Sie sind auf Seite 1von 195

THE NATIONAL ACADEMIES PRESS

This PDF is available at http://nap.edu/25701 SHARE


   

Concrete Technology for Transportation Applications (2019)

DETAILS

194 pages | 8.5 x 11 | PAPERBACK


ISBN 978-0-309-48101-4 | DOI 10.17226/25701

CONTRIBUTORS

GET THIS BOOK Jamshid Armaghani, Global Sustainable Solutions, LLC and Tara Cavalline,
University of North Carolina at Charlotte; National Cooperative Highway
Research Program; Transportation Research Board; National Academies of
FIND RELATED TITLES Sciences, Engineering, and Medicine

SUGGESTED CITATION

National Academies of Sciences, Engineering, and Medicine 2019. Concrete


Technology for Transportation Applications. Washington, DC: The National
Academies Press. https://doi.org/10.17226/25701.

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

– Access to free PDF downloads of thousands of scientific reports


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

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

Copyright © National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

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

NCHRP SYNTHESIS 544


Concrete Technology for
Transportation Applications

A Synthesis of Highway Practice

Jamshid Armaghani
Global Sustainable Solutions, LLC
Gainesville, FL

Tara Cavalline
University of North Carolina at Charlotte
Charlotte, NC

Subscriber Categories
Highways  •  Maintenance and Preservation

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

2020

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

NATIONAL COOPERATIVE HIGHWAY NCHRP SYNTHESIS 544


RESEARCH PROGRAM
Systematic, well-designed, and implementable research is the most Project 20-05, Topic 49-09
effective way to solve many problems facing state departments of ISSN 0547-5570
transportation (DOTs) administrators and engineers. Often, highway ISBN 978-0-309-48101-4
problems are of local or regional interest and can best be studied by Library of Congress Control Number 2019956812
state DOTs individually or in cooperation with their state universities
© 2020 National Academy of Sciences. All rights reserved.
and others. However, the accelerating growth of highway transporta-
tion results in increasingly complex problems of wide interest to high-
way authorities. These problems are best studied through a coordinated
program of cooperative research. COPYRIGHT INFORMATION
Recognizing this need, the leadership of the American Association Authors herein are responsible for the authenticity of their materials and for obtaining
of State Highway and Transportation Officials (AASHTO) in 1962 ini- written permissions from publishers or persons who own the copyright to any previously
tiated an objective national highway research program using modern published or copyrighted material used herein.
scientific techniques—the National Cooperative Highway Research Cooperative Research Programs (CRP) grants permission to reproduce material in this
Program (NCHRP). NCHRP is supported on a continuing basis by publication for classroom and not-for-profit purposes. Permission is given with the
funds from participating member states of AASHTO and receives the understanding that none of the material will be used to imply TRB, AASHTO, FAA, FHWA,
FMCSA, FRA, FTA, Office of the Assistant Secretary for Research and Technology, PHMSA,
full cooperation and support of the Federal Highway Administration,
or TDC endorsement of a particular product, method, or practice. It is expected that those
United States Department of Transportation. reproducing the material in this document for educational and not-for-profit uses will give
The Transportation Research Board (TRB) of the National Academies appropriate acknowledgment of the source of any reprinted or reproduced material. For
of Sciences, Engineering, and Medicine was requested by AASHTO to other uses of the material, request permission from CRP.
administer the research program because of TRB’s recognized objectivity
and understanding of modern research practices. TRB is uniquely suited Cover photo caption: Select concrete technologies used in transportation projects.
for this purpose for many reasons: TRB maintains an extensive com-
mittee structure from which authorities on any highway transportation
subject may be drawn; TRB possesses avenues of communications and
NOTICE
cooperation with federal, state, and local governmental agencies, univer-
sities, and industry; TRB’s relationship to the National Academies is an The report was reviewed by the technical panel and accepted for publication according to
procedures established and overseen by the Transportation Research Board and approved
insurance of objectivity; and TRB maintains a full-time staff of special- by the National Academies of Sciences, Engineering, and Medicine.
ists in highway transportation matters to bring the findings of research
The opinions and conclusions expressed or implied in this report are those of the
directly to those in a position to use them.
researchers who performed the research and are not necessarily those of the Transportation
The program is developed on the basis of research needs identified by Research Board; the National Academies of Sciences, Engineering, and Medicine; or the
chief administrators and other staff of the highway and transportation program sponsors.
departments, by committees of AASHTO, and by the Federal Highway The Transportation Research Board; the National Academies of Sciences, Engineering,
Administration. Topics of the highest merit are selected by the AASHTO and Medicine; and the sponsors of the National Cooperative Highway Research Program
Special Committee on Research and Innovation (R&I), and each year do not endorse products or manufacturers. Trade or manufacturers’ names appear herein
R&I’s recommendations are proposed to the AASHTO Board of Direc- solely because they are considered essential to the object of the report.
tors and the National Academies. Research projects to address these
topics are defined by NCHRP, and qualified research agencies are
selected from submitted proposals. Administration and surveillance of
research contracts are the responsibilities of the National Academies
and TRB.
The needs for highway research are many, and NCHRP can make
significant contributions to solving highway transportation problems
of mutual concern to many responsible groups. The program, however,
is intended to complement, rather than to substitute for or duplicate,
other highway research programs.

Published reports of the

NATIONAL COOPERATIVE HIGHWAY RESEARCH PROGRAM


are available from

Transportation Research Board


Business Office
500 Fifth Street, NW
Washington, DC 20001

and can be ordered through the Internet by going to


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

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

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

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

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

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

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

The Transportation Research Board is one of seven major programs of the National Academies of Sciences, Engineering, and Medicine.
The mission of the Transportation Research Board is to provide leadership in transportation improvements and innovation through
trusted, timely, impartial, and evidence-based information exchange, research, and advice regarding all modes of transportation. The
Board’s varied activities annually engage about 8,000 engineers, scientists, and other transportation researchers and practitioners from
the public and private sectors and academia, all of whom contribute their expertise in the public interest. The program is supported by
state transportation departments, federal agencies including the component administrations of the U.S. Department of Transportation,
and other organizations and individuals interested in the development of transportation.

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

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

COOPERATIVE RESEARCH PROGRAMS

CRP STAFF FOR NCHRP SYNTHESIS 544


Christopher J. Hedges, Director, Cooperative Research Programs
Lori L. Sundstrom, Deputy Director, Cooperative Research Programs
Velvet Basemera-Fitzpatrick, Senior Program Officer
Demisha Williams, Senior Program Assistant
Eileen P. Delaney, Director of Publications
Natalie Barnes, Associate Director of Publications

NCHRP PROJECT 20-05 PANEL


Joyce N. Taylor, Maine DOT, Augusta, ME (Chair)
Socorro “Coco” Briseno, California DOT, Sacramento, CA
Anita K. Bush, Nevada DOT, Carson City, NV
Joseph D. Crabtree, University of Kentucky, Lexington, KY
Mostafa “Moe” Jamshidi, Nebraska DOT, Lincoln, NE
David M. Jared, Georgia DOT, Forest Park, GA
Cynthia L. Jones, Ohio DOT, Columbus, OH
Jessie X. Jones, Arkansas DOT, Little Rock, AR
Brenda Moore, North Carolina DOT, Raleigh, NC
Ben T. Orsbon, South Dakota DOT, Pierre, SD
Randall R. “Randy” Park, Avenue Consultants, Bluffdale, UT
Jack Jernigan, FHWA Liaison
Stephen F. Maher, TRB Liaison

TOPIC 49-09 PANEL


Norb Delatte, Oklahoma State University, Stillwater, OK
Daniel R. Dennis, New York State DOT, Albany, NY
Paul D. Krauss, Wiss, Janney, Elstner Associates, Inc., Falls Church, VA
Andy Naranjo, Texas DOT, Austin, TX
Lucy Priddy, U.S. Army Corps of Engineers, Pittsburgh, PA
Tyson D. Rupnow, Louisiana DOTD, Baton Rouge, LA
Julie M. Vandenbossche, University of Pittsburgh, Pittsburgh, PA
Robert A. Younie, Iowa DOT, Ames, IA
Ahmad A. Ardani, FHWA Liaison
Michael Praul, FHWA Liaison
Nelson H. Gibson, TRB Liaison
Amir N. Hanna, TRB Liaison

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

ABOUT THE NCHRP SYNTHESIS PROGRAM


Highway administrators, engineers, and researchers often face problems for which information
already exists, either in documented form or as undocumented experience and practice. This infor-
mation may be fragmented, scattered, and unevaluated. As a consequence, full knowledge of what has
been learned about a problem may not be brought to bear on its solution. Costly research findings
may go unused, valuable experience may be overlooked, and due consideration may not be given to
recommended practices for solving or alleviating the problem.
There is information on nearly every subject of concern to highway administrators and engineers.
Much of it derives from research or from the work of practitioners faced with problems in their day-
to-day work. To provide a systematic means for assembling and evalu­ating such useful information
and to make it available to the entire highway community, the American Association of State High-
way and Transportation Officials—through the mechanism of the National Cooperative Highway
Research Program—authorized the Transportation Research Board to undertake a continuing study.
This study, NCHRP Project 20-05, “Synthesis of Information Related to Highway Problems,” searches
out and synthesizes useful knowledge from all available sources and prepares concise, documented
reports on specific topics. Reports from this endeavor constitute an NCHRP report series, Synthesis
of Highway Practice.
This synthesis series reports on current knowledge and practice, in a compact format, without the
detailed directions usually found in handbooks or design manuals. Each report in the series provides
a compendium of the best knowledge available on those measures found to be the most successful
in resolving specific problems.

FOREWORD
By Velvet Basemera-Fitzpatrick
Staff Officer
Transportation Research Board

This synthesis documents how state departments of transportation select and deploy concrete
technologies in the construction of transportation facilities. It includes a review of domestic and
international literature on concrete technology, survey responses from 40 states, and five state case
examples illustrating the implementation and use.
Information used in this study was gathered through a literature review, a survey of state depart-
ments of transportation, and follow-up interviews with selected agencies, resulting in case examples
of concrete technologies. The literature review included standard practices, reports, and technical
papers. These represent, but are not limited to, the American Concrete Institute’s standard practices,
along with reports from the Federal Highway Administration, state departments of transportation,
and industry, as well as research papers from domestic and international universities and research
organizations.
Jamshid Armaghani, of Global Sustainable Solutions, LLC, and Tara Cavalline, at the University of
North Carolina at Charlotte collected and synthesized the information and wrote this report. This
synthesis is an immediately useful document that records practices that were acceptable within
the limitations and available knowledge at the time of its preparation. As progress in research and
practice continues, new knowledge will be added to that now at hand.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

CONTENTS

1 Summary
5 Chapter 1 Introduction
5 Background
6 Objectives and Focus
7 Scope of Work
9 Goal of the Synthesis Report

10 Chapter 2  Overview of Concrete Technologies


10 Introduction
11 High-Strength Concrete
17 Self-Consolidating Concrete
26 Internally Cured Concrete
35 Ultrahigh-Performance Concrete
41 Temperature Control of Mass Concrete
47 Precast Concrete Pavement
54 Roller-Compacted Concrete
61 Pervious Concrete
68 Recycled Concrete Aggregate
76 High Early Strength Concrete
81 Very High Early Strength Concrete Repair Materials and Alternative
Cementitious Materials
87 Performance-Engineered Concrete Mixtures

92 Chapter 3  Survey of State Practices


92 Introduction
92 Responses to Survey Questionnaire
101 Lessons Learned from Survey Results

102 Chapter 4  Case Examples of State Practices


102 Florida Department of Transportation: Temperature Control of Mass Concrete
105 Illinois Department of Transportation: Reducing Concrete Shrinkage
in Bridge Decks
107 Missouri Department of Transportation: Precast Concrete Pavement
Demonstration Project on I-57—Lessons Learned
109 New York State Department of Transportation: Implementation of
Performance-Engineered Mixtures
111 Tennessee Department of Transportation: Barriers and Solutions to Concrete
Technology Implementation

113 Chapter 5  Conclusions and Technology Information Gaps


113 Conclusions
114 Gaps in Concrete Technology Information

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

116 References
126 Appendix A  Survey Questionnaire
145 Appendix B  Responses to Survey Questionnaire
162 Appendix C  State DOT Specifications/Special Provisions

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

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

SUMMARY

Concrete Technology for


Transportation Applications

The past few years have seen some significant advances in concrete technology. For
example, newer concrete incorporating advances in admixtures and cementitious materi-
als has emerged. High-strength concrete (HSC) and ultrahigh-strength concrete (UHSC),
and self-consolidating concrete (SCC) have found widespread use in bridge members to
facilitate and accelerate construction. Ultrahigh-performance concrete (UHPC) is becom-
ing an essential material for joints connecting prefabricated elements in accelerated bridge
construction and has also seen uses in bridge overlays. High early strength concrete (HESC),
very high early strength concrete (VHESC) and other rapid-repair materials are increasingly
being used for accelerated construction and repairs of pavements and bridge decks. Roller-
compacted concrete (RCC), which has been used for military, industrial, and port facilities,
is now also becoming a growing part in state pavement construction. Pervious concrete
(PC) has also found uses in stormwater management applications, in street pavement con-
struction, and in base layers and drainage systems for highway pavements. Internally cured
concrete (ICC), precast concrete pavements (PCPs), and performance-engineered mixtures
(PEMs) are subjects of increasing interest and implementation efforts by a growing number
of state departments of transportation (DOTs). Also, the DOTs are more aware of the need
for the temperature control of mass concrete (TCMC) to preserve integrity and long-term
durability of the massive element.
Concrete technology is also facing some emerging challenges that need to be addressed.
These include the present or future depletion of high-quality aggregates in some parts of the
country, changes to power generating plants that will reduce the supply and consistency of
acceptable fly ashes, and the incorporation of reclaimed or traditionally landfilled materials
such as recycled concrete aggregate (RCA) into concrete. Challenges may also include
having to overcome barriers to the incorporation of new concrete technologies into
transportation projects. Such barriers may include institutional and cultural resistance at
the agency and by industry and insufficient training.
Although information about these concrete technologies is available in published
literature, there do not appear to be resources for the DOTs and industry on the advanced
and emerging concrete technologies and their applications and best practices. Beneficial to
DOTs would be information on the need for the technology, benefits, types of applications,
performance, experiences from state and industry implementations, as well as limitations
and gaps in the technology information and barriers to wider implementation of some
technologies.
The objective of this synthesis study is to provide (1) an overview of advanced and
emerging concrete technologies suitable for transportation applications, (2) state DOT
applications and practices of the technologies and performance of these technologies, and

1  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

2   Concrete Technology for Transportation Applications

(3) information on gaps in the technologies and in implementation efforts that may be
addressed to expand their use.
In this synthesis, information has been collected from a review of the literature and from
a survey of state DOTs. This report presents technical information that has been synthe-
sized from review of standard practices, reports, and technical papers. These include, but
are not limited to, American Concrete Institute standard practices; FHWA state DOT and
industry reports; and research papers from universities and research organizations, domes-
tic and international. A survey questionnaire was prepared and electronically transmitted
to the AASHTO Committee on Materials and Pavements members of the 50 states and the
District of Columbia. Forty state DOTs responded to the survey questionnaire and five
states prepared case examples of technologies implemented in their states as well as barriers
and solutions to technology implementation.
Specific technologies covered in this report include HSC, SCC, ICC, UHPC, TCMC, PCP,
RCC, PC, RCA, HESC and repair materials, and PEMs. The synthesized information from
literature review on each technology includes need for the technology, types of applications,
benefits and limitations, materials and mixtures, properties and characteristics, construction
guidelines and specifications, implementation, and performance.
The report also presents results of the survey responses from 40 states: AL, AZ, AR,
CO, CT, DE, FL, GA, ID, IL, KS, KY, LA, ME, MA, MI, MN, MS, MO, MT, NE, NH,
NJ, NY, NC, ND, OH, OR, PA, RI, SC, SD, TN, TX, UT, VT, WA, WV, WI, and WY. The
survey information identifies the states using the specific technologies and whether or not
the states have developed specifications for the technologies. The information also includes
types of applications in pavement and structural construction and repairs, level and extent
of the experience in implementation, and specific challenges and limitation of the tech-
nologies. The report also identifies many states that are experimenting with and/or have
implemented other technologies not covered in the report, as well as states that are utilizing
reclaimed materials such as RCAs and other landfilled materials. On the issue of shortage of
quality aggregates—availability of fly ash in the United States—the information is presented
in the form of number of states affected or unconcerned about these issues. Barriers to
implementation of advanced and emerging concrete technologies is another topic included
in the discussion of the state DOT responses.
Case examples from Florida, Illinois, Missouri, New York, and Tennessee DOTs of
technology implementation or of technology barriers are also presented to be shared with
other DOTs. Florida presented their experience and specification requirements on mass
concrete and temperature control requirements. Illinois described the effort to mitigate
bridge deck cracks through the use of shrinkage-compensating or -reducing concrete and
the use of ICC for bridge decks. Missouri presented a case example of a PCP project includ-
ing lessons learned and gaps in the technology that require attention. New York described
their use of performance tests in their bridge construction projects as part of their efforts to
use the concept of PEMs. Tennessee discussed the barriers to implementation of new concrete
technologies and presented some actions that would remove or lessen those barriers.
From the results of the literature review and of the survey responses, a number of key
conclusions have been drawn:
1. The literature review showed that implementation of advanced, emerging, and new con-
crete technologies has resulted in major benefits to the transportation infrastructure.
The benefits include accelerated construction, replacement and repairs of pavements
and bridges (UHPC, HESC, VHESC, RCC, and PCP), better performance and improved
durability (HSC, SCC, and ICC, PEM), control of temperature in massive structural

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Summary  3  

members to mitigate thermal cracking and improve durability (TCMC), and enhanced
sustainability and environmental benefits (RCA and PC).
2. The survey responses from 40 states showed that the top three most implemented con-
crete technologies are HSC (40 states), SCC (37), HESC (37), and lightweight concrete
(33), and the least implemented technologies are UHSC ≥ 10,000 psi (69 MPa) (15),
PC (13), and ICC (9).
3. The survey also showed the following results:
a. Fourteen states (Delaware, Florida, Georgia, Illinois, Kansas, Louisiana, Maine,
Missouri, New York, Ohio, Pennsylvania, Vermont, West Virginia, and Wyoming)
have experimented with or implemented technologies other than those discussed
in this report.
b. Only three states reported depletion in quality aggregates (Florida, Maine, and Kansas).
However, another 13 states predicted shortages in the future (Idaho, Louisiana,
Minnesota, Montana, New Jersey, New York, North Dakota, Oregon, Pennsylvania,
South Carolina, Texas, Utah, and Vermont). The remaining 24 states reported no
shortages.
c. Thirteen states reported current shortages in the availability of fly ash (Alabama,
Florida, Illinois, Maine, Massachusetts, Michigan, Missouri, New Jersey, New York,
North Dakota, Oregon, Rhode Island, and Texas), 16 other states (Arizona, Arkansas,
Colorado, Connecticut, Delaware, Georgia, Idaho, Minnesota, Mississippi,
Montana, Nebraska, North Carolina, South Dakota, Tennessee, Vermont, and
Washington) predicted future shortages, and the remaining 11 did not report
any shortages.
d. The top solutions offered to address the shortage of fly ash include the expanded use
of slag, use of alternative pozzolans such as metakaolin, and import of foreign ash
and use, after reprocessing, to achieve a lower loss on ignition, or importing ash from
other states.
e. The most widely used recycled/reclaimed material in concrete applications is
RCA. Fifteen states (Alabama, Colorado, Connecticut, Florida, Illinois, Michigan,
Minnesota, Missouri, New York, Ohio, Texas, Washington, West Virginia, Wisconsin,
and Wyoming) use RCA in pavements, and 5 states (Alabama, Connecticut, Illinois,
New York, and Texas) use RCA in structural applications as well.
f. Seven states have experimented with the use of other reclaimed materials in concrete
mixtures. Five states (Florida, Georgia, North Carolina, Oregon, and Wisconsin) have
conducted research on the use shredded/crumbed tire rubber. Four states (Florida,
New York, North Carolina, and Wisconsin) have reported research on the use of
bottom ash as an ingredient in concrete mixtures. Three states (Florida, New York, and
Wisconsin) have experimented with the use of granulated glass in concrete. Two states
(Florida and New York) have experimented with municipal waste ash, and Georgia
and Rhode Island have used plastic bottle fibers. Florida has also conducted research
on biomass ash.
g. The top five barriers to implementation of concrete technologies by the state include
– Technology not sufficiently proven to be adopted (30 states),
– Too expensive to use (28 states),
– Lack of experience and not enough training (27 states),
– No specifications or construction guidelines available (23 states), and
– Industry resistance (21 states).
h. Other notable responses on the issue of barrier to technology implementation include
the following:
– Lack of experience by agency and local industry,
– Concern about potential reduction in concrete mixture quality,

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

4   Concrete Technology for Transportation Applications

– Ability to assess long-term concrete durability with some technologies,


– Implementation challenges, and
– Time and cost-effectiveness.
Gaps in the information pertaining to specific technologies were also identified and
provided for further attention. Some of the information gaps are the following:
1. State DOTs reported that they need more training to successfully implement new
technologies.
2. The need for air entrainment in HSC and UHSC to resist freeze-thaw actions has not
been completely settled by the research community.
3. Unanswered questions remain about the use of SCC in pavement repairs and slab
replacements.
4. Some states indicated that sources of lightweight aggregates for ICC are not available at
convenient locations to make them cost-effective. Also, there seems to be uncertainty
about the expected ICC performance using lightweight aggregate from different
sources.
5. Most mixtures used to produce UHPC are proprietary. This causes an increase in
construction cost, according to the survey responses.
6. With respect to TCMC, there does not seem to be a consensus among the states, indus-
try groups, or published research on the maximum core temperature and temperature
differential between the core and surface of the structure.
7. There does not seem to be well-defined, acceptable procedures for maintenance, reser-
vation, and panel replacement in a posttensioned PCP.
8. Two main barriers to expanding the use of RCC in highway pavements are control of
surface smoothness and absence of effective load transfer or dowel bars.
9. The two major gaps in the PC technology are lack of a specialized paving machine to
place and uniformly compact the material without damaging its void structure and
availability of guidelines for design, construction, quality control testing and mainte-
nance of PC pavements.
10. The most common distress problem when using HESC in replacement panels and slabs
is premature cracking from thermal and nonuniform shrinkage stresses. Many states do
not seem to have effective measures to mitigate the problem or guidelines for when to
repair or remove the damaged slabs.
11. The long-term performance of VHESC and repair materials is not well understood,
including impact of the type of application and weather conditions. Issues such as
premature setting, excessive shrinkage, and cracking are areas of concern.
12. The states have shown interest in the feasibility of using alternative pozzolans to sup-
plement the expected shortages in traditional fly ashes. However, they are concerned
about the impact of the alternative pozzolans on short- and long-term performance
of concrete.
This synthesis report can benefit DOT engineers, consultants, and construction pro-
fessionals as well as university researchers by providing information on the purpose,
applications, and performance of concrete technologies currently used by the DOTs. The
references provide more detailed information on various aspects of each technology that
may be of interest to engineers, practicing professionals and the research community.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

CHAPTER 1

Introduction

Background
In recent years many advances have been made in concrete ingredients and mixtures, includ-
ing admixtures, and supplementary cementitious materials (SCMs) that allowed the develop-
ment of new concrete technologies and improvement of existing technologies for deployment
in transportation infrastructure. New and more effective admixtures have enabled the reduc-
tion of cement content and water–cement ratio and increased efficiency of cement hydration
to produce high early and later strengths of concrete for accelerated construction and repair of
pavements and bridges. Other admixtures have aided in retaining mix workability to increase
flowability and mitigate segregation of self-consolidating concrete (SCC); and have facilitated
placement, compaction, and smoother surface finish of roller-compacted concrete (RCC) and
pervious concrete pavements (PCPs).
Combined with better and more versatile admixtures, the inclusion of SCMs such as regular
and ultrafine fly ash, slag, silica fume, and metakaolin at regular and high cement replacement
proportions, has allowed reduction of the cement content and increased reaction efficiency
of the SCMs. This has contributed to high-strength concrete (HSC) and ultrahigh-strength
concrete (UHSC), with improved durability for better corrosion protection and freeze-
thaw resistance and more sustainable structures and pavements. Also, incorporating steel
fibers in concrete mixtures has contributed to the development of ultrahigh-performance
concrete (UHPC) technology and application for accelerated bridge construction (ABC) and
overlay methods.
The use of prewetted fine lightweight aggregate in concrete has shown good potential
for providing internal curing that supplements surface curing for enhanced durability
and shrinkage reduction in bridge decks. Also, recycled concrete aggregate (RCA) is becom-
ing increasingly available for use as base material or in concrete mixtures for pavements and
drainage structures. This has contributed to more sustainability due to reduction in disposal of
demolished concrete in landfills.
Placement of mass concrete in piers, pile caps, and other large members generates significant
heat that substantially increases concrete temperatures and temperature differential with poten-
tial for induced cracking. Many state highway agencies require the development of an effective
mass concrete control plan to control the rise in concrete core temperature and differential
temperatures between the core and surface to mitigate cracking.
With all the success in concrete technologies, potential challenges have also emerged, includ-
ing shortages in quality aggregates and the availability of fly ash in some states. Depletion of
quality aggregates is driving states to import aggregates, conserve resources, and modify specifi-
cations to allow lesser-quality aggregates to be used in nonstructural applications. Also, because

5  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

6   Concrete Technology for Transportation Applications

many power generating plants are converting from coal to gas as fuel, the potential shortage of
fly ash is prompting departments of transportation (DOTs) to modify their concrete specifica-
tions to allow more use of alternative SCMs such as slag, silica fume, and metakaolin. This also
presents a research opportunity to enable, after second processing, the use of imported fly ash
with high loss on ignition (LOI) and imported bottom ash as well as ashes from wood burning
and rice husks.
Utilizing landfill materials in concrete to enhance its properties will contribute to longer
availability of landfill storage space and provide greater sustainability and environmental pro-
tection. Municipal waste, rubber tires, and bottom ash have been used as sources of fuel in
some cement plants. Also, research has been conducted on the use of granulated glass, crumb
rubber, and solid waste from treatment plants in concrete applications to reduce and reuse
these reclaimed products.
The experience and implementation of the proven concrete technologies varies from state
to state and among different regions of the country. The main barrier to the adoption and
implementation is insufficient knowledge of these technologies with respect to specification,
construction guides, quality assurance/quality control (QA/QC) testing requirements, and
performance. Loss of experience due to retirement and heavy workload of engineers, internal
and industry resistance to change, and limited funding for research in many DOTs are some of
the other barriers to adoption of successful technologies.
Although information about new and established concrete technologies is available in pub-
lished literature, there is no single document available on existing applications and practices
for DOTs. Beneficial to the DOTs would be the awareness and sharing of information on
the application of appropriate concrete materials and technologies to encourage their wider
implementation.

Objectives and Focus


Synthesis Objective
The synthesis objectives were to
1. Provide an overview of concrete technologies suitable for transportation applications,
2. Report on the state DOT applications of the technologies, and
3. Identify information gaps in the technologies.

Definitions
The synthesis focused on new and traditional concrete materials and technologies defined
below:
High-Strength Concrete: HSC is defined by the American Concrete Institute (ACI), as con-
crete having a specified compressive strength of 8,000 psi (55 MPa) or greater, and it
does not include polymer-impregnated concrete, epoxy concrete, or concrete made with
artificial normal-weight and heavyweight aggregates (1).
Self-Consolidating Concrete: SCC is highly flowable, nonsegregating concrete that can spread
into place, fill the formwork, and encapsulate the reinforcement without mechanical
consolidation (2).
Internally Cured Concrete: ICC is a concrete mixture that uses prewetted, highly absorptive
material that releases moisture inside the concrete to enhance and maximize the cement
hydration without increasing the water–cement ratio (3).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Introduction  7  

Ultrahigh-Performance Concrete: UHPC is a mixture that includes portland cement, SCMs,


well-graded fine sand, a high dosage of fiber reinforcement (usually steel), as well as super-
plasticizers and other admixtures, and has a very low water–cement ratio. It is a highly
flowable and self-consolidating concrete that develops very high compressive and tensile
strengths and exhibits durable performance (4).
Temperature Control of Mass Concrete: TCMC is a construction plan for massive concrete
members. The purpose of the plan is to control the rise in internal temperature of concrete
and maintain the temperature differential between the interior and outside surfaces below
threshold levels that may cause cracking and loss of durability (5).
Precast Concrete Pavement: PCP systems are composed of concrete panels that are fabricated
off-site, transported to the project site, and installed rapidly on an existing pavement or
prepared base (6).
Roller-Compacted Concrete: RCC is a stiff concrete mixture with very low or no slump that is
used for road and highway pavements. RCC pavements are designed similarly to traditional
concrete pavements, but are constructed similarly to asphalt pavements, without the use of
forms, dowel and tie bars, or other reinforcement (7).
Pervious Concrete: PC is an open-graded concrete mixture with only a small amount of, or
no, fine aggregates, has a near-zero slump, and can contain 20% to 35% voids to allow rapid
passage of water through the body of the concrete in pavement surfaces, base layers, and
drainage structures (8).
Recycled Concrete Aggregate: RCA is produced by crushing and grading concrete debris from
infrastructure demolition, or by crushing returned “leftover” concrete from ready-mix
concrete supplied to infrastructure projects. RCA is used to produce concrete mixtures for
pavement construction and as a road base material (9).
High Early and Very High Early Strength Concrete: These concrete mixes are used in acceler-
ated construction and overlay and repair of pavements and bridges. The high early strength
concrete (HESC) is a concrete mixture that is designed to achieve a specified high early
strength within 24 hours, and, in many cases, in less than 12 hours (10). The very high early
strength concrete (VHESC) is a concrete mixture that can achieve a specified high strength
in less than 4 hours.
Performance-Engineered Mixture: PEM is a new concrete technology used by a growing
number of state DOTs to engineer mixtures based on specified performance indicators
assessed using performance-related test methods (11).

Synthesis Focus
The synthesis focused on information related to the nature of the above technologies, their
applications, aspects of materials and construction, benefits, performance, state users, suc-
cesses and limitations in implementation, as well as any gaps in information. The information
is intended to assist DOTs in making informed decisions about the appropriateness and appli-
cations of these technologies in their transportation projects and in identifying resources to
facilitate implementation of the technologies.

Scope of Work
The synthesis included three tasks:
• Task 1: Literature review
• Task 2: Survey of DOT practices and prepare case examples
• Task 3: Synthesis report

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

8   Concrete Technology for Transportation Applications

Task 1: Literature Review


The literature review synthesized information on the above technologies. Literature covered
in the review included DOT, FHWA, industry and military publications, ACI reports and stan-
dards, as well as research papers on work performed in university laboratories, at test tracks or
at new and in-service projects. Where relevant, international publications and practices were
also reviewed to supplement the information on the state-of-the-practice in the United States.
Information derived from the literature review is presented in Chapter 2. For each technol-
ogy, the following information is provided: an overview of the technology, applications, aspects
of materials, mixtures and constructions, benefits, performance, users’ experience, and limita-
tions in technology implementation, as well as any gaps in the technology.

Task 2: Survey of DOT Practices


A draft survey questionnaire was prepared and submitted to NCHRP for review, comments
and approval by the Topic Panel. Upon approval, the final version of the questionnaire was
formatted, using a web survey tool, and sent electronically to the DOT representative serving on
the AASHTO Committee on Materials and Pavements (COMP).
The survey questionnaire is presented in Appendix A. It included 22 questions, inquiring the
state DOTs about their implementation and performance records of the concrete technologies
covered in the literature review. Additional established technologies, not covered in the litera-
ture review, were also included in the survey questionnaire. These technologies included UHSC
greater than 10,000 psi (69 MPa), lightweight concrete, lightweight cellular concrete, latex modi-
fied concrete and polymer concrete.
The survey questionnaire also sought information on technologies not addressed in the
synthesis, shortages in quality aggregates, availability status of fly ash, extent of use of RCA, use
of other landfilled/reclaimed materials in concrete, and barriers to wider implementation of
established and new concrete technologies.
The goal was to ensure at least 80% (40 states) response rate from the DOT COMP representa-
tives. With follow up by the authors, 40 state DOT responses were received. The questionnaire
results were analyzed and are presented with commentary in Chapter 3. Details of the responses
are shown in Appendix B. Conclusions and gaps in the information derived from the survey
responses along with the synthesized information from the literature survey were presented in
Chapter 5 as conclusions and information gaps.

Case Examples
From the responses of state agencies, eight DOT agency representatives were invited to
participate in preparing case examples on notable technologies or barriers to technology imple-
mentation in their respective states. Five case examples were received from Florida, Illinois,
Missouri, New York and Tennessee. The case examples included temperature control of mass
concrete (Florida DOT), shrinkage control on bridge decks (Illinois DOT), precast concrete
pavement (Missouri DOT), performance engineered mixtures (New York State DOT), and bar-
riers to technology implementation (Tennessee DOT). Details of the case examples are pre-
sented in Chapter 4. Specifications provided by the respective states are included in Appendix C.

Task 3: Synthesis Report


This report is prepared based on information collected from the literature review, survey
responses, and case examples. The report is organized as follows:

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Introduction  9  

• Chapter 1: Introduction
• Chapter 2: Overview of Concrete Technologies
• Chapter 3: Survey of State Practices
• Chapter 4: Case Examples of State Practices
• Chapter 5: Conclusions and Technology Information Gaps
• References
• Appendix A: Survey Questionnaire
• Appendix B: Responses to Survey Questionnaire
• Appendix C: State DOT Specifications/Special Provisions

Goal of the Synthesis Report


The report is prepared for the benefit of DOT engineers, consultants, and construction and
industry professionals, as well as university students and researchers. The report provides infor-
mation on a number of established and new concrete technologies to encourage more states to
use these technologies and assist in removing barriers to wider implementation of the technolo-
gies. Useful references are listed. They include state DOT, FHWA, and ACI reports, and relevant
research papers in the specific technologies. These documents provide more detailed information
about specific aspects pertaining to the concrete technologies.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

CHAPTER 2

Overview of Concrete Technologies

Introduction
This chapter includes an overview of established, emerging, and new concrete technologies
and materials used by state highway agencies in construction, rehabilitation, and repairs of
pavements, bridges, and other structures. The technologies include the following:

• High-strength concrete (HSC);


• Self-consolidating concrete (SCC);
• Internally cured concrete (ICC);
• Ultrahigh-performance concrete (UHPC);
• Temperature control of mass concrete (TCMC);
• Precast concrete pavement (PCP);
• Roller-compacted concrete (RCC);
• Pervious concrete (PC);
• Recycled concrete aggregate (RCA);
• High early strength concrete (HESC);
• Very high early strength concrete (VHESC), repair materials, and alternative cementitious
materials; and
• Performance-engineered concrete mixtures (PEM).

Most technologies listed above have been developed for some time and many are being
specified and used by state DOTs. However, technologies such as ICC and PCP may be con-
sidered as emerging technologies. In recent years, there has been growing interest by some
state DOTs in PCP and ICC technologies. Some states have constructed and are evaluating
the technologies in demonstration projects. Some states have even developed specifica-
tions and construction guidelines and have implemented these technologies in roadway
and bridge projects.
PEM is a new concept in specifying concrete. A number of states in conjunction with
FHWA and academia are developing performance criteria for acceptance based on spe-
cific tests that measure or evaluate performance indicators rather than specific material
properties.
In discussing each technology an introduction is provided to briefly describe the technology.
This is followed by benefits and needs for the technology. Also, information is provided
on materials, mixture parameters, construction, tests and/or basic properties, a brief dis-
cussion of specifications and implementation examples, and technology limitations and
challenges.

10

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   11  

High-Strength Concrete
Introduction
HSC is defined by ACI 363R-10 (1) as concrete having a specified compressive strength of
8,000 psi (55 MPa) or greater, and it does not include polymer-impregnated concrete, epoxy con-
crete, or concrete made with artificial normal-weight and heavyweight aggregates. ACI recognizes
that as technology improves and higher compressive strengths are successfully demonstrated, it is
likely that the definition of HSC will continue to be revised.
Compared with normal-strength mixtures, HSC requires higher quality materials, the use
of water-reducing admixtures, and more rigorous specification requirements for QA/QC tests
and procedures to ensure that the mixture consistently meets the required workability, volume
stability and strength development.
The highest specified concrete compressive strength, as reported by ACI, is 14,700 psi
(101 MPa) in Texas (1). In fact, a survey of state DOTs conducted in conjunction with this syn-
thesis study showed that 15 of the 40 DOT responses indicated specifying UHSC with strengths
exceeding 10,000 psi (69 MPa) (Appendix B). (UHSC is defined in this synthesis as concrete with
compressive strength exceeding 10,000 psi.)

Benefits
HSC has demonstrated significant benefits in transportation projects in terms of optimizing
structural design and accelerating construction and repairs of structures and pavements. With
respect to design, the use of HSC allows the design of thinner and longer bridge girders (Figure 1)
placed at wider spacings and/or requiring less steel reinforcement. Pavements constructed with
HSC can also be optimized to handle heavier traffic volume over an extended service life without
increasing their thicknesses.
HSC may be specified in other concrete technologies such as SCC, UHPC, HESC, and ICC.
These technologies are used in accelerated construction and repairs.
SCMs such as fly ash, slag, and silica fume are often included in HSC mixtures. Their addition
in conjunction with the use of water-reducing admixtures and low water–cement ratio not only
produces high strength but also enhances concrete durability and longevity of the structure.

Applications
Since the 1990s, the use of HSC in prestressed concrete bridge girders has significantly
increased (12). In the mid-1990s, many demonstration projects were constructed as a result

Figure 1.   Eighty-foot high-strength concrete


prestressed girder.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

12   Concrete Technology for Transportation Applications

of the FHWA initiative to implement the use of HSC in bridges. Since then, the use of HSC
has been expanded in all states as well as in more diverse applications such as bridge and
tunnel components that are cast-in-place, precast/prefabricated, and prestressed and in
pavement construction and repairs.

HSC Mixture Materials


Cementitious Materials
Type II cement and SCMs such as slag or fly ash (Classes F and C), or blended hydraulic
cements incorporating slag or fly ash, can be used in the production of HSC, provided they are
in appropriate quantities that meet the strength and heat-of-hydration requirements (1). The
type and amount of cement and SCMs can have a significant effect on temperature development
within the concrete. Unless high early strength is required, such as in prestressed concrete, there
is no need to use Type III HESCs that generate high heats of hydration.
SCMs such as fly ash (Classes C and F), slag, silica fume, and metakaolin are widely used in
binary or ternary concrete mixtures to produce HSC. In finely divided form and in the presence of
moisture, they will chemically react with calcium hydroxide released by cement hydration to form
additional calcium silicate hydrate gel and thus contribute to the high-strength properties (1).
It is important that all cementitious materials be tested for acceptance and uniformity
and be carefully investigated for strength-producing properties and compatibility with the
other materials in the mixture, particularly chemical admixtures, before they are used in
the work (1).

Admixtures
Chemical admixtures are widely used in the production of HSC. Selection of type, brand, and
dosage rate of all admixtures needs to be based on performance with the other materials being
considered or selected for use on the project. Significant increases in compressive strength, con-
trol of rate of hardening, accelerated strength gain, improved workability, and durability can be
achieved with the proper selection and use of chemical admixtures (1).
A reliable track record of good performance in previous work and compatibility with the pro-
posed cementitious materials and between chemical admixtures needs to be considered when
selecting admixtures. Specifications for chemical admixtures and air-entraining admixtures are
covered under ACI 212.3R, ASTM C494/C494M, and C260. The most commonly used admix-
tures (1) are discussed.

Retarding Admixtures (ASTM C494/C494M, Types B and D).   Retarding admixtures


are highly beneficial in controlling early hydration, particularly as it relates to strength
(ACI 212.3R) (13). With all else being equal, increased hydration time results in increased
long-term strength.
Retarding admixtures are also beneficial in improving mixture workability. They can control
the rate of hardening in the forms to eliminate cold joints and provide more flexibility in place-
ment size and duration.
The dosage of a retarding admixture can be adjusted to give the desirable rate of hardening
under anticipated temperature conditions. Retarding admixtures frequently provide a strength
increase proportional to the dosage rate, although the selected dosage rate is significantly affected
by ambient temperature conditions (13). In hot weather, higher dosage of retarders in the mix-
tures would mitigate temperature-induced strength loss. In winter months, lower dosage will
prevent delayed setting time.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   13  

Normal-Setting Admixtures (ASTM C494/C494M, Type A).   These are water-reducing


admixtures, commonly called normal-setting admixtures, that can provide strength increases
while having minimal effect on setting time and hardening of concrete.

High-Range Water-Reducing Admixtures (ASTM C494/C494M, Types F and G).   High-


range water reducers (HRWRs) reduce the mixture water and allow lower water–cement ratio,
improve efficiency of the cement hydration and thus aid in developing HSC, particularly at early
ages (24 hours) (14). The use of HRWRs tends to increase concrete strength through reduc-
tion in the water–cement ratio while maintaining a consistent slump, or increasing slump while
maintaining equal water–cement ratio, or a combination of both effects.
HRWR admixtures frequently perform better in HSC mixtures when used in combination
with conventional water-reducing admixtures or water-retarding admixtures. This is because of
the increased slump retention and control of the hydration (1).

Accelerating Chemical Admixtures (ASTM C494/C494M, Types C and E).   Accelerating


admixtures are not normally used in HSC unless early form removal or early strength development
is absolutely critical. HSC mixtures can usually be proportioned to provide strengths adequate for
vertical form removal on walls and columns at an early age. In pavement repair applications they
contribute to rapid strength gain in HESC to allow opening the pavement to traffic (1).

Air-Entraining Admixtures (ASTM C260).   Air entrainment is needed to enhance mixture


workability in the plastic stage and the freeze-thaw resistance in hardened concrete. However,
entrained air can significantly reduce the strength of HSC mixtures and increase the potential for
strength variability. This means that extreme caution needs to be exercised with respect to the use
of air entrainment in HSC. Even though many DOTs require entrained air in prestressed precast
HSC bridge girders, their use needs to be carefully evaluated for cast-in-place applications (1).

Chemical Admixture Combinations.   Water-reducing, set-retarding, and HRWR admixtures,


or a combination of these types, have been used effectively to control water demand, rate of hydra-
tion, slump loss, and strength increase. Combining HRWRs with water-reducing or -retarding
chemical admixtures has become common practice to achieve optimum performance at lowest
cost. Self-consolidating HSC mixtures are frequently produced using HRWRs in conjunction
with viscosity-modifying admixtures. With optimized admixture combinations, improvements in
strength development and control of setting times and workability are achievable (1).

Aggregates
Production of HSC requires proper selection of quality aggregates for the specific application.
Coarse aggregate mineralogical characteristics, grading, shape, surface texture, elastic modulus
(stiffness), and cleanliness can influence concrete properties. Both fine and coarse aggregates
used for HSC need to, at a minimum, meet the requirements of ASTM C33/C33M.
HSC often uses higher strength and higher quality aggregates to generate the targeted com-
pressive strength level. Using normal-strength or low-quality aggregates will result in fracture of
the aggregate before fully developing the strength potential of the paste matrix or bond strength
of the aggregate–paste transition zone.
Coarse aggregate may have a more significant effect in HSC than in conventional concrete (15).
In conventional concrete, compressive strength is typically limited by the cement paste capacity
or by the capacity of the bond between coarse aggregate and cement paste. In HSC, where the
cement paste and coarse aggregate–cement paste bond are enhanced by the use of SCMs and
low water–cement ratio, ultimate strength potential may still be limited by the strength of the
aggregate particles (16).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

14   Concrete Technology for Transportation Applications

Maximum aggregate sizes of 1/2 in. (13 mm), or smaller sizes of coarse aggregate and crushed
coarse aggregate, are needed for use in HSC. Smaller sizes of coarse aggregate have greater sur-
face area for a given aggregate content, which improves bond between coarse aggregate and
cement paste and thus enhances the ultimate strength of concrete.

Mixture Proportioning
Concrete mixture proportions for HSC have varied widely. Factors influencing mixture pro-
portions include the required strength level, material characteristics, and type of application.
Mixture proportioning of HSC is a more critical process than proportioning normal-strength
concrete mixtures. The use of SCMs, admixtures, and low water–cement ratio are considered
essential in high-strength mixture proportioning. Many trial batches are often required to
generate and enable optimum mixture proportions to be identified (1).
HSCs made with SCMs may gain considerable strength at later ages. Therefore, they may
be evaluated at later ages, such as at 56 or 90 days, when construction requirements allow the
concrete more time to develop strength before loads are applied.

Water–Cementitious Material Ratio


Higher cementitious material contents and lower water contents have produced higher
strengths. In many cases, however, using larger amounts of cementitious material increases
water demand. The use of HRWRs has enabled concrete to be placed at flowing and self-
consolidating consistencies with lower water–cement ratio. Water–cementitious material
ratios by mass for HSCs have typically ranged from 0.25 to 0.40.

Supplementary Cementitious Materials


Typical SCMs include fly ash, slag, silica fume, and metakaolin and other pozzolans and
cementitious materials. The total cementitious content, including cement and SCMs, in HSC
has ranged from 650 to 1,000 lb/yd3 (386 to 593 kg/m3). Incorporating SCMs and admixtures
can significantly increase concrete strength and durability (17). Today, UHSC with specified
compressive strengths up to 16,000 psi (110 MPa) at 56 days have been produced successfully.
Another important benefit of the SCMs in HSC is their impact on reducing the heat gener-
ated by the cement hydrations. The rise in temperature and, subsequently, the peak temperature
resulting from high cement content in HSC mixtures can be reduced by using slag cement and
pozzolans. Reduction in the peak temperature also reduces the potential for cracking and loss of
durability, especially in hot weather and mass concrete.

Aggregate
Low fine aggregate contents with high coarse aggregate contents have resulted in a reduction
in paste requirements and have typically been more economical. Also, such proportions have
made it possible to produce higher strengths for a given amount of cementitious materials.
However, if the proportion of fine aggregate is too low, there may be serious problems in mixture
workability.
In HSC, it has been found that the highest strengths for a given water–cement ratio are
obtained by using smaller maximum size coarse aggregate. However, the selection of coarse
aggregate size and content for HSC may be influenced by mixture requirements such as modulus
of elasticity, creep, shrinkage, and heat of hydration. For these cases, larger aggregate sizes may
be more desirable.
In general, the least amount of fine aggregate consistent with necessary workability gives the
best strength for a given paste. Mixtures with objectionably high coarse aggregate contents,

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   15  

however, may exhibit poor pumpability or may be significantly more prone to segregation
during placement and consolidation (1).

Construction
Batching and Placement
HSC may be mixed entirely at the batch plant, in a central or truck mixer, or by a combina-
tion of the two. Batching, mixing, transporting, placing, and control procedures for HSC are not
essentially different from procedures used for normal-strength concretes. Special consideration,
however, needs to be given to minimizing the length of time between concrete batching and final
placement in the forms. Delays in concrete placement can result in a subsequent loss of long-
term strength or difficulties in concrete placement (1).

Curing
The potential strength and durability of HSC will fully develop only if the concrete is properly
cured for an adequate period. Cast-in-place HSC needs to be water cured because of the low
water–cement ratio in the mixture. At a water–cement ratio below 0.40, the ultimate degree of
hydration is significantly reduced if an external supply of water is not provided. ACI 308-16
“Guide to External Curing of Concrete” provides guidance for curing methods and materials,
curing for different types of construction, and means for monitoring curing procedures and
effectiveness (18).
The most effective method of water curing is ponding the horizontal surfaces with water.
In many projects, however, this is not feasible or practical. Other methods include fogging,
misting, or spraying at very early ages. Lawn sprinklers, applied continuously, are effective
where water runoff is of no concern. Soaker hoses are useful, especially on surfaces that are
vertical. Burlap, cotton mats, rugs, and other coverings of absorbent materials will hold water
on the surface, whether horizontal or vertical.
Liquid membrane-forming curing compounds assist in retaining the original moisture in
the concrete but do not provide additional moisture or completely prevent moisture loss.
Monomolecular film-forming agents have been effectively employed for interim curing before
deployment of final curing procedures for exposed surfaces susceptible to drying during
finishing. These so-called “evaporation retarders” are not to be used as an aid to finishing (1).

Tests and Properties


Compressive Strength
The compressive strength in HSC is usually much higher than 8,000 psi (55 MPa). Also, accord-
ing to Florida DOT research, HSC of 8,000 psi or higher can also be considered high-performance
concrete with respect to low permeability and corrosion resistance (19). This has been attributed
to the low water–cement ratio and the use of SCMs. With compressive strength, some exceeding
10,000 psi (UHSC), the concrete is often sampled and tested using 4- by 8-in. (10- by 20-cm)
cylinders, which are more compatible with loading limits of most testing machines than 6- by
12-in. (15- by 30-cm) samples. Also, proper grinding of the sample’s ends, or using capping
compounds or rubber caps, will ensure plainness of the sample’s ends and reduce the variability
in the compressive strength test results (20).
HSC gains a higher rate of strength at early ages compared with lower-strength concrete. The
typical ratios of 7- to 95-day strength average 0.60 for low-strength concrete, 0.65 for medium-
strength concrete, and 0.73 for HSC (21). It seems likely that the higher rate of strength devel-
opment of HSC at early ages is caused by an increase in the internal concrete temperature due

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

16   Concrete Technology for Transportation Applications

to a higher rate of heat of hydration, tighter spacing between the hydrated cement, and a lower
water–cement ratio.

Resistance to Freeze-Thaw
The use of air entrainment to increase resistance to freeze-thaw actions can contribute
to lowering the strength of HSC. Research results are inconclusive about the role and the appro-
priate dosage rate of air-entraining admixtures in HSC (1). Some studies have indicated that
with the low water–cement ratio and the use of SCMs, the HSC will have very low permeability
and high tensile strength, making the concrete structure more resistant to freeze-thaw actions.
This may negate the use of a large dosage of air entrainment (22, 23).

Shrinkage
For normal-strength concrete with highwater–cement ratio, the change in volume due to
evaporation of the unbound portion of the mixture water, commonly termed drying shrinkage,
is the predominant mechanism. For HSCs that have a low water–cement ratio and high binder
content, other volume-change mechanisms influence the overall magnitude and rate of shrink-
age and possible cracking. Most important among these are chemical shrinkage and autogenous
shrinkage.
Chemical shrinkage refers to the reduction in absolute volume of solids and liquids in paste
resulting from cement hydration (1). The absolute volume of hydrated cement products is less
than the absolute volume of cement and water before hydration (24). Autogenous shrinkage
is that portion of chemical shrinkage that starts at initial set and results in volume change in
concrete. It has been reported that autogenous shrinkage can be significant for HSC with water–
cement ratio less than 0.40 and silica fume content greater than 10% (1).

Permeability
Moist curing not only influences the strength development, but also contributes to lowering
the permeability of concrete. As the moist-curing period is increased, the strength development
will increase, and the permeability will be lower (25). The use of highly porous aggregates will
increase the permeability of the concrete because substances can flow more easily through
aggregate pores than through smaller pores of the cement paste (25). However, a Florida DOT
research project showed that HSC mixtures using Florida limestone aggregate with high absorp-
tion capacity, compared to dense limestone, river gravel, and granite, can achieve a similar low
permeability, using mixtures that include a combination of fly ash, slag, and silica fume and low
water–cement ratio (26). Another study by the Virginia DOT (27) demonstrated the effectiveness
of using silica fume, fly ash and slag to reduce concrete permeability and increase its durability.

Specifications
Most state DOTs have specification requirements for classes of concrete that are considered
high strength and high performance (28). The specifications apply mostly to bridge structural
members. The strength requirement is not necessarily set at 8,000 psi, but rather, at lower
strengths, with no limits on how high the concrete strength may reach during trial batches or
production.
The specification requirements for the concrete generally include allowable cement types and
quantity, minimum water–cement ratio, SCM types, and proportions. Also included are plastic
properties such as slump and air content as well as hardened properties such as strength and
durability. Some states provide special requirements for mass concrete and environmental con-
ditions with respect to temperature control, monitoring, and testing needs.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   17  

Challenges of HSC
The initial cost of HSC is higher than that of normal-strength concrete. This is due to the use
of higher quantities of cementitious materials and admixtures in the mixture as well as higher
cost of QA/QC to achieve a higher and more consistent quality of concrete during production
and in various stages of construction from placement to curing.
Another challenge is personnel experience in producing HSC and maintaining the speci-
fied concrete properties such as workability, strength, and durability. This requires more
training and greater availability of experienced personnel at the project site and in the con-
crete plants.
Also, with the use of high cement content, the heat generated from cement hydration can
result in significant rise in concrete temperature, causing cracks and delamination in new struc-
tures, pavements, overlays, and repairs.
However, the benefits of optimizing the design and accelerating construction tend to offset
the disadvantages of higher initial cost and other challenges with HSC. Also, using HSC with its
enhanced durability will most likely result in less maintenance and longer service life.

Self-Consolidating Concrete
Introduction
SCC (Figure 2) is a highly flowable, nonsegregating concrete that can spread into place,
fill the formwork, and encapsulate the reinforcement without mechanical consolidation (2).
SCC is used primarily in cast-in-place or precast structural members with highly congested
reinforcement.
In general, SCC mixtures are produced with conventional concrete materials, and incor-
porate admixtures such as high-range water reducers, viscosity modifiers, set retarders, and
workability retainers to achieve a high workability maintained for an extended period of
time. SCC has normal setting time and develops strength similar to that of conventional
concrete (29).

Figure 2.   Self-consolidating concrete mixture—


VSI rating 1 (30, pp. 37–38).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

18   Concrete Technology for Transportation Applications

Benefits
SCC mixtures have become more widely used in construction due to their favorable char-
acteristics, which include increased construction productivity, improved work environment
and safety, and higher strength and durability (2). With its high flow rate, the SCC mixture is
discharged in narrow and/or deep forms with congested reinforcement. The concrete mixture
rapidly fills the form without leaving voids and fully encapsulates the reinforcing bars and self-
consolidates without the need for external vibration. In precast plants, SCC mixtures can be dis-
charged from one point and allowed to travel some distance to fill a large portion of the precast
member without segregations.
The SCC mixture is likely to include SCMs, a blend of workability and water-reducing admix-
tures and low water–cement ratio. This mixture would produce HSC with enhanced durability
suitable for use in structures subjected to highly aggressive environments. Also, in precast plants,
the SCC mixture can be designed to achieve the necessary strength in less than 24 hours to allow
early stripping of the forms and release of the prestressing strands in the cast member.

Applications
Among precast applications for SCC are prefabricated bridge construction, including decks
and girders, and tunnel-lining segments and building structures (31). Cast-in-place applications
include columns, walls, and bridge piers and girders. Virginia recently used SCC for bridge sub-
structure repairs (32). Application of SCC in pavements has been limited to research and small
experimental projects. In Iowa, SCC was used in an experimental project to evaluate its applica-
tion in concrete pavement using slipform paving, and in Florida, research was conducted on the
use of high early strength SCC mixtures for accelerated slab replacement (30, 33, 34).

Materials
SCC mixtures include the same basic ingredients used in conventional concrete mixtures.
However, the SCC mixtures must achieve three important characteristics: (1) sustained fluidity
over a long period, (2) stability to travel without segregation, and (3) ability to consolidate and
self-level in the form without the aid of external vibration.

Cementitious Materials
In addition to portland cement, other cementitious materials such as fly ash, silica fume,
and ground-granulated blast-furnace slag (GGBFS), and fillers such as limestone powder, can
be incorporated in the mixture as partial replacement of the cement content to produce high-
performance and durable SCC (35, 36, 37). These cementitious materials benefit the SCC by
enhancing both its plastic and hardened properties, as well as controlling the heat of hydration.
For example, silica fume and GGBFS with their finer particles (compared to cement) tend to
increase the stability of SCC mixtures. Fly ash, with its spherical and smooth particles, can act
as a ball bearing in the mixture to enhance SCC workability, facilitate its spread in forms, and
improve the compressive strength (35, 36).
The replacement of a portion of cement with finely ground limestone filler has also been
shown to improve packing density and stability of the mixture. The concrete may exhibit up to
10% lower 28-day strength compared with similar concrete without the filler, according to ACI
Report 237R-07 (2). However, a recent study showed slightly different results. SCC mixtures
with 15% ground-limestone filler produced similar or marginally higher compressive strengths
at 28, 90, and 180 days than those of the reference concrete (37). These slight improvements,
which ranged between 2% and 8%, were partially attributed to the limestone powder’s filler
effect, which improved the packing density of the studied SCC mixtures. The replacement of

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   19  

part of the cement with a less reactive powder may prove beneficial when project requirements
limit the heat of hydration.

Aggregate
The intended type of application of the SCC mixture dictates the appropriate shape, nominal
maximum size, and gradation of coarse and fine aggregates (2). The nominal maximum size of
the coarse aggregate needs to be chosen to achieve an acceptable spread within the form, a pass-
ing ability around the congested reinforcement, and a stability to resist segregation of the SCC
mixture. ACI 237R-07 (2) suggests that the nominal maximum size of the coarse aggregate be
one size smaller than that recommended by ACI 301, “Specifications for Structural Concrete,”
to improve the passing ability. If the coarse aggregate is greater than 1/2 in. (12.5 mm) nominal
maximum size, then the absolute volume of coarse aggregate needs to be in the range of 28%
to 32% of the volume. For slab placements without reinforcement to obstruct the flow, the
nominal maximum size and percentage of total volume of coarse aggregate in the SCC mixture
may be increased (38).
The particle shape of the coarse aggregate will affect the workability of SCC. For the same
water content, a rounded coarse aggregate will provide greater mixture workability to fill and
consolidate within the formwork compared with a crushed stone of similar size. Also, blending
of different aggregate sizes can often improve the overall characteristics of the mixture. The fine
aggregate component needs to be well graded and preferably prepared from a blend of natural
and manufactured sand to improve the mixture’s plastic properties (2).

Admixtures
Polycarboxylate-based HRWRs are the most typical admixtures used for developing and
proportioning SCC mixtures. They tend to maintain workability of SCC mixtures to allow
longer transportation and construction time, as well as allowing the reduction of mixture water
to produce concrete with higher strength and lower permeability. Additionally, some, but not
all, HRWRs enhance stability and cohesiveness of the mixture (2).
Viscosity-modifying admixtures (VMAs) are used to adjust the viscosity and to improve the
ability of SCC mixtures to resist segregation. A VMA used with a compatible HRWR improves
the viscosity of the mixture and increases its ability to tolerate water adjustments between
batches. The use of a VMA is not always necessary, but a VMA can be advantageous when using
lower powder content and gap-graded aggregates (2).
In addition to HRWRs and VMAs, other admixtures and additives are often utilized in SCC.
These may include air-entraining admixtures, normal and mid-range water reducers, accelera-
tors, retarders, extended set-control admixtures, corrosion inhibitors, shrinkage reducers, and
liquid and dry color. Fibers can also be specified for use in SCC (2).
Guidelines have been developed for the use of SCC in precast and prestressed concrete bridge
elements (39). These guidelines address the selection of constituent materials, proportioning
of concrete mixtures, testing methods, fresh and hardened concrete properties, production
and quality control issues, and other aspects of SCC. Many of these recommendations can be
adopted for cast-in-place applications of SCC.

Mixture Proportioning
SCC mixtures need to be both fluid and stable to be successfully cast to achieve the desired
structural characteristics and durability performance. The required level of fluidity is greatly
influenced by the specific application. The fresh properties of SCC have a much higher
degree of workability and self-consolidation than any conventional concrete. The workability

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

20   Concrete Technology for Transportation Applications

characteristics of SCC include filling ability, passing ability, and stability (segregation resistance).
These characteristics need to be present in concrete mixtures to be considered SCC. To achieve
these characteristics, the SCC mixture needs to be carefully proportioned to account for the
application type and placement technique. Combining finely divided cementitious and
filling powders with portland cement and incorporating a blend of admixtures can enhance
the behavior of fresh SC in terms of its form filling ability, passing ability, and stability (2).
The intended application of the SCC can significantly affect the appropriate mixture propor-
tions. The proportions of fine and coarse aggregate and powder content (cement, SCMs, and
finely divided powders) for SCC need to be balanced to achieve the desired fresh and hardened
properties. For noncongested footings and in plain slab applications, SCC mixtures can include
a larger size and higher percentage of coarse aggregate and lower slump flow compared to an
SCC being used in a congested girder or column applications (2, 30). Columns and wall forms
congested with reinforcing steel will require an SCC mixture with greater passing ability to fully
encapsulate the reinforcement, enhanced stability to minimize segregation, and a well-balanced
ingredient to develop sufficient strength to meet the load requirement (2). The quantity, size, and
spacing of steel reinforcement in a structure, if any, and method of SCC delivery and discharge
play a major role in determining the filling ability, passing ability, and stability requirements (29).
Examples of successful SCC mixture designs, as suggested by ACI 237 (2), are shown in
Table 1. ACI emphasizes that the mixture proportions in Table 1 should not be copied or used
in a project without first performing field trials, because local materials may have a considerable
effect on the proportioning of SCC mixtures. The desired SCC performance needs to be verified
for local materials and admixtures, application type, construction methods, and weather condi-
tions. Table 1 mixtures can be used as a starting point, but then adjusted for the local condi-
tions. It is important to perform trial batches of the candidate SCC mixture to make any needed
adjustments in water and admixture contents to arrive at the proper slump flow and stability for
the specific application.

Construction
Mixture Production
Similar to other types of concrete, strict control of the materials characteristics and moisture
conditioning of the aggregates is paramount for successful production of SCC mixtures. It is
important that a consistent source of raw materials be used throughout the duration of a
project. Causes of variability in performance among the batches of the SCC production mixture
are changes in material characteristics and moisture condition of aggregates. Also, controlling
the mixing temperature and using appropriate dosage of admixtures have been found to be
important factors in controlling mixture set time and heat of hydration (40).

Transport.  SCC can be delivered to a job site by truck mixers. A concrete truck is an effective
method of transporting and placing SCC mixtures with slump flow levels of 18 to 30 in. (455 to
760 mm) (41). However, because of the fluidity of SCC mixtures, the volume of SCC placed into
a truck should not exceed 80% of the capacity of the drum in areas where the truck travels on
steep inclines to avoid spilling the fluid concrete. Also, to avoid drastic loss of fluidity, it is advised
that the revolving drum be turned in the mixing mode direction while in transport.
Note that use of some HRWRs will result in slump flow loss more rapidly than others
during transportation. This needs to be taken into consideration when preparing the production
mixture to ensure proper fluidity and stability upon arrival of the concrete truck to the casting
site. Alternatively, some suppliers have delivered the mixtures to the project at a conventional
concrete consistency and then added an HRWR to bring the mixture to an SCC consistency
prior to placement (2).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   21  

Table 1.   Examples of successful SCC mixturesa (2).

Slump flow 26 in. 26 in. 26 in. 33 in. 27 in. 26 in.


Polycarboxylate Yes Yes Yes Yes Yes Yes
Air entraining Yes Yes Yes Yes Yes Yes
Water reducer Yes — — Yes — —
VMA — — — Yes Yes Yes
Total cementitious material, 750 680 780 797 700 700
3
lb/yd
Cement 600 680 620 345 700 600
Fly ash 150 — — 140 — 100
GGBFS — — 160 312 — —
Water–cement ratio 0.37 0.42 0.39 0.34 0.41 0.40
Paste fraction (%) 37.1 36.5 38.1 36 34.7 35
Mortar fraction (%) 64.6 68.3 63.4 64 59.5 65.6
Volume of coarse
aggregate (%) 35.6 31.7 36.6 36 31 33.5
Total gradation (sieve size), % retained
1.0 in. (25 mm) — — — — — —
3/4 in. (19 mm) 2.3 — 0.7 0.75 3 9
1/2 in. (12.5 mm) 9.2 — 11.3 5.6 15 19
3/8 in. (9.5 mm) 5.1 — 6.5 11.8 14 8
No. 4 (4.75 mm) 25.4 26.6 23.6 26.2 15 14
No. 8 (2.36 mm) 14.4 23.3 16.9 12.4 16 4
No 16 (1.18 mm) 9.5 10 5.7 12.5 14 12
No. 30 (600 µm) 11.1 12.5 8.2 20.1 10 13
No. 50 (300 µm) 12.2 14.2 18.4 8.5 8 14
No. 100 (150 µm) 7 11.2 7.1 1.5 3 6
Pan 3.9 2.3 1.4 0.24 1 1
3 3
Note: 1 lb/yd = 0.593 kg/m . Dash = no information provided in reference to indicate that the item
was not used or applied.
a
Must perform trial batches to verify properties and performance for local materials, application,
construction methods, and weather conditions.

Placement.  SCC is placed in the forms by concrete truck, pump, hopper, bucket trans-
porters, or other specialized devices such as tremies. A properly designed SCC mixture can easily
be pumped into place without segregation. The height of discharge in tall walls and columns
needs to be shortened to avoid mixture segregation. The fluid SCC mixture can flow long dis-
tances without any mechanical consolidation. In practice, the flow distance is typically limited to
33 ft (10 m) to mitigate segregation of the concrete while ensuring self-consolidating properties.
When used in cast-in-place slabs and precast forms (Figures 3 and 4), the SCC flow and
space filling of the open pit, mold, or formwork is affected by the placement method and the
concrete fluidity and stability. These characteristics need to be considered when designing
SCC mixtures for slabs and precast elements. Trial batching is imperative to ensure successful
placement characteristics.
When placing SCC, the formwork needs to be watertight (nonleaking) and grout-tight, espe-
cially when the mixture has relatively low viscosity. Also, there is a need to design the formwork
for water tightness compared to conventional formwork so as to avoid honeycombs and sur-
face defects. Also, the highly fluid nature of SCC may lead to higher formwork pressure than

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

22   Concrete Technology for Transportation Applications

Figure 3.   Placement of SCC in a slab pit (38).

conventional concrete, especially when the casting rate is high. Therefore, the formwork designs
must accommodate the expected liquid head pressures. This will allow unrestricted placement
rates and permit the contractor to take full advantage of a fast casting rate of the SCC.
During construction, SCC needs to be discharged at one point and allowed to flow into place
before moving the point of placement. Also, when possible, SCC needs to be discharged in the
direction of desired flow to maximize the distance of travel. A fresh layer of SCC can be placed
onto recently placed SCC that has not yet achieved initial set. In this case, it is acceptable to use
an internal or external vibrator for a 2- to 3-second duration. However, the concrete needs to be
deposited continuously and in layers of such thickness that no fresh layer is placed on concrete
that has hardened enough to cause a seam or plane of weakness. If a section cannot be placed
continuously, construction joints need to be provided (2).
For casting new or replacement slabs using SCC, it is advised to have sufficient workers avail-
able to accomplish surface strikeoff and finish in a timely manner. This is due to faster SCC
placement that yields a larger concrete surface area at a given time interval ready to be finished
as compared to conventional concrete (2). Also, research at FAMU-FSU College of Engineering
showed that when using accelerators in high early strength SCC for slab replacements, surface

Figure 4.   Placement of SCC in precast member


(courtesy of Dura-Stress, Florida).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   23  

strikeoff and finish need to be executed simultaneously with concrete placement because the
mixture tends to set much faster when stationary than conventional concrete (38).

Curing.  There should not be a difference in curing requirements for SCC compared with
curing for conventional concrete mixtures. Curing is essential, and early protection of exposed
surfaces is key to preventing rapid moisture loss that could lead to plastic shrinkage cracking.

Tests and Properties


Quality Assurance/Quality Control Tests
The four main characteristics of SCC are ability to fill the form under its own weight, resist
segregation, flow through reinforcing bars or other obstacles without segregation and without
mechanical vibration, and achieve a quality surface finish (2).
To evaluate these characteristics of SCC, the following tests can be performed.

Slump Flow.   This test is based on ASTM C143/C143M and is used to determine the hori-
zontal free-flow or spread characteristics of SCC in the absence of obstructions (Figure 5). The
common range of slump flow for SCC is 18 to 30 in. (450 to 760 mm). The higher the slump flow,
the farther the SCC can travel under its own mass from a given discharge point and the faster it
can fill a form or mold.

Visual Stability Index.   The VSI test involves visual evaluation of the SCC slump flow spread
resulting from performing the slump flow test. It is used to evaluate the relative stability of
batches of the same or similar SCC mixtures. A VSI rating of 0 or 1 is an indication that the SCC
mixture is stable and should be suitable for placement. A VSI rating of 2 or 3 indicates possible
segregation potential and that the SCC mixture materials and/or proportions need to be adjusted
to ensure stability. An example of VSI rating of 1 (stable mixture) is shown in Figure 2.

Figure 5.   Slump flow test.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

24   Concrete Technology for Transportation Applications

J-Ring.  This test (ASTM C1621) is used to characterize the ability of SCC to pass through
reinforcing steel (Figure 6). The higher the J-ring slump flow, the farther the SCC can travel
through a reinforcing steel obstacle under its own mass from a given discharge point, and
the faster it can fill a form or mold with steel reinforcement. The difference between the J-ring
slump flow and the unconfined slump flow is an indication of the degree to which the passage
of SCC through reinforcing bars is restricted.

T50.  The flow rate of an SCC mixture is influenced by its viscosity. When developing an SCC
mixture in the laboratory, a relative measure of viscosity is useful. The time it takes for the outer
edge of the concrete spread (resulting from the procedure described in the slump flow test) to
reach a diameter of 20 in. (500 mm) from the time the slump cone is first raised provides a rela-
tive measure of the unconfined flow rate of the concrete mixture (see Figure 2). This time period,
termed T50, gives an indication of the viscosity of the SCC mixture. T50 time of 2 seconds or less
typically characterizes a low-viscosity SCC mixture, and a T50 time of greater than 5 seconds is
generally considered a high-viscosity SCC mixture.

Column Segregation.   This is a laboratory test (ASTM C1610) to evaluate SCC mixture
stability and resistance to aggregate segregation. The test is used to develop SCC mixtures with
segregation not exceeding specified limits. The degree of segregation can indicate if a mixture is
suitable for the application.
The static segregation of SCC is determined by measuring the coarse aggregate content in the
top and bottom portions of a cylindrical specimen (or column). This test consists of filling a
26-in.-high (610-mm-high) column with concrete (Figure 7). After 15 minutes from casting the
sample in the mold, the concrete is removed in top and bottom sections and is washed over a
No. 4 (4.75 mm) sieve. The retained aggregate in both sections on the sieve is weighed. A non-
segregating mixture will have consistent aggregate mass distribution between the top and bottom
sections. A segregating mixture will have a higher concentration of aggregate in the lower section.

Strength
Quality SCC mixtures are required to be highly flowable yet stable and cohesive enough to resist
segregation. To achieve such characteristics, the mixture proportioning would include cementi-
tious materials, low water–cement ratio, and the use of a combination of high water-reducing

Figure 6.   J-ring test (courtesy of A. Schindler,


Auburn University).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   25  

Figure 7.  Column
segregation test
tube (42).

and viscosity-retaining admixtures. As a result of a low water–cement ratio and efficient cement
hydration process, much higher compressive and flexural strengths are achieved compared to
conventional concretes. SCC mixtures typically used for fabrication of high-strength precast
members are proportioned with a water–cement ratio of 0.32 to 0.40. Mixtures with water–
cement ratio higher than 0.40 are sometimes employed for cast-in-place and repair applications,
and have strength characteristics similar to conventional concrete (2).

Shrinkage
Autogenous Shrinkage.   Autogenous shrinkage can be high in SCC mixtures made with
a relatively low water–cement ratio, high content of cement, and SCMs that exhibit a high
rate of pozzolanic reactivity at an early age. In particular, the fineness of the cementitious
materials can impact the rate of autogenous shrinkage for the first 28 days. For example, the
finer the slag particles, the larger is the surface area for pozzolanic reaction. This leads to a
faster reaction and greater autogenous shrinkage from loss of moisture and self-desiccation.
Special attention needs to be given to moist-cure SCC at early ages to minimize autogenous
shrinkage (2).

Drying Shrinkage.   Drying shrinkage is related to the water and paste contents as well as
aggregate volume, size, and stiffness. High paste volumes and reduction in aggregate content can
lead to greater potential for drying shrinkage. Paste volumes can be optimized during the SCC
mixture-proportioning process (2).

Plastic Shrinkage.   SCC can be prone to plastic shrinkage cracking given the fact that
these mixtures may exhibit little or no surface bleeding. SCC needs to be protected from rapid
moisture loss, similar to conventional concrete that exhibits little or no surface bleeding, by using
external curing techniques such as fogging, misting, and other curing provisions.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

26   Concrete Technology for Transportation Applications

The higher the water–cement ratio of the mixture, the lower the autogenous shrinkage, and
the higher is the drying shrinkage. Proper engineering and proportioning of the SCC mixture
and early curing would minimize shrinkage at the fresh and hardened stages.

Bond to Reinforcing Steel and Prestressed Strand


The bonding characteristics of properly designed SCC are equal to or better than conven-
tional concrete. SCC flows easily around reinforcing steel and generally bonds well to the
reinforcement. In fact, the bond strength of reinforcing bars in SCC may be up to 40% higher
when compared with conventional concrete (43). This may be due to the lower water content
and the higher powder volume in the SCC mixtures relative to the conventional concrete
mixtures, which reduces the accumulation of bleed water under horizontally embedded
reinforcing bars. However, this bond can be reduced by excessive bleeding and segregation
in poorly designed SCC mixtures, especially in upper sections containing reinforcement (44).

Long-Term Durability
SCC mixtures with SCMs, and low water–cement ratio, are dense and relatively impermeable,
providing long-term durability and resistance to corrosion (20). Also, when a proper air-void
system is developed in the mixture, SCC can exhibit excellent resistance to freezing and thawing
and to deicing salt scaling (45). SCC is also expected to have the same resistance to carbonation
as conventional concrete. With appropriate design and construction provisions, the durability
performance of SCC is not expected to differ from other types of concrete mixtures.

Issues and Challenges


Cost of materials and production of SCC is higher than in conventional concrete mixtures.
This is due to the fact that higher quantities of cementitious materials may be used, specific
aggregate size may be required, and several types of admixtures may be incorporated in the mix-
ture to attain favorable SCC plastic and rigid properties. Also, level of experience and training
could make a difference in the quality of the SCC mixtures and structures.
Mixture segregations can be a major challenge during construction. This may be due to poor
mixture design or variation in the mixture properties among different batches. Poor quality con-
trol during various construction activities may cause cracking and other surface blemishes such
as “bugholes.” Unstable SCC mixtures that are prone to segregation can also exhibit poor surface
quality and surface durability problems when excessive bleeding and surface foaming occur.
With the power of SCC discharge and high rate of flow inside the form, there is a concern
that the reinforcing cage may be shifted from its original position or pushed against the sides of
forms. Also, with the high SCC fluid pressure, the forms may yield or collapse. It is important
that reinforcement be well secured to prevent any shifting and the formwork be designed to
handle SCC pressure during discharge. It is important to provide training to the agency pro-
fessionals and utilize experienced project and concrete producer personnel to make a positive
impact on quality of SCC mixtures and structures.

Internally Cured Concrete


Introduction
Adequate curing using external curing methods will improve concrete strength, volume sta-
bility, permeability, abrasion resistance, and durability (46). Loss of water due to inadequate
curing can slow or end hydration, resulting in lower strength development, more permeable

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   27  

concrete, and shrinkage that may lead to cracking (47). The objectives of external curing include
preventing loss of moisture through evaporation and hydration, replenishing the moisture lost
to evaporation, and maintaining favorable heat to allow continuation of the hydration process
(46, 47). Because of the need to meet construction schedule requirements, external curing is
often discontinued prematurely, resulting in a negative impact on the concrete quality.
Internal curing is another method of providing moisture inside the concrete to support
hydration without increasing the water–cement ratio and also has the benefit of reducing self-
desiccation and the accompanying stresses that can result in early age cracking (3, 48, 49, 50).
ICC mixtures include prewetted lightweight aggregate, superabsorbent polymers (SAPs), or
other agents that release moisture to the concrete paste to facilitate hydration after traditional
curing measures have been terminated. An illustration of the differences between internal curing
and external curing is shown in Figure 8 (51).
As illustrated in Figure 8, conventional curing methods provide moisture only to a limited
zone of paste within the upper few millimeters of the concrete surface. The distribution of the
prewetted agents throughout the concrete allows for wide dispersion of the additional moisture
to facilitate curing. Note that internal curing does not provide enough moisture to counter-
act loss of moisture through evaporation and therefore is not a substitute for external curing.
Appropriate provisions for external curing need to always be utilized for ICC mixtures (52).

Need for the Technology


The need for longer lasting, more durable transportation infrastructure requires design and
construction of concrete components with low permeability and more resistance to cracking.
In recent decades, concrete mixtures have been increasingly utilizing lower water–cement ratio
and SCMs. HSC mixtures often combine a low water–cement ratio with high cementitious
materials contents. These concrete mixtures are ideal candidates to benefit from the additional
moisture provided by internal curing, since mixtures with lower water–cement ratio tend to be
more susceptible to autogenous shrinkage due to less water available to support the hydration

Figure 8.   Illustration of the mechanisms of internal and


external curing (51).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

28   Concrete Technology for Transportation Applications

(52, 53, 54). In fact, research suggests that mixtures with water–cement ratio less than 0.42
do not have adequate water to fully hydrate the cementitious materials (25). Cementitious
binder systems that utilize SCMs can also benefit from increased hydration provided by the
longer curing durations supported by internal curing (52). The additional hydration supplied
by internal curing moisture has been shown to provide further benefits that include reduced
cracking potential, enhanced durability, and sometimes increased strength (3, 49, 56, 57).

Mechanisms of Internal Curing


Internal curing can be accomplished using a number of materials that have a high absorption,
including prewetted crushed concrete fines, SAPs, perlite, prewetted wood fibers, and light-
weight aggregates (52). Lightweight aggregates provide structural strength to a mixture, while
many other materials capable of providing internal curing benefits do not (49, 55), and have
therefore been the material most commonly used in research and implementation of internal
curing. For these reasons, use of prewetted lightweight aggregates to support internal curing is
the focus of this report.
To provide internal curing moisture, prewetted lightweight fine aggregate can replace a por-
tion of the mixture fine aggregate, and/or prewetted coarse lightweight aggregate can replace a
portion of natural coarse aggregates. The effectiveness of lightweight aggregates in facilitating
internal curing is dependent on several factors, including (1) the amount of water included in
the prewetted lightweight aggregate, (2) the spacing of the aggregate particles, (3) the character-
istics of the pore structure of the aggregates, and (4) the strength and shape of the lightweight
aggregate (54, 58).
Lightweight fine aggregate has been demonstrated to be more effective in promoting inter-
nal curing than lightweight coarse aggregate because [according to ACI (308-213)R-13] “it is
distributed more fully throughout the concrete mixture and, therefore, the particle surfaces
are closer to the hydrating cement particles” (52). In Figure 9, a representative image from an
internal curing simulation model shows schematically the paste volume potentially protected
via internal curing using prewetted lightweight fine aggregate.
As cement hydrates, the products of hydration occupy less volume than the reacting materials,
resulting in a net chemical shrinkage. This chemical shrinkage, which increases proportionally

Figure 9.   Example of two-dimensional image from


internal curing simulation program showing predicted
protected paste, using 70% aggregates by volume and
20% replacement of the fine aggregate with prewetted
lightweight aggregate (54).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   29  

with the degree of hydration, results in self-desiccation as pores within the microstructure lose
water below the level of saturation (52). If curing water is not available, these pores in the cement
paste will instead be filled with vapor. Curing water supplied internally from the pores in the
lightweight aggregates provides moisture to support the cement hydration until the equilibrium
point is reached between moisture in the lightweight aggregates and the surrounding paste (59).
Key to ensuring that water delivered into the concrete by the prewetted lightweight aggregate
is that the water release is delayed. The characteristics of the internal pore structures of manu-
factured lightweight aggregates vary based upon source geology and manufacturing process.
However, in general, the pores of lightweight aggregates are of a size that will allow water to be
held within the aggregate during mixing (absorption) but released from the aggregate back into
the paste after setting (desorption) (60). If reservoirs of water within the lightweight aggregates
are available and well dispersed, the water within the reservoirs will be desorbed into the cement
paste via capillary suction, restoring the internal humidity of the concrete to saturated condi-
tions (61). Progression of hydration will continue to narrow the capillary pores as new hydra-
tion products are formed. This will result in increased capillary suction from the lightweight
aggregate reservoirs, allowing the internal curing to progress until all cement is hydrated or
equilibrium relative humidity is reached between the capillary pores and the lightweight aggre-
gate reservoirs (61).

Mixture Proportioning of Internally Cured Concrete


ACI (308-213)R-13, “Report on Internally Cured Concrete Using Prewetted Absorptive
Lightweight Aggregate,” provides extensive guidance on development and implementation of
ICC mixtures (52). The quantity of lightweight aggregates utilized to provide adequate moisture
to effectively support internal curing is a function of the type of lightweight aggregates utilized
(size, degree of saturation, desorption characteristics), the amount of lightweight aggregates
utilized, the concrete water–cement ratio and type of binders utilized, and the type and extent of
external curing provided (52). When proportioning concrete mixtures, an equation for the mass
of lightweight aggregates or nomographs provided in the document can be used.
Studies have indicated that virtually all lightweight aggregates produced in the United States
demonstrate adequate desorption characteristics to support internal curing (60), and informa-
tion such as desorption can be obtained from the manufacturers. A review of the literature
indicates that most internally cured mixtures used in the laboratory and field utilize substitution
percentages less than 50%, with most field studies tending to utilize volumetric substitution per-
centages ranging from 20% to 35%. Regardless of the proportioning approach, ACI (308-213)
R-13 emphasizes preparing trial batches and performing tests to verify that the required concrete
properties can be obtained using the selected absorptive material and mixture proportions (52).

Construction
Prewetting of the Lightweight Aggregates
Batching, mixing, transporting, and placing of ICC do not differ significantly from the pro-
cess used for conventional concrete mixtures (52). Key to successful implementation of ICC is
adequate prewetting of the lightweight aggregates prior to batching. Most lightweight aggregates
with open textured surfaces can be prewetted by sprinkling stockpiles with water for a sufficient
duration of time. The rate at which the lightweight aggregates will absorb the water will depend
on a number of factors including the characteristics of the lightweight aggregates (in particular,
aggregate absorption), the water application rate, stockpile characteristics, and environmental
conditions including temperature and humidity (52). Prewetting needs to be performed for a
duration to ensure that the lightweight aggregate has exceeded the saturated-surface-dry (SSD)

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

30   Concrete Technology for Transportation Applications

condition. Use of lightweight aggregate that has not achieved the SSD condition in the concrete
mixture may have adverse effects, including slump loss and placement issues during pumping
and finishing (62).
The SSD state is difficult to define for lightweight aggregates, and it is therefore consid-
ered “best if the lightweight aggregates can be provided in a known and maintainable state of
moisture equilibrium” (3). Similar to use of other types of aggregates in concrete, the moisture
content and absorption of the prewetted lightweight aggregate needs to be accounted for during
the batching process. Provisions to adequately prewet the lightweight aggregates will ensure that
variability with slump loss and unit weight as well as issues with pumpability, segregation, and
finish quality will not occur (62).
In the laboratory, methods to prewet and drain the lightweight aggregate include submerging
lightweight aggregate in a container of water and allowing it to soak for at least 24 hours (63, 64, 65).
In the field, various methods have been used to prewet the aggregate, with most suppliers choos-
ing to stockpile the material and provide sprinklers for the specified duration of time. The stock-
pile needs to be configured in a manner that allows adequate drainage (Figure 10), and a uniform
moisture state needs to be achieved by turning the stockpile. Testing needs to be performed to
ensure that moisture corrections can be made to the batch weights.
Historically, ASTM C1761 has been utilized to define the SSD state for lightweight aggregate
and to compute moisture corrections. This method, also known as the “paper towel method” is
time-consuming and can be subject to variability due to differences in paper towel characteris-
tics and operator inconsistency (Figure 11). An alternative method to characterize the moisture
state of prewetted lightweight aggregate is the centrifuge method, in which a sample of the
material is spun in a centrifuge for 3 minutes at 2,000 rpm (Figure 12). This method has been
shown to provide rapid, accurate results that correlate well with ASTM C1761 (66).

Performance Properties
Fresh Concrete Properties
When proper proportioning considerations have been made and the lightweight aggregate
adequately prewetted, fresh concrete properties tend to be relatively unaffected by internal curing.
Impact on slump is generally negligible, although some studies have indicated increased work-
ability when prewetted aggregate was used (67). It has been suggested that a portion of absorbed
water may leave the aggregate and become integrated into the concrete mixture during mixing,

Figure 10.   Lightweight aggregate being prewetted


by sprinkler (circled), with floor sloped to drain water
from the stockpile.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   31  

Figure 11.   Paper-towel drying method.

hauling, and placement, thus aiding in maintaining workability and surface finish quality (62).
Field studies have not reported any complaints from placing or finishing personnel (68, 69).

Hardened Concrete Properties


The most significant improvement in performance of hardened concrete with the use of ICC
is in the reduction of autogenous shrinkage at early ages (59, 61, 65), as shown in Figure 13.
In some mixtures, use of prewetted lightweight aggregates for internal curing was shown to
virtually eliminate autogenous shrinkage (49).

Figure 12.   Centrifuge drying method (66).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

32   Concrete Technology for Transportation Applications

Sealed
33.0%k
29.3%k
23.7%k
18.3%k
14.3%k
11.0%k
7.3%k
0.0%

Figure 13.   Reduction in autogenous shrinkage for internally


cured mortars (50).

An important benefit of ICC is the reduced shrinkage and the associated internal strains,
which, in turn, reduce the early age cracking potential (48, 52, 56, 58, 70). A study for the Colorado
DOT indicated that the reduced autogenous and drying shrinkage reduced residual stress
buildup in restrained shrinkage testing of mixtures with low water–cement ratios, although this
effect was decreased in mixtures with higher water–cement ratios (71). Using the ASTM C1581
restrained ring test, Henkensiefken et al. (63) found that cracking was essentially eliminated when
prewetted lightweight aggregate was used at replacement rates greater than 23.7% (Figure 14).
Use of prewetted lightweight aggregates for internal curing has been shown to reduce the
permeability of concrete, enhancing resistance to deleterious substances such as chlorides and
sulfates. Although improvements in permeability may not be observed at early ages, increased
hydration of cementitious materials at later ages, likely supported by the internal curing moisture,
improves the quality of the interfacial transition zone (ITZ) between paste and aggregate, leading
to lower permeability (47, 67, 70). Overall, the potential durability of ICC using prewetted light-
weight aggregates has been shown to be improved due to the increased hydration of cementitious
materials, reduced shrinkage, and potentially increased compressive strength (3).
An increase in compressive strength is typically not the goal of using ICC. However, some
researchers have found slightly higher compressive strengths in ICCs and mortars (49, 56, 72). The

Sealed
23.7%k
14.3%k
11.0%k
7.3%k
3.8%k
0.0%k

Figure 14.   ASTM C1581 restrained shrinkage test results,


showing a correlation between age at cracking and
prewetted lightweight aggregate replacement rate (63).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   33  

potential for increased strength will depend on factors including strength and degree of saturation
of the lightweight aggregates and improved strength of the paste surrounding the lightweight aggre-
gates that has benefited from the internal curing moisture. Use of prewetted absorptive materials
other than lightweight aggregates has been shown to decrease compressive strength (49).
Flexural strength has been shown to be improved by internal curing (73, 74). The modulus of
elasticity of lightweight aggregates is typically lower than that of natural aggregates. However, when
using prewetted lightweight aggregates for internal curing, the change in concrete’s modulus is
not linear. At low replacement percentages, the modulus of elasticity has been shown to increase
slightly. However, at higher replacement levels [greater than 100 lb/yd3 (59.3 kg/m3)], a reduction in
concrete’s modulus of elasticity has been demonstrated (75). Studies on creep behavior of concrete
internally cured with lightweight aggregates have not been conclusive. A study found that creep
often decreases but could potentially increase (76). ACI (308-213)R-14 recommends trial batching
and testing of mixtures for creep when creep is a critical performance criterion (77).
The unit weight of ICC will generally be slightly lower than that of conventional mixtures.
However, at replacement levels typically used in internal curing applications, the unit weight does
not fall below 135 lb/ft3 (2,160 kg/m3), the density at which the mixture is no longer classified as
normal-weight concrete and additional structural considerations are generally required (77, 78).
With respect to freeze-thaw resistance, a study by the Colorado DOT on ICC mixtures focused
on evaluating the potential for enhanced freeze-thaw durability (71). Findings of this study indi-
cated that ICC can exhibit freeze-thaw performance comparable to that of conventional mixtures,
provided that excessive internal curing water and high water–cement ratios are avoided (69).

Implementation
A number of state agencies are moving forward with implementation of ICC for bridge decks.
Internally cured pavements have been constructed on several municipal streets (primarily in
Texas) and, perhaps most notably, the Dallas Intermodal Terminal constructed in 2005 (75),
which has been cited as providing excellent performance (79). Interest has also been shown in
the use of ICC for overlays (80, 81) and cementitious repair materials (82).
To date, ICC has been used on bridge decks in New York, Indiana, Utah, Virginia, Georgia, Ohio,
Colorado, and North Carolina (Figure 15) (68, 69, 71, 72, 80, 83). Studies of field installations

Figure 15.   Internally cured concrete bridge deck in


Durham, North Carolina, being placed by pumping.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

34   Concrete Technology for Transportation Applications

of ICC bridge decks largely report success. The Indiana DOT constructed four bridges utiliz-
ing prewetted lightweight aggregates to internally cure high-performance concrete mixtures
and found that early-age autogenous shrinkage was reduced by 80% compared to non-ICC
mixtures (69).
Two internally cured bridge decks were constructed in northern Utah, along with two com-
panion conventional bridge decks. At 28 days, the two types of concretes had similar compressive
strengths; however, the ICCs exhibited lower permeability when tested using the rapid chloride
permeability test. After several months, cracking was observed in the conventional decks while
no visible cracks were found in the internally cured bridge decks (72).
Most laboratory studies indicate that ICC members would have the potential to require less
maintenance and provide extended service life compared to conventional concrete. However,
a more complete assessment of its true potential will become evident over time as field imple-
mentation sites age. Remember that although internal curing is a promising tool that can
improve concrete properties, it cannot be expected to compensate for deficiencies in design and
construction that result in cracking and other issues (68).

Specifications and Guidelines


Extensive guidance on development and implementation of ICC mixtures using prewetted
lightweight aggregates is presented in ACI (308-213)R-13 (52). This document provides infor-
mation on the materials, mechanisms, and benefits of internal curing, along with recommenda-
tions for mixture proportioning and batching procedures. Sustainability benefits associated with
ICC are also discussed. The National Institute of Standards and Technology maintains a robust
online database of resources on internal curing. An extensive bibliography is provided, along
with resources including mixture proportioning tools and internal curing simulation results (84).
The Expanded Shale, Clay, and Slate Institute (ESCSI) has published a guide specification for
ICC (85). This guide specification, prepared by the industry, suggests the need for prequalifica-
tion of the lightweight aggregate (meeting ASTM C330 and ASTM C1761), provides guidance on
prewetting and testing of the lightweight aggregate, and outlines a procedure for computing the
quantity of prewetted lightweight aggregate required for internal curing. Construction require-
ments outlined in this guide specification typically refer the reader to the contract documents.
Some states have developed specifications for internal curing. The New York State DOT has had
extensive experience in implementing ICC mixtures in bridge decks and has prepared several spec-
ifications as well as a specified procedure for determining the moisture content of the fine aggre-
gate on site (Test Method No. NY 703-19, “Moisture Content of Lightweight Fine Aggregate”).
New York State DOT Specifications 557.51XX0018, 557.52XX0018, and 557.54XX0018 cover ICC
for superstructure slabs as well as approach slabs (86). A consistent substitution proportion of 30%
prewetted lightweight aggregates has worked well for the many ICC bridge decks constructed in
New York, as has the stockpile prewetting provisions provided in the specifications (70).
The West Virginia DOT has a special provision “Structural Concrete Internal Curing,” which
provides guidance on material properties as well as guidance for mixture proportioning and use
of shrinkage-reducing admixtures to enhance performance of these mixtures (Appendix C).
Modified procedures for determining the absorption of the lightweight fine aggregates include
testing from an independent laboratory to prequalify an aggregate for use. Procedures for com-
puting the amount of dry lightweight fine aggregate are provided in the special provision, as well
as the target cement factor, maximum water content, and entrained air content. Requirements
for handling, measuring, and batching of materials are also provided, including detailed guid-
ance on stockpile management of the lightweight fine aggregate (87).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   35  

Requirements for use of internal curing in the Virginia DOT’s bridge concrete are provided
in a Special Provision entitled, “Low Cracking Bridge Deck Concrete.” ICC mixtures can be
used in both bridge decks and substructures. The Virginia DOT special provisions’ guidance
and requirements are minimal and allow for input from the manufacturer of the lightweight
aggregate.

Issues and Challenges


Availability of lightweight aggregate locally is a challenge in many states, because the lightweight
aggregate is primarily a manufactured material. This will affect the price of the aggregate and
will result in a higher cost to implement ICCs. From a concrete producer’s standpoint, internal
curing requires introduction of an additional material to their operations, often requiring addi-
tional space, storage bins, and handling equipment. This may influence the capability of some
producers to supply ICC without some modifications to their plant or operation. Additional
QA/QC associated with development and control of mixtures may also affect the cost of ICC.
Agency awareness, experience, and guidance on the proper prewetting of the lightweight fine
aggregates can be seen as an issue requiring additional training and oversight. Development of
material handling guidelines by state agencies to assist concrete producers and project personnel
would be helpful.

Ultrahigh-Performance Concrete
Introduction
UHPC is a portland cement–based product that includes SCMs, well-graded fine sand, a
high dosage of fiber reinforcement (usually steel) as well as superplasticizers and other admix-
tures, and a very low water–cement ratio (4). Some mixtures may also include coarse aggregates,
with small particle sizes, in quantities that are much less than that those used in conventional
concrete. The ingredients may be mixed on site or batched and delivered using truck mixers.

Benefits
UHPC is highly flowable and self-consolidating when discharged, develops very high com-
pressive and tensile strengths, and exhibits durability. UHPC mixtures have developed ultra-
high compressive strengths greater than 21,700 psi (150 MPa) and have exhibited postcracking
tensile strengths as high as 720 psi (5 MPa). These mixtures also have very low permeability,
which reduces penetration of harmful liquids and greatly enhances durability compared to
conventional concrete (88). Such favorable mechanical and durability properties have made
UHPC an ideal product for use in joints connecting prefabricated components of bridges in
ABC projects (Figure 16) and in bridge deck construction, overlay, and repairs (4).
Field casting of UHPC connections results in stronger and more durable connections and
provides better long-term performance compared to connections cast with conventional con-
crete. The superior mechanical properties of UHPC allow more optimized design of connection
dimensions and reinforcement that promotes both speed and ease of construction.

Applications
UHPC became commercially available in the United States in 2000. Since that time, research
projects and field deployment have demonstrated the material capabilities in many applica-
tions related to bridge construction, replacement, and overlays (88). These applications included

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

36   Concrete Technology for Transportation Applications

Figure 16.   UHPC placement in connections between


prefabricated bridge components (courtesy of
New York State DOT) (4).

connections between precast or prestressed bridge elements including girders, pile cap closure
pours, and prefabricated deck components (Figures 17 and 18). Other applications included
thin overlays of deteriorated bridge decks and fabrication of precast structural elements.
The majority of applications have been in connection joints between prefabricated bridge
components during construction or deck replacement, as shown in Figures 17 and 18. In these
applications, UHPC mixtures have demonstrated superior mechanical and durability properties
compared to conventional concrete mixtures.
UHPC mixtures are used in joints as closure pours to accelerate connection of bridge com-
ponents. They require shorter splicing of reinforcing bars of adjoining elements, develop high
early strength for early form removal, and allow minimum shrinkage for long-term durability.
The connection design using UHPC includes simple lap splicing of reinforcing bars where the
tension development length of the reinforcement is much shorter than in conventional concrete
connections, as shown in Figure 19 (4). In this application, the overall cost of using UHPC is
relatively small because of the small quantity of material needed to cast the connections (88).
This is important when considering that the cost of UHPC mixes is very high compared to con-
ventional or HSC mixes (89).

Materials and Mixtures


The dry ingredients of UHPC include a large quantity of cementitious materials which may
exceed 1,500 lb/yd3 (890 kg/m3), a blend of well-graded fine and coarse sand, and, in some
mixtures, a small quantity of coarse aggregate of small particle sizes (88). A very high volume

Figure 17.   Schematic showing cross section of


precast beams with UHPC closure pours (4).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   37  

Figure 18.   Prefabricated bridge deck components with UHPC connections (4).

of steel fibers (1% to 2% of mix volume) is an essential ingredient to provide the necessary
reinforcement for enhanced compressive and tensile strengths and reduced shrinkage (4, 89,
90, 91). HRWRs and a variety of other admixtures are included to produce a highly flowable and
self-consolidating mixture capable of filling narrow forms with congested reinforcement. The
water–cement ratios for UHPC mixtures are typically not greater than 0.25 (4).
Most of the UHPC mixtures are proprietary with the dry ingredients prepackaged by the pro-
ducers. At the project site the dry ingredients are mixed with water, fibers, and admixtures using
a portable mixer. The mixture ingredients may also be batched in truck mixers and transported
to the project site. The cementitious contents of most mixtures include more than 1,000 lb/yd3
of portland cement with low tricalcium aluminate (C3A) (to control the mixture temperature)
and silica fume powder at 10% or more by weight of the portland cement (4). Other pozzola-
nic materials, such as slag, metakaolin, fly ash, and limestone powder, have also been used in
addition to or as a replacement of silica fume and cement to control the mixture temperature,
enhance durability and sustainability, and reduce cost (4, 89, 90, 91).

Conventional Detail UHPC Detail

Figure 19.   Conventional and UHPC connections between prefabricated deck elements (4).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

38   Concrete Technology for Transportation Applications

Fine aggregates including quartz, limestone, and basalt aggregates are proportioned and sized
in the dry constituents to facilitate the flowability of UHPC mixtures and increase the compressive
strength. As mentioned previously, coarse aggregates are sometimes included in UHPC mixes to
provide cost savings. The coarse aggregate tends to be relatively small [0.25 in. (6 mm) or less] (4).
However, in a recent study of fiber balling in a nonproprietary UHPC mixture, 0.5-in. (13-mm)
coarse aggregate was used to break up fiber balling in the mixture with good results (91).
The steel fiber most commonly used in UHPC applications is fine fiber with 0.008-in.
(0.2-mm) diameter and a 0.5-in. (19-mm) long straight fiber (4). However, it has been reported
that a high volume of fine fibers can create high potential for fiber balling in UHPC mixtures (91).
This challenge can be more pronounced in large-volume mixtures produced using drum mixers
in concrete trucks. The proposed solution to the balling problem was to use a blend of 50% each
of fine and medium-size fibers when batching large UHPC mixtures in concrete trucks. This
solution does not seem to have had an effect on fresh and hardened properties of the mixtures.
A variety of admixtures are used in UHPC mixtures to provide favorable fresh and hardened
properties. These admixtures commonly include accelerators, polycarboxylate, and phosphonate-
based superplasticizers (4). These and other admixtures provide the mixture’s high flowability,
self-consolidation, and early strength properties.

Construction
Mixing and Placement
UHPC mixtures can be batched in conventional rotary pans or drum mixers, including con-
crete truck mixers. However, in most applications, UHPC mixtures are field-batched in a rotary
pan mixer (Figure 20) (4, 88). Prepackaged dry components are first discharged into the mixer.
Then water, admixtures, and fibers are added to the rotating mixer until the proper UHPC for-
mulation is produced that meets the specific project requirement. This is a process similar to that
of many of the proprietary grouts and patching materials used in bridge repairs and overlays (4).
The high cementitious material contents and longer mixing time, compared with those of
conventional concrete mixtures, will cause the temperature of the UHPC mixture to rise during
mixing and placement. This temperature increase may require the use of chilled water or ice
cubes to control the mixture temperature (4). Lowering the mixture temperature increases the
flow rate and self-consolidation of UHPC during placement.
Nonproprietary UHPC mixtures have been developed and deployed in the field in Michigan and
New Mexico (89, 90, 91). These were engineered mixtures that used materials and products com-
monly available to the concrete producers and contractors. The UHPC mixtures included portland
cement with low C3A contents, silica fume, slag, and fly ash, as well as a blend of well-graded fine

Figure 20.   On-site mixing (92).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   39  

quartz sand. In some mixtures, a small quantity of coarse aggregate, small and medium-size fibers,
and HRWR admixture were also added. Prior to their deployment in the field, trial batches were
prepared to optimize the mixture ingredients and ensure proper flow characteristics when placed,
and also to verify that the specified strength and durability could be achieved (91).

Tests and Properties


Flow Characteristic
QC tests for UHPC are similar to those used for conventional concrete or mortar. Both fresh
and hardened concrete properties are measured (93). The flow of UHPC mixtures is frequently
measured using ASTM C1437, Standard Test Method for Flow of Hydraulic Cement Mortar
(94). The test is performed immediately after mixing to assess UHPC mixture appropriateness
for casting. The favored flow range for UHPC mixtures used in bridge applications is 7 to 10 in.
(178 to 254 mm) (4). Dryer UHPC mixes used in pavement overlays would be tested using the
conventional concrete slump test according to ASTM C143.

Strength
Similar to most other concrete mixtures, compressive strength testing is used to evaluate
the quality of hardened UHPC. The test is performed according to a modified version of
ASTM C39, “Standard Test Method for Compressive Strength of Cylindrical Concrete.”
Because of the high compressive strength of UHPC, the test method is modified to include an
increased load rate of 150 psi/s (1 MPa/s). Also, since higher failure loads are anticipated for these
high-strength mixtures, 3-in. by 6-in. (75-mm by 150-mm) test cylinders are used to accom-
modate capabilities of most testing machines (88). Based on FHWA assessment, preblended
UHPC mixtures with 2% fibers and cured in field-type conditions have exhibited compressive
strengths exceeding 14,000 psi (97 MPa) after 4 days and 21,000 psi (145 MPa) after 28 days (4).
ASTM C109 “Standard Test Method for Compressive Strength of Hydraulic Cement Mortars”
[using 2-in. (50-mm)-cube specimens] can also be applied to UHPC. For UHPC qualification,
the New York State DOT requires compressive strength test results for a minimum of sixty-
four 2-in. (50-mm) cubes, tested at 4, 7, 14, and 28 days. The minimum compressive strength
requirements are 14,300 psi (100 MPa) at 4 days and 21,800 psi (150 MPa) at 28 days (95).

Durability and Shrinkage


Durability and shrinkage tests are also performed during the mixture verification stage but
are not necessarily required by the highway agency as QC tests (84). These tests include chloride
ion penetration (ASTM C1202), freeze-thaw resistance (ASTM 666, procedure A) and shrink-
age (ASTM C 157) tests. In bridge applications, the acceptance criteria are as follows: chloride
penetration of 250 coulombs or less after 28 days; freeze-thaw resistance durability factor of at
least 95% after 300 freeze-thaw cycles; and shrinkage of 800 macrostrain or less after 28 days (4).

DOT Implementation
As of December 2016, five states including Idaho, Iowa, New Jersey, New York, and
Pennsylvania have made UHPC a standard practice on bridge projects for connections between
pre­fabricated components. Nineteen other states and Washington, DC, are either using UHPC
in bridge construction projects or making plans to standardize the use of the technology (93).
States in which UHPC connection projects are in various application stages, from research
to deployment, include California, Connecticut, Florida, Idaho, Iowa, Massachusetts, Ohio,
Oregon, and South Carolina. In addition to the United States, UHPC has been used in high-
way infrastructure in Australia, Austria, Croatia, France, Germany, Italy, Japan, Malaysia, the
Netherlands, New Zealand, Slovenia, South Korea, and Switzerland (88).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

40   Concrete Technology for Transportation Applications

Specifications
Use of UHPC requires the development of material and construction specifications. The
material specifications define the constituent properties, testing protocol(s), testing frequen-
cies, performance criteria, and unit of payment. It is important that construction specifications
provide guidance on field-related activities that may affect the performance of UHPC, including
considerations such as material storage, adequacy of form work, mixing, and in-field testing,
placement, and curing (4).

Issues and Challenges


Issues
The joint faces of precast concrete components must be intentionally roughened to a 0.25-in.
(6.4-mm) amplitude prior to filling with UHPC. Roughening of precast concrete allows increased
UHPC bond at the joints and reduces the shear stresses carried by the discrete steel reinforce-
ment crossing the interface (4).
It is important that flow distances of field-cast UHPC mixtures in long joint connection be
limited to 10 ft (3 m). Long flow distances around the reinforcement can interrupt the disper-
sion of the fiber reinforcement, which could reduce the mechanical resistance of the UHPC (84).
Also, to avoid possible mixture segregation, adjustments to length of casting must be made when
coarse aggregate is used.
A proprietary-based product can only be used if it undergoes the highway agency’s appro-
priate acceptance testing protocols and is placed on a preapproved materials/products list.
Sometimes an “approved equal” provision is included in the contract plans to allow unidentified
suppliers an opportunity to demonstrate whether their product meets the criteria of the contract
plans and specifications (4).
Successful performance of UHPC connections requires that each stage of construction be
completed in a timely and appropriate manner. The owner, the material supplier, the inspectors,
and the contractor must work together to ensure success (4).
At present, very few producers have experience with UHPC for precast or cast-in-place appli-
cations. Information needs to be made available so that they become aware of the issues involved
in dealing with UHPC mixtures and construction. For example, precast plant personnel need to
be aware that longer mixing times than in conventional concrete mixers are necessary due to the
high cementitious materials content and low water–cement ratio, longer set times due to high
admixture dosage, and modified curing treatments (4).

Challenges
In an international conference on UHPC in 2016, a poll was conducted to gauge partici-
pants’ responses to a question related to obstacles that were preventing wider implementation
of UHPC. The responses included lack of design guidance, industry resistance to change or lack
of understanding of the technology, high cost compared to conventional concrete, and other
minor issues (96).
To encourage greater implementation of UHPC in the highway infrastructure, the following
have been identified as needs:
• Studies showing the cost-effectiveness of UHPC in various applications,
• Design and construction guide specifications for structures made with UHPC,
• Research to address some of the missing information needed in the structural design,
• Standard test methods and material specifications for UHPC,

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   41  

• Production procedures for precast and cast-in-place construction,


• Broader geographic distribution of demonstration projects, and
• Wider distribution of technical information.

Temperature Control of Mass Concrete


Introduction
Mass concrete is defined as any volume of concrete with dimensions large enough to
require that measures be taken to manage the generation of heat from hydration of the
cement and associated volume change to minimize cracking (97). Mass concrete is used in
dams, navigation locks, nuclear plants, and power houses. In transportation structures, mass
concrete is used in large footings, mat foundations, bridge piers and columns, and some
maritime structures (98).
The generally large size of mass concrete structures creates the potential for significant rise
in internal concrete temperature in the structure and significant temperature differentials
between the interior and the outside surface of the structure during the early stages of hydration.
The accompanying volume-change differentials and restraint result in tensile strains and stresses
that may cause cracking that is detrimental to the integrity and durability of the structure and
thus can shorten its service life (99).
A mass concrete temperature control plan is required by many state DOTs for massive
structural members in bridges and other transportation structures, and by specifiers of other
mass structures such as dams, power plants, and navigation locks. The TCMC is developed
and implemented during construction to control the core temperature and temperature dif-
ferential in the mass concrete structure to avoid the negative consequences of mass concrete
heat (28, 98).

Objective
The objective of temperature control plans in mass concrete members is to control the
rise in internal temperature of concrete and maintain the temperature differential between
the interior and outside surface below threshold levels that may cause cracking (28).
To accomplish such objective, these plans address the selection of type and quantities of
cementitious materials, admixtures, and aggregates; precooling mixture ingredients; cooling
during the mixing process; postcooling of in-place concrete by embedded pipes; and surface
insulation (5).

Applications
TCMC is used in dams, navigation locks, nuclear and other power plants, and foundations
for large building structures. In transportation projects, a TCMC plan is developed and executed
during construction of large footings, mat foundations, and bridge piers and columns that meet
the definition of mass concrete. The purpose of the plan is to ensure that the core tempera-
ture and temperature differential between the core and edges of the structure remain below the
thresholds specified by the state DOT.
Most state DOTs require that a TCMC plan be developed for projects that involve construc-
tion of mass structural members. This is based on the survey responses by DOTs (Appendix B).
The Florida DOT, for example, requires that the plan be administered by a specialty engineer to
ensure proper execution of the TCMC plan in mass concrete projects (28).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

42   Concrete Technology for Transportation Applications

Elements of TCMC Plan


The TCMC plans may achieve effective temperature control by incorporating four elements:
(a) proper selection and proportioning of mixture materials, (b) precooling of aggregates
and mix water, (c) cooling of the concrete during the batching process and placement, and
(d) postplacement cooling and insulation (5). These elements may be used collectively, or,
in many cases, only a select number of these elements may be chosen to accomplish the tem-
perature control. The extent and duration of the plan depends on the member shape and
dimensions, ambient weather conditions, aggressiveness of the surrounding environment,
and specification provisions.
The elements of TCMC include but are not limited to, the following details:
• Selection of cement type and content to avoid high heat of hydration. The mixture is also
likely to incorporate SCMs such as fly ash and slag to replace portions of the cement to lower
heat generation in the mass concrete structure.
• Precooling of aggregates and mixture water to achieve a lower concrete temperature when
cast in the structure.
• Construction management and controls, whereby concrete placement timing, scheduling,
and procedures are adjusted and construction actions and equipment that introduce addi-
tional heat in the freshly placed concrete are avoided.
• Postcooling, where embedded cooling tubes and coils are installed to limit the temperature
rise in the structure core, and insulation is applied on the edges and exposed surfaces to pre-
vent rapid heat escape and steep decrease in surface concrete temperature thereby controlling
the temperature differential.

Materials and Mixture Proportioning


Mass concrete (MC) is composed of the same ingredients as conventional concrete. The
objective of the MC mixture design is to produce concrete that results in low temperatures
during early ages while achieving the agency-required strength at the specified ages and
durability.

Cement
Portland cement Types I, II, I/II, and V (ASTM C150), and blended cement Types P, IP, S,
IS, I(PM), and I(SM), covered by ASTM C595 are commonly used in MC mixtures. One effec-
tive measure to lower the heat of hydration is to limit the quantity of cement to a relatively low
amount. This will result in lower heat of cement hydration and will contribute to lower rise in
concrete temperature.

Supplementary Cementitious Materials


The use of SCMs such as Class F fly ash and GGBFS in MC mixtures as partial replacement
of portland cement not only contributes to low permeability and long-term durability but also
lowers the amount of heat liberated during hydration of the cement. This requirement of SCMs
in MC is illustrated in DOT specifications. For example, Florida DOT Specification 346 requires
that MC mixtures include up to 50% fly ash or 70% GGBFS as replacement by mass of the
cement quantity (28).
Class F fly ash with high silica content and low calcium content reacts chemically with the
calcium hydroxide liberated during the hydration of cement. This is a very slow pozzolanic
reaction that produces more hydration products, but at a lower rate of heat liberation, and thus
contributes to a lower overall concrete temperature (14). However, Class C fly ashes contain

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   43  

high calcium content and tend to react in a manner similar to that of cement, and thus
contribute to early concrete strength gain, but they are not as effective as Class F ash in
lowering concrete temperature (5).
Slag added to MC mixtures in large quantities as cement replacement reacts similarly to
cement but at a slower rate. The slower hydration of slag reduces the rate of heat generation.
Also, by substituting a large proportion of the cement, the use of slag further contributes to
reducing the rise in overall temperature of the concrete (5, 14).

Chemical Admixtures
Admixtures provide important benefits to mass concrete in its plastic state by increasing
workability, reducing water content, or both. Also, chemical admixtures can be used for retard-
ing initial setting by slowing the hydration process and thus lowering rate of heat generation.
However, their impact in heat reduction is minimal and is limited to early hours prior to final
setting of the concrete.

Construction Practices for Temperature Control


Temperature control measures for a small structure may be as simple as restricting placement
operations to cool periods at night or during cool weather. At the other extreme, some projects
can be large enough to justify a wide variety of separate, but complementary, control measures
in a well-prepared and -executed temperature control plan. Practices that have evolved to con-
trol temperatures and consequently minimize thermal stress and cracking are listed below (98):
• Cooling batch water;
• Replacing a portion of the batch water with ice;
• Shading aggregates in storage and in conveyors;
• Spraying aggregate stockpiles for evaporative cooling;
• Injecting liquid nitrogen into the mixture;
• Scheduling placements when ambient temperatures are lower, such as at night or during
cooler times of the year;
• Controlling rapid surface cooling of the concrete with insulation;
• Avoiding thermal shock during form and insulation removal;
• Protecting exposed edges and corners from excessive heat loss; and
• Effectively monitoring concrete and ambient temperatures.

Precooling Mixture Materials


Minimizing the temperature of the fresh concrete at placement is one of the most important
and effective ways to minimize the maximum core and differential temperatures and reduce
thermal stresses and cracking. In mass structural members, each 10°F (6°C) reduction of the
placing temperature below average air temperature will lower the peak temperature of the hard-
ened concrete by approximately 4°F to 6°F (2°C to 3°C) (99).
The temperature of the mixed concrete is influenced by each component of the mixture, and
the degree of influence depends on the individual component’s temperature, specific heat, and
proportion of the mixture. Because the amount of cement in a typically lean mass concrete mix-
ture is relatively small, cooling it may not be significant in a temperature control program (98).
With respect to aggregates, which comprise the greatest part of a concrete mixture, a change in
their temperature will produce the greatest change in the temperature of the concrete.
Methods for cooling coarse aggregates can include sprinkling stockpiles with water to pro-
mote evaporation, which helps to cool the aggregate. The moisture condition of the aggregates

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

44   Concrete Technology for Transportation Applications

needs to be considered not only for adjustments in amount of batch water, but also in the heat
balance calculations for control of the placing temperature (98). If stockpiles are sprinkled,
adequate drainage needs to be provided beneath the stockpiles to assist in achieving a more
uniform moisture throughout the stockpile.
With respect to the batch water, a unit change in the temperature of the batch water has
approximately five times the effect on concrete temperature as a unit change in the temperature;
ice chips can be added to substitute for a portion of the batch water.
The use of ice is one of the basic and most efficient methods to lower the concrete placing
temperatures. It is important that all of the ice melts before the conclusion of mixing and
that sufficient mixing time is allowed to adequately blend the last of the ice meltwater into
the mixture.
Liquid nitrogen is another effective method used for cooling batch water and creating an
ice–water mixture. It can be injected into the water in a specially designed mixer just before
the water enters the concrete mixer, whereby the liquid nitrogen causes a portion of the water
to freeze. Liquid nitrogen has also been injected directly into mixers. This approach typically
requires that the mixing time be prolonged and, preferably, that the mixer is at least partially
sealed to minimize the loss of gas to the atmosphere.

Placement Area
During hot weather, precooled concrete can absorb ambient heat, mechanical heat, and
solar radiation during placement, which will increase the effective placing temperature and the
resulting peak temperature. This increase in temperature can be minimized or eliminated by
reducing the temperature in the immediate placing area using fog sprayers, shading measures,
or both. Placing concrete at night can also reduce the impact of hot weather and radiant heat.

Scheduling
In many parts of the United States, the temperature of concrete placed during the spring
and early summer will be cooler than the average daily ambient temperature. Conversely, the
temperature of concrete placed during the fall and early winter will often be warmer than the
average daily ambient temperature. Scheduling placements during nighttime in the summer
and during the daytime in the winter can take advantage of cooler and warmer temperatures
and supplement other methods of reducing concrete temperature.

Temperature Modeling and Measurement


For critical projects, thermal modeling, using the Schmidt’s method (ACI 207.2R) (99) or
finite element models (FEM), has been performed during design and planning stages to predict
concrete temperatures and evaluate different control strategies. However, the prediction accu-
racy is not very good due to multiple assumptions made in the calculations. Modeling accuracy
can be greatly improved by measuring the adiabatic or semiadiabatic heat from the cement or
cementitious materials used in the concrete mix. ASTM 1702 is the test method used to measure
the heat of hydration of cementitious materials using isothermal calorimetry. Results of this test
can then be used as input values in models.
The modeling can aid in the optimization of mixture proportions, determining proper
cooling and insulation needs, accommodation of predicted ambient conditions, and test-
ing the effectiveness of construction provisions (100, 101). During construction, monitoring
of the early age temperature of mass concrete is often required for QA/QC purposes. Temperature
measurements can be obtained using a variety of embedded temperature sensors, including
commercially available maturity measurement systems. In addition to monitoring internal and

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   45  

surface temperatures, embedded sensors can also be used to apply the maturity method to
predict concrete strength (102).

After Concrete Placement

Thermal and Stress Development After Placement


Once cement and water interact during mixing, the hydration process starts and the mixture
temperature begins to rise. At placement with lifts of 4 ft (1.2 m) or higher and lateral form dimen-
sions of more than 5 ft (1.5 m), the temperature will start to rise in the central part of the mass of
fresh concrete without the opportunity to dissipate. In contrast, the heat generated at the exposed
surfaces (formed or unformed) is dissipated into the surrounding air at a rate dependent on the
temperature differential with surrounding temperature. The net temperature increase in concrete
adjacent to the surface (or forms) is less than in the interior (99).
Before initial set of concrete, the rise of the interior temperature of the structural member
causes little or no stress or strain. Also, the development of steep thermal gradients near exposed
surfaces during early ages while the modulus of elasticity is very low is usually not a serious con-
dition. After the concrete hardens and acquires elasticity, decreasing ambient temperatures and
rising internal temperatures work together to increase the temperature gradient and widen the
stress difference between the interior and the surface. This results in a high tensile stress in the
region of the exposed surfaces and a comparatively low compressive stress over the extensive
interior areas. Also, changes in the temperature gradient result in length and volume changes
that are partially restrained due to reinforcement and boundary effects, causing cracks that may
affect structural integrity, durability, and long-term service life of the structure. This necessitates
executing a plan for deploying postcooling systems to limit the rise in the concrete interior and
the temperature differential with the exposed surfaces.

Postcooling Systems
The postcooling systems described below require installation of temperature-sensing devices
(thermocouples or resistance thermometers) at key locations within the massive member to
provide special information for the control of concrete cooling rates.

Internal Cooling.   An effective system to reduce the peak temperature of mass concrete
placements is by circulating cool water through small-diameter aluminum or PVC pipes
embedded in the concrete (Figure 21). Water circulated through cooling pipes in concrete
sections such as massive mat foundations, piers, columns and beams can reduce the core
temperatures and differential temperatures between the core and surface. This can sometimes
be an economical alternative to pre-cooling the concrete during hot weather. After the cooling
period is completed, the embedded pipes must be pressure grouted (103).

Surface Insulation.   Applying insulation to formed and exposed concrete surfaces is an


effective method to maintain a manageable thermal differential between the core and surfaces
of the concrete. An insulation factor (R-value) of –2.0 or higher will be necessary in many cases.
Rigid synthetic cellular material in sprayed, board, or sheet form as well as thermal blankets
containing closed-cell material can be practical methods of insulating.
For example, a minimum 3-in (75-mm) wood form may be necessary to provide the desired
level of protection while the forms remain in place. Steel forms offer virtually no insulation
protection and need to be supplemented with suitable insulation materials before concrete is
placed. Another practical solution is to coat the exterior of reusable steel forms with a spray-on
synthetic foam of the necessary thickness.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

46   Concrete Technology for Transportation Applications

Figure 21.   Cooling pipes in mass concrete sections such as piers, beams, and mat foundations (103).

Also, delaying the removal of forms until a balance of temperature is reached between
the concrete and ambient temperatures that meets the specification requirements may save
additional efforts and cost of additional monitoring and insulation. However, if construction
scheduling requires early removal of forms, insulation needs to be promptly installed against
the exposed concrete surface. When practical, form removal and insulating activities need to
be planned for the warmest time of the day. For unexposed formed surfaces, an alternative
procedure is to install insulation on the inside of the forms before concrete placement. The
insulation then is held in place against the concrete surface when the forms are removed. This
method has not been successful on exposed concrete because of surface imperfections caused
by the relatively flexible insulation. It is also suggested that insulation be increased along the
edges and at corners of massive concrete structures. This action has effectively reduced the rate
and magnitude of heat escape and temperature decline during the cold-weather season (98). In
no event should the gradual surface temperature drop when protection is removed exceed the
values recommended in ACI 306R or specified by the highway agency.

Specifications
The negative impact of temperature rise in mass concrete necessitates executing a TCMC
plan to limit the rise in the concrete interior and the temperature differential with the exposed
surfaces. Thirty of the 40 states that responded to the synthesis survey indicated that they have
implemented TCMC. Also, 23 of the 40 states have developed specifications and/or construction
guidelines (Chapter 3).
For example, the Florida DOT Specification 346 requires the development of a mass concrete
control plan in accordance with ACI 207 to ensure that concrete core temperatures for any mass
concrete element do not exceed the maximum allowable core temperature of 180°F and that
the temperature differential between the element core and surface do not exceed the maximum
allowable temperature differential of 35°F (28). It needs to be emphasized that the tempera-
ture requirements and temperature control plans may vary among the DOT specifications. In
some cases, the maximum allowed temperature is 160°F. However, the common goal among all
specifications is to control the core and differential temperatures to mitigate cracking and loss
of durability in massive concrete structures.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   47  

Issues and Challenges


Experience and cost are two major challenges when implementing TCMC plans. Training of
construction personnel and utilization of knowledgeable professionals to prepare the TCMC
and supervise its execution will produce an optimized plan and minimize failures in meeting
the specification requirements. Proper execution of the TCMC plans will ultimately save cost
and construction time and will produce quality and highly durable structures.
Also, it is critically important that the agency designer or the project specifier identify in the
project plans those members that are considered mass concrete. This will allow contractors
to account for the additional cost in their bids and plan ahead for construction of the mass
members (104).
The agency may also set the required strength at 56 days instead of the traditional 28 days
for foundation members that will not be loaded before 6 months after construction (104). This
provision will allow the increased use of SCMs and lower the cement content in the mixture,
which will limit the rise in concrete temperature without affecting the structural integrity and
long-term durability of the member.

Precast Concrete Pavement


Introduction
Traditional slab replacement in concrete pavement rehabilitation projects involves several
intense and time-consuming construction activities that must be completed in a short lane-
closure time. In most urban areas the nighttime lane closure is limited to 8 or 9 hours, although
some daytime and weekend closures may be possible if traffic conditions permit. This results
in reduced productivity, early age distresses, and possible shorter service life of the replacement
slabs compared to normal paving (105).
One solution to increase production and to maintain good quality and longevity of replaced
slabs is to use PCP instead of cast-in-place (CIP) panels or slabs. Precast concrete technology
in pavement construction and rehabilitation is relatively new when compared to its tradi-
tional use in structural members (106). In fact, most U.S. projects have been in service for less
than 20 years. However, PCP systems are gaining wider acceptance in the United States for
rapid repair and reconstruction of concrete and asphalt pavements in highways, roads, and
intersections.

Benefits and Advantages


PCP systems are used in roads and highways with high traffic volume and where lane
closures are problematic. The work is performed during the night, typically from about
8 p.m. to 6 a.m., or when short closures during daytime are possible. The production rate per
lane closure is 15 to 20 segments or slab replacements or 300- to 600-feet-long sections of
continuous slab replacements (6).
Below are benefits and advantages of PCP as compared with CIP pavements (6):
• Better quality concrete—Panels are prepared, cast, and cured in a precast plant under more
controlled conditions. This eliminates potential CIP construction problems such as concrete
delivery issues, paving equipment malfunction, early setting and poor mix consolidation, and
poor surface finish.
• Minimum weather restrictions on placement—Construction may continue during very
cold nights or in rainy conditions that would otherwise restrict CIP construction.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

48   Concrete Technology for Transportation Applications

• Reduced delay before lane opening to traffic—PCP can be installed during nighttime lane
closures and be ready to be opened to traffic in early morning.
• Elimination of premature cracking or joint raveling—Early age distresses due to shrinkage
and thermal stresses or raveling sawed joints are almost eliminated when PCP is used.
• Control of structural cracks by panel reinforcement or prestressing—During its service
life, any fatigue cracking will be maintained tight and confined by the reinforcement, simi-
larly to bridge decks. This will allow the slab to function for a long time before replacement
will be needed.

Applications
PCP technology is being applied in intermittent repairs of full slabs and for continuous
paving of long uninterrupted road sections in rehabilitation projects. The service life is pre-
dicted to be at least 20 years for individual slab replacements and 40 years for continuous
paving projects.
Dozens of projects have been constructed using different PCP systems and designs. Cali-
fornia (107, 108), Florida (109), Missouri (110), New York (106), Texas (111), Virginia (112),
and Illinois Tollway (113), among other state agencies, have constructed PCP on sections of
their roads and highways. Progress continues to be made through research and demonstration
projects in all aspects of the technology, including panel design, fabrication, installation, and
performance (106, 109).

Implementation of PCP Systems


The implemented PCP systems include proprietary as well as nonproprietary systems (106).
The panels are prepared in precasting plants, and then transported to the project site to
be placed individually on prepared surfaces to replace the distressed slabs. They can also be
assembled in groups of multiple adjoining panels to form a continuous pavement. Depend-
ing on the pattern of the pavement assembly, the PCP fabrication may include prestressing
the panels at the precasting plant and posttensioning them in groups of multiple slabs during
the pavement reconstruction.
The applications of PCP technology can be classified as follows:
• Intermittent segment or full slab replacements—This application is for isolated individual
segments or full slabs. The segment replacements are made by installing a precast panel on
a prepared leveling course to replace a removed area of the slab. Full-slab replacements with
precast slabs are also made to remove severely cracked, settled, or shattered single slabs or
several consecutive slabs. The process is similar to CIP slabs. The precast panels are typically
a full single lane or two lanes wide.
• Continuous paving—In this application, a long and uninterrupted section of the road or
highway is overlaid or reconstructed on a prepared leveling course. Prestressed precast con-
crete panels or slabs are transported to the project site, assembled in groups of several panels,
and then posttensioned at expansion joints between adjoining groups.

PCP for Intermittent Replacements


The two precast systems that have often been used in the United States for individual seg-
ments or full slabs, or short sections of jointed slabs, include the Illinois Tollway PCP generic
system (106, 112, 113), and the Fort Miller Super-Slab® system (112, 113).

Illinois Tollway PCP System.   In 2007 and 2008, Illinois Tollway implemented the Super-
Slab® system on a section of I-294 and I-88 ramps. However due to the proprietary nature of

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   49  

the Super Slab® system and to encourage competition, the Tollway worked with local industry
to develop their own generic precast system (112, 113). The generic system was developed in
2009 and has been used for intermittent replacement of individual segments or full slabs in their
tollway system.
The generic PCP system uses precast slabs with preformed dovetail slots that are fitted with
loosely positioned dowel bars ready to be pushed in the holes in the adjacent slab, as shown in
Figure 22. The panels have been designed and fabricated in preset dimensions to fit in the slab-
replacement pit dimensions with respect to length, width, and thickness. The most widely used
panels are 6 ft (1.8 m) long and a lane-width wide. The panel thickness equals the average thick-
ness of the existing slabs in the project as determined from average thickness of multiple core
samples from the project
The dowels are concentrated at the wheel path zones, with four dowels in each zone. During
installation, the precast panel is lowered into the replacement pit, ensuring that slots in the panels
match the dowel holes in the opposing side to facilitate insertion of dowel bars in sockets in the
opposing side. The dowels are pushed from each slot into the opposite epoxied hole (Figure 22).
The slots are then filled with fast-setting nonshrink grout. The PCP surface is later diamond-
ground to achieve the required surface smoothness.

Fort Miller Super-Slab® System.   The Super-Slab® fabricated concrete slab system is a pro-
prietary product developed by the Fort Miller Co., Inc., with assistance from the New York State
Thruway Authority and New York State DOT (106). This system was designed as a precast, non-
prestressed concrete slab with prepared slots at the edges for housing tie bars and dowels across
the joints as shown in Figure 23. Foam gaskets are placed underneath the slab to aid in proper
grout distribution beneath the slab. The design of the system is most suitable for intermittent
replacement of segments and full slabs. The precast panels can be fabricated to the specific
project-designed dimensions.
Installation involves unique construction methods, including the use of special bedding
material (leveling course), a laser-guided grader for proper finishing of the bedding material,
and grout to anchor the dowels and tie bars and provide proper slab support.
The super-slab system includes grouting the dowel bar slots and the interface between the
slab and the bedding layer (leveling course). Grout ports are provided on the top of the slab

Figure 22.   Illinois Tollway PCP system for individual slab replacement (113).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

50   Concrete Technology for Transportation Applications

Figure 23.  Super-Slab® precast system and its installation (111, 112).

for the dowel grout and the bedding grout. Additionally, for even distribution of the bedding
grout, grout distribution channel and foam gaskets are provided underneath the slab as shown
in Figure 23.
Key requirements for successful intermittent PCP applications include
• Good and uniform support condition under the panels;
• Adequate load transfer at transverse joints;
• Ensuring minimum, if any, elevation differences between the panel and the surrounding
pavement; and
• Acceptable long-term performance of the PCP replacement panels (112).

PCP for Continuous Paving


The PCP systems have been used to pave continuous short and long sections of roads and
highways as part of pavement rehabilitation and reconstruction. The most common system used
by several states is the FHWA generic prestressed precast concrete pavement (PPCP) system,
originally developed at The University of Texas at Austin (111). The PPCP system has been
modified in recent projects. The Florida DOT adopted a slightly different PPCP design for the
reconstruction of a section of US-92 in central Florida (109).
The continuous PCP sections have generally been longer than 500 ft (150 m), covering a
single or multiple lane, including shoulders. The technology has been implemented on ramps,
intersections, as wells as on highways in states such as California, Florida, Illinois, Missouri,
Texas, and Virginia (107, 108, 109, 110, 111, 112). The PCP is designed using procedures similar
to those used in design of CIP pavements.
The Super-Slab and Illinois Tollway Generic Precast systems have also been used for
continuous paving. The paving involves preparing a well-profiled base, using laser-guided
grading equipment. The precast panels are transported to the project site and installed according
to the design plans in a single lane or tied multiple lanes (106). A typical installation of the
Super-Slab system for continuous paving is shown in Figure 24 (114).

FHWA Prestressed and Posttensioned Concrete Pavement.   The prestressed and postten-
sioned concrete pavement design simulates the design of CIP paving with respect to traffic loads
and jointing requirements, and type of base. One of the most common posttensioning designs
is shown schematically in Figure 25. The panels are fabricated in special forms at a precasting
plant. Reinforcement is placed in the forms. In addition, prestressing strands are also fitted in the

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   51  

Figure 24.   Installation of Super-Slab for


continuous paving (114).

forms and tensioned. The forms also include posttensioning ducts. After casting the panels, the
prestressing strands are released when the concrete reaches the required strength. Some panel
designs also include keyway edges to facilitate interlocking the joints between adjacent panels
(115). In other designs, the ends are flat but are coated with epoxy to bond to adjacent panels
during the pavement assembly to maximize load transfer across the joints.
The fabricated prestressed panels are transported and installed at the project site in segments
composed of 10–15 consecutive panels. Posttensioning strands are inserted through the ducts
in the panels, and then posttensioned with sufficient force to press the panels against each
other to establish good load transfer across the joints (Figure 26). The ends of each segment
are connected to the next segment at a doweled expansion joint to maintain continuity for the
intended pavement.
The length of a segment may vary from 150 to 250 ft (46 to 76 m). The individual panel width
may extend across single or multiple lanes. The panel length can vary from 8 to 12 ft (2.4 to
3.6 m), and the width can vary from 12 to 36 ft (3.6 to 10.8 m) for single or multiple lanes with
shoulders.

Florida’s US-92 PPCP Project.   In 2012, the Florida DOT constructed a 793-ft demonstra-
tion PPCP section on a US-92 rehabilitation project near Deland (109). It was part of an 8.2-mile
concrete pavement rehabilitation project, which also included concrete overlay and other
conventional rehabilitation such as slab replacement, grinding, joint resealing, and crack
sealing. PPCP panels were fabricated in a precasting plant and then transported, installed,
and posttensioned into a pavement structure.

(a) (b) (c)

Ducts for Ducts for


Posttensioning Posttensioning
Ducts for
Posttensioning
Continuous Continuous
Shear Key Shear Key
Continuous
Shear Key Stressing Pockets
Expansion Joint Detail
Pretensioning Strands Pretensioning Strands Pretensioning Strands

Figure 25.   Schematic of the FHWA prestressed and posttensioned concrete pavement system (115).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

52   Concrete Technology for Transportation Applications

Figure 26.   Posttensioning panels on site (110).

The design included an asphalt interlayer to change the transverse profile of the pavement
from a centerline crown to a 2% cross slope toward the outside edge and also to provide a
smooth base to place the PPCP panels, as shown in Figure 27. Plastic sheet was used to cover
the asphalt layer prior to placement of the PPCP panels to minimize friction during the post-
tensioning process.
The panels were 12 ft long, 24 ft wide, and 9 in. thick (3.65 m × 7.3 m × 23 cm) and included a
keyway-pattern transverse joint. The PPCP panels were reinforced, fitted with six posttensioning
ducts, and then cast and cured; they were prestressed transversely in four locations at a precast­
ing plant. After fabrication, the panels were transported and placed along the paving track, where
they covered two 11-foot lanes (3.35 m) and a 2-foot (0.6 m) shoulder, as shown in Figure 27.
The PPCP panels were grouped in three 264-ft (80-m) segments. Each segment comprised
22 PPCP panels that where placed with the posttensioning ducts of adjacent panels, aligned, and
connected at the joints (109). Panels of each full segment were posttensioned at the end panel.
The ends of two consecutive segments were connected at a doweled expansion joint (Figure 28).
The posttensioning was performed at the end panels and were patched with a quick-set patch-
ing material. The posttensioning ducts were grouted through ports at the end panels. Also,
the intermediate joints between adjoining panels were patched using the quick-set patching
material, as shown in Figure 29.

Figure 27.   Placement of PPCP panel. Notice the six


posttensioning ducts.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   53  

Figure 28.   PPCP unit end panel with prefabricated


dowel slots, posttensioning end blocks, and duct
grouting ports.

The expansion joints were sealed, and dowel slots and posttensioning blocks were patched.
The pavement was later diamond-ground to establish a smooth riding surface (Figure 29). The
PPCP section was constructed in 14 days. The pavement has been in service for 4 years under
local traffic. It provides a smooth ride and has performed well structurally with no signs of
distress (116).

Limitations and Challenges


Note that because of the short experience with the use of PCP technology in the United States,
information and guidance on PCP design, practices, and in-service performance are continu-
ously evolving with every demonstration project and performance study. At this time, there are
not sufficient PCP projects available with long service time to allow direct field validation. As a
result, the design of PCP systems needs to be based on current design procedures for conven-
tional CIP jointed concrete pavement, such as the AASHTO Maine software (107).

Figure 29.   All PPCP joints were patched with


quick-set patching material and the pavement was
diamond-ground.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

54   Concrete Technology for Transportation Applications

It must be emphasized that PCPs are not “super pavements” and are not be expected to perform
significantly better than CIP concrete pavements. Once installed, precast concrete pavements can
be expected to behave similarly to CIP pavements under traffic loading and environmental effects.
The primary difference between the two technologies is how each system is constructed (6).
Some key limitation and disadvantages of PCP are the following (6):
• Higher cost compared to CIP;
• Lack of adequate long-term performance history for PCP;
• Lack of experience of local contractor and precasting operators with the technology;
• Proprietary products;
• Newness of the PCP technology to DOT designers and consultants;
• Construction delays and challenges, including
– Establishment of standard dimensions for the slab replacement pit,
– Panel dimension tolerances to fit the standard replacement pit with acceptable joint
openings that must be sealed,
– Preparation of the bedding layer (leveling course) and its proper grading to ensure even
surface grade of panels with surrounding pavement,
– Space limitation in the closed lanes to accommodate the PCP transporting trucks as well as
cranes and other lifting equipment, which will require special maintenance of traffic plans,
– Ensuring the patching of wide dowel slots prior to lane opening to prevent accidents
involving motorcycles,
– Treatment of special pavement areas such as gores, horizontal curves, or turning at inter-
sections that would require either the use of CIP concrete or special precast panels that
would fit in these areas;
• Long-term performance of expansion joints;
• Lack of national well-developed selection criteria for use of precast concrete technologies
(criteria are generally based on desire of a highway agency to experiment with PCP); and
• Lack of knowledge about maintenance needs and repairs, considered a gap in the PCP
technology.

Roller-Compacted Concrete
Introduction
RCC is a stiff concrete mixture with very low or no slump. In recent years, RCC pavements
have been used for roads and highways. They are designed similarly to conventional concrete
pavements, but are constructed without the use of forms, dowel and tie bars, or other steel
reinforcement. RCC pavements use conventional asphalt or high-density paving equipment
instead of slipform paving machines. The pavement is compacted into its final form using a
combination of heavy vibratory steel-drum and rubber-tired rollers (7, 117).

Benefits
RCC pavements are typically more cost-effective compared to conventional concrete pave-
ments (7). The cost savings are due to reduction in cement content, lack of forms, lower
placement costs, faster construction, as well as absence of dowels and tie bars.
RCC can also be constructed rapidly without the need to set up forms or stringlines or to install
dowel baskets ahead of the paving operation. After placement and compaction, RCC pavements
are well consolidated, where stable aggregate interlock develops to support movement of workers
and occasional light vehicle traffic without damaging its fresh surface (Figure 30). Cement hydra-
tion and strength gain continue in a manner similar to that of conventional concrete.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   55  

Figure 30.   Freshly placed RCC pavement (courtesy of


G. Dean, American Concrete Pavement Association).

With well-graded aggregates, proper cement and water content, and dense compaction, RCC
pavements can achieve strength properties equal to those of conventional concrete, and with low
permeability. In fact, a key advantage to this type of concrete paving is that the pavement may be
opened to traffic shortly after construction, as soon as the strength reaches 3,000 psi (21 MPa) (7).

Applications
The modern use of RCC started in the 1970s in Canada when the logging industry switched
to RCC as a heavy-duty paving material for log-sorting yards to sustain the massive loads and
heavy equipment (117). Since then, RCC has been used in many other types of applications
throughout North America. One of the main applications of RCC in the United States is in the
construction of dams (118). Other primary applications of RCC include ports, industrial park-
ing and storage areas, intermodal and military facilities, and low-speed roads and intersections
carrying heavy truck traffic (119).
In recent years, RCC pavement has been used in commercial areas, on local streets and roads,
and on highway shoulders (Figure 31), as well as in base layers in two-lift concrete or asphalt

Figure 31.   RCC shoulder next to existing concrete


pavement (courtesy of G. Dean, American Concrete
Pavement Association).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

56   Concrete Technology for Transportation Applications

pavements (120, 121, 122). Success of these applications has been a result of improvements in RCC
mixtures, paving equipment, and construction practices, as well as an increase in the availability of
specialty admixtures that improve RCC placement speed and surface smoothness (7, 123, 124, 125).

RCC Mixtures
RCC mixtures consist of the same basic ingredients as conventional concrete mixtures, but
with less cementitious material, lower water–cement ratio and coarse aggregate contents, and
higher percentage of fine aggregates to allow for tight packing and compaction (7). A typical
RCC mixture includes coarse and fine aggregates as well as cementitious materials including
cement, fly ash, water, and, when appropriate, chemical admixtures. Aggregates make up to
85% of the volume of RCC. Proper attention to aggregate gradation will help ensure concrete
workability at placement, proper compaction, and good surface finish.
Fresh RCC mixtures have low or no slump and are stiff enough to remain stable under vibra-
tory rollers yet wet enough to permit adequate mixing without segregation. RCC mixtures typi-
cally provide similar strength properties and less shrinkage compared to conventional concrete
mixtures in similar supplications (117).
RCC mixtures are stiffer than conventional concrete mixtures because of their higher fines
content and lower cement and water contents. The mixtures are placed with a heavy-duty, self-
propelled asphalt paving machine, using a high-density single or double tamper bar screed to
initially consolidate the mixture to a slab of uniform thickness. This is followed by a combina-
tion of passes with rollers for proper compaction and fine surface texturing (7).

Mixture Production
The RCC mixtures are prepared in rotary concrete mixers at concrete batch plants or in pug-
mills at plants or project sites (Figures 32 and 33). The RCC is then charged into dump trucks to
be transported to the placement site.

Pavement Construction
Placement
RCC pavements are constructed without forms or reinforcing steel because the mixture is dry
and stiff. Dowels, tie bars, and reinforcement bars are not incorporated into RCC pavements

Figure 32.   RCC drum mixer (126).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   57  

Figure 33.   RCC pugmill (courtesy of Morgan Corp.).

because the dry consistency of the mixture does not allow bonding with steel and the roller com-
paction will result in dowel/tie bar misalignment (7). This fact must be considered in pavement
design, since load transfer at the joints would be dependent on aggregate interlock with aid from
a firm base support to minimize deflections under heavy vehicles.
RCC is typically placed with an asphalt paver. The pavements are usually placed in lifts of 6 to
8 in. (150 to 200 mm) with 4 in. (100 mm) as a minimum and 10 in. (250 mm) as a maximum
thickness (7). Conventional asphalt machines may be used for lifts that are 6 in. (150 mm) thick
or less. High-density asphalt paving machines have been used successfully for pavements up
to 10 in. (250 mm) thick, although 6- to 8-in. (150- to 200-mm)-thick lifts are more common
(117). When placing multiple lifts, the top lift needs to be placed within 60 minutes of the lower
lift to allow for adequate bonding between the layers (7).

Compaction
RCC is typically compacted with a 10-ton (9-metric ton) dual-drum vibratory roller imme-
diately after placement. Typically, four to six passes of a dual-drum 10-ton vibratory roller
will achieve the desired density of at least 98% for RCC lifts in the range of 6 to 10 in. (150 to
250 mm). Rubber-tired rollers or lighter steel rollers (4 tons) have also been used successfully,
especially for a final pass, to remove surface cracks and tears and to provide a smooth and
tight surface. The RCC pavement density is tested and verified using the field nuclear density
gauge (Figure 34), according to ASTM C1040/C1040M.

Figure 34.   Field density test.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

58   Concrete Technology for Transportation Applications

RCC needs to be placed and compacted while it is still fresh and workable, usually within
60 minutes of delivery. If the RCC is too dry, the surface will appear dusty or grainy and may
even shear (tear) horizontally. Excessive vibration can lead to edge collapse, which will disturb
the profile of the road and its riding quality. Longitudinal construction cold joints are formed
when a lane is placed more than an hour after the placement of the adjacent lane. Construction
joints are formed by trimming away the outer uncompacted edge of the paving lane with a
concrete saw and then paving against the resulting clean vertical edge (7).

Curing
Curing of RCC pavements is as important as in conventional pavements. Application of a
heavy dosage of curing compound on the pavement surface and sides will preserve the mixture
water to sustain the hydration process and achieve proper strength and a durable surface (7).
Applying a fine mist or fogging prior to curing will prevent evaporation from the dry pavement
surface. Also, using intermittent water spray in addition to the application of curing compound
may add surface durability.

Joints
The joint saw cut is one-third of the slab thickness to minimize uncontrolled cracks (125, 127, 128).
Early entry “green” saws have been used to initiate a 1-in. (25-mm) deep joint cut in the freshly
compacted RCC pavement to prevent uncontrolled cracks. That is followed by a deeper cut to
the proper depth using conventional saw machines (124). The joints are later cleaned and sealed
similarly to conventional pavements.

Texture
RCC pavement has an open surface texture similar to that of asphalt pavement. The surface
may have a rough texture and may exhibit poor ride quality in higher-speed roadways. The use
of smaller aggregates, adding more cement, or using specialty admixtures can produce a denser
and more favorable surface texture (123). Surface smoothness is further improved with diamond
grinding (124).

Opening to Traffic
The pavement can be opened to traffic within hours after construction. Compressive strengths
of 2,500 to 3,000 psi (17 to 21 MPa) are used as criteria for opening the pavement to traffic (7).

Properties
Strength
In general, RCC pavement mixtures can have compressive and flexural strengths compara-
ble to those of conventional concrete mixtures. RCC compressive strengths at 28 days typically
range from 4,000 to 7,000 psi (28 to 48 MPa) (129). The densely graded aggregates and the
use of low water–cement ratio in the mixtures help to achieve higher compressive strengths in
the concrete (7). Although strength tests for RCC are similar to those used for conventional
concrete, unlike in conventional concrete mixtures, cylinders and beam samples used for
RCC strength tests are compacted using a vibrating hammer (Figure 35).

Bond Strength and Multiple Lifts


When placing RCC in multiple lifts, bond strength at the interface of RCC layers becomes a
critical engineering property. Good bond strength must be achieved to produce a monolithic
pavement that supports the designed traffic loads. Insufficient bond or bond failure will result
in delamination and premature failures.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   59  

Figure 35.   Preparing RCC field-test cylinders


(courtesy of Georgia DOT).

Freeze-Thaw Performance
RCC is not typically air entrained from conventional air-entraining admixtures. However,
field performance studies in freeze-thaw environments have indicated that RCC performed
well for more than three decades, whether air entrained or not (112, 130). Test data clearly
demonstrate that very little entrained air is required to adequately protect RCC against frost-
induced microcracking and deicing-salt scaling (7).

Specifications
ACPA has published a guide specification (124) useful for developing project specifications
for RCC pavements. The guide specification includes materials and mixture requirements as
well as recommendations for construction and QC processes and procedures.

Implementation Examples in Streets and Highways


RCC has been used in pavements for local streets and roads and for highway shoulders. Speed
of construction, economy, and early opening to traffic are key reasons to use RCC for local streets
and roads (7, 124). For example, RCC may be used for inlays to replace rutted asphalt pavements.
In some cases, light traffic has been allowed on the RCC pavement within 24 hours after construc-
tion to accommodate local traffic, or when concrete reaches the strength of 2,500 psi (7, 120).
Because of the tendency of RCC to produce a rough surface texture, advancements have
been made to improve this aspect of RCC paving. Most projects use high-density pavers,
mixtures with smaller top-size aggregate, and workability-aiding admixtures to improve surface
smoothness (123). In streets with speed limits exceeding 30 mph (48 km/h), diamond grinding
has often been applied to improve surface smoothness (124). Applying a thin asphalt surface
course on top of the RCC is another option to improve surface smoothness.
For roadways carrying traffic at highway speeds, RCC is currently used primarily as a base that
is overlaid with a thin asphalt wearing course for better rideability. Another possibility for high-
ways is the use of RCC as a base for conventional concrete pavement. RCC provides an excellent
construction platform in unbonded concrete overlays. This allows for the conventional concrete

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

60   Concrete Technology for Transportation Applications

pavement surface thickness to be reduced. A separation layer (generally asphalt) between the
RCC base and concrete overlay is required as a bond breaker to allow for separate movement
of the layers and to prevent cracks from reflecting upward from the base into the conventional
concrete pavement. In all applications, joint spacings similar to those used in conventional pave-
ments have been used in RCC pavements. In unbonded concrete overlays on RCC, the joints
must coincide to prevent reflective cracking (121, 126, 127).
Some examples of street and road applications include reconstruction of a street in Columbus,
Ohio. This four- to six-lane project was constructed under traffic. It consists of 8 in. (200 mm) of
RCC overlaid with 3 in. (75 mm) of asphalt to provide smoothness for the higher traffic speeds.
Another example is US-78 in Aiken, South Carolina, where a 10-in. (250-mm) RCC pavement
replaced an existing full-depth asphalt pavement. The RCC surface was diamond ground for
this four-lane section to improve its surface smoothness and rideability (128).
RCC has also been used in highway shoulder applications. For example, the Georgia DOT
used RCC to reconstruct a 34-mile (55-km) shoulder on I-285 (121, 122) (Figure 36). The exist-
ing shoulder was milled and replaced with a 10-ft (3-m)-wide RCC pavement at a thickness
of 6 to 8 in. (150 to 200 mm). Rumble strips were ground into the RCC surface to conform to
highway safety requirements. No surfacing was placed on the RCC. The transverse joints were
sawed to align with the joints in the mainline pavement.
A directory of RCC pavement projects in the United States is available from the ACPA (131).
Also, the ACPA recently developed pavement design software, called Pavement Designer, that
includes provisions for design of RCC pavements (132). The software is a web-based pavement
design tool that can be used for design of roadways, overlays, parking areas, and industrial and
intermodal pavements.

RCC Pavement Issues and Challenges


Possible limitations and challenges associated with RCC pavements include the following:
1. Surface profile and smoothness. An RCC pavement may not meet the rideability require-
ments for roads and streets carrying high-speed traffic. Surface grinding has significantly

Figure 36.   RCC paving on an Interstate shoulder


(courtesy of G. Dean, American Concrete Pavement
Association).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   61  

improved the ride quality of RCC pavements (127). Also, recent advances in specialty admix-
tures and surface finishing aids have produced smoother RCC surfaces (123). The use of
smaller coarse aggregates and modifications in roller compaction patterns and weights have
also helped contractors to achieve better surface smoothness.
2. Paving in multiple lifts and adjacent lanes. Placing multiple lifts of RCC must be accom-
plished within 1 hour of each other to ensure good bond between the layers. Also, the ini-
tially paved layer must be clean and kept damp prior to placement of the next lift. In case of
multiple lane construction, if adjacent lanes are not placed within 1 hour of each other, a
longitudinal joint must be formed between lanes (7).
3. Pavement edges are more difficult to compact. It is suggested that the specifications allow
96% modified Proctor density along pavement edges instead of the 98% required on pave-
ment interior. History has shown that properly prepared and consolidated edges compacted
to lower density perform very well (7).
4. Hot-weather paving. Precautions and good practices followed when paving RCC in
hot weather are not much different than those used in conventional pavement. In hot
paving environments, extra attention is required to minimize loss of RCC mixture water by
evaporation. Use of admixtures and misting, fogging, or spraying of newly placed pavement
can minimize water loss and protect its surface integrity (7). Timely curing with application
of a heavy dose of curing compound on pavement surface and edges as well as intermittent
water spraying will provide the necessary curing for strength development and produce a
durable surface.

Pervious Concrete
Introduction
Pervious concrete (PC) is an open-graded concrete mixture with a near-zero slump, comprising
portland cement, coarse aggregate, a small amount or no fine aggregate, admixtures, and water.
In its hardened state, it contains high percentages (20% to 35%) of interconnected voids, which
allow rapid passage of water through the body of the concrete (Figure 37). The typical compressive
strengths of PC are from 400 to 4,000 psi (2.8 to 28 MPa). The drainage rate of PC pavement will
vary with aggregate size and mixture density from 2 to 18 gal/min/ft2 (81 to 730 L/min/m2) or
192 to1,724 in./h (0.14 to 1.22 cm/s) (8).

Figure 37.   PC parking during a rainstorm.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

62   Concrete Technology for Transportation Applications

Figure 38.   PC base for concrete pavement.

Applications and Benefits


PCP is often used as a surface layer in streets and parking areas to reduce or eliminate storm-
water runoff (133). It is also used as a rigid subbase beneath concrete pavements in roads and
highways (Figure 38) to drain surface water and provide stiff support for the pavement (134). PC
is also commonly used in edge drains of pavements (Figure 39) to collect and pass the drained
water into a perforated pipe and discharge it at specific distances to the side trenches (135).

Materials
A PC mixture is composed of cement or a combination of cement and other cementitious
materials, coarse aggregate, water, and admixtures. A limited quantity of fine aggregate may
be added to increase the compressive strength and density of the mixture and achieve higher

Figure 39.   PC edge drain.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   63  

load-carrying capacity of PC pavements. However, the flow rate of water through the pavement
is reduced with increase in the amount of fine aggregate (8). Color may also be included in the
mixture for enhanced esthetics and/or to distinguish the surface from adjacent conventional
concrete or asphalt surfaces.
Portland cement is used as the primary aggregate binder. However, SCMs such as fly ash,
GGBFS, and silica fume can also be used in PC mixtures as partial replacements of the cement
content. Use of SCMs has been, in some cases, shown to improve the strength and durability of
PC, although the results vary depending on SCM type used and replacement rates (135, 136, 137).
Rounded and crushed aggregates, meeting requirements ASTM C33/C33M, have been used
in PC mixtures. Aggregate sizes No. 7 (1/2 in. to No. 4), No. 8 (3/8 in. to No. 8), No. 67 (3/4 in. to
No. 4), and No. 89 (3/8 in. to No. 16) have been used in PC mixtures. Aggregates in PC mixtures
can be normal-weight or lightweight aggregates. The aggregate must be clean and free from dust
or clay that might adversely affect bond with the cementitious paste. Prior to mixing the PC
ingredients, the aggregate must be in a saturated-surface-dry (SSD) moisture condition. This is
achieved by applying water sprinklers on the aggregate stockpile. Dry porous aggregate will absorb
mixture water and reduce the workability necessary for proper placement and compaction (8).
With respect to admixtures, water-reducing admixtures (high range or medium range) are
commonly used to achieve proper workability. Also, the use of cement hydration stabilizers
and VMAs aids in extending the placement and compaction time of the mixture. The three
admixtures are also helpful in producing and maintaining proper consistency of the PC mixture
during transport, placement, and compaction, and also to mitigate segregation (138, 139).
Retarding admixtures may be used to stabilize and control the cement hydration to allow
longer transport and to extend the compaction window of the PC mixture. They can also act
as lubricants to help discharge concrete from a mixer and improve placement and compaction,
especially in hot weather. Accelerators can be used when PC is placed during cold weather (8).
Air-entraining admixtures are not commonly used in PCs but may be included in pavement
susceptible to freezing and thawing (8). Also, incorporating fibers in mixtures exposed to freez-
ing and thawing has shown success in improving durability and abrasion resistance in cold
climates. However, fibers do net seem to improve compressive strength and may reduce PC
permeability and infiltration (140, 141).
Dry and windy conditions create high evaporation rates that reduce PC workability during
placement and compaction and may case raveling of surface aggregates. The use of evaporation
retarders and other chemicals is beneficial in windy and dry paving conditions to maintain sur-
face moisture prior to applying plastic sheet cover (8).

Mixture Proportioning
The goal of mixture proportioning for PC is to achieve balance between permeability, strength,
and workability. The mixture has to be proportioned to meet the main objective of the application
type. For example, to maximize the water percolation rate, the PC mixture has to be proportioned
to produce high void content including the use of gap-graded aggregate and lower paste content.
Conversely, to maximize a pavement load-carrying capacity, the PC mixture may include higher
cement content, lower water–cement ratio, well-graded coarse aggregate, and possibly, the addi-
tion of some fine aggregate. The Virginia DOT has developed special provisions for PC parking
areas that include provisions for materials, mixture design, and construction (142, 143). The
Florida DOT has specifications for pervious base layers and pavement edge drains (134, 135).
Higher voids in PC will lead to a higher percolation rate and lower strength. In contrast, lower
void content results in a reduced percolation rate but higher PC strength. The addition of fine

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

64   Concrete Technology for Transportation Applications

Figure 40.   Virginia DOT’s falling head


permeameter (143).

aggregate will decrease the void content and increase strength (135). Density and void content of
freshly mixed PC mixtures can be determined using ASTM C1688/C1688M-14a “Standard Test
Method for Density and Void Content of Freshly Mixed Pervious Concrete.” The infiltration rate
of water determined using 4-in. × 6-in. (10-cm × 15-cm) samples obtained from trial batches or
field core samples may be tested using ASTM D5084—Method B, “Standard Test Methods for
Measurement of Hydraulic Conductivity of Saturated Porous Materials Using a Flexible Wall
Permeameter,” as shown in Figure 40 (143).
Experience has shown that a water–cement ratio in the range of 0.26 to 0.45 will provide
paste with a consistency supporting the best aggregate coating, paste stability, and overall mix-
ture cohesiveness. The conventional relationship between water–cement ratio and compressive
strength for normal concrete does not apply to PC (8). This is because other factors may also
affect compressive strength development such as paste–aggregate bond strength, void content,
and strength of the aggregate particles.
A necessary range of quantities for the main PC mixture ingredients is shown in Table 2. Admix-
tures are also used in dosage rates recommended by the manufacturer or concrete producer.

Table 2.   Necessary quantities of basic PC ingredients


[adapted from ACI 522 (8)].

Component Proportions
3 3
Cementitious materials, lb/yd (kg/m ) 450 to 700 (270 to 415)

Aggregate, lb/yd3 (kg/m3) 2,000 to 2,500 (1,190 to 1,480)

Water–cement ratio, by mass 0.27 to 0.34

Aggregate–cement ratio, by mass 4:1 to 4.5:1

Fine aggregate–coarse aggregate ratio, by mass 0 to 1:1

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   65  

Experience of local PC producers and paving contractors with track records of successful PC
projects will be helpful in designing appropriate PC mixtures for the application.

Design and Construction


Pavement Design
For thickness determination of a pervious pavement, structural adequacy and drainage
efficiency are two important design considerations. These two characteristics influence each
other so both need to be addressed with care. The modulus of elasticity and flexural strength
of PC are typically lower than those of conventional pavement concrete. However, with prop-
erly designed PC thickness on a thick and stiff base layer, PC pavements have demonstrated
acceptable structural capacity for low truck traffic (144).
ACPA’s design software for concrete pavements, Pavement Designer (132), includes PCP
design as one of the software modules. Pavement Designer software pertains to the structural
requirement of the pavement. However, the pavement designer must also be concerned with
the drainage aspect of the pavement, including pavement storage capacity, need for any
backup stormwater drainage system, and the percolation efficiency of the pavement and its
support layers (145).

Construction
The construction process needs to result in a PC pavement possessing both adequate
strength and efficient percolation. It is paramount that the agency be assured that the
contractor and concrete producer are adequately qualified and experienced to construct,
provide construction QC, and produce successful mixtures for PC pavement projects.
Variability in PC pavement performance can be attributed to poor construction practices,
inadequate QC and/or lack of consistency in the properties of PC mixtures among the
production batches (142, 146).
The PC mixture is batched and/or transported to the construction site in truck mixers. The
mixture is discharged in stationary forms, distributed with steel rakes and struck off, with
some compaction, using low-frequency vibratory screeds. This is followed closely with final
compaction and surface finishing using a heavy-pipe roller, roller screed, or frame-mounted
roller (Figure 41). The necessary average roller weight is approximately 500 lb (227 kg). Over-
compaction needs to be avoided to prevent paste from migrating to the surface and closing the

Figure 41.   Placement, compaction, and finishing of PC pavement.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

66   Concrete Technology for Transportation Applications

surface voids (145). Also, because of the nature of the surface compaction, some variation in
strength, void ratio, and permeability can be expected (146, 147, 148).
PC placement needs to be completed as quickly as possible, because the mixture has almost
no excess water. Exposure to adverse weather conditions such as high wind and/or high ambient
temperature for a significant time period will result in rapid loss of hydration water, which will
reduce the final strength and my also cause surface raveling (8).
The pavement surface needs to be covered as soon as the final compaction is completed. Plastic
sheets are generally used to cover the surface and edges of the freshly placed pavement. Applica-
tion of the plastic sheets must be within 20 minutes after compaction and surface finish. This
time may be significantly reduced depending on ambient temperature and wind velocity. Curing
compounds need to be avoided because they may seal the surface voids. For proper curing, the
pavement typically needs to remain covered for at least 7 days. Poor curing practices will result
in surface raveling. No traffic should be allowed on the pavement during curing (8).
Contraction joints need to be installed in the fresh PC pavement using special grooving tools.
Saw cutting the pavement is another method used to form the joints. Saw-cut depths need to be
one-fourth to one-third of the pavement thickness. The saw-cutting procedure needs to begin
as soon as the pavement has hardened sufficiently to prevent surface raveling and damage. The
exposed area around the sawed joint must be recovered with a plastic sheet as soon as saw cuts
have been made (8). The sawed joint must be flushed out with water to remove fines and debris.
However, no sealant is required to fill the joint grooves.
In cold weather conditions, PC construction may be suspended and curing blankets be used
when ambient temperature is expected to fall below 40°F (4°C) during and 24 hours after place-
ment. The PC pavement must be protected from freezing conditions during the curing period.
In hot weather, transporting, placing, and compacting need to be handled as quickly
as possible. An evaporation retarder may be applied to the surface of the concrete follow-
ing the strike-off process to retard loss of moisture from the surface. After consolidation and
before placing the plastic sheet, the surface may be lightly misted, fogged, or sprayed with an
evaporation retardant if the surface appears to be losing its sheen appearance.

Properties
Strength of PC is a function of the aggregate strength, aggregate-paste bond strength, and
strength of the cement paste. Compressive and other strengths increase as the paste proportion
in the mixture increases, the size aggregate decreases, and when the aggregate is well graded.
Adequate compaction and effective curing techniques, which preserve mixture moisture for more
efficient hydration, are important construction-related factors that produce higher compressive
and tensile strengths. Four-inch (10-cm)-diameter core samples are obtained from the pavements
and tested for compressive strength using ASTM C39 (143). It is also suggested that laboratory
test cylinders for compressive strength be prepared with a shape factor of 1:1 (diameter to length).
This shape factor produces test results more representative of the composition of the PC structure
than the cylinder samples with a 1:2 shape factor used in conventional concrete.

Issues and Challenges


Infiltration Capacity
The performance of PC pavements is affected by changes in infiltration rates, structural
distress, surface erosion, and degree of resistance to freeze-thaw cycles. Also, dust, sand, clay,
and other debris will gradually clog the voids in the surface of PC pavements. By reducing the

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   67  

permeability of the PC in order to increase its strength, the risk of surface clogging is further
increased (141).
To avoid surface clogging and loss of drainage capacity, PC pavements in streets and parking
areas require frequent vacuum sweeping or power washing (146). Vacuum sweeping seems to
be the preferred method to dislodge and remove accumulated debris from the surface voids
and to restore or maintain good percolation rate (149). Also, a backup stormwater drain system
may be required in areas of high potential for surface clogging. Infiltration rate at the project
site can be measured using ASTM C1701/C1701M-17, “Standard Test Method for Infiltration
Rate of In Place Pervious Concrete.”

Joint Raveling
Raveling and erosion of the pavement surface and especially along the joints is a major
problem in PC pavements (133). This functional distress may due to poor construction prac-
tices or deficient mixtures. Surface abrasion and erosion may also be caused by heavy traf-
fic wheels exerting repeated pressure on the pavement surface upon breaking and turning.
A useful test to evaluate the abrasion resistance of PC samples is ASTM C1747, “Standard Test
Method for Determining Potential Resistance to Degradation of Pervious Concrete by Impact
and Abrasion.” Selection criteria for PC streets have to consider traffic including volume of
truck traffic. In general, PC pavements are suitable for streets with light truck traffic only.

Compaction
The compaction energy during PC pavement construction affects the pavement load-carrying
capacity, infiltration efficiency, and durability of the pavement. Undercompaction of the pave-
ment will cause low unit weight, high void ratio, and low strength and durability. Overcompac-
tion tends to push the cement paste up to the surface and cause clogging of the voids. A useful
test to determine density and void content of PC is ASTM C1754.
A study was conducted to evaluate the effects of compaction energy on PC void ratio, compres-
sive strength, tensile strength, unit weight, and freeze-thaw durability (150). The study concluded
that low compaction energy reduces PC compressive and split tensile strengths and unit weight
and reduces the freeze-thaw durability of the pavement. Another study evaluated the laboratory
performance of polymer-modified PC with a focus on the abrasion and freeze-thaw durability
(151). Various laboratory tests were conducted to evaluate the physical properties (air voids,
permeability), mechanical properties (compressive and split tensile strengths), and durability
performance (abrasion and freeze-thaw resistance) of PC mixtures. The test results showed that
the use of polymer in PC improves the strength and abrasion resistance of the PC.

Structural Distresses
Structural distress includes cracking or subsidence due to loss of subgrade support. Sources of
the distress include materials and mixture problems, weak subgrade or voids beneath the pave-
ment, and heavy traffic loads. Proper attention to pavement design, good QC of materials and
mixtures, and following good construction practices will reduce the rate of structural distresses
and surface erosion (8). Improved test methods to support QC/QA of PC are the focus of a
recently published study (152).

Freeze-Thaw
When the PC is completely saturated and subjected to freezing, excessive stresses from frozen
water may rupture the thin cement paste coating the aggregates and cause deterioration of the
pavement. It has been recommended that PC pavements not be used in freeze-thaw environments
where the groundwater table rises to less than 3 ft (0.9 m) from the surface of the subgrade (8).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

68   Concrete Technology for Transportation Applications

Fibers seem to improve resistance to freeze-thaw. A study in Maryland concluded that using
cellulose fibers in PC mixtures resulted in a significant improvement in freeze-thaw durability,
increase in abrasion resistance, and improvement in tensile strength (138).

Recycled Concrete Aggregate


Introduction
RCA is produced by crushing and processing demolished concrete infrastructure. In trans-
portation applications, RCA is typically sourced from highway bridges and pavements, but is
also available from commercial crushing of concrete buildings, parking areas, and drainage
facilities. In addition, returned concrete from ready-mixed concrete trucks can also be crushed
and used as RCA. As existing transportation infrastructure is decommissioned, pavements,
bridges, and other concrete elements are becoming increasingly available for and utilized as RCA.
Although there are some limitations, RCA is commonly used in new concrete pavements
and drainage structures. The most popular use of RCA by state highway agencies is in unbound
applications such as pavement base material, fill, or for erosion control at bridges and drainage
structures (9, 153). Agencies have also used RCA from crushed concrete as partial or complete
replacement of aggregates in bound applications including concrete mixtures for pavements
and lean concrete bases (154, 155). Use of RCA in concrete for bridges and other nonpave-
ment structural applications is still rare. However, in recent years, driven by economic and
sustainability factors, as well as the success of research studies and field implementation, concrete
recycling has become more desirable and has been gaining greater acceptance in light structures.

Need for the Technology


Recycling is generally viewed as one of the most sustainable end-of-life alternatives for existing
concrete infrastructure (156). Sustainability benefits associated with use of RCA include reduc-
tion of landfill space, energy savings, fewer impacts from quarrying activities, and a reduced
carbon footprint associated with the production of the new infrastructure. If RCA is produced
on site (or near the site), haul times can be reduced, providing environmental and social benefits
including fuel savings, reduced air emissions, and lower community impacts such as traffic and
noise (155, 156, 157).
Agencies are increasingly pressured to build transportation infrastructure more rapidly, at
reduced cost, and with minimal impact to the traveling public. Use of RCA has been shown to
provide savings in both cost and time. Economic benefits associated with reuse of RCA include
reduced unit cost per ton, time and fuel savings (with reduced hauling), elimination of landfill
tipping fees, and improved contractor production efficiencies (155, 157). Availability of a wide
range of stationary and movable recycling equipment has made concrete recycling an economi-
cal, time-saving option for many projects (155). Depletion of aggregate resources in existing
quarries, difficulty commissioning new quarries, and reduced availability of an ample supply of
high-quality aggregates are also viewed as drivers of RCA use (153).
A recent survey of state highway agencies indicates that the most common use of RCA is
in unbound base material for pavements. This decision appears to result from considerations
including specification provisions, risk, and contractor advantages associated with use of
on-grade recycling (9). When used as an unbound base material in pavements, RCA has
demonstrated performance superior to that of virgin aggregates, primarily due to increased
stiffness associated with RCA’s angular particles and the secondary cementing action of unhydrated
cement (158). However, economic and sustainability factors may increasingly support the

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   69  

use of RCA in new concrete mixtures. With proper design considerations and testing, concrete
containing RCA has been shown to provide performance equivalent to concrete containing
conventional aggregates (154, 155, 159, 160). For the purpose of this report, further discussion
will primarily focus on use of RCA in new portland cement concrete (“RCA concrete”), and the
reader is referred to other resources for guidance on use of RCA in other bound and unbound
applications (153, 155, 161, 162, 163).

Materials, Mixture Parameters, and Common Test Procedures


Source Concrete and RCA Properties
RCA can be used as a full or partial replacement of coarse aggregate or fine aggregate in a
concrete mixture. Concrete produced using RCA needs to meet the same performance require-
ments as concrete produced using conventional (virgin) aggregates. Mindful of this goal, some
specifications treat RCA the same as other types of aggregates. Other agencies have developed
a separate specification for RCA or have modified requirements to accommodate RCA (155).
Regardless of approach, it is important to bear in mind that RCA can have different properties
than conventional aggregates, and the effect of these differences needs to be considered during
design and proportioning of concrete mixtures. Effective characterization of RCA will support
successful use in new infrastructure (153, 155, 162).
Characteristics of RCA will be influenced by (1) the characteristics and quality of the source
concrete and (2) the production process. Most concrete that has demonstrated satisfactory
durability performance can be used to produce RCA for use in new concrete (162). Concrete
sourced from projects with known materials and performance history, such as existing highway
infrastructure, to produce RCA generally brings less risk than use of concrete from unknown
sources or RCA-sourced construction and demolition waste, which may include material from
a variety of sources within a single stockpile (155). Existing infrastructure generally comprises
known materials, and the concrete will have met agency specifications for acceptance.
Source concrete affected by materials-related distress, such as alkali-aggregate reactivity and
D-cracking, needs to receive additional scrutiny prior to making the decision for use as RCA
in new concrete. The potential for alkali-silica reactivity (ASR) due to RCA produced from
ASR-affected source concrete in new concrete is a function of the potential reactivity remain-
ing in the RCA as well as the alkali contents of both the source concrete and the new concrete
mixture. However, ASR damage in new concrete produced using RCA from ASR-affected source
concrete has been shown to be mitigated using the same techniques typically utilized for other
reactive aggregates. These techniques include the use of low-alkali cements, SCMs, admixtures,
low water–cement ratio, and other measures to reduce permeability and moisture exposure
(162). In fact, several case examples exist of concrete pavements produced using RCA from
D-cracking and ASR-affected source concrete, where the pavement performed satisfactorily for
several decades (154, 155, 160, 162).
RCA is often produced from pavements and bridge components that have been exposed
to chlorides and other deicing or anti-icing chemicals. The presence of chlorides may cause
concerns related to potential accelerated set time and corrosion. However, field studies of
RCA concrete have shown no problems attributable solely to chloride content of RCA (162).
Nevertheless, the chloride content of the RCA that may have been exposed to excessive salt needs
to be evaluated, and epoxy-coated reinforcing steel needs to be used if the predicted chloride
content of the new concrete mixture is high enough to cause concern. Since chlorides are typi-
cally contained in the fine RCA particles produced from the near-surface portions of the source
concrete, reducing or eliminating the fines or washing the RCA to remove material passing a
No. 200 sieve is also an option (162).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

70   Concrete Technology for Transportation Applications

Figure 42.   Typical crushing and grading operation


for producing RCA on site (courtesy of G. Fick, Trinity
CM Services).

During the recycling process, provisions need to be made to reduce the amount of contami-
nant material included in the RCA (155, 162). Reinforcing steel and other materials such as
joint sealant need to be removed. Concrete containing low amounts of asphalt cement con-
crete patching material has been utilized to produce RCA, but limiting its presence is advisable.
Crushing and processing equipment (particularly the type of crusher) used to produce RCA
will influence the yield, properties, and characteristics of RCA produced (155). RCA needs to be
generally free of contaminants, and provisions to protect stockpiles from contamination need
to be implemented (162). A typical on-site crushing and grading process is shown in Figure 42.
RCA particles comprise aggregates and reclaimed cement mortar from the source concrete
(Figure 43), and therefore properties of RCA will be a function of both components. Relative
proportions depend on the source concrete materials, mixture proportions, and properties, as
well as the crushing and processing methods. The properties of the coarse aggregates contained
in the source concrete, the aggregate–paste bond, and the type(s) of crusher(s) used have been
shown to influence the ultimate composition of RCA (155, 162). Particle composition will also

Figure 43.   Coarse RCA (courtesy of M. Adams, New Jersey Institute of Technology).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   71  

vary by size, with larger sizes of RCA containing more aggregates from the source concrete, and
finer sizes of RCA containing a greater fraction of mortar. The crushing process will also influ-
ence the shape, texture, and gradation of RCA particles. In general, RCA particles tend to be
angularly shaped with a rough surface texture (162).
RCA typically has a higher absorption than conventional aggregate. This is driven by the
absorption of the reclaimed mortar, which is more porous and has additional surface area.
Smaller RCA particles tend to contain a greater mortar fraction with higher impact on absorp-
tion (164). The porosity of the reclaimed mortar also results in a relatively lower specific gravity
and unit weight of RCA concrete (162). RCA also tends to have a slightly higher abrasion loss
than conventional aggregates, attributed to the adhered mortar fraction and partially fractured
particles (155). The presence of fines (or crusher dust) on RCA particles can increase water
demand in new concrete mixtures and, in unbound base applications, the potential for pre-
cipitate formation in drain systems. Washing RCA will aid in reducing fines and the associated
issues (162). Field studies of pavement applications have shown that RCA concrete that includes
higher mortar content often exhibits more distress, and the reclaimed mortar content needs to
be considered and accounted for during mixture proportioning and trial batching (160).

RCA Concrete Production and Properties


In research and in practice, it has been shown that concrete mixtures containing RCA can
meet performance expectations for strength and durability (154, 155, 162, 165). The fundamen-
tals of mixture design and proportioning for RCA concrete mixtures do not differ from those
used for a conventional concrete mixture, and production of RCA concrete, including batch-
ing, mixing, placement, and finishing can also be very similar to production of conventional
concrete. However, certain characteristics of RCA that differ from conventional aggregates can
affect the properties of fresh and hardened RCA concrete differently, and these differences need
to be considered and accounted for in the design process (155, 162, 165).
Mixture proportioning procedures for RCA concrete mixtures can follow the same methodol­
ogies used for conventional mixtures. Guidance on the effects of RCA on the properties of
fresh and hardened concrete is presented in subsequent sections of this report as well as in a
number of guidance documents (155, 162, 165). General needs include ensuring that proper test-
ing to characterize RCA be performed, admixtures be used to address increased water demand
without adding water, mixture proportions be adjusted to accommodate the potential higher
variability in strength, and air content be adjusted to account for air voids in RCA mortar fraction.
Trial batching and adequate testing are key to achieving the desired performance (155).

Impact of RCA on Fresh Concrete Properties


Typical concerns associated with use of RCA are increased water demand and premature
stiffening, driven by the increased absorptivity of RCA. The angular particle shape and rough
texture of the crushed RCA can also result in reduced workability of a concrete mixture. ACI 555
indicates that for a given slump, RCA concrete will have a 5% additional free water requirement,
although an even greater increase in water demand can be expected if fine RCA is used (165).
Without the need to add water, workability and premature stiffening issues can generally
be addressed by using chemical and mineral admixtures, prewetting or presoaking the RCA,
adjusting past content, optimizing aggregate gradations, and limiting or eliminating use of
RCA as fine aggregate (155, 162, 165). ACPA recommends limiting inclusion of RCA as a fine
aggregate to 30% replacement level in order to reduce the potential for loss of workability (162).
Use of RCA in high amounts can result in mixtures that are difficult to finish by hand, although
mechanical finishing equipment is generally not affected (155, 164). When used as coarse
aggregate, RCA has minimal effect on bleeding, although bleeding is often reduced in mixtures

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

72   Concrete Technology for Transportation Applications

using RCA as fine aggregates (162). Also, the setting time of RCA concrete could be shorter than
that of a conventional concrete mixture, particularly when RCA fines are included (155, 166).
Although RCA should not affect the action of air-entraining admixtures, the air content of a
new mixture containing RCA will be affected by the air void system introduced to the new con-
crete through the RCA’s reclaimed mortar. Air content tests of fresh RCA-concrete mixtures
will be influenced by the air void system of the RCA, and thus the results need to be adjusted
using the aggregate correction factor in ASTM C231. Alternatively, the volumetric method
(ASTM C173) could be utilized (162).

Effect of RCA on Hardened Concrete Properties


The effect of RCA use on concrete mechanical properties and durability performance will
depend on the characteristics of the source concrete, RCA characteristics as produced, the type
of RCA used (coarse RCA, fine RCA, or both), and the level of replacement of conventional
aggregate. Use of RCA as coarse aggregate in a concrete mixture will generally reduce the com-
pressive strength, tensile strength, and modulus of elasticity compared to equivalent mixtures
using conventional aggregates (up to 24%, 10%, and 33%, respectively). When RCA is used as
fine aggregate, greater reductions in test values for these properties can be observed (up to 40%, 20%,
and 40%, respectively), compared to use of RCA as coarse aggregate only. However, decreases
in strength and the potential for increased variability can be accommodated by reducing the
water–cement ratio, use of admixtures, and other mixture proportioning considerations (162, 165).
RCA concrete tends to have higher values of the coefficient of thermal expansion (CTE), drying
shrinkage, creep, and permeability than equivalent conventional concrete mixtures. The effects
will be highly influenced by the characteristics of the RCA and the replacement type (coarse aggre-
gate only or coarse and fine aggregate), and percentage replacement of natural aggregate. Use of
RCA as fine aggregate tends to result in up to 15% reduction in unit weight, and significantly
affects shrinkage and permeability. This is a result of the increasing amount of paste needed with
the fine RCA (158, 159, 162, 167).
The permeability of RCA concrete can be higher than equivalent conventional concrete
mixtures, but the impact can range from negligible to very large (162). Use of RCA produced
from high-quality, low-permeability source concrete has been shown to mitigate this predicted
increase in permeability, along with measures to reduce water content and paste volume dis-
cussed in the previous section (165). Limiting RCA use as a coarse aggregate replacement can
significantly limit the anticipated increase in shrinkage of the concrete (162).
Similar to conventional concrete, a properly entrained air void system has been shown to sup-
port freeze-thaw durability of RCA concrete, and use of sulfate-resistant cements should aid in
prevention of issues with sulfate environments (165). The increased permeability of RCA concrete
can lead to a higher potential for carbonation, although this can be offset by use of a lower water–
cement ratio for the new mixture (165). ACPA (162) indicates that this increased permeability can
be “offset by reducing the water–cement ratio by 0.05 to 0.10” and/or by replacing some portion
of cement with SCMs. Also, effect of ASR in the RCA can be mitigated in the new concrete by
conventional mitigation approaches such as the use of SCMs and low-alkali cements (165, 168).
Overall, research has shown that “RCA concrete can be highly durable, even when the RCA
is produced from concrete with durability problems, provided that the mixture proportioning
(including chemical and mineral admixtures) is done properly and the construction (including
concrete curing) is of good quality” (162). The RCA user needs to be mindful that the impact of
RCA use on the durability performance of concrete is dependent on the characteristics of
the RCA source concrete, the RCA as produced, other mixture materials, and mixture proportions
(162, 167). In fact, some research has shown favorable changes in RCA concrete performance (169).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   73  

Construction
Practical and economic factors determine whether a project is a candidate for use of RCA.
A key consideration is whether the potential amount of RCA produced and utilized on a project
warrants the mobilization of a mobile crushing and grading plant to the site. The cost of
conventional aggregates, the project’s proximity to available stationary crushing and grading
plants, the amount of RCA to be utilized on the project, permitting requirements, project phas-
ing, and the space requirements of a mobile plant on the site (public vs. private right-of-way
land) each play a role in project scoping and determination of the potential use of RCA (153, 155,
157, 167). Alternatively, if RCA is to be utilized, the project costs and the environmental impact
of hauling material to a stationary crushing plant are also considerations (153).
Technologies to support in-place or on-site recycling, such as crack-and-seat, rubblization,
and on-grade crushing and processing are available, and economic and time-saving benefits
have supported increasing use of RCA in unbound applications or as a stabilizer in full-depth
reclamation techniques in recent years (155). Other commonly utilized unbound applications
for RCA include fill material, erosion control, and embankments. Despite the fairly widespread
use of RCA in unbound applications, use of RCA concrete in pavement applications is relatively
uncommon in the United States, and use of RCA concrete in other structural applications in
transportation, such as bridge components, is quite rare. Concerns with the consistency or
quality of the RCA, the increased water demand of the RCA due to high absorption, the potential
for chemical contamination, and the possibility of introducing preexisting durability issues into
the new concrete components are often cited (155, 157, 165).
Because of risk aversion, structural considerations, and other concerns, pavement applica-
tions appear most promising for increased use of RCA concrete. However, RCA can and has
been successfully used in new concrete pavements, and over 100 field sites exist where RCA
concrete has been utilized. Reza and Wilde (154) present an extensive list of existing RCA con-
crete projects, with supporting information on type of RCA, how RCA was used in the mixture,
pavement design, and performance. Field performance of RCA pavement projects is reported
by Gress and coworkers (160).
One successful project where RCA was used as a 100% replacement for both coarse and fine
aggregates is a section of I-10 near Houston, Texas (Figure 44). In 1995, 5.8 miles of continu-
ously reinforced concrete pavement (CRCP) were reconstructed using RCA concrete. RCA was
required to meet the same specifications for virgin fine and coarse aggregates. During a 2013
inspection, with the pavement approaching 20 years of age, no punchouts were present and
crack widths appeared tight (170).

Figure 44.   CRCP RCA pavement near Houston,


Texas (courtesy of A. Naranjo, Texas DOT).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

74   Concrete Technology for Transportation Applications

Use of RCA in concrete for structural applications such as bridges continues to be rare,
with concerns often expressed about issues affecting structural performance, including durabil-
ity and corrosion potential (171, 172). However, many research studies have been performed
demonstrating the potential for RCA concrete to perform satisfactorily in structural members
(165). These studies have provided strategies for mitigating the risks (perceived or real) in use
of RCA and have offered guidance and strategies to support increased use in structural appli-
cations. Economic and sustainability drivers may result in the increased appeal of RCA for
structural applications. In one notable recent example of use of RCA concrete in a transporta-
tion structure, RCA sourced from a demolished temporary detour bridge has been used as a
30% replacement of conventional aggregate in a portion of the foundations (mass concrete
shaft caps) of the new Willamette River bridges (173).

Performance Properties
For successful use in bound or unbound applications, RCA needs to be treated as an engi-
neered material (155). From a specification standpoint, some state agencies treat RCA similarly
to conventional aggregates, with specification requirements for RCA identical to those for
other aggregates. Several states have separate specifications for RCA or require additional test-
ing if material is not sourced from agency infrastructure. QC/QA tests for RCA are generally
the same as those used for conventional aggregates, and appropriate specification provisions
often include gradation, limits on deleterious substances, abrasion resistance requirements,
absorption, and specific gravity. For unbound uses, specifications for RCA are provided in
AASHTO M 319-02, which provides guidance for use of RCA sourced from both agency
infrastructure and from construction and demolition debris. Additional guidance is provided
by ACPA (162). Of note, use of the sodium and magnesium sulfate soundness tests is not
appropriate for testing RCA because of inconsistencies in the tests. Other soundness testing
procedures have been found appropriate by other agencies, some of which are described in
AASHTO M 319 and by Snyder et al. (155).
Specific gravity and absorption tests can be performed according to AASHTO T 85. How-
ever, since the specific gravity and absorption of RCA can be more variable than conventional
aggregates, AASHTO MP 16 provides limits on total variability for both of these tests, along with
additional guidance. AASHTO MP 16 presents limits for soundness tests, but also states that
alternative tests can be used. Testing for alkali-aggregate reactivity is suggested in accordance
with AASHTO T 303 and ASTM C586 for ASR and ACR, respectively, with alternative methods
also discussed in an appendix. Testing of RCA for D-cracking is suggested using AASHTO T 161.

Implementation
Conclusions of many field studies indicate that RCA concrete can be used to construct pave-
ments capable of providing performance equivalent to pavements constructed with conven-
tional aggregates. Appropriate design considerations, such as the use of load transfer devices,
shorter joint spacing to account for the increased CTE and other thermal properties, and con-
sideration of RCA properties in mixture proportioning have been recommended (154, 160).
Use of production processes that reduce the amount of reclaimed mortar in the RCA would
ensure that the RCA performs close to conventional aggregates in pavements (160). RCA con-
crete pavements demonstrating undesirable performance have failed due to mid-panel cracks and
design issues such as poor support layers, excessive slab lengths, and undoweled joints (159, 160).
Implementation of RCA concrete in the lower lift of two-lift pavements has been a practice in
Europe, with RCA concrete in the lower lift and virgin aggregate in the upper lift. Several two-
lift concrete pavements with RCA concrete in the lower lift have been constructed in the United

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   75  

States (174). These include a reconstructed section of US-75 in Iowa that provided over 40 years
of service, and more recently a section of I-70 in Kansas (155). In these applications, use of RCA
in the lower lift was found to be a viable alternative to conventional pavement construction from
both economical and sustainability standpoints, with cost savings associated with the RCA in the
lower lift helping to offset expenses associated with construction of a two-lift pavement.

Specifications and Guidelines


Specifications for use of RCA in new concrete mixtures is presented in AASHTO MP 16
and in ACPA (162). AASHTO MP 16 states that when using RCA for new concrete, enhanced
QA/QC procedures should be followed to ensure that deleterious materials are not present in
the RCA. The standard provides limits for deleterious substances as well as requirements for
physical property. The specified maximum Los Angeles (LA) abrasion for RCA is 50%, which is
slightly higher than the LA abrasion specified for virgin aggregates by some agencies. Gradation
requirements for RCA are typically the same as those required of conventional aggregates for
the same use. Guidance on use of RCA for concrete mixtures has been published by a number
of organizations (155, 162, 165, 166, 167).
Many state specifications also address construction-related concerns associated with RCA.
Appropriate and attentive aggregate stockpile management is necessary when using RCA,
because stockpiles must be protected from contamination and moisture application must be
monitored. Stockpile management has been developed for several states, including Michigan
(153, 155). Environmental concerns related to use of RCA include high pH and potential pol-
lutants in stockpile runoff and air quality and community impacts associated with crushing and
handling (155). In unbound base applications, drainage structures can be clogged by precipi-
tates, but can be mitigated by minimizing RCA fines and use of effective drainage designs (158).
Guidance on mitigating environmental concerns associated with RCA use during planning,
design, and construction is presented by Snyder et al. (155).

Limitations and Challenges


RCA is produced from crushed concrete that can be obtained from a variety of sources, where
the parent concrete may have been of high, medium, or low quality. Variation in the quality of the
parent concrete can produce RCA with unacceptably variable properties with respect to strength,
absorption, and particle stability. Low- or nonuniform-quality source concrete will produce RCA
that may not be suitable for reuse in concrete mixtures for pavements and structures (155, 165).
State agencies need to use RCA only in new concrete mixtures for pavements and structures if
they can be assured that the properties of the source concrete will support production of RCA of
acceptable quality. Concrete sourced from agency projects with known, acceptable performance
history is typically recommended for use as RCA. Properties of the source concrete can be tested
and verified to be adequate, if characteristics are not already known from historical records of
the source concrete. RCA sourced from a variety of projects or from unknown projects needs to be
used only as subgrade or fill material, and only if QA/QC testing indicates sufficient performance
characteristics and an acceptable amount of contaminant material. A detailed discussion of project
selection and scoping for concrete recycling in pavement applications, including a flowchart to
aid in characterization of the source concrete for use as RCA, is presented by Snyder et al. (155).
Using RCA in new concrete mixtures as a partial or full replacement for virgin coarse aggre-
gate typically requires prewetting the RCA, since its absorption can be high due to the paste com-
ponent of the aggregate particle. Without prewetting, the RCA can make it difficult to control
mixture workability and strength.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

76   Concrete Technology for Transportation Applications

Figure 45.   HESC for slab replacement.

When using RCA in jointed concrete pavements, the designer needs to include dowel bars
at transverse joints to provide adequate load transfer between adjoining slabs. In pavements
with undoweled joints, the recycled coarse aggregate can deteriorate due to the friction between
interlocking slabs at the transverse joint, leading to loss of load transfer and joint failure issues.

High Early Strength Concrete


Introduction
HESC is a concrete mixture that is designed to achieve a specified high early strength within
24 hours, and in many cases, in less than 12 hours (10, 175). HESC is commonly used to accel-
erate construction and repair of highway pavements (Figure 45) and repair of airport runways
to allow early opening of the facility to traffic (10, 176). In precasting plants (Figure 46), HESC
mixtures are used to achieve the required strength for early release of the prestressed strands
(177) and removal of forms in non-prestressed members.
HESC uses materials similar to those used in conventional mixtures. However, for the mixture to
achieve high early strength, several changes in the mixture design are required based on the specified
strength to open the pavement to traffic or to release the prestressed strands. Among mixture ingre-
dients and proportions that produce rapid strength gain are higher quantities of Type I/II cement,
lower water–cement ratio, high dosage rate of accelerating admixtures and HRWRs, and the use of

Figure 46.   High early strength SCC placement in


prestressed beam (courtesy of Dura-Stress, Florida).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   77  

Table 3.   Common ranges of mixture ingredients for HESC


in accelerated pavements (179).

20- to 24-Hour
Mix Characteristic 4- to 5-Hour Concrete 6- to 8-Hour Concrete Concrete
Cement type I or III I or III I or III
3
Cement content, lb/yd 560–895 (333–532) 715–885 (425–525) 675–800 (400–475)
(kg/m3)

Water–cement ratio 0.38–0.40 0.36–0.40 0.40–0.43

Accelerator Yes Yes None to Yes

Type III cement instead of Types I/II (175). Insulating blankets may be required to aid in developing
high early strength in cold weather or to accelerate strength gain (10, 178).
For pavement projects in cold environments, HESC must not only gain strength rapidly but
must also achieve long-term durability (175). In all projects requiring HESC, design of the pave-
ment and the concrete mixture, as well as construction methods must be compatible to ensure
that premature distresses including cracking and spalling are mitigated.

Concrete Materials and Mixtures


Mixture Proportioning
In proportioning HESC mixtures for accelerated paving, factors such as cement type and
content, aggregate size distribution, use of entrained air, concrete, and ambient tempera-
tures may influence early and long-term concrete strength, durability, and performance (178).
Table 3 shows combinations of mixture materials utilized for projects with different lane-
opening times (179). Table 4 shows a mixture design for high early strength SCC used to cast
a prestressed member (177). The 14-hour strength gain allows release of the tensioned strands
in prestressed beams.
A thorough laboratory investigation using trial batches and performing tests of key perfor-
mance characteristics is critical before utilizing an HESC mixture. The laboratory work needs
to determine plastic and hardened concrete properties using project materials and needs to aim
to verify the compatibility of all chemically active ingredients in the mixture. For example, not
every cement will gain strength rapidly (14). Also, with respect to the compatibility issue, an
NCHRP study concluded that it might be difficult to control the air content in a HESC mix-
ture that contains a large quantity of Type III cement, low water–cement ratio, and multiple

Table 4.   High early strength SCC mixture design


for a prestressed member (177).

Material Quantity
Cement, lb/yd3 (kg/m3) 740 (440)
Class F fly ash 175 (104)
Water–cement ratio 0.27
HRWR As necessary

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

78   Concrete Technology for Transportation Applications

admixtures (175). Therefore, proper testing of laboratory samples is needed not only to verify
strength, but also to evaluate the concrete’s performance using other appropriate tests such as
the test for resistance of concrete to rapid freezing and thawing (AASHTO T 161).

Cement
With proper proportioning, concrete mixtures using Type I, Type II, or AASHTO I/II, and
Type III portland cements can produce the required high early strength and durability perfor-
mance for accelerated concrete paving and precast members. The HESC mixture often requires
multiple admixtures such as HRWR, air entraining, and accelerating admixtures to provide the
needed fresh and hardened properties.
Use of cement with high levels of tricalcium silicate (C3S) and finely ground cement particles
will result in rapid strength gain. Tricalcium aluminate (C3A), although not contributing to
strength gain, is a catalyst to enhance the rate of hydration of C3S. Cement with high fineness has
an increased surface area, allowing more cement contact with mixing water and, consequently,
contributing to faster hydration and rapid strength gain (14).

Fly Ash
Classes C and F fly ash have been used in HESC with the purpose of improving workability
and flow, as well as contributing to long-term strength and durability. Class F fly ash has been
used in high early strength mixtures for prestressed members, while Class C ash has been mostly
used in accelerated pavement construction (14). Strength evaluation of HESC with Class F fly
ash must be performed using trial batches to verify achieving early strength development. This
is due to the lower heat of hydration generated from the pozzolanic reaction which can delay
early strength gain. Use of accelerators needs to be considered when the strength of concrete is
below that specified for lane opening (10).

Ground-Granulated Blast-Furnace Slag


Because of its cementitious properties and contribution to long-term strength and durability,
the use of GGBFS (slag) in accelerated paving is generally acceptable (10). However, the slow nature
of the cementitious reaction generates lower overall heat not conducive to very rapid strength
development. Therefore, in designing HESC mixtures with slag, tests must be performed on trial
batches to determine the optimum dosage of slag and the needed accelerating admixture to ensure
early strength development that meets agency specifications. Also, use of curing blankets may be
needed to aid in rapid strength development.

Admixtures
Air-entraining admixtures meeting ASTM C260 requirements are necessary for HESC
mixtures in freeze-thaw environments. However, care needs to be exercised when selecting
the dosage rate of the admixture. A high dosage of air entrainment will reduce the rate of
strength development and a low dosage may reduce the resistance to freeze-thaw action.
Therefore, control of air content is necessary for successful projects. According to ACPA, the
concrete mix should have between 4.5% and 7.5% entrained air, depending on the maximum
coarse aggregate size and the local climate (178).
Normal and high-range water-reducing admixture Types A, E, and F (ASTM C494) generally
provide the necessary properties for accelerated concrete paving. These admixtures tend to reduce
mixture water and increase efficiency of the cement hydration, thus contributing to workability at
placement and early strength gain. However, laboratory testing is essential to determine if a concrete
containing the admixture will develop the desired properties. For example, using excessive
dosage of high-range water-reducing admixtures may delay concrete setting and strength gain (10).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   79  

Accelerating admixtures aid in early strength development and reduce initial setting times
by increasing the reaction rate of C3A. Energizing this reaction generates additional heat to
increase the cement hydration to form more hydrated gel and achieve high early strength.
Accelerating admixtures generally consists of soluble inorganic salts or soluble organic
compounds and needs to meet requirements of ASTM C494, Type C or Type E (10). Cal-
cium chloride is used as an accelerator in pavement repairs and slab replacement when no
reinforcement is present. However, in reinforced pavements the use of calcium chloride may
be prohibited due to its corrosive effect. Agencies generally specify the noncorrosive or
corrosion-inhibiting accelerators such as calcium nitrite in patching or replacement of
reinforced slabs or in bridge repairs.

Construction
Curing and Temperature Management
The key to achieving rapid strength gain in HESC is to have an efficient cement hydration
that is continuously energized by internal and external heat. Therefore, it is important to con-
trol and reduce heat loss, especially in cold weather, and to mitigate moisture loss in hot and
windy weather conditions. This is accomplished by implementing effective curing provisions to
maintain satisfactory moisture and temperature conditions in concrete for a period sufficient to
ensure proper hydration and rapid strength development (10).
The most effective and practical curing provisions for concrete pavements and replacement
slabs are curing compounds and the use of curing and insulating thermal blankets (176). In
precast plants the use of an external heat source or steam curing of forms has proven effective in
achieving the required high concrete strength to allow for early release of the tensioned strands
and/or removal of the formwork (175, 180).

Curing Compound.   Liquid membrane-forming curing compounds for HESC need to meet
ASTM C309 requirements. Typically, a white-pigmented compound (Type 2, Class A) is applied
to the surface and exposed edges of the concrete pavement. In mountainous and arid climates,
agencies often specify a slightly heavier dosage rate of resin-based curing compound meeting
ASTM C309, Type 2, Class B requirements (10).
The recommended application rate of curing compound for accelerated paving projects
ranges from 100 to 150 ft2/gal (2.5 to 3.75 m2/L). Thinner pavement requires a thicker coating of
the curing material (181). To reduce loss of mix water from the paved surface, the curing com-
pound must be applied as soon as the final surface finish is completed. Evaporation retarders
may also help to prevent very early loss of mix water when applied immediately after concrete
placement and initial strikeoff.

Curing and Insulating Blankets.   Some states also require the use of curing or insulating
blankets to supplement the action of the curing compound. Curing and insulating blankets
reduce the heat loss and moisture evaporation from the finished pavement surface (182). The
purpose of the insulation is to aid early strength gain in cool ambient temperatures by reducing
heat loss from the concrete. The insulating blanket needs to consist of a layer of closed-cell poly-
styrene foam with another protective layer of plastic film. Additional blankets may be necessary
for temperatures below about 40oF (4oC).
In warm weather with mild cool nights, thinner curing blankets are sufficient to aid in pre-
serving heat and moisture in the pavement. Florida, for example, requires the use of one and, if
necessary, two layers or more of white burlap-polyethylene blankets (183).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

80   Concrete Technology for Transportation Applications

Joint Sawing
HESC sets fast and gains strength early and rapidly. Therefore, there is a short window for
saw cutting joints in an accelerated pavement project. Timely joint sawing is important to
avoid uncontrolled cracks that develop due to tensile stresses induced by drying shrinkage
and thermal effects (10).
Concrete placed in early morning often reaches higher maximum temperatures than con-
crete placed in the late morning or afternoon, because the former receives more radiant heat
throughout the day. This suggests that concrete placed early in the morning will generally have
a shorter sawing window. Nighttime paving allows a longer window for joint sawing in the early
morning since the maximum concrete temperature will not coincide with the maximum ambi-
ent temperature (10).
Proper timing of joint sawing is important to avoid joint spalling when sawing prematurely
or uncontrolled cracking when sawing too late. A sacrificial slab placed at the project site
and then sawcut at different time intervals may determine proper timing of the sawcut for the
paving project. The nondestructive maturity test (ASTM C1074 and C918), which is normally
used to determine strength of concrete in pavements and structures, may be performed at
different time intervals to identify the beginning and end of the joint sawing window.

Strength Requirement for Lane Opening


The critical issue in accelerated pavement construction is determining the lane opening time
when traffic is allowed to use the newly placed pavement. This decision needs to be based on achiev-
ing a specified concrete strength and not arbitrarily on a set time from concrete placement (184).
Some state highway agencies use flexural strength test results and requirements for lane
opening, but most states use compressive strength (10). There is no clear consensus on what
strength is required for opening accelerated concrete pavements to traffic. Factors such as
expected traffic loadings, edge conditions, pavement geometry and project type, new construc-
tion or rehabilitation affect the pavement strength required for traffic (179).
A review of state highway practices suggests that a range of values is often specified for lane
opening of accelerated pavements. Compressive strength values of 2,000 to 3,000 psi (13.8 to
20.7 MPa) and flexural strength values of 290 to 400 psi (2.0 to 2.8 MPa) have been reported
(185). These values are conservative and represent the minimum strength needed to open the
repaired/replaced slabs to traffic. The required ultimate strength of the HESC would normally
be reached from 24 hours to several days. In Table 5 minimum opening-strength requirements
for various slab thicknesses are shown (186). Thicker slabs require lower strength.

Table 5.   Minimum opening strength for full-depth slab


replacements (186).

Strength for Opening to Traffic, psi (MPa)


Repair Length < 3 m (10 ft) Slab Replacements

Slab Thickness 3rd-Point 3rd-Point


in. (mm) Compressive Flexural Compressive Flexural
6 (150) 3,000 (20.7) 490 (3.4) 3,600 (24.8) 540 (3.7)
7 (175) 2,400 (16.5) 370 (2.6) 2,700 (18.6) 410 (2.8)
200 (8.0) 2,150 (14.8) 340 (2.3) 2,150 (14.8) 340 (2.3)
225 (9.0) 2,000 (13.8) 275 (1.9) 2,000 (13.8) 300 (2.1)
250+ (10.0) 2,000 (13.8) 250 (1.7) 2,000 (13.8) 300 (2.1)

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   81  

A few highway agencies, including Florida and Georgia, have revised their required opening
strengths to values lower than 2,000 psi (13.8 MPa). Florida specifies 1,600 psi as a minimum
opening strength requirement for replacement slabs (183). However, state DOTs are encour-
aged to evaluate the appropriateness of using lower opening strength values that are suitable for
their climatic conditions and project types (179). The use of maturity meters or pulse-velocity
devices for monitoring the in-place concrete strength is recommended as part of that process
(178, 182). Also, HIPERPAV computer software (www.hiperpav.com) may be helpful in identi-
fying conditions that may be potential contributors to random cracking in HESC mixtures used
in accelerated paving (187).

Very High Early Strength Concrete Repair Materials


and Alternative Cementitious Materials
Introduction
The demand to maintain and extend the service life of existing infrastructure has driven devel-
opment of a wide variety of VHESC or cementitious repair materials that gain the required
strength (often to support lane opening) from 1 to 4 hours. Selection of the most suitable
approach will depend on many factors, including the condition of the existing facility, the type
and extent of pavement deterioration, service life goals, and other site-specific and agency con-
siderations. Although identification of the appropriate approach is beyond the scope of this
synthesis, this section presents a basic overview of the cementitious materials often utilized for
such purposes.

Need for Technology


In transportation applications, cementitious repair materials are utilized in pavements and
bridges subjected to heavy traffic with challenges in closing lanes to perform repairs with con-
ventional concrete that requires longer cure time. Typical application of such materials includes
rapid repair of pavements in highways and airports and in bridge deck patching, as well as other
structural and nonstructural applications. In airfield applications, rapid repairs of distressed
areas are necessary to reduce the potential damage to aircraft from concrete debris and/or to
allow quick opening of runways to resume military flights (188).
Repairs to concrete infrastructure can be performed with conventional concrete mixtures, pro-
vided that time, weather, traffic, and construction conditions allow for batching, mixing, place-
ment, surface finish, and early strength gain. However, transportation agencies, airports, and the
military are increasingly required to perform the repairs more rapidly and open the pavement,
bridge, or runway to traffic much sooner to keep the traffic disruption period to an absolute mini-
mum (4 hours or less). Very often, the most optimal lane closure window to perform the work is
during nighttime. The short time frames allowed for work generally require that repairs be per-
formed using VHESC materials that cure and gain strength very rapidly, often in less than 1 hour.

Materials, Mixture Parameters, and Common Test Procedures


Concrete repair often requires addressing a variety of engineering challenges. For example,
repairs may need to be performed in cold or hot weather, which affects the fresh and hardened
properties of the repair materials with respect to setting time, strength gain, shrinkage, bonding
to substrate, and overall longevity of the repair. In addition, construction challenges associ-
ated with the repairs need to be addressed. These include achieving proper installation condi-
tions, such as adequate removal of existing materials and achieving the appropriate substrate

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

82   Concrete Technology for Transportation Applications

conditions (clean, dry, appropriate bonding texture and/or bonding agent) to achieve adequate
bonding of the repair material (189, 190, 191).
Concrete repair has been called “as much an art as a science,” and engineers and contractors
traditionally receive minimal formal training, relying more upon experience and lessons learned
from previous trials (189). The conditions associated with repairs often require rapid setting
and very high strength gain of the materials, while allowing adequate time and workability for
proper placement and finish. Because of the rapid and often high strength gain of many repair
materials, durability performance can be compromised. Many cementitious repair materials
have historically been associated with high heat of hydration, issues with dimensional stability or
shrinkage, bond failure, and distresses resulting from high stress concentrations due to the high
stiffness of the hardened material (190). In fact, the rapid hydration of early strength cements
using latex admixtures has been linked to significant shrinkage (up to 80% of the total volume
change) during the first several hours of curing (191).
Preferred performance criteria for repair materials for bridge decks and pavements are sum-
marized as follows (190):
• Very high early strength,
• Long-term durability,
• Installation efficiency,
• Exhibiting no surface damage,
• No internal cracking,
• No separation from underlying pavement or bridge deck (debonding), and
• No other indications of distress.
Repair materials can be classified into the following general types (179, 189, 190, 192):
• Cementitious concrete including Type III portland cement, ultrafine portland cement,
high-alumina cement, and expansion-producing grouts;
• Polymer-modified concrete including additives such as styrene butadiene rubber, vinyl
acetate, acrylic, and magnesium phosphate; and
• Polymer concrete and resinous mortars including epoxy, polyester, acrylic, and polyurethane
compounds.
As can be surmised from the above list, the variety of types and formulations for concrete
repair material presents a wide range of fresh, early age, and late age performance character-
istics. These repair materials can have proportions and early age performance characteristics
that are (1) similar to conventional concrete, (2) gain early high strength during the first day or
over several days, or (3) set very quickly and gain strength early and rapidly to levels that can
support traffic within a few hours (such as high-alumina cements and magnesium phosphate
materials). For rapid-hardening repair materials, traffic opening times vary by product and
installation conditions, but can range from 0.5 hour to 24 hours (179, 190). The rapid hard-
ening of some materials requires expedient placement, since hardening and set can occur as
quickly as 10 to 30 minutes (193). Recent initiatives to support improving the sustainability
of construction have resulted in increased interest in the use of alternative (non-portland)
cements, which can provide early age properties and good to superior durability performance
in a variety of applications while also reducing environmental impacts associated with their
production (194).
ASTM C1600/C1600M-17, “Standard Specification for Rapid Hardening Hydraulic Cement,”
provides the requirements for strength, volume stability, and durability properties of rapid-
hardening hydraulic cements, as shown in Tables 6 and 7. Rapid-hardening hydraulic cement
is defined according to ASTM C1600 as a hydraulic or blended hydraulic cement that exhibits

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   83  

Table 6.   Standard physical requirements for rapid-hardening


cements (from ASTM C1600).
Cement Type
Property Age URH VRH MRH GRH
Compressive strength (ASTM C109), 1½ hours 3,000 1,700 — —
minimum psi (MPa) (21) (12)
3 hours 4,100 2,200 1,500 1,000
(28) (15) (10) (7)
6 hours — — 2,000 1,500
(14) (10)
1 day 5,100 3,500 2,500 2,000
(35) (24) (17) (14)
7 days 6,000 4,100 4,100 3,500
(41) (28) (28) (24)
28 days 8,300 5,100 4,500 4,100
(57) (35) (31) (28)
Drying shrinkage (ASTM 596), 7 days 0.06 0.06 0.08 0.10
max %
28 days, air 0.07 0.07 0.09 0.12
storage

Minimum time of final set (ASTM C191),a apparatus (min) 10 10 10 10

Autoclave (ASTM C151), max expansion (%) 0.8 0.8 0.8 0.8

Note: Dash = no information provided in reference to indicate that the item was not used or applied.
a
The initial setting time typically ranges from 10 to 45 minutes for rapid-hardening cements of various types
and composition.

rapid strength gain during the first 24 hours of hydration, with or without other constituents,
processing additions, and functional additions.
The specification covers four types of rapid-hardening cement, shown in Tables 6 and 7.
They are as follows:
Type URH. Ultrarapid hardening for use where ultrahigh early strength is desired.
Type VRH. Very rapid hardening for use where very high early strength is desired.
Type MRH. Medium rapid hardening for use where mid-range rapid-hardening high early
strength is desired.
Type GRH. General rapid hardening for use when the higher strength properties of a
Type VRH or a Type MRH cement is not required.

Table 7.   Optional requirements for rapid-hardening cements


(from ASTM C1600).
Cement Type
Property Age URH VRH MRH GRH
Sulfate expansion 6 months, max % 0.05 0.05 0.05 0.05
(ASTM C1012)a 1 year, max % 0.10 0.10 0.10 0.10
ASR expansion 14 days, max % 0.020 0.020 0.020 0.020
(ASTM C441)b 56 days, max % 0.060 0.060 0.060 0.060
Heat of hydration 7 days, max kJ/kg 250 (60) 250 (60) 250 (60) 250 (60)
(ASTM C186) (kcal/kg)
28 days, max kJ/kg 290 (70) 290 (70) 290 (70) 290 (70)
(kcal/kg)
Expansion in water 14 days, max % 0.10 0.10 0.10 0.10
(ASTM C1038)
a
In the testing of these cements, testing at 1 year shall not be required when the cement meets the 6-month
limit. Cement failing the 6-month limit shall not be rejected unless it also fails the 1-year limit.
b
The test for mortar expansion is an optional requirement to be applied only at the purchaser’s request and is
not required unless the cement will be used with alkali-reactive aggregate.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

84   Concrete Technology for Transportation Applications

Alternative Cementitious Materials


There has been increasing interest in use of alternative cementitious materials (ACMs) for
a variety of sustainability and performance-related reasons (194). The rapid hardening and
strength gain of several types of ACMs such as calcium sulfoaluminate cements (CSAs), calcium
aluminate cements (CACs) and alkali-activated (AA) binders has generated interest for use in
slab replacements and accelerated concrete pavements (195). Several ACMs have been linked
to enhanced durability performance, although limited field studies exist, and claims of supe-
rior durability have not been investigated to a full extent (196). Although recent studies have
provided insight into the mechanisms and performance of ACMs, only limited use has been
observed to date (194, 195, 196).
Guidance on use of ACMs is presented in ACI ITG-10R-18, “Practitioner’s Guide for
Alternative Cements,” which presents an introduction to ACMs, provides information on
the properties and applications of these materials, presents selected case studies, and provides
guidelines for use (194). Performance measures discussed as part of ACI ITG-10R-18 include
the environmental impacts of ACMs, life-cycle cost considerations, initial cost considerations,
and functional performance. A comparison of commercially available alternative cement tech-
nologies, including CAC, CSA, AA, and several other types of ACMs, is provided in Table 8,
as presented in ACI ITG-10R-18 (194). The publication provides a detailed description of
selected types of ACM technologies, along with testing requirements to support specifications
and QA/QC.

Construction
As can be surmised from the above list, the variety of types of concrete repair material for-
mulations present a wide range of fresh, early age, and late-age performance characteristics,
which can also affect construction. Project characteristics drive selection of materials that
provide the required strength (within the required time to opening) and durability perfor-
mance, while also addressing constructability and cost constraints. Many cementitious repair
materials are sold in packaged form that requires preparation of small batches on site. For
larger quantities of repairs, some materials can be batched in mobile volumetric mixers or at
ready-mix plants.
ACI provides a summary of repair and overlay material properties, selection, and a list
of essential steps for repair in ACI 546.3R-14, “Guide to Materials Selection for Concrete
Repair” (189). Guidance specifically on polymer-modified concrete mixtures is available in
ACI 548.3R-09, “Report on Polymer-Modified Concrete” (197). This report provides mixture
proportioning guidance for these types of concretes, which are a common choice for larger
repairs where cementitious ready-mix repair material can be batched (197). Guidance for
materials used for pavement maintenance and preservation activities, as well as approaches for
different repair types including slab stabilization or jacking, partial-depth repairs (patches),
and dowel bar retrofitting, is presented in the “Concrete Pavement Preservation Guide” (179).
Use of bonding agents for partial-depth repairs is also discussed in ACI 546.3R-14 and in other
publications (179, 198). Other references on partial-depth repairs include those published by
FHWA, the National Highway Institute (NHI), and ACPA (198, 199, 200).

Performance Properties
Mechanical properties and durability performance characteristics vary by formulation, place-
ment conditions, and for prepackaged products, whether the material is extended with fine
aggregate or small-size coarse aggregate (often “pea gravel”). For larger repairs, some materials

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   85  

Table 8.   Comparison of commercially available alternative cement technologies


(from ACI ITG-10.1R-18).

CAC CSA MOC MPC AAFA AAS SCC


Hydraulic Yes Yes No No No No No
Main Calcium Ettringite, C- Hydro- Insoluble Variable, C- Variable, C- Ettringite,
hydration or aluminate S-H, CH magnesite ammonium A-S-H, N-A- A-S-H C-S-H
reaction hydrates and S-H
products C3AH6 and phosphate
AH8 phases
Set time Rapid Rapid and Same or Rapid Rapid Same or Same or
relative to expansive longer faster faster
portland
cement
Strength High early, High early, High early, Low early, High early, Low early, Low early,
late strength slower rate late strength moderate late late strength late strength late strength
comparable of late comparable strength comparable comparable comparable
to portland strength gain to portland compared to to portland to portland to portland
compared to portland
portland
Key Good sulfate, Good sulfate Good fire, Good sulfate, Good sulfate, Good sulfate, Good
durability ASR resistance abrasion, ASR acid, ASR corrosion, sulfate
attributes resistance ASR performance resistance acid, ASR resistance
Carbonation resistance resistance
Abrasion rates high Good fire Good fire
resistant Good resistance resistance Good fire
freeze-thaw resistance
resistance
Concerns Conversion Carbonation Loses Significant Performance Performance Limited
reactions affecting strength heat varies with varies with field
increase corrosion when evolution fly ash slag source experience
porosity and resistance exposed to source
reduce water at
strength over Possible early ages
time thaumasite
formation Significant
Significant heat
heat evolution evolution

Applications Refractory Structural, Patching, Fireproof Same as Same as Same as


concrete, precast, wallboard coatings, portland portland portland
sulfate and cold-weather patching cement cement cement
acid
resistance
Note: C-S-H = calcium silicate hydrate, C-A-S-H = calcium aluminate hydrate, N-A-S-H = sodium aluminate hydrate, CH = calcium
hydroxide, ASR = alkali-silica reactivity.

can be provided in “super sacks” of several thousand pounds of preblended material. Agencies
typically require prequalification of prepackaged products in order to appear on an approved
materials list, which can be developed independently and/or with the support of documentation
from the National Transportation Product Evaluation Program (NTPEP).
The cost of repair materials varies greatly, and some products can be quite expensive and
cost-prohibitive for larger repairs (179, 190). For economic and performance reasons, high-
quality conventional concrete is generally the most appropriate material for repairs when
given reasonable time to cure and gain the lane opening strength (179). However, in many
cases the lane closure time is limited to less than 4 hours and thus requires very high early
strength repair products.
Because of the variety of repair conditions and products that exist, specifiers can find it
challenging to identify the appropriate performance criteria and tests to ensure good results

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

86   Concrete Technology for Transportation Applications

(179, 197). Testing of these materials will provide confidence to the design engineer that the
material will meet performance needs.
ACI 548.3R-09 provides an extensive table of information on the performance require-
ments of repair materials replacing portland cement, including a summary of available test
methods and test values (197). Also provided in this document is a summary of the changes
in material properties of cementitious repair materials when a variety of modifications are
performed, such as addition of a variety of chemical and mineral admixtures, SCMs, and
fibers (197). The U.S. Army Corps of Engineers Engineer and Research Development Center
(ERDC) recommendations for QA/QC testing for cementitious materials includes compres-
sive strength, bond strength, modulus of elasticity, volumetric expansion, shrinkage poten-
tial, CTE, and time of setting (201, 202). Other testing recommendations have been suggested
by the NTPEP and Delatte et al. (190), Lesak (203), and Susinskas (204).
Although many types of repair materials have been successfully utilized for several decades,
relatively few studies on performance of repair materials exist, and a better understand-
ing of the long-term performance of repair materials has been cited as a need by several
researchers (190, 194, 201). For proprietary products, issues have been caused by changes in
their formulation, and occasionally procurement difficulties have been encountered (188).
Other challenges include the inability to batch large quantities of the products at one time,
resulting in the potential for cold joints to form in the patch between consecutive batches
(190, 201, 202). Issues facing the industry, including test methods and reporting, curing
procedures, product limitations and warnings, standardized industry acceptance, bond to
substrate material, corrosion potential, and structural repairs, are discussed in Appendix A
of ACI 546.3R-14 (189).
In recent studies, ERDC performed laboratory and field testing of products for mechani-
cal properties and durability performance to evaluate their use in rapid repairs for airfield
pavements. This project resulted in development of a recommended testing protocol to guide
specifications, including a table of proposed test requirements and performance thresholds for
emergency (temporary) repairs and permanent repairs for airfield pavements (201, 202). Results
of field studies using full-scale traffic tests indicated that a wide variety of repair material types
provided suitable performance for rapid repair of small, full-depth sections of PC (188, 193). To
address concerns about changes in material properties due to product reformulations, periodic
retesting is suggested (188). Additional requirements for polymeric repair materials have been
published by the Department of Defense (205).
A study for the Ohio DOT (190) found that all materials selected for the study (includ-
ing the lowest cost material) performed quite well, indicating that the use of the higher cost
materials may not be necessary for most installation conditions. A 2-year field study revealed
that measured and observed distresses tended to be attributable to substrate conditions, and
not to the repair materials (203, 204). Laboratory testing indicated that there may be perfor-
mance differences between the materials, but this was not observed in the field (206). Materials
investigated in that study did not perform well when used to repair asphalt pavements, and
therefore bituminous repair products were recommended for this application (190). Specifi-
cation recommendations and a draft specification based on this work are presented in Delatte
et al. (190) and Woods (207).
Project descriptions and findings of field studies of highway infrastructure constructed using
ACMs is presented in Burris et al. (195). Case studies include two Los Angeles, California,
installations of pavement constructed using CSA and CSA–portland cement blend (85% CSA/
15% PC) which have shown good performance, and a CAC concrete pavement repair in
Chicago, Illinois, that also appears to be performing well.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   87  

Performance-Engineered Concrete Mixtures


Introduction
State highway agencies are faced with the challenge of maintaining an aging infrastructure
with increasingly limited resources. A key to ensuring the integrity of concrete highway infra-
structure includes construction of new infrastructure and repairs with concrete mixtures that
provide durable performance and extended service life. Historically, concrete has been speci-
fied using three criteria (slump, air content, and compressive strength) which are only loosely
correlated with successful, long-term performance. Highway engineers are focusing more on
enhanced concrete durability as a means of reducing maintenance and replacement costs.
Enhancing durability can be achieved and assured by careful selection of concrete ingredients
as well as utilizing advanced performance testing for meaningful acceptance criteria. Although
concrete science and testing technology has advanced over the past decades, state specifications
have often not been modified to reflect these advancements (11).
Consistent with the focus of Moving Ahead for Progress in the 21st Century (MAP-21) legis-
lation on performance, as well as increased interest in use of performance specifications, there
is a desire by FHWA, public agencies, and industry to move toward performance-engineered
construction materials. One area of focus on materials performance is the work toward the
development of performance-engineered concrete mixtures.
PEMs include optimized mixture designs (materials selection, gradation, cement content,
etc.) that are engineered to meet or exceed design requirements, and are predictable, durable,
and have increased sustainability (208). The key features of PEM include (209):
• Design and field control of concrete mixtures around engineering properties related to
performance;
• Development of practical, performance-based specifications;
• Incorporation of this knowledge into an implementation system (Design/Materials/
Construction/Maintenance); and
• Validation and refinement by performance monitoring.

Concept and Benefits


In the past century, many changes have been made to concrete mixtures and materials.
Concrete materials have become more complex, using a wide variety of admixtures and SCMs,
sometimes in combinations (210). Exposure conditions, including increased traffic, adverse
chemical reactions from chlorides in sea environments, and increased use of deicing chemicals
in winter environments have affected long-term durability and service lives of many concrete
structures (211). Advances in construction technologies have resulted in changes in the way
concrete is batched, placed, consolidated, and finished. Also, accelerated schedules are placing
increased demands on the early age performance of concrete. Because of impacts of these
various factors on long-term concrete performance of concrete infrastructure, FHWA, state
highway agencies, and many industry stakeholders agree that there is a need to re-address
the way that concrete mixtures are specified and tested.
Extensive research in recent decades has led to new understanding of concrete deteriora-
tion mechanisms, advancements in concrete mixture design, and development of better field
and laboratory tests for QA/QC. With this new knowledge, an FHWA initiative is under way
to improve performance of concrete infrastructure through PEMs. This initiative has resulted
in development of AASHTO specification and commentary, AASHTO PP 84, “Standard Prac-
tice for Developing Performance Engineered Concrete Pavement Mixtures,” which provides

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

88   Concrete Technology for Transportation Applications

a framework and guidance for state highway agencies to develop a specification for PEMs
that focuses on measurement and acceptance of concrete based on characteristics that have
been linked to satisfactory long-term durability performance of the concrete (212). Although
developed for pavement concrete mixtures, the approach outlined in AASHTO PP 84 could be
extended to include specifications for PEMs utilized for other infrastructure, such as bridges,
barriers, and lower-grade uses, as well.
Performance-related specifications provide agencies the ability to obtain the desired con-
struction quality while allowing contractors greater control and flexibility (208). For instance,
current prescriptive specifications for minimum cement content and rate of strength gain may
preclude the acceptance of mixtures that have superior economy, durability, and satisfactory
mechanical performance, but contain high proportions of SCMs. The provisions included in
AASHTO PP 84 are presented in a format that allows state agencies flexibility in selecting the
tests and requirements most applicable to their states. Recommended uses of the PEM tests, such
as for mixture qualification or for acceptance, are also suggested. An appendix to the standard
provides additional context, technical information, and guidance.
Performance-related specifications require measurement of key properties and perfor-
mance characteristics. For performance specifications to be successfully utilized, QA/QC
tests need to be performance related, rapid, effective, reliable, and inexpensive (209). Recent
advancements in testing technologies have provided means of more directly measuring the
properties of concrete mixtures that have been linked to successful field performance (210).
A number of state agencies are using and evaluating new, rapid, early age testing technologies
such as resistivity, sorptivity, and air void system analysis that support development and
use of PEMs. Ongoing concrete materials research is providing state highway agencies data to
support the use of PEMs. However, additional work is needed to identify appropriate per-
formance measures, performance goals, and QA/QC protocol. The capabilities of these tests
to evaluate the durability performance of concrete mixtures is improving as state highway
agencies build sufficient data to correlate the test results with field performance.

Materials, Mixture Parameters, and Specifications


As mentioned earlier, tests used for specification and acceptance of concrete mixtures have
typically been tests for slump, air content, and strength. It has generally been believed that
strength could be used as a “quasi-indicator of durability” (210). However, research and field
experience has shown that slump, air, and strength are not reliable predictors of long-term
performance (212).
To endure stresses related to field exposure, concrete needs to exhibit characteristics that
indicate good resistance to freezing and thawing and to chemical attack from the corrosive
deicing salts and chlorides to the extent to which it will be exposed during its service life (211).
Additionally, the concrete cannot be susceptible to deleterious alkali–aggregate interaction
(such as ASR) (213).
Although none of these characteristics is easily measured directly, it has long been known that
durable concrete is associated with several performance characteristics measurable in a labora-
tory setting. Low permeability, resistance to cracking and transport of deleterious substances,
aggregate stability, and an adequate air void system are examples of these performance charac-
teristics (11, 214). Additionally, to mitigate placement defects that result in the poor durability
of what would otherwise be a satisfactorily performing concrete mixture, workability needs to
be considered (11, 209).
Conventional mixture designs using ordinary portland cements and quality aggregates can
provide good durability performance if they are properly proportioned with low water–cement

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   89  

ratio, have good workability, and take advantage of admixtures to create an adequately dispersed
air void system (214). Additionally, SCMs such as fly ash and slag have been shown to provide
enhanced durability performance (reduced permeability and mitigation of ASR). Established
and emerging mixture proportioning and test methods included in AASHTO PP 84 (212) will
help to ensure that mixtures will meet performance expectations.
A key feature of the AASHTO PP 84 specification is that it provides a menu of potential speci-
fication provisions that address six key performance-related properties (shown below), with rec-
ommended test methods that state highway agencies can select (or omit) as they desire. This
approach allows state highway agencies to incorporate knowledge of local historical performance,
risk tolerance, and agency preference into a durability-based specification. For many performance
requirements, an agency can select from either a prescriptive or a performance approach.
The six key performance requirements included in AASHTO PP 84 (212) include:
1. Concrete strength. Despite not always being directly indicative of long-term performance,
the strength of concrete continues to be an important specification parameter. AASHTO PP
84 suggests use of either flexural or compressive strength (or both) for mixture qualification
and for acceptance.
2. Reducing cracking due to shrinkage. AASHTO PP 84 suggests several specification provi-
sions to reduce cracking, including a prescriptive measure of limiting the volume of paste in
a paving mixture to 25%. A performance test that could be selected includes the unrestrained
volume change (AASHTO T 160). Other conventional and emerging test methods such as
the restrained ring tests and a probability of cracking method are discussed in the appendix
of AASHTO PP 84.
3. Durability of hydrated cement paste for freeze-thaw durability. AASHTO PP 84 suggests the
use of a prescriptive water–cement ratio limit (0.45) or acceptable performance using one of
several other currently utilized or emerging rapid test methods. These methods include fresh
air content using the conventional pressure or volumetric air meter (AASHTO T 152 and
T 196), the Super Air Meter (SAM) (AASHTO TP 118), and tests related to time of critical
saturation (ASTM C1585) and deicing-salt damage (215). Other prescriptive specification
provisions suggested for protecting concrete from deicing salts include use of SCMs at a sug-
gested replacement rate of 30%, and application of sealers (AASHTO M 224). Measures to
protect joints from damage caused by calcium oxychloride formation include tests to quantify
the amount of calcium oxychloride in the cement paste (AASHTO T 365) (216, 217).
4. Transport properties. A prescriptive measure of maximum water–cement ratio (limiting to
less than 0.45 or 0.50) is suggested, based upon freeze-thaw conditions. Results from several
research projects showed that test results of the surface resistivity meter (AASHTO T 358)
correlate with the well-established but time-consuming rapid chloride ion permeability test
(ASTM C1202) (218). However, both of these electrical tests have limitations associated with
pore solution ionic concentration, temperature effects, sample geometry, degree of satura-
tion, and storage. AASHTO PP 84 extends the use of resistivity meter by suggesting use of a
formation factor (F-factor) to assist in normalizing the results of surface resistivity testing.
5. Aggregate stability. Prevention of deleterious aggregate-related issues such as D-cracking,
ASR, and ACR are addressed in AASHTO PP 84. ASTM T161 and ASTM C1646 are suggested
for screening aggregates for D-cracking. AASHTO PP 84 suggests use of the approaches out-
lined in AASHTO R 80 to prevent and mitigate ASR and ACR.
6. Workability. Although not a measure of durability performance, considerations for assessing
workability are included in the AASHTO PP 84 guide specification due to the linkage between
inappropriate workability and construction-induced issues such as poor consolidation, edge
slump, segregation, and degraded air void system quality. Two emerging methods of assessing
concrete workability of low-slump paving mixtures suggested in AASHTO PP 84 are the Box
test (219) and the Modified V-Kelly test (AASHTO TP 129) (220).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

90   Concrete Technology for Transportation Applications

Construction
AASHTO PP 84 provides mixture proportioning guidance to assist in developing PEMs.
Commentary provided within the standard and appendix provide insight into strategies
to meet performance goals, such as optimizing aggregate gradation to assist in reducing
paste volume, and meeting strength requirements while simultaneously economizing the
mixture. Requirements for mixture qualification and mixture acceptance are also presented
in AASHTO PP 84 (212).
Performance specifications tend to shift risk from the agency to the contractor, with the
contractor in turn benefiting from the opportunity to innovate. AASHTO PP 84 details the
required QC activities to be performed by the producer, which include development, approval,
and implementation of a QC plan. This QC plan needs to include details on the methods and
frequency of monitoring and testing as well as data management and reporting tools such as
control charts. The QC plan will communicate to the agency how the contractor intends to meet
the specification requirements (210).
A QC plan to support PEMs will reduce risk for all parties and can maximize the economic
and performance benefits associated with the mixture. Education and training (of both agency
and contractor personnel), use of shadow and pilot projects, a mixture qualification/verification
procedure, and QC tools such as control charts are important parts of a QA program for
PEMs (210).
With proper mixture design, control, and testing, construction considerations for use of
PEMs in transportation infrastructure components need not differ significantly from con-
struction considerations for conventional concrete mixtures. As with other types of concrete,
appropriate construction techniques must be utilized and adequate curing must be performed
in order to ensure development of the desired properties.

Performance Properties
Properties of interest for PEMs will depend on the agency’s goals, preferences, and risk
tolerance, as well as the project constraints. Performance specifications allow the contractor
flexibility to meet contract requirements, encouraging innovation and potential cost savings
for the owner (210). Agencies will need to continue to review and approve PEMs as they
would other types of concrete. Guidance for establishing a QA program for PEMs is outlined
in Cackler et al. (210). This publication also describes the importance of development and use
of contractor QC plans, which serve as a means for the contractor to alert the agency about
how specification provisions will be met.

Implementation
The approach outlined in AASHTO PP 84 was developed for pavement concrete mixtures.
However, PEMs could be extended to include mixtures of other classes of concrete as well. For
example, use of PEM specifications could be used to ensure placement of low-permeability
concrete in bridge decks, girders, piers, and foundations, if desired.
Development and implementation of PEM specifications is an extensive undertaking, and
the shift will affect all stakeholders in the construction process. The menu of specification pro-
visions suggested by AASHTO PP 84 has provided guidance for a number of state agencies to
make initial movement toward PEMs in a variety of means (examples include shadow testing
using emerging test methods, pilot projects, and enhanced QC plans). At the time of publica-
tion of this report, a number of states, including Colorado, Indiana, Iowa, Michigan, Minnesota,
New York, North Carolina, South Dakota, and Wisconsin, were performing PEM tests.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Overview of Concrete Technologies   91  

Although some states are currently utilizing some performance-related or performance-


based specification provisions, implementation of PEMs is an ongoing effort to improve
specifications. Research is being performed at a number of universities and DOTs to
enhance the knowledge of the basic science, emerging tests, and field performance data to
support the specification. Evaluation of new technologies and equipment under field working
conditions is ongoing, with a goal of providing feedback to researchers and agencies, as well as
refining the technologies (221).
Challenges facing agencies interested in moving toward PEMs include ensuring that stake-
holders are aware of and capable of performing the new tests as well as properly interpreting the
results. Some tests require purchase of new testing equipment by both agencies and contrac-
tors. Ongoing technology transfer efforts are focused on preparing guidance for specification
approaches, tests, and QC (221). AASHTO PP 84 also requires an approved QC plan. This
provision, although likely to improve the quality of concrete produced, may be viewed as an
additional burden by some contractors.
Overall, implementation of PEMs should result in “higher quality concrete and more efficient
construction practices, reducing long-term costs to the agency” (210). AASHTO PP 84 is the
specification for PEMs. Additional guidance is presented in Cackler et al. (11, 210).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

CHAPTER 3

Survey of State Practices

Introduction
A survey was prepared and sent electronically using online survey software to all members
of the ASSHTO Committee on Materials and Pavements (COMP) representing the 50 state
highway agencies and the District of Columbia. The survey was a questionnaire composed of
24 questions about DOT practices related to implementation of concrete technologies in
transportation projects, possible shortages in quality aggregates and fly ash, use of recyclable
materials, and barriers to concrete technology implementation. The survey questions were
divided into eight main topic areas including:

1. General Information.
2. Concrete Technology Implementation.
3. Other Concrete Technologies (not covered in the synthesis).
4. Depletion of Quality Aggregates.
5. Availability of Fly Ash.
6. Use of Recycled and Reclaimed Materials.
7. Barriers to Technology Implementation.
8. Case Examples.

The questionnaire is shown in Appendix A. Forty state DOTs responded to the questionnaire.
These states included Alabama, Arizona, Arkansas, Colorado, Connecticut, Delaware, Florida,
Georgia, Idaho, Illinois, Kansas, Kentucky, Louisiana, Maine, Massachusetts, Michigan,
Minnesota, Mississippi, Missouri, Montana, Nebraska, New Hampshire, New Jersey, New York,
North Carolina, North Dakota, Ohio, Oregon, Pennsylvania, Rhode Island, South Carolina,
South Dakota, Tennessee, Texas, Utah, Vermont, Washington, West Virginia, Wisconsin, and
Wyoming. Their detailed responses are shown in Appendix B. This chapter summarizes the state
responses to the survey questions.

Responses to Survey Questionnaire


General Information
In this section the name and contact information of the state DOT respondent was provided.
The names of respondents are not included in Appendix B. In response to the questions whether
the respondent was a COMP member, 17 responded “Yes” and 23 responded “No.” Because
of the broad nature of the questions, which covered many technologies, both established and
new, the responses may have been composed by the COMP state member or designee with
assistance from other agency professionals.

92

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey of State Practices   93  

Concrete Technologies
This topic included seven questions concerning 14 concrete technologies, including those
that were the main focus of the synthesis, and others that were not but were still considered
important technologies. The surveyed technologies were SCC, UHSC, HESC, PC, RCC, UHPC,
ICC, TCMC, lightweight concrete (LC), lightweight cellular/foamed concrete (LCC), PCP,
VHESC or rapid-hardening concrete, latex-modified concrete (LMC), and polymer concrete
(Poly C). The questions were seeking information on implementation status, availability of
specifications, number of projects and average project age, application types (pavements and/or
structures), and implementation challenges.

Responses to Technology Use and Specifications Availability


The responses from the 40 states on technology use and availability of specifications for the
technologies are shown in Table 9. HSC seems to be used in all 40 responding states. This was
inferred based on the survey results of technology use including HSC, UHSC, SCC, ICC, UHPC,
and HESC.
The top three most implemented technologies according to the survey are SCC (38 states),
HESC (37 states), and LMC (31 states). The least implemented technologies are UHSC
(15 states), PC (14 states), and ICC (9 states).
With respect to the availability of specifications for the implemented technologies, the top
three technologies are SCC (34 states), HESC (31 states), and LMC (25 states). The lowest
three technologies are UHSC (9 states), PC (8 states), and ICC (7 states).

Technology Applications
The states reported a variety of applications for the 14 technologies in both pavements
and structures. High early strength technologies such as HESC, VHESC, LMC, and Poly C
are mainly used for rapid or emergency repairs of pavements and bridges and for overlays.
Other technologies such as SCC, UHSC, UHPC, and ICC are used in cast-in-place or in pre-
cast structural applications, while PCP, RCC, and PC are used in pavement construction
and overlays. The TCMC technology is used to control heat generation in mass elements of
bridges, LC is used as a lightweight filler in movable bridge decks and in overlays. LCC is used
for filling trenches, geotechnical voids, and gaps and as a backfill material for retaining walls.

Project Information
The number of states that have constructed projects using concrete technologies is shown
in Table 10. Each state respondent indicted, based on their estimate, the number of projects
where the specific technology was used. HESC and SCC seem to be well established in many
states. For example, 29 states indicated that they have constructed between 11 and more than
25 HESC projects, and 27 states indicated having between 11 and more than 25 SCC projects.
The UHPC seems to have gained more acceptance, as evident by the number of states (20)
that have implemented the technology (1 to 10 projects). A similar trend can be observed in
the use of LC (16 states), RCC (15 states), and PCP (14 states). These fairly young technologies
(except LC) are gaining more acceptance in transportation projects as a result of more appli-
cations, experience, training, and availability of construction guidelines.
The age of the projects is also an indication of the maturity in knowledge and use of technol-
ogy as well as demonstrated performance in various applications. Table 11 shows the range of
service lives (to present day) of the projects as estimated by the respondents. The projects with
technologies that have been in service from 6 to more than 10 years are SCC (in 29 states),

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

94   Concrete Technology for Transportation Applications

Table 9.   Technology use by the states and availability of specifications


and guidelines.

Availability of
Technology Users Specifications/Guidelines
No. of No. of
Technology States State DOT States State DOT
High-strength 40 AL, AZ, AR, CO, CT, DE, FL, GA, 40 AL, AZ, AR, CO, CT, DE, FL,
concrete (HSC) (Not ID, IL, KS, KY, LA, ME, MA, MI, GA, ID, IL, KS, KY, LA, ME,
surveyed but inferred MN, MS, MO, MT, NE, NH, NJ, NY, MA, MI, MN, MS, MO, MT,
from responses to NC, ND, OH, OR, PA, RI, SC, SD, NE, NH, NJ, NY, NC, ND, OH,
other technologies) TN, TX, UT, VT, WA, WV, WI, WY OR, PA, RI, SC, SD, TN, TX,
UT, VT, WA, WV, WI, WY
Ultrahigh-strength 15 CT, DE, GA, KY, ME, MI, MO, NE, 9 DE, GA, MI, MO, NE, NJ, NY,
concrete (UHSC) >10,000 NJ, NY, OR, PA, RI, TX, VT PA, RI
psi
Self-consolidating 38 AL, AZ, CO, CT, DE, FL, GA, ID, IL, 34 AL, AZ, CO, DE, FL, GA, ID,
concrete (SCC) KS, KY, LA, ME, MA, MI, MN, MS, IL, KS, KY, LA, ME, MA, MN,
MO, MT, NE, NH, NJ, NY, NC, OH, MS, MO, MT, NE, NH, NJ, NY,
OR, PA, RI, SC, SD, TN, TX, UT, NC, OH, PA, RI, SC, SD, TN,
VT, WA, WV, WI, WY TX, UT, VT, WA, WV, WI, WY
Internally cured 9 IL, KS, LA, MN, NY, NC, OH, UT, 7 IL, LA, MN, NY, OH, UT, WV
concrete (ICC) WV

Ultrahigh- 22 AL, CT, DE, FL, GA, ID, IL, ME, 19 AL, CT, DE, GA, ID, IL, ME,
performance concrete MA, MI, MT, NE, NJ, NY, OH, OR, MA, MI, MT, NE, NJ, NY, OH,
(UHPC) PA, RI, UT, VT, WI, WY OR, PA, RI, UT, VT, WI, WY

Temperature control of 30 AR, CO, CT, DE, FL, GA, IL, KS, 23 CO, CT, DE, FL, GA, IL, KY,
mass concrete (TCMC) KY, LA, ME, MA, MI, MN, MT, NJ, LA, MA, MI, MN, NJ, NY, ND,
NY, ND, OH, PA, RI, SC, SD, TN, OH, PA, RI, SC, TX, VT, WA,
TX, VT, WA, WV, WI, WY WV, WY
Precast concrete 21 AL, CO, CT, DE, FL, GA, IL, KS, 18 AL, CO, CT, DE, FL, GA, IL,
pavement (PCP) LA, MI, MN, MO, NJ, NY, NC, PA, LA, MI, MO, NJ, NY, PA, TX,
TX, UT, VT, WV, WI UT, VT, WV, WI
Roller-compacted concrete 17 AL, AR, CO, DE, GA, IL, KS, LA, 12 AL, AR, CO, GA, IL, LA, MO,
(RCC) MN, MO, NH, NC, PA, SC, TN, TX, PA, SC, TN, TX, WV
WV
Pervious concrete (PC) 14 DE, FL, IL, KY, ME, MN, NH, 8 DE, MN, NH, NY, PA, VT, WA,
NY, OR, PA, RI, VT, WA, WY WY
Recycled concrete 16 AL, AR, CO, CT, FL, IL, MI, MN, 12 AL, CO, CT, FL, IL, MI, MN,
aggregate (RCA) MO, NH, NY, OH, TX, WA, WV, NY, TX, WA, WV, WY
WY
High early strength 37 AL, AZ, AR, CO, CT, DE, FL, GA, 31 AL, AZ, AR, CO, CT, DE, FL,
concrete (HESC) ID, IL, KS, KY, LA, MA, MI, MN, GA, IL, KS, KY, LA, MA, MI,
MS, MO, NE, NH, NJ, NY, NC, OH, MO, NE, NH, NJ, NY, NC, OH,
OR, PA, RI, SC, SD, TN, TX, UT, OR, PA, SD, TN, TX, UT, VT,
VT, WA, WV, WI, WY WA, WV, WI, WY
Very high early strength 24 AL, CO, CT, DE, FL, IL, KY, LA, 16 CT, FL, IL, KY, ME, MA, MI,
concrete (VHESC) ME, MA, MI, MN, MO, NH, NY, MN, NH, NY, OH, RI, TX, UT,
NC, OH, RI, SD, TX, UT, WA, WI, WA, WY
( rapid-hardening
WY
concrete)
Lightweight concrete (LC) 29 CT, DE, FL, IL, KS, KY, ME, MA, 22 CT, DE, FL, KY, ME, MI, MN,
MI, MN, MO, NE, NH, NJ, NY, NC, MO, NE, NH, NJ, NY, NC, OH,
OH, OR, PA, RI, SC, SD, TN, TX, PA, RI, TN, TX, UT, VT, WV,
UT, VT, WA, WV, WI WI
Lightweight cellular/ 24 AZ, CO, CT, DE, FL, ID, IL, KY, 14 DE, FL, ID, IL, KY, ME, MI,
foamed concrete (LCC) ME, MA, MI, MN, NH, NY, NC, OH, MN, NH, NY, OH, SD, WA,
OR, RI, SD, UT, VT, WA, WI, WY WY
Latex-modified 31 AR, CO, CT, DE, FL, IL, KS, KY, 25 AR, CO, CT, DE, FL, IL, KY,
concrete (LMC) LA, ME, MA, MI, MN, MO, MT, NE, LA, ME, MI, MO, NJ, NY, NC,
NJ, NY, NC, OH, OR, PA, RI, SC, OH, OR, PA, RI, SC, TN, TX,
SD, TN, TX, UT, WA, WV, WY UT, WA, WV, WY
Polymer concrete (Poly C) 20 AL, AZ, CO, CT, DE, FL, IL, KS, 14 AL, AZ, CO, DE, FL, IL,
KY, MA, MN, NE, NJ, NY, NC, OR, KS, NJ, NY, OR, PA, TX,
PA, TX, WA, WI WA, WI

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey of State Practices   95  

Table 10.   Number of states that have constructed


projects using the technologies.

No. of States
1–5 6–10 11–25 >25
Technology Projects Projects Projects Projects
Ultrahigh-strength Concrete 10 1 4 2
10,000 psi (UHSC)
Self-consolidating concrete 9 2 7 20
(SCC)
Internally cured concrete 4 3 — 1
(ICC)
Ultrahigh-performance 13 7 — 2
concrete (UHPC)
Temperature control of mass 9 4 6 10
concrete (TCMC)
Precast concrete pavement 10 4 1 2
(PCP)
Roller-compacted concrete 11 4 — —
(RCC)
Pervious concrete (PC) 8 2 2 1

High early strength concrete 4 1 7 22


(HESC)
Very high early strength 3 2 1 9
concrete (VHESC) (rapid-
hardening concrete)
Lightweight concrete (LC) 12 4 4 7
Lightweight cellular/foamed 7 1 4 4
concrete (LCC)
Latex-modified concrete (LMC) 7 3 7 10
Polymer concrete (Poly C) 5 3 2 7

Note: Dash = no information provided in reference to indicate that the item was
not used or applied.

HESC (in 27 states), LC (in 27 states), TCMC (in 25 states), and LMC (in 23 states). Conversely,
PC, a technology developed in the early 1980s, has only six projects with ages between 6 to more
than 10 years. This may be due to the fact that cities and counties build more PC projects than
the state DOTs because PC is mostly used for parking pavements and much less in roadways.

Challenges with Technology Use


In response to a question related to challenges, problems, and distresses associated with the
use of each technology or material, the states identified several key challenges; some were unique
to a specific technology and others were common among several technologies. Experience of
agency personnel, inspectors, and contractors was the most common challenge reported by a
majority of the responding states. Other important challenges included higher cost of construc-
tion, availability of materials locally, and mixture design and properties. Important challenges
and problems with specific technologies are shown in Table 12.

Other Concrete Technologies


In this section, the survey questions pertained to other technologies not covered in the scope
of the synthesis. Examples of “other” technologies in the questionnaire included innovative

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

96   Concrete Technology for Transportation Applications

Table 11.   Number of states reporting project service lives.

No. of States
Technology <2 Years 2–5 Years 6–10 Years >10 Years
Ultrahigh-strength concrete 1 4 1 9
10,000 psi (UHSC)
Self-Consolidating Concrete 3 3 4 25
(SCC)
Internally cured concrete (ICC) 3 2 — 2

Ultrahigh-performance 1 9 1 9
concrete (UHPC)
Temperature control of mass 1 1 10 15
concrete (TCMC)
Precast concrete pavement 2 4 3 10
(PCP)
Roller-compacted concrete 2 2 6 5
(RCC)
Pervious concrete (PC) 3 3 3 3

High early strength concrete 1 3 9 18


(HESC)
Very high very early strength — 3 5 6
concrete (VHESC) (rapid-
hardening concrete)
Lightweight concrete (LC) — 6 4 23

Lightweight cellular/foamed 1 3 4 12
concrete (LCC)
Latex-modified concrete (LMC) — 1 4 19

Polymer concrete (Poly C) — 2 4 5

Note: Dash = no information provided in reference to indicate that the item was not used
or applied.

admixtures and nanomaterials to improve concrete properties and/or reduce shrinkage and
cracking, performance-based mix designs, fibers, innovative deicing technologies, high-friction
surface treatments, nontraditional SCMs, and nontraditional cementitious overlay and repair
materials.
Among the 40 responding DOTs, 14 states have worked on other technologies in experimen-
tal or implementation stage, and indicated the technology use in pavements and/or structural
applications. In addition, a number of those states have also developed specifications for the
implemented “other” technologies.
The following are some examples of other technologies in the experimental stage or that have
been implemented in demonstration or planned projects:
• Delaware, Illinois, West Virginia, and Wyoming are working with shrinkage-reducing
materials to mitigate shrinkage cracking on bridge decks.
• Florida, Maine, and Missouri are developing mixtures with steel and/or synthetic fibers for
structural applications.
• New York State has implemented the recently developed PEM guidance and has prepared a
specification for use in projects (see Case Example in Chapter 4).
• Florida, New York State, and Pennsylvania are working to optimize mixture ingredients
using SCMs.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey of State Practices   97  

Table 12.   Important challenges and problems with technologies.

Technology Important Challenges and Problems


Ultrahigh-strength concrete QA/QC issues and cracking
10,000 psi (UHSC)
Self-consolidating concrete Segregation, QA/QC issues, and appropriate aggregate size
(SCC)
Internally cured concrete Availability of lightweight aggregate and moisture conditioning
(ICC)
Ultrahigh-performance Availability of nonproprietary systems; surface preparation and
concrete (UHPC) joint forming
Temperature control of mass Insulation and other temperature control requirements and
concrete (TCMC) implementation; inspection and data interpretation
Precast concrete pavement Base-layer preparation; long-term performance and
(PCP) maintenance
Roller-compacted concrete Achieving proper density and surface roughness
(RCC)
Pervious concrete (PC) Mixture design issue; surface smoothness and freeze-thaw
durability
High early strength concrete Premature cracking; maintaining mixture workability;
(HESC) balancing durability needs with early strength
Very high early strength Premature setting; heat generation and shrinkage; production
concrete (VHESC) (rapid- handling and placement issues
hardening concrete)
Lightweight concrete (LC) Aggregate availability; storage and conditioning; and curing and
longevity
Lightweight cellular/foamed Availability of suppliers and equipment; mixture variability;
concrete (LCC) maintaining required density
Latex-modified concrete Bonding and delamination issue and surface smoothness
(LMC)
Polymer concrete (Poly C) Surface separation; poor ride quality

• Delaware and West Virginia have implemented, with success, the high surface friction
technology.
• Florida is developing concrete mixtures using recycled asphalt, while Missouri is using RCAs
in pavement mixtures.
• Delaware is evaluating some proprietary materials for use in UHPC mixtures.
• Kansas has implemented the air void analyzer to evaluate air content in concrete for freeze-
thaw resistance. In a similar effort, Vermont is experimenting with the SAM to determine if the
internal air structure is sufficient to protect the concrete from adverse effects of freeze-thaw.
• Louisiana has implemented the surface resistivity test to evaluate the long-term durability of
mixtures used in bridge decks.
• New York State is testing alternative deicing salts.

Most of the concerns and challenges that states reported when using the above technologies
were related to cost, industry acceptance, and lack of information on the potential long-term
performance of the technology.

Depletion of Quality Aggregates


In this section the states were asked whether they had a problem with availability of quality
aggregates presently or would have in the future. Florida, Kansas, and Maine reported currently
experiencing shortage of quality aggregates. Thirteen other states predicted that there could be

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

98   Concrete Technology for Transportation Applications

a shortage of quality aggregates in the future. The remaining 24 states did not have an issue with
quality aggregate availability now nor predicted a shortage in the future. Table 13 shows how
the various states responded.
Some of the solutions to address shortage of quality aggregates offered by the states included
conducting research to extend service life of their transportation facilities (12 states), import-
ing quality aggregates (9 states), modifying specifications and mixtures to allow the use of
noncomplying or reactive aggregates in nonstructural applications (9 states), and using RCAs
in concrete mixtures (7 states). Other suggestions included restricting the use of limestone and
other aggregates to high-quality mixtures and encouraging more mining of aggregates locally.

Shortage of Fly Ash


With the abundance of the less-expensive natural gas alternatives to power generating plants,
many coal-powered plants are gradually being phased out. This has resulted in a shortage of fly

Table 13.   Shortage of quality aggregate and fly ash.

Shortage of Quality
Aggregate Shortage of Fly Ash
Yes In the Yes In the
State DOT Now Future No Now Future No
Alabama X X
Arizona X X
Arkansas X X
Colorado X X
Connecticut X X
Delaware X X
Florida X X
Georgia X X
Idaho X X
Illinois X X
Kansas X X
Kentucky X X
Louisiana X X
Maine X X
Massachusetts X X
Michigan X X
Minnesota X X
Missouri X X
Mississippi X X
Montana X X
Nebraska X X
New Hampshire X X
New Jersey X X
New York X X
North Carolina X X
North Dakota X X
Ohio X X
Oregon X X
Pennsylvania X X
Rhode Island X X
South Carolina X X
South Dakota X X
Tennessee X X
Texas X X
Utah X X
Vermont X X
Washington X X
West Virginia X X
Wisconsin X X
Wyoming X X
Total 3 13 24 13 16 11

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey of State Practices   99  

ash that is used in concrete mixtures to enhance the long-term durability of structures. In this
section of the questionnaire, states were asked whether they are experiencing a shortage of fly
ash, presently or predict future shortages. Table 13 shows that 13 states are presently experienc-
ing a shortage of fly ash, another 16 states predicted future shortages, and the remaining 11 states
do not anticipate shortages.
Among the solutions suggested in the responses, 22 states agreed on the need to expand the
use of slag and increase its quantity in concrete mixtures. Fifteen states would expand the use
of other pozzolans such as metakaolin. Six states would import fly ash and apply, if needed,
a second-stage carbon removal process to bring the LOI content to within acceptable limits
before their use in concrete. Other notable responses included conducting research to use
alternative pozzolans such as rice husk ash, further processing of bottom ash, and allowing
contractors to switch fly ash supplies between approved sources in ongoing projects.

Use of Recyclable and Reclaimed Materials


Table 14 shows the responses to the questions related to the use of RCA and other landfilled
or reclaimed materials. Fifteen states use RCA in pavements and only five states use it in
structural applications as well. The majority of the responding states do not use RCA in pave-
ments (22 states). Among the states that use RCA in pavement mixtures, the RCA replacement
percentage used ranges from 10% to 50%. Responses on RCA proportions in structural concrete
applications were inconclusive. With respect to specifications, two states had stand-alone RCA
specifications, six states incorporate RCA in their aggregate specifications, and one state uses
project special provisions when incorporating RCA in concrete mixtures.
In addition, a number of states have conducted research or applied in experimental projects
other reclaimed materials such as granulated glass, tire rubber, bottom ash, construction debris,
biomass ash, solid waste ash, and plastic bottle fibers, as shown in Table 14. However, these
materials have had very limited, if any, use in concrete application, except ground tire rubber,
which is routinely used in asphalt mixes for roadway applications.

Barriers to Technology Implementation


The survey asked the state DOTs about their opinions on a list of perceived barriers to imple-
mentation of concrete technologies and any others the barriers they may be concerned with.
The state responses and number responding were as follows:
• Technology not sufficiently proven to be adopted (30).
• Too expensive to use (28).
• Not enough training to use the technology (27).
• No specifications or construction guidelines available (23).
• Industry resistance (21).
• Not sufficient time available to be devoted to new technologies (16).
• Bad experience with their construction or performance (13).
• Resistance to change or to explore new technologies (12).
• Weather conditions not permitting (5).
Other notable responses include the following:
• Lack of experience by agency and local industry.
• Concern about potential reduction in concrete mixture quality.
• Ability to assess long-term concrete durability with some technologies.
• Implementation challenges.
• Time and cost-effectiveness.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

100   Concrete Technology for Transportation Applications

Table 14.   Use of RCA and landfilled or reclaimed materials.

Use of Other Landfilled or


Use of Recycled Aggregate Reclaimed Materials
State DOT Pavements Structures None Materials None
Alabama X X X
Arizona X X
Arkansas X X
Colorado X X
Connecticut X X X
Delaware X X
Florida X Glass, tire rubber, solid
waste ash, bottom ash,
biomass ash
Georgia X Tire rubber, construction
debris, plastic bottle fibers
Idaho X X
Illinois X X X
Kansas X X
Kentucky X X
Louisiana X X
Maine X X
Massachusetts X X
Michigan X X
Minnesota X X
Missouri X X
Mississippi X X
Montana X X
Nebraska X X
New Hampshire X X
New Jersey X X
New York X X Glass, solid waste ash,
bottom ash
North Carolina X Tire rubber, bottom ash
North Dakota X X
Ohio X X
Oregon X Tire rubber
Pennsylvania X X
Rhode Island X Plastic bottle fibers
South Carolina X X
South Dakota X X
Tennessee X
Texas X X X
Utah X
Vermont X
Washington X X
West Virginia X X
Wisconsin X Glass, tire rubber, bottom
ash
Wyoming X X
Total 15 5 22 7 33

The responses point to important gaps in technology transfer to assist some states in imple-
mentation of new and innovative concrete technologies. Other barriers to implementation
include fragmented training efforts, negative perception of reliability, concerns about mixture
integrity and long-term performance, industry and agency reluctance, as well as cost and time
availability to pay attention to new technologies.

Case Examples
At the end of the survey the respondents were asked if they would agree to be interviewed
on certain aspects of their responses to formulate a case example from their state DOT.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey of State Practices   101  

An encouraging 24 respondents indicated willingness for additional interviews to prepare


case examples for their respective states. Five case examples were received from Florida, Illinois,
Missouri, New York, and Tennessee, and are presented in Chapter 4.

Lessons Learned from Survey Results


The survey was very helpful in collecting useful information from the DOTs on the state
of the practice with various technologies, considering that the DOTs are the main users of
these technologies. Results of the survey provided an understanding of the states’ practices,
including level of implementation, applications, performance, challenges, and barriers to wider
implementation.
Based on analysis of the survey responses, several issues emerged. For example, most state
DOTs are primarily interested in technologies to accelerate construction and/or repairs of
bridges and pavements, to achieve the required consolidated strength in heavily reinforced
forms, and to reduce cracking in mass concrete. Technologies such as SCC, HESC, VHESC,
LMC, TCMC, and ICC serve these purposes. Some DOTs are considering use of fairly new
technologies developed in recent years, such as UHPC, PCP, and ICC in demonstration or in
actual projects. Other states have used more traditional technologies such as RCC, PC, and LC
in new highway applications. However, with implementation of the technologies in transpor-
tation projects, the states have also experienced challenges related to inadequately qualified or
experienced project personnel, QA/QC issues, and some short- and long-term poor perfor-
mance on projects.
Fourteen states are experimenting with or have implemented technologies other than those
covered by this synthesis. Some of the technologies are proprietary but serve an important
purpose; others are considered traditional but are being used in new applications.
With respect to recycling, only 15 of the 40 states use RCAs in concrete mixtures. Also, only
seven states have experimented with or used, on a limited basis, other landfilled or reclaimed
materials.
The states also pointed out many barriers to implementation of traditional (but new to the
state) or new technologies. The majority of the respondents cited three main barriers: (1) tech-
nology not proven, (2) high cost, and (3) inadequate training. There were many states that also
cited other important barriers such as industry and agency reluctance, time availability to devote
to new technologies, and questions about long-term performance.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

CHAPTER 4

Case Examples of State Practices

Florida Department of Transportation:


Temperature Control of Mass Concrete
Criteria to Be Considered as Mass Concrete
ACI 301, “Specifications for Structural Concrete for Buildings,” states that a concrete place-
ment should be designated as mass concrete when the maximum temperatures and temperature
differences must be controlled due to factors including the content and type of cementitious
materials, environment surrounding placement, and minimum dimension of placement. To
put this statement in context, ACI 301 contends that, in general, a placement of structural con-
crete with a “minimum dimension equal to or greater than 4 feet should be considered mass
concrete.” ACI 301 also indicates that the minimum dimension could be reduced by concrete
placements that produce higher rates of heat generation at early ages, such as those that contain
Type III cement, accelerating admixtures, or have high contents of cementitious materials.
The Florida DOT Structures Design Guidelines (https://www.fdot.gov/structures/structures
manual/currentrelease/structuresmanual.shtm) require that, for all bridge components except
drilled shafts and segmental superstructure pier and expansion joint segments, “when the
minimum dimension of the concrete exceeds 3 feet and the ratio of volume of concrete to the
surface area is greater than 1 foot, provide for Mass Concrete.”

Mass Concrete Control Plan Content


Florida DOT Standard Specifications for Road and Bridge Construction (January 2018),
Section 346-3.3 states that
When mass concrete is designated in the Contract Documents, use a Specialty Engineer to develop
and administer a Mass Concrete Control Plan (MCCP). Develop the MCCP in accordance with
Section 207 of the ACI Manual of Concrete Practice to ensure concrete core temperatures for any
mass concrete element do not exceed the
• maximum allowable core temperature of 180°F and . . .
• maximum allowable temperature differential of 35°F [between the element core and surface].
Submit the MCCP to the Engineer for approval at least 14 days prior to the first anticipated mass
concrete placement. Ensure the MCCP includes and fully describes the following:
1. Concrete mix design proportions,
2. Casting procedures,
3. Insulating systems,
4. Type and placement of temperature measuring and recording devices,
5. Analysis of anticipated thermal developments for the various mass concrete elements for all anticipated
ambient temperature ranges,

102

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Case Examples of State Practices   103  

6. Names and qualifications of all designees who will inspect the installation of and record the output of
temperature measuring devices, and who will implement temperature control measures directed by
the Specialty Engineer,
7. Measures to prevent thermal shock, and
8. Active cooling measures (if used).
The Specialty Engineer or approved designee shall:
• Inspect and approve installation of temperature measuring devices
• Verify that the process for recording temperature readings is effective for first placement of each size
and type of mass component
• Be available for immediate consultation during monitoring period of any mass concrete element
• Record temperature measuring device readings at intervals no greater than six hours, beginning
at the completion of concrete placement and continuing until decreasing core temperatures and
temperature differentials are confirmed
• Leave temperature control mechanisms in place until the concrete core temperature is within 50°F of
the ambient temperature
• Within three days of the completion of temperature monitoring, submit a report to the Engineer which
includes all temperature readings, temperature differentials, data logger summary sheets and the maxi-
mum core temperature and temperature differentials for each mass concrete element.

Mass Concrete Mix Designs


Mass concrete is normally Class IV 5,500 psi, with a minimum total cementitious material
content of 658 lb/yd3, and a maximum water–cement ratio of 0.41. Mixtures typically contain
fly ash (18%–50% by weight), slag (50%–70%), or a combination of fly ash (10%–20%) and slag
(50%–60%) as partial replacements for portland cement.

Remediation Procedures When Temperature Limits Are Exceeded


If either the maximum allowable core temperature or temperature differential of any mass
concrete element is exceeded:
• Implement immediate corrective action as directed by Specialty Engineer.
• Approval of the MCCP shall be revoked, and no mass concrete elements can be placed until a
revised MCCP has been approved by the Engineer.
• Submit an Engineering Analysis Scope that addresses the structural integrity and durability
of any mass concrete element that is not cast in compliance with the approved MCCP or that
exceeds the allowable core temperature or temperature differential.
• Submit all analyses and test results requested by the Engineer for any noncompliant mass
concrete element to the satisfaction of the Engineer.

Exceptions to the Temperature Monitoring Requirements


of Mass Concrete
Reduced Monitoring.   When there are multiple mass elements that are similar, reduced
monitoring can be initiated by the following procedure:
• Submit a Reduced Monitoring request to the Engineer at least 14 days prior to the requested
date of reduced monitoring;
• If approved, the Specialty Engineer shall monitor the initial element of the group of similar
elements; and
• If the performance of the initial element meets all the requirements of the MCCP, then the
remainder of the similar elements may not be monitored if they meet all of the following
requirements:
– All elements have the same least cross-sectional dimension,
– All elements have the same concrete mix design,

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

104   Concrete Technology for Transportation Applications

– All elements have the same insulation R value and active cooling measures (if used), and
– Ambient temperatures during concrete placement for all elements is within −10°F or +5°F
of the ambient temperature during placement of the initial element.

No Monitoring.   Mass concrete control provisions are not required for drilled shafts sup-
porting sign, signal, lighting, or intelligent transportation (ITS) structures. At the contractor’s
option, instrumentation and temperature measuring may be omitted for any mass concrete
substructure element meeting all of the following requirements:
• Minimum cross-sectional dimension of 6 feet or less,
• Insulation R value of at least 2.5 provided for at least 72 hours following the completion of
concrete placement,
• Environmental classification of the concrete element is Slightly Aggressive or Moderately
Aggressive, and
• The concrete mix design meets the mass concrete proportioning requirements of
Section 346-2.3, and the total cementitious content of the concrete mix design is 750 lb/yd3
or less.

Problems with the Use of Reduced Monitoring and No Monitoring.   Section 346 states that
for Reduced Monitoring it is necessary to:
• Install temperature measuring devices for all mass concrete elements.
• Resume the recording of temperature monitoring device output for all elements if “directed
by the Engineer.”
However, in the field, the following have become common practices:
• For “Reduced Monitoring,” only the first element is being instrumented.
• For “No Monitoring,” MCCPs are not being submitted.

These procedures have been in place since January 2016 and have resulted in very few
structures being instrumented. Two of the most important means of reducing the potential
for thermal cracking involve limiting the core–surface temperature differential to 35°F until
the core temperature has cooled sufficiently, and keeping forms and thermal control measures
in place until the core–ambient temperature differential is ≤50°F. Limiting the core tempera-
ture and core–surface temperature differential to specified values cannot be done without
recording and monitoring temperatures.

Potential Future Changes to Mass Concrete Specifications

• Instrumentation of mass concrete elements that are not being monitored under the Reduced
Monitoring provision will need to be enforced.
• Research, which will likely lead to modifications of specifications, is needed to answer the
following questions:
– What are safe maximum temperatures for each combination of portland cement and
supplemental cementitious materials (SCMs)?
– Are particular portland cement-SCM combinations more susceptible to cracking?
– What should the maximum core-surface and core-ambient temperature differentials
(gradients) be?
– How should maximum concrete temperatures at placement be determined to mitigate
cracking?
– Should analyses using finite element modeling programs replace the use of physical and
compositional characteristics to indicate likelihood of mass concrete behavior and how to
mitigate?

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Case Examples of State Practices   105  

Illinois Department of Transportation:


Reducing Concrete Shrinkage in Bridge Decks
Introduction
The Illinois DOT has for at least the last 20 years been working to mitigate the early age
transverse cracking that is a typical problem on its bridge decks. A variety of high-performance
concrete mix designs were tried in the late 1990s. However, many of these mixes incorporated
silica fume, which seemed to exacerbate the cracking problem. Thus, in 2010, research work
started on mixture- or materials-based methods for reducing bridge deck cracking. Earlier
efforts of this research identified shrinkage-compensating mineral additives (Type K and Type G
additive products) and shrinkage-reducing chemical admixtures as potentially viable technolo-
gies. Research has proceeded now into a third, and final, phase wherein the scope has expanded
to include internal curing with prewetted lightweight fine aggregate, as well as considering the
potential benefits of using the identified mitigation technologies in combination (e.g., Type K
additive with internal curing). Additionally, a component of this ongoing research has been
trial implementation. As of the writing of this report, there have been six field trials using special
provisions drafted for each of the three mitigation strategies. Two bridge decks each have been
constructed using Type K mineral additive, SRA, or internal curing, and each has been successful
in reducing early age cracking in bridge decks.

Research on Additives, Admixtures, and Curing


Shrinkage-Compensating Additives
Shrinkage-compensating mineral additives were developed to compensate for the shrink-
age characteristics of ordinary portland cement by causing the concrete to expand and induce
compressive stress during hydration. This induced compressive stress essentially self-stresses
the concrete to counteract tensile stresses, such as those due to shrinkage. For Type K mineral
additives, the expansion is achieved by ettringite forming as a result of the reaction of gypsum
and calcium sulfoaluminate (CSA). On the other hand, the expansive properties of Type G
mineral additives are based on the formation of calcium hydroxide crystals resulting from
the reaction of calcium oxide and water. While both Type K and Type G additives have been
investigated as part of this research, the Illinois DOT has focused primarily on Type K, or CSA,
cement products, since the second phase.

Shrinkage-Reducing Admixtures
SRAs are liquid chemical admixtures added to concrete like other typical liquid admixtures.
SRAs can help mitigate shrinkage-related cracking by reducing the surface tension of the water
in concrete. This is because as concrete dries, the capillary stresses induced by its pore water
seeking escape are directly proportional to the pore water’s surface tension. Thus, as surface
tension decreases, capillary stresses decrease as well, leading to lower shrinkage strains within
the concrete. Furthermore, Sant et al. (222) concluded that SRAs may also enhance concrete’s
durability by reducing its sorptivity and moisture diffusivity, thereby reducing chloride and
other deleterious ion absorption and migration.

Internal Curing
In addition to mineral additives and chemical admixtures, the Illinois DOT began investigat-
ing internal curing to help address concrete shrinkage. Internal curing is achieved by substituting
a portion of the conventional fine aggregate volume of a concrete mix with prewetted light-
weight fine aggregate made from expanded shale, clay, slate, or slag. The prewetted lightweight
aggregate (LWA) is able to provide additional water to the concrete because its internal relative

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

106   Concrete Technology for Transportation Applications

humidity decreases during hydration or drying. Thus, by reducing moisture gradients within
the concrete, internal curing can alleviate capillary stresses and associated shrinkage strains.
This can be particularly beneficial to concretes placed with water–cement ratios less than 0.42,
such as Illinois DOT’s bridge decks, by hydrating more of the cementitious material without the
potential consequences of batching at higher water–cement ratios.

Conclusions from First Two Phases


Conclusions of the first two phases of the research included the following:
• Replacing 15% portland cement with Type K resulted in minimal shrinkage strain after
100 days.
• Including Class F fly ash increased the restrained expansion of Type K concrete specimens,
whereas silica fume resulted in a decrease in the extent of expansion.
• Slump loss was rapid in Type K concrete, and expansion was reduced with extended mixing
time; both factors need to be considered for long haul times.
• Including Type K reduced the total heat of hydration of paste specimens.
• Compressive strengths of Type K concretes at 28 days were found to be similar to or higher
than the compressive strengths of plain concrete. Including Class C or F fly ash resulted in
lower 28-day strengths, whereas silica fume increased 28-day strengths.
• High dosages of SRA lengthened the time to cracking in the ring test, but also seem to have
reduced compressive strength.
• Shrinkage rate after expansion of Type K concrete is comparable to that of conventional
concrete, whereas the shrinkage rate of SRA-dosed concrete is less than that of conventional
concrete.
• Even when the same amount of internal curing water is provided, the type and source of
LWA can have different effects on autogenous and drying shrinkage of mortar mixtures,
which is believed to be a result of the LWA’s desorption properties, gradation, and modulus.
• Increasing the amount of internal curing water, whether by increasing the replacement rate
of LWA or by increasing the initial moisture content of LWA, reduced autogenous shrinkage.
• In a sealed condition, the total shrinkage of mortar mixtures decreased as the replacement
with LWA increased.
• External curing is still necessary. When no curing method is applied, LWA was not beneficial
in reducing drying shrinkage of mortar specimens.
Practical lessons learned from trial implementation of these shrinkage mitigation technolo-
gies are as follows:
• Class C fly ash may reduce or be incompatible with Type K mineral additives. The period and
extent of expansion may be reduced due to the Class C fly ash reacting with the ye’elimite
component of the Type K additive. Early results suggest that adding a small amount of
gypsum to the system will counteract this problem.
• In a ready-mix operation using bagged Type K mineral additives, they appear best incor-
porated into the batching process as a slurry. Otherwise, if adding it via bags to a ready-mix
truck, there is the risk that the product will clump or ball. These balls may not be noticeable
during the placing and finishing of the concrete, but because they are less dense than the con-
crete, they can rise to the surface of the deck, resulting in pockets of unhydrated product that
will blister the surface when exposed to water, for example, during wet curing.
• Type K mineral additives require a large amount of water. It is important to ensure that
enough water is added to completely hydrate the Type K additive; otherwise, unhydrated
material will attempt to expand once exposed to water, resulting in essentially a delayed
ettringite formation distress.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Case Examples of State Practices   107  

• Working closely with the SRA manufacturer’s technical representative is advisable. For
example, if freeze-thaw deterioration is a concern, there are some SRAs that can make it
more difficult to achieve a stable, satisfactory air void system.
• Sufficient drain-down time helps prevent issues when batching prewetted lightweight fine
aggregate; otherwise, the material can be sticky and may “bridge” in bins or hoppers not
equipped with vibrators. For example, Illinois DOT’s initial specification required a mini-
mum 48-hour wetting period followed by 12 to 15 hours of drain down; the current specifica-
tion requires at least 72 hours of wetting and 20 to 24 hours of drain down.
• When discussing internal curing with DOT personnel and contractors, it is necessary to
emphasize that internal curing is not a substitute for conventional external curing practices.
Internal curing supplements external curing, which can only provide water to the near-surface
concrete. Internal curing helps provide water to the entire cross section of the concrete during
curing. For example, in laboratory tests, internal curing did not significantly change the drying
shrinkage characteristics of concrete without some external method to prevent moisture loss.

Developing Shrinkage Mitigation Strategies


The goal of the final phase of this research was to develop shrinkage mitigation strate-
gies and performance criteria for statewide implementation. Doing so will require providing
robust options that account for issues such as material availability, local experience and exper-
tise, material incompatibilities, and so on. Additionally, there is the potential optimization
of concrete mixes using a combination of strategies. For example, because Type K mineral
additives require so much water, it may be possible to provide water via prewetted LWA
instead of increasing the mix water, foregoing consequences typical of concretes with high
water–concrete ratios (e.g., lower strength, increased permeability). So far, results have been
promising, showing that internal curing increases the early age expansion of Type K mortar
specimens, partly due to a reduction in bulk modulus of elasticity (from the LWA portion),
but also due to increased Type K hydration. Meanwhile, adding SRA to internally cured mixes
reduced drying shrinkage considerably, made the mixture more volumetrically stable and, at
a water–concrete ratio of 0.34, reduced autogenous shrinkage.
For more information, please see the research reports from the first and second phases of this
research available for download from the Illinois Center for Transportation (223, 224).

Missouri Department of Transportation:


Precast Concrete Pavement Demonstration
Project on I-57—Lessons Learned
Introduction
In 2005 the Missouri DOT constructed a precast posttensioned concrete pavement (PPCP)
on I-57 in the southeast part of the state. The project was initiated by funding support from
the FHWA. The existing pavement was a badly distressed 8-in. (20-cm) jointed reinforced con-
crete pavement (JRCP) with 61.5 ft (18.75 m) joint spacing that was 46 years old at the time of
the PPCP construction. Extensive pumping and slab failures were prevalent. The local terrain
was flat and consisted of sandy-silt soils. Daily traffic was approximately 18,000 with 30% trucks.

Construction
The PPCP panels were to replace a 1,000-ft tangent section of the old JRCP. Both ends of the
PPCP would have new cast-in-place 15-ft jointed plain concrete pavement (JPCP) transitions.
The proposed dimensions of the PPCP panels were 38 ft (11.58 m) wide [including two 12-ft

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

108   Concrete Technology for Transportation Applications

(3.65-m) lanes, 4-ft (1.22-m) inside shoulder, and 10-ft (3-m) outside shoulder] and 10 ft (3 m)
long. Panels were lightly pretensioned across the width with No. 4 rebar to resist cracking during
transit from the precast plant.
To simplify the installation of the panels, a 4-in. (10-cm) permeable asphalt-treated base
(PATB) course was constructed as a subbase beneath the PPCP. The PATB was designed with
a flat profile, rather than the standard center crown profile, which would allow the bottom
of the panels to also be flat. The PATB was placed on a 4-in. (10-cm) dense graded aggregate
base. To maintain a minimum 8-in. (20-cm) structural thickness across both driving lanes and
provide a 2% surface cross slope for drainage, the panel thickness at the centerline crown had
to be 10.875 in. (27.63 cm) thick. The 2% cross slope also resulted in reducing the thickness
of the inside and outside shoulders to 7 in. (17.8 cm) and 5.625 in. (14.3 cm), respectively.
The panels were divided into four 250-ft (76.2-m) posttensioned sections, called “super
slabs,” which comprised 24 base panels. Two additional joint panels were installed at the two
ends of each super slab. Joint panels were designed as two 38-ft × 5-ft (11.6-m × 1.5-m) panels.
The panels were connected by dowel bars spaced every 12 in. (30 cm). The joint panels had post-
tensioning pockets on 2-ft centers through which 5/8-in. (1.6-cm) steel strands were threaded
and passed through the length of the super slab. Strands were posttensioned in the pockets after
all panels were in place.

Challenges
The perceived advantage to precast pavement technology is speed in construction. Posttension-
ing panels, however, added a layer of complexity that extended the construction period to
4 weeks. The contractor was not familiar with the technology to begin with. In addition, he also
encountered other unexpected difficulties.
The first challenge was that the joint panels were locked and could not slide open, even when
subjected to small hydraulic jacks. It is suspected that the dowel bars might have been mis-
aligned during fabrication. This led to a succession of other problems. Since the contractor had
to forego opening a specified gap in the joint panel to account for future anticipated expansion,
based on the ambient temperature range, they instead left small gaps between the base panels
in the super slab. The intent was that the joint panels would finally be forced to open during
posttensioning and the gaps between the base panels would then close.
The base panels could not close cleanly though, because the epoxy material that was brushed
on their vertical faces, which was supposed to provide a waterproof seal after posttensioning,
dried and created uneven texture before the base panels could be tensioned together. Another
related issue was that the panels were drifting off course, whether from the uneven gaps between
panels or some other reason. To compensate and bring the panels back in alignment, the con-
tractor inserted several small wooden wedges in the joints, thus creating more opportunity for
water to infiltrate after posttensioning.
During the actual posttensioning, the joint panels did indeed finally open; however, one of
the joints did not open properly and ruptured the concrete in tension failure a few inches away
from the joint. This led to a chronic maintenance problem to keep the poured joint header
material intact. The joint header and silicone filler in the other expansion joints required
periodic maintenance to a lesser degree.
Filling the grout ducts after posttensioning required an unexpected quantity of grout. The
first three ducts in one super slab used as much grout as was anticipated for all the strand ducts
in the four super slabs. It is likely that grout escaped through the slightly open base panel joints
and infiltrated the PATB.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Case Examples of State Practices   109  

Performance
Within several months after construction, hairline horizontal cracks appeared in some
panels. The thinking was that they were present since fabrication because shrinkage cracks
had not opened enough to become visible until later. It is highly doubtful that these cracks
could have formed after posttensioning.
The University of Missouri at Columbia had a research contract to instrument several panels
with strain gauges and monitor the sensors during fabrication and after installation and opening
to traffic. An important finding was that the tensile strains developed during the curing process
were two orders of magnitude higher than the barely perceptible 1–2 macrostrains measured
from truck loading. This was a strong indication that the posttensioned panels were virtually
assured of never experiencing critical load-induced stresses.

Conclusion
Overall, the PPCP is a success in terms of providing more than adequate structural support on
an Interstate highway with heavy truck traffic. However, the high cost of the time of installation
and posttensioning and the periodic maintenance required at expansion joints conspire to make
this an unlikely design option for future pavement replacement. Nonstressed modular precast
pavement panels seem to be a much more practical option.

New York State Department of Transportation:


Implementation of Performance-Engineered Mixtures
Introduction
PEM specifications are different from prescriptive specifications in that the latter com-
municate how a material or product is to be formulated and constructed, whereas perfor-
mance specifications communicate the desired characteristics of the material or product. PEM
specifications allow for greater partnership between the owner (the DOTs) and the contrac-
tor to allow greater opportunities for innovation and improved long-term performance. The
New York State DOT, like many other state DOTs, is moving to the use of PEM to take advan-
tage of newer technologies to progress projects in a way that provides greater performance,
often at reduced cost. The availability of resources has decreased over the years making it imper-
ative to develop and use more effective strategies for progressing projects. Using materials more
effectively with the ability to perform less direct inspection while still achieving and ensuring
quality is a key part of PEM implementation.

Background
The New York State DOT began using performance specifications many years ago. Initially
the performance requirements were focused on early strength gain or for specialty concrete
mixtures such as lightweight concrete or SCCs. Later, pavement mixtures were modified to
allow the use of well-graded aggregates to reduce cement content and to use lower water–
cement ratios. These changes achieved the strength and durability characteristics of the mix-
tures while providing more economical mixtures and maintaining workability.
Technological advances in materials testing capabilities have allowed further advancement of
performance specifications. The key is having the appropriate tools in place to provide QA for
concrete. Moving forward from the above, the use of both the surface resistivity (SR) and the
super air meter (SAM) represents those tools. By specifying certain resistivity requirements for
different applications (deck versus substructure versus pavement . . .) the quality of the concrete

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

110   Concrete Technology for Transportation Applications

can be better controlled. SR is directly affected by water content, pozzolan content, and, to a
lesser extent, by aggregate gradation. A quality mixture for a structural application will need to
have a typical water–cement ratio (0.40–0.42), pozzolan content of at least 20% (if not more for
ASR concerns), and well-graded aggregate (likely blended aggregate or No. 3 size). By specify-
ing resistivity requirements, along with the above-listed requirements, we can be assured of the
quality of the concrete for the application. Lesser SR requirements can be specified for other ele-
ments that have less risk and/or do not need to be as high a quality to perform for the expected
service life. The SAM, which is an enhanced Type B pressure meter, can measure not only total
air content, but can also assess the quality of the air void structure, similar to a linear traverse.
Having an adequately dispersed air void system provides the assurance toward long-term dura-
bility. The biggest benefit of SAM use is that information can be determined while the concrete
is still plastic, and so adjustments to subsequent batches can be made as needed.

Specification
The New York State DOT has evaluated a number of standard concrete mixtures using SR
and SAM, comparing the performance characteristics to newer PEMs. The PEMs typically
exceed the performance of existing department mixtures. Through the use of special specifi-
cations, PEMs are being used on a number of high-profile projects where characteristics are
desired to achieve longer service life and improved durability, or where special performance
is necessary. Because of the success with these special specifications, the New York State DOT
is now in the process of replacing standard specifications using prescriptive concrete mix-
tures with PEM specifications containing performance provisions. The proposed specification
changes include:
• Contractors/producers develop QC plan for mixture design and plant production. DOT
performs QA with random audits at batching facilities.
• Approved-list materials are still required to be used (why disregard what we know already
works), but unique materials may be considered for use on a case-by-case basis, based on past
successful experience demonstrated by the contractor.
• Contractors/producers would develop and provide PCC mixtures that meet the performance
criteria defined in specifications for a given use or application.
• Performance requirements include
– Compressive strength (4,000 psi minimum),
– Aggregate gradation to meet specific project applications,
– Air content,
– Aggregate friction requirement for pavements and bridge decks,
– ASR mitigation requirements,
– SAM use to determine freeze-thaw durability requirements for flatwork applications,
– Concrete permeability requirements determined by use of surface resistivity or RCP tests.
This requirement ensures that the water–cement ratio is kept in check and that pozzolans
are properly used as needed. The criteria vary depending on application, for example, decks
have the highest requirements.

Implementation
This initiative to implement PEMs for different types of concrete has been ongoing for a
number of years. The biggest challenge toward implementation is the migration toward and
acceptance of change. The prescriptive approach has been in place for many decades, and people
are used to how it works and what the normal expectations are. Getting people to understand
the PEM concept in order to become willing participants in trial projects has been difficult.
Therefore, extensive training is necessary.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Case Examples of State Practices   111  

With the performance approach, the contractor/producer now assumes the responsibility of
the production and performance of the concrete mixture. With this transfer of risk comes an
increased material cost to the owner. This increased cost, however, can be misunderstood at
times as it is generally balanced out by the reduced staffing and resource needs of the department
and frequently tied to improved and longer expected performance of the concrete. A significant
fiscal investment is also necessary in order to procure the needed new SR and SAM testing tools,
combined with becoming familiar with their use from trial work. The collection of data associ-
ated with locally available materials is also important in order to ensure the appropriate selection
of specification criteria.

Key to Success
The key to success is familiarizing people with the concept and developing a proper under-
standing of how it is intended to work. Identification of pilot project applications is important
to be able to learn from their advantages and shortcomings and apply this knowledge to further
specification or policy implementation.

Tennessee Department of Transportation:


Barriers and Solutions to Concrete
Technology Implementation
Introduction
The Tennessee DOT has experienced many barriers to implementing new concrete
technologies. Highway agencies, such as the Tennessee DOT, are tasked with being good
stewards of taxpayers’ money, and so cost is always an important issue but is not the most
critical factor in the decision to implement new technologies. Typically, new technologies and
materials are promoted to either improve efficiency with time savings or create better quality
products with a longer service life, and so the associated cost is just one of the factors that would
be a part of the decision-making process to implement the technology on state projects.

Examples of Barriers and Solutions


An important potential barrier to technology implementation is the lack of sufficient train-
ing to use the technology. Training was an important issue when SCC became increasingly
popular in transportation projects. At that time, there were not many training options for the
agency inspectors to be certified to test SCC, since ACI had not yet developed a training class or
included the SCC tests in the ACI Concrete Field-Testing Technician Grade I course (https://
www.concrete.org/certification/certificationprograms.aspx?m=details&pgm=Field%20Concrete
%20Testing&cert=Concrete%20Field%20Testing%20Technician%20-%20Grade%20I).
The Tennessee DOT’s approach to overcoming this barrier was to develop a training course
as part of the scope of work of an ongoing Tennessee DOT-funded research project. The course
was developed to help technicians understand potential uses of and best practices with SCC as
well as to focus on how to perform the ASTM/AASHTO test methods for SCC mixes. Initially,
the agency developed and provided pilot classes throughout the Tennessee DOT regions. Once
the training was completed, the course was refined and adopted as part of the department’s
Concrete Field-Testing Technician Course (https://www.tn.gov/tdot/materials-and-tests/
field-operations/training.html).
Resistance to change is another one of the most prevalent obstacles to implementing new
technologies. This resistance does not always come from industry, but also internally within the

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

112   Concrete Technology for Transportation Applications

DOT. One effective method that the Tennessee DOT has found to be useful in overcoming
this resistance is by having a good working relationship with the individual industry liaison
groups. The industry liaison groups are effective in communicating to their members the
information on potential program changes that the DOT is considering and getting feedback
from the affected industry. This feedback is valuable for the DOT. It allows full understanding
of how changes in the program might affect the industry, including any negative consequences
of the changes. This may lead to more communications and possible modifications. Also,
through this good working relationship, the industry has often proposed new technologies
or materials for Tennessee DOT consideration, which effectively eliminates one barrier of
resistance to change.

Concluding Remarks
Internally within the Tennessee DOT, a myriad of methods and tools are used to inform
the staff of the potential program changes to alleviate internal resistance to change. For
internal awareness and education, the agency utilizes webinars and presentations by industry
or product experts. The department considers actual demonstration of the proposed new
technology or material as the most effective educational and training tool for both agency
engineers and industry professionals to assess the true potential of the technology or material
and promote wide acceptance.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

CHAPTER 5

Conclusions and Technology


Information Gaps

Conclusions
From the results of the literature review and of the survey responses, a number of key
conclusions have been drawn:

1. The literature review showed that implementation of traditional, emerging, and new concrete
technologies has resulted in major benefits to the transportation infrastructure. The benefits
include accelerated construction, replacement and repair of pavements and bridges (UHPC,
HESC, VHESC, RCC, and PCP), better performance, and improved durability (HSC, SCC,
ICC, and PEM), control of temperature in massive structural members to mitigate thermal
cracking and improve durability (TCMC), and enhanced sustainability and environmental
benefits (RCA and PC).
2. The survey responses from 40 states showed that the top three most implemented concrete
technologies are HSC (40 states), SCC (38), HESC (37) and LMC (31), and the least imple-
mented technologies are UHSC ≥ 10,000 psi (69 MPa) (15), PC (14), and ICC (9).
3. The survey also showed the following results:
a. Fourteen states (Delaware, Florida, Georgia, Illinois, Kansas, Louisiana, Maine, Missouri,
New York, Ohio, Pennsylvania, Vermont, West Virginia, and Wyoming) have experimented
with or implemented technologies other than those discussed in this report.
b. Only three states (Florida, Maine, and Kansas) reported depletion in quality aggregates.
However, another 13 states (Idaho, Louisiana, Minnesota, Montana, New Jersey, New York,
North Dakota, Oregon, Pennsylvania, South Carolina, Texas, Utah, and Vermont) predicted
shortages in the future. The remaining 24 states reported no shortages.
c. Thirteen states (Alabama, Florida, Illinois, Maine, Massachusetts, Michigan, Missouri,
New Jersey, New York, North Dakota, Oregon, Rhode Island, and Texas) reported current
shortages in the availability of fly ash, 16 other states (Arizona, Arkansas, Colorado,
Connecticut, Delaware, Georgia, Idaho, Minnesota, Mississippi, Montana, Nebraska,
North Carolina, South Dakota, Tennessee, Vermont, and Washington) predicted future
shortages, and the remaining 12 did not report any shortages.
d. The top solutions offered to address the shortage of fly ash include expansion of the use
of slag, use of alternative pozzolans such as metakaolin, and import of foreign ash and use
after reprocessing for a lower LOI, or import ash from other states.
e. The most widely used recycled or reclaimed material in concrete applications is RCA.
Fifteen states (Alabama, Colorado, Connecticut, Florida, Illinois, Michigan, Minnesota,
Missouri, New York, Ohio, Texas, Washington, West Virginia, Wisconsin, and Wyoming)
use RCA in pavements, and 5 states (Alabama, Connecticut, Illinois, New York, and Texas)
use RCA in structural applications as well.

113  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

114   Concrete Technology for Transportation Applications

f. Seven states have experimented with the use of other reclaimed materials in concrete
mixtures. Five states (Florida, Georgia, North Carolina, Oregon, and Wisconsin) have
conducted research on the use of shredded or crumbed tire rubber. Four states (Florida,
New York, North Carolina, and Wisconsin) have reported research on the use of bottom
ash as an ingredient in concrete mixtures. Three states (Florida, New York, and Wisconsin)
have experimented with the use granulated glass in concrete. Two states (Florida and
New York) have experimented with municipal waste ash, and Georgia and Rhode Island
have used plastic bottle fibers. Florida has also conducted research on biomass ash.
g. The top five barriers to implementation of concrete technologies by the state include:
– Technology not sufficiently proven to be adopted (30 states),
– Too expensive to use (28 states),
– Lack of experience and not enough training (27 states),
– No specifications or construction guidelines available (23 states), and
– Industry resistance (21 states).
h. Other notable responses on the issue of barriers to technology implementation include the
following:
– Lack of experience by agency and local industry,
– Concern about potential reduction in concrete mixture quality,
– Ability to assess long-term concrete durability with some technologies,
– Implementation challenges, and
– Time constraints and cost-effectiveness.

Gaps in Concrete Technology Information


Also, gaps in the information pertaining to specific technologies were identified and are
provided for further attention. Among the information gaps are the following:
1. The level of training needed to successfully implement new technologies is not to the
expectations of many state DOTs.
2. The need for air entrainment in HSC and UHSC to resist freeze-thaw actions has not been
completely settled in the research community.
3. Unanswered questions remain about the use of SCC in pavement repairs and slab
replacements.
4. Some states indicated that sources of lightweight aggregates for ICC are not available at
convenient locations to make them cost-effective. Also, there seems to be uncertainty about
the expected ICC performance using lightweight aggregate from different sources.
5. Most mixtures used to produce UHPC are proprietary. This causes an increase in construc-
tion costs according to the survey responses.
6. With respect to TCMC, there does not seem to be consensus among the states, industry
groups, or published research on the limit of the maximum core temperature and tem-
perature differential between the core and the surface of the structure. Also, the question of
accuracy of numerical modeling is an area of concern.
7. There does not seem to be well-defined, acceptable procedures for maintenance, reservation,
and panel replacement in a posttensioned PCP.
8. Two main obstacles remain that limit expanding the use of RCC in highway pavements.
These are control of surface smoothness and absence of effective load transfer or dowel bars
to increase the load-carrying capacity at joints and prolong their performance.
9. The two major gaps in the PC technology are lack of a specialized machine to place
and uniformly compact the material without damaging its void structure, and avail-
ability of guidelines for design, construction, QC testing, and preventive maintenance of
PC pavements.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Conclusions and Technology Information Gaps   115  

10. The most common distress problem when using HESC in replacement panels and slabs is
premature cracking from thermal and nonuniform shrinkage stresses. Many states do not
seem to have effective measures to mitigate the problem or guidelines to assist in deciding
when to repair or to remove the damaged slabs.
11. The long-term performance of VHESC repair materials is not well understood, particularly
impact of type of application and weather conditions. Issues such as premature setting,
excessive shrinkage and cracking are also areas of concern.
12. The states have shown interest in the feasibility of using alternative pozzolans to
supplement the expected shortages in traditional fly ashes. However, they are concerned
about the impact of the alternative pozzolans on short- and long-term performance
of concrete.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

References

1. American Concrete Institute, Report on High-Strength Concrete, ACI 363R-10, Farmington Hills, MI,
March 2010.
2. American Concrete Institute, Self-Consolidating Concrete, ACI 237R-07, Farmington Hills, MI, April 2007.
3. Bentz, D., and J. Weiss, Internal Curing: A 2010 State-of-the-Art Review, Report NISTIR-7765, National
Institute of Standards and Technology, 2011.
4. Graybeal, B., TechNote: Design and Construction of Field-Cast UHPC Connections, FHWA-HRT-14-084,
U.S. Department of Transportation, October 2014.
5. American Concrete Institute, Guide to Mass Concrete, ACI 207.1R-05, Farmington Hills, MI, 2006.
6. Tayabji, S., D. Ye, and N. Buch, SHRP 2 Report S2-R05-RR-1: Precast Concrete Pavement Technology,
Transportation Research Board of the National Academies, Washington, DC, 2013.
7. American Concrete Institute, Guide to Roller-Compacted Concrete Pavements, ACI 327R-14, Farmington
Hills, MI, February 2015.
8. American Concrete Institute, Report on Pervious Concrete, ACI 522R-10, Farmington Hills, MI, March 2010.
9. Cackler, T., Recycled Concrete Aggregate Usage in the US, National Concrete Pavement Technology
Center, Iowa State University. Ames, 2018. https://intrans.iastate.edu/app/uploads/sites/7/2018/08/
RCA_US_usage_summary_w_cvr.pdf (accessed November 5, 2019).
10. American Concrete Institute, Accelerated Techniques for Concrete Paving, ACI 325.11R-01, Farmington
Hills, MI, 2001.
11. Cackler, T., D. Harrington, and P. C. Taylor, Performance Engineered Mixtures (PEM) for Concrete
Pavements, MAP Brief, Concrete Pavement Technology Center, Ames, IA, April 2017.
12. Russell, H. G., High-Performance Concrete—From Buildings to Bridges, Concrete International, Vol. 19,
No. 8, August 1997, pp. 62–63.
13. American Concrete Institute, Report on Chemical Admixtures for Concrete, ACI 212.3R-10, Farmington
Hills, MI, March 2010.
14. Mindess, S., and J. F. Young, Concrete, Prentice-Hall Inc., Englewood Cliffs, NJ, 1981.
15. Mokhtarzadeh, A., and C. French, Time-Dependent Properties of High-Strength Concrete with Consider-
ation for Precast Applications, ACI Structural Journal, Vol. 97, No. 3, 2000, pp. 263–271.
16. deLarrard, F., and A. Belloc, The Influence of Aggregate on the Compressive Strength of Normal and
High Strength Concrete, ACI Materials Journal, Vol. 94, No. 5, 1997, pp. 417–426.
17. Myers, J., and R. Carrasquillo, Influence of Hydration Temperature on the Durability and Mechanical
Property Performance of HPC Prestressed/Precast Beams, Transportation Research Record: Journal of the
Transportation Research Board, No. 1696, Vol. 1, TRB, National Research Council, Washington, DC, 2000,
pp. 131–142.
18. American Concrete Institute, Guide to External Curing of Concrete, ACI 308-16, Farmington Hills, MI, 2016.
19. Armaghani, J., T. Larsen, and D. Romano, Aspects of Concrete Strength and Durability, Transportation
Research Record 1335, TRB, National Research Council, Washington, DC, 1992, pp. 63–69.
20. Torres, S., and J. Eggers, Capping Systems for High-Strength Concrete, Transportation Research Record:
Journal of the Transportation Research Board, No. 1979, Transportation Research Board of the National
Academies, Washington, DC, 2006, pp. 46–53.
21. Carrasquillo, R., F. Slate, and A. Nilson, Microcracking and Behavior of High Strength Concrete Subject to
Short-Term Loading, ACI Journal Proceedings, Vol. 78, No. 3, 1981, pp. 179–186.
22. Hale, W., and B. Russell, The Need for Air Entrainment in High Performance Concrete, Symposium Pro-
ceedings: PCI/FHWA/FIB International Symposium on High Performance Concrete, L. S. Johal, Ed., Precast/
Prestressed Concrete Institute, Chicago, IL, September 2000.

116

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

References  117  

23. Cohen, M., Y. Zhou, and W. Dolch, Non-Air-Entrained High Strength Concrete—Is It Frost Resistant?
ACI Materials Journal, Vol. 89, No. 4, 1992, pp. 406–415.
24. Kosmatka, S. H., B. Kerkhoff, and W. C. Panarese, Design and Control of Concrete Mixtures, 14th ed.,
Portland Cement Association, Skokie, IL, 2001.
25. Neville, A., Properties of Concrete, Pitman Books, 779 pp., 2011.
26. Armaghani, J., M. Romano, M. Bergin, and J. Moxley, High Performance Concrete in Florida Bridges, in
SP 140-1: High Performance Concrete in Severe Environments, Paul Zia, Ed., American Concrete Institute,
1993, pp. 1–24.
27. Ozyildirim, C., Effects of Temperature on the Development of Low Permeability in Concretes, VTRC 98-R14,
Virginia Transportation Research Council, Charlottesville, 1998.
28. Section 346: Portland Cement Concrete, Standard Specifications for Road and Bridge Construction,
Florida Department of Transportation, 2017.
29. Yang, F., CE 241—Self-Consolidating Concrete, Report #1, University of California, Berkeley, March 2004.
30. Armaghani, J., K. Tawfiq, S. Squillacote, and M. Bergin, Accelerated Slab Replacement Using Self-Consolidated
Concrete, Transportation Research Record: Journal of the Transportation Research Board, No. 2508, Transporta-
tion Research Board of the National Academies, Washington, DC, 2015.
31. Hodgson, D., A. K. Schindler, D. A. Brown, and M. Stroup-Gardiner, Self-Consolidating Concrete for Use
in Drilled Shaft Applications, Journal of Materials in Civil Engineering, Vol. 17, No. 3, 2005, pp. 363–369.
32. Ozyildirim, C., and G. Moruza, Using Self-Consolidating Concrete for Bridge Repairs—SCC Mix-
tures Prove to Be Effective for Substructure Repairs, Concrete International, Vol. 37, No. 4, 2015,
pp. 42–46.
33. Wang, K., S. Shah, D. White, J. Gray, T. Voigt, L. Gang, J. Hu, C. Halverson, and B. Pekmezci, Self-Consolidating
Concrete—Application for Slip-Form Paving: Phase I (Feasibility Study), Report No. TPF-5(098), Center for
Portland Cement Concrete Paving Technology, Iowa State University, Ames, November 2005.
34. Wang, K., S. Shah, J. Grove, J. Gray, P. Taylor, P. Weigand, J. Hu, B. Steffes, G. Lomboy, Z. Quanji,
L. Gang, and N. Tregger, Self-Consolidating Concrete—Application for Slip-Form Paving: Phase II,
Report No. DTFH61-06-H-00011, Center for Portland Cement Concrete Paving Technology, Iowa
State University, Ames, May 2011.
35. Fang, W., C. Jianxiong, and Y. Changhui, Studies of Self-Compacting High-Performance Concrete with
High Volume Mineral Additives, in Proceedings of the First International RILEM Symposium on SCC,
Stockholm, Sweden, September 1999, pp. 569–578.
36. Ghafoori, N., R. Spitek, and M. Najimi, Transport Properties of Limestone-Containing Self-Consolidating
Concrete, ACI Materials Journal, Vol. 114, No. 4, 2017, pp. 527–536.
37. Ghezal, A., and K. Khayat, Optimizing Self-Consolidating Concrete with Limestone Filler by Using Statisti-
cal Factorial Design Methods, ACI Materials Journal, Vol. 99, No. 3, 2002, pp. 264–272.
38. Tawfiq, K., J. Armaghani, Accelerated Slab Replacement Using Temporary Precast Panels and Self-Consolidating
Concrete, Final Report BDV30TWO 977-02, Florida Department of Transportation, 2016.
39. Khayat, K., and D. Mitchell, NCHRP Report 628: Self-Consolidating Concrete for Precast, Prestressed Concrete
Bridge Elements, Transportation Research Board of the National Academies, Washington, DC, 2009.
40. Brown, D., and A. Schindler, High Performance Concrete and Drilled Shaft Construction, in ASCE GSP
158: Contemporary Issues in Deep Foundations, W. Camp, R. Castelli, D. F. Laefer, and S. Paikowsky, Eds.,
ASCE Press, 2005, pp. 1–12.
41. Bury, M., and E. Bühler, Methods and Techniques for Placing Self-Consolidating Concrete—An Overview
of Field Experiences in North American Applications, in Proceedings of the First North American Conference
on the Design and Use of SCC, ACBM, Chicago, IL, 2002, pp. 253–258.
42. Naito, C., G. Brunn, G. Parent, and T. Tate, Comparative Performance of High Early Strength and Self Consoli-
dating Concrete for Use in Precast Bridge Beam Construction: Final Report, ATLLS Report No. 05-03, National
Center for Engineering Research on Advanced Technology for Large Structural Systems, Lehigh University,
Bethlehem, PA, May 2005.
43. Chan, Y., and Y. Liu, Development of Bond Strength of Reinforcement Steel in Self-Consolidating Concrete,
ACI Structural Journal, Vol. 100, No. 4, 2003, pp. 490–498.
44. Sonebi, M., and P. Bartos, Performance of Reinforced Columns Cast with Self-Compacting Concrete,
SP-200: Fifth CANMET/ACI International Conference on Recent Advances in Concrete Technology, Proceed-
ings, American Concrete Institute, Farmington Hills, MI, 2001, pp. 415–432.
45. Khayat, K. H., Optimization and Performance of Air-Entrained Self-Consolidating Concrete, ACI Materials
Journal, Vol. 97, No. 5, 2000, pp. 526–535.
46. Kosmatka, S., and M. Wilson, Design and Control of Concrete Mixtures, 16th ed., Portland Cement Association,
Skokie, IL, 2016.
47. American Concrete Institute, Guide to External Curing of Concrete, ACI 308R-16, Farmington Hills,
MI, 2016.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

118   Concrete Technology for Transportation Applications

48. Bentur, A., S.-I. Igarashi, and K. Kovler, Prevention of Autogenous Shrinkage in High Strength Concrete by
Internal Curing Using Wet Lightweight Aggregates, Cement and Concrete Research, Vol. 31, No. 11, 2001,
pp. 1587–1591.
49. Geiker, M., D. Bentz, and O. Jensen, Mitigating Autogenous Shrinkage by Internal Curing, in SP-218: High-
Performance Structural Lightweight Concrete, Proceedings of American Concrete Institute Fall Convention,
J. P. Ries and T. A. Holm, Eds., 2004, pp. 143–154.
50. Bentz, D., Early Age Cracking Review: Causes, Measurements, and Mitigation Strategies, National Institute of
Standards and Technology. Gaithersburg, MD, 2009. https://www-pub.iaea.org/MTCD/Publications/PDF/
TE-1701_add-CD/PDF/USA%20Attachment%2004.pdf (accessed November 5, 2019).
51. Castro, J., I. De la Varga, M. Golias, and W. Weiss, Extending Internal Curing Concepts to Mixtures Con-
taining High Volumes of Fly Ash, in Proceedings of the International Concrete Bridge Conference, Phoenix,
AZ, 2010.
52. American Concrete Institute, Report on Internally Cured Concrete Using Prewetted Absorptive Lightweight
Aggregate, ACI (308-213)R-13, Farmington Hills, MI, 2013.
53. Powers, T., L. Copeland, and H. Mann, Capillary Continuity or Discontinuity in Cement Pastes, Portland
Cement Association R&D Lab Bulletin, Vol. 110, 1959.
54. Bentz, D. P., P. Lura, and J. Roberts, Mixture Proportioning for Internal Curing, Concrete International,
Vol. 27, No. 2, 2005, pp. 35–40.
55. Byard, B., and A. Schindler, Cracking Tendency of Lightweight Concrete, Research Report, Highway Research
Center, Auburn University, Auburn, AL, 2010.
56. Delatte, N., E. Mack, and J. Cleary, Evaluation of High Absorptive Materials to Improve Internal Curing of
Low Permeability Concrete, Ohio Department of Transportation Office of Research and Development and
Federal Highway Administration, Report No. FHWA/OH-2007/06, 2007.
57. Henkensiefken, R., D. Bentz, T. Nantung, and J. Weiss, Volume Change and Cracking in Internally Cured
Mixtures Made with Saturated Lightweight Aggregate Under Sealed and Unsealed Conditions, Cement and
Concrete Composites, Vol. 31, 2009, pp. 427–437.
58. Hammer, T., O. Bjontegaard, and E. J. Sellevold, Internal Curing—Role of Absorbed Water in Aggregate,
in SP-218: High-Performance Structural Lightweight Concrete, Proceedings of American Concrete Institute
Fall Convention, J. P. Ries and T. A. Holm, Eds., American Concrete Institute, 2004, pp. 131–142.
59. Holm, T., O. Ooi, and T. Bremner, Moisture Dynamics in Lightweight Aggregate and Concrete, in Theodore
Bremner Symposium on High-Performance Lightweight Concrete, J. P. Ries and T. A. Holm, Eds., American
Concrete Institute, 2003, pp. 167–184.
60. Castro, J., L. Keiser, M. Golias, and J. Weiss, Absorption and Desorption Properties of Fine Lightweight
Aggregate for Application to Internally Cured Concrete Mixtures, Cement & Concrete Composites, Vol. 33,
2011, pp. 1001–1008.
61. Weber, S., and H. Reinhardt, Modeling the Internal Curing of High Strength Concrete Using Lightweight
Aggregates, in Theodore Bremner Symposium on High-Performance Lightweight Concrete: Sixth CANMET/
ACI International Conference on Durability, Thessaloniki, Greece, 2003, pp. 45–64.
62. Villareal, V., Internal Curing, Real-World Ready-Mix Production and Applications—A Practical Approach
to Lightweight Modified Concrete, in SP-256: Internal Curing of High-Performance Concretes: Laboratory
and Field Experiences, D. Bentz and B. Mohr, Eds., American Concrete Institute, Farmington Hills, MI,
2008, pp. 45–56.
63. Henkensiefken, R., J. Castro, D. Bentz, T. Nantung, and J. Weiss, Water Absorption in Internally Cured
Mortar Made with Water-Filled Lightweight Aggregate, Cement and Concrete Research, Vol. 29, 2009,
pp. 883–892.
64. Indiana Department of Transportation, Office of Materials Management, Specific Gravity Factor and
Absorption of Lightweight Fine Aggregate, Test Method ITM 222-14T, 2014.
65. Ardeshirilajimi, A., D. Wu, P. Chaunsali, P. Mondal, Y. T. Chen, M. M. Rahman, A. Ibrahim, W. Lindquist,
and R. Hindi, Bridge Decks: Mitigation of Cracking and Increased Durability, FHWA-ICT-16-016, U.S.
Department of Transportation, 2016.
66. Miller, A., T. Barrett, R. Zander, and J. Weiss, Using a Centrifuge to Determine Moisture Properties of
Lightweight Fine Aggregate for Use in Internal Curing, Advances in Civil Engineering Materials, Vol. 3,
No. 1, pp. 142–157.
67. Bentz, D., and P. Stutzman, Internal Curing and Microstructure of High-Performance Mortars, in SP-256:
Internal Curing of High-Performance Concrete: Laboratory and Field Experiences, D. Bentz and B. Mohr, Eds.,
American Concrete Institute, Farmington Hills, MI, 2008, pp. 81–90.
68. Streeter, D., W. Wolfe, and R. Vaughn, Field Performance of Internally Cured Concrete Bridge Decks in
New York State, in SP-290: The Economics, Performance and Sustainability of Internally Cured Concrete,
A. K. Schindler, J. G. Grygar, and W. J. Weiss, Eds., American Concrete Institute, Farmington Hills, MI,
2012, pp. 1–16.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

References  119  

69. Barrett, T., A. Miller, and W. J. Weiss, Documentation of the INDOT Experience and Construction of the
Bridge Decks Containing Internal Curing in 2013, Joint Transportation Research Program Publication
No. FHWA/IN/JTRP-2015/10. Purdue University, West Lafayette, IN, 2015.
70. Bentz, D., Internal Curing of High-Performance Blended Cement Mortars, ACI Materials Journal, Vol. 104,
2007, pp. 408–414.
71. Jones, W., M. House, and J. Weiss, Internal Curing of High-Performance Concrete Using Lightweight
Aggregates and Other Techniques: Final Report, Colorado Department of Transportation, CDOT-2014-3,
February 2014.
72. Guthrie, W. S., and J. Yaede, Internal Curing of Concrete Bridge Decks in Utah: Preliminary Evaluation,
Transportation Research Record: Journal of the Transportation Research Board, No. 2342, Transportation
Research Board of the National Academies, Washington, DC, 2013, pp. 121–128.
73. Roberts, J., Internal Curing in Pavements, Bridge Decks, and Parking Structures Using Absorptive Aggre-
gates to Provide Water to Hydrate Cement not Hydrated by Mixing Water, presented at the Annual Meeting
of the Transportation Research Board, Washington, DC, 2005.
74. Roberts, J., The 2004 Practice and Potential of Internal Curing of Concrete Using Lightweight Sand, in
1st International RILEM Symposium on Advances in Concrete Through Science and Engineering, RILEM
Publications SARL, Bagneux, France, 2004, pp. 442–449.
75. Hoff, G., Internal Curing of Concrete Using Lightweight Aggregate, in Theodore Bremner Symposium
on High-Performance Lightweight Concrete: Sixth CANMET/ACI International Conference on Durability,
Thessaloniki, Greece, 2003, pp. 185–203.
76. Lopez, M., Creep and Shrinkage of High-Performance Lightweight Concrete: A Multi-Scale Investigation,
PhD dissertation, Georgia Institute of Technology, Atlanta, GA, 2005.
77. American Concrete Institute, Guide for Structural Lightweight Aggregate Concrete, ACI 213R-14,
Farmington Hills, MI.
78. Louisiana Transportation Research Center, Research Project Capsule 18-6C: Influence of Internal
Curing on Measured Resistivity, 2018. http://www.ltrc.lsu.edu/pdf/2018/capsules_18-6C.pdf (accessed
November 5, 2019).
79. Villareal, V., and D. Crocker, Better Pavements Through Internal Hydration: Taking Lightweight Aggregates
to the Streets, Concrete International, February 2007, pp. 32–36.
80. Rao, C., and M. I. Darter, Evaluation of Internally Cured Concrete for Paving Applications, 2013.
https://www.escsi.org/wp-content/uploads/2017/12/Eval-of-ICC-for-Paving-Apps-Report.pdf (accessed
November 5, 2019).
81. Cleary, J., and N. Delatte, Implementation of Internal Curing in Transportation Concrete, in Transportation
Research Record: Journal of the Transportation Research Board, No. 2070, 2008, pp. 1–7.
82. Bentz, D., S. Jones, and P. Peltz, Influence of Internal Curing on Properties and Performance of Cement-Based
Repair Materials, Report NISTIR 8076, National Institute of Standards and Technology, 2015.
83. Cavalline, T., B. Tempest, J. Leach, G. Loflin, M. Fitzner, and R. Newsome, Internally Cured Concrete Using
Prewetted Lightweight Aggregate: Final Report (Draft), Project FHWA/NC/RP 2016-06, North Carolina
Department of Transportation. Raleigh, 2018.
84. National Institute of Standards and Technology, Menu for Internal Curing with Lightweight Aggregates.
https://concrete.nist.gov/lwagg.html (accessed November 5, 2019).
85. Expanded Shale, Clay, and Slate Institute, Guide Specifications for Internally Cured Concrete, 2012.
https://www.escsi.org/wp-content/uploads/2017/10/4001.1-IC-Guide-Specification-.pdf (accessed
November 5, 2019).
86. New York State Department of Transportation, ITEM 557.5101-18: Internal Curing Specifications
for Superstructure Slabs, Approach Slabs, Sidewalks, and Safety Walks, 2018. https://www.dot.ny.gov/
spec-repository/557.5101—18.pdf (accessed November 5, 2019).
87. West Virginia Department of Transportation, Division of Highways, Special Provision: Section 601,
Structural Concrete, Internal Curing, 2013.
88. Russell, H. G., and B. A. Graybeal, Ultra-High-Performance Concrete: A State-of-the-Art Report for the Bridge
Community, FHWA-HRT-13-060, June 2013.
89. Weldon, B., D. Jáuregui, C. Newtson, C. Taylor, K. Montoya, S. Allena, J. Muro, M. Talahat, E. Lyell,
and E. Visage, Feasibility Analysis of Ultra-High-Performance Concrete for Prestressed Concrete Bridge
Applications—Phase I and II, Final Report NM09MCS-01, New Mexico Department of Transportation
Research Bureau, June 2012.
90. El-Tawil, S., Y.-S. Tai, and J. A. Belcher II, Field Application of Nonproprietary Ultra-High-Performance
Concrete, Concrete International, Vol. 40, No. 1, January 2018, pp. 36–42.
91. Chen, Y., F. Matalkah, R. Weerasiri, A. Balachandra, and P. Soroushian, Dispersion of Fibers in Ultra-High-
Performance Concrete, Concrete International, Vol. 39, No. 12, December 2017, pp. 45–50.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

120   Concrete Technology for Transportation Applications

92. Corvez, D., Mixture Proportioning and Formulating for Robustness of UHPC Production on Site, presen-
tation, Interactive Panel Proceedings, Panel 6, 1st International Symposium on Ultra-High-Performance
Concrete, Des Moines, IA, July 2016.
93. Federal Highway Administration, Every Day Counts: An Innovation Partnership with States: EDC-3 Final
Report, FHWA-17-CAI-005, McLean, VA, May 2017.
94. Standard Test Method for Flow of Hydraulic Cement Mortar, ASTM C1437, ASTM International,
Vol. 04.01, West Conshohocken, PA, 2007.
95. New York State Department of Transportation, 557.21.16—Field Cast Joints Between Precast Concrete
Units, July 2010. https://www.fhwa.dot.gov/hfl/resources/webinar/nysdot_uhpc_spec.cfm (accessed
November 5, 2019).
96. Panel-Audience Interaction Results of Poll Questions on UHPC, Interactive Panel Proceedings, Panel 2,
1st International Symposium on Ultra-High-Performance Concrete, Des Moines, IA, July 2016.
97. American Concrete Institute, Cement and Concrete Terminology, ACI 116R-00, Farmington Hills, MI.
98. American Concrete Institute, Cooling and Insulating Systems for Mass Concrete, ACI 207.4R-05 (Reapproved
2012), Farmington Hills, MI.
99. American Concrete Institute, Report on Thermal and Volume Change Effects on Cracking of Mass Concrete,
ACI 207.2R-07, Farmington Hills, MI.
100. Emborg, M., and S. Bernander, Assessment of Risk of Thermal Cracking in Hardening Concrete, Journal
of Structural Engineering, Vol. 120, No. 10, 1994, pp. 2893–2912.
101. Ruiz, J., A. Schindler, R. Rasmussen, P. Nelson, and G. Chang, Concrete Temperature Modeling and Strength
Prediction Using Maturity Concepts in the FHWA Hiperpav Software, in Proceedings of the 7th International
Conference on Concrete Pavements, Orlando, Florida, 2001, pp. 97–111.
102. Yikici, T., and H. Chen, Use of Maturity Method to Estimate Compressive Strength of Mass Concrete,
Construction and Building Materials, Vol. 95, 2015, pp. 802–812.
103. Roush, K., and J. O’Leary, Cooling Concrete with Embedded Pipes, Concrete International, Vol. 27, No. 5,
May 2005, pp. 30–32.
104. Suprenant, B., and W. Malisch, Contractors’ Guide to Mass Concrete—Pre- and Post-Bid Considerations for
Meeting Performance and Cost Demands, Concrete International, Vol. 30, No. 1, January 2008, pp. 37–40.
105. Tawfiq, K., and J. Armaghani, Accelerated Slab Replacement Using Temporary Precast Panels and Self-
Consolidating Concrete, Final Report to Florida Department of Transportation, BDV30TWO 977-02,
June 2016.
106. Tayabji, S., and W. Brink, Precast Concrete Pavement Implementation by U.S. Highway Agencies, Tech Brief,
FHWA-HIF-16-007, October 2015.
107. Fugro Consultants, Inc., Proposed Process for Design of Precast Concrete Pavements, Sacramento, California,
Final Report to Caltrans, Task Order No. 56a0315-003, December 2012.
108. Rao, C., P. Littleton, S. Sadasivam, and G. Ullman, California Demonstration Project: Pavement Replacement
Using a Precast Pavement System on I-15 in Ontario, Final Report, FHWA, Washington, DC, June 2013.
109. Littleton, P., and J. Mallela, Florida Demonstration Project: Precast Concrete Pavement System on US 92, Final
Report, FHWA Highways for LIFE, April 2014.
110. Gopalaratnam, V., B. Davis, C. Dailey, and G. Luckenbill, Performance Evaluation of Precast Pre-
stressed Concrete Pavement, Report No. OR08-008/RI03-007, Missouri Department of Transportation,
November 2007.
111. Merritt, D., F. McCullough, and N. Burns, Precast Prestressed Concrete Pavement Pilot Project Near
Georgetown, Texas, in Transportation Research Record: Journal of the Transportation Research Board,
No. 1823, Transportation Research Board of the National Academies, Washington, DC, 2003, pp. 11–17.
112. Rao, C., S. Sadasivam, P. Littleton, G. Ullman, and J. Mallela, Virginia I-66 Concrete Pavement Replacement
Using Precast Concrete Pavement Systems, Final Report, FHWA, Washington, DC, June 2013.
113. Gillen, S., Overview of Illinois Tollway Precast Panel System, Power Point Presentation, Illinois Tollway,
August 2016.
114. Tayabji, S., D. Ye, and N. Buch, Precast Concrete Pavements: Technology Overview and Technical
Considerations, PCI Journal, Vol. 58, No. 1, Winter 2013, pp. 112–128.
115. Merritt, D., B. McCullough, N. Burns, and A. Shindler, The Feasibility of Using Precast Concrete Panels to
Expedite Highway Pavement Construction, Report No. 9-1517-3, Prepared for Federal Highway Adminis-
tration; Report No. 1517-S, Project Summary Report, Prepared for University of Texas, 2000.
116. BCC Engineering Inc., Pavement Design Package—FPIDs: 432584-1-32-01 and 432584-2-32-01US 92/
SR 600/Hillsborough Ave. From E. of N. Central Ave. to W. of SR 583/N. 56th St and US 92/SR 600/
Hillsborough Ave. from N. 9th St. to N. 15th St, Prepared for the Florida Department of Transportation,
July 2016.
117. Portland Cement Association, RCC Pavement, Roller Compacted Concrete, PL397, Skokie, IL.
118. Bass, R., and G. Horninger, RCC Dam Construction—Examples, Project Details and Design Consider-
ations, Concrete International, Vol. 40, No. 2, February 2018, pp. 31–36.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

References  121  

119. Piggott, R., Roller Compacted Concrete Pavements—A Study of Long-Term Performance, RP366, Portland
Cement Association, 1999.
120. Roller-Compacted Concrete Pavements for Highways and Streets, IS 328, Portland Cement Association, 2005.
121. Yeonsoo, K., Roller-Compacted Concrete Construction on Interstate Highway in Georgia, Transportation
Research Record: Journal of the Transportation Research Board, No. 2040, Transportation Research Board of
the National Academies, Washington, DC, 2007, pp. 71-79.
122. Section 440—Roller Compacted Concrete Shoulder Pavement, Special Provision, Georgia Department of
Transportation, April 2004.
123. Reed, P., Next-Generation Roller-Compacted Concrete, Concrete Construction, December 15, 2017.
124. ACPA Guide Specification: Roller-Compacted Concrete Pavement as Exposed Wearing Surface, Version 1.2,
American Concrete Pavement Association, September 2014.
125. Damrongwiriyanupap, N., Y. Liang, and Y. Xi, Application of Roller Compacted Concrete in Colorado Road-
ways, Report No. CDOT-2012-11, Colorado Department of Transportation—Research, October 2012.
126. Harrington, D., F. Abdo, W. Adaska, and C. Hazaree, Guide for Roller-Compacted Concrete Pavements,
Portland Cement Association, Skokie, IL, 2010.
127. Hossain, S., and C. Ozyildirim, Use of Roller-Compacted Concrete Pavement in Stafford, Virginia, Report
No. FHWA/VCTIR 15-R19, Virginia Center for Transportation Innovation and Research, May 2015.
http://www.virginiadot.org/vtrc/main/online_reports/pdf/15-r19.pdf (accessed November 5, 2019).
128. Harrington, D., F. Abdo, W. Adaska, and C. Hazaree, Guide for Roller-Compacted Concrete Pavements, Iowa
State University, National Concrete Pavement Technology Center, 2010.
129. American Concrete Institute, Report on Roller-Compacted Concrete Pavements, ACI 325.10R-95(01),
Farmington Hills, MI.
130. Piggott, R., Roller Compacted Concrete Pavements—A Study of Long-Term Performance, Publication
No. RP366, Portland Cement Association, Skokie, IL, 1999.
131. American Concrete Pavement Association, The National RCC Explorer. http://rcc.acpa.org/webapps/
rccexplorer/index.html (accessed November 5, 2019).
132. American Concrete Pavement Association, Pavement Designer software, Version 1-18, 2018. http://
pavementdesigner.org (accessed November 5, 2019).
133. Delatte, N., D. Miller, and A. Mrkajic, Portland Cement Pervious Concrete Pavement—Field Performance
Investigation on Parking Lot and Roadway Pavements, Final Report, RMC Research & Education Founda-
tion, December 2007.
134. Florida Department of Transportation, Section 288: Cement Treated Permeable Base, in Standard
Specifications for Road and Bridge Construction, July 2017.
135. Florida Department of Transportation, Section 446: Edge Drains, in Standard Specifications for Road and
Bridge Construction, July 2017.
136. Anderson, I., and M. Dewoolkar, Laboratory Freezing-and-Thawing Durability of Fly Ash Pervious Con-
crete in a Simulated Field Environment, ACI Materials Journal, Vol. 112, No. 5, 2015, pp. 603–611.
137. Pent, G., T. Boming, Z. Hongzhou, F. Min, and Z. Yibo, The Pavement Performance of Steel Slag Pervious
Concrete, Proc., Third International Conference on Transportation Engineering, Chengdu, China, 2011,
pp. 1654–1659.
138. Amade, A., and S. Rogge, Development of High Quality Pervious Concrete Specifications for Maryland
Conditions, Report No. MD-13-SP009B4F, Maryland State Highway Administration, February 2013.
139. McCain, G., and M. Dewoolkar, Porous Concrete Pavements: Mechanical and Hydraulic Properties, in
Transportation Research Record: Journal of the Transportation Research Board, No. 2164. Transportation
Research Board of the National Academies, Washington, DC, 2010, pp. 66–75.
140. Kevern, J., D. Biddle, and Q. Cao, Effects of Macrosynthetic Fibers on Pervious Concrete Properties, Journal
of Materials in Civil Engineering, Vol. 27, No. 9, 2015.
141. Joung, Y., and Z. Grasley, Evaluation and Optimization of Durable Pervious Concrete for Use in Urban Areas,
Report No. SWUTC/08/167163-1, Texas Transportation Institute, Texas A&M University, February 2008.
142. Ozyildirim, H. C., M. S. Hossain, and M. Sharifiasl, Preliminary Investigation of Pervious Concrete for
a VDOT Parking Lot, Extended Abstract and Presentation, Transportation Research Board Annual
Meeting, 2018.
143. Moruza, G., C. Ozyildirim, and T. Culver, Development of a Special Provision on the Use of Pervious Concrete
as a Stormwater Management Tool in Parking Lots, VTRC 18-R15, Virginia Department of Transportation,
November 2017.
144. Vancura, M., K. MacDonald, and L. Kazanovich, Structural Analysis of Pervious Concrete Pavement, in
Transportation Research Record: Journal of the Transportation Research Board, No. 2226, Transportation
Research Board of the National Academies, Washington, DC, 2001, pp. 13–20.
145. Armaghani, J., Structural Evaluation of Pervious Concrete on Cellular Lightweight Permeable Concrete
(CLPC) Base, Final Report prepared for Cellular Concrete, LLC, April 2011.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

122   Concrete Technology for Transportation Applications

146. Delatte, N., A. Mrkajic, and D. Miller, Field and Laboratory Evaluation of Pervious Concrete Pavements,
in Transportation Research Record: Journal of the Transportation Research Board, No. 2113, Transportation
Research Board of the National Academies, Washington, DC, 2009, pp. 132–139.
147. Schaefer, V., K. Wang, M. Suleiman, and J. Kevern, Mix Design Development for Pervious Concrete in Cold
Weather Climates, Iowa State University, Ames, IA, 2006.
148. Haselbach, L., and R. Freeman, Vertical Porosity Distributions in Pervious Concrete Pavement, ACI Materials
Journal, Vol. 103, No. 6, 2008, pp. 452–458.
149. Chopra, M., M. Wanielista, C. Ballock, and C. Spence, Construction and Maintenance Assessment of Pervious
Concrete Pavements, Final Report, Florida Department of Transportation, January 2007.
150. Suleiman, M., J. Kevern, V. Schaefer, and K. Wang, Effect of Compaction Energy on Pervious Concrete
Properties, Iowa State University, Ames, IA, January 2006.
151. Huang, B., X. Shu, H. Wu, and Q. Dong, Laboratory Evaluation of Polymer-Modified Pervious Concrete,
Report No. FHWA-GA-10-0806, Office of Materials and Research, Georgia Department of Transportation,
August 2010.
152. Rehan, T., A. Werner, and Y. Qi, Pervious Concrete Test Methods for the Lab and Field, in Proceedings,
Construction Research Congress 2018: Construction Information Technology, C. Wang, C. Harper, Y. Lee, R.
Harris, and C. Berryman, Eds., ASCE, 2018, pp. 644–653.
153. Van Dam, T., J. Harvey, S. Muench, K. Smith, M. Snyder, I. Al-Qadi, H. Ozer, J. Meijer, P. Ram, J. Roesler,
and A. Kendall, Towards Sustainable Pavement Systems: A Reference Document, FHWA-HIF-15-002, U.S.
Department of Transportation, 2015.
154. Reza, F., and W. Wilde, Evaluation of Recycled Aggregates Test Section Performance, Final Report, MN/RC
2017-06, Minnesota Department of Transportation, 2017.
155. Snyder, M., T. Cavalline, G. Fick, P. Taylor, and J. Gross, Recycling Concrete Pavement Materials:
A Practitioner’s Reference Guide, National Concrete Pavement Technology Center, Iowa State University,
Ames, IA, 2018. https://intrans.iastate.edu/app/uploads/2018/09/RCA_practioner_guide_w_cvr.pdf
(accessed November 5, 2019).
156. Van Dam, T., K. Smith, C. Truschke, and S. Vitton, Using Recycled Concrete in MDOT’s Transportation
Infrastructure, RC-1544, Michigan Department of Transportation, 2011.
157. Federal Highway Administration, Transportation Applications of Recycled Concrete Aggregate: FHWA State
of the Practice National Review, U.S. Department of Transportation, 2004.
158. Chai, L., C. Monismith, and J. Harvey, Re-cementation of Crushed Material in Pavement Bases, UCPRC-
TM-2009-04, Caltrans, Sacramento, CA, 2009.
159. Cuttell, G., M. Snyder, J. Vandenbossche, and M. Wade, Performance of Rigid Pavements Containing
Recycled Concrete Aggregates, in Transportation Research Record 1574, TRB, National Research Council,
Washington, DC, 1997, pp. 89–98.
160. Gress, D., M. Snyder, and J. Sturtevant, Performance of Rigid Pavements Containing Recycled Concrete
Aggregate—2006 Update, presented at Annual Meeting of the Transportation Research Board, Washington,
DC, January 13–17, 2008.
161. Saeed, A., NCHRP Report 598: Performance-Related Tests of Recycled Aggregates for Use in Unbound
Pavement Layers, Transportation Research Board of the National Academies, Washington, DC, 2008.
162. American Concrete Pavement Association, Recycling Concrete Pavements, Engineering Bulletin EB043P,
2009.
163. Jones, C., Recycled Concrete Aggregate: A Sustainable Choice for Unbound Base, Construction Materials
Recycling Association, 2012.
164. Yrjanson, W., NCHRP Synthesis of Highway Practice 154: Recycling Portland Cement Concrete Pavements,
Transportation Research Board, National Research Council, Washington, DC, 1989.
165. American Concrete Institute, Removal and Reuse of Hardened Concrete, ACI 555R-01, Farmington Hills, MI.
166. Obla, K., H. Kim, and C. Lobo, Crushed Returned Concrete as Aggregates for New Concrete, Final Report,
Project 05-13, RMC Research & Education Foundation, 2007.
167. FHWA, Use of Recycled Concrete Pavement as Aggregate in Hydraulic-Cement Concrete Pavement,
Technical Advisory, 2007.
168. Ideker, J., M. Adams, J. Tanner, and A. Jones, Durability Assessment of Recycled Concrete Aggregates for Use in
New Concrete, Phase II Final Report, Research Report OTREC-RR-13-01, Oregon Transportation Research
and Education Consortium, Portland, OR, 2014.
169. Smith, J., and S. Tighe, Recycled Concrete Aggregate Coefficient of Thermal Expansion: Characteriza-
tion, Variability, and Impacts on Pavement Performance, in Transportation Research Record: Journal of
the Transportation Research Board. No. 2113, Transportation Research Board of the National Academies,
Washington, DC, 2009, pp. 53–61.
170. Naranjo, A., IH 10 100% RCA in CRCP Section, presented at the 11th International Conference on Con-
crete Pavements, San Antonio, TX, August 30, 2016.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

References  123  

171. Behera, M., S. K. Bhattacharyya, A. K. Minocha, R. Deoliya, and S. Maiti, Recycled Aggregate from C&D
Waste & Its Use in Concrete—A Breakthrough Towards Sustainability in The Construction Sector:
A Review, Construction and Building Materials, Vol. 68, 2014, pp. 501–516.
172. De Brito, J., and R. V. Silva, Current Status on the Use of Recycled Aggregates in Concrete: Where Do We
Go from Here? RILEM Technical Letters, Vol. 1, 2016, pp. 1–5.
173. Bollman, J., I-5 Willamette River Bridges, Longest Concrete Arch Spans in Oregon, Aspire: The Concrete
Bridge Magazine, Summer 2012, pp. 36–38.
174. Hu, J., M. S. Siddiqui, D. W. Fowler, and D. Whitney, Two-Lift Concrete Paving—Case Studies and Reviews
from Sustainability, Cost Effectiveness, and Construction Perspectives, TRB 93rd Annual Meeting Compen-
dium of Papers, Transportation Research Board, Washington, DC, January 2014.
175. Van Dam, T., K. Peterson, L. Sutter, A. Panguluri, J. Sytsma, N. Buch, R. Kowli, and P. Desaraju,
NCHRP Report 540: Guidelines for Early-Opening-to-Traffic Portland Cement Concrete for Rehabilitation,
Transportation Research Board of the National Academies, Washington, DC, 2005.
176. American Concrete Pavement Association, Fast Track Concrete Paving, Technical Bulletin 004.02, Skokie,
IL, 1994.
177. Armaghani, J., Precast Concrete for Highway Structures: Practices & Standards, presented at 360° in
Concrete—Solution Without Frontiers, Argos Forum, Bogota, Colombia, August 2016.
178. American Concrete Pavement Association, Concrete Pavement Technology—Guidelines for Full Depth
Repair, Technical Bulletin 002.02 P, Skokie, IL, 1995.
179. Smith, K., D. Harrington, P. Pierce, P. Ram, and K. Smith, Concrete Pavement Preservation Guide, 2nd ed.,
Report No. FHWA-HIF-14-014, U.S. Department of Transportation, Washington, DC, September 2014.
180. Lee, S., N. Nguyen, T. S. Le, and C. Lee, Optimization of Curing Regimes for Precast Prestressed Members
with Early-Strength Concrete, International Journal of Concrete Structures and Materials, Vol. 10, No. 3,
September 2016, pp. 257–269.
181. American Concrete Pavement Association, Guidelines for Bonded Concrete Overlays, TB007P, Skokie, IL.
182. Grove, J., Blanket Curing to Promote Early Strength Concrete, in Transportation Research Record 1234,
Transportation Research Board, National Research Council, Washington, DC, 1989, pp. 1–7.
183. Florida Department of Transportation, Section 353: Concrete Pavement Slab Replacement, in Standard
Specifications for Road and Bridge Construction, 2017.
184. Okamoto, P. A., C. L. Wu, S. M. Tarr, and L. W. Cole, Early Opening of PCC Pavements to Traffic, in Proc.,
Fifth International Conference on Concrete Pavement Design and Rehabilitation, 1993, pp. 141–152.
185. Whiting, D., M. Nagi, P. Okamoto, T. Yu, D. Peshkin, K. Smith, M. Darter, J. Clifton, and L. Kaetzel, SHRP-C-373:
Optimization of Highway Concrete Technology, National Research Council, Washington, DC, 1994.
186. American Concrete Pavement Association, Concrete Pavement Field Reference—Preservation and Repair,
Engineering Bulletin EB239P, Skokie, IL, 2006.
187. Xu, Q., J. Ruiz, G. Chang, S. Dick, S. Garber, and R. Rasmussen, Computer-Based Guidelines for Concrete
Pavements: HIPERPAV® III User Manual, FHWA-HRT-09-048. Federal Highway Administration, McLean,
VA, 2009.
188. Priddy, L., S. Jersey, and R. Freeman, Determining Rapid-Setting Material Suitability for Expedient
Pavement Repairs: Full-Scale Traffic Tests and Laboratory Testing Protocol, in Transportation Research
Record: Journal of the Transportation Research Board, No. 2113, Transportation Research Board of the
National Academies, Washington, DC, 2009, pp. 140–149.
189. American Concrete Institute, Guide to Materials Selection for Concrete Repair, ACI 546.3R-14, Farmington
Hills, MI, 2014.
190. Delatte, N., R. Miller, M. Asghar, A. Sommerville, A. Lesak, K. Amini, L. Susinskas, and J. Woods, Evaluation
of High-Performance Pavement and Bridge Deck Wearing Surface Materials, Report No. FHWA/OH-2016/15,
2016.
191. Choi, P., and K. Yun, Experimental Analysis of Latex-Solid Content Effect on Early-Age and Autogenous
Shrinkage of Very-Early Strength Latex-Modified Concrete, Construction and Building Materials, Vol. 65,
2014, pp. 396–404.
192. Sommerville, A., Selection of High-Performance Repair Materials for Pavements and Bridge Decks, MSCE
thesis, Cleveland State University, May 2014.
193. Priddy, L., H. Bell, L. Edwards, W. Caruth, and J. Rowland, Evaluation of the Structural Performance of
CTS Rapid Set Concrete Mix, U.S. Army Corps of Engineers Engineer Research and Development Center,
Report ERDC/GSL TR-16-20, 2016.
194. American Concrete Institute, Practitioner’s Guide for Alternative Cements, ACI ITG-10R-18. Farmington
Hills, MI, 2015.
195. Burris, L., K. Kurtis, and T. Morton, Novel Alternative Cementitious Materials for Development of the
Next Generation of Sustainable Transportation Infrastructure, Tech Brief, FHWA-HRT-16-017, 2015.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

124   Concrete Technology for Transportation Applications

196. Juenger, M., F. Winnefeld, J. Provis, and J. Ideker, Advances in Alternative Cementitious Binders, Cement
and Concrete Research, Vol. 41, 2011, pp. 1232–1243.
197. American Concrete Institute, Report on Polymer-Modified Concrete, ACI 548.3R-09, Farmington Hills, MI,
2015.
198. Wilson, T., K. Smith, and A. Romine, Materials and Procedures for Rapid Repair of Partial-Depth Spalls in
Concrete Pavement, Manual of Practice, FHWA-RD-99-152, Federal Highway Administration, McLean, VA,
1999.
199. Hoerner, T., K. Smith, H. T. Yu, D. Peshkin, and M. Wade, PCC Pavement Evaluation and Rehabilitation,
Reference Manual for NHI Course No. 131062, National Highway Institute, Arlington, VA, 2001.
200. American Concrete Pavement Association, Concrete Pavement Field Reference: Preservation and Repair,
Report EB239P, Skokie, IL, 2006.
201. Priddy, L., Development of Laboratory Testing Criteria for Evaluating Cementitious, Rapid Setting Repair
Materials, ERCD/GLS TR-11-13, U.S. Army Corps of Engineers, Engineer and Research Development
Center, Vicksburg, MS, 2011.
202. Priddy, L., and T. Rushing, Development of Laboratory Testing Protocol for Rapid-Setting Cementitious
Material for Airfield Pavement Repairs, in Transportation Research Record: Journal of the Transportation
Research Board, No. 2290, Transportation Research Board of the National Academies, Washington, DC,
2012, pp. 89–98.
203. Lesak, A., Installation and Field Testing of High-Performance Repair Materials for Pavements and Bridge
Decks, MSCE thesis, Cleveland State University, December 2014.
204. Susinskas, L., Field Observation of Installation and Performance of Repair Materials, MSCE thesis, Cleveland
State University, August 2016.
205. Department of Defense, Tri-Service Pavements Working Group, Testing Protocol for Polymeric Spall
Repair Materials, Tri-Service Pavements Working Group (TSPWG) Manual, TSPWG M 3-270-01.08-4,
February 28, 2017.
206. Amini, K., Laboratory Testing of High-Performance Repair Materials for Pavements and Bridge Decks, MSCE
thesis, Cleveland State University, May 2015.
207. Woods, J., Specification Recommendations for Use of High Performance Repair Material, MSCE thesis,
Cleveland State University, August 2016.
208. Ahlstrom, G., Update: Performance Engineered Concrete Mixtures and Quality Assurance Program,
presented at National Concrete Consortium meeting, September 2015. https://intrans.iastate.edu/app/
uploads/2018/07/Ahlstrom-Update-Performance-Engineered-Concrete-and-QA-for-NCC-9-2015.pdf
(accessed November 5, 2019).
209. Van Dam, T., Performance Engineered Mixtures—The Key to Predictable Long-Life Pavement Perfor-
mance. Concrete Pavement Association of Minnesota. March 9, 2017. http://www.concreteisbetter.com/
wp-content/uploads/2017/04/1.PerformanceEngineeredMixes_VanDam.pdf (accessed November 5, 2019).
210. Cackler, T., M. Praul, and R. Duval, Developing a Quality Assurance Program for Implementing Perfor-
mance Engineered Mixtures for Concrete Pavements, MAP Brief, Concrete Pavement Technology Center,
Ames, IA, July 2017.
211. Cavalline, T., M. Ley, W. Weiss, T. Van Dam, and L. Sutter, A Road Map for Research and Implementation
of Freeze-Thaw Resistant Highway Concrete, presented at 11th International Conference on Concrete
Pavements, San Antonio, TX, August 28–31, 2016.
212. AASHTO, Standard PP 84-18: Standard Practice for Developing Performance Engineered Concrete
Pavement Mixtures, West Conshohocken, PA, 2018.
213. Thomas, M., B. Fournier, and K. Folliard, Selecting Measures to Prevent Deleterious Alkali-Silica Reaction in
Concrete: Rationale for the AASHTO PP65 Prescriptive Approach, FHWA-HIF-13-002, 2012.
214. Transportation Research Circular E-C171: Durability of Concrete, 2nd ed., Transportation Research Board
of the National Academies, Washington, DC, 2013.
215. Weiss, J., Relating Transport Properties to Performance in Concrete Pavements, Concrete Pavement Technology
Center, Ames, IA, 2014.
216. Monical, J., C. Villani, Y. Farnam, E. Unal, and J. Weiss, Using Low Temperature Differential Scanning
Calorimetry to Quantify Calcium Oxychloride Formation for Cementitious Materials in the Presence of
Calcium Chloride, Advances in Civil Engineering Materials, Vol. 5, No. 2, 2016, pp. 142–156.
217. Weiss, J., and Y. Farnam, Concrete Pavement Joint Deterioration: Recent Findings to Reduce the Potential
for Damage, MAP Brief, CP Road Map, Concrete Pavement Technology Center, Ames, IA, 2015.
218. Rupnow, T., and P. Icenogle, Evaluation of Surface Resistivity Measurements as an Alternative to the Rapid
Chloride Permeability Test for Quality Assurance and Acceptance, Final Report, FHWA/LA.11/479, Louisiana
Transportation Research Center, Baton Rouge, 2011.
219. Cook, M., M. Ley, and A. Ghaeezada, A Workability Test for Slipformed Concrete Pavements, Construction
and Building Materials, Vol. 68, 2014, pp. 376–383.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

References  125  

220. Taylor, P., X. Wang, and X. Wang, Concrete Pavement Mixture Design and Analysis (MDA): Development
and Evaluation of the Vibrating Kelly Ball Test (V Kelly Test) for the Workability of Concrete, Report No. TPF
5(205), FHWA and Pooled Fund Partners, U.S. Department of Transportation, Washington, DC, 2015.
221. Praul, M., Quality Assurance and the Move to Performance Concrete Specifications, presented at the
11th International Conference on Concrete Pavements, San Antonio, TX, August 28–31, 2016.
222. Sant, G., A. Eberhardt, D. Bentz, and J. Weiss, Influence of Shrinkage-Reducing Admixtures on Moisture
Absorption in Cementitious Materials at Early Ages, Journal of Materials in Civil Engineering, Vol. 22,
No. 3, 2010, pp. 277–286.
223. Chaunsali, P., S. Li, P. Mondal, D. Foutch, D. Richardson, Y. Tung, and R. Hindi, Phase I Report: Bridge
Decks: Mitigation of Cracking and Increased Durability, FHWA-ICT-13-023, 2013. https://apps.ict.illinois.
edu/projects/getfile.asp?id=3099 (accessed November 5, 2019).
224. Ardeshirilajimi, A., D. Wu, P. Chaunsali, P. Mondal, Y. Tung Chen, M. Mahfuzur Rahman, A. Ibrahim,
W. Lindquist, and R. Hindi, Phase II Report: Bridge Decks: Mitigation of Cracking and Increased Dura-
bility, FHWA-ICT-16-016, 2016. https://apps.ict.illinois.edu/projects/getfile.asp?id=4980 (accessed
November 5, 2019).

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

APPENDIX A

Survey Questionnaire

The survey below was mailed electronically to 50 state departments of transportation using

SurveyGizmo software. The purpose of the survey was to obtain information on the use of

concrete technologies for transportation applications and practices by the states. Responses were

received from 40 states.

NCHRP SYNTHESIS 49-09

CONCRETE TECHNOLOGY FOR TRANSPORTATION APPLICATIONS

SURVEY QUESTIONNAIRE

KEY DEFINITIONS (for the purpose of this survey)

• COMP – AASHTO Committee on Materials and Pavements

• SCC – Self-consolidating concrete – SCC is highly flowable, non-segregating

concrete that can spread into place, fill the formwork, and encapsulate the

reinforcement without any mechanical consolidation.

• UHSC – Ultra-high strength concrete – UHSC is concrete with strength equal to or

higher than 10,000 psi.

126

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey Questionnaire   127  

• HESC – High-early strength concrete – HESC is used in accelerated construction,

repair and rehabilitation. It normally has short setting and hardening times.

• PC – Pervious concrete – PC is concrete with little or no fine aggregate used in

parking areas, base layers, some streets and in edge drains.

• RCC – Roller compacted concrete – RCC is a very dry concrete mix placed with

special or asphalt paves and is compacted with steel rollers.

• UHPC – Ultra-high-performance concrete – UHPC is an advanced cementitious

material composed of high cement and silica fume contents, very low w/c, high

dosage of steel fibers and optimized gradation of fine aggregate. UHPC has been

used in connections between prefabricated bridge elements.

• ICC – Internally cured concrete – ICC is concrete with saturated lightweight

aggregates or polymer particles as replacements of conventional aggregate to

provide internal curing, and supplements externally applied curing.

• TCMC – Temperature control of mass concrete – TCMC is a process to mitigate

the impact of temperature in mass concrete members.

• LC – Lightweight concrete – LC is concrete using lightweight aggregates as partial

or full replacement of normal-weight aggregate.

• LCC – Lightweight cellular/foamed concrete – In LCC, foaming agent is used to

expand the volume of cement slurry or mortar and is used as a filler material.

• PCP – Precast concrete pavement – PCP is composed of prefabricated concrete

panels that are assembled on-site and used for new construction, overlay, and/or

repairs.

• PRHC – Proprietary rapid hardening concrete repair materials – These are

patching materials used for rapid repair of pavements and bridges.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

128   Concrete Technology for Transportation Applications

• LMC – Latex modified concrete – LMC is polymer modified portland cement–based

material used for patching and overlays.

• PolyC – Polymer concrete – PolyC is polymer resin and aggregate without

portland cement. It is used in patching and overlays and can be placed and then

opened to traffic swiftly. It provides high wear resistance.

• RCA – Recycled concrete aggregate.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey Questionnaire   129  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

130   Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey Questionnaire   131  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

132   Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey Questionnaire   133  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

134   Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey Questionnaire   135  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

136   Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey Questionnaire   137  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

138   Concrete Technology for Transportation Applications

11. Has your agency evaluated or implemented other new or promising concrete
technologies/materials.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey Questionnaire   139  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

140   Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey Questionnaire   141  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

142   Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Survey Questionnaire   143  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

144   Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

APPENDIX B

Responses to Survey Questionnaire

The following 40 state departments of transportation responded to the survey questionnaire:


1. Alabama (AL)
2. Arizona (AZ)
3. Arkansas (AR)
4. Colorado (CO)
5. Connecticut (CT)
6. Delaware (DE)
7. Florida (FL)
8. Georgia (GA)
9. Idaho (ID)
10. Illinois (IL)
11. Kansas (KS)
12. Kentucky (KY)
13. Louisiana (LA)
14. Maine (ME)
15. Massachusetts (MA)
16. Michigan (MI)
17. Minnesota (MN)
18. Mississippi (MS)
19. Missouri (MO)
20. Montana (MT)
21. Nebraska (NE)
22. New Hampshire (NH)
23. New Jersey (NJ)
24. New York (NY)
25. North Carolina (NC)
26. North Dakota (ND)
27. Ohio (OH)
28. Oregon (OR)
29. Pennsylvania (PA)

145  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

146   Concrete Technology for Transportation Applications

30. Rhode Island (RI)


31. South Carolina (SC)
32. South Dakota (SD)
33. Tennessee (TN)
34. Texas (TX)
35. Utah (UT)
36. Vermont (VT)
37. Washington (WA)
38. West Virginia (WV)
39. Wisconsin (WI)
40. Wyoming (WY)

Questions and Responses

1. Are you the agency representative serving on AASHTO Committee on Materials and Pavements
COMP)?

Yes 17

No 23

2. Has your agency used the following new concrete technologies/materials in transportation
applications?

3. Does your agency have specification or guidelines for following technologies/materials?

No. of states
Technology Availability of
Users
specification

Self-Consolidating Concrete (SCC) 38 34

Ultra-High strength Concrete > 10,000 psi


15 9
(UHSC)
High Early Strength Concrete
37 31
(HESC)

Pervious Concrete (PC) 14 8

Roller Compacted Concrete (RCC) 17 12

Ultra-High-Performance Concrete
22 19
(UHPC)

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Responses to Survey Questionnaire   147  

Internally Cured Concrete


9 7
(ICC)
Temperature Control of Mass Concrete
30 23
(TCMC)

Lightweight Concrete (LC) 29 22

Lightweight Cellular/Foamed Concrete


24 14
(LCC)

Precast Concrete Pavement (PCP) 21 18

Very High Early Strength Concrete


24 16
(VHESC) {Rapid-Hardening Concrete}

Latex Modified Concrete (LMC) 31 25

Polymer Concrete (Poly C) 20 14

4. Please identify the most the common application(s) in pavement construction, overlay, or repair.

Technology Application 1 Applications 2 Application 3 Application 4

Bridge approach Precast Pavement


SCC
slabs panels
UHSC — — — —
Intersection
HESC Pavement repair Slab replacement Overlay
construction
PC Parking areas Sidewalks Alleyways

RCC New construction Shoulder paving Industrial parking Local roads

UHPC — — — —
Pavement
ICC
construction
TCMC — — — —

LC Pavement overlay

LCC Embankment fill Culvert repair


Intersection
PCP New construction Overlay Pavement repairs
construction

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

148   Concrete Technology for Transportation Applications

VHESC Partial depth Maintenance


Spall repairs
(PRHC) patches repairs
LMC Overlays

Poly C — — — —

5. Please identify application(s) in construction, overlay or repair of bridges, tunnels and other
structure.

Technology Application 1 Applications 2 Application 3 Application 4

Precast bridge
Retaining and
SCC members & pier Precast inlets Drilled shafts
noise walls
caps
Bridge Design build Voids between
UHSC Piers
construction projects box beams
HESC Bridge repairs Emergency repairs Closure pours Deck overlay

PC — — — —

RCC — — — —
Joint filler in
UHPC Shear key bridge overlays ABC Construction
prefab. bridges
ICC Bridge decks Approach slabs
Bridge pier, pier Members thicker
Mass footings and
TCMC caps and Drilled shafts than 4ft, another
spread foundations
abutments state > 6 ft
Filling grates in
LC Bridge decks movable bridge
decks
Fill lining of
Filler behind Drilled caissons in Fill old culverts
LCC pipes and
abutments slope failures and tunnels
culverts
PCP — — — —
Emergency
VHESC
Partial deck repair Precast patching bridge joint Deck overlay
(PRHC)
repairs
Deck and other Substructure
LMC Deck overlays
member repairs spall repairs
Expansion joint
Poly C Thin deck overlay Patching spalls
repair

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Responses to Survey Questionnaire   149  

6. Please indicate, in your estimate, the number of projects where the technology/material has
been used.

1–5 6–10 11–25 More than 25


Technology
projects projects projects projects
SCC 9* 2 7 20

UHSC 10 1 4 2

HESC 4 1 7 22

PC 8 2 2 1

RCC 11 4 — —

UHPC 13 7 — 2

ICC 4 3 — 1

TCMC 9 4 6 10
LC 12 4 4 7

LCC 7 1 4 4
PCP 10 4 1 —
VHESC
3 2 1 9
(PRHC)
LMC 7 3 7 10

Poly C 5 3 2 7

* No. of states.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

150   Concrete Technology for Transportation Applications

7. Please indicate the range of service lives (years) of the projects where the
technology/material has been used. (Service life refers to age from construction to present
time).

Less than More than 10


Technology 2–5 yrs. 6–10 yrs.
2 yrs. yrs.
SCC 3* 3 4 25

UHSC 1 4 1 9

HESC 1 3 9 18

PC 3 3 3 3

RCC 2 2 6 5

UHPC 1 9 1 9

ICC 3 2 — 2

TCMC 1 1 10 15

LC — 6 4 23

LCC 1 3 4 12

PCP 2 4 3 10
VHESC
— 3 5 6
(PRHC)
LMC — 3 5 6

Poly C 0 2 4 5

* No. of states.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Responses to Survey Questionnaire   151  

8. What, in your opinion, are the most common problems, distresses or challenges, when using
the technology/material?

Technology Challenge 1 Challenge 2 Challenge 3 Challenge 4

Familiarity of agency & Segregation & Proper form Appropriate


SCC
construction personnel other Q/C issues tightness aggregate sizes
UHSC Experience Q/C issues Cracking High cost

Agency familiarity Balancing F/T Maintain slump Early age


HESC durability needs for consolidation cracking
with early strength and finishing.
Experience of local Mix design issues Surface Freeze/Thaw
PC
contractors smoothness durability
Experience and Compaction & Surface roughness Long term
RCC availability of special achieving proper durability
paving machine density
Experience and Surface High cost of Availability of
UHPC qualifications preparation and proprietary non-proprietary
Joint forming systems systems
Moisture conditioning Mix design issues
ICC
of lightweight aggregate
Experience of Requirement for Contractors’ Inspection and
contractors & proper insulation and complaints data
TCMC
pre-construction other temperature interpretation
planning. control measures.
Accurate mixture Storage and Aggregate Curing and
LC proportioning conditioning of availability locally Longevity
aggregate
Availability of suppliers Mix variability Maintaining Experience and
LCC
and equipment and poor Q/C required density Cost
Preparation of base Ensuring panel Long term Cost
PCP layer surface level with maintenance and
surrounding performance
High cost Heat generation, Production Setting time
VHESC
shrinkage and handling and
(PRHC)
cracking placement
Experience and material Mix design issues Bonding Surface
LMC
cost delamination smoothness
Surface preparation Contractor Poor ride quality Cost
Poly C
experience effectiveness

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

152   Concrete Technology for Transportation Applications

Other Technologies

9. Has your agency evaluated or implemented other new or promising concrete technologies/materials?

Yes 14

No 26

10. Please fill out the information on any other new or promising concrete technology/material
your agency has evaluated or implemented in construction projects.

Technology/ Used For Experimental Number Uses/Benefits


Material pavements, or of
(state) structures or both? Implemented? Projects
Shrinkage-
reducing Less shrinkage in
Both Implemented 6-10
admixtures concrete
(DE)
High Friction
Surface Improved skid
Pavements Implemented 6-10
treatment resistance
(DE)
Fly ash
Compensate for fly
alternatives Both Experimental —
ash shortage
(FL)
Fibers in
Crack reduction and
concrete Structures Implemented More than 10
steel replacement
(FL)
Optimizing
aggregate
gradation and Reduce heat and
Both Experimental —
cement cracking
content
(FL)
Recycled
asphalt
pavement Reduce the use of
Pavements Experimental —
(RAP) in virgin aggregate
concrete
(FL)
Prevent C
(GA) Structures Implemented 2 Control cracking

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Responses to Survey Questionnaire   153  

EdenCrete Increases strength


(GA) Pavements Implemented More than 5
and Durability
Shrinkage
reducing To mitigate deck
Structures Experimental 2
admixtures cracking
(IL)
Shrinkage
Compensating
To mitigate deck
cementitious Structures Experimental 2
cracking
materials
(IL)
Air Void Structures Implemented 50 Produces better air
Analyzer spacing factors
(KS)
Surface Structures Implemented All projects Increase service life
resistivity
(LA)
“Lafarge Structures Implemented 3 To produce 20,000
Ductil” psi deck panel joint
(ME) fill
Steel fibers Structures Implemented 3 To produce non-
(ME) shrink deck panel
joint filler
CTS Structures Implemented More than 20 Non-shrink deck
Komponent Panel/keyway grouts
(ME)
Prevent C Structures Experimental 1 Bridge crack control
(ME)
Steel & Structures Experimental 17 Deck overlay and
synthetic repairs. Reduce
fibers cracking
(MO)
Recycled Pavements Experimental 2 Sustainability. Utilize
concrete closer aggregate
aggregate sources
(MO)
Two-lift Pavements Experimental 1 Excellent wearing
pavement surface on layer with
(MO) soft aggregates or
RCC
Performance Both Implemented More than 20 Lower cement
based pavements & content &
mixtures 5 structures permeability. High
(NY) strength & durability

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

154   Concrete Technology for Transportation Applications

Alternative Both Experimental Multiple Cost effective de-


deicers projects icing
(NY)
Alternative Structures Experimental 5 Reduce cement
SCMs/additiv content. (No real
es performance
(NY) improvements
noticed)
Macro Fibers Pavements Experimental 6-10 Mitigate cracking
(OH)
Ultra - thin Pavements Experimental 1 Heavy truck in rural
Overlay areas
(OH)
Mixture Pavement Implemented 5 —
design
optimization
(PA)
Super Air Structures Experimental More than 20 Determines if
Meter internal air structure
(VT) is sufficient for
freeze/thaw
resistance
High Friction Pavements Implemented 10 High skid resistance
surface
treatment
(WV)
Shrinkage Structures Implemented 8 Reduce deck
reducing cracking
(WV)
Shrinkage Structures Implemented 3 Reduce cracking in
reducing UHPC mixes.
admixtures
(WV)
Shrinkage Structures Implemented 3 Crack reduction in
resistance UHPC bridge deck
admixture overlays.
(SRA)
(WY)

11. Please indicate whether or not your agency has developed specifications and/or construction
guidelines.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Responses to Survey Questionnaire   155  

12. What in your opinion are the top two concerns, challenges or problems your agency has
with these technologies/materials?

Technology / Material Specification Challenges


(state) (Yes, No)
Shrinkage reducing Yes High cost
admixtures
(DE)
High Friction Surface Yes High cost
treatment
(DE)
Fly ash alternatives Yes Availability of quality alternative, price, and
(FL) availability of large volume of raw material.
Fibers in concrete Yes Distribution and orientation of fibers in
(FL) pavement. Buy American limitations.
Optimizing aggregate No Potential industry resistance
gradation and cement
content.
(FL)
Recycled asphalt pavement No Concrete plants not willing to use another
(RAP) in concrete. aggregate source not from materials mined
(FL) and supplied by their parent company.
Prevent C No
(GA)
EdenCrete No Cost is extremely high
(GA)
Shrinkage reducing Yes Negative impact on mixture air content
admixtures
(IL)
Shrinkage Compensating Yes May result in limiting expansion of fresh
cementitious materials concrete when used in mixtures with Class C
(IL) fly ash. Also, if concrete not mixed
sufficiently, blisters may form on the deck
surface.
Air Void Analyzer Yes High cost of test equipment
(KS)
Surface Resistivity Yes
(LA)

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

156   Concrete Technology for Transportation Applications

“Lafarge Ductil” Yes High Cost


(ME)
Steel fibers No Safety concerns handling the fibers.
(ME)
CTS Komponent No
(ME)
Prevent C No
(ME)
Steel & synthetic fibers Yes Criteria on the use of fibers, determining
(MO) fiber dosage
Recycled concrete aggregate Yes Economical consideration when deciding to
(MO) use RCA vs. availability of virgin aggregate
Two-lift pavement Yes Cost
(MO)
Performance engineered Yes Industry acceptance to new testing protocols
mixtures (PEM) for PEM, and capability of some plants to
(NY) handle additional materials.
Alternative deicers Yes Damage to deck surface and variability of
(NY) materials.
Alternative SCMs/additives No
(NY)
Macro Fibers Yes Even distribution in the concrete mixture
(OH)
Ultra - thin Overlay Yes Cost
(OH)
Mixture design optimization Yes Getting all involved up to speed with the
(PA) technology and producing materials meeting
specifications
Super air meter Yes More expensive device
(VT)
High Friction surface Yes Good results
treatment
(WV)
Shrinkage reducing Yes Good results
admixture
(WV)
Shrinkage reducing Yes
admixture (SRA)
(WY)

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Responses to Survey Questionnaire   157  

Depletion of Quality Aggregates

13. Does your state have a problem of depletion of quality aggregates?

Yes, presently 3

Yes, in the future 13

No 24

14. Please select from the following, the possible agency actions to address this challenge.

Agency action No. of states


Conduct research to optimize concrete mixes, improve design and
12
construction quality to extend service life of pavements and structures.
Import quality aggregates. 9
Modify specifications to relax some stringent aggregate requirements for
9
use in non-structural concrete.
Use more pozzolanic materials in concrete mixes to prevent the adverse
9
effects of reactive aggregates being used in non-structural application.
Expand the use of recycled concrete aggregates in paving mixes. 7

Other responses:
• Support continued mining of local aggregates.
• Be more restrictive of limestone aggregate use in concrete.
• Use quality aggregates only when needed.

Availability of Fly Ash

15. The conversion to gas as fuel in power plants has or will reduce the availability of fly ash. Is
the shortage of fly ash an issue in your state?

Yes, presently 13

Yes, in the future 15

No 12

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

158   Concrete Technology for Transportation Applications

16. How does or will your agency address the problem? Please select all that apply from the
following:

Agency Action No. of states


Expand the use of slag and increase its proportion in the mix. 22
Allow and expand the use of alternative pozzolans such as metakaolin. 15
Allow the use of imported high LOI fly ash after second stage burning to
6
lower LOI.
Use volcanic ash. 1
Conduct research on feasibility of using wood burning ash from biomass
1
plants and paper mills.

Other responses:
• Allow easy switch between approved fly ash sources.
• Conduct research on feasibility of using rice hall ash.
• Import ash from overseas.
• Conduct research on a variety of ash alternatives.
• Do not allow the use of reactive aggregates to reduce the need for fly ash.
• Smart use of flay ash by avoiding the use of large projects that can cause supply problems.
• Use Slag instead of flay ash in ASR mitigation.
• Allow flexibility in switching fly ash sources in active projects.
• Further process bottom ash for reuse.
• Need industry innovation for alternatives.
• States should encourage research and innovations.
• Use straight cement in mixtures.

Use of Recycled, Landfilled/Reclaimed Materials

17. Does your agency allow the use of recycled concrete aggregate (RCA) in concrete mixes for
pavements?

Yes 15

No 25

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Responses to Survey Questionnaire   159  

18. Does your agency allow the use of recycled concrete aggregate (RCA) in concrete mixes for
structural applications?

Yes 7

No 33

19. Do you have RCA specification?

Availability of RCA Specification No. of states


Yes, stand-alone specification 2
Yes, incorporated in aggregate specification 6

No, special provisions are used for RCA 1

No 19

20. Select the most common replacement proportion of RCA for coarse aggregate in the concrete
mix.

Responses:

0 5%– 10% 10%– 20% 20%– 30% 30%– 40% 40%– 50%
Pavements 1 1 1 1 1

Structures — — — — 1

21. Has your agency allowed the use of the following land-filled/reclaimed materials in concrete
mixes?

22. Is the concrete containing the land-filled/reclaimed material used in the applications below?

Land filled/reclaimed Used in Applications


material Research Projects Nonstructural Pavements Structures
*
Granulated glass 3 — 2 — —

Shredded/crumbed tire rubber 2 2 — 4 —

Fibers from plastic bottles 1 1 1 — —

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

160   Concrete Technology for Transportation Applications

Ash from municipal solid


2 — 2 — —
waste
Recycled construction debris — 2 1 — —

Bottom ash 2 — 2 — —
Ash from Biomass plants or
1 — 1 — —
paper mills
* No. of states.

Barriers to Technology Implementation

23. Please select/write-in what you think are possible barriers to implementation of new concrete
technologies/materials. (check all that apply).

Barriers to Technology Implementation No. of states


Technology not sufficiently proven to be adopted 30
Too expensive to use 28
Not enough training to use the technology 27

No specifications or construction guidelines available 23


Industry resistance 21
Not sufficient time available to be devoted to new technologies 16
Bad experience with construction or performance of the
13
technology/material
Resistance to change or to explore new technologies 12

Weather conditions not permitting 5

Other responses:
• Lack of experience by agency and local industry.
• Concern about potential reduction in concrete mixture quality.
• Ability to assess long term concrete durability with some technologies.
• Time and cost effectiveness.
• Implementation challenges.
• Different environmental conditions of jobsite compared to research phase.
• Finding an application to justify time and cost to use the technology.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Responses to Survey Questionnaire   161  

Case Examples

24. The synthesis will include “Case Examples” of five agencies with extensive experience in
new concrete technologies, and two other agencies with little or no experience with most
technologies. Would you be interested in participating on behalf of your agency in this
effort?

Yes 24

No 16

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

APPENDIX C

State DOT Specifications/


Special Provisions

1. Florida DOT
Specification Section 346 – Item 3.3 – Mass Concrete
2. Illinois DOT
a. Specification Item – Internally Curing Concrete with Lightweight Aggregate for Bridge
Deck Internal Curing.
b. Special Provision for Shrinkage Reducing Admixture in Bridge Deck Concrete.
c. Special Provision for Shrinkage-Compensating Concrete in Bridge Deck Concrete.

3. New York DOT


Special Provision – Concrete Mixture Performance Requirement
4. West Virginia DOT
Special Provision – Structural Concrete Internal Curing

162

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

State DOT Specifications/Special Provisions   163  

Florida Department Of Transportation

Section 346
Item 3.3 – Mass Concrete

346-3.3 Mass Concrete: When mass concrete is designated in the Contract Documents, use a
Specialty Engineer to develop and administer a Mass Concrete Control Plan (MCCP). Develop
the MCCP in accordance with Section 207 of the ACI Manual of Concrete Practice to ensure
concrete core temperatures for any mass concrete element do not exceed the maximum allowable
core temperature of 180°F and that the temperature differential between the element core and
surface do not exceed the maximum allowable temperature differential of 35°F. Submit the
MCCP to the Engineer for approval at least 14 days prior to the first anticipated mass concrete
placement. Ensure the MCCP includes and fully describes the following:

1. Concrete mix design proportions,


2. Casting procedures,
3. Insulating systems,
4. Type and placement of temperature measuring and recording devices,
5. Analysis of anticipated thermal developments for the various mass concrete elements for all
anticipated ambient temperature ranges,
6. Names and qualifications of all designees who will inspect the installation of and record the
output of temperature measuring devices, and who will implement temperature control measures
directed by the Specialty Engineer,
7. Measures to prevent thermal shock, and
8. Active cooling measures (if used).

Fully comply with the approved MCCP. The Specialty Engineer or approved designee shall
personally inspect and approve the installation of temperature measuring devices and verify that
the process for recording temperature readings is effective for the first placement of each size
and type mass component. The Specialty Engineer shall be available for immediate consultation
during the monitoring period of any mass concrete element. Record temperature measuring
device readings at intervals no greater than six hours, beginning at the completion of concrete
placement and continuing until decreasing core temperatures and temperature differentials are
confirmed in accordance with the approved MCCP. Leave temperature control mechanisms in
place until the concrete core temperature is within 50°F of the ambient temperature. Within three
days of the completion of temperature monitoring, submit a report to the Engineer which
includes all temperature readings, temperature differentials, data logger summary sheets and the
maximum core temperature and temperature differentials for each mass concrete element.

Upon successful performance of the MCCP, reduced monitoring of similar elements may be
requested. Submit any such requests to the Engineer for approval at least 14 days prior to the
requested date of reduced monitoring. If approved, the Specialty Engineer may monitor only the
initial element of concrete elements meeting all of the following requirements:

1. All elements have the same least cross-sectional dimension,


2. All elements have the same concrete mix design,
3. All elements have the same insulation R value and active cooling measures (if used), and
2

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

164   Concrete Technology for Transportation Applications

4. Ambient temperatures during concrete placement for all elements is within minus 10°F or plus
5°F of the ambient temperature during placement of the initial element.

Install temperature measuring devices for all mass concrete elements. Resume the recording of
temperature monitoring device output for all elements if directed by the Engineer. The
Department will make no compensation, either monetary or time, for any impacts associated
with reduced monitoring of mass concrete elements.

Mass concrete control provisions are not required for drilled shafts supporting sign, signal,
lighting or intelligent transportation (ITS) structures. At the Contractor’s option, instrumentation
and temperature measuring may be omitted for any mass concrete substructure element meeting
all of the following requirements:

1. Least cross-sectional dimension of six feet or less,


2. Insulation R value of at least 2.5 provided for at least 72 hours following the completion of
concrete placement,
3. The environmental classification of the concrete element is Slightly Aggressive or Moderately
Aggressive,
4. The concrete mix design meets the mass concrete proportioning requirements of 346-2.3, and
5. The total cementitious content of the concrete mix design is 750 lb./cy or less.

If either the maximum allowable core temperature or temperature differential of any mass
concrete element is exceeded, implement immediate corrective action as directed by the
Specialty Engineer to remediate. The approval of the MCCP shall be revoked. Do not place any
mass concrete elements until a revised MCCP has been approved by the Engineer. Submit an
analysis prepared by a Specialty Engineer to the Engineer for approval which addresses the
structural integrity and durability of any mass concrete element which is not cast in compliance
with the approved MCCP or which exceeds the allowable core temperature or temperature
differential. Submit all analyses and test results requested by the Engineer for any noncompliant
mass concrete element to the satisfaction of the Engineer. The Department will make no
compensation, either monetary or time, for the analyses and tests or any impacts upon the
project.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

State DOT Specifications/Special Provisions   165  

Illinois Department of Transportation

Internally Curing Concrete With Lightweight Aggregate For Bridge Deck (BMPR)
Effective: January 1, 2016

Description. This item shall consist of the materials and mix design for internally curing concrete with
lightweight aggregate according to the requirements of Section 1020 of the Standard Specifications, and
the following.

Materials. Revise Article 1020.02(c) as follows:

Item Article/Section
(c) Fine Aggregate (Note 1) .................................................................................................................... 1003

Note 1. The Contractor shall replace a portion of the normal weight fine aggregate with pre-wetted
lightweight fine aggregate, pre-wetted lightweight coarse aggregate, or a combination thereof to provide
water for internally curing the concrete as specified herein. The lightweight aggregate shall be an expanded
shale, expanded blast furnace slag, expanded slate, or expanded clay product according to ASTM C 1761.
The lightweight fine aggregate shall be Gradation FA 1, FA 2, FA 20, or FA 21, and the lightweight coarse
aggregate shall be Gradation CA 14 or CA 16. Storage of lightweight aggregate shall be according to Article
1003.01(e), except the stockpile shall be on a sloped surface. Lightweight aggregate stockpiles shall be
uniformly wetted with a sprinkler system for a minimum 72 hours, and then allowed to drain for 22 ± 2
hours immediately prior to use. Lightweight aggregate from different sources shall not be mixed without
permission of the Engineer.

Proportioning and Mix Design. Proportioning and mix design shall be for Class BS concrete and as
follows.

(a) Water/Cement Ratio. The water/cement ratio shall not be less than 0.36.

(b) Paste Content. The total cement plus finely divided minerals and water content shall not exceed
26% by volume of the mix design. The minimum cement factor may be reduced to 5.80 cwt/cu yd
(345 kg/cu m).

(c) Volume of Lightweight Aggregate. The pre-wetted lightweight aggregate shall replace a minimum
30 percent, by volume, of the normal weight fine aggregate.

(d) Batching. Immediately prior to batching, the pre-wetted and drained lightweight aggregate shall
have a field absorbed moisture content value not less than 15 percent. The field absorbed moisture
content shall be determined according to ITP ICC-1. Stockpiles that do not achieve the minimum
degree of absorption shall receive additional wetting and be allowed to drain for a minimum 12
hours prior to determining field absorbed moisture content again.

Trial Batch. For a new mix design to be verified, the Engineer will require the Contractor to provide a trial
batch at no cost to the Department. The trial batch shall be scheduled a minimum 30 calendar days prior to
anticipated use and shall be performed in the presence of the Engineer. A minimum of 2 cu yd (1.5 cu m)
trial batch shall be produced and placed offsite. The trial batch shall be produced with the equipment,
materials, and methods intended for construction. The trial batch will be evaluated and tested by the
Engineer according to the “Portland Cement Concrete Level III Technician” course manual. The Engineer
may require the Contractor to provide a sample of the lightweight aggregate, at no cost to the Department,
to verify the specific gravity, absorbed moisture content, and desorption of the material.
4

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

166   Concrete Technology for Transportation Applications

Verification of the mix design will include trial batch test results and other criteria as determined by the
Engineer. The Contractor will be notified in writing of verification. Verification of a mix design shall in no
manner be construed as acceptance of any mixture produced. Tests performed at the jobsite will determine
if a mix design can meet specifications.

Quality Control Sampling and Testing of Lightweight Aggregate by the Contractor. The Contractor
shall sample and test the lightweight aggregate as follows.

(a) Gradation. The gradation shall be tested a minimum once per day prior to pouring, unless the stockpile
has not received additional aggregate material since the previous test. The gradation shall be determined
according to ITP 27.

(b) Moisture. The field absorbed moisture content and surface moisture of the lightweight aggregate
stockpile shall be determined daily at the start of production for that day, and then as needed to
control production throughout the day, according to ITP ICC-1.

Quality Assurance Sampling and Testing of Lightweight Aggregate by the Engineer. The Engineer
reserves the right to perform quality assurance tests on independent and split samples of the lightweight
aggregate. An independent sample is a field sample obtained and tested by only one party. A split sample
is one of two equal portions of a field sample, where two parties each receive one portion for testing. The
Engineer may request the Contractor to obtain a split sample. The results of all quality assurance tests by
the Engineer will be made available to the Contractor. However, Contractor split sample test results shall
be provided to the Engineer before Department test results are revealed. The Engineer’s quality assurance
independent sample and split sample testing for placement or acceptance will be as follows:

(a) Gradation. One independent or split sample test at the beginning of the project. Thereafter, independent
testing frequency will be as determined by the Engineer, and split testing frequency will be a minimum
of 10 percent of the total tests required of the Contractor.

(b) Moisture. One independent or split sample test at the beginning of the project, and as determined by the
Engineer thereafter.

Comparing Lightweight Aggregate Test Results. Differences between the Engineer’s and the
Contractor’s split sample test results will be considered reasonable if within the following limits:

Test Parameter Acceptable Limits of Precision


Gradation See “Guideline for Sample Comparison” in
Appendix “A” of the Manual of Test Procedures
for Materials.
Moisture 0.5%

Action shall be taken when either the Engineer’s or the Contractor’s test results are not within specification
limits. Action may include, but is not limited to, immediate retests on a split sample; investigation of the
sampling method, test procedure, equipment condition, equipment calibration, and other factors; or the
Contractor being required to replace or repair test equipment as determined by the Engineer.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

State DOT Specifications/Special Provisions   167  

State of Illinois
Department of Transportation

SPECIAL PROVISION
FOR
SHRINKAGE REDUCING ADMIXTURE IN BRIDGE DECK CONCRETE

Effective: December 5, 2011

This Special Provision requires the use of a shrinkage reducing admixture to lower the bridge deck concrete
drying shrinkage. The Contractor shall provide a technical representative to assist in mix design,
proportioning, batching, placement, finishing, and curing of the concrete. The technical representative shall
be present for the jobsite trial batch and the first day of the bridge deck pour.

Shrinkage Reducing Admixture Concrete Requirements

The Contractor shall provide test data that the shrinkage reducing admixture meets the freeze/thaw
requirements according to ASTM C 494, Type S (specific performance), or other freeze/thaw test data to
show the shrinkage reducing admixture does not harm the concrete. The Department will maintain an
approved list of shrinkage reducing admixtures.

The bridge deck concrete shall be Class BS and shall meet Section 1020 with the following additions or
modifications.

(a) The cement shall be Type I or II. The coarse aggregate shall be crushed limestone or dolomite or
gravel.

(b) When determining water/cement ratio, the Contractor shall calculate 70 percent of the shrinkage
reducing admixture as water.

(c) The air content range shall be 6.0 to 8.5 percent.

(d) The mix design mortar factor range shall be 0.70 to 0.86.

(e) Only admixtures and finely divided minerals compatible with the shrinkage reducing admixture
shall be used, and alkali-silica reaction shall be addressed when specified in the contract plans.

(f) Microsilica and high reactivity metakaolin shall not be employed as a finely divided mineral.

(g) The Contractor shall be responsible for determining material proportions. The mix design will be
verified by the Bureau of Materials and Physical Research. Verification of a mix design shall in
no manner be construed as acceptance of any mixture produced.

(h) The batch sequence of materials shall be per the Manufacturer’s recommendation.

(i) Truck mixers providing truck-mixed or shrink-mixed concrete shall be limited to a volume at least 2
cubic yards less than the rated maximum mixing capacity as determined according to Article
1020.11(a)(6).

(j) The shrinkage reducing admixture dosage shall be determined by the Contractor, and a jobsite trial
batch will be required. The trial batch shall be performed according to the current “Portland Cement
6

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

168   Concrete Technology for Transportation Applications

Concrete Level III Technician Course – Manual of Instructions for Design of Concrete Mixtures”.
The shrinkage reducing admixture may be added at the plant or the jobsite. If added at the plant,
sampling for the trial batch shall be delayed a period of time that considers transportation from the
plant to the jobsite.

The Contractor will be required to perform shrinkage testing according to ASTM C 157. The
laboratory performing this testing shall have been inspected by the Cement and Concrete Reference
Laboratory (CCRL). The concrete shrinkage shall be determined after 7 days of cure plus 28 days
of drying, and shall be -0.030 percent. Verification of the mix design by the Engineer shall be
according to the current “Portland Cement Concrete Level III Technician Course – Manual of
Instructions for Design of Concrete Mixtures”.

Measurement and Payment


The concrete will be paid for at the contract unit price per cubic yard (cubic meter) for BRIDGE DECK
(SHRINKAGE REDUCING ADMIXTURE). This work will be measured according to Article 503.21.

State of Illinois
Department of Transportation

SPECIAL PROVISION
FOR
SHRINKAGE-COMPENSATING CONCRETE IN BRIDGE DECK CONCRETE

Effective: November 28, 2011

This Special Provision requires the use of a dry expansive component to produce a shrinkage-compensating
concrete mixture for the bridge deck. The Contractor shall provide a technical representative to assist in
mix design, proportioning, batching, placement, finishing, and curing of the shrinkage-compensating
concrete. The technical representative shall be present for the jobsite trial batch and the first day of the
bridge deck pour.

Dry Expansive Component Requirements

Revise Section 1010.01 of the Standard Specifications to read:

“1010.01 Description. Finely divided minerals shall include fly ash, microsilica (silica fume), high-
reactivity metakaolin (HRM), ground granulated blast furnace slag (GGBF), and dry expansive
components. The finely divided minerals will be approved according to the current Bureau of Materials and
Physical Research Policy Memorandum, “Acceptance Procedure for Finely Divided Minerals Used in
Portland Cement Concrete and Other Applications”. The Department will maintain an approved list of
suppliers for finely divided minerals.”

Different sources or types of finely divided minerals shall not be mixed or used alternately in the same item
of construction, unless approved by the Engineer.

Add Section 1010.06 to the Standard Specifications to read:

“1010.06 Dry Expansive Component. The dry expansive component material shall be Type K or Type G
and shall be defined according to ACI 223R. The expansive component shall be used in combination with
Type I or II cement. The minimum restrained expansion shall be 0.04 percent at seven days according to
ASTM C 806. The maximum restrained expansion shall be 0.18 percent.”

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

State DOT Specifications/Special Provisions   169  

Shrinkage-Compensating Concrete Requirements

The shrinkage-compensating concrete for the bridge deck shall be Class BS and shall be according to the
special provision, “Portland Cement Concrete”, with the following revisions.

(k) The cement shall be Type I or II. The coarse aggregate shall be crushed limestone or dolomite.

(l) The water-cement ratio shall be defined according to the special provision, “Portland Cement
Concrete”. When either the Type K or Type G expansive component is used, it shall be included
with the cement plus finely divided mineral in the calculation of water/cement ratio.

(m) The maximum water-cement ratio for concrete mixture shall be 0.50 for the Type K expansive
component and 0.44 for the Type G expansive component.

(n) For the Type K expansive component, the slump shall be 5-7 in. (125-175 mm). The slump range
includes the use of a high range water-reducing admixture.

(o) For the Type K or G expansive component, the concrete mixture shall have a minimum of 605
lbs./cu yd (360 kg/cu m) of cement, finely divided mineral, and expansive component summed
together. The maximum shall be 705 lbs./cu yd (418 kg/cu m). For concrete mixtures utilizing the
Type G expansive component, fly ash shall not be employed as a finely divided mineral. For
concrete mixtures utilizing the Type K or G expansive component, microsilica and high reactivity
metakaolin shall not be employed as a finely divided mineral. The amount of cement may be a
minimum of 455 lbs./cu yd (270 kg/cu m) in the mix design. For concrete mixtures utilizing the
Type K expansive component, the amount of either fly ash or ground granulated blast-furnace slag
may be a maximum 25 percent of the cement, finely divided mineral, and expansive component
summed together. For concrete mixtures utilizing the Type G expansive component, the amount of
ground granulated blast-furnace slag may be a maximum 25 percent of the cement, finely divided
mineral, and expansive component summed together. Article 1020.05(c)(1)d shall not apply.

(p) The mix design mortar factor range shall be 0.70 to 0.86.

(q) Only admixtures and finely divided minerals compatible with the expansive component shall be
used, and alkali-silica reaction shall be addressed when specified in the contract plans.

(r) The batch sequence of materials shall be per the Manufacturer’s recommendation.

(s) Truck mixers providing truck-mixed or shrink-mixed shrinkage-compensating concrete shall be


limited to a volume at least 2 cubic yards less than the rated maximum mixing capacity as
determined according to Article 1020.11(a)(6).

(t) The Contractor shall be responsible for determining material proportions. The mix design will be
verified by the Bureau of Materials and Physical Research. Verification of a mix design shall in no
manner be construed as acceptance of any mixture produced.

(u) The amount of expansive component shall be determined by the Contractor, and a jobsite trial batch
will be required. The trial batch shall be performed according to the current “Portland Cement
Concrete Level III Technician Course – Manual of Instructions for Design of Concrete Mixtures”.
The Type K or G expansive component may be added at the plant or jobsite. If added at the plant,

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

170   Concrete Technology for Transportation Applications

sampling for the trial batch shall be delayed a period of time that considers transportation from the
plant to the jobsite.

The Contractor will be required to perform restrained expansion testing according to ASTM C 878.
The laboratory performing this testing shall have been inspected by the Cement and Concrete
Reference Laboratory (CCRL). The maximum restrained concrete prism expansion shall be a
minimum of 0.05% and a maximum of 0.09% according to ASTM C 878. Verification of the mix
design by the Engineer shall be according to the current “Portland Cement Concrete Level III
Technician Course – Manual of Instructions for Design of Concrete Mixtures”.

Measurement and Payment

Shrinkage-compensating concrete will be paid for at the contract unit price per cubic yard (cubic meter) for
BRIDGE DECK (SHRINKAGE-COMPENSATING CONCRETE). This work will be measured according
to Article 503.21.

New York Department of Transportation


Special Provision

ITEM 557.00000211 – 5000 PSI CONCRETE FOR STRUCTURAL APPROACH


SLAB – TYPE 1 FRICTION

DESCRIPTION
Furnish and place structural approach slab with a minimum compressive strength of 5,000 psi
meeting the performance requirements defined herein where specified in the contract documents.
The provisions of §557 shall apply except as noted herein.

MATERIALS

Reinforcement
The provision of §556-2 shall apply. All Reinforcement shall meet the requirements of §709-13,
Stainless Steel Bar Reinforcement.

Concrete
The provisions of §557-2 shall apply, except as modified herein.

1. Design a concrete mixture proportioned according to the American Concrete Institute


Manual of Concrete Practice, ACI 211.1, Standard Practice for Selecting Proportions for
Normal, Heavyweight, and Mass Concrete. Produce a homogeneous mixture of cement,
pozzolan (fly ash or GGBFS), fine aggregate, coarse aggregate, air entraining agent, water-
reducing and set-retarding admixture, and water as designed. Other NYSDOT Approved List
materials may be used as approved by the Director, Materials Bureau.

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

State DOT Specifications/Special Provisions   171  

2. The designed concrete mixture shall meet the following requirements:


• Strength: 28-day minimum compressive strength of 5,000 psi.
• Slump: 6ʺ to 9ʺ.
• Entrained Air: 5 to 8%.
• Water/Total Cementitious Material Ratio: 0.40 maximum.
• Use Type I, I/II, or II cement. Use 15% to 35% fly ash or 30% to 70% GGBFS by weight
of cementitious materials.
• Resistivity >37(kΩ-cm) (AASHTO T358) or Permeability <1000 Coulombs (AASHTO
T 277) at 28 days of age. The time frame may be extended to 56 days moist cure for high
pozzolan content mix designs or the Accelerated Moist Curing alternative may be used.

3. Perform mix development testing in accordance with ASTM C143, C231, C192 and C39, to
assure all performance criteria can be achieved during production and placement.

4. The maximum aggregate size used in a concrete mixture shall be dependent on the size and
shape of the concrete member and on the amount and distribution of reinforcing steel. The
Contractor shall select the largest available nominal maximum size of aggregate which does
not exceed the following:
• three-quarters of the clear distance between reinforcing bars and between the reinforcing
bars and the forms; and
• one-third the thickness of the placement.

5. At least 1 month prior to the start of any concrete placement, provide a copy of the proposed
mixture design(s) and trial batch test results to the Director, Materials Bureau, submitted
through the Regional Materials Engineer, for evaluation. Submit sufficient data to permit the
Director to offer an informed evaluation. Include at least the following:
• Concrete mix proportions.
• Material sources. Also include fineness modulus and specific gravity for all aggregates.
• Air content of plastic concrete.
• Slump of plastic concrete.
• Compressive strength at 7, 14, 28, and 56 days, and at any other age tested or deemed
necessary.
• Resistivity or Permeability test data showing results of >37(kΩ-cm (AASHTO T358) or
<1000 Coulombs (AASHTO T 277) respectively.

Do not interpret having a valid mixture design as approval of the mixture. Also, resubmit any
proposed mixture design change to the Director, Materials Bureau, for evaluation. Multiple
mixture designs may be used to address performance and placement issues as deemed necessary
by the Contractor. Submit each mixture for evaluation, as indicated above, prior to use.

CONSTRUCTION DETAILS
The provisions of §556-3 and §557-3 shall apply, except as modified herein:

A. Bar Reinforcement
Placement details and bar lists are not included in the contract plans, the following provisions
apply:
10

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

172   Concrete Technology for Transportation Applications

1. The Contractor shall submit a minimum of two copies of the bar lists and placement
drawings showing the bar locations to the Engineer. The details of the bar list
drawing and placement shall meet the requirements of the current edition of the
Concrete Reinforcing Steel Institute’s publication Reinforcing Bar Detailing.
Drawings submitted for the review possess the same size and layout as the Plans.
Electronic submission is required. Drawings and bar lists shall be clear and legible.

2. The Engineer will transmit the documents to the designer for review for conformance
with the design requirements and in accordance with §105-16. The designer will not
check lengths, number of bars, weights or bar marks. Corrections will be returned to
the Contractor. A review time of two days per placement drawing submitted with a
minimum of 15 days for each submission will be allowed upon receipt of the
submission. When the documents are satisfactory they will be returned to the
Contractor stamped “Approved in Conformance with Design Requirements”. The
Contractor shall supply the Engineer with five (5) copies of the approved documents.
No reinforcement shall be placed until copies of the approved documents are received
by the Engineer.

3. The reinforcement shall be of the type indicated in the contract documents.

4. Partial submissions that require coordination with other drawings will not be
accepted.

B. Concrete
Prior to placing any concrete required by this specification, perform a trial placement of at
least 8 cubic yards using the proposed mixture design(s). This trial placement(s), when
approved by the Engineer, may be incorporated into the project as a substitute for the
placement of another Class of concrete shown on the plans. If used in another element as a
trial placement, the entire placement for that element on the day of the trial must use the
same concrete. The Department will make and test concrete cylinders from the trial
placement(s) to verify laboratory test results.

The loading limitations of §555-3.10 apply, except that concrete cylinder sets designated for
early loading must attain an average compression strength of 5,000 psi, or greater, with no
individual cylinder less than 4,500 psi.

1. To evaluate 28-day strength of the concrete, the Department will cast cylinders
following the requirements and frequency of Materials Method 9.2 for each
placement, with a minimum of two (2) 6ʺ x 12ʺ cylinders for each day. The results of
all test cylinder specimens representing an element placed, or part thereof, on a given
day will be averaged to determine the ultimate compressive strength for each
placement. The average shall be 5,000 psi with no individual cylinder less than 4,500
psi

11

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

State DOT Specifications/Special Provisions   173  

If any strength test falls below the criteria established above, the Engineer will
determine if investigation is required. The investigation may consist of, but is not
limited to, review of the foll owing:
• Sampling and testing of plastic concrete,
• Handling of cylinders,
• Cylinder curing procedures, or
• Compressive strength testing procedures.

If necessary, coring may be required to determine in-place strength. The contractor


shall perform all coring at locations directed by the Engineer.

Make any repairs as per the provisions of §555-3.13, Damaged or Defective


Concrete. The Engineer will reject any concrete represented by a 28-day cylinder set
with an average compressive strength less than 5,000 psi, or an individual cylinder
with a compressive strength less than 4,500 psi. Proposed repairs require Deputy
Chief Engineer, Structures approval.

2. To evaluate the Resistivity of the concrete, the Department will cast cylinders at the
same frequency and from the same sample(s) of concrete used to cast compressive
strength specimens, with a minimum of one (1) set per placement. A set consists of
three (3) 4″ x 8″ cylinders. Cylinders will be cured for 28 days. The time frame may
be extended to 56 days moist cure for high pozzolan content mix designs or the
Accelerated Moist Curing alternative may be used (7 day normal cure at 73 degrees F,
21 days wet cure at 100 degrees F). The results of all test cylinder specimens
representing an element placed, or part thereof, on a given day will be averaged to
determine the Resistivity for each placement. The average shall be >37 k -cm
(AASHTO T358).

Permeability will be considered an alternative method to measure durability and will


require only two (2) 4″ x 8″ cylinders sampled at the same frequency as for
compressive strength. Cylinders will be cured for 28 days. The time frame may be
extended to 56 days moist cure for high pozzolan content mix designs or the
Accelerated Moist Curing alternative may be used (7 day normal cure at 73 degrees F,
21 days wet cure at 100 degrees F). The results of all test cylinder specimens
representing an element placed, or part thereof, on a given day will be averaged to
determine the Permeability for each placement. The average shall be ≤1000 Coulombs
(AASHTO T 277).

If any Resistivity / Permeability test data falls outside the criteria established above,
the Engineer will determine if an investigation is required. The investigation may
consist of, but is not limited to, review of the following:
• Sampling and testing of plastic concrete,
• Handling of test cylinders,
• Cylinder curing procedures, or
• Permeability testing procedures.

12

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

174   Concrete Technology for Transportation Applications

METHOD OF MEASUREMENT:

The provisions of §557-4 shall apply.

BASIS OF PAYMENT:

The provisions of §557-5 shall apply and the cost of the mix design, trial placement and all
laboratory testing shall also be included.

Pay adjustments will be made for cast-in-place concrete that does not meet specified
performance characteristics and shall be computed on the actual quantity of concrete
representing an element placed, or part thereof, on a given day. The concrete pay adjustment
(CPA) will be made for non-conforming material according to the formulas defined as follows:

For concrete not meeting strength requirement, but allowed to remain in place, the payment
representing the quantity of concrete for a given day / element’s placement shall be reduced as
follows:

Compressive Strength Pay Factor (PF)


>100% of fc′ The Department will pay 100%
>95.0% and <100.0% of fc′ The Department will pay 87.5%
>90.0% and <95.0% of fc′ The Department will pay 75%
< 90.0% of fc′ Reject concrete

For concrete not meeting resistivity / permeability requirements, but allowed to remain in place,
the payment representing the quantity of concrete for a given day / element’s placement shall be
reduced as follows:

Surface Resistivity Permeability Pay Factor (PF)


(kΩ-cm) Coulombs (C)
>37 <1000 The Department will pay 100%
<37 and >27 >1000 and <1500 The Department will pay 87.5%
< 27 and >19 >1500 and <2500 The Department will pay 75%
<19 >2500 Reject concrete
When a concrete mix contains corrosion inhibitor, all resistivity values will be decreased by
20% or permeability values will be increased by 20%

The total concrete pay adjustment for compressive strength and resistivity for a given day’s
placement / element shall be computed as

CPA = [(Compressive strength PF) (0.60)] + [(Resistivity PF) (0.40)]

13

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

State DOT Specifications/Special Provisions   175  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

176   Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

State DOT Specifications/Special Provisions   177  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

178   Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

State DOT Specifications/Special Provisions   179  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

180   Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

State DOT Specifications/Special Provisions   181  

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

Abbreviations and acronyms used without definitions in TRB publications:


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

Copyright National Academy of Sciences. All rights reserved.


Concrete Technology for Transportation Applications

ADDRESS SERVICE REQUESTED

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

ISBN 978-0-309-48101-4
NON-PROFIT ORG.
COLUMBIA, MD
PERMIT NO. 88

90000
U.S. POSTAGE
PAID

9 780309 481014

Copyright National Academy of Sciences. All rights reserved.

Das könnte Ihnen auch gefallen