penetration (using PG 64- performance depended on the type of
22) was selected to be the the RA used.
optimum dosage.

Hajj et al.  Virgin mixture, recycled TSRST  The addition of the RA improved the
(2013) mixtures (15, 50, and 100% low-temperature resistance where a
RAP), and recycled mixtures decrease in the fracture stress and
with RA (1.5% by weight of microcracking stress was observed,
RAP). with a shift to the colder side in the
fracture temperature.

41
Table 2.8. Previous Research on the Effect of RAs on Low-Temperature Cracking
Resistance of Recycled Asphalt Mixtures (continued).
Laboratory
Authors Experimental Plan Main Findings
Tests
Yan et al.  Virgin mixture and 30, 40, and Three-Point  The addition of the RAs improved the
(2014) 50% RAP mixtures. Bending Test low-temperature cracking resistance.
 Three types of RAs used: RA-1,  At a fixed 40% RAP content, the
RA-2, with viscosity of 0.493 and mixture with RA-2 had better cracking
0.254 Pa.s at 60°C, and RA-3, resistance compared to mixtures with
which consists of 40% RA-1 and RA-1 and RA-3.
60% RA-2.  With using RA-2, failure strain value
decreased with increasing RAP
contents.

2.3.1.5 Moisture Susceptibility


Moisture damage is one of the primary distresses in asphalt pavements in some climates
and for some material sources and combinations. For asphalt mixtures that are susceptible to
moisture, the internal bond between the binder and aggregate is weakened in the presence of
water, which may lead to stripping. Several studies have indicated that HMA and WMA
mixtures containing recycled materials have acceptable moisture resistance (Shen, Amirkhanian,
and Miller 2007; Hajj et al. 2009: Zhao et al. 2012). The adequate moisture susceptibility
performance may be related to the fact that the aggregate in the recycled materials is covered and
protected by the aged binder; hence, the bond between aggregate and aged binder is stronger than
the bond between aggregate and virgin binder, making the recycled mixture less vulnerable to
moisture damage (Zhao et al. 2012). Other studies have also shown that incorporating RAs into
asphalt mixtures with high RBRs improves their moisture resistance (Yan et al. 2014; Im and
Zhou 2014). This same conclusion was reached by Tran et al. (2012), who found that the use of
RAs in RAP/RAS mixtures slightly increased the tensile strength ratio (TSR) of the recycled
asphalt mixtures.
Table 2.9 summarizes selected research studies on the effect of the RAs on the moisture
susceptibility of recycled asphalt mixtures.

42
Table 2.9. Previous Research on the Effect of RAs on Moisture Susceptibility of Recycled
Asphalt Mixtures.
Laboratory
Authors Experimental Plan Main Findings
Tests
Shen,  Virgin mixture with PG 64-22, Tensile  Mixtures containing RAP (with soft
Amirkhanian, control mixtures with two Strength Ratio binder or with RA) had the same level
and Miller different sources of RAPs (each (TSR) Test of moisture susceptibility as the virgin
(2007) with 15 and 38% RAP) with AASHTO T283 mixture.
PG 52-28, and the same RAP
mixtures with PG 64-22 and
with 12.5% oil-type RA.
Tran et al.  Virgin mixture, control mixture TSR  The addition of the RAs in the recycled
(2012) with 50% RAP, and control AASHTO T283 mixtures did not negatively affect the
mixture with 20% RAP + 5% TSR values but slightly increased them.
RAS.
 For each recycled mixture, one
type of RA added to the virgin
binder with dosage of 12% by
the total weight of recycled
binder.
Hajj et al.  Virgin mixture, recycled Dynamic  The addition of the RA to the RAP
(2013) mixtures (15, 50, and 100% Modulus* mixtures improved their resistance to
RAP), and recycled mixtures moisture damage after three freeze-
with RA (1.5% by weight of thaw cycles.
RAP).
Yan et al.  Virgin mixture and 30, 40, and TSR  TSR value decreased with increasing
(2014) 50% RAP mixtures. AASHTO T283 RAP contents.
 Three types of RAs used: RA-1,  At a fixed 40% RAP content, the
RA-2, and RA-3 based on mixture with RA-2 had higher moisture
ASTM classification (D4552). stability compared to mixtures with
 The viscosity of RA-1 and RA-2 RA-1 and RA-3.
is 0.493 and 0.254 Pa.s at 60°C.
RA-3 consists of 40% RA-1 and
60% RA-2.
* Testing and comparison of the dynamic modulus at the moisture-conditioned and unconditioned stages.

2.3.2 Characterizing High RBR Mixture Performance with RAs

As discussed in the previous sections, the characterization of aged and blended binders
and RAs is an important component in understanding the behavior of mixtures with high RBRs.
However, in order to have a better understanding of the performance and properties of recycled
asphalt mixtures, several studies have conducted laboratory evaluations and correlated these
results to field performance. This section describes some of the efforts that have focused on
assessing the distresses that typically affect recycled asphalt mixtures.

2.3.2.1 Aging Evaluation


Research on the effect of long-term aging in asphalt mixtures with high RAP HMA (i.e.,
20 percent, 30 percent, and 40 percent) and corresponding mechanical characterization has been
studied by Daniel et al. (2013). Their research focused on the assessment of fatigue and complex
modulus via the asphalt mixture performance tester (AMPT) in uniaxial tension and unconfined
uniaxial compression, respectively. For data analysis, the simplified viscoelastic continuum
damage (S-VECD) model developed by Underwood et al. (2010) was used. This model allows

43
for a rapid characterization and prediction of the response to fatigue in asphalt mixtures.
Additionally, the method is relevant due to the combination and improvement of previous
models of controlled-stress, controlled-strain, and controlled-crosshead push-pull tests that
composed the previous S-VECD model and allowed for concluding that the damage
characteristic curve is a material property, independent of temperature and test type. Due to the
importance of aging on the performance of recycled mixtures, this variable was addressed in a
separate study with the Global Aging System (GAS) method that correlates the changes in
viscosity in the mixture to the dynamic modulus. This method was developed by Mirza and
Witzak (1995) and is currently part of the MEPDG software to account for temporal changes in
the physical properties of pavements.

2.3.2.2 Rutting and Moisture Sensitivity


Mechanical characterization of WMA mixtures containing RAS was addressed by Buss
et al. (2014). Use of chemical additives and foaming was included in the methodology, as well as
the construction of test sections in 2009 and 2010 in Indiana and Iowa, respectively. According
to the test results, the WMA mixtures with RAS showed improved rutting performance, similar
or superior dynamic modulus, and better moisture damage resistance. Their recommendations
included the construction of new pavement sections with WMA and RAS in order to gather
additional evidence to support those conclusions.
Additional rutting test studies employed Marshall-compacted mixtures. Hussain and
Yanjun (2012) evaluated the performance of HMA mixtures containing high RAP percentages
(i.e., 30, 45, 60, and 100 percent) and 19.0 mm nominal maximum aggregate size. The
conclusions established good or superior overall performance of the recycled mixtures as
compared with control mixtures in terms of the Marshall flow and stability values. In particular,
the stability for the 100 percent RAP mixture was more than double that of the virgin mixture.
Nevertheless, some of the mixtures containing more than 45 percent RAP reported
noncompliance with the flow criteria, which may suggest problems related to the stiffness of the
aged binder. Further research was recommended on trial sections to compare laboratory and field
performance.
Xiao et al. (2007) studied HMA with RAP by means of the IDT strength test, leading to
results that highlight the positive effect of RAP, as supported from higher stiffness and IDT
strength values that indicate better rutting resistance as compared with conventional HMA
mixtures. Additional improvements were reported on the workability of the recycled mixtures as
observed during the mixing process as well as better rutting resistance (i.e., lower rut depths as
compared to the virgin mixtures) when using asphalt rubber and 25 percent RAP in the APA and
IDT strength tests.

2.3.2.3 Fatigue Cracking


Due to the fact that many factors influence the cracking performance of recycled asphalt
mixtures, the establishment of a single cracking requirement is not trivial (Im and Zhou
2014).With regard to the improved cracking resistance of several types of mixtures (virgin, 50
percent RAP, and 20 percent RAP − 5 percent RAS), Tran et al. (2012) obtained improved
performance when a specific RA (i.e., Cyclogen® L) was added. In addition, no negative effects
on the moisture susceptibility or rutting, measured through tensile strength ratio (TSR-AASHTO
T 283) and asphalt pavement analyzer (APA-AASHTO TP 63-07), respectively, were observed.
The authors concluded that 12 percent by total weight of recycled binder of this particular RA

44
was the optimum dosage. Nevertheless, it is important to conduct additional testing with other
RAs and recycled materials.

2.3.2.4 Low-Temperature Cracking


Performance test results based on the Overlay Test (OT) and Thermal Stress Restrained
Specimen Test (TSRST) showed improved reflective and low-temperature cracking performance
for the mixtures containing RAs. Nevertheless, some of the tests conducted to evaluate
performance in terms of rutting and moisture susceptibility, measured through the Hamburg
Wheel Tracking Test (HWTT), showed a negative effect after the addition of RAs to mixtures
with 40 percent RAP and 5 percent RAS (Mogawer and Booshehrian et al. 2013).
An additional study by Shen et al. (2004) validated the previous findings (Shen and Ohne
2002) through laboratory performance tests including HWTT and TSRST. The results indicated
that performance of HMA mixtures containing RAs and fabricated following the Marshall
procedure had similar performance as compared to the virgin asphalt mixtures. More
specifically, the dynamic stability (DS) at 60°C decreased as a function of increased RA content.
Moreover, low-temperature fracture properties of the mixtures were positively influenced by the
addition of RA, yielding an overall performance equivalent to that of virgin mixtures. The RA
dosage also had a significant effect on the mixture properties. Finally, the effects of the RA on
compaction were assessed, indicating nonsignificant changes on the compaction properties for
recycled and virgin mixtures.

2.3.2.5 Field Studies


Evaluation of field sections with recycled materials is important, especially to compare
their performance against conventional HMA mixtures. Johnson and Hesp (2014) assessed the
cracking performance of binders typically used in Ontario highway projects by means of regular
binder grading via AASHTO M320 grading, DSR, and BBR analyses. In their study, the
incorporation of waste engine oil was evaluated to determine its influence on the cracking
performance of already-in-place pavement sections. For the 20 recovered binders, cracking was
mainly caused by physical hardening of the binder, which eventually turned into cracking. The
chemical evaluation indicated that this hardening was mainly due to the presence of zinc,
molybdenum, and other common metals typically encountered in WEO. As a result, cautious use
of WEO for modification of low asphaltenes binders was emphasized, which restricts the use of
this by-product as a countermeasure for cracking performance.
Anderson et al. (2011) compared the long-term performance of 19 high RAP field
sections subjected to varying climatic and loading conditions across the United States and
Canada and reported an overall good performance with time. After both quantitative and
qualitative assessments, most of the field sections were concluded to have similar or even better
performance as compared to virgin field sections. The distresses that were more frequently
observed on the high RAP field sections included slight rutting and fatigue cracking, but they did
not diminish the overall performance. An economical evaluation was also included in the study
and indicated that maintenance costs were not significantly higher on the RAP sections as
compared to the virgin field sections.
Daniel et al. (2010) conducted a study in cooperation with the New Hampshire
Department of Transportation that observed the effect of varying RAP percentages on binder
properties that included the high and low PG grades, as well as failure and critical cracking
temperatures. The failure temperatures are those at which the binder does not comply with the
specifications at high and low temperatures. The critical cracking temperature is determined

45
through the Direct Tension Test and corresponds with the low temperature at which the binder
fails as a result of the tensile stress exceeding its tensile strength. The conclusions from the study
were (a) high- and low-temperature PG grades were the same or one grade higher for all the RAP
binders; (b) failure temperatures and critical cracking temperatures slightly changed for the
different RAP percentages; (c) the increase in RAP binder content allowed for better critical
cracking temperature, but more testing was recommended to validate this finding; and (d) a
calculation for the percent binder replacement was suggested for normalizing the RAP asphalt
content and mixture asphalt content.

46
CHAPTER 3.0 STATE DOT, CONTRACTOR, AND SUPPLIER WEB-BASED
SURVEYS

Web-based surveys were launched to assess the current state of the practice on the use of
RAs in HMA and WMA containing high RBRs. The questionnaires, presented in Appendix A,
were sent to representatives of state DOTs, contractors, and RA (RA) suppliers after review and
approval from the NCHRP panel. Contacts for state DOT representatives were obtained from the
membership list of the AASHTO Subcommittee on Materials. A list of contractors that had
collaborated on previous NCHRP projects (i.e., 9-49 and 9-52) was compiled to generate the
contacts for this group. The RA supplier contact list was generated using representatives from
various companies known to the research team or that had communicated with the research team
in an effort to make them aware of their product. At least one company producing each type of
RA was included in the contact list (NCAT 2014a).

3.1 State DOT Web-Based Survey

The survey was divided into two main topics: (a) the use of high RBRs, and (b) the use of
RAs. The first topic included separate questions for RAP and RAS on current use, typical
content, standards and test methods for characterization, and barriers to employing higher levels
of the recycled materials. In addition, questions about RAP/RAS mixtures with regard to RBR
requirements, mix design, performance tests, and pavement distress were included in the first
topic. The second topic inquired about current use of RAs, types employed, barriers to utilizing
them, standards and test methods for characterization of RAs and binders containing RAs, and
performance tests for mixtures with RAs. Finally, interviewees were asked about upcoming field
projects or knowledge of instances of premature pavement distress or failure and their
availability to participate further in the study. Based on the response to this question, a follow-up
email or phone call will be sent to selected individuals.
The survey was sent to representatives from all 50 states in addition to Washington, D.C.
and Puerto Rico on October 31, 2014, with a response deadline of November 14, 2014. A
reminder was sent at the beginning of December to those state DOTs that had not submitted a
response, with a revised deadline of December 19, 2014. The total number of responses received
by the end of January 2015 was 35, which is equivalent to a 70 percent response rate.
More than 80 percent of the respondents indicated that they use RAP more frequently
than RAS or a combination of RAP and RAS, as shown in Figure 3.1 and Figure 3.2. The current
practice for most respondents was to allow RAP as an option in surface mixtures; only a few did
not allow it, as illustrated in Figure 3.3 and Figure 3.4.

47
Figure 3.1. Predominant Use of RAP and/or RAS in Surface Mixtures.

Figure 3.2. Predominant Type of Recycled Materials Used in Surface Mixtures.

48
Figure 3.3. Current RAP Practice in Surface Mixtures.

Figure 3.4. RAP Practice in Surface Mixtures.

49
Respondents indicated that the typical RAP content in surface asphalt mixtures (when no
RAS was used) was between 11 and 20 percent, as shown in Figure 3.5. Only two agencies
indicated RAP amounts higher than 30 percent. The main barriers identified to using higher
levels of RAP in asphalt mixtures included mainly (a) specification limits, (b) poor pavement
performance, and (c) RAP variability. The concerns regarding pavement performance included
dry and stiff mixtures, rapid aging, cracking, and unknown long-term performance.

Figure 3.5. Typical RAP Content in Surface Mixtures (no RAS).

With respect to RAS, about half of the respondents allowed its use as an option and the
other half did not allow it, as shown in Figure 3.6 and Figure 3.7. The agencies that did not allow
RAS in surface mixtures indicated as reasons:

 Durability or long-term performance concerns.


 No prior experience.
 Shortage of RAS materials.
 RAS stiffness.

Typical RAS content for the agencies that allowed its use was 3 to 5 percent, as
illustrated in Figure 3.8. The maximum percentage allowed by most of the agencies was
5 percent; only three agencies indicated allowing more than 5 percent RAS in surface mixtures.

50
Figure 3.6. Current RAS Practice in Surface Mixtures.

Figure 3.7. RAS Practice in Surface Mixtures.

51
Figure 3.8. Typical RAS Content in Surface Mixtures.

The main barriers identified to using higher levels of RAS in surface mixtures included
(a) specification limits, (b) RAS availability, and (c) poor pavement performance. Besides most
of the agencies not allowing or limiting the amount of RAS, respondents identified the following
concerns:

 RAS acting as a Black rock and not contributing to the effective binder content.
 Mixtures becoming too stiff (i.e., workability and compactability issues).
 Premature cracking and raveling.
 Issues with RAS management such as clumping and excess moisture.
 Differences in estimated RAS binder content from chemical extraction vs. ignition
oven.
 Lack of viable RAS sources.

About 53 percent of the respondents indicated having a RBR requirement for surface
mixtures, with values ranging from 0.2 to 0.3. About 40 percent of the respondents said that they
had standard test methods or procedures for characterizing RAP and/or RAS, and almost
70 percent had mix design procedures for asphalt mixtures with RAP and/or RAS.
With respect to performance tests required for asphalt mixtures with RAP and/or RAS,
more than 60 percent of the respondents indicated those were included in their specifications, the
most common being IDT strength and/or TSR, APA, and HWTT, as shown in Figure 3.9.
Respondents noted if they had observed any distress in surface mixtures with RAP and/or
RAS. As illustrated in Figure 3.10, most common failures included various types of cracking
(i.e., fatigue, thermal, reflective) and raveling.

52
Figure 3.9. Performance Tests for Surface Mixtures with RAP and/or RAS.

Figure 3.10. Common Distresses Observed on Surface Mixtures with RAP and/or RAS.

The second topic of the survey related to the use of RAs. First, the respondents were
asked about their current practice with regard to the use of RAs in surface mixtures with RAP
and/or RAS. More than 80 percent indicated that RAs were not used or not allowed; only four
agencies noted their use was allowed as illustrated in Figure 3.11.

53
Figure 3.11. Current RA Practice in Surface Mixtures.

The respondents that indicated the use of RAs listed Hydrogreen® as the most common
commercial product; other products mentioned included Evoflex™, Reclamite®, SonneWarmix
RJ™, and an emulsion-based recycling agent (ERA). The main barriers to utilizing RAs in
asphalt mixtures with RAP and/or RAS that were identified included (a) lack of expertise in
using RAs, (b) lack of tests and criteria to determine dosage rates and/or performance, and
(c) limited amounts of RAP and/or RAS allowed under current DOT specifications (and
therefore no need to incorporate RAs). All 35 respondents indicated they did not have a standard
test method or procedure for characterizing RAs, and the majority (i.e., 91 percent) also did not
follow a standard for mix design of mixtures with RAP and/or RAS and RAs.
With respect to laboratory tests required to characterize the properties of binders
containing RAs, 40 percent of respondents indicated not having standard tests or procedures.
From the ones that did, the most common laboratory tests included measuring S and m-value
with the BBR, G* and sin() with the DSR, and the MSCR, as illustrated in Figure 3.12.
Forty percent of the agencies also responded that no performance tests were required for
surface mixtures with RAP and/or RAS containing RAs; the ones that did indicated primarily
IDT/TSR, APA, and HWTT. This agreed with the response given previously regarding the
performance tests used on mixtures with RAP and/or RAS and no RAs (Figure 3.9). In other
words, the presence of or absence of RAs in the mixture did not seem to determine the type of
mixture performance tests used.

54
Figure 3.12. Laboratory Tests to Characterize Properties of Binders with RAs.

Finally, when asked about upcoming field projects that used RAs in surface mixtures
with RAP and or/RAS and their willingness to participate in the study by sharing information
about mix design, construction, materials, or performance monitoring, the majority of agencies
responded negatively, as indicated in Figure 3.13. Only two agencies answered positively and
indicated they were available to participate.

Figure 3.13. Upcoming Field Projects Using RAs and Availability.

3.2 Contractor Web-Based Survey

The questionnaire for contractors was similar to the one sent to state DOTs and covered
the same two main topics: (a) use of high RBRs, and (b) use of RAs. The survey was sent to 15
contractors from various states on October 31, 2014, with a response deadline of November 14,
2014. A reminder was sent at the beginning of December to those contractors who had not

55
submitted a response with a revised deadline of December 19, 2014. The total number of
responses received by the end of January 2015 was 11 (two of them from the same company, but
in a different state), which is equivalent to a 67 percent response rate.
Similar to the state DOTs, most of the contractors indicated a predominant use of RAP in
surface mixtures and RAP being allowed as an option on projects, as shown in Figure 3.14 and
Figure 3.15. The typical RAP content used in surface mixtures is between 20 and 30 percent, as
illustrated in Figure 3.16.

Figure 3.14. Predominant Use of RAP and/or RAS in Surface Mixtures.

Figure 3.15. Current RAP Practice in Surface Mixtures.

56
Figure 3.16. Typical RAP Content in Surface Mixtures (no RAS).

Some of the main barriers indicated by the contractors with regard to the use of higher
levels of RAP in surface mixtures included (a) specification limits, (b) mix design criteria,
(c) RAP availability, and (d) mix production difficulties.
With respect to RAS, contractors’ current practice was similar to what was indicated by
state DOTs, with about half of the respondents indicating that the inclusion of RAS is allowed as
an option and the rest stating it is not allowed or it is unknown, as shown in Figure 3.17. Among
the ones that indicated RAS use was allowed, the typical content was predominantly 5 percent,
as illustrated in Figure 3.18. The barriers preventing use of higher levels of RAS in asphalt
mixtures identified by the respondents included, in order of importance, (a) specification limits,
(b) RAS availability, (c) mix production difficulties, and (d) poor pavement performance.

Figure 3.17. Current RAS Practice in Surface Mixtures.

57
Figure 3.18. Typical RAS Content in Surface Mixtures.

When the contractors were asked if they were required to meet a specific RBR for surface
mixtures with RAP and/or RAS, only 36 percent answered positively and indicated a required
RBR ratio between 0.2 and 0.3. More than half of the contractors (i.e., 55 percent) said they
followed standard test methods to characterize RAP and/or RAS, while around 64 percent of the
respondents indicated following standard test methods or procedures for mix design of asphalt
mixtures with RAP and/or RAS.
More than 70 percent of the respondents indicated using performance tests to characterize
asphalt mixtures with RAP and/or RAS. Like indicated by the state DOTs, the most common
performance tests included IDT/TSR, APA, and HWTT, as illustrated in Figure 3.19.

Figure 3.19. Performance Tests for Surface Mixtures with RAP and/or RAS.

58
Contractors indicated that the most common premature distresses or failures observed in
surface mixtures with RAP and/or RAS included fatigue cracking, reflective cracking, and
raveling, as shown in Figure 3.20.

Figure 3.20. Common Distresses Observed on Surface Mixtures with RAP and/or RAS.

Regarding the use of RAs, contractors were asked about their current practice with
respect to including RAs in asphalt mixtures with RAP and/or RAS. The majority of the
respondents indicated that RAs were not allowed or not used as shown in Figure 3.21. From the
respondents that used RAs, the commercial product employed with more frequency was
Hydrogreen® followed by Evoflex™, BITUTECH RAP, and Cargill (Anova).

Figure 3.21. Current RA Practice in Surface Mixtures.

59
Contractors identified the following two main barriers preventing the utilization of RAs
in surface mixtures with RAP and/or RAS: (a) lack of expertise in using RAs, and (b) lack of
tests and criteria to determine dosage rate and/or performance.
None of the contractors indicated they followed a standard test method or procedure for
characterizing RAs, and only one of them indicated he was required to follow a standard test
method or procedure for mix design of mixtures with RAP and/or RAS containing RAs.
With regard to laboratory tests to characterize the properties of binders with RAs,
60 percent of the respondents listed at least one laboratory test method, which included S and
m-value with the BBR, G* and sin() with the DSR, and MSCR, as illustrated in Figure 3.22.

Figure 3.22. Laboratory Tests to Characterize Properties of Binders with RAs.

The performance tests for asphalt mixtures with RAP and/or RAS containing RAs
identified by the respondents included IDT/TSR, HWTT, APA, and, to a lesser extent, overlay
cracking test and dynamic modulus. This response was similar to the one provided for mixture
characterization of RAP and/or RAS mixtures without RAs (Figure 3.19). Therefore, as in the
case of the state DOTs, the inclusion of RAs in the mixture did not seem to impact the type of
performance tests performed on the mixtures.
Last, contractors were asked about upcoming field projects using RAs in surface mixtures
with RAP and/or RAS and their willingness to participate in the study by sharing information
about mix design, construction, materials, and/or performance monitoring. Only one contractor
responded positively, indicating both knowledge of upcoming field projects and availability to
participate in the study, as shown in Figure 3.23.

60
Figure 3.23. Upcoming Field Projects Using RAs and Availability.

3.3 RA Supplier Web-Based Survey

The questionnaire found in Appendix A was sent to 10 suppliers of the most common
RAs available on the market. In contrast with the state DOTs and contractors, the survey was
much shorter and focused on the types, test methods, blending protocols prescribed for field and
laboratory operations, and knowledge about upcoming field products employing RAs. The
survey was sent on October 31, 2014, with a response deadline of November 14, 2014. A
reminder was sent at the beginning of December to those RA suppliers who had not submitted a
response with a revised deadline of December 19, 2014. A total number of six responses were
received by the end of January 2015, which is equivalent to a 60 percent response rate.
The types of RAs produced by the suppliers are indicated in Figure 3.24. As shown, the
most common were tall oils and bio oils. RA suppliers were also asked to identify if they had
standard test methods or procedures for characterizing RAs, and more than 80 percent of the
respondents answered positively. The standard test methods noted by the respondents varied
significantly and included penetration test, kinematic viscosity, PG grading, AASHTO M 320
and AASHTO M 323, HWTT, and Marshall Stability.
With regard to blending protocols, the one used for plant or field operations as well as
laboratory mixture production included primarily blending the RA with the virgin asphalt (only
one RA supplier indicated blending the RA with RAP/RAS), as shown in Figure 3.25.
Last, all six RA suppliers indicated knowledge of upcoming field projects using RAs in
surface mixtures with RAP and/or RAS and their willingness to participate in the study by
sharing information on mix design, construction, materials, and performance monitoring.

61
Figure 3.24. Types of RAs Produced.

Figure 3.25. Prescribed Blending Protocol for Plant and Laboratory Operations.

Other comments provided by the RA suppliers included the following recommendations:

 Quantify differences between extracted binder rheology and mixture rheology to


assess blending of the RAP/RAS binder with the virgin binder in the mixing process.
 Differentiate between rejuvenators and softeners by prescribing blended PG grade
requirements.
 Provide method for RA dosage estimation based on the RAP and/or RAS
characterization by following the blended PG grade requirement.
 Balance high RAP and/or RAS applications with traffic loading and temperature
sweeps in colder climates.

62
CHAPTER 4.0 PHASE II WORK PLAN DEVELOPMENT

The literature review (Task 1, Chapter 2) and survey results (Task 1, Chapter 3) identified
the following key points and remaining issues that will be addressed to some extent in this study.
This chapter builds on these points and issues and describes the Phase II work plan for the
expanded laboratory experiments that include materials associated with field projects planned for
use in Phase III.

 In addition to an overall RBR, the contribution from RAP and RAS should be
identified separately as RAP binder ratio (RAPBR) and RAS binder ratio (RASBR)
as follows (NCAT 2014b):

100

100

100
where PbRAP is the binder content of the RAP, PRAP is the percentage of RAP by
weight of mix, PbRAS is the binder content of the RAS, PRAS is the percentage of RAS
by weight of mix, and Pbtotal is the binder content of the combined mixture.
 State DOTs predominantly use RAP at high RBRs, and RAP is more readily available
as compared to RAS. Guidelines are needed on the blending of virgin and recycled
binders and characterization of these blends.
 At high RBRs, state DOTs are most concerned with long-term performance.
 The use of RAs that restore or rejuvenate the binder and associated mixture in terms
of both stiffness and phase angle as indicated in Black space, as shown in Figure 2.1,
are preferred over RAs that only soften the blended binder.RA
 State DOTs are predominantly exploring the use of tall oil (T) type RAs.
 The mortar procedure defined by the latest draft of AASHTO T XXX-12, Estimating
Effect of RAP and RAS on Blended Binder Performance Grade without Binder
Extraction (www.arc.unr.edu/Outreach.html), provides a more realistic method for
estimating the effectiveness of RAs because complete blending is not assumed.
 The effectiveness of RAs in restoring blended binder rheology and improving mixture
performance can be captured through binder and mortar rheology and mixture
stiffness and cracking performance after long-term aging. These can then be
subsequently tied to field project locations in terms of environment and construction
date that play a role in the blending of the binder components through diffusion.
 The evolution with aging of the effectiveness of RAs in improving mixture cracking
performance is unknown, but there should be a dependence on environment and
construction date in terms of the diffusion process and subsequent blending of the
binder components. The effect of RAs on binder oxidation kinetics is also unknown
and important to understanding longer-term effects.
 Differences in laboratory specimen fabrication and field production and construction
of mixtures with RAs should be considered before making recommendations for mix
design and quality assurance testing.

63
This chapter provides the Phase II work plan for the laboratory experiments expanded
from selected field projects (planned for use in Phase III) and describes analysis methods and
preliminary results to:

 Assess the effectiveness of RAs in restoring blended binder rheology by binder


testing and more representative mortar evaluation with incomplete blending.
 Assess the effectiveness of RAs in improving mixture cracking performance at
optimum dosage rates using selected laboratory tests.
 Evaluate the evolution of RA effectiveness in improving mixture cracking
performance.
 Recommend evaluation tools for assessing the effectiveness of RAs and their
evolution for specific material combinations at specific locations.

Based on the literature review and survey results, this Phase II work plan focuses on the
following:

 0.4 RBR with all RAP and balanced RAP/RAS combinations.


 Traditional aromatic and greener alternative tall oil RAs.
 PG 64-22, PG 64-28, and polymer-modified PG 64-28P virgin binder grades.

In addition, an extreme 0.5 RBR (with a balanced RAP/RAS combination) and an all-RAS
combination will also be explored for limited material combinations. All proposed experiments
will be revised as results are obtained, with some material combinations possibly removed and
some added based on the continuous analysis of the binder blend, mortar, and mixture results.

4.1 Field Projects and Expanded Materials

In selecting field projects to be associated with the Phase II expanded laboratory


experiments and the Phase III field validation, consideration was given to obtaining a range in
each of the following factors to make the conclusions of this study as comprehensive as possible:

 RA by category as defined in Table 2.1 for comparison of traditional aromatic and


paraffinic agents with alternative green tall oil technologies.
 RA handling procedure including temperature of mixing, time of mixing, and location
where it is introduced in the production process.
 Environment of wet and dry, freeze and no-freeze (Figure 4.1).
 Traffic volume.

64
Figure 4.1. SHRP-LTPP Environmental Zones.

Since the laboratory testing program will be tied to actual field projects to facilitate Phase III,
selection of a project in a specific environment will simultaneously result in selection of the
aggregate and binder types (and additives, including RAs) based on the materials selected by the
respective DOT. Other eligibility requirements for field projects to be further evaluated in Phase
II include:

 Include a high RBR between 0.3 and 0.5.


 Located on a highway, arterial, or collector facility in North America.
 Include a virgin section (with no recycled materials).
 Include a control section (with recycled materials but without RA).
 Include multiple RAs (if possible).
 With a minimum number of WMA and anti-stripping additives.
 Not yet constructed with materials available for testing according to the work plan
described in this section.

New construction is beneficial to this study, as it provides the only opportunity for fabrication of
reheated plant-mixed, laboratory-compacted (RPMLC) specimens that capture field blending of
virgin binders, recycled materials, and RAs and a starting point for tracking performance of cores
to validate laboratory aging protocols critical to evaluating mixture cracking performance and its
evolution.
Table 4.1 provides the field projects that are most likely to be used in this study and
where most of the eligibility requirements are met. These projects are proposed based on the
cooperation of state DOTs, contractors, and asphalt paving associations. A second field project in
California in a dry, freeze environment is also under discussion. For many of these projects, the

65
specific RAs are not yet selected, but the best information to date is provided. Table 4.1 also
shows the associated binder, mortar, and mixture testing to be completed on all selected field
projects (including associated virgin and control sections) to meet the objectives of this study.
Further details on the selected laboratory tests are included in the following section.
Aging states are also shown in Table 4.1 to provide an overview of the proposed testing plan.
Critical representative aging states are proposed for each laboratory test across the pavement
temperature spectrum, with short-term (S) aged binders (RTFO) or mixtures (short-term oven
aging [STOA]) evaluated for stiffness at high temperatures and long-term (L) aged binders
(RTFO and PAV) or mixtures (STOA and long-term oven aging [LTOA]) for cracking resistance
or stiffness at intermediate and low temperatures. Short-term and long-term aging protocols for
binders and mixtures will be defined first, with mixture protocols based on a preliminary
specimen fabrication experiment and ongoing NCHRP Project 9-52 as described subsequently.
Procedures used to date for specimen fabrication are described in Appendix B, and ongoing
results from NCHRP Project 9-54 will also be considered with respect to the aging protocols.

66
Table 4.1. Proposed Field Projects and Associated Laboratory Testing Plan.

Aging Test$ Performance


RA
Material State(s)% Property
Field Project RB RAPBR @ Field
(Env) Mix Type R |RASBR Dosage Virgin U,S PGH Rheology
0.1 | 0.2 T1 Binder L PGL & Tc Rheology
Texas RAP U,S PGH Rheology
w/RA 0.3 0.1 0.2 T2
SH 31 Binder L PGL & Tc Rheology
0.1 0.2 Emulsn
(Wet, No- RAS U,S PGH Rheology
Control 0.3 0.1 | 0.2 —
Freeze) Binder L PGL & Tc Rheology
Virgin — — —
Indiana/ w/RA TBD TBD TBD BINDER DSR Master
U,S Rheology
Michigan Control TBD TBD — RA Curve
(Wet, Freeze) Virgin — — — U,S  @ 60°C Viscosity
w/RA TBD TBD TBD U,S PGH Rheology
Nevada
Control TBD TBD — Binder Lx2 PGL Rheology
(Dry, Freeze)
Virgin — — — Blend DSR Master Rheology &
w/RA 0.5 TBD TBD S,Lx2
Curve, FTIR Aging
67

Florida w/RA 0.5 TBD TBD U,S PGH Rheology


(Wet, No- w/RA 0.5 TBD TBD MORTAR PGI, PGL, &
Freeze) Control 0.5 TBD — L Rheology
Tc
Virgin — — — S,L,R,Cx2 MR Stiffness
0.4 | 0.0 Fatigue
0.4 0.2 0.1 A2 S,L EBM
Resistance
0.3 0.2 Fatigue
w/RA MIXTURE AR or L#, Cx2 E*+S-VECD
0.4 | 0.0^ Resistance
California
0.4 0.4 0.1 T1 Thermal
(Dry, No-
Freeze)
0.2 0.2 AR or L#, Cx2 E*+UTSST Cracking
0.4 | 0.0 Resistance
Control 0.4 0.3 0.1 —
0.2 | 0.2
Virgin — — —
^
Mixtures with a high RA dosage based on PGH grade will also be considered.
%
S = short-term aged in lab, L = long-term aged in lab (x2 = two states), R = reheated plant mix, AR = long-term aged plant mix, Cx2 = cores @ construction and after
approximately 1 year of field aging.
$
PGH = high-temperature PG grade, PGI = intermediate-temperature PG grade, PGL = low-temperature PG grade.
#
For limited number of mixtures.
The Texas field project with five field sections, as shown in Table 4.1 (with additional
details provided in Table 4.2), was the first constructed—in June 2014—and included a 0.3 RBR
with a 0.1 RAPBR and a 0.2 RASBR.

Table 4.2. Texas Field Project.


Mixture 1 Mixture 2 Mixture 3 Mixture 4 Mixture 5
(Virgin) (Control) (w/RA1) (w/RA2) (w/RA3)
Binder Grade 70-22 64-22 64-22 64-22 64-22
RAP / MWAS — 10/5 10/5 10/5 10/5
(wt% by mix) |
RAPBR | RASBR 0.1 | 0.2 0.1 | 0.2 0.1 | 0.2 0.1 | 0.2
Other Additive — WMA — WMA —
Chemical Chemical
Additive 1 Additive 2
RA — — Tall Oil 1 Tall Oil 2 Emulsion
(T1) (T2) Product

These materials from the Texas field project were expanded with the following supplemental
materials to further develop the Phase II laboratory experiments:

 An unmodified PG 64-28 (S-controlled) binder from the Northeast.


 A polymer-modified PG 64-28P binder from the West.
 A second RAP source from New Hampshire.
 Two TOAS sources.
 Three rejuvenating RAs:
o An aromatic extract RA (A1).
o A paraffinic oil RA (P1).
o A re-refined lube oil RA (R1).

Thus, the materials proposed for use in Phase II were selected based on input from panel
comments and careful selection of efficient material and testing combinations to focus on
commonly used and available TOAS sources, four rejuvenating RAs (including one tall oil RA
[T1]) that include traditional aromatic extract and paraffinic oil RAs to compare with alternative
re-refined lube oil and green tall oil RAs, and 0.3 and 0.4 RBR values with all RAP or RAP/RAS
combinations. If possible in the laboratory, a very high 0.5 RBR or a RAS-only combination will
also be examined with a few selected material combinations as shown in the proposed
experiments.
Three other RAs (a low carcinogenic aromatic extract RA [A2], a modified triglyceride
product [F], and a third tall oil RA [T3]) are also available for evaluation and use at the request
of specific DOTs (as currently proposed for CA) and thus included in the overall testing plan
associated with field projects (Table 4.1). Results from the associated Texas Department of
Transportation (TxDOT) Project 0-6738 for a second tall oil (T2) used with a WMA additive, the
uncommon emulsion product, and two softeners (softer binder grades AC 0.6 and AC 5) will
also be reviewed by the research team but not utilized in the expanded Phase II laboratory
experiments.

68
Preliminary results for some of the material combinations in the proposed Phase II
expanded laboratory experiments are presented after the field activities and the selected
laboratory tests are introduced.

4.2 Field Activities

Field activities will include gathering original component materials and plant mix for
laboratory-mixed, laboratory-compacted (LMLC) and RPMLC specimen fabrication and
procuring cores at construction and after approximately 1 year to verify specimen fabrication and
aging protocols and validate relationships between binder and mixture properties that are
discussed in the next section. These materials will be available through cooperation with the state
DOTs, contractors, and state asphalt paving associations. Cores at the time of construction and
after 1 year of aging during the study will be requested from state DOTs. A construction report
will be produced for each field project (as shown in Appendix C for the Texas field project), and
field performance assessment by visual survey in cooperation with the associated DOTs will also
be completed. In addition, cumulative degree days (CDDs, 0°C base) based on daily average
temperatures since construction will be gathered for each selected field project location to
quantify field aging and normalizing data with respect to both environment and date of
construction as proposed and utilized in NCHRP Project 9-52 and discussed in Section 4.6.2.
Additional details on these proposed field activities are discussed in the following subsections.

4.2.1 Material Sampling during Production and Construction

The sampling plan proposed for use in this study will follow the guidelines established in
Research Results Digest 370, Guidelines for Project Selection and Materials Sampling,
Conditioning, and Testing in WMA Research Studies. The following general procedures will be
observed:

 Asphalt binders, WMA additives, RAS, RAP, any other additives (anti-stripping,
etc.), and low-viscosity RAs if utilized will be sampled at the plant location during
production of the field sections to be utilized to fabricate LMLC specimens.
 Plant mix will be sampled, and some of the mixtures may be compacted at the plant
site to produce PMLC specimens while others will be reheated at a central laboratory
prior to compaction to produce RPMLC specimens.
 Plant production information such as plant type and characteristics, baghouse fines
handling, production rate, mixture design inputs, plant temperatures, silo storage
times, and haul times will be recorded.
 Field laydown and compaction process information will be recorded, including
equipment, rolling pattern, mat thickness, mat temperature, temperature uniformity,
and climate conditions at the time of construction.

4.2.2 Coring Plan after Construction

One critical issue is to define the evolution of stiffness and cracking resistance of
mixtures over time. Pavement cores will be obtained from the field sections associated with each
field project during the duration of this study. It is anticipated that cores will be taken from the
field projects at two different times—immediately after construction and after 1 year. Additional

69
coring times are desirable; however, it is likely that the study budget and timeline will not allow
for additional coring and testing, with the exception of the field projects tied to the materials
used in the laboratory experiment including the Texas field project constructed in June 2014. For
this field project, three coring periods may be possible.

4.2.3 Prior- and Post-Construction Assessment

If allowed and possible, a prior-construction distress assessment will be performed,


which will include a visual survey and cracking map of existing distresses. If needed, cores will
also be procured prior to and after placement of an overlay to assess the depth of top-down
cracking, particularly in those states where this distress is prevalent. Additionally, other
pavement design information (e.g., pavement structure layer thickness, modulus, and traffic data)
will be collected if it is available. After construction, it is anticipated that two field surveys for
some field projects will be performed after 1 and 2 years of service. A visual condition survey
will be performed according to the Long-Term Pavement Performance (LTPP) program
methodology, and crack maps will be prepared.

4.3 Laboratory Testing

The results from Task 1 (Chapter 2 and Chapter 3) were also utilized to select the laboratory tests
listed in

70
Table 4.3 to evaluate the performance of binders and mixtures with high RBRs for the
materials associated with field projects for Phase III (Table 4.1) and the expanded material
combinations for the Phase II laboratory experiments presented in the following sections.
Different tests will be utilized for binders, mortars, and mixtures to:

 Assess the effectiveness of RAs in restoring blended binder rheology by binder


testing and more representative mortar evaluation with incomplete blending.
 Assess the effectiveness of RAs in improving mixture cracking performance at
optimum dosage rates using selected laboratory tests.
 Evaluate the evolution of RA effectiveness in improving mixture cracking
performance.
 Recommend evaluation tools for assessing the effectiveness of RAs and their
evolution for specific material combinations at specific field locations.

For all tests, a minimum of two replicate specimens will be utilized with at least three replicates
for MR testing. An additional third replicate for other mixture tests is also desired due to higher
variability associated with the use of recycled materials, and the use of three or four replicates
will be considered and balanced against budget and time constraints. Air voids for all mixture
specimens will be determined by AASHTO T 312, and other mixture volumetrics including
effective binder content (Pbe) and absorbed binder content (Pba) will be calculated and utilized in
analyzing laboratory test results.

71
Table 4.3. Selected Laboratory Tests.

Performance Property Binder Test Mortar Test Mixture Test

Compatibility Exudation Droplet Test


NA NA
with and without Aging DSR Master Curve Parameters

G*,  @ Thigh after S Aging


G*,  @ Tint after L Aging
S, m-value @ Tlow after L MR @ Thigh, Tint, &
Aging Tlow after S, L Aging
PGH after S
DSR Master Curve Parameters G-R Mixture
Rheology Evolution Aging
including Parameter from E*
with Aging PGI, PGL after L
G-R = G* (cos δ)2/sin δ @ Tint Master Curve after
Aging
with Aging @ multiple t, T S, L Aging
CA by FTIR with Aging @
multiple t, T
Oxidation Kinetics
Specimen Fabrication
NA NA MR @ Tint
Protocol
PGH after S
Rutting Resistance G*/sin @ Thigh after S Aging HWTT^
Aging
S-VECD after L
Fatigue Resistance
G*sin δ after L Aging PGI after L Aging Aging
after Aging
EBM after L Aging
Predicted S @ Tlow after L
Low-Temperature Aging
PGL after L UTSST after L
Cracking Resistance Predicted m-value @ Tlow after
Aging Aging
after Aging L Aging
 Tc
^
For limited number of mixtures with high RA dosage rate or incompatible binder blends.

4.3.1 Binder Testing

For binder testing, the RTFO will be utilized for short-term aging. A single long-term
aging protocol of 20 hr in the PAV will be used for the PG grading tests, and a second long-term
aging protocol of 60 or 80 hr in the PAV will be used for DSR master curves and determination
of the G-R parameter for Black space analysis. Based on the results shown in Figure 2.1, 60 hr
may be sufficient, but different binder blends may need to be aged for 80 hr in the PAV to move
beyond the G-R damage zone. Recent work by Ryan et al. (2014) suggested that 40 hr might be
sufficient in extreme climates such as Minnesota.
Previous research by Glaser and Porot (2014) utilized the exudation droplet test originally
developed by Shell Bitumen to test binder blends with an RA T type and a softening agent
petroleum product at dosage rates from 1 to 20 percent and found that all blends were
compatible. This test utilizes a marble slab (with 10 mm diameter indentations) to simulate an
aggregate, and blended binder droplets are placed in the indentations and kept in an oven at 60°C
for 4 days under nitrogen. After aging, the width of any light-colored ring is measured under

72
ultraviolet light and a microscope. If the ring width is greater than 1.5 mm, the binder blend is
considered incompatible.
Traditional DSR and BBR binder testing will be used to determine rheological
parameters of stiffness (G* and S) and phase angle or stress relaxation ability ( and m-value)
and the associated PG grade at high, intermediate, and low temperatures, or PGH grade, PGI
grade, and PGL grade (AASHTO M 320 with AASHTO T 315 and AASHTO T 313). To reduce
overall testing time and eliminate BBR binder testing, S and m-values will be predicted after
validation of the use of DSR frequency sweeps with binder blends including RAs (Farrar et al.
2014).
The influence of RAs on the rheology restoration and aging characteristics of binders will
be assessed using chemical (FTIR) and physical (rheology) parameters. The binders with and
without RAs will be aged in a forced draft oven at different temperatures and for multiple
durations and will be evaluated through FTIR spectroscopy measurements to determine the
carbonyl area, which is an indicator of oxidation of the binder through quantification of the
carbonyl functional group. The binders will also be tested in a DSR to determine the master
curves of G* and  by conducting isothermal frequency sweeps at different temperatures.
Rheological indices such as low shear viscosity (LSV), G-R parameter, crossover frequency and
modulus (c and G*c), and rheological index R-value will be calculated as a function of aging
for the various binders to study the effect of RAs. Furthermore, the tests will also be utilized to
develop and assess the effects of RAs on binder oxidative aging kinetics and resulting hardening
susceptibility, which are key inputs for the modeling of binder aging during the in-service life of
a pavement.
The intermediate-temperature DSR parameter, called the G-R parameter (G* x (cos
δ)2/(sin δ)), as originally defined by Glover et al. (2005) and reformulated for greater practical
use by Rowe (2011), in a discussion of Anderson et al. (2011), will be utilized to assess the
effectiveness of RAs in restoring aged recycled binders. This G-R parameter is ideal for this
purpose because it measures both fundamental rheological properties using a DSR at an
intermediate temperature. The results can be compared on Black space diagrams, as shown in
Figure 2.1, to cracking resistance as defined by inadequate ductility of 5 cm to 3 cm that
correlates to G-R parameter values between 180 and 450 kPa, respectively, and relates to surface
raveling and cracking (Kandhal 1977). At low temperatures, the difference between the S and m-
controlled PGL grades ( Tc) that Anderson et al. (2011) found correlated with the G-R
parameter will also be determined.

4.3.2 Mortar Testing

For mortars, the laboratory testing procedure will follow the latest draft of test method
AASHTO T XXX-12, Estimating Effect of RAP and RAS on Blended Binder Performance
Grade without Binder Extraction (www.arc.unr.edu/Outreach.html). In this procedure, mortar
and binder samples are tested in the DSR and BBR to quantify the effect of recycled binder on
the continuous grade profile of virgin binder, allowing for an estimation of blended binder
properties at critical pavement temperatures with commonly available equipment and without the
need for the time-consuming and hazardous binder extraction and recovery process that may
impact binder properties. The following three samples are each tested at low, intermediate, and
high critical PG temperatures after appropriate or critical aging in the RTFO or PAV: (a) virgin
binder; (b) void-less Mortar A with the same virgin binder and a single size RAP (or RAS) from

73
a single source; and (c) void-less Mortar B with the same virgin binder, the same total binder
content as Mortar A, and recovered aggregate from the same RAP (or RAS) material (using the
ignition oven). With known virgin binder properties, this procedure determines the change in
continuous PG grade of the composite binder with the addition of recycled materials, which then
allows for an estimation of mixture performance for a specific RBR. The effect of RAs on
selected binder properties will also be evaluated by adding this component to the virgin binder
and both mortars.

4.3.3 Mixture Testing

Prior to mixture performance testing, specimen fabrication and aging protocols will be
established for use in the laboratory to simulate blending of the RA, recycled binder, and virgin
binder that occurs in the field during early life of the pavement. The research team has recent and
ongoing experience in NCHRP Projects 9-49 and 9-52 toward developing these protocols for
laboratory conditioning and aging to evaluate performance evolution (Epps Martin et al. 2014;
Yin et al. 2013; Yin and Garcia Cucalon et al. 2014; Yin et al. 2015). Preliminary results and
analyses are presented subsequently, but the selected protocols based on completed research are:

 5 days at 85°C for long-term oven aging for all mixtures.


 2 hr at 135°C (275°F) for short-term oven aging for HMA, all mixtures with RAs, and
all mixtures prior to LTOA.
 2 hr at 116°C (240°F) for WMA without RAs.

Ongoing results from NCHRP Project 9-54 will also be considered with respect to long-term
oven aging protocols for mixtures.

4.3.3.1 Stiffness
Mixture rheology and its evolution with aging will also be evaluated through the use of
resilient modulus (MR) and dynamic modulus (E* and ) testing. The nondestructive MR
(ASTM D7369) test utilized successfully in recent NCHRP Projects 9-49 and 9-52 will be used
to measure stiffness at three temperatures (4, 25, and 40°C) to assess the spectrum of
performance across the temperature spectrum for comparison with binder and mortar results. In
addition, as described subsequently, MR at 25°C will be utilized to define specimen fabrication
protocols for short-term loose-mix oven aging prior to compaction for mixtures and long-term
compacted mixture oven aging. The MR test was selected in addition to the commonly used E*
because it provides a repeatable, cost-effective measurement of tensile stiffness that is likely
more sensitive to binder properties and is of great concern in mixtures with high RBRs as
compared to the compressive E* stiffness. Dynamic modulus characterization, however, is
required for the S-VECD and the Uniaxial Thermal Stress and Strain Test (UTSST), a modified
TSRST, mixture cracking tests and provides both rheological parameters (stiffness E* and phase
angle ) needed to completely assess the effectiveness of RAs in these mixtures with high RBRs.
In addition, E* master curve parameters will be utilized in the aging experiments described
subsequently.

4.3.3.2 Restoration of Rheology


The restoration of mixture rheology will also be evaluated using mixture Black space and
the Glover-Rowe-based parameter for low-temperature performance indication of mixtures

74
currently under development (Mensching et al. 2015). In this work, a separation of different
mixtures in Black space has been observed near the inflection point. Figure 4.2 shows an
example of this for a set of mixtures tested as part of pooled fund study TPF 5(230), Evaluation
of High RAP Plant Produced Mixtures, in the Northeast; the mixtures with the softer base binder
(PG 52-34 designated by “e”) show separation from the mixtures with the PG 58-28 (designated
by “a”) base binder.

Figure 4.2. Mixture Black Space Curves for Vermont Pooled Fund Phase I Materials.

Calculation of thermal stress and relaxation modulus during representative cooling events
has resulted in a range of interest that corresponds to 0.1 to 10 sec on the reduced time axis for
plant-mixed, laboratory-compacted and field cores. Figure 4.3 (Mensching et al. 2015) shows the
mixture Glover-Rowe parameter calculated at a frequency of 0.005 rad/s and a temperature of
15°C plotted against the critical cracking temperature determined from two laboratory low-
temperature cracking tests for a group of the TPF 5(230) mixtures. These results indicate that
performance groupings based on laboratory tests are possible with the Glover-Rowe parameter
for mixtures.

75
Figure 4.3. Modified Mixture Glover-Rowe Values (0.005 rad/s and 15oC) against
Laboratory-Measured Critical Cracking Temperature for Pooled Fund Mixtures.

More recently, work has been done to identify a commonly measured temperature-
frequency combination for use as the Black space evaluation point for ease of use. This has been
identified as 5 Hz at 20°C, and initial results show there is a separation in Black space at this test
combination that correlates with the poorly performing mixtures. This methodology will be
further explored with the laboratory and field mixtures in this study to evaluate the effect of RAs
on the low-temperature mixture performance.

4.3.3.3 Fatigue Resistance


For mixtures with high RBRs, cracking resistance is critical, and two approaches were
selected to analyze mixture fatigue response in repeated uniaxial DTTs. The energy-based
mechanistic (EBM) approach developed at TTI will be used to evaluate the effect of aging on
cracking performance of LMLC specimens with and without RAs based on controlled-strain
repeated direct tension testing and analysis to determine the evolution of the following index,
which represents cracking damage density:
Sc
  100%
S0

where  is the cracking damage density, Sc is the lost area due to cracking on a cross-section of
the material, and S0 is the original cross-sectional area before cracking. This index is then used
to formulate a modified Paris’ law to simulate the growth of a multitude of cracks in mixtures
(Luo et al. 2013a):
d
 A  J R 
n

dN
where A and n  are modified Paris’ law coefficients associated with the evolution of the
cracking damage density, and J R is the pseudo J integral. These fracture parameters A and n 
are the output of this approach, and Luo et al. (2013b) proved that they are interrelated and can
successfully discriminate the fatigue resistance of different types of mixtures. They can also be

76
predicted using performance-related properties by the following equations (Luo et al. 2013b) that
are also being utilized by NCHRP Project 1-52:
log E1
 log A  56.824  4.920  3.166a  1.131PGhigh  0.810PGlow  100.435
m
1
n  36.135  3.007  1.323a  0.448PGhigh  0.301PGlow  24.521
m
where E1 and m are relaxation modulus model coefficients that change with aging; a is the
mean air void content from mix design; PGhigh and PGlow are the PGH and PGL grades,
respectively; and  is an aggregate gradation shape parameter from a power model. Figure 4.4
compares the predicted values to the measured values for 20 different mixtures with a strong
correlation.

77
40

35

30
Measured (‐logA')
25

20

15

10

0
0 5 10 15 20 25 30 35 40
Predicted (‐logA')

(a) A′

16

14

12
Measured n'

10

0
0 2 4 6 8 10 12 14 16
Predicted n'

(b) n′
Figure 4.4. Comparison of Measured and Predicted Modified Paris’ Law Coefficients (Luo
et al. 2013b).

78
The second approach (S-VECD) was developed originally by Kim and Little (1990) and
extended for both controlled stress and strain cyclic loading conditions and to include the effects
of healing (Lee and Kim 1998a, 1998b), temperature, and monotonic loading (Daniel and Kim
2002). The viscoelastic continuum damage model produces a damage characteristic curve for a
mixture that can be used to determine mixture response to any uniaxial loading history. Testing
time was reduced when several researchers (Chehab et al. 2002; Underwood et al. 2009)
confirmed the validity of the time-temperature superposition principle that allows
characterization at a single temperature. Recently, the S-VECD approach was developed to
reduce analysis time and establish compatibility with the AMPT (Underwood et al. 2010). This
approach that includes the E* measurement is now AASHTO TP 107. This approach can also
now be used on specimens cut from cores based on results from Li and Gibson (2013) that
demonstrated that E* and S-VECD fatigue behavior measured on 38 mm diameter and 110 mm
tall specimens are comparable to those measured on the traditional 100 x 150 mm geometry. This
second approach will be used to evaluate the evolution of RA effectiveness in improving mixture
cracking performance.
In addition to these two approaches selected based on completed research, ongoing
results from NCHRP Project 9-57 will also be considered with respect to mixture cracking
performance tests.

4.3.3.4 Thermal Cracking Resistance


The evolution of RA effectiveness in improving mixture cracking performance will also
be evaluated with respect to low-temperature cracking since these mixtures with high RBRs will
be stiff and exhibit significant brittle behavior with an inability to relax stress. The TSRST
measures thermal stress buildup under a constant cooling rate in a restrained mixture specimen
until fracture. Recently, modifications to the traditional TSRST setup (in the UTSST) were
developed to improve and facilitate sample preparation, repeatability, and reproducibility of test
data, measure thermal strain in an unrestrained specimen, and analyze test measurements (Alavi
et al. 2013). From the concurrent measurements of thermal stress and strain, the relaxation
modulus of the mixture is calculated using linear viscoelastic theory. Subsequently, the thermo-
viscoelastic properties at the following stages of the material behavior are identified based on the
change in the mixture relaxation modulus with temperature: viscous softening, viscous-glassy
transition, glassy hardening, crack initiation, and fracture. Moreover, the coefficient of thermal
contraction is determined by analyzing the thermal strain measurements as a function of
temperature. Recently, the UTSST test was used successfully to investigate the influence of
cooling rates (Alavi and Hajj 2014), long-term oxidative aging (Morian et al. 2014), WMA
technology (Hajj et al. 2013), high RBRs (Alavi and Hajj, 2014), and a RA (Hajj et al. 2013) on
thermo-viscoelastic and fracture properties of mixtures. A draft AASHTO standard method has
been developed, and specimens cut from cores can now be evaluated.

4.3.3.5 Rutting Resistance and Moisture Susceptibility


While mixture cracking resistance is the biggest concern with high RBRs and will thus be
the focus of this study, addition of RAs to address this issue must be balanced by ensuring
adequate mixture resistance to rutting and moisture susceptibility during early life. For a limited
number of mixtures with a high RA dosage rate defined based on restoring PGH grade (or
possibly for incompatible binder blends), the HWTT (AASHTO T 324) with traditional
parameters of rut depth at a specific number of load cycles and stripping inflection point,
respectively, will be utilized to check mixture resistance to rutting and moisture susceptibility,

79
respectively. In addition, three new parameters defined by Yin and Arambula et al. (2014) that
separate mixture resistance to these two forms of distress will be determined from the same set of
laboratory results.

4.4 Mortar Experiment

The mortar experiment is described prior to the binder and mixture experiments because
it provides a more realistic blending of the binder components and a complete characterization
across the entire temperature spectrum.

4.4.1 Preliminary Mortar Results

A preliminary mortar experiment (Table 4.4) to quantify the effect of recycled materials
and a tall oil RA (T1) on the blended binder continuous PG grade using the Texas field project
materials and the supplemental PG 64-28 unmodified binder was completed. The mortar
procedure followed the latest draft of standard test method AASHTO T XXX-12, Estimating
Effect of RAP and RAS on Blended Binder Performance Grade without Binder Extraction
(www.arc.unr.edu/Outreach.html), which utilizes the DSR and BBR to test virgin binders (with
and without RAs) and mortars consisting of these binders and R100 (passing sieve #50 and
retained on sieve #100) RAP and RAS materials (Figure 4.5). The effect of RAP and RAS on the
blended binder continuous PG grade is first identified separately at the highest recycled binder
ratios that maintain sufficient workability for specimen fabrication, and then the combined effect
for the RAP and RAS blend is predicted from the linear combination of the RAP-alone and RAS-
alone blends (Bahia and Swiertz 2011). A selected blend of RAP and RAS materials was also
created and tested, again at the highest recycled binder ratio possible while still maintaining
workability, to verify the predictions.

Table 4.4. Preliminary Mortar Experiment for Texas Field Project Materials and
Supplemental Virgin Binder.
RAP RAS RAP + RAS
Virgin
With RA With RA With RA
Binder PG Without RA Without RA
(T1) (T1) (T1)
70-22 X X
X X
64-22
X X X
64-28 X X
RA

80
Figure 4.5. RAP and RAS Materials (#50 X #100) from the Texas Field Project.

Figure 4.6 summarizes the high (PGH), intermediate (PGI), and low (PGL) continuous
PG grades for the various virgin and recycled binders. To simulate the common practice of in-
line blending of the small quantity of RA with virgin binder prior to introduction into the drum,
the PG 64-22 virgin binder with 2.65 percent RA (T1) was also tested, and the results are shown
in Figure 4.6. The selected dosage for the RA (T1) was established from the field application
dosage rate of 0.13 percent by total weight of mix and 4.9 percent optimum binder content for
the Texas mixtures. While the use of RA (T1) resulted in 4.7C decrease in the continuous PGH
grade of the PG 64-22 asphalt binder, only a 0.7C softening in the continuous PGL grade was
observed. The recovered RAP and RAS binders were only tested at high temperatures, and
continuous PGH grades of 108.6C and 133.9C were determined, respectively. Limitations in
the load capacity of the DSR and BBR equipment prevented testing of the recycled binders at
intermediate and low temperatures, respectively.

81
Figure 4.6. Continuous Grades for Virgin and Recycled Binders.

The estimated influence of each of the RAP and RAS binders on the PGH, PGI, and PGL
grades for the blended binders using the mortar procedure is shown in

82
Table 4.5 and Table 4.6, respectively. For this analysis, the RAP-only mortar mix was
blended at 9.2, 10.1, and 11.4 percent RAPBR for the PG 70-22, PG 64-22, and PG 64-28 virgin
binders, respectively. The RAS-only mortar mix was blended at 20.7 percent RASBR for the PG
70-22 and PG 64-28 virgin binders and at a ratio of 25.0 percent for the PG 64-22 virgin binder.
Table 4.5 and Table 4.6 also summarize, at each of the high, intermediate, and low temperatures,
the grade change rate, which is defined as the change in the binder grade with respect to the
recycled binder replacement and is calculated as follows:

83
Table 4.5. Effect of RAP on Continuous PG Grade.
Continuous PG Grade (C)
Virgin Grade Change
Temperature RAPBR Blended
Binder RA Virgin Binder Rate
Range (%) Binder
PG (+ RA) (C/%RAP BR)
(w/RAP)
High 73.6 76.2 0.28
Intermediate 18.8 24.2 0.33
70-22 9.2 None
−24.5 −20.6
Low 0.42
(m-controlled) (m-controlled)
High 67.6 70.4 0.28
Intermediate 18.8 21.5 0.26
64-22 10.1 None
−24.5 −20.2
Low 0.42
(m-controlled) (m-controlled)
High 62.9 65.5 0.26
Intermediate 17.1 21.7 0.46
64-22 10.1 T11
−25.2 −21.4
Low 0.38
(m-controlled) (m-controlled)
High 66.0 71.6 0.49
Intermediate 20.3 23.3 0.27
64-28 11.4 None
−29.7 −27.3
Low 0.22
(S-controlled) (S-controlled)
1
Virgin binder replacement of 2.65% by total weight of asphalt binder.

Table 4.6. Effect of RAS on Continuous PG Grade.


Virgin Continuous PG Grade (C) Grade Change
Temperature RASBR
Binder RA Virgin Binder Blended Binder Rate
Range (%)
PG (+ RA) (w/RAS) (C/%RASBR)
High 73.6 77.7 0.20
Intermediate 18.8 19.9 0.10
70-22 20.7 None
−24.5 −19.0
Low 0.27
(m-controlled) (m-controlled)
High 67.6 74.4 0.27
Intermediate 18.8 22.6 0.15
64-22 25.0 None
−24.5 −14.8
Low 0.39
(m-controlled) (m-controlled)
High 62.9 70.1 0.18
Intermediate 17.1 21.3 0.17
64-22 25.0 T11
−25.2 −20.6
Low 0.29
(m-controlled) (m-controlled)
High 66.0 72.3 0.30
Intermediate 20.3 23.2 0.14
64-28 20.7 None
−29.7 −23.2
Low 0.32
(S-controlled) (m-controlled)
1
Virgin binder replacement of 2.65 percent by total weight of asphalt binder.

84
Figure 4.7, Figure 4.8, Figure 4.9, Figure 4.10, Figure 4.11, Figure 4.12, Figure 4.13, and
Figure 4.14 show a surface plot for the expected effect of the RAP and RAS combination on the
continuous PGH and PGL grades of the blended binders for the three different virgin binders
(PG 70-22, PG 64-22, and PG 64-28) and the PG 64-22 with the RA (T1) (Figure 4.11 and
Figure 4.12). The horizontal and vertical axes represent the percent binder ratio for RAP and
RAS, respectively. The shaded area represents the resulting continuous PGH and PGL grades of
the RAP, RAS, and virgin binder blends (with or without RA). The combined effect for the RAP
and RAS blend is predicted from the linear combination of the RAP-alone and RAS-alone blend
test results and the grade change rates. To verify this assumption, a combination of RAP and
RAS mortar mix was blended and tested at 15.9 percent recycled binder ratio (the highest
possible while maintaining workability) with the PG 64-22 virgin binder and 2.65 percent RA
(T1) virgin binder replacement. Table 4.7 summarizes the test results for the measured and
predicted blended binder continuous PG grades and shows that the predictions for the blended
binder continuous PG grades were within 1.4C for PGH grade and 0.8C for both PGI and PGL
grades from the measured continuous PG grades at the same recycled binder ratio.

Figure 4.7. Effect of RAP and RAS Continuous PGH Grade with PG 70-22 Virgin Binder.

85
Figure 4.8. Effect of RAP and RAS on Continuous PGL Grade with PG 70-22 Virgin
Binder.

Figure 4.9. Effect of RAP and RAS on Continuous PGH Grade with PG 64-22 Virgin
Binder.

86
Figure 4.10. Effect of RAP and RAS on Continuous PGL Grade with PG 64-22 Virgin
Binder.

Figure 4.11. Effect of RAP and RAS on Continuous PGH Grade with PG 64-22 Virgin
Binder and RA (T1).

87
Figure 4.12. Effect of RAP and RAS on Continuous PGL Grade with PG 64-22 Virgin
Binder and RA (T1).

Figure 4.13. Effect of RAP and RAS on Continuous PGH Grade with PG 64-28 Virgin
Binder.

88
Figure 4.14. Effect of RAP and RAS on Continuous PGL Grade with PG 64-28 Virgin
Binder.

Table 4.7. Verification of the Mortar Procedure for RAP and RAS Combination.
Virgin Continuous PG Grade (C) Abs.
Temperature RBR
Binder RA Virgin Binder Blended Binder Diff.2
Range (%)
PG + RA Estimated1 Measured (C)
High 62.9 66.2 67.6 1.4
Intermediate 17.1 21.5 22.3 0.8
64-22 15.91 T13
−25.2 −20.1 −20.9
Low 0.8
(m-controlled) (m-controlled) (m-controlled)
1
5.7 percent RAP binder ratio and 10.2 percent RAS binder ratio.
2
Absolute difference between measured and estimated continuous PG grades.
3
Virgin binder replacement of 2.65 percent by total weight of binder.

Using the determined grade change rates (Table 4.5 and Table 4.6), the expected
continuous PGH, PGI, and PGL grades for the various blended binders were calculated at the
RBR of 28.6 percent used in the Texas field project (with 10.2 percent RAPBR and 18.4 percent
RASBR). The results are shown in Figure 4.15 and Figure 4.16 and lead to the following
observations when rounded to the nearest °C:

 The virgin mixture in the Texas field project utilized a PG 70-22 (74-25) virgin
binder without recycled materials or RAs, and thus this PG grade is the target PG
grade.
 The control mixture in the Texas field project utilized a 0.3 RBR (0.1 RAPBR + 0.2
RASBR with MWAS) with no RAs, and the virgin binder grade was reduced in terms
of PGH grade to a PG 64-22 (68-25) based on standard practice. At this RBR, the

89
resulting grade is a PG 70-10 (75-13). Thus, the standard practice restored the PGH
grade, but the PGL grade was significantly inadequate.
 When a RA (T1) was added to the control mixture in the Texas field project, the
resulting grade was a PG 64-16 (69-16). Thus, reduction of the PGH grade of the
virgin binder and addition of a specific RA (T1) at the selected dosage rate (within
the range to restore the target PGL grade (optimum or minimum dosage) and the
target PGH grade (high dosage) by binder blend testing) provided an almost adequate
PGH grade. However, the RA did improve the PGL grade but did not restore it
completely.
 If the virgin binder had not been changed in the control mixture without RAs (i.e.,
retained use of the PG 70-22), the resulting grade would have been a PG 76-10 (80-
15) with the same inadequate PGL grade and a more than adequate PGH grade.
 If the virgin binder had instead been reduced in terms of both PGH and PGL grades in
the control mixture without RAs (i.e., used the supplemental S-controlled PG 64-28
[66-30]), the resulting grade would have been a PG 76-22 (77-22) with restoration of
the PGL grade and a more than adequate PGH grade.

Figure 4.15. Estimated Continuous Grades for Blended Binders at 28.6 Percent RBR.

90
80
70
60
50
Continuous PG grade (°C)

64‐22 PGH
40
64‐22 PGL
30
64‐28 PGH
20
64‐28 PGL
10
70‐22 PGH
0 70‐22 PGL
‐10
‐20
‐30
virgin control with RA

Figure 4.16. Estimating the Restoration of Blended Binders at 28.6 Percent RBR.

In summary, the mortar procedure was capable of characterizing the effects of RAP,
RAS, and RA materials on virgin binders without the use of chemical extraction and unrealistic
complete blending required for binder experiments. However, the test variability was high in
some cases, which necessitated the testing of additional replicates. Blending charts in the form of
surface plots were developed for the various material combinations evaluated to demonstrate the
blending possibilities. A summary plot showing the predicted change in continuous PG binder
grade from virgin mortars to control mortars to mortars with RAs for a specific RBR and
combination of materials was developed to assess the different factors that contribute to restoring
binder rheology.

4.4.2 Expanded Mortar Experiment

The next step in understanding the more realistic blended binders using the mortar
procedure is to complete the preliminary experiment by examining both the PG 70-22 and the
PG 64-28 for mixtures with the RA (T1) to determine if these blends can meet the target
performance grade of the virgin mixture (PG 70-22) at the field RA dosage rate. The effects of
polymer modification will also be explored for a PG 64-28P virgin binder. Thus, the expanded
Phase II mortar experiment, as shown in Table 4.8, will build on the preliminary results shown in
the previous section (for the shaded material combinations) and utilize the more common RAS
source (TOAS1) to better understand the effect of these stiffer recycled materials with and
without RAs on the blended binder continuous PG grade at the RA dosage level used in the
Texas field project.
Table 4.8 also shows that the effect of RA dosage will be explored for a limited number
of material combinations with the PG 64-28 virgin binder. In addition, an analogous experiment
with another RA type (the traditional A1 type) will be completed as shown in Table 4.9 to
compare the two different RA types. For this second expanded mortar experiment, MWAS and

91
the PG 70-22 virgin binder will not be included and the optimum RA dosage will be determined
based on blended binder testing to restore the target PG grade of the virgin mixture in the Texas
field project.
With the testing completed as shown in Table 4.8 and Table 4.9 for the two selected RA
types (T1 and A1), the continuous PG grades for blended binders at the optimum RA dosage
used in the Texas field project (and a high RA dosage for one virgin binder) will be predicted
over a range of the following RBR values, each with one or more different RAP and RAS
percentages or RAPBR and RASBR values:

 0.3 RBR with 0.3 | 0.0 and 0.1 | 0.2 RAPBR | RASBR combinations.
 0.4 RBR with 0.4 | 0.0, 0.0 | 0.4, 0.2 | 0.2, and 0.3 | 0.1 RAPBR | RASBR
combinations.
 0.5 RBR with 0.25 | 0.25 RAPBR | RASBR combination.

As time and budget allow, fractions of the experiments shown in Table 4.9 will be
completed for other RA types (P or R) that may not be as compatible (P type) or are associated
with recent poor performance (R type; Mooney 2015).

92
Table 4.8. Expanded Mortar Experiment with RA T1.
Virgin PGH, PGI, PGL, &
Binder RAP RAS RA# Tc
TX — — √ (PGH only)
— — MWAS — √ (PGH only)
— TOAS1 — √ (PGH only)
— — — √√√
— — T1 √√√
TX — — √√√
— MWAS — √√√
— TOAS1 — √√√
64-22
TX — T1 √√√
— MWAS T1 √√√
— TOAS1 T1 √√√
TX MWAS T1 Calculated
TX TOAS1 T1 Calculated
— — — √√√
— — T1 √√√
TX — — √√√
— MWAS — √√√
— TOAS1 — √√√
70-22
TX — T1 √√√
— MWAS T1 √√√
— TOAS1 T1 √√√
TX MWAS T1 Calculated
TX TOAS1 T1 Calculated
— — — √√√
— — T1 √√√^
TX — — √√√
— MWAS — √√√
— TOAS1 — √√√
64-28
TX — T1 √√√^
— MWAS T1 √√√^
— TOAS1 T1 √√√^
TX MWAS T1 Calculated^
TX TOAS1 T1 Calculated^
— — — √√√
— — T1 √√√^
TX — — √√√
— MWAS — √√√
— TOAS1 — √√√
64-28P
TX — T1 √√√^
— MWAS T1 √√√^
— TOAS1 T1 √√√^
TX MWAS T1 Calculated^
TX TOAS1 T1 Calculated^
^
Mortars with a high RA dosage based on restoration of PGH grade will also be included.
#
At Texas field dosage.

93
Table 4.9. Expanded Mortar Experiment with RA A1.
Virgin PGH, PGI, PGL,
#
Binder RAP RAS RA & Tc
TX — — √ (PGH only)

— TOAS1 — √ (PGH only)
— — — √√√
— — A1 √√√
TX — — √√√
64-22 — TOAS1 — √√√
TX — A1 √√√
— TOAS1 A1 √√√
TX TOAS1 A1 Calculated
— — — √√√
— — A1 √√√^
TX — — √√√
64-28 — TOAS1 — √√√
TX — A1 √√√^
— TOAS1 A1 √√√^
TX TOAS1 A1 Calculated^
— — — √√√
— — A1 √√√^
TX — — √√√
64-28P — TOAS1 — √√√
TX — A1 √√√^
— TOAS1 A1 √√√^
TX TOAS1 A1 Calculated^
^
Mortars with a high RA dosage based on restoration of PGH grade will also be included.
#
At Texas field dosage.

The blended binder continuous PGH, PGI, and PGL grades will be presented as the
preliminary results are in the previous section. They will be determined based on the mortar
procedure in the latest draft of standard test method AASHTO T XXX-12, Estimating Effect of
RAP and RAS on Blended Binder Performance Grade without Binder Extraction
(www.arc.unr.edu/Outreach
.html).

4.5 Binder Experiments

4.5.1 Preliminary Binder Results

Figure 4.17 and Figure 4.18 show preliminary results for PGH and PGL grades,
controlled by m-value and S, respectively, for the binder blends of PG 64-22, TX RAP, MWAS,
and the three TX RAs (Table 4.2) at different RA dosage rates. Based on these preliminary
results, the assumption of regional or piecewise linear blending appears to hold, and these
blended binder results can be utilized to determine an optimum (or minimum) RA dosage rate
based on restoring PGL grade to the target grade in the virgin mixture and a high (or maximum)

94
RA dosage rate based on restoring PGH grade. For the Texas field project materials, these RA
dosage rates are shown in Table 4.10 and are required to verify field quantities and for further
mortar and mixture testing.

RA T1(%)  RA T1(%) 

RA T2(%)  RA T2(%) 

Emulsion (%)  Emulsion (%) 
Figure 4.17. Validation of Regional Linear Blending Concept for Combined Blends of
PG 64-22, 10 percent TX RAP, 5 percent MWAS, and TX RAs Using M-Controlled Values
for PGL Grade.

95
RA T1(%)  RA T1(%) 

RA T2(%)  RA T2(%) 

Emulsion (%)  Emulsion (%) 
Figure 4.18. Validation of Regional Linear Blending Concept for Combined Blends of
PG 64-22, 10 percent TX RAP, 5 percent MWAS, and TX RAs Using S-Controlled Values
for PGL Grade.

Table 4.10. Dosage Rates for Texas Field Project.


Dosage
RA High to Restore Optimum to
PGH Grade Restore PGL Grade
T1 5.6 2.5 (m-controlled)
T2 6.7 2.5 (m-controlled)
Emulsion 9.6 4.7 (m-controlled)

96
Unlike for mortars, these blended binder continuous PG grades cannot be predicted for
the entire temperature spectrum and for different RBR values and RAP/RAS combinations
because in general, the recycled binders extracted from the RAP and RAS cannot be
characterized in terms of intermediate- and low-temperature properties due to their high stiffness
and associated equipment limitations. An evaluation of changes in rheology due to recycled
materials at different RAPBR and RASBR values can be completed for the PGH grade as shown
in Table 4.11 and Figure 4.19 for the Texas field project materials and supplemental PG 64-28
virgin binder. Blending tables such as those shown in Table 4.11 will be utilized to efficiently
select the material combinations (and RAPBR and RASBR values) for mixture testing.

97
Table 4.11. Binder Blending Table for Restoring PGH Grade^
(Estimated by Binder | Mortar Analysis with NA Results Not Available Yet).
RA T1
TX @ Field PG 64- PG 64- PG 70- PG 70- PG 64- PG 64-
RBR RAPBR RASBR RAP MWAS Dosage 22 22 w/T1 22 22 w/T1 28 28 w/T1
0.0 0.0 0.0 109 134 −9 68 63 74 NA 66 NA
0.3 0.0 80 | 76 79 | 71 85 | 82 83 | NA 79 | 81 78 | NA
0.3
0.1 0.2 85 | 75 84 | 69 90 | 80 88 | NA 84 | 77 83 | NA
0.4 0.0 84 | 79 83 | 73 88 | 85 87 | NA 83 | 86 82 | NA
0 0.4 94 | 78 93 | 70 98 | 82 97 | NA 93 | 78 92 | NA
0.4
0.2 0.2 89 | 79 88 | 72 93 | 83 92 | NA 88 | 82 87 | NA
0.3 0.1 87 | 79 86 | 73 91 | 84 89 | NA 86 | 84 85 | NA
0.5 0.25 0.25 95 | 81 94 | 74 98 | 86 97 | NA 94 | 86 93 | NA
^
For each combination, PGH grade provided based on binder estimation and then from mortar analysis (after |) with NA shown for results not available yet
98
Figure 4.19. Estimating the Restoration of PGH Grade for Binder Blends with MWAS.

These estimated PGH grades based on binder testing (with complete blending) and those
determined from the mortar procedure described previously will also be compared, as shown in
Figure 4.20, to further understand blending of the binder components in these mixtures. For the
Texas field project RAP, MWAS, and T1 materials at the field dosage rate of 2.65 percent by
weight of virgin binder replacement; the PGH grade determined by the more realistic blending in
the mortar procedure is on average 5 to 10°C lower than that determined by binder testing, with
ranges from 3°C higher to 16°C lower. In addition, the expected higher degree of blending of the
RAP binder as compared to the RAS binder is reflected by the mortar procedure but the opposite
is shown for the estimated grades based on complete blending of the binder.

Figure 4.20. Comparing Restoration of PGH Grade Based on Binder and Mortar Testing
for Blends with MWAS.

99
The expected effect of utilizing stiffer TOAS is shown in Figure 4.21, and this effect will
be evaluated further in the expanded binder and mortar experiments. A comparison of the effect
of the different RAS sources is shown in Figure 4.22. Finally, G-R parameters at intermediate
temperatures on short-term aged and long-term aged binders will also be plotted in Black space
for these blends to quantify the restoration properties of the RA, as shown in Figure 4.23, based
on Zhou et al. (2015) for some of the Texas field project materials after 20, 40, and 80 hr in the
PAV. They found that the virgin PG 70-22 and the binder blend with 5 percent RA T1 required
approximately the same number of hours in the PAV to reach both the G-R damage onset curve
(at G-R = 180 kPa) and the G-R significant cracking curve (at G-R = 450 kPa). However, the
binder blend at a lower 2 percent RA T1 dosage required approximately half the number of hours
as the virgin PG 70-22 to reach these two durability curves.

Figure 4.21. Comparing Restoration of PGH Grade Based on Binder and Mortar Testing
for Blends with TOAS (Expected Results).

Figure 4.22. Comparing Restoration of PG Grade Based on Different RAS Sources.

100
RA T1
RA T1

Figure 4.23. Evaluating Restoration of PG Grade in Black Space (Zhou et al. 2015).

4.5.2 Expanded Binder Rheology Experiments

Prior to extensive Phase II laboratory testing of binder blends, some combinations of the
supplemental materials (including different RAP and RAS sources) for each different RA will be
utilized to further validate the assumption of regional linear blending. Thus, the proposed
expanded binder experiments for Phase II that consider the materials collected at the Texas field
project and the selected supplementary materials are shown in Table 4.12, Table 4.13, and Table
4.14 for three virgin binder types and two RA types to characterize component materials and
combined blends, respectively, using the selected laboratory tests described previously.
Table 4.12 will be utilized to characterize the component binders and RAs and identify
key properties that produce optimal binder blends. These key properties include those currently
contained in ASTM D4552, which defines six RA grades based on viscosity at 60°C and DSR
master curve parameters (including G-R parameter, rheological index R-value, G*c, and c),
where larger DSR plates of 50 mm in diameter are needed to characterize the RAs. Preliminary
results by Zhou et al. (2015) indicate that the RA T1 utilized in the Texas field project is an RA1
grade by this specification. Results from the experiment shown in Table 4.12 and the
compatibility experiment described subsequently will be utilized to formulate a specification for
RAs to replace ASTM D4552.

101
Table 4.12. Expanded Component Binders and RA Experiment
RAP
Virgin Binder RAS Binder RA
Binder

Aromatic Oil
Aging State%

Tall Oil (T1)

Re-Re fined
PG 64-28P

Paraffinic
PG 70-22

PG 64-22

PG 64-28

Lube Oil
NH RAP
TX RAP

Oil (P1)
MWAS

TOAS1

TOAS2

(R1)
(A1)
Test

U √ √ √ √ √ √ √ √ √
PGH
S √ √ √ √ √ √ √ √ √
PGL & Tc L √ √ √ √ √ √ √ √ √
U √ √ √ √
 @ 60°C
S √ √ √ √
DSR Master U √ √ √ √
Curve (with
50 mm
S √ √ √ √
plates for
RA)
%
U = original, unaged; S = short-term aged in lab; L = long-term aged in lab.

102
Table 4.13. Expanded Binder Blend Experiment for RA Type T1.
DSR
Master
Curve
with G-R
Virgin RAPBR RASBR & PGH & PGL (S & 2 L
Binder RBR & Source Source RA @ 0, 2, & 10% RA Aging)
0 — — — √ (0% only) √√√
0.1 TX 0.2 MWAS √√√ √√√
0.3
0.1 TX 0.2 TOAS √√√ √√√
0.4 TX 0.0 √√√ √√√
64-22
0.4 NH 0.0 T1 √√√ √√√
0.4
0.2 TX 0.2 TOAS √√√ √√√
0.0 0.4 TOAS √√√# √√√
0.5 0.25 TX 0.25 TOAS √√√# √√√
0 — — — √ (0% only) √√√
0.4 TX 0.0 √√√ √√√
0.4 NH 0.0 √√√ √√√
0.2 TX 0.2 TOAS √√√ √√√
64-28 0.4
0.2 NH 0.2 TOAS T1 √√√ √√√
0.2 NH 0.2 TOAS2 √√√ √√√
0.0 0.4 TOAS √√√# √√√
0.5 0.25 TX 0.25 TOAS √√√# √√√
0 — — — √ (0% only) √√√
0.4 TX 0.0 √√√ √√√
64-28P 0.4
0.2 TX 0.2 TOAS T1 √√√ √√√
0.5 0.25 TX 0.25 TOAS √√√# √√√
#
Selected combinations to be tested first to validate assumption of regional linear blending.

103
Table 4.14. Expanded Binder Blend Experiment for RA Type A1.
DSR
Master
Curve
with G-R
Virgin RAPBR RASBR & PGH & PGL (S & 2 L
Binder RBR & Source Source RA @ 0, 2, & 10% RA Aging)
0 — — — √ (0% only) √√√
0.4 TX 0.0 √√√ √√√
0.4 NH 0.0 √√√ √√√
64-22 0.4
0.2 TX 0.2 TOAS A1 √√√ √√√
0.0 0.4 TOAS √√√# √√√
0.5 0.25 TX 0.25 TOAS √√√# √√√
0 — — — √ (0% only) √√√
0.4 TX 0.0 √√√ √√√
0.4 NH 0.0 √√√ √√√
0.2 TX 0.2 TOAS √√√ √√√
64-28 0.4
0.2 NH 0.2 TOAS A1 √√√ √√√
0.2 NH 0.2 TOAS2 √√√ √√√
0.0 0.4 TOAS √√√# √√√
0.5 0.25 TX 0.25 TOAS √√√# √√√
0 — — — √ (0% only) √√√
0.4 TX 0.0 √√√ √√√
64-28P 0.4
0.2 TX 0.2 TOAS A1 √√√ √√√
0.5 0.25 TX 0.25 TOAS √√√# √√√
#
Selected combinations to be tested first to validate assumption of regional linear blending.

For the experiments shown in Table 4.13 and Table 4.14 (without the shaded
combination associated with the Texas field project that included MWAS and its corresponding
TOAS blend at 0.3 RBR), the optimum RA dosage will be determined based on restoring the
PGL grade and used when assessing the Glover-Rowe parameter for short-term aged blends and
long-term aged binder blends and mixture performance properties. As shown, the effects of the
following factors will be explored for some material combinations:

 RAS type (MWAS vs. TOAS).


 RAP source.
 RAS source.
 RAP only vs. balanced RAP/RAS vs. RAS only.
 Virgin binder type (including modified PG 64-28P).
 RBR.
 RA type.

Estimation of low-temperature PG binder properties from DSR frequency sweep testing


will also be utilized for efficiency with the use of extracted and recovered binders from RAP and
RAS based on preliminary validation results for binder blends with RAs. These testing protocols

104
have been successfully utilized by Hanz and Bahia (2010) and Johnson (2010) to measure
viscoelastic properties in the DSR at low load levels over a range of loading frequencies
(typically from 0.1 to 30 Hz) and intermediate temperatures to predict low-temperature PG
properties using interconversion techniques (Anderson et al. 1994). Preliminary results for the
Texas field project materials with the T1 RA, as shown in Figure 4.24 for 8 mm DSR frequency
sweeps conducted at 6°C, indicate consistent rankings and some correlation between PGL grades
based on measured and predicted m-values. To further explore the possibility of obtaining
satisfactory correlations for both m-value and S-value, 4 mm DSR frequency sweep testing at a
temperature 10°C higher than the PGL grade of the virgin binder (e.g., −22°C for a PG 64-22)
based on the procedures by Farrar et al. (2014) will be explored. Selected material combinations
with the PG 64-28 virgin binder will also be utilized at a high RA dosage rate limited by
restoring the PGH grade (as currently proposed for CA for a RAP-only mixture and proposed in
the expanded mortar experiments (Table 4.8 and Table 4.9)).

(a) (b)
Figure 4.24. Comparison of Low-Temperature PG Grade from BBR Measurements to
Those Predicted from 8 mm DSR Frequency Sweeps Based on (a) S-Controlled PGL, and
(b) m-Controlled PGL.

4.5.3 Binder Compatibility Experiment

The exudation droplet test currently utilized by a tall oil RA supplier and originally
developed by Shell will also be explored as a screening test for compatibility of the binder
blends. Table 4.15 shows proposed blends to be tested with the exudation droplet test for three
different RA types plus some expected incompatible blends of TOAS with RA types P and R.
The selected material combinations are focused on the stiffer high RBR- and RAS-only
combinations and the combination utilized in the Texas field project (highlighted). If the
exudation droplet test does not prove useful, the spot test (AASHTO T 102) may be evaluated
for the same combinations shown in Table 4.15.

105
Table 4.15. Blends for Compatibility Experiment.
DSR Master
Exudation Droplet Curve
Virgin RAPBR & RASBR & @ Opt & High (S & 2 L
Binder RBR Source Source RA Dosage Aging)
0.4 — TOAS T1,
— A1, √√√√√√√√
P, R
0.4 — TOAS2 T1,
— A1, √√√√√√√√
P, R
0.3 0.1 TX 0.2 MWAS T1, √√ (T1 only)
64-22 0.4 0.0 0.4 TOAS A1, √√√√√√
0.5 0.25 TX 0.25 TOAS R √√√√√√
Available from
0.4 0.0 0.4 TOAS T1, √√√√√√
Expanded
64-28 0.5 0.25 TX 0.25 TOAS A1,
√√√√√√ Binder Blend
R
Experiment
0.4 0.4 TX 0.0 T1, √√√√√√
64-28P 0.5 0.25 TX 0.25 TOAS A1,
√√√√√√
R

For combinations that are incompatible with the selected screening test, the Automated
Heithaus Test (ASTM D6703) or SARA analysis (ASTM D4124) will be considered to
determine if additional chemical properties are required to predict binder blends that should not
be utilized. This additional analysis will only be pursued if incompatible blends cannot be
identified through rheological properties determined from DSR master curves and PG grading
(Tc). Recent research by Ryan et al. (2014) has shown that Tc is able to screen for
compatibility, and this same rheological parameter is tied to cracking performance in MN with a
threshold of approximately 5 based on work by Anderson et al. (2011). With the ultimate goal of
acceptable mixture performance, incompatible binder blends will be examined to determine if
rheological properties can preclude their use and limited mixture testing will be considered to
define inadequate mixture performance.

106
4.5.4 Binder Aging Experiments

To build from existing knowledge presented in Chapter 2 on aging of binders and


predicting carbonyl area growth and subsequent hardening susceptibility from oxidation kinetics,
a binder aging experiment is proposed as shown in Table 4.16. This experiment first screens the
virgin binder with added RA to determine if these materials affect the rheological properties and
oxidation kinetics. Next, the virgin binders will be aged for 60 or 80 hr in the PAV to simulate
binders in RAP (but with known initial properties) based on the Black space plot in Figure 2.1
prior to adding RA. Finally, the blends that will be utilized in field projects will be evaluated. For
all of these steps, the following times and temperatures will be used to age the materials prior to
chemical characterization (FTIR CA determinations of oxidative aging) and physical
characterization (DSR measurements of the rheological master curve including the G-R
parameter):

 60°C for 4, 8, 15, 30, 60, 100, and 160 days.


 85°C for 0.5, 1, 4, 8, 15, 25, and 40 days.
 100°C for 0.08, 0.25, 0.5, 1, 4, 8, and 15 days (or 2, 6, 12, 24, 96, 192, and 360 hr).

107
Table 4.16. Binder Aging Experiment.
FTIR, DSR
Master Curve
Virgin after Aging @
Binder RAP RAS RA# multiple t, T
— — — √√
^
— — T1 √√
64-22 — — A1^ √√
— — P √√
— — R √√
64-22 — — — √√
after 60 — — T1 √√
or 80 hr — — A1
√√
PAV
TX MWAS T1 √√
TX TOAS T1^ √√
TX MWAS A1^ √√
64-22
TX TOAS A1 √√
TX TOAS P √√
TX TOAS R √√
— — — √√
64-28 — — T1 √√
— — A1 √√
64-28 — — — √√
after 60 — — T1 √√
or 80 hr — — A1
√√
PAV
TX MWAS T1 √√
TX TOAS T1 √√
64-28
TX MWAS A1 √√
TX TOAS A1 √√
— — — √√
64-28P — — T1 √√
— — A1 √√
64-28P — — — √√
after 60 — — T1 √√
or 80 hr — — A1
√√
PAV
TX MWAS T1 √√
TX TOAS A1 √√
64-28P
TX MWAS A1 √√
TX TOAS A1 √√
^
Limited combinations with a high RA dosage based on PGH grade will also be considered.
#
At Texas field dosage.

108
Isothermal frequency sweeps tests at different temperatures will be conducted to
determine the master curves of G* and . Table 4.17 summarizes the anticipated testing protocol
based on the virgin binder grade. The LSV, G-R parameter, and crossover modulus will be
calculated as a function of aging for the various evaluated binders at different aging levels. The
effect of RAs on the binder oxidative aging kinetics will also be studied.

Table 4.17. Proposed Testing Protocol for Developing Binder DSR Master Curves.
High Temp. Int.–Low Temp.
(25 mm plate; 0.01–100 rad/s)# (8 mm plate; 0.1–100 rad/s)^
46, 40, 34, 28,
Virgin 85, 95, 100°C 60, 70, 80°C 60, 64, 70°C 22°C (2 mm 15, 10, 4, −2°C
Binder (0.5 mm gap) (1 mm gap) (1 mm gap) gap) (2 mm gap)
64-22 √ √ √
64-28 √ √ √
64-28P √ √ √ √ √
^
Total of 10 data points per isotherm.
#
Total of 19 data points per isotherm.

Aging results at a single temperature of 85°C will be evaluated first in a preliminary


screening experiment to detect possible effects of different types of RAs on binder oxidative
aging kinetics and hardening susceptibility parameters in an effort to reduce the scale of the
experiments to be done and determine the most efficient tools to recommend at the conclusion of
this study. This preliminary screening experiment will be conducted for four RA types (T1, A1,
P, and R), one virgin binder type (PG 64-22), a single RAP (TX) source, and a single RAS
(TOAS) source. In these experiments, instead of measuring extensive data for all three
temperatures as described and up to 160 days, binder oxidation reaction rates at only 85°C will
be determined. Based on the discussion in Chapter 2, this single temperature was selected as the
farthest from the isokinetic temperature but practical in terms of the capabilities of the equipment
and time required to obtain results. Figure 4.25 shows an example of CA growth at 85°C to be
measured in this preliminary screening experiment. From these CA growth data, the oxidation
reaction rate of each binder-blend combination will be evaluated. It is expected that if RA
changes the binder oxidation kinetics significantly, as discussed in Chapter 2, this change will be
reflected in different reaction rates at a given temperature. Thus, by comparing the reaction rate
data of virgin binders and those blended with RA (and also with RAP/RAS), any combinations
where there is a significant effect can be identified for further evaluation with more precise but
time-consuming testing at multiple temperatures.

109
Figure 4.25. Example CA Growth over Time at 85°C.

In addition, a shortcut method with faster PAV oxidative aging of binders, rather than the
longer testing at atmospheric pressure, is also being evaluated and will be used as appropriate.
While indications so far are that it is possible to estimate a binder’s (or a blend’s) oxidation
kinetics parameters using PAV data (as described in Chapter 2 [Cui et al. 2014]), the accuracy of
such determinations is not yet well established. Furthermore, a binder’s hardening susceptibility
(degree of rheological hardening in response to oxidation as measured by FTIR CA growth) is
significantly different at PAV (elevated pressure) conditions versus atmospheric measurements;
so far, the only way to determine hardening susceptibility at atmospheric conditions is to
measure it at the slower atmospheric conditions.

4.6 Mixture Experiments

4.6.1 Preliminary Specimen Fabrication Results

The preliminary mixture experiment to determine the specimen fabrication protocol


(conditioning time and temperature) for LMLC specimens using the Texas field project materials
(Table 4.2) was completed, and the procedures followed are detailed in Appendix B. The
objective was to establish a conditioning time and temperature to prepare specimens in the
laboratory to match the initial stiffness in the field. MR testing for all specimen types including
LMLC with 2 hr of conditioning at 135°C (275°F) for all mixtures except Mixture 2 (which was
conditioned based on NCHRP 9-52 recommendations for WMA), cores at construction, PMLC
specimens (i.e., compacted onsite), and RPMLC specimens (i.e., compacted off site) was
conducted. These proposed short-term aging protocols were evaluated and validated in NCHRP
Projects 9-49 and 9-52 (Epps Martin et al. 2014; Yin et al. 2013; Yin and Garcia Cucalon et al.
2014; Yin et al. 2015).
The results were analyzed, and the corresponding results are presented in Figure 4.26
along with AV and production and construction temperature data. From the trends, it is apparent
that the LMLC specimen with 2 hr of conditioning at 135°C (275°F) for Mixture 1 (virgin
mixture) showed higher stiffness than the cores at construction, but lower stiffness than the
PMLC and RPMLC specimens. This follows trends shown in NCHRP 9-52 for a wide variety of

110
mixtures, including specifically those with recycled materials and WMA additives. For
Mixture 2 (control mix with RAP, RAS, and WMA additive), the stiffness of the LMLC
specimens with 2 hr of conditioning at 116°C (240°F) showed the same trends. For Mixtures 3,
4, and 5 with RAs, the LMLC specimens with 2 hr of conditioning at 135°C (275°F) showed
statistically equivalent stiffness to the cores at construction but lower stiffness than that for
RPMLC specimens. For these mixtures without a WMA additive (Mixtures 3 and 5), the LMLC
specimens with 2 hr of conditioning at 135°C (275°F) also had lower stiffness than that for
PMLC specimens. However, for Mixture 4 with an additional WMA technology, equivalent
stiffnesses were shown for the LMLC specimens and the PMLC specimens.
These results provide justification for selecting an STOA protocol of 2 hr at 135°C
(275°F) for HMA and mixtures with RAs, either HMA or WMA. The STOA protocol of 2 hr at
116°C (240°F) for WMA recommended by NCHRP 9-49 and utilized and confirmed in
NCHRP 9-52 will continue to be utilized when RAs are not present.

111
1100
CORES PMLC R-PMLC LMLC 2 hrs
1000

900

800 AV
AV
Resilient Modulus (ksi)

6.5% AV
6.9%
700 7.1%
AV
6.9%
600 AV
AV
7±1%
AV 6.5%
500 AV AV 7±1%
AV 7.1% AV AV
9.6% AV AV
7.3% AV 8.2% 6.6% AV
7%
400 AV 9.8% 6.9% AV
7.2%
7±1%
9.9%
112

AV
300 6.8%

200 AV
9.3%

100

  0        
Mixture 1 (Virgin) Mixture 2 (Control) Mixture 3 (w/RA1) Mixture 4 (w/RA2) Mixture 5 (w/RA3)

Mixing
Temperature 325–327 310 275–280 280 280–290 280 275–280 280 280–287 280
(⁰F)
Compaction
285– 255– 265– 240– 260–
Temperature 290
270 270 275
260
250 250 240
270
250 250 275
250
250 250 275
265
250 250 275
(⁰F)

Mixture Type
Figure 4.26. MR Results for the Mixtures Used in the Texas Field Project.
4.6.2 Aging Analysis

Cumulative degree days (0°C base) are proposed for use in this study to provide a
quantitative basis for field aging that accounts for differences in construction dates and
environments at different field projects. As an example, the CDD values for seven field projects
from NCHRP 9-52 are presented in Figure 4.27 with coring dates indicated by a Black point.
Differences between field projects are shown with:

 Steeper overall slopes for warmer climates (Texas I, New Mexico, and Florida) as
compared to colder climates (Wyoming, South Dakota, Iowa, and Indiana).
 Flat initial slopes for fall or winter construction (as shown for South Dakota, which
was constructed in the fall).

A similar plot will be formulated for the field projects in this study to provide a distinct
indication of the individual climate and its cumulative effect since construction.

Texas I New Mexico Wyoming South Dakota

Iowa Indiana Florida


40000
Cumulative Degree Days (°F-days)

35000

30000

25000

20000

15000

10000

5000

0
Dec-11 Jul-12 Jan-13 Aug-13 Mar-14 Sep-14 Apr-15
Coring Date

Figure 4.27. Cumulative Degree Days for NCHRP 9-52 Field Projects.

Figure 4.28 shows the CDD values for post-construction cores from NCHRP Project 9-52
versus their associated MR ratios (aged/unaged) and an exponential trendline with a high
coefficient of determination (R2). Figure 4.28 also shows the corresponding average (1.76) MR
ratio value for all LMLC specimens with an STOA protocol of 2 hr at 135°C (275°F) plus LTOA
protocol of 5 days at 85°C (185°F) plotted as a marker where this value crosses the exponential
trendline for MR ratio versus CDD values. The vertical and horizontal error bars represent one
standard deviation from the average MR ratio values and corresponding CDD values,
respectively. A laboratory STOA protocol of 2 hr at 135°C (275°F) plus LTOA protocol of
5 days at 85°C (185°F) was able to produce mixture aging equivalent to an average of 16,000
CDD in the field. Based on the CDD curves shown in Figure 4.27, the in-service time for each
field project corresponding to 16,000 CDD was determined and summarized in Table 4.18. As
shown, the laboratory STOA protocol of 2 hr at 135°C (275°F) plus LTOA protocol of 5 days at

113
85°C (185°F) was equivalent to approximately 11 months in service in warmer climates and
22 months in service in colder climates.

Figure 4.28. MR Ratio versus CDD for NCHRP 9-52 Post-Construction Cores and
Correlation of LTOA Protocols with Field Aging.

Table 4.18. Correlation of Field Aging in Terms of In-Service Time and Laboratory LTOA
Protocol of 5 days at 85°C (185°F) for NCHRP 9-52 Field Projects.
MR Ratio for 2 hr at 135°C
Field Project Climate (275 °F) + 5 days at 85°C
(185°F)
Texas I 10 months
Warmer
New Mexico 12 months
Climate
Florida 11 months
Average 11 months
Wyoming 22 months
South Dakota Colder 22 months
Iowa Climate 22 months*
Indiana 21 months*
Average 22 months
*Predicted in-service time based on historical climatic information.

Similar plots will be formulated for the field projects in this study, but the exponential
trendline to be utilized from NCHRP 9-52 will be that from only RAP mixtures, as shown in
Figure 4.29.

114
Figure 4.29. MR Ratio versus CDD for NCHRP 9-52 Post-Construction Cores for RAP
Mixtures.

In addition to the tie of laboratory long-term aging to specific field projects/locations and
construction dates, binder kinetics measured for blends used in field projects will be utilized to
predict mixture stiffness (E*) through a relationship between E* master curve parameters and
CA growth from kinetics (Alavi et al. 2013). E* will be measured at a minimum of three
different aging levels: after short-term aging, after 5 days at 185F, and after a longer duration at
185F determined from the binder kinetics data for the location of interest. The binder will also
be extracted and recovered from the E* samples for carbonyl measurements. Using the measured
data, the E* will be determined following the 2S2P1D (2 springs, 2 parabolic, 1 dashpot) model
(Olard and Benedetto 2003). Consequently, the relationship between the 2S2P1D model
parameters and the increase in CA will be developed, hence allowing for the prediction of the E*
at any oxidative aging level. Predicted E* values will then be validated using measured E* for
aged cores (at approximately 1 year after construction). For virgin mixtures, binders will be
extracted and recovered to validate the CA growth predictions. Finally, the CDD estimates for
laboratory aging will be validated using these same relationships.

4.6.3 Expanded Mixture Experiments

Proposed Phase II mixture experiments will focus on all material combinations utilized in
2014 or 2015 field sections (Table 4.1) plus expansion to include high RBR values of 0.5 and
high RA dosage rate for selected combinations. For each field project, expanded material
combinations for mixture testing will be selected based on the results of the expanded mortar and
binder experiments toward efficient use of materials, time, and budget in this study. For example,
Table 4.19 shows expanded material combinations proposed for the Texas field project with
expansion to include the PG 64-28 virgin binder grade based on the corresponding promising
mortar results.

115
Table 4.19. Expanded Mixture Experiment for Texas Field Project.
RAPBR
Virgin & RASBR & MR E* + E* +
Binder RBR Source Source RA# @ 4, 25, 40°C EBM S-VECD UTSST

ARPMLC

ARPMLC
ULMLC

ALMLC

ULMLC

ALMLC

ALMLC

ALMLC
RPMLC

Cores

Cores

Cores
Aging State

or

or
70-22 0.0 — — — √ √ √ √√ √ √ √ √√ √ √√
64-22 0.3 0.1 TX 0.2 MWAS — √ √ √ √√ √ √ √ √√ √ √√
0.3 0.1 TX 0.2 MWAS √ √ √ √√ √ √ √ √√ √ √√
64-22+ 0.5 0.25 TX 0.25 MWAS T1+ √ √ √ √ √ √ √
0.5^ 0.25 TX 0.25 TOAS √ √ √ √ √ √ √
^
Mixtures with a high RA dosage based on restoration of PGH grade will also be included, and HWTT testing will be added to this experiment.
116

#
Optimum or field dosage.
+
Additional virgin binder PG 64-28 and/or RA type A1 will be evaluated based on binder and mortar results.
The MR testing will be used to verify specimen fabrication (conditioning) and aging
protocols for use in the remainder of the study by comparing mixture stiffness for unaged LMLC
(ULMLC), aged LMLC (ALMLC), and aged reheated PMLC (ARPMLC) specimens and field
cores at construction and after 1 year of field aging. Based on the intermediate aging results of
NCHRP 9-52 that included almost 30 different mixtures, a long-term oven aging protocol of
5 days at 85°C (185°F) will be considered for the ALMLC specimens to simulate 1–2 years of
field aging, and the specimen fabrication protocols discussed previously will be utilized for the
ULMLC specimens to match the initial field state. EBM testing will be used to evaluate the
effect of aging on cracking performance of LMLC specimens with and without RAs. Fatigue
parameters from the EBM approach for mixtures with expanded material combinations may also
be predicted using stiffness (E*) measurements from master curves (or a single temperature),
AV, PG binder grade, and aggregate gradation based on equations presented previously. Finally,
the S-VECD and UTSST testing will be used to evaluate the long-term laboratory and field
cracking performance of mixtures with and without RAs. This preliminary experiment will be
revised and refined based on the results of the binder blend and mortar laboratory results, and
analogous experiments will be developed for each field project to select material and testing
combinations that most efficiently meet the study objectives while considering the limitations of
materials, time, and budget.

4.7 Statistical Analysis, Technical Reports, and Phase III

4.7.1 Statistical Comparisons

As an example of the types of statistical comparisons that will be utilized to meet the
objectives of this study, Table 4.13 and Table 4.14 show the expanded binder blend experiment
for two RA types. The main study factor of interest in the overall study is RA type with four
levels (T, A, P, R), but the focus of most of the experiments is on two types (A and T). The other
factors considered for the blended binders are virgin binder type with three or four levels
(including PG 64-22, PG 70-22, PG 64-28, and PG 64-28P), RAP source with two levels (TX,
NH), RAS source with three levels (MWAS, TOAS, TOAS2), and RBR with up to four levels
(0, 0.3, 0.4, 0.5). Researchers are interested in assessing the effects of the individual factors and
possible interaction effects among them (most importantly, interactions between RA type and
other factors). Once the values for the response variables (laboratory test results from the
experiments) are obtained, the resulting data will be analyzed by employing analysis of variance
(ANOVA) or analysis of covariance (ANCOVA) depending on the presence of any covariates in
the experiment. Restrictions (constraints) in randomization in the experiment will be taken into
account in the analysis by utilizing variations of ANOVA such as randomized block designs or
split-plot design analysis. It is important to consider potential interaction effects between RA
type and other factors to examine whether the effectiveness of RA depends on other factors such
as virgin binder type, RAP source, RAS source, and/or RBR. Multiple comparison procedures—
Tukey-Kramer Honestly Significant Difference test or Fisher’s Least Significant Difference
test—will be performed following detection of any significant main effects or interactions to
identify which levels or factor level combinations are statistically different from others while
controlling an overall (experiment-wise) Type I error rate to be under 0.05.
Statistical analyses of the test results from the experiments for assessing effectiveness of
RA in mortars with high RBRs and for assessing effectiveness of RA in mixtures with high
RBRs will be conducted in a similar way, using ANOVA (or ANCOVA), incorporating main

117
effects and important two-way interaction effects (e.g., interactions of other factors with RA
type) into the analysis first, followed by multiple comparison tests.
The results from statistical analyses of various test results will be presented by various
plots of predicted values (least square means plots including interaction plots) as well as by least
square means (predicted values) along with their uncertainty estimates (standard errors) at
various factor levels/combinations.

4.7.2 Technical Reports

Based on the results produced in Phase II, a second draft interim report will be developed
to document the results of the extensive laboratory investigation described in this interim report
with the objectives listed at the beginning of this chapter. End products from Phase II of this
study are envisioned to include the following:

 Evaluation tools for assessing the effectiveness of RAs and its evolution for specific
material combinations at high RBRs and at specific field locations.
 Better understanding of the effectiveness of traditional aromatic and greener
alternative tall oil RAs in restoring blended binder rheology and improving mixture
cracking performance at high RBRs.
 Recommendations for evaluating the evolution of RA effectiveness in improving
mixture cracking performance.

The proposed evaluation tools in the form an overall mix design and performance
evaluation guidelines that accounts for the evolution of RA effectiveness will be a detailed
version of the example shown in Figure 4.30. In addition, the internal advisory panel and any
volunteer external advisory panel members will review the draft second interim report, and it
will be submitted to the NCHRP panel for review and approval.

118
Figure 4.30. Example Mix Design and Performance Evaluation Guidelines and Evaluation
Tools.

4.7.3 Phase III

While the Phase II laboratory experiments are being conducted, coordination will
continue toward procuring field projects as proposed in Table 4.1. Selection of material
combinations to explore further in Phase III will be based on the results from the Phase II work
plan contained in this report.

119
REFERENCES

Advanced Asphalt Technologies (2011) NCHRP 9-33: A Manual for Design of Hot Mix Asphalt
with Commentary, NCHRP Report 673, Transportation Research Board, National Research
Council, Washington, DC.

Alavi, M.Z., and E.Y. Hajj (2014) “Effect of Cooling Rate on the Thermo-Volumetric, Thermo-
Viscoelastic, and Fracture Properties of Asphalt Mixtures.” 12th ISAP Conference on Asphalt
Pavement, International Society for Asphalt Pavements, Raleigh, NC (June).

Alavi, M.Z., E.Y. Hajj, and N.E. Morian (2013) “Approach for Quantifying Effect of Binder
Oxidative Aging on Viscoelastic Properties of Asphalt Mixtures.” Transportation Research
Record No. 2373, Transportation Research Board, National Research Council, Washington, DC,
pp. 109-120.

Anderson, D.A., D.W. Christensen, H.U. Bahia, R. Donger, M.G. Sharma, C.E. Antle, and J.
Button (1994) “Binder Characterization and Evaluation, Volume 3: Physical Characterization.”
SHRP Report A-369, National Research Council, Washington, DC, 475 p.

Anderson, R.M., G.N. King, D.I. Hanson, and P.B. Blankenship (2011) “Evaluation of the
Relationship between Asphalt Binder Properties and Non-Load Related Cracking.” Journal of
the Association of Asphalt Paving Technologists, Vol. 80, pp. 615-649.

Asli, H., E. Ahmadinia, M. Zargar, and M. R. Karim (2012) “Investigation on Physical


Properties of Waste Cooking Oil—Rejuvenated Bitumen Binder.” Construction and Building
Materials, Vol. 37, pp. 398-405.

Asphalt Institute (1986) “Asphalt Hot-Mix Recycling.” The Asphalt Institute Manual, Series No.
20 (MS-20), Second Edition, 46 p.

Bahia, H., and D. Swiertz (2011) “Design System for HMA Containing a High Percentage of
RAS Material.” Final Report for RMRC Project 66, Recycled Materials Resource Center,
Madison, WI, 40 p.

Bennert, T., J.S. Daniel, and W. Mogawer (2014) “Strategies for Incorporating Higher Recycled
Asphalt Pavement Percentages: Review of Implementation Trials in Northeast States.”
Transportation Research Record No. 2445, Transportation Research Board, National Research
Council, Washington, DC, pp. 83-93.

Booshehrian, A., W.S. Mogawer, and R. Bonaquist (2013) “How to Construct an Asphalt Binder
Master Curve and Assess the Degree of Blending between RAP and Virgin Binders.” Journal of
Materials in Civil Engineering, Vol. 25, No. 12, pp. 1813-1821 (December).

Burke, K., and S.A. Hesp (2011) “Penetration Testing of Waste Engine Oil Residue Modified
Asphalt Cements.” 1st Conference of Transportation Research Group of India (CTRG),
Bangalore, India, Proceedings.

120
Buss, A., A. Cascione, and R.C. Williams (2014) “Evaluation Of Warm Mix Asphalt Containing
Recycled Asphalt Shingles.” Construction and Building Materials, Vol. 61, pp. 1-9.

Carpenter, S.H., and J.R. Wolosick (1980) “Modifier Influence in the Characterization of Hot-
Mix Recycled Material.” Transportation Research Record No. 777, Transportation Research
Board, National Research Council, Washington, DC, pp. 15-22.

Chaffin, J.M., R.R. Davison, C.J. Glover, and J.A. Bullin (1995) “Viscosity Mixing Rules for
Asphalt Recycling.” Transportation Research Record No. 1507, Transportation Research Board,
National Research Council, Washington, DC, pp. 78-85.

Chehab, G.R., Y.R. Kim, R.A. Schapery, M. Witczack, and R. Bonaquist (2002) “Time-
Temperature Superposition Principle for Asphalt Concrete Mixtures with Growing Damage in
Tension State.” Journal of the Association of Asphalt Paving Technologists, Vol. 71, pp. 559-
593.

Coffey, S., E. Dubois, Y. Mehta, and C. Purdy (2013) “Determining Impact of Degree of
Blending Between Virgin and Reclaimed Asphalt Binder on Predicted Pavement Performance
Using Mechanistic-Empirical Pavement Design Guide.” 92nd Annual Meeting of the
Transportation Research Board, Washington, DC, Proceedings.

Copeland, A. (2011) Reclaimed Asphalt Pavement in Asphalt Mixtures: State of the Practice,
Report No. FHWA-HRT-11-021, Federal Highway Administration, Washington, DC.

Copeland, A., J. D’Angelo, R. Dongre, S. Belagutti, and G. Sholar (2010) “Field Evaluation of a
High Reclaimed Asphalt Pavement-Warm Mix Asphalt Project in Florida-A Case Study.”
Transportation Research Record No. 2179, Transportation Research Board, National Research
Council, Washington, DC, pp. 93-101.

Cui, Y., X. Jin, R. Han, and C.J. Glover (2014). “An Accelerated Method for Determining
Asphalt Oxidation Kinetics Parameters for Use in Pavement Oxidation and Performance
Modeling.” Petroleum Science and Technology, Vol. 32, pp. 2691-2699.

Daniel, J.S., N. Gibson, S. Tarbox, A. Copeland, and A. Andriescu (2013) “Effect of long-Term
Ageing on RAP Mixtures: Laboratory Evaluation of Plant-Produced Mixtures.” Road Materials
and Pavement Design, Vol. 14, No. S2, pp. 173-192.

Daniel, J.S., J.L. Pochily, and D.M. Boisvert (2010) “Can More Reclaimed Asphalt Pavement be
Added? Study of Extracted Binder Properties from Plant-Produced Mixtures with up to 25%
Reclaimed Asphalt Pavement.” Transportation Research Record No. 2180, Transportation
Research Board, National Research Council, Washington, DC, pp. 19-29.

Daniel, J.S., and Y.R. Kim (2002) “Development of a Simplified Fatigue Test and Analysis
Procedure Using a Viscoelastic Damage Model.” Journal of the Association of Asphalt Paving
Technologists, Vol. 71, pp. 619-650.

121
Defeng, Q., Y. Jianying, and H. Xin (2014) “Effects of Diffusible Rejuvenator on Properties of
Recycled Asphalt Mixture.” Key Engineering Materials, Vol. 599, pp. 145-149 (February).

Epps, J.A., and R.J. Holmgreen (1980) Engineering, Economy and Energy Consideration in
Design, Construction and Materials: Design of Recycled Asphalt Concrete Mixtures.
Cooperative Research Project 2-9-74-214, Texas State Department of Highways and Public
Transportation and Texas Transportation Institute, College Station, TX (August).

Epps, J.A., D.N. Little, R.J. Holmgreen, and R.L. Terrel (1980) Guidelines for Recycling
Pavement Materials, NCHRP Report 224, National Cooperative Highway Research Program,
Washington, DC (September).

Epps Martin, A., E. Arambula, F. Yin, L. Garcia Cucalon, A. Chowdhury, R. Lytton, J. Epps, C.
Estakhri, and E.S. Park (2014) NCHRP 9-49: Evaluation of the Moisture Susceptibility of WMA
Technologies, NCHRP Report 763, Transportation Research Board, National Research Council,
Washington, DC.

Farrar, M., C. Sui, S. Salmans, and Q. Qin (2014) “Determining the Low-Temperature
Rheological Properties of Asphalt Binder Using a Dynamic Shear Rheometer (DSR).” Western
Research Institute, Laramie, WY (November).

FHWA (1993) A Study of the Use of Recycled Paving Materials: A Report to Congress, Report
No. FHWA-RD-93-147, Federal Highway Administration, Washington, DC.

Glaser, R., and L. Porot (2014) “Miscibility and Microstructure Property Changes in RAP Binder
/Rejuvenator Blends.” International Society for Asphalt Pavements Workshop, Raleigh, NC
(June), Technical Presentation.

Glover, J.C., R.R. Davison, C.H. Domke, Y. Ruan, P. Juristyarini, D.B. Knorr, and S.H. Jung
(2005) Development of a New Method for Assessing Asphalt Binder Durability with Field
Evaluation. Report No. FHWA/TX/05-1872-2, Texas Transportation Institute, College Station,
TX, 334 pp (August).

Hajj, E.Y., P.E. Sebaaly, and R. Shresta (2009) “Laboratory Evaluation of Mixes Containing
Recycled Asphalt Pavement (RAP).” International Journal of Road Materials and Pavements
Design, Vol. 10, No. 3.

Hajj, E.Y., M.I. Souliman, M.Z. Alavi, and L.G.L. Salazar (2013) “Influence of Hydrogreen
Bioasphalt on Viscoelastic Properties of Reclaimed Asphalt Mixtures.” Transportation Research
Record No. 2371, Transportation Research Board, National Research Council, Washington, DC,
pp. 13-22.

Han, R., X. Jin, and C.J. Glover (2011) “Modeling Pavement Temperature for Use in Binder
Oxidation Models and Pavement Performance Prediction.” Journal of Materials in Civil
Engineering, Vol. 23, No. 4, pp 351-359 (April).

122
Hanson, D.I., P.B. Blankenship, G.N. King, and R.M. Anderson (2010) Techniques for
Prevention and Remediation of Non-Load-Related Distresses on HMA Airport Pavements—
Phase II. Final Report, Airfield Asphalt Pavement Technology Program, Project 06-01.

Hanz, A., and H.U. Bahia (2010) “Initial Evaluation of an Approach to Estimate Low
Temperature Creep Stiffness Properties using Intermediate Temperature DSR Testing.” White
paper submitted to the FHWA Emulsion Task Force, Federal Highway Administration,
Washington, DC.

Hesp, S.A., and H.F. Shurvell (2010) “X-Ray Fluorescence Detection of Waste Engine Oil
Residue in Asphalt and Its Effect on Cracking in Service.” International Journal of Pavement
Engineering, Vol. 11, No. 6, pp. 541-553 (December).

Holmgreen, R.J., J.A. Epps, D.N. Little, and J.A. Button (1982) RAs for Recycled Bituminous
Binders—Executive Summary. Report No. FHWA-RD-82-122, Texas Transportation Institute,
Texas A&M University System, Arlington, TX, 7 pp.

Hussain, A., and Q. Yanjun (2012) “Evaluation of Asphalt Mixes Containing Reclaimed Asphalt
Pavement for Wearing Courses.” International Conference on Traffic and Transportation
Engineering (ICTTE 2012), pp. 43-48.

Im, S., and F. Zhou (2014) Field Performance of RAS Test Sections and Laboratory
Investigation of Impact of Rejuvenators on Engineering Properties of RAP/RAS Mixes. Report
No. FHWA/TX-14/0-6614-3, Texas A&M Transportation Institute, College Station, TX (April),
102 pp.

Im, S., F. Zhou, R. Lee, and T. Scullion (2014) “Impacts of Rejuvenators on Performance and
Engineering Properties of Asphalt Mixtures Containing Recycled Materials.” Construction and
Building Materials, Vol. 53, pp. 596-603 (February).

Irving, J.B. (1997) Viscosities of Binary Liquid Mixtures: The Effectiveness of Mixture
Equations, Report No. 631, National Engineering Laboratory, Department of Industry, U.K.

Jin, X., Y. Cui, and C.J. Glover (2013) “Modeling Asphalt Oxidation in Pavement with Field
Validation.” Petroleum Science and Technology, Vol. 31, No. 13, pp. 1398-1405.

Jin, X., R. Han, Y. Cui, and C.J. Glover (2011) “Fast-Rate-Constant-Rate Oxidation Kinetics
Model for Asphalt Binders.” Industrial and Engineering Chemistry Research, Vol. 50, pp.
13373-13379.

Johnson, C.M. (2010) “Estimating Asphalt Binder Fatigue Resistance Using an Accelerated Test
Method.” PhD. dissertation, Civil and Environmental Engineering, University of Wisconsin,
Madison, 132 pp.

Johnson, K. N., and S.A. Hesp (2014). “Effect of Waste Engine Oil Residue on Quality and
Durability of SHRP Materials Reference Library Binders.” Transportation Research Record No.

123
2444, Transportation Research Board, National Research Council, Washington, DC, pp. 102-
109.

Juristyarini, P., R.R. Davison, and C.J. Glover (2011) “Oxidation Hardening Kinetics of the
Rheological Function G′/(η′/G′) in Asphalts.” Petroleum Science and Technology, Vol. 29, pp.
2027-2036.

Kandhal, P.S. (1977) “ASTM STP 628: Low-Temperature Properties of Bituminous Materials
and Compacted Bituminous Paving Mixtures.” C.R. Marek (Ed.), American Society for Testing
and Materials, Philadelphia, PA.

Kandhal, P.S., and R.B. Mallick (1997) Pavement Recycling Guidelines for State and Local
Governments—Participant’s Reference Book. Report No. FHWA-SA-97, Chapter 7: Hot Mix
Asphalt Recycling (Materials and Mix Design), National Center for Asphalt Technology,
Auburn, AL.

Kari, W.J., N.E. Andersen, D.D. Davidson, H.L. Davis, R.N. Doty, S.J. Escobar, D.L. Kline, and
T.K. Stone (1980) “Prototype Specifications for RAs Used in Hot-Mix Recycling.” Journal of
the Association of Asphalt Paving Technologists, Vol. 49, pp. 177-192.

Karlsson, R., and U. Isacsson (2003) “Application of FTIR-ATR to Characterization of Bitumen


Rejuvenator Diffusion.” Journal of Materials in Civil Engineering, Vol. 15, pp. 157-165.

Kim, Y.R., and D.N. Little (1990) “One-Dimensional Constitutive Modeling of Asphalt
Concrete.” Journal of the Engineering Mechanics, Vol. 116, No. 4, pp. 619-650 (April).

King, G., M. Anderson, D. Hanson, and P. Blankenship (2012) “Using Black Space Diagrams to
Predict Age-Induced Cracking.” 7th RILEM International Conference on Cracking in
Pavements, Vol. 4, pp. 453-463, Delft, Netherlands (June).

Kriz, P., D.L. Grant, B.A. Veloza, M.J. Gale, A.G. Blahey, J.H. Brownie, R.D. Shirts, and S.
Maccarrone (2014) “Blending and Diffusion of Reclaimed Asphalt Pavement and Virgin Asphalt
Binders.” Journal of the Road Materials and Pavement Design, Vol. 15, No. S1, pp. 78-112.

Lee, H.J., and Y.R. Kim (1998a) “A Uniaxial Viscoelastic Constitutive Model for Asphalt
Concrete under Cyclic Loading.” Journal of the Engineering Mechanics, Vol. 124, No. 1, pp. 32-
40 (January).

Lee, H.J., and Y.R. Kim (1998b) “A Viscoelastic Continuum Damage Model of Asphalt
Concrete with Healing.” Journal of the Engineering Mechanics, Vol. 124, No. 41, pp. 1224-1232
(November).

Lee, T., R.L. Terrel, and J.P. Mahoney (1983) “Test for Efficiency of Mixing of Recycled
Asphalt Paving Mixtures.” Transportation Research Record No. 911, Transportation Research
Board, National Research Council, Washington, DC, pp. 51-60.

124
Li, X., M.O. Marasteanu, R.C. Williams, and T.R. Clyne (2008) “Effect of RAP (Proportion and
Type) and Binder Grade on the Properties of Asphalt Mixtures.” Transportation Research
Record No. 2051, Transportation Research Board, National Research Council, Washington, DC,
pp. 90-97.

Li, X., and N. Gibson (2013) “Using Small Scale Specimens for AMPT Dynamic Modulus and
Fatigue Tests.” Journal of the Association of Asphalt Paving Technologists, Vol. 82, pp. 579-
616.

Lin, J., J. Hong, C. Huang, J. Liu, and S. Wu (2014) “Effectiveness of Rejuvenator Seal
Materials on Performance of Asphalt Pavement.” Construction and Building Materials, Vol. 55,
pp. 63-68.

Lin, P., T.L. Wu, C. Chang, and B. Chou (2011) “Effects of RAs on Aged Asphalt Binders and
Reclaimed Asphalt Concrete.” Materials and Structures, Vol. 44, No. 5, pp. 911-921 (June).

Little, D.N., R.J. Holmgreen, and J.A. Epps (1981) “Effect of RAs on the Structural Performance
of Recycled Asphalt Concrete Materials.” Journal of the Association of Asphalt Paving
Technologists, Vol. 50, pp. 32-63.

Luo, X., R. Luo, and R. Lytton (2013a) “Modified Paris’s Law to Predict Entire Crack Growth in
Asphalt Mixtures.” Transportation Research Record No. 2373, Transportation Research Board,
National Research Council, Washington, DC, pp. 54-62.

Luo, X., R.L. Lytton, and R. Luo (2013b) NCHRP Project 1-52: A Mechanistic-Empirical Model
for Top-Down Cracking of Asphalt Pavement Layers. Phase I Interim Report, NCHRP.

Mallick, R., K.A. O’Sullivan, M. Tao, and R. Frank (2010) “Why Not Use Rejuvenator for 100%
RAP Recycling?” 89th Annual Meeting of the Transportation Research Board, Washington, DC,
Proceedings.

Martin, K.L., R.R. Davison, C.J. Glover, and J.A. Bullin (1990) “Asphalt Aging in Texas Roads
and Test Sections.” Transportation Research Record No. 1269, Transportation Research Board,
National Research Council, Washington, DC, pp. 9-19.

McDaniel, R., and R.M. Anderson (2001) Recommended Use of Reclaimed Asphalt Pavement in
the Superpave Mix Design Method: Technician’s Manual, NCHRP Report 452, National
cooperative Highway Research Program, Washington, D.C. 49 p.

Mensching, D.J., G.M. Rowe, J.S. Daniel, and T. Bennert (2015) “Exploring Low Temperature
Performance in Black Space.” Journal of the Association of Asphalt Paving Technologists,
accepted for publication.

Mirza, M.W., and M.W. Witczak (1995) “Development of a global aging system for short and
long term aging of asphalt cements” Journal of Association of Asphalt Paving Technologists,
Vol. 64, pp. 393-430.

125
Mogawer, W., A. Austerman, L. Mohammad, and M.E. Kutay (2013) “Evaluation of High RAP-
WMA Asphalt Rubber Mixtures.” Road Materials and Pavement Design, Vol. 14, pp. 129-147.

Mogawer, W., T. Bennert, J. Daniel, R. Bonaquist, A. J. Austerman, and A. Booshehrian (2012)


“Performance Characteristics of Plant Produced High RAP Mixtures.” Journal of the Road
Materials and Pavement Design, Vol. 13, pp. 183-208 (June).

Mogawer, W.S., A. Booshehrian, S. Vahidi, and A.J. Austerman (2013) “Evaluating the Effect
of Rejuvenators on the Degree of Blending and Performance of High RAP, RAS, and RAP/ RAS
Mixtures.” Journal of the Road Materials and Pavement Design, Vol. 14, Supplement 2, pp.
193-213.

Mooney, K. (2015) “Re-Refined Engine Oil Bottoms: Don’t Jump on the “Banned”-wagon Too
Quickly.” The Association of Modified Asphalt Producers, AMAP Newsletters and Bulletins,
February 2015, http://www.modifiedasphalt.org/news-bulletin-reob. (As of February 24, 2015).

Morian, N.E., M.Z. Alavi, E.Y. Hajj, and P.E. Sebaaly (2014) “Evolution of Thermo-viscoelastic
Properties of Asphalt Mixtures with Oxidative Aging.” Transportation Research Record No.
2447, Transportation Research Board, National Research Council, Washington, DC, pp. 1-12.

Nahar, S.N., J. Qiu, A.J.M. Schmets, E. Schlangen, M. Shirazi, M.F.C. van de Ven, G. Schitter,
and A. Scarpas (2014) “Turning Back Time: Rheological and Microstructural Assessment of
Rejuvenated Bitumen.” Transportation Research Record No. 2444, Transportation Research
Board, National Research Council, Washington, DC, pp. 52-62.

NCAT (2009) “Hot-Mix Asphalt Materials, Mixture Design and Construction.” NAPA Research
and Education Foundation, Lanham, MD, 720 pp.

NCAT (2014a) “How Should We Express RAP and RAS Contents?” Asphalt Technology News,
Vol. 26, No. 2 (Fall), http://www.eng.auburn.edu/research/centers/ncat/info-
pubs/newsletters/fall-2014/recycledcontents.html. (As of February 24, 2015).

NCAT (2014b) “NCAT Researchers Explore Multiple Uses of Rejuvenators.” Asphalt


Technology News, Vol. 26, No. 1 (Spring), http://www.ncat.us/info-pubs/newsletters/spring-
2014/rejuvenators.html. (As of February 24, 2015).

Newcomb, D.E., and J.A. Epps (1981) Asphalt Recycling Technology: Literature Review and
Research Plan. Report No. ESL-TR-81-42, Air Force Engineering and Services Center, Florida,
(June).

Newcomb, D.E., B.J. Nusser, B.M. Kiggundu, and D.M. Zallen (1984) “Laboratory Study of the
Effects of Recycling Modifiers on Aged Asphalt Cement.” Transportation Research Record No.
968, Transportation Research Board, National Research Council, Washington, DC, pp. 66-77.

126
Olard, F., and H. Benedetto (2003) “General ‘2S2P1D’ Model and Relation Between the Linear
Viscoelastic Behaviours of Bituminous Binders and Mixes.” Road Materials and Pavement
Design, Vol. 4, No. 2, pp. 185-224.

Oliver, J.W.H. (1975) Diffusion of Oils in Asphalts. Australian Road Research Board, Report No.
9, Vermont, Victoria, Australia 3733.

O’Sullivan, K.A. (2011) “Rejuvenation of Reclaimed Asphalt Pavement (RAP) in Hot Mix
Asphalt Recycling with High RAP Content.” M.Sc. thesis, Worcester Polytechnic Institute,
University of Worcester, MA (April), 45 pp.

Painter, S. (1993). “The Characterization of Asphalt and Asphalt Recyclability.” Washington,


DC: Strategic Highway Research Program Press. In: Wang, F. L., Long, J., Shen, B. X., & Ling,
H. (2014). A Study of the Regenerating Effects of RAs on Aged Asphalts. Petroleum Science
and Technology, Vol. 32, pp. 1160-1167.

Peterson, G.D., R.R. Davison, C.J. Glover, and J.A. Bullin (1994) “Effect of Composition on
Asphalt RA Performance.” Transportation Research Record No. 1436, Transportation Research
Board, National Research Council, Washington, DC, pp. 38-46.

Robinett, C.J., and J.A. Epps (2010) “Energy, Emissions, Material Conservation, and Prices
Associated with Construction, Rehabilitation, and Material Alternatives for Flexible Pavement.”
Transportation Research Record No. 2179, Transportation Research Board, National Research
Council, Washington, DC, pp. 10-22.

Romera R., A. Santamaría, J.J. Peña, M.E. Muñoz, M. Barral, E. García, and V. Jañez (2006)
“Rheological Aspects of the Rejuvenation of Aged Bitumen.” Rheologica Acta, Vol. 45, No. 4
(April), pp. 474-478.

Rose, A.A., E. Arambula, T. Howell, and C.J. Glover (2014) “An X-Ray CT Validated
Laboratory Measurement Method for Air Voids Distribution over Depth in Asphalt Pavement: A
Step toward Simplified Oxidation Modeling.” Petroleum Science and Technology, Vol. 32, No.
24, pp. 3020-3028.

Rostler, F.S., and R.M. White (1959) “Influence of Chemical Composition of Asphalts on
Performance, Particularly Durability.” American Society of Testing and Materials, Special
Technical Publication.

Rowe, G.M. (2011) “Prepared Discussion for the AAPT Paper by Anderson et al.: Evaluation of
the Relationship between Asphalt Binder Properties and Non-Load Related Cracking.” Journal
of the Association of Asphalt Paving Technologists, Vol. 80, pp. 649-662.

Ruan, Y., R.R. Davison, and C.J. Glover (2003) “An Investigation of Asphalt Durability:
Relationships between Ductility and Rheological Properties for Unmodified Asphalts.”
Petroleum Science and Technology, Vol. 21, pp. 231-254.

127
Ryan, M., D. Herlitzka, S. Engber, A. Engstler, S. Veglahn, A. Hanz, and G. Reinke (2014) “A
Discussion of Some Factors Impacting Performance of Binders Blended With Additives for
Reducing Low Temperature Properties of Asphalt Binders and Their Impact on Mix
Performance” FHWA Asphalt Mix Expert Task Group (ETG), Technical Presentation.

Shah, A., R.S. McDaniel, G.A. Huber, and V.L. Gallivan (2007). “Investigation of Properties of
Plant-Produced Reclaimed Asphalt Pavement Mixtures.” Transportation Research Record No.
1998, Transportation Research Board, National Research Council, Washington, DC, pp. 103-
111.

Shen, J., S. Amirkhanian, and S. Lee (2005) “The Effects of Rejuvenating Agents on Recycled
Aged CRM Binders.” The International Journal of Pavement Engineering, Vol. 6, No. 4, pp.
273-279 (December).

Shen, J., S. Amirkhanian, and S. Lee (2007) “HP-GPC Characterization of Rejuvenated Aged
CRM Binders.” Journal of Materials in Civil Engineering, Vol. 19, pp. 515-525.

Shen, J., S. Amirkhanian, and J.A. Miller (2007) “Effects of Rejuvenating Agents on Superpave
Mixtures Containing Reclaimed Asphalt Pavement.” Journal of Materials in Civil Engineering,
Vol. 19, No. 5, pp. 376-384 (May).

Shen, J., S. Amirkhanian, and B. Tang (2007) “Effects of Rejuvenator on Performance-Based


Properties of Rejuvenated Asphalt Binder and Mixtures.” Construction and Building Materials,
Vol. 21, No. 5, pp. 958-964.

Shen, J., B. Huang, and Y. Hachiya (2004) “Validation of Performance-Based Method for
Determining Rejuvenator Content in HMA.” The International Journal of Pavement
Engineering, Vol. 5, No. 2, pp. 103-109 (June).

Shen, J., and Y. Ohne (2002) “Determining Rejuvenator Content for Recycling Reclaimed
Asphalt Pavement by SHRP Binder Specifications.” The International Journal of Pavement
Engineering, Vol. 3, No. 4, pp. 261-268.

Soleymani, H.R., H.U. Bahia, and A.T. Bergan (1999) “Time-Temperature Dependency of
Blended and Rejuvenated Asphalt Binders.” 1999 Annual Meeting of the Association of Asphalt
Paving Technologists, Proceedings.

Tabatabaee, H.A. (2015) “Bio-based Rejuvenation of Highly Aged Asphalt: Rejuvenation vs.
Softening.” Presentation to TRB AFK20 Committee, Washington, DC (January).

Tran, N.H., A. Taylor, and R. Willis (2012) Effect Of Rejuvenator on Performance Properties of
HMA Mixtures with High RAP and RAS Contents. NCAT Report 12-05, National Center for
Asphalt Technology, Auburn, AL, 70 pp.

Underwood, B.S., Y.R. Kim, M. Guddati, S. Thirunavukkarasu, and S. Savadatti (2009)


“Response and Fatigue Performance Modeling of ALF Pavements Using 3-D Finite Element

128
Analysis and a Simplified Viscoelastic Continuum Damage Model.” Journal of the Association
of Asphalt Paving Technologists, Vol. 78, pp. 743-776.

Underwood, B.S., M.N. Guddati, and Y. R. Kim (2010) “Improved calculation method of
damage parameter in viscoelastic continuum damage model” International Journal of Pavement
Engineering, Vol. 11, No. 6, pp. 459-476.

Valdés, G., F. Pérez-Jiménez, R. Miró, A. Martínez, R. Lee, and R. Botella (2011)


“Experimental Study of Recycled Asphalt Mixtures with High Percentages of Reclaimed Asphalt
Pavement (RAP).” Construction and Building Materials, Vol. 25, (March), pp. 1289-1297.

Vallerga, B.A. (1981) “Pavement Deficiencies Related to Asphalt Durability.” Journal of the
Association of Asphalt Paving Technologists, Vol. 50, pp. 481-491.

Wang, F.L., J. Long, B.X. Shen, and H. Ling (2014) “A Study of the Regenerating Effects of
RAs on Aged Asphalts.” Petroleum Science and Technology, Vol. 32, pp. 1160-1167.

West, R., A. Kvasnak, N. Tran, B. Powell, and P. Turner (2009) “Testing of Moderate and High
RAP Content Mixes: Laboratory and Accelerated Field Performance at the National Center for
Asphalt Technology Test Track.” Transportation Research Record No. 2126, Transportation
Research Board, National Research Council, Washington, DC, pp. 100-108.

West, R., Willis, J.R., and M. Marasteanu (2013) NCHRP 9-46: Improved Mix Design,
Evaluation, and Materials Management Practices for Hot Mix Asphalt with High Reclaimed
Asphalt Pavement Content, NCHRP Report 752, Transportation Research Board, National
Research Council, Washington, DC.

Xiao, F., S. Amirkhanian, and C. Juang (2007) “Rutting Resistance of Rubberized Asphalt
Concrete Pavements Containing Reclaimed Asphalt Pavement Mixtures.” Journal of Materials
in Civil Engineering, pp. 475-483 (June).

Yan, J., Z. Zhang, H. Zhu, F. Li, and Q. Liu (2014) “Experimental Study of Hot Recycled
Asphalt Mixtures with High Percentages of Reclaimed Asphalt Pavement and Different RAs.”
ASTM Journal of Testing and Evaluation, Vol. 42, No. 5 (September).

Yin, F., E. Arambula, R. Lytton, A. Epps Martin, and L. Garcia Cucalon (2014) “A Novel
Method to Evaluate Moisture Susceptibility and Rutting Resistance of Asphalt Mixtures Using
the Hamburg Wheel Tracking Test.” Transportation Research Record, Transportation Research
Board, National Research Council, Washington, DC, in press.

Yin, F., A. Epps Martin, E. Arambula, and D.E. Newcomb (2015) “Short-Term Aging of Asphalt
Mixtures.” Journal of the Association of Asphalt Paving Technologists, Vol. 84, in press.

Yin, F., L. Garcia Cucalon, A. Epps Martin, E. Arámbula, A. Chowdhury, and E.S. Park (2013)
“Laboratory Conditioning Protocols for Warm-Mix Asphalt.” Journal of the Association of
Asphalt Paving Technologists, Vol. 82, pp. 177-211.

129
Yin, F., L. Garcia Cucalon, A. Epps Martin, E. Arámbula, and E.S. Park (2014) “Performance
Evolution of Hot-Mix and Warm-Mix Asphalt with Field and Laboratory Aging.” Journal of the
Association of Asphalt Paving Technologists, Vol. 83, pp. 109-142.

Yu, X., M. Zaumanis, S. Santos, L.D. Poulikakos (2014) “Rheological, Microscopic, and
Chemical Characterization of the Rejuvenating Effect on Asphalt Binders.” Fuel, Vol. 135, pp.
162-171.

Zaumanis, M., R. Mallick, and R. Frank (2013) “Evaluation of Rejuvenator’s Effectiveness with
Conventional Mix Testing for 100% Reclaimed Asphalt Pavement Mixtures.” Transportation
Research Record No. 2370, Transportation Research Board, National Research Council,
Washington, DC, pp. 17-25.

Zaumanis, M., R. Mallick, and R. Frank (2014a) “Determining Optimum Rejuvenator Dose for
Asphalt Recycling Based on Superpave Performance Grade Specifications.” Construction and
Building Materials, Vol. 69, pp. 159-166.

Zaumanis, M., R. Mallick, and R. Frank (2014b) “Evaluation of Different RAs for Restoring
Aged Asphalt Binder and Performance of 100 % Recycled Asphalt.” Materials and Structures,
14 p (May).

Zhao, S., B. Huang, X. Shu, X. Jia, and M. Woods (2012). “Laboratory Performance Evaluation
of Warm-Mix Asphalt Containing High Percentages of Reclaimed Asphalt Pavement.”
Transportation Research Record No. 2294, Transportation Research Board, Washington, DC,
pp. 98-105.

Zhou, F., J.W. Button, and J.A. Epps (2012) Best Practice for Using RAS in HMA. Report No.
FHWA/TX-12/0-6614-1, Texas Transportation Institute, College Station, TX, 72 pp.

Zhou, F., L. Hongsheng, S. Hu, J. Button, and J. Epps (2013) Characterization and Best Use of
Recycled Asphalt Shingles in Hot-Mix Asphalt. Report No. FHWA/TX-13/0-6614-2, Texas
Transportation Institute, College Station, TX (July), 88 pp.

Zhou, F., S. Hu, G. Das, and T. Scullion (2011) High RAP Mixes Design Methodology with
Balanced Performance. Report No. FHWA/TX-11/0-6092-2, Texas A&M Transportation
Institute, College Station, TX, 35 p.

Zhou, F., S. Im, D. Morton, R. Lee, S. Hu, and T. Scullion (2015) “Rejuvenator Characterization,
Blend Characteristics, and Proposed Mix Design Method.” Journal of the Association of Asphalt
Paving Technologists, accepted for publication.

130
LIST OF ACRONYMS

AASHTO American Association of State Highway and Transportation Officials


AFM Atomic Force Microscopy
ALMLC Aged Lab Mixed Lab Compacted
AMPT Asphalt Mixture Performance Tester
ANCOVA Analysis of Covariance
ANOVA Analysis of Variance
APA Asphalt Pavement Analyzer
ARPMLC Aged Reheated Plant Mixed, Laboratory Compacted
ASTM American Society for Testing and Materials
AV Air Voids
BBR Bending Beam Rheometer
CA Carbonyl Area
CDD Cumulative Degree Days
CRM Crumb Rubber Modified
DLV Dimensionless Log Viscosity
DOT Department of Transportation
DS Dynamic Stability
DSR Dynamic Shear Rheometer
DTT Direct Tension Test
EBM Energy-Based Mechanistic
E* Dynamic Complex Modulus
FHWA Federal Highway Administration
FT Fischer-Tropsch
FTIR Fourier Transform Infrared
G* Shear Complex Modulus
G-R Glover-Rowe Parameter
HMA Hot-Mix Asphalt
HP-GPC High-Pressure Gel Permeation Chromatography
HWTT Hamburg Wheel Tracking Test
IDT Indirect Tensile Strength
LAS Linear Amplitude Sweep
LMLC Lab Mixed Lab Compacted
LMS Large Molecular Size
LSV Low Shear Viscosity
LTOA Long-Term Oven Aging
LTPP Long-Term Pavement Performance Program
MEPDG Mechanistic-Empirical Pavement Design Guide
MMS Medium Molecular Size
MR Resilient Modulus
MSCR Multiple Stress Creep Recovery
MWAS Manufacturer Waste Asphalt Shingles
NCAT National Center for Asphalt Technology
NCHRP National Cooperative Highway Research Program
OT Overlay Test

131
PAV Pressure Aging Vessel
PG Performance Grade
PGH High-Temperature PG
PGI Intermediate-Temperature PG
PGL Low-Temperature PG
PI Penetration Index
PMLC Plant Mixed, Laboratory Compacted
POV Pressure Oxidation Vessel
PVN Penetration-Viscosity Number
RA RA
RAP Reclaimed Asphalt Pavement
RAPBR RAP Binder Ratio
RAS Recycled Asphalt Shingle
RASBR RAS Binder Ratio
RBR Recycled Binder Ratio
RPMLC Reheated Plant Mix Lab Compacted
RTFO Rolling Thin Film Oven
S Asphalt Binder Stiffness at Low Temperatures
SARA Saturate, Aromatic, Resin, and Asphaltene Analysis
SHRP Strategic Highway Research Program
SMS Small Molecular Size
STOA Short-Term Oven Aging
S-VECD Simplified Viscoelastic Continuum Damage
TOAS Tear-Off Asphalt Shingles
TSR Tensile Strength Ratio
TSRST Thermal Stress Restrained Specimen Test
TTI Texas A&M Transportation Institute
TxDOT Texas Department of Transportation
ULMLC Unaged Lab Mixed Lab Compacted
UTSST Uniaxial Thermal Stress and Strain Test
WMA Warm-Mix Asphalt
X-Ray CT X-Ray Computed Tomography
δ Binder Phase Angle
 Phase Angle
c Crossover Frequency

132
APPENDIX A. SURVEYS

NCHRP 9-58: The Effects of RAs on Asphalt Mixtures with High RAS and RAP Binder
Ratios

A.1 State DOT Web-Based Survey

Email-mail Introduction
The Texas A&M Transportation Institute is conducting National Cooperative Highway
Research Program (NCHRP) Project 9-58 The Effects of RAs on Asphalt Mixtures with High RAS
and RAP Binder Ratios. This study includes a web-based survey (http://tti.tamu.edu/state-dot-
survey/) to assess the current state of practice on the use of RAs in surface asphalt mixtures with
reclaimed asphalt pavement (RAP) and/or recycled asphalt shingles (RAS) at high recycled
binder ratios.
The survey is being sent to state departments of transportation. We estimate that the
survey will take approximately 15-30 minutes to complete. We would appreciate your response
by November 14, 2014. If you feel you are not the appropriate person to complete this survey,
please send alternate contact information to the principal investigator Amy Epps Martin (a-
eppsmartin@tamu.edu). Any questions or comments about the survey can also be directed to
her. For questions about your rights as a research participant; or if you have questions,
complaints, or concerns about the research, you may contact the Texas A&M University Human
Subjects Protection Program at 979.458.4067, toll-free at 1.855.795.8636, or email at
irb@tamu.edu.
Your participation is entirely voluntary, and you will receive no direct benefit from
completing the survey. Responses will be confidential, but results will only be released in
aggregate form and possibly by organization/company without identification of specific
individual respondents. Questions marked with an asterisk (*) are necessary to make your survey
usable. Please use the survey link above to start the survey if you consent to participate in this
study. If at any point you change your mind and decide not to participate, you can simply close
the browser. Only complete responses will be recorded.
Thank you in advance for your participation. Your response will help formulate the
laboratory and field experiments for this study that will ultimately provide guidance to pavement
professionals and practitioners.
Web-Based Survey Introduction
You are invited to complete an online survey as part of National Cooperative Highway
Research Program (NCHRP) Project 9-58 The Effects of RAs on Asphalt Mixtures with High RAS
and RAP Binder Ratios. The purpose of this web-based survey is to assess the current state of
practice on the use of RAs in surface asphalt mixtures with high reclaimed asphalt pavement
(RAP) and/or recycled asphalt shingles (RAS) at high recycled binder ratios. Your response will
help formulate the laboratory and field experiments for this study. A follow-up phone interview
may be necessary to clarify some responses to the survey.
Your participation is entirely voluntary, and you will receive no direct benefit from
completing the survey. We would appreciate your response by November 14, 2014. Responses
will be confidential, but results will only be released in aggregate form and possibly by
organization/company without identification of specific individual respondents. Questions
marked with an asterisk (*) are necessary to make your survey usable. Please press NEXT

A-1
SCREEN if you consent to participate in this study by completing the survey. If at any point you
change your mind and decide not to participate, you can simply close the browser. Only
complete responses will be recorded.
Contact Info
Please provide your contact information.
*Name:
*Organization:
*Address:
*City: *State: *Zip:
*Phone Number:
*Email Address:

Use of High Recycled Binder Ratios


Recycled binder ratio in this section is defined as the ratio of the RAP and/or RAS binder content
in the mixture divided by the mixture’s total binder content, analogous to RAP binder ratio
defined in NCHRP Report 752. Assuming typical values of total binder content, RAP binder
content, and RAS binder content of 5%, 5%, and 20%, respectively; a high recycled binder ratio
in the range of 0.3-0.5 can be produced in mixtures with: (a) 30% RAP, (b) 8% RAS, or (c) 10%
RAP and 5% RAS as an example.
1. What is your predominant use of RAP and/or RAS in surface asphalt mixtures?
o Predominant use of RAP [please list typical percentage]
o Predominant use of RAS [please list typical percentage]
o Predominant use of RAP and RAS [please list typical percentages]

2. What is your current practice with regard to using RAP in surface asphalt mixtures?
o Not allowed by current specification [please provide reason(s) for restriction]
o Required [please specify maximum percentage by weight of mix]
o Allowed as an option [please specify maximum percentage by weight of mix]
o Allowed as a separate bid item [please specify maximum percentage by weight of
mix]
o Used previously [please provide reason(s) for discontinuation of use]
o Don’t know

3. What is your typical RAP content in surface asphalt mixtures (with no RAS)?
o Do not use
o 0-10 % by weight of mix
o 11-20% by weight of mix
o 21-30% by weight of mix
o >30% by weight of mix
o Don’t know

4. What are the main barriers, if any, to using higher levels of RAP in surface asphalt
mixtures? (Check all that apply)
□ No significant barriers

A-2
□ Current specification limits [please describe briefly]
□ Lack of a mix design procedure
□ Problems meeting mix design criteria
□ RAP variability
□ RAP availability
□ RAP management difficulties (e.g., segregation, stockpiling, crushing, fractionation,
etc.) [please list]
□ Mix production difficulties (e.g., moisture content, changes in plant operation,
workability, etc.) [please list]
□ Poor pavement performance (i.e., premature distress or failure) [please describe
briefly]
□ Others [please describe briefly]

5. What is your current practice with regard to using RAS in surface asphalt mixtures?
o Not allowed by current specifications [please provide reason(s) for restriction]
o Required [please specify maximum percentage by weight of mix]
o Allowed as an option [please specify maximum percentage by weight of mix]
o Allowed as a separate bid item [please specify maximum percentage by weight of
mix]
o Used previously [please provide reason(s) for discontinuation of use]
o Don’t know

6. What is your typical RAS content in surface asphalt mixtures (with no RAP)?
o Do not use
o 0-3 % by weight of mix
o 4-6% by weight of mix
o 7-9% by weight of mix
o >9% by weight of mix
o Don’t know

7. What are the main barriers, if any, to using higher levels of RAS in surface asphalt
mixtures? (Check all that apply)
□ No significant barriers
□ Current specification limits [please describe briefly]
□ Lack of a mix design procedure
□ Problems meeting mix design criteria
□ RAS variability
□ RAS availability
□ RAS management difficulties (e.g., segregation, stockpiling, crushing, fractionation,
etc.) [please list]
□ Mix production difficulties (e.g., moisture content, changes in plant operation,
workability, etc.) [please list]
□ Poor pavement performance (i.e., premature distress or failure) [please describe
briefly]
□ Others [please describe briefly]

A-3
8. Do you have a requirement for recycled binder ratio for surface asphalt mixtures with
RAP and/or RAS?
o Yes [please specify minimum and/or maximum values]
o No
o Don’t know

9. Do you have standard test methods or procedures for characterizing RAP and/or RAS?
o Yes [please list standard(s) by name/#]
o No
o Don’t know

10. Do you have standard test methods or procedures for designing surface asphalt mixtures
with RAP and/or RAS?
o Yes [please list standard(s) by name/#]
o No
o Don’t know

11. What performance tests, if any, are required for surface asphalt mixtures with RAP
and/or RAS?
(Check all that apply)
o No performance tests are required
o Hamburg Wheel Tracking Test (HWTT)
o Asphalt Pavement Analyzer (APA)
o Flow Number Test
o Simplified Viscoelastic Continuum Damage (S-VECD)
o Overlay Cracking Test
o Thermal Stress Restrained Specimen Test (TSRST)
o Indirect Tensile (IDT) / Tensile Strength Ratio (TSR)
o Dynamic Modulus
o Resilient Modulus
o Creep Compliance
o Others [please list]
o Don’t know

12. What premature distresses or failures, if any, have you observed for surface asphalt
mixtures with RAP and/or RAS at high recycled binder ratios?
(Check all that apply)
□ No premature distress or failure has been observed to date
□ Fatigue cracking
□ Thermal cracking
□ Reflective cracking
□ Other distress [please identify]
□ Don’t know

A-4
Use of RAs
RA in this section is defined as an additive that is incorporated in the surface asphalt mixture to
soften or rejuvenate the RAP and/or RAS materials.

13. What is your current practice with regard to using RAs in surface asphalt mixtures with
RAP and/or RAS at high recycled binder ratios?
o Do not use / Not allowed by current specification [please provide reason(s) for
restriction]
o Require [please specify how dosage rate is determined]
o Allow [please specify how dosage rate is determined]
o Used previously [please provide reason(s) for discontinuation of use]
o Don’t know

14. What RAs to you use in surface asphalt mixtures with RAP and/or RAS at high recycled
binder ratios?
(Check all that apply)
□ Do not use
□ BITUTECH RAP [please provide typical dosage rate]
□ CPR™ [please provide typical dosage rate]
□ CYCLOGEN® [please provide typical dosage rate]
□ Ergon HyPrene® [please provide typical dosage rate]
□ Hydrogreen® [please provide typical dosage rate]
□ Hydrolene® [please provide typical dosage rate]
□ Reclamite® [please provide typical dosage rate]
□ RejuvaSeal [please provide typical dosage rate]
□ SonneWarmix RJ™ [please provide typical dosage rate]
□ Storbit® [please provide typical dosage rate]
□ SYLVAROAD™ RP1000 [please provide typical dosage rate]
□ ValAro 130A® [please provide typical dosage rate]
□ Valero VP 165® [please provide typical dosage rate]
□ Others [please list and include typical dosage rate]
□ Don’t know

15. What are the main barriers, if any, to utilizing RAs in surface asphalt mixtures with RAP
and/or RAS at high recycled binder ratios?
(Check all that apply)
□ No significant barriers
□ Current specification limits [please describe briefly]
□ Lack of tests and criteria to determine dosage rate and/or performance
□ RA availability
□ Lack of expertise in using RAs
□ Poor performance (i.e., premature distress or failure) [please describe briefly]
□ Others [please describe briefly]

A-5
16. Do you have standard test methods or procedures for characterizing RAs?
o Yes [please list standard(s) by name/#]
o No
o Don’t know

17. Do you have standard test methods or procedures for designing surface asphalt mixtures
with RAP and/or RAS at high recycled binder ratios and RAs (including determining the
dosage rate)?
o Yes [please list standard(s) by name/#]
o No
o Don’t know

18. What laboratory tests, if any, are required for characterizing the properties of binders
containing RAs?
(Check all that apply)
□ No laboratory tests are required
□ G*, by Dynamic Shear Rheometer (DSR)
□ S, m-value by Bending Beam Rheometer (BBR)
□ Multiple Stress Creep Recovery (MSCR) by Dynamic Shear Rheometer (DSR)
□ Linear Amplitude Sweep (LAS) by Dynamic Shear Rheometer (DSR)
□ Others [please describe briefly]
□ Don’t know

19. What performance tests, if any, are required for surface asphalt mixtures with RAP
and/or RAS at high recycled binder ratios and RAs?
(Check all that apply)
o No performance tests are required
o Hamburg Wheel Tracking Test (HWTT)
o Asphalt Pavement Analyzer (APA)
o Flow Number Test
o Simplified Viscoelastic Continuum Damage (S-VECD)
o Overlay Cracking Test
o Thermal Stress Restrained Specimen Test (TSRST)
o Indirect Tensile (IDT) / Tensile Strength Ratio (TSR)
o Dynamic Modulus
o Resilient Modulus
o Creep Compliance
o Others [please describe briefly]
o Don’t know

Other

20. Do you have upcoming field projects using RAs in surface asphalt mixtures with RAP
and/or RAS at high recycled binder ratios OR surface asphalt mixtures with high recycled
binder ratios that exhibited premature distress or failure? AND Are you willing to
participate in NCHRP Project 9-58 by sharing information about mix design,

A-6
construction, materials, and/or performance monitoring? Please note that follow-up email
and/or phone contact will be necessary.
o Yes, willing to participate
o Yes, but not able to participate
o No
o Don’t know

Additional Comments

Please provide any additional comments, experiences, or challenges you wish to share
about using RAs in surface asphalt mixtures with RAP and/or RAS at high recycled
binder ratios. We sincerely appreciate your time in helping us complete this research
project that will ultimately provide guidance useful to you as a pavement professional.

A-7
NCHRP 9-58: The Effects of RAs on Asphalt Mixtures with High RAS and RAP Binder
Ratios

A.2 Contractor Web-Based Survey

E-mail Introduction
The Texas A&M Transportation Institute is conducting National Cooperative Highway
Research Program (NCHRP) Project 9-58 The Effects of RAs on Asphalt Mixtures with High RAS
and RAP Binder Ratios. This study includes a web-based survey (http://tti.tamu.edu/contractor-
survey/) to assess the current state of practice on the use of RAs in surface asphalt mixtures with
reclaimed asphalt pavement (RAP) and/or recycled asphalt shingles (RAS) at high recycled
binder ratios.
The survey is being sent to asphalt paving contractors. We estimate that the survey will
take approximately 15-30 minutes to complete. We would appreciate your response by
November 14, 2014. If you feel you are not the appropriate person to complete this survey,
please send alternate contact information to the principal investigator Amy Epps Martin (a-
eppsmartin@tamu.edu). Any questions or comments about the survey can also be directed to her.
For questions about your rights as a research participant; or if you have questions, complaints, or
concerns about the research, you may contact the Texas A&M University Human Subjects
Protection Program at 979.458.4067, toll-free at 1.855.795.8636, or email at irb@tamu.edu.
Your participation is entirely voluntary, and you will receive no direct benefit from
completing the survey. Responses will be confidential, but results will only be released in
aggregate form and possibly by organization/company without identification of specific
individual respondents. Questions marked with an asterisk (*) are necessary to make your survey
usable. Please use the survey link above to start the survey if you consent to participate in this
study. If at any point you change your mind and decide not to participate, you can simply close
the browser. Only complete responses will be recorded.
Thank you in advance for your participation. Your response will help formulate the
laboratory and field experiments for this study that will ultimately provide guidance to pavement
professionals and practitioners.
Web-Based Survey Introduction
You are invited to complete an online survey as part of National Cooperative Highway
Research Program (NCHRP) Project 9-58 The Effects of RAs on Asphalt Mixtures with High RAS
and RAP Binder Ratios. The purpose of this web-based survey is to assess the current state of
practice on the use of RAs in surface asphalt mixtures with high reclaimed asphalt pavement
(RAP) and/or recycled asphalt shingles (RAS) at high recycled binder ratios. Your response will
help formulate the laboratory and field experiments for this study. A follow-up phone interview
may be necessary to clarify some responses to the survey.
Your participation is entirely voluntary, and you will receive no direct benefit from
completing the survey. We would appreciate your response by November 14, 2014. Responses
will be confidential, but results will only be released in aggregate form and possibly by
organization/company without identification of specific individual respondents. Questions
marked with an asterisk (*) are necessary to make your survey usable. Please press NEXT
SCREEN if you consent to participate in this study by completing the survey. If at any point you
change your mind and decide not to participate, you can simply close the browser. Only
complete responses will be recorded.
Contact Info

A-8
Please provide your contact information.
*Name:
*Company:
*Address:
*City: *State: *Zip:
*Phone Number:
*Email Address:

Use of High Recycled Binder Ratios


Recycled binder ratio in this section is defined as the ratio of the RAP and/or RAS binder content
in the mixture divided by the mixture’s total binder content, analogous to RAP binder ratio
defined in NCHRP Report 752. Assuming typical values of total binder content, RAP binder
content, and RAS binder content of 5%, 5%, and 20%, respectively; a high recycled binder ratio
in the range of 0.3-0.5 can be produced in mixtures with: (a) 30% RAP, (b) 8% RAS, or (c) 10%
RAP and 5% RAS as an example.
1. What is your predominant use of RAP and/or RAS in surface asphalt mixtures?
o Predominant use of RAP [please list typical percentage]
o Predominant use of RAS [please list typical percentage]
o Predominant use of RAP and RAS [please list typical percentages]

2. What is your current practice with regard to using RAP in surface asphalt mixtures?
o Not allowed by current specifications [please provide reason(s) for restriction]
o Required [please specify maximum percentage by weight of mix]
o Allowed as an option [please specify maximum percentage by weight of mix]
o Allowed as a separate bid item [please specify maximum percentage by weight of
mix]
o Used previously [please provide reason(s) for discontinuation of use]
o Don’t know

3. What is your typical RAP content in surface asphalt mixtures (with no RAS)?
o Do not use
o 0-10 % by weight of mix
o 11-20% by weight of mix
o 21-30% by weight of mix
o >30% by weight of mix
o Don’t know

4. What are the main barriers, if any, to using higher levels of RAP in surface asphalt
mixtures? (Check all that apply)
□ No significant barriers
□ Current specification limits [please describe briefly]
□ Lack of a mix design procedure
□ Problems meeting mix design criteria

A-9
□ RAP variability
□ RAP availability
□ RAP management difficulties (e.g., segregation, stockpiling, crushing, fractionation,
etc.) [please list]
□ Mix production difficulties (e.g., moisture content, changes in plant operation,
workability, etc.) [please list]
□ Poor pavement performance (i.e., premature distress or failure) [please describe
briefly]
□ Others [please describe briefly]

5. What is your current practice with regard to using RAS in surface asphalt mixtures?
o Not allowed by current specifications [please provide reason(s) for restriction]
o Required [please specify maximum percentage by weight of mix]
o Allowed as an option [please specify maximum percentage by weight of mix]
o Allowed as a separate bid item [please specify maximum percentage by weight of
mix]
o Used previously [please provide reason(s) for discontinuation of use]
o Don’t know

6. What is your typical RAS content in surface asphalt mixtures (with no RAP)?
o Do not use
o 0-3 % by weight of mix
o 4-6% by weight of mix
o 7-9% by weight of mix
o >9% by weight of mix
o Don’t know

7. What are the main barriers, if any, to using higher levels of RAS in surface asphalt
mixtures? (Check all that apply)
□ No significant barriers
□ Current specification limits [please describe briefly]
□ Lack of a mix design procedure
□ Problems meeting mix design criteria
□ RAS variability
□ RAS availability
□ RAS management difficulties (e.g., segregation, stockpiling, crushing, fractionation,
etc.) [please list]
□ Mix production difficulties (e.g., moisture content, changes in plant operation,
workability, etc.) [please list]
□ Poor pavement performance (i.e., premature distress or failure) [please describe
briefly]
□ Others [please describe briefly]

8. Are you required to meet a minimum and/or maximum recycled binder ratio for surface
asphalt mixtures with RAP and/or RAS?

A-10
o Yes [please specify minimum and/or maximum values]
o No
o Don’t know

9. Are you required to follow standard test methods or procedures for characterizing RAP
and/or RAS?
o Yes [please list standard(s) by name/#]
o No
o Don’t know

10. Are you required to follow standard test methods or procedures for designing surface
asphalt mixtures with RAP and/or RAS?
o Yes [please list standard(s) by name/#]
o No
o Don’t know

11. What performance tests, if any, are required for surface asphalt mixtures with RAP
and/or RAS?
(Check all that apply)
o No performance tests are required
o Hamburg Wheel Tracking Test (HWTT)
o Asphalt Pavement Analyzer (APA)
o Flow Number Test
o Simplified Viscoelastic Continuum Damage (S-VECD)
o Overlay Cracking Test
o Thermal Stress Restrained Specimen Test (TSRST)
o Indirect Tensile (IDT) / Tensile Strength Ratio (TSR)
o Dynamic Modulus
o Resilient Modulus
o Creep Compliance
o Others [please describe briefly]
o Don’t know

12. What premature distresses or failures, if any, have you observed for surface asphalt
mixtures with RAP and/or RAS at high recycled binder ratios?
(Check all that apply)
□ No premature distress or failure has been observed to date
□ Fatigue cracking
□ Thermal cracking
□ Reflective cracking
□ Other distress [please list]
□ Don’t know

Use of RAs

A-11
RA in this section is defined as an additive that is incorporated in the surface asphalt mixture to
soften or rejuvenate the RAP and/or RAS materials.

13. What is your current practice with regard to using RAs in surface asphalt mixtures with
RAP and/or RAS at high recycled binder ratios?
o Do not use / Not allowed by current specification [please provide reason(s) for
restriction]
o Required [please specify how dosage rate is determined]
o Allowed [please specify how dosage rate is determined]
o Used previously [please provide reason(s) for discontinuation of use]
o Don’t know

14. What RAs to you use in surface asphalt mixtures with RAP and/or RAS at high recycled
binder ratios?
(Check all that apply)
□ Do not use
□ BITUTECH RAP [please provide typical dosage rate]
□ CPR™ [please provide typical dosage rate]
□ CYCLOGEN® [please provide typical dosage rate]
□ Ergon HyPrene® [please provide typical dosage rate]
□ Hydrogreen® [please provide typical dosage rate]
□ Hydrolene® [please provide typical dosage rate]
□ Reclamite® [please provide typical dosage rate]
□ RejuvaSeal [please provide typical dosage rate]
□ SonneWarmix RJ™ [please provide typical dosage rate]
□ Storbit® [please provide typical dosage rate]
□ SYLVAROAD™ RP1000 [please provide typical dosage rate]
□ ValAro 130A® [please provide typical dosage rate]
□ Valero VP 165® [please provide typical dosage rate]
□ Others [please list and include typical dosage rate]
□ Don’t know

15. What are the main barriers, if any, to utilizing RAs in surface asphalt mixtures with RAP
and/or RAS at high recycled binder ratios?
(Check all that apply)
□ No significant barriers
□ Current specification limits [please describe briefly]
□ Lack of tests and criteria to determine dosage rate and/or performance
□ RA availability
□ Lack of expertise in using RAs
□ Poor performance (i.e., premature distress or failure) [please describe briefly]
□ Others [please describe briefly]

16. Are you required to follow standard test methods or procedures for characterizing RAs?
o Yes [please list standard(s) by name/#]
o No

A-12
o Don’t know

17. Are you required to follow standard test methods or procedures for designing surface
asphalt mixtures with RAP and/or RAS at high recycled binder ratios and RAs (including
determining the dosage rate)?
o Yes [please list standard(s) by name/#]
o No
o Don’t know

18. What laboratory tests, if any, are required for characterizing the properties of binders
containing RAs?
(Check all that apply)
□ No laboratory tests are required
□ G*, by Dynamic Shear Rheometer (DSR)
□ S, m-value by Bending Beam Rheometer (BBR)
□ Multiple Stress Creep Recovery (MSCR) by Dynamic Shear Rheometer (DSR)
□ Linear Amplitude Sweep (LAS) by Dynamic Shear Rheometer (DSR)
□ Others [please describe briefly]
□ Don’t know

19. What performance tests, if any, are required for surface asphalt mixtures with RAP
and/or RAS at high recycled binder ratios and RAs?
(Check all that apply)
o No performance tests are required
o Hamburg Wheel Tracking Test (HWTT)
o Asphalt Pavement Analyzer (APA)
o Flow Number Test
o Simplified Viscoelastic Continuum Damage (S-VECD)
o Overlay Cracking Test
o Thermal Stress Restrained Specimen Test (TSRST)
o Indirect Tensile (IDT) / Tensile Strength Ratio (TSR)
o Dynamic Modulus
o Resilient Modulus
o Creep Compliance
o Others [please describe briefly]
o Don’t know

Other

20. Do you have upcoming field projects using RAs in surface asphalt mixtures with RAP
and/or RAS at high recycled binder ratios OR surface asphalt mixtures with high recycled
binder ratios that exhibited premature distress or failure? AND Are you willing to
participate in NCHRP Project 9-58 by sharing information about mix design,
construction, materials, and/or performance monitoring? Please note that follow-up email
and/or phone contact will be necessary.

A-13
o Yes, willing to participate
o Yes, but not able to participate
o No
o Don’t know

Additional Comments

Please provide any additional comments, experiences, or challenges you wish to share about
using RAs in surface asphalt mixtures with RAP and/or RAS at high recycled binder ratios. We
sincerely appreciate your time in helping us complete this research project that will ultimately
provide guidance useful to you as a pavement professional.

A-14
NCHRP 9-58: The Effects of RAs on Asphalt Mixtures with High RAS and RAP Binder
Ratios

A.3 RAs Supplier Web-Based Survey

Email-mail Introduction
The Texas A&M Transportation Institute is conducting National Cooperative Highway
Research Program (NCHRP) Project 9-58, The Effects of RAs on Asphalt Mixtures with High
RAS and RAP Binder Ratios. This study includes a web-based survey
(http://tti.tamu.edu/recycling-agents-suppliers-survey/) to assess the current state of practice on
the use of RAs in surface asphalt mixtures with reclaimed asphalt pavement (RAP) and/or
recycled asphalt shingles (RAS) at high recycled binder ratios.
The survey is being sent to RA suppliers. We estimate that the survey will take
approximately 15-30 minutes to complete. We would appreciate your response by November 14,
2014. If you feel you are not the appropriate person to complete this survey, please send alternate
contact information to the principal investigator Amy Epps Martin (a-eppsmartin@tamu.edu).
Any questions or comments about the survey can also be directed to her. For questions about
your rights as a research participant; or if you have questions, complaints, or concerns about the
research, you may contact the Texas A&M University Human Subjects Protection Program at
979.458.4067, toll-free at 1.855.795.8636, or email at irb@tamu.edu.
Your participation is entirely voluntary, and you will receive no direct benefit from
completing the survey. We would appreciate your response by November 14, 2014. Responses
will be confidential, but results will only be released in aggregate form and possibly by
organization/company without identification of specific individual respondents. Questions
marked with an asterisk (*) are necessary to make your survey usable. Please use the survey link
above to start the survey if you consent to participate in this study. If at any point you change
your mind and decide not to participate, you can simply close the browser. Only complete
responses will be recorded.
Thank you in advance for your participation. Your response will help formulate the
laboratory and field experiments for this study that will ultimately provide guidance to pavement
professionals and practitioners.
Web-Based Survey Introduction
You are invited to complete an online survey as part of National Cooperative Highway
Research Program (NCHRP) Project 9-58 The Effects of RAs on Asphalt Mixtures with High RAS
and RAP Binder Ratios. The purpose of this web-based survey is to assess the current state of
practice on the use of RAs in surface asphalt mixtures with high reclaimed asphalt pavement
(RAP) and/or recycled asphalt shingles (RAS) at high recycled binder ratios. Recycled binder
ratio is defined as the ratio of the RAP and/or RAS binder content in the mixture divided by the
mixture’s total binder content, analogous to RAP binder ratio defined in NCHRP Report 752.
Assuming typical values of total binder content, RAP binder content, and RAS binder content of
5%, 5%, and 20%, respectively; a high recycled binder ratio in the range of 0.3-0.5 can be
produced in mixtures with: (a) 30% RAP, (b) 8% RAS, or (c) 10% RAP and 5% RAS as an
example.
Your response will help formulate the laboratory and field experiments for this study. A
follow-up phone interview may be necessary to clarify some responses to the survey.

A-15
Your participation is entirely voluntary, and you will receive no direct benefit from
completing the survey. We would appreciate your response by November 14, 2014. Responses
will be confidential, but results will only be released in aggregate form and possibly by
organization/company without identification of specific individual respondents. Questions
marked with an asterisk (*) are necessary to make your survey usable. Please press NEXT
SCREEN if you consent to participate in this study by completing the survey. If at any point you
change your mind and decide not to participate, you can simply close the browser. Only
complete responses will be recorded.
Contact Info
Please provide your contact information.
*Name:
*Organization:
*Address:
*City: *State: *Zip:
*Phone Number:
*Email Address:

RAs
RA in this section is defined as an additive that is incorporated in the surface asphalt mixture to
soften or rejuvenate the RAP and/or RAS materials.

1. What types of RAs do you produce?


(Check all that apply)
□ Paraffinic Oils [please specify brand name and typical dosage rate]
□ Aromatic Extracts [please specify brand name and typical dosage rate]
□ Napthenic Oils[please specify brand name and typical dosage rate]
□ Tryglycerides & Fatty Acids[please specify brand name and typical dosage rate]
□ Tall Oils[please specify brand name and typical dosage rate]
□ Other [please specify brand name and typical dosage rate]
□ Don’t know

2. Do you have standard test methods or procedures for characterizing RAs?


o Yes [please list standard(s) by name/#]
o No
o Don’t know

3. What is the blending protocol followed during plant/field operations for your RA(s)?
□ Blending with virgin asphalt binder [please describe briefly]
□ Blending with RAP and/or RAS [please describe briefly]
□ Blending with virgin aggregate [please describe briefly]
□ Other [please describe briefly]
□ Don’t know

A-16
4. What is the blending protocol followed during laboratory mixture production for your
RA(s)?
□ Blending with virgin asphalt binder [please describe briefly]
□ Blending with RAP and/or RAS [please describe briefly]
□ Blending with virgin aggregate [please describe briefly]
□ Other [please describe briefly]
□ Don’t know

Other

5. Are you participating in any upcoming field projects using RAs in surface asphalt
mixtures with RAP and/or RAS at high recycled binder ratios OR do you know of any
surface asphalt mixtures with high recycled binder ratios and RAs that exhibited
premature distress or failure? AND Are you willing to participate in NCHRP Project 9-58
by sharing information on mix design, construction, materials, and/or performance
monitoring for surface asphalt mixtures using your RAs? Please note that follow-up email
and/or phone contact will be necessary.
o Yes, willing to participate
o Yes, but not able to participate
o No
o Don’t know

Additional Comments

Please provide any additional comments, experiences, or challenges you wish to share
about using RAs in surface asphalt mixtures with RAP and/or RAS at high recycled
binder ratios. We sincerely appreciate your time in helping us complete this research
project that will ultimately provide guidance useful to you as a pavement professional.

A-17
APPENDIX B. SPECIMEN FABRICATION

Preliminary laboratory experiments were conducted to determine a specimen fabrication


protocol for LMLC specimens using the Texas field project materials. The objective of this task
was to determine the best conditioning time and temperature to prepare LMLC specimens in the
laboratory to match the initial stiffness of the field core specimens and to assess differences
between these types of specimens and PMLC specimens and RPMLC specimens. This section
describes the detailed procedure used to fabricate and test LMLC, PMLC, and RPMLC
specimens for five different mixtures.

B.1 LMLC Specimens

B.1.1 Materials Used and Mix Design

Five different mixtures were used in the Texas field project. These mixtures employed
two grades of asphalt binder: PG 64-22 unmodified binder and PG 70-22 modified with styrene-
butadiene-styrene (SBS), and TxDOT Class A (granite) and Class B (limestone) aggregates. One
source of RAP, which came from several different highway sections, was used, while the RAS
was produced by shredding and grinding manufacturer waste shingles. The binder content in the
recycled materials was 5.3 percent and 18.0 percent for the RAP and RAS, respectively. Figure
B.1 and B.2 illustrate the RAP and RAS materials, while Figure B.3 shows the gradation of the
virgin aggregate and the recycled materials.
Two chemical-based WMA additives—Cecabase® and Evotherm® 3G—and three types
of RAs—HydroGreen, Evoflex, and ERA-1—were used. HydroGreen (henceforth labeled as
RA1) is a combination of long chain and tricyclic organic acids, resin acids, fatty acids, sterols,
and esterified fatty acids (PVS Chemicals, Inc.). Evoflex (henceforth labeled as RA2) consists of
fatty acid derivatives, and ERA-1 (henceforth labeled as RA3) is an emulsion-based product.
RA1 and RA2 were added to the mixtures as a replacement. This means that the virgin binder
content in the mix was reduced by the specified amount of RA. RA3 was incorporated in the
mixture as an addition, which means that the specified amount of RA was combined with the
other components without adjusting the virgin binder content. Similarly, all WMA additives
were incorporated in the mixture as additions.
The five mixtures used in the Texas field project consisted of one mixture with only
virgin materials and four mixtures with a combination of virgin and recycled materials. A
summary of the characteristics of the mixtures is presented in Table B.1. The materials and mix
design complied with TxDOT’s specifications and procedures.

Figure B.1. Texas Field Project RAP Material . Figure B.2. TX Field Project RAS Material.

B-1
Gradation for Mixture 1 Gradation for Mixtures 2-5
RAP Gradation RAS Gradation
100.0
90.0
80.0
70.0
Percent Passing

60.0
50.0
40.0
30.0
20.0
10.0
0.0
0.075 0.3 0.6 2.36 4.75 9.5 19 25
Sievev Size (mm)

Figure B.3. Virgin and Recycled Aggregate Gradation


Table B.1. Mixture Characteristics for the Texas Field Project.
Mixture 1 Mixture 2 Mixture 3 Mixture 4 Mixture 5
(Virgin) (Control) (w/RA1) (w/RA2) (w/RA3)
Materials Used

Binder Grade 70-22 64-22 64-22 64-22 64-22

RAP / RAS No Yes Yes Yes Yes

WMA Additive — Cecabase® — Evotherm® 3G —

RA — — HydroGreen Evoflex ERA-1

Mixture Proportions
Optimum Asphalt
4.9 4.9 4.9 4.9 4.9
Content (%)
RAP / RAS — — 10% / 5.0% 10% / 5.0% 10% / 5.0%
WMA Additive (0.5) of total (0.5) of total
Dosage (%) binder binder
(0.13) of total (3.7) of total (1.3) of total
RA Dosage (%)
mix binder binder

B-2
B.1.2 Recycled Material Handling

One of the critical steps in producing high-quality HMA and/or WMA recycled mixtures
is handling of the recycled materials. Collecting, drying, and heating recycled materials for
preparing samples in the lab can have an effect on the properties of the recycled mixtures.
However, no clear guidance is provided in current standards with regard to proper handling of
recycled materials (West and Marasteanu 2013). Therefore, actual practices used in this study for
handling recycled materials in the laboratory are detailed next.

B.1.2.1 RAP/RAS Collecting


For collecting RAP/RAS materials from the Texas field project, samples were acquired
from the stockpiles illustrated in Figures B.4 and B.5 by using standard-size barrels. These
barrels were labeled and kept inside a storage unit at TTI’s Riverside Campus.
When needed, the required amount of RAP and/or RAS was transferred to TTI’s McNew
laboratory and placed in large, shallow pans and then mixed thoroughly to reduce variability
before any further action. The materials at the bottom of each barrel were removed since they
contained large quantities of clay and dust.

Figure B.4. RAP Stockpile Figure B.5. RAS Stockpile

B.1.2.2 RAP/RAS Drying


It is common for RAP/RAS stockpile field samples to have high moisture contents, which
may reach 5 percent or more in most cases (West and Marasteanu 2013). Higher moisture
contents can definitely affect the proportion of the RAP/RAS in the recycled mixtures. Plus,
given the fact that higher moisture contents mean higher temperatures during mixing because the
RAP/RAS must be completely dried before the aged binder can be heated sufficiently to be able
to be blended with the virgin binder, it is imperative to dry the recycled materials thoroughly.
In the laboratory, the RAP/RAS materials were placed into separate shallow pans with an
even depth of approximately 2 inches (5 cm) and then stored in a temperature-controlled room at
60°C. Many trials were conducted to determine the adequate period for drying the RAP/RAS
without excessive heating and further aging the materials. Basically, the recycled materials were
stored at 60°C, and each sample was weighed periodically every hour. After 6 and 7 hr of drying,
the weight was practically constant, which indicated stable moisture content. Thus, 6 hr was
selected as the standard drying time to prevent additional aging of the recycled materials.

B.1.2.3 RAP/RAS Heating


As mentioned previously, since no clear guidance is provided in current standards
regarding RAP/RAS heating for mixing, practical guidance for heating the recycled materials in

B-3
the laboratory was followed. Zhou et al. (2013) recommended manually blending the virgin
aggregates with RAS and then storing them together for 2 hr in the oven at the required mixing
temperature. They concluded that 2 hr was adequate to heat the RAS and assure that the RAS
and the virgin aggregates had the same preset temperature. Therefore, for each type of mixture,
the RAP and RAS were manually blended with the virgin aggregates and placed in the oven for
2 hr before mixing at the specified mixing temperature.

B.1.3 Specimen Fabrication

To fabricate the LMLC specimens, the virgin aggregate and binder along with the
recycled materials were heated to the specified mixing temperature of 154°C (310°F) for Mixture
1 and 138°C (280°F) for Mixtures 2–5. The virgin aggregate was heated overnight at the mixing
temperature, while the recycled materials were combined with the virgin aggregate and placed in
the oven 2 hr prior to mixing. The virgin binder was also heated at the mixing temperature 2 hr
prior to mixing.
A portable mixer was used for preparing the mixtures. Afterwards, the loose mixture was
conditioned for 2 and 4 hr at 116°C (240°F) for Mixture 2 and 135°C (275°F) for the other
mixtures. The reason for conditioning Mixture 2 at 116°C (240°F) is that a WMA additive was
employed in the mixture. Specimens with a 6.0 inch (150 mm) diameter and 2.4 inch (61 mm)
height were molded using the Superpave gyratory compactor (SGC) at the compaction
temperature. Trial specimens were fabricated initially to ensure that the specimens had an air
void content in the range of 7.0 ± 0.5 percent.

B.2 PMLC and RPMLC Specimens

During the construction of the Texas field project, onsite PMLC specimens were
prepared at the laboratory located within the plant facilities. For each mixture type, plant mix
was collected from the trucks at the plant before leaving for the construction site and stored in
5-gal buckets. These loose mixtures were quickly placed in an oven for about 1–2 hr to stabilize
to the required compaction temperature: 132°C (270°F) for Mixture 1 and 121°C (250°F) for
Mixtures 2–5. Afterwards, the mixtures were molded into specimens 6.0 inches (150 mm) in
diameter and 2.4 inches (61 mm) in height using the SGC with air void content in the range of
7.0 ± 1.0 percent. After the specimens cooled, they were labeled and transferred, along with the
buckets of loose mix, to TTI’s McNew laboratory for further testing.
A few days later, samples of loose mix were delivered to the Federal Highway
Administration (FHWA) Mobile Asphalt Pavement Mixture Laboratory (MAPML) for the
fabrication of the RPMLC specimens. The loose mix was reheated to the previously mentioned
compaction temperatures and compacted to target air voids of 7.0 ± 1.0 percent. Two sets of
specimens were fabricated: 6 × 2.4 inch (150 mm × 61 mm) specimens for the resilient modulus
(MR) test, and 4 × 6 inch (100 mm × 150 mm) specimens for the dynamic modulus test.

B.3 Resilient Modulus Test

In order to establish a conditioning protocol (time and temperature) to prepare laboratory


specimens to match the stiffness of the field cores acquired right after construction, MR tests
were performed on field cores; LMLC, PMLC, and RPMLC specimens of asphalt mixtures with
and without RAP/RAS; RAs; and WMA additives.

B-4
The MR test was conducted in accordance with ASTM D7369 by applying a repetitive
haversine compressive load pulse, in the vertical diametral plane of the cylindrical specimen (6.0
× 2.4 inches [150 × 61 mm]), every 0.1 sec with a 0.9-sec rest period. During testing, a set of two
linear variable differential transformers aligned in the diametral plane (perpendicular to the load)
recorded the horizontal deformation occurring in the specimen due to the repeated load and
saved the data in the computer attached to the device. For each specimen type, three replicates
per mixture type were considered, and all MR measurements were performed by the same
individual to reduce testing variability. Figure B.6 shows the MR testing device.

Figure B.6. Resilient Modulus Testing Device and Specimen Setup.

MR was calculated using the following formula in accordance with ASTM D7369 with an
assumed Poisson’s ratio of 0.35 for the test temperature of 25°C:



Where:
MR = resilient modulus of elasticity, MPa (psi).
Pcyclic = cyclic load applied to the specimen, N (lb).
t = thickness of the specimen, mm (in.).
δh = recoverable horizontal deformation, mm (in.).
I1, I2 = constant values; for gauge length as fraction of specimen diameter = 1, I1 = 0.27
and I2 = −1.00.

B-5
μ = Poisson’s ratio, 0.35

B.4 Test Results

Complete MR results for the field cores and the LMLC, PMLC, and RPMLC specimens
are provided in Table B.2 through Table B.5, respectively. A discussion of the results and
selected conditioning protocol can be found in Chapter 4 of this interim report.

Table B.2. MR Results for the Field Cores.


Aging Stage/ Air Voids MR
Mixture Type Statistics
Conditioning Protocol (%) (ksi)
Average 9.3 251.5
Mixture 1 (Virgin) At construction Standard Deviation 0.5 13.7
COV% 5.8 5.5
Average 9.8 498.5
Mixture 2 (Control) At construction Standard Deviation 0.9 32.0
COV% 8.9 6.4
Average 9.6 576.0
Mixture 3 (w/RA1) At construction Standard Deviation 0.3 50.3
COV% 3.2 8.7
Average 8.2 540.3
Mixture 4 (w/RA2) At construction Standard Deviation 1.6 37.8
COV% 18.9 6.4
Average 9.9 458.3
Mixture 5 (w/RA3) At construction Standard Deviation 0.9 48.9
COV% 8.9 10.7

B-6
Table B.3. MR Results for the LMLC Specimens.
Aging Stage/ Air Voids MR
Mixture Type Statistics
Conditioning Protocol (%) (ksi)
Average 6.8 354.2
2 hr at 275°F (135°C) Standard Deviation 0.1 8.3
COV% 2.7 2.3
Mixture 1 (Virgin)
Average 7.0 393.9
4 hr at 275°F (135°C) Standard Deviation 0.2 25.3
COV% 3.2 6.4
Average 7.0 770.5
2 hr at 275°F (135°C) Standard Deviation 0.4 57.4
COV% 6.1 7.4
Average 6.5 955.2
Mixture 2 (Control) 4 hr at 275°F (135°C) Standard Deviation 0.0 51.5
COV% 0.6 5.4
Average 7.1 596.8
2 hr at 240°F (116°C) Standard Deviation 0.3 40.7
COV% 4.1 6.8
Average 7.0 515.7
2 hr at 275°F (135°C) Standard Deviation 0.3 37.5
COV% 4.8 7.2
Mixture 3 (w/RA1)
Average — —
4 hr at 275°F (135°C) Standard Deviation — —
COV% — —
Average 6.9 489.9
2 hr at 275°F (135°C) Standard Deviation 0.3 27.1
COV% 4.7 5.5
Mixture 4 (w/RA2)
Average — —
4 hr at 275°F (135°C) Standard Deviation — —
COV% — —
Average 7.2 503.9
2 hr at 275°F (135°C) Standard Deviation 0.2 45.4
COV% 3.2 9.0
Mixture 5 (w/RA3)
Average 6.7 669.2
4 hr at 275°F (135°C) Standard Deviation 0.2 26.6
COV% 3.6 3.9

B-7
Table B.4. MR Results for the PMLC Specimens.
Aging Stage/ Air Voids MR
Mixture Type Statistics
Conditioning Protocol (%) (ksi)

1 to 2 hr to achieve the Average 7±1 442.3


Mixture 1 (Virgin) required compaction Standard Deviation — 10.3
temperature 270°F (132°C) COV% — 2.3
1 to 2 hr to achieve the Average 7±1 667.9
Mixture 2 (Control) required compaction Standard Deviation — 36.0
temperature 250°F (121°C) COV% — 5.3
1 to 2 hr to achieve the Average 7±1 603.7
Mixture 3 (w/RA1) required compaction Standard Deviation — 22.1
temperature 250°F (121°C) COV% — 3.6
1 to 2 hr to achieve the Average 6.6 498.8
Mixture 4 (w/RA2) required compaction Standard Deviation — 3.3
temperature 250°F (121°C) COV% — 0.6
1 to 2 hr to achieve the Average 6.5 624.0
Mixture 5 (w/RA3) required compaction Standard Deviation — 4.1
temperature 250°F (121°C) COV% — 0.6

Table B.5. MR Results for the RPMLC Specimens.


Aging Stage/ Air Voids MR
Mixture Type Statistics
Conditioning Protocol (%) (ksi)

1 to 2 hr to achieve the Average 7.3 541.6


Mixture 1 (Virgin) required compaction Standard Deviation 0.0 27.7
temperature 270°F (132°C) COV% 0.0 5.1
1 to 2 hr to achieve the Average 6.5 876.3
Mixture 2 (Control) required compaction Standard Deviation 0.3 62.4
temperature 250°F (121°C) COV% 4.7 7.1
1 to 2 hr to achieve the Average 7.1 858.2
Mixture 3 (w/RA1) required compaction Standard Deviation 0.2 84.8
temperature 250°F (121°C) COV% 3.5 9.8
1 to 2 hr to achieve the Average 6.9 849.7
Mixture 4 (w/RA2) required compaction Standard Deviation 0.0 53.0
temperature 250°F (121°C) COV% 0.0 6.2
1 to 2 hr to achieve the Average 6.9 754.2
Mixture 5 (w/RA3) required compaction Standard Deviation 0.1 48.9
temperature 250°F (121°C) COV% 2.1 6.4

B-8
APPENDIX C. TEXAS FIELD PROJECT CONSTRUCTION REPORT

The Texas Department of Transportation (TxDOT) executed the Texas State Highway 31
(SH 31) reconstruction project (Project ID CSJ 064-01-068) in the summer of 2014. This
reconstruction project, located in northeast Texas, included an approximately 1.4-mi-long asphalt
overlay placement. This project used five test sections to study and evaluate the effects of
different rejuvenators on the performance of asphalt mixtures with high RAP and RAS content.
Originally, this project was initiated under a research study sponsored by TxDOT. The overlay
was constructed in the first week of June 2014. APAC-Texas Inc. was the general contractor for
this project. The same contractor provided the materials and paved the overlay.
SH 31, at this project site, is a divided rural highway with two lanes in each direction.
Typical roadbed width, in each direction, was 30 ft including two 12-ft travel lanes and a 3-ft
shoulder on each side. The contractor placed five test sections on the eastbound outside travel
lane and right shoulder. These test sections were located between the east side of the city of
Murchison and the west side of the city of Brownsboro, Texas (Figure C.1). The new
construction of these test sections included 1-inch crack attenuating mix (CAM) followed by one
layer of hot rubber seal coat and 2-inch dense-grade Type C mix as a surface course. The inner
part of the existing pavement structure before this reconstruction included a 4-inch HMA layer
and 6-inch cement concrete layer. The outside (6 ft) of the existing pavement structure included a
4-inch HMA layer, then another 6-inch HMA layer, and 4 inches of iron core base at the bottom.
Figure C.2 shows the typical existing pavement structure. Pavement widening at some point in
the past attributed to this unusual pavement structure. Annual average daily traffic in each
direction measured in 2013 was approximately 5000 with 18 percent truck traffic.
Each of the five test sections had a different surface mix design for the Type C mixture,
as shown in Table C.1.

Table C.1. Test Sections with Five Different Mixtures.


Section Section Description Additive/Rejuvenator Dosage
No. Name
1 Virgin Mix Only virgin aggregate with N/A
PG 70-22 binder
2 Control Mix 10% RAP and 5% MRAS with Cekabase: 0.5% of total AC by
PG 64-22 weight
3 HydroGreen 10% RAP and 5% MRAS with HydroGreen: 0.2% of total mix by
PG 64-22 + Hydrogreen weight (0.75% weight of RAP +
1.0% weight of RAS)
4 Evoflex 10% RAP and 5% MRAS with Evoflex: 3.7% of total AC by
PG 64-22 + Evoflex + weight
Evotherm Evotherm 3G: 0.3% of total AC by
weight
5 ERA-1 10% RAP and 5% MRAS with 1.3% of total AC by weight
PG 64-22 + HydroGreen

C-1
Section 5
Section 4
Section 2
Section
Section 1

Figure C.1. Layout of Five Test Sections on SH 31.

Figure C.2. Typical Existing Section.

C.1 Mixtures and Materials

Figure C.3 and C.4 present the mixture designs used for this construction project. The
binders came from the Lion Oil terminal located at Longview, Texas. PG 64-22 was an
unmodified binder, whereas PG 70-22 was a modified by SBS supplied by Kraton Inc. Lion Oil
used crude oil from various sources, primarily from the local area and areas along its pipelines.
The source of the RAP is unknown, but it came from several different highway sections. RAS
was produced by shredding and grinding manufacturer waste shingles from a shingle
manufacturer in Corsicana, Texas. The TxDOT Class A (granite) aggregate came from the Smith
Buster quarry in Oklahoma, and the TxDOT Class B (limestone) aggregate came from the
Hanson quarry located in Bridgeport, Texas. All mixtures for this project were produced at
APAC’s hot-mix plant located just west side of Tyler, Texas. Table C.2 summarizes the
production, placement, and ambient temperatures during laydown for different mixes.

C.2 Production of Mix and Paving

For each test section, 350 tons of asphalt mix was produced and placed. Sections 1, 2, and
3 were constructed on June 3, 2014. All the area was paved with the control mix except

C-2
Sections 1, 3, 4, and 5, each approximately 1800 ft in length. Note that Section 2 was established
between Sections 1 and 3. Sections 4 and 5 were paved on June 4, 2014. The Section 4 mix was
placed right next to the Section 3 mix. Besides the sections, other areas of the roadbed were
paved with the control mix. Production began with the mixture used in Section 1 (virgin
aggregate with PG 70-22 binder) at 7:15 a.m. on June 03, 2014, at a rate of 200 tons/hr. Table
C.2 presents the mixture production schedule with temperature.

Table C.2. Production, Paving, and Ambient Temperatures.


Section Mixture Date of Plant Mix Paving Ambient
Production Temp, F Temp, F Temp, F
1 Virgin Mix 06/03/2014 325–327 285–290 75–80
2 Control Mix 06/03/2014 275–280 255–260 80–84
3 HydroGreen 06/03/2014 280–290 265–270 86–88
4 Evoflex 06/04/2014 275–285 240–250 74–78
5 ERA-1 06/04/2014 280–287 260–265 81–84

C.2.1 Section 1: Virgin Mix

The Section 1 virgin mix was paved on the morning of June 3, 2014. It started at
Station 472+37 and ended at Station 490+00 (global positioning system [GPS] coordinate N
32.28094/W 095.73956 to N 32.28194/W 095.73400. The ambient temperature was 75°F during
the paving of this section. The virgin mix was produced at around 310°F and then hauled to the
paving side in 30 min. The hauling truck (a “flow boy”) directly dumped the mix into the
RoadTec shuttle buggy (material transfer device), and then the mix was transferred to the paver
(Figure C.5). The temperature behind the paver was around 290ºF, measured using infrared
temperature gun. Also, an infrared Pave-IR bar was used to measure the asphalt mat temperature
behind the paver.
A steel-wheel vibratory (breakdown) roller closely followed the paver. The compaction
was achieved by two passes at vibrating mode and two passes at static mode (Figure C.5a),
followed by four passes of a pneumatic roller (Figure C.6b) and then two passes of a static steel
finish roller (Figure C.6c). Figure C.7 shows the compacted mat. Table C.3 shows the list of
equipment used.

Table C.3. Paving Equipment Used for SH 31 Construction Project.


Equipment Type Manufacturer
Shuttle Buggy RoadTec
Paver Bomag
Vibratory Steel-Wheeled Roller Sakai
Pneumatic Roller Sakai
Steel-Wheel Finish Roller Caterpiller

C-3
C-4

Figure C.3. Mixture Design of Virgin Mix Placed on Section 1 (Virgin Mix).
C-5

Figure C.4. Mixture Design of Control Mix Placed on Section 2 (Control Mix).
Figure C.5. RoadTec Shuttle Buggy and Paver.

(a) Vibratory Roller (b) Pneumatic Roller (c) Steel Finish Roller
Figure C.6. Rollers Used for Compaction.

Figure C.7. Finished Test Section 1: Virgin Mix.

C-6
C.2.2 Section 2: Control Mix with 10 Percent RAP and 5 Percent RAS

Section 2 was the control mix section with 10 percent RAP and 5 percent RAS. This
section was a warm-mix asphalt section produced with the CekaBase additive. It started at
Station 501+50 and ended at Station 520+00 (GPS coordinate N 32.28265/W 095.73025 to
N 32.28371/W 095.72421). The mix temperature measured at the dump truck at the project site
was around 275–280°F, and the mat surface temperature behind the paver was around 260°F.
The rolling pattern was similar to that of Section 1 except that the breakdown roller had three
passes in vibratory mode and one pass in static mode followed by four passes with the pneumatic
roller and then two passes with the static steel finish roller. Some minor segregations were
observed after the laydown at the center of the mat (Figure C.8a), but they were not noticeable
on the compacted mat (Figure C.8b).

(a) (b)
Figure C.8. (a) Observed Segregation on Loose Mat, and (b) Finished Surface.

C.2.3 Section 3: Control Mix with Rejuvenator-Hydrogreen

Section 3 started at Station 536+00 and ended at Station 557+00 (GPS coordinate
N 32.28460/W 095.71922 to N 32.28585/W 095.72226). The mix temperature measured at the
dumping truck was around 290°F, although it was supposed to be a warm-mix temperature of
275F or below according to the experimental design. The mat temperature behind the paver was
around 270°F. Section 3 employed the same rolling pattern as Section 2. Again, some
segregation was observed on the loose mat (Figure C.9a), but the finished surface did not show
any sign of segregation (Figure C.9b).

C-7
(a) (b)
Figure C.9. (a) Observed Segregation on Loose Mat of Section 3, and (b) Finished Surface.

C.2.4 Section 4: Control Mix with Rejuvenator-Evoflex and Evotherm

Section 4 started at Station 557+00 and ended at Station 580+00, (GPS coordinate
N 32.28585/W 095.72226 to N 32.287081/W 095.70523). Note that the end point of Section 3
was the beginning of Section 4. Section 4 and Section 2 had the same rolling pattern. Section 4
was paved in the early morning at around 8:30 a.m. In the beginning, the mix temperature
measured at the dumping truck was around 270°F. The mat temperature behind the paver was
around 230–240°F. However, obvious segregation was observed on the loose mat (Figure C.10).
The paving crew asked to increase the production temperature. After the speed limit sign
(Figure C.11), the mix dumping temperature measured from the truck was 285°F. However,
segregation never disappeared in the whole test section (Figure C.12).

Figure C.10. Segregation behind the Paver—Section 4.

C-8
Figure C.11. Mix with Higher Production Temperature Placed beyond Speed Limit Sign.

Figure C.12. Observed Segregation Even after Higher Production Temperature.

C.2.5 Section 5: Control Mix with Rejuvenator-ERA-1

Section 5 started at Station 607+00 and ended at Station 527+00, (GPS N 32.28858/W
095.69669 to N 32.28970/W 095.69034). Section 5 was paved on the afternoon of June 4 starting
at around 1:00 p.m. The mix temperature measured at the dumping truck was around 287°F,
although it was supposed to be a WMA temperature of 275F or below. The mat temperature
behind the paver was around 265ºF. A rolling pattern similar to Section 2 was used to compact
the mat. No obvious segregation (Figure C.13) was observed in Section 5.

C-9
Figure C.13. No Obvious Segregation in Section 5.

C.3 Description of Asphalt Mix Plant

All the mixtures were produced at APAC’s asphalt mix plant located on the west side of
Tyler, Texas. The average distance between the plant and the test sections was about 20 mi, or a
25-min drive. This plant is somewhat unique in that it has two separate drums: one for drying the
aggregates and the other for mixing. Figure C.14 shows the overview of this hot-mix plant. RAP
and RAS were added with hot aggregate just outside the drying drum before they entered into the
mixing drum. The binder was directly injected into the mixing drum. The admixtures and
rejuvenators were injected into the AC line. This StanSteel brand plant, manufactured in 1997,
has a capacity of 400 tons of mixture per hour. The parallel mixing drum has a dimension of
18 ft in length and 6 ft in diameter with a 15-ft-long mixing zone. The counterflow drying drum
has a dimension of 9 ft in diameter and 38 ft in length. This natural-gas-fuelled plant has a
conventional baghouse emission system where part of the fines is returned to the drum. The plant
has six bins for virgin aggregates and three bins for RAP and RAS. The three insulated silos each
have a storage capacity of 200 tons of mixture. The plant has three horizontal binder storage
tanks. The plant produced typically 200 tons per hour during the construction of the test sections.
The binder temperatures at the storage tanks were 304F and 294F for PG 70-22 and
PG 64-22, respectively. Additives and rejuvenators were kept at an ambient temperature.

C.4 Sample Collection

Plant mix was collected from the trucks at the plant by climbing on scaffolding.
Figure C.15 shows the sample collection at the plant. Due to the demand from multiple research
projects and universities involved, a huge amount of plant mixtures were collected in 5-gal
buckets. Due to the large quantity of mixtures required for a given section, mixtures were
collected from multiple trucks. Small amounts of mixtures were immediately brought back to the
onsite laboratory for immediate compaction. Loose plant mix samples were collected usually
after 200 tons of production for any given section. The material sampling scheme is presented in
Table C.4. With the help of the paving contractor, the research team also collected forty 6-inch-
diameter road cores from four test sections. Road cores were obtained from the outside shoulder.
The team also collected the quality-control cores from random locations.

C-10
Figure C.14. Overview of the Asphalt Mix Plant.

Table C.4. Material Sampling Scheme.


Sample Type Material Point of Sampling
Lab Mixed, Lab Compacted Fine Aggregate Stockpile
Coarse Aggregate Stockpile
RAP Stockpile
RAS Stockpile
Admixture/Rejuvenators Storage Tank (plastic tote in
metal cage) at Plant
PG 64-22 Asphalt Terminal
PG 76-22 Asphalt Terminal
Plant Mixed, Lab Compacted Loose Mix Truck at Plant
Plant Mixed, Field Compacted Road Cores Shoulder

C.5 Onsite Specimen Compaction

Twenty specimens 6.0 inches (150 mm) in diameter and 2.4 inches (61 mm) in height
were compacted onsite using plant mix at the lab located within the plant premises. Loose plant
mix collected from trucks was quickly brought to the contractor’s lab located inside the plant and
placed in the oven between 1 to 2 hr to achieve respective compaction temperature for that
particular mix. The researchers compacted these specimens using a Troxler Superpave gyratory
compactor to 7 ± 1 percent air voids. Specimens were compacted at 270F and 250F for
mixtures with PG 70-22 (Section 1) and PG 64-22 binders, respectively.

C-11
Figure C.15. Collection of Loose Mix from Truck at Asphalt Mix Plant.

C.6 Stockpiles and Plant Details

Figure C.16 shows the separate conveyor belts carrying virgin aggregates and RAP/RAS
to the different parts of the asphalt plant. Figure C.17 depicts the entry point of
admixture/rejuvenator into the AC line. Figure C.18 through C.20 depict RAS, RAP, and
aggregate stockpiles. None of the stockpiles was covered. Moisture content of RAP, RAS,
screenings, and sand was 5.5 percent, 6.4 percent, 5.4 percent, and 8.1 percent, respectively.
Average moisture content of 5.0 percent was input during the production. RAPs were screened
over a 2-inch screen before mixing with aggregate. RAS was produced by shredding the
manufacturer waste shingles using a shredder located on the same premises.

C-12
Figure C.16. Aggregate and RAP/RAS on Separate Conveyor Belts.

Figure C.17. Admixture/Rejuvenator Injected Directly into the AC Line.

C-13
Figure C.18. RAS Stockpile.

Figure C.19. RAP Stockpile.

C-14
Figure C.20. Aggregate (One of Several) Stockpile.

C-15

You might also like