Sie sind auf Seite 1von 223

ACKNOWLEDGMENTS

A STUDY LIKE THIS involves the research and help of countless organizations.
In particular, we would like to acknowledge the help of the following:

American Association of State Highway & Transportation Officials

American Concrete Institute

American Concrete Pavement Association

American Society of Civil Engineers

Environmental Protection Agency

National Ready Mix Concrete Association

National Stone Association

Portland Cement Association

The National Concrete Pavement Technology Center (National CP Tech Center)


at Iowa State University

United States Department of Transportation – Federal Highway Administration

Wisconsin Department of Transportation

1
EXECUTIVE
SUMMARY

THIS ANALYTICAL STUDY PROVIDES AN OVERVIEW of, and reference to, the cur-
rent practices in the European and North American concrete pavement industries. It offers
opinions of the Editorial Technical Advisory Committee with regard to best practices, as
well as opinions about those items and practices it deems out-of-date, even though they
are still in place within either market.
Transfer of technology is always a risky endeavor to be considered, and this analytical
study provides some guidance with regard to current practices on both continents.
The study attempts to analyze the current differences in the European and North
American marketplaces based on the authors’ firsthand and combined experiences, as well
as on the results of tests, studies, and other publications as referenced. The reader will find
cross-referenced publications, association and professional society information, specifica-
tions, test results, opinions, and editorials herein. Even so, the analytical study is designed
to offer the most unbiased and objective look possible at the ongoing specifications, stan-
dards, constituent materials, practices, and methodologies that make up the very best in
concrete paving, and which unfortunately, in some cases, make up the very worst in con-
crete paving practices.
At the time of publication, the study makes comment on current trends and types of
equipment, methodology and practices in constituent materials design, and level of devel-
opments in the European market and in the North American market. The reader can take
away an idea of what both markets may offer regarding the trends and developments in the
growing Russian concrete pavement marketplace and sectors.

About Robert Ober & Associates, LLC.


The mission of Robert Ober & Associates, LLC., and its related firms is to provide concrete
technology information through research, technological transfer, and technology imple-
mentation. As an international design and consulting firm involved in turnkey project
development and management of advanced and high-performance concrete production

2
E X E CUT I V E S UM MA RY 3

organisms, the company is often involved in the transfer of technology from mature
markets to emerging markets, and more often than might be expected, from emerging
markets back to more-mature markets.
Robert Ober & Associates, LLC., greatly encourages employee support for the global
efforts of Engineers Without Borders (EWB), as sponsored by the American Society of
Civil Engineers (ASCE), and for the Concrete Industry Management (CIM) degree pro-
gram at various American universities and colleges, as supported by American concrete
associations and societies; it also supports local endeavors involving continuing education
and volunteerism. Matching donations of time and other gifts are provided to employees
for their participation in such programs.
The company has a worldwide pedigree of high-performance concrete production
installations, and provides design-build services to governments, utilities, production
facility owners, construction managers, and contractors.
“In globally sharing learned experiences and technology concerning concrete products
and construction practices, the world’s infrastructures are greatly enhanced; where there is
advancement in infrastructure, there is advancement of mankind.”

—Robert Ober, Founder and Chairman

Nondiscrimination Statement
Robert Ober & Associates, LLC., does not discriminate on the basis of race, color, age, religion,
national origin, sexual orientation, gender identity, sex, marital status, disability, or status as a U.S.
veteran, or as a foreign services veteran. Inquiries can be directed to the Director of Equal Oppor-
tunity and Diversity, Human Resources Department, at: info@robertober.com

Disclaimer Notice
The contents of this analytical study reflect the views of the authors, who are responsible for the
facts and the accuracy of the information presented herein. The opinions, findings, and conclusions
expressed in this study are those of the authors and not necessarily those of the sponsors. The spon-
sors assume no liability for the contents or use of the information contained in this document. This
analytical study does not constitute a standard, specification, or regulation.
The sponsors do not endorse any products or manufacturers. Trademarks or manufacturers’
names appear in this report only because they are considered essential to the objective of the
document.
STUDY
OBJECTIVES

THE OBJECTIVE OF THIS ANALYTICAL REPORT on concrete pavement equipment


and methodology is to provide the reader with an overview of present-day practices for
concrete pavements and road surfaces within North America and Europe.
The general purpose of this particular report, prepared for the Russian marketplace, is
to provide a basis for considered personal analysis by the Russian reader while providing
analyses of different methodology and equipment practices, as gauged between the North
American marketplace and the European marketplace. The object of the study is thus to
provide ongoing practices as generally found in normal circumstances in the North Amer-
ican and European markets, and an overview of what is considered out-of-date or obsolete
about such practices, and also to address upcoming trends and new innovations develop-
ing within these markets. Secondarily the report covers all current specifications and
industry-accepted products and constituents, in order to allow the reader to discern those
that are practical for possible transfer of technology from either marketplace to the Rus-
sian market.
The comparison of the European and North American concrete pavement markets’
equipment and methodology allows the Russian reader a review of the best practices and
of the trends in each market, so as to judge which transfers of technologies from Europe
to North America, and from North America to Europe, have been successful and which
have not.
The Russian marketplace has much in common with the North American market-
place, with long distances between major metropolitan areas, and varying climatic and
road conditions. There is the need for quantities and qualities of concrete for pavement
which, in many respects, are alike in both markets. On the other side of this comparison,
it is also known that many current practices and some specifications have been developed
based entirely on European standards. This dichotomy presents interesting choices for the
Russian marketplace as it further develops with regard to concrete pavements, and offers
many opportunities. This report is developed to help those with firsthand knowledge of
the Russian marketplace to apply ideas and operational methods and concrete science
from each market.
5
S T UDY O BJ E CT IV ES 6

Concrete technologies, while not always developed in one marketplace entirely or


another, vary greatly between Europe and North America. Today there is a convergence of
best practices, but not at a fast rate, and not in an agreed-upon fashion. Therefore one of
the key purposes of this study is to allow the Russian reader an overview of what is consid-
ered state-of-the-art in each marketplace and note what might be the trend toward one
type of practice or methodology or another, even though the practice may not yet have
resulted in an actual transfer of technology in practice as of this writing.
Specifically with regard to equipment and with regard to emerging technologies and
practices, one market can learn from the other market—while the upcoming and still-
developing Russian market can learn from each of the North American and the European
markets—both the best and worst ongoing practices.
While the study was not commissioned to make recommendations per se, it does
draw some obvious conclusions from the reported data—and makes no attempt to dis-
guise what are believed to be the current best practices in concrete pavements and roads.

The study provides an in-depth look at the following areas:


N Materials
N Concrete Mix Designs
N Standards
N Equipment for Concrete Pavements
N Practice and Methodology
N Trends and New Technologies

The study also provides:


N Associations and Standards References
N Affiliated and Referenced Studies and Reports
N Glossary of Terminology

The 2010 edition study is the first of a several-part series to be undertaken for private
industry in Russia, including but not limited to concrete production and practices
specifically for:
N Pavement and Roads
N Tunnels and Hydro Structures
N Prestressed Bridges and Spanning Structures
N Precast Housing and Commercial Structural Construction
N Architectural Precast Concrete Cladding
N Rail Sleeper and Rail Concrete Production
N Reinforced and Non-Reinforced Concrete Pipe
N Concrete Block and Paver Production/Construction
N Spun-Cast Concrete Utility Poles and Structures
GLOSSARY

A
AASHTO: American Association of State Highway and Transportation Officials.
Absorption: The process of water infiltrating into concrete. The amount of water absorbed
is expressed as a percentage of the dry weight of the concrete.
Accelerating admixture, or accelerator: An admixture that speeds the rate of hydration,
shortens the set time, or increases the rate of hardening.
Accelerator: An admixture added to concrete to increase the rate of early-strength
development or to shorten the time of setting, or both.
Acceptable Quality Level (AQL): A statistical definition of “good” quality in a lot of concrete
that gives the maximum percentage of allowable defective materials.
Acceptance Quality Characteristics (ACQs): Measurements of characteristics such as
strength, slab thickness, smoothness, air content, and consolidation around dowels,
used to predict future pavement performance.
ACI: American Concrete Institute.
ACPA: American Concrete Pavement Association.
ACR: Alkali-Carbonate Reaction.
Admixture: A material, other than water, aggregate, and Portland cement, that is added to
a batch of concrete before and during mixing. Common admixtures include air-
entraining and accelerators. S ee also Mineral Admixture.
Agglomeration: An unwanted build-up, producing lumps of uneven hydration, that
sometimes occurs when mixing lightweight and flaky materials in concrete.
Aggregate: Granular minerals, primarily sand and gravel, that form a part of a concrete
mix. Other granular materials include manufactured sand, crushed stone, and air-
cooled blast-furnace slag. S ee also Aggregate, Fine, and Aggregate, Coarse.
Aggregate Gradation: The distribution of granular materials of various sizes, expressed in
terms of cumulative percentages of the different sizes. Various methods are used to
determine gradation, but all use sieve openings to determine the sizes.
Aggregate Interlock: The spanning of larger aggregate particles, mostly gravel, across the
two (2) sides of a joint or crack in concrete. Aggregate interlock joins the two (2) sides
of a joint or crack to a certain extent, and so helps maintain mutual alignment and
load transfer.

7
GL OSS A RY 8

Aggregate, Air-Dry: Aggregate with a dry surface but with some water in its pores.
Aggregate, Angular: Aggregate particles that are sharp, with well-defined edges, rather
than rounded.
Aggregate, Coarse: Particles composed of natural gravel, crushed stone, or blast-furnace
slag, typically ranging in size between 9.5 mm and 37.5 mm.
Aggregate, Dense-graded: Aggregates graded to produce minimal void content and greater
weight.
Aggregate, Fine: Aggregate that passes through a 9.5-mm sieve, almost entirely passes
through a 4.75-mm (No. 4) sieve, and is predominantly retained on a 75-µm (No. 200)
sieve.
Aggregate, Gap-Graded: Aggregate that is missing intermediate sizes, resulting in an
inferior concrete mix.
Aggregate, High-Density: Concrete aggregate of heavy weight.
Aggregate, Lightweight: Aggregate composed of low-density materials such as expanded
or sintered clay, slate, shale, or slag, pumice, volcanic cinders, diatomite, sintered fly
ash, or industrial cinders.
Aggregate, Normal-Weight: Average-density aggregates, usually crushed stone or gravel
with a density of approximately 2,400 kg/m3.
Aggregate, Open-Graded: An aggregate mix that produces relatively large voids.
Agitating Truck: A truck that agitates concrete, usually with a drum roller, either while
moving or while stationary, so the concrete can maintain its plasticity.
Agitation: The process of providing gentle motion in mixed concrete just sufficient to
prevent segregation or loss of plasticity.
Air Content: The amount of air in concrete, including entrained and entrapped air but not
including pore space in the aggregate particles.
Air-Entrainment: An addition that incorporates minute bubbles of air in a concrete mix
during mixing. The resulting concrete has increased workability and resistance to
cracking due to frost.
Air Void: An entrapped air pocket or an entrained air bubble in concrete. Entrapped air
voids typically measure 1 mm or more and are irregular in shape; entrained air voids
are between 10.25 µm to 1 mm in diameter and are spherical.
Alkali-Carbonate Reaction (ACR): An unwanted reaction between the alkalies (sodium and
potassium) in Portland cement and certain carbonate rocks, particularly calcite
dolomite and dolomitic limestones, present in some aggregates, especially volcanic
aggregates. Excess expansion can result, which can cause the concrete to crack.
Alkali-Silica Reaction (ASR): An unwanted reaction between alkalies in Portland cement
and certain siliceous rocks or minerals, present in some aggregates. Excessive expan-
sion and the cracking of concrete can be the result.
Angle of Repose: The angle at which a pile of aggregate naturally rests when left
undisturbed.
Asphalt: A dark bituminous substance, composed mostly of hydrocarbons, that is usually
a residue of petroleum refining. The term often is used to mean Asphaltic Concrete.
Asphaltic Concrete: A paving surface composed of asphalt and aggregate.
ASTM: American Society for Testing and Materials.
Autoclave Expansion Test: A standard cement test that detects potential unsoundness in
concrete as a result of expansion due to free lime or magnesia in the cement.
GL OSS A RY 9

Average Absolute Deviation (AAD): A statistical measure of how much a particular sample
deviates from the average sample.

B
Backer Rod: A foam cord that is inserted into a joint channel to keep liquid sealant from
flowing out of the channel or to keep it from adhering to the channel.
Bag of Cement: An amount of cement that varies from country to country, but
that usually equals about 36 kg. A “six-bag” concrete mixture has six (6) bags
of cement per cubic yard (0.76 cubic meter) of concrete. Also referred to as
a Sack.
Base, or Base Course: The underlying surface, usually tamped gravel but sometimes
asphalt, that supports a concrete slab. The base functions to distribute loads, lessen the
effects of frost, and provide drainage. Sometimes called Subbase, though that term
often refers to the soil that underlies the gravel base.
Batch: An amount of concrete mixed at one time.
Batch Mixer: A concrete mixer that mixes and unloads a fixed amount of concrete at a time.
Batch Plant: A series of components that operate together in the production of concrete.
The plant provides for the storage feed, storage, storage discharge, weighing, transfer
of weighed materials, mixing of materials with water and chemical admixtures,
mixed material discharge, and recording of batch weights, mixing time, and
inventories.
Batching: Weighing or volumetrically measuring each of the concrete ingredients and
placing them in the mixer.
Bitumen: A liquid or semisolid (tar-like) binding material usually derived from crude
petroleum or from bituminous coal, serving as the adhesive agent in a bituminous
mixture.
Blast-Furnace Slag: S ee Ground, Granulated Blast-Furnace Slag.
Blast-Furnace Slag Cement: Hydraulic cement consisting of granulated blast-furnace slag,
either without additions or mixed with ground Portland cement clinker or Portland
cement.
Bleed Water: Water from a newly placed concrete mixture that appears on the surface due
to the settling of solid materials in the mixture.
Blended Cement: S ee Cement, Blended.
Blowup: A localized upward movement of a section of concrete, often occurring at a joint
or crack during hot weather when expansion creates compressive forces high enough
to cause buckling.
Bond: The adhesion of concrete to adjoining surfaces or to metal reinforcement. Also,
within a concrete mix, the adhesion of cement paste to aggregate.
Bonded Concrete Overlay: A resurfacing method in which a thin layer of new concrete (5
to 10 cm) is placed directly onto an existing concrete slab that has been cleaned and
prepared, so that the two bond together.
Bonding Agent: A liquid or paste product applied to an existing concrete or asphalt
surface to enhance a bond between it and a succeeding layer of concrete.
Bulk Density: The mass or weight of material that fills a container of a specified unit
volume, including both the weight of the particles and the air between them.
GL OS SA RY 10

Bulking: An unwanted condition in which a thin layer of water causes sand particles to
clump together so they aren’t readily compacted. This leads to an increase in volume,
which means a decrease in bulk density.
Bulk Supplements: Cementitious materials often added to a concrete mix along with
Portland cement, including fly ash and blast-furnace slag. Also called Supplementary
Cementitious Material (SCM).
Burlap: A coarse fabric, often used as a water-retaining cover to slow curing of concrete.

C
Calcium Chloride: CaCl2. Used as a drying agent, as a concrete accelerator, and for
de-icing of roads.
Cement, Blended: Cement that combines Portland cement with Supplementary Cementi-
tious Material (SCM) such as pozzolans, fly ash, slag, or other hydraulic cement.
Cement, High-Early-Strength: Cement that produces earlier strength in concrete than
regular cement. Often referred to as Type III cement.
Cement, Hydraulic: Cement capable of setting and hardening under water. The most
common type of hydraulic cement is Portland cement.
Cement, Portland: A calcium silicate cement produced by finely pulverizing clinker that
was produced by burning a defined mixture of iron, alumina, and silica at high
temperature.
Cement Paste: A combination of cement and water, which forms a constituent of concrete.
Cementitious Material: A material that contributes to the formation of hydrated calcium
silicate compounds, and so sets and hardens in the presence of water. Cementitious
materials include Portland cement; blended hydraulic cement; fly ash; ground, granu-
lated blast-furnace slag; silica fume; calcined clay; metakaolin; and calcined shale.
Central-Mixed: Concrete that is completely mixed in a stationary mixer, and is then
transported to the job.
Central Mixer: A stationary concrete mixer, usually very large, from which the fresh
concrete is loaded onto trucks and transported to the job.
Chemical Admixture: S ee Admixture.
Chemical Bond: A bond between materials developed by chemical reaction, producing
cohesion and adhesion.
Chloride Penetration: The infiltration of chemical compounds containing chloride ions
into a concrete surface, which promotes the corrosion of steel reinforcement. Chloride
de-icing chemicals are the most common source.
Clinker: The raw cementitious material produced by heating in a Portland cement kiln.
The clinker is then ground and mixed with other ingredients like gypsum to produce
Portland cement.
Coarse Aggregate: S ee Aggregate, Coarse.
Coefficient of Thermal Expansion (CTE): A slab’s change in length or volume per degree of
temperature change.
Colloidal Mixing: A mixing process that separates particles from one another to uniformly
distribute them in the concrete mix.
Combined Aggregate Grading: In a concrete mix, the distribution of fine and coarse
aggregates.
GL OS SA RY 11

Compaction: Reducing the volume of freshly mixed concrete to by inducing a closer


arrangement of solid particles and reducing voids (other than air-entrainment
bubbles). The most common compaction methods are vibrating and tamping. Also
called Consolidation.
Compression Seal: Joint sealant that is preformed by manufacture; it is compressed into
the joint and further held in place by liquid sealant poured or injected over it.
Compressive Strength: The resistance of concrete to compression, or axial loading. Typi-
cally a specimen of the concrete is tested, and the results are given as force per unit
of cross-sectional area, such as megapascals (MPa) or pounds per square inch (psi).
Concrete: A material that consists essentially of a binding medium, usually Portland
cement and water, in which particles or fragments of natural or artificial aggregates
are embedded.
Concrete, Reinforced: Concrete construction that contains wire mesh, steel bars, or
reinforcing fibers embedded in it.
Consistency: The relative ability of fresh concrete to flow. The usual measure of consis-
tency is slump. S ee also Slump and Workability.
Consolidation: A synonym for Compaction.
Construction Joint: A joint created at a stopping point when placing concrete. The two (2)
successive concrete placements of concrete are joined, typically with a keyway or
reinforcement across the joint. Construction joints are often located at contraction
(control) joint locations and are reinforced to perform as contraction joints.
Continuously Reinforced Pavement: A pavement with longitudinal steel reinforcement that
has no gaps, and no transverse expansion or contraction joints.
Contraction: A decrease in length or volume of a concrete pavement.
Contraction Joint: A joint designed to prevent the formation of shrinkage cracks elsewhere
in the concrete. Reinforcing steel does not traverse a contraction joint. The joint may
be grooved, sawed, or formed. Also known as a Control Joint.
Control Joint: S ee Contraction Joint.
Core Sample: A cylindrical specimen drilled and removed from a concrete slab so that it
can be tested for strength.
Corner Crack: A crack that usually occurs where a transverse joint meets a longitudinal
joint, with the crack connecting them to form a triangle. It may be only surface-deep,
or it may go through the slab.
Corrosion: Deterioration of metal that can be caused by chemical, electrochemical, or
electrolytic reaction.
Course: A horizontal layer of concrete, usually one of several making up a lift. S ee also
Lift.
Crazing: A pattern of thin pattern cracks in concrete, usually caused by unequal shrink-
age or contraction while drying or cooling. S ee also Hairline Cracking, Map Cracking,
and Pattern Cracking.
CRC Pavement (CRCP): Continuously reinforced concrete pavement.
Cross-Stitching: Reinforcing a crack or joint by drilling angled small-diameter holes
through both sides of the crack and then filling them with tie bars and grout.
Crushed Gravel: Gravel with stones that have been reduced in size by crushing, with a
specified minimum percentage of fragments having one (1) or more faces resulting
from fracture. S ee also Aggregate, Coarse.
GL OS SA RY 12

Crushed Stone: An aggregate resulting from crushing of rocks, boulders, or large stones,
in which most of the faces have well-defined edges resulting from the crushing
operation.
Cullet: Crushed and screened waste glass sometimes used as a substitute for sand in
concrete (including types of glass not considered suitable for bottle recycling). Fine
cullet can cause an Alkali-Silica Reaction, which compromises concrete strength and
durability.
Curing: The process of maintaining freshly placed concrete at a favorable temperature for
a suitable period of time during its early stages to ensure satisfactory hydration and
hardening of the cementitious materials.
Curing Compound: A chemical liquid applied to the surface of newly placed concrete to
retard the loss of water and so slow the curing process.

D
D-Cracks: Cracks that usually start at the bottom of a slab of concrete and move up toward
the surface, often accompanied by dark discoloration and a series of hairline cracks.
They are attributed to the use of aggregate stones that absorb too much moisture. Their
presence usually means the entire slab is compromised and needs to be replaced.
Deformed Bar, or Deformed Tie Bar: A concrete reinforcing bar (rebar) with a pattern of
surface ridges that enhances anchorage with the surrounding concrete.
Density: The mass or weight of concrete or aggregate per volume. Also referred to as unit
weight. Dry Density refers to the mass per volume of a dry material at a stated tem-
perature. S ee also Specific Gravity.
Deterioration: Failure of a concrete slab, such as cracking, delamination, flaking, pitting,
scaling, and spalling.
Diamond-Grinding: Use of equipment with numerous rotating diamond-impregnated
blades to remove a top surface, thereby smoothing bumps and improving readability.
Dolomite: A mineral with a specific crystal structure, composed of calcium carbonate and
magnesium carbonate in specific chemical amounts (54.27% and 45.73% by weight,
respectively); a rock containing dolomite as the principal constituent.
Dowel: A steel reinforcing bar, either plain round or deformed, extending into two (2)
adjoining portions of concrete so as to transfer shear loads. Also called a Load-
Transfer Device.
Dowel Retrofit: Placing dowels at transverse joints or cracks to tie slab sections together.
S ee also Load-Transfer Restoration.
Drainage: The removal of water from an area under, on, or along a roadway.
Dry-Mixing: Blending the solid components of concrete prior to adding the mixing water.
Drying Shrinkage: Contraction of concrete caused by drying.
Dynamic Load: Load that is imposed by moving or vibrating machinery producing
stresses in excess of those imposed by their dead load.

E
Early Strength: The strength developed by concrete soon after placement, often during the
first 72 hours.
GL OS SA RY 13

Edge Drains: Drains placed vertically adjacent to pavement to catch and remove water
from the subbase and subgrade.
Emulsification: The process of converting two (2) or more immiscible liquids (that is,
substances that are not capable of being combined so as to attain homogeneity) into a
single mixture.
Entrained Air: S ee Air-Entrainment.
Epoxy Resin: A class of organic chemical bonding agents used to coat reinforcing metals,
or as adhesives for concrete, or as binders in epoxy-resin concretes.
Expansion Joint: S ee Isolation Joint.
Exposed Aggregate: Surface texture where cement paste is washed away from a concrete
slab surface to expose durable chip-sized aggregates for the riding surface.

F
Falling Weight Deflectometer (FWD): A stationary device that drops a large weight on the
road surface and measures the deflection at the point of impact, as well as at specific
intervals; used to assess the structural integrity of pavement.
False Set: When freshly mixed Portland cement concrete prematurely develops rigidity
without gaining heat, which would indicate true setting. Plastiticy can be regained by
mixing without adding water. Other terms for this condition include premature
stiffening, hesitation set, early stiffening, and rubber set.
Fast-Track Concrete: Concrete produced using a series of techniques to accelerate con-
crete construction and reopen a roadway.
Faulting: When one section of concrete is raised or lowered to a level noticeably higher
or lower than an adjoining section on the other side of a joint or crack.
FHWA: Federal Highway Administration; part of the U.S. Department of Transportation.
Fiber Reinforcing: Fibrous material added to a concrete mix to increase strength.
Materials range from 0.05 to 4 mm in diameter and from 10 to 150 mm in length.
Polypropylene microfibers are the most common choice, but polyester and steel
fibers are also used.
Fibrous Concrete: Concrete containing dispersed, randomly oriented fibers.
Field-Cured Cylinder or Specimen: Samples of concrete poured into test cylinders and
cured in the same way as the roadway concrete so they can be used for testing for
strength. A field-cured cylinder can indicate when forms may be removed and when
the roadway can be reopened to traffic, as well as whether additional construction
load can be imposed.
Final Set: A degree of stiffening of a concrete mixture sufficient to resist to an established
degree the penetration of a weighted test needle. S ee also Initial Set.
Fine Aggregate: S ee Aggregate, Fine.
Fines: A general term referring to fine aggregate, mostly sand, or ultra fines such as dust,
fly ash, and cement.
Finishing of Concrete: Screeding, consolidating, floating, troweling, or texturing of
concrete to establish the final appearance and texture of a concrete surface. Most
highway finishing is performed by power-operated machines.
Fixed-Form Paving: Using fixed forms on the edges of a pavement to uniformly control the
edge and alignment. S ee also Slip-Form Paving.
GL OS SA RY 14

Flash Set: A rapid development of rigidity in freshly placed concrete, usually accompa-
nied by considerable heat, in which plasticity be regained by further mixing or the
addition of water. Also referred to as quick set or grab set.
Flexural Strength: The ability of concrete or reinforcement to resist bending, or to resist
failure while bending.
Float: A finishing hand tool made of wood, aluminum, or magnesium, used to impart a
relatively even but still-open texture to an unformed fresh concrete surface.
Flow or Flowability: The ability of freshly mixed concrete to seek out a level surface and
encapsulate reinforcing metal.
Fly Ash, or Flyash: A fine residue resulting from the combustion of ground or powdered
coal, used as a concrete admixture. Class F fly ash is typically created from the burn-
ing of harder, older anthracite and bituminous coal. It is pozzolanic in nature. Class C
fly ash is typically created from the burning of younger lignite or subbituminous coal.
It has pozzolanic properties, and also has some self-cementing properties: In the
presence of water, it will harden and gain strength over time.
Fresh Concrete: Concrete that has been recently mixed and is still workable and plastic.
Full-Depth Repair or Patch: Removal and replacement of a portion of a concrete slab to the
bottom of the concrete in order to restore areas of deterioration.

G
GGBF: Ground, granulated blast-furnace slag.
Gap-Graded Concrete: Concrete containing gap-graded aggregate. S ee Aggregate,
Gap-Graded.
Geocomposite Drains: Prefabricated drains installed in a sand-filled trench under the
shoulder; they are easier to install and less expensive than pipe drains.
Gravel: Granular material, natural rather than machine-crushed, predominantly retained
on a 4.75-mm (#4) sieve.
Green Concrete: Concrete that has set but has not hardened sufficiently for restoring
traffic.
Green Sawing: The process of sawing uniformly spaced joints in early-age concrete, in
order to prevent random cracking without tearing or dislocating the aggregate.
Grooving: Cutting slots into a concrete pavement to promote skid resistance and to create
channels along which water can escape beneath tires.
Ground, Granulated Blast-Furnace Slag: The non-metallic molten byproduct created by
steel production, consisting of silicates and aluminosilicates of calcium. When the slag
is still hot, it is granulated by chilling it in water, and it is then ground to less than 45
microns.
Grout: A mixture of cementitious material, with or without aggregate or admixtures, to
which sufficient water is added to produce a pouring or pumping consistency without
segregation of the constituent materials.

H
Hairline Cracking: Small-width cracks, often barely visible, often in a web-like pattern, in a
concrete surface. Hairline cracks are usually due to drying shrinkage, and are not deep.
GL OS SA RY 15

Hardening of Concrete: The process whereby compounds of Portland cement react with
mix water to form cementitious products that adhere to each other and to the inter-
mixed sand and stone particles and become very hard. If moisture remains present,
the reaction may continue for years, continually strengthening the concrete.
High-Early-Strength Cement: S ee Cement, High-Early-Strength.
High Range Water-Reducing Admixture: S ee Water-Reducing Admixture.
High-Shear Mixing: A mixing process using an impeller-type mixer that breaks up cement
ingredients in a cement slurry into their primary particles, using a higher level of
speed and power.
High-Strength Concrete: Concrete with a design strength of at least 70 MPa.
Hopper, Stationary: A container used to temporarily store freshly mixed concrete.
Hot-Pour Sealant: Joint-sealing materials such as PVC-coal tars, silicone sealants, polysul-
fides, and polyurethanes that require heating for installation.
Hydrated Lime: A dry powder obtained by treating quicklime with water to convert it to
calcium hydroxide, a mixture of calcium hydroxide and magnesium oxide, or magne-
sium hydroxide.
Hydration: The chemical reaction between hydraulic cement and water that produces new
compounds with strength-producing properties and so causes concrete to harden.
Hydroplaning: A dangerous condition when a vehicle goes out of steering control because
it skims the surface of a wet road.

I
Initial Set: A degree of stiffening of a mixture of cement and water less than final set,
generally stated as an empirical value indicating the time in hours and minutes
required for cement paste to stiffen sufficiently to resist to an established degree the
penetration of a weighted test needle; also applicable to concrete or mortar with use of
suitable test procedures. S ee also Final Set.
International Roughness Index (IRI): A method for evaluating pavement quality by quanti-
fying key distresses—transverse joint faulting, spalling, transverse fatigue cracking,
and smoothness—over time.
Isolation Joint: A cut along a straight line, partway through the concrete, to control
contraction cracking. Cracking that occurs due to movement of the slab is contained
in the joint, rather than being allowed to spread onto the pavement. A joint may be
cut in green concrete or in fully hardened concrete.

J
Joint: A cut along a straight line, partway through the concrete, to control contraction
cracking. Cracking that occurs due to movement of the slab is contained in the joint,
rather than being allowed to spread onto the pavement. A joint may be cut in green
concrete or in fully hardened concrete. Sometimes called an Isolation Joint.
Joint Deterioration: Concrete failure at or near a joint, most often spalling or D-cracking.
Joint Filler: Compressible material such as backer rod used to partway fill a joint to
prevent the infiltration of debris and to provide support for the joint sealant.
GL OS SA RY 16

Joint Sealant: Poured or injected material such as PVC-coal tar, silicone sealant, polysul-
fide, or polyurethane used to minimize water and solid-debris infiltration into the joint.
Jointed Plain Concrete Pavement (JPCP): Pavement with sufficient joints to control all
natural cracks expected in the concrete. In most JPCP, steel tie bars are installed at
longitudinal joints to prevent joint opening, and dowel bars are used to enhance load
transfer at transverse contraction joints.
Jointed Reinforced Concrete Pavement (JRCP): Pavement with joints and embedded steel-
mesh reinforcement to control expected cracks.

K
Kiln: In cement manufacture, a furnace that heats and chemically combines raw materials,
such as limestone, sand, and clay, into calcium silicate clinker.

L
Life-Cycle Cost (LCC): A means of analysis used to compare projects based on their initial
cost, future cost, and salvage value, in order to determine which projects and repairs
are most cost-effective in the long run.
Lift: A layer of concrete placed over or under another layer. S ee Two-Lift Paving.
Lightweight Aggregate: S ee Aggregate, Lightweight.
Lime: A general term that encompasses various chemical and physical forms of quick-
lime, hydrated lime, and hydraulic lime. Lime may be high-calcium, magnesian, or
dolomitic.
Liquid Sealant: Sealant materials that are heated and poured or injected in liquid form
and cool or cure to gain their final strength and adhesive properties.
Load-Transfer Assembly: A basket or carriage, used during Load-Transfer Restoration
(LTR), designed to support or link dowel bars at the correct height and position to
hold them in place during pouring of concrete.
Load-Transfer Device: S ee Dowel.
Load-Transfer Efficiency: The relative ability of a joint or crack to transfer part of a load
applied on the one side of the joint or crack to the other side of the joint or crack.
Load-Transfer Restoration (LTR): The process of cutting slots and installing smooth dowels
to transfer loads across a joint and so tie a slab together. Also called Dowel
Retrofitting.
Longitudinal Joint: A joint running lengthwise along a roadway.
Lot: A defined quantity.
Low-Shear Mixing: A mixing process using an impeller-type mixer that gently mixes
ingredients into a cement slurry, using a lower level of speed and power to limit
damage to the particles.

M
Map Cracking: A pattern of fine cracks often caused by improper finishing or curing too
quickly. Also called Hairline Cracks. S ee also Crazing and Pattern Cracking.
Mass Concrete: Concrete that is cast in place in volumes so large that the heat of hydra-
tion can cause volume change that could cause cracking.
GL OS SA RY 17

Membrane Curing: Application of a sealing compound, also called an evaporation reducer,


to create a membrane that slows evaporation rates and thus ensures slow curing. Some
membranes are sprayed on, including bituminous and paraffinic emulsions, coal tar
cut-backs, pigmented and non-pigmented resin suspensions, or suspensions of wax
and drying oil. Sheet plastics and waterproof paper can also be used as membranes..
Mesh Reinforcement: S ee Wire-Mesh Reinforcement.
Metakaolin: Highly reactive pozzolan made from kaolin clays.
Mineral Admixtures: Powdered or pulverized materials added to concrete before or during
mixing to improve or change some of the plastic or hardened properties of Portland
cement concrete. These materials are either natural or by-products of other manufac-
turing processes. S ee also Supplementary Cementitious Materials.
Mix Design: The selection of concrete mix ingredients and their proportions with the
intent of making the most economical use of available materials while still producing
concrete of sufficient strength.
Mixing Time: The length of time it takes for a mixer to combine ingredients for a batch of
concrete. For stationary mixers, this is measured in actual clock time. For truck
mixers, mix time is calculated in number of drum revolutions.
Mixing Water: The water added to concrete while mixing, not including any moisture
previously absorbed by the aggregate.
Modulus of Elasticity: The ratio of tensile stress to tensile strain, used in order to deter-
mine how elastic a material is; denoted by the symbol E. Also referred to as elastic
modulus, Young’s modulus, and Young’s modulus of elasticity.
Modulus of Rupture: A measure of the flexural strength and so the ultimate load-carrying
capacity of a structure, sometimes referred to as rupture modulus or rupture strength.
Moisture Content of Aggregate: The percentage of the weight of water in an aggregate mass
to the dry weight of the mass.

N
Non-Agitated Box or Truck: A container for transporting central-mixed concrete that does
not mix the concrete during transport.
Normal-Weight Aggregate: S ee Aggregate, Normal-Weight.
NRMCA: National Ready Mixed Concrete Association.

O
Oven-Dry Aggregates: Aggregates that have been kiln-dried at a temperature between 105°
and 115°C and so contain no water either in the pores or on the surfaces.
Overlay: The addition of a new layer of concrete or asphalt onto an existing pavement
surface. S ee also Resurfacing; Bonded Concrete Overlay; Unbonded Concrete Over-
lay; and White-Topping.

P
Particle Size Distribution: The division of aggregate particles by various sizes, expressed in
terms of cumulative percentages of various sizes. Particle size is determined by sieving.
Paste: A combination of cement and water, which is a constituent of a concrete mix.
GL OS SA RY 18

Pattern Cracking: Cracks on a concrete surface that form a weblike pattern. S ee also
Crazing, Hairline Cracking, and Map Cracking.
Pavement: A highway, road, street, or parking lot surfaced with concrete or asphalt. Also
the top surface of a highway, road, street, or parking lot.
Pavement Structure: The total structure of a highway, road, street, or parking lot, includ-
ing the subgrade, the base/subbase, and the surface course.
PCA: Portland Cement Association.
PCC: Portland Cement Concrete.
Pea Gravel: Gravel with round-edged particles that range from 0.5 to 1 cm in diameter.
Performance-Related Specification (PRS): A quantitative method of determining accep-
tance quality of concrete. Includes specifications for desired levels of key materials and
construction quality characteristics that have been found to predict future perfor-
mance. The characteristics can be tested at the time of concrete placement.
Permeability: In concrete construction, the property of allowing water to pass through.
Permeable Subbase: A subbase layer of crushed aggregates with a reduced amount of
fines to promote drainage; often stabilized with cement.
Pervious Concrete: Also called no-fines or porous concrete. Concrete that contains no or
few fines, so the voids between coarse aggregate particles are left unfilled. The coarse
aggregate particles are coated with a cement paste, so they bond together at contact
points. The resulting concrete has interconnected openings that allow storm water to
drain through the concrete to the subbase below.
pH: The chemical symbol for a measure of acidity and alkalinity, on a scale of 0 to 14; less
than 7 represents acidity, and more than 7 represents alkalinity.
Placing: The depositing, distribution, and consolidation of freshly mixed concrete in the
place where it is to harden; sometimes referred to as Pouring.
Plain Concrete: Concrete without reinforcement.
Plastic Consistency: A condition of freshly mixed concrete that makes it moldable and
workable. The wet concrete is cohesive, with an ample content of cement and fines,
but is not so wet as to be soupy.
Plasticity: A condition of concrete such that a certain amount of deformation will be
sustained without cracking; most often, concrete with slump of 80 to 100 mm.
Plasticizer: An admixture that increases the plasticity of Portland cement concrete.
Plastic Shrinkage Cracking: Surface cracks that result from rapid moisture loss on the
surface of wet concrete due to evaporation. Cracks are usually parallel and only a few
centimeters deep and a meter or so long.
Pneumatic: Powered or worked by air pressure.
Polymeric (or Polymer-Based) Concrete: Portland cement concrete to which a polymer is
added for improved durability and adhesion. Often used for patches and in overlays
for bridges.
Pop-Out: A pit on a concrete surface, often part of a grouping, that results from cracking
of the mortar; may be caused by poor finishing, overly absorbent aggregate, or a
contaminating material, such as wood or glass.
Porosity: The ratio of the volume of voids in a concrete slab to the total volume of the
slab, including voids.
Portland Cement: A calcium silicate cement produced by finely pulverizing clinker that was
produced by burning a defined mixture of iron, alumina, and silica at high temperature.
GL OS SA RY 19

Portland Cement Concrete: A composite material composed of a paste of Portland cement


and water, with embedded particles of fine and coarse aggregate.
Pouring: The deposit, spreading, and consolidation of freshly mixed concrete in the place
where it is to harden; generally referred to as Placing.
Pozzolan: A siliceous or siliceous-and-aluminous material such as fly ash or silica fume,
which itself is not cementitious but which will, if finely divided and in the presence of
moisture, chemically react with calcium hydroxide to form compounds possessing
cementitious properties.
Pozzolan-Cement Grout: A grout commonly used for slab stabilization, composed of
Portland cement, water, and pozzolan (usually fly ash). S ee Pozzolan.
Precast Concrete: Concrete cast in forms and cured in a controlled environment, then
transported to the placement location.
Preservation: Maintaining a roadway in its present condition and forestalling further
deterioration. S ee also Rehabilitation, Repair, and Restoration.
Prestressed Concrete: Concrete that has been subjected to induced compressive stresses
before loading, usually by means of high-strength steel tendons or bars. Prestressing
balances the tensile stresses that will occur during service.
Process Control: Steps taken to assure a high level of quality in the final concrete slab,
including sampling and testing.
PSI: Pounds per square inch; a measure of the compressive, tensile, or flexural strength.
Multiply by 0.0703 to convert PSI to kilograms per square centimeter.
Pumping: The outward displacement of subsurface soil and water that occurs when a
pavement is repeatedly depressed and released under heavy, high-speed loads, espe-
cially heavy trucks. Pumping indicates that the subsurface materials are too erod-
able—most often, fine-grained and plastic soils.
Punchout: A type of structural distress that occurs in continuously reinforced concrete
pavement, where a small area enclosed by transverse cracks, or a crack and the edge of
the pavement, exhibits spalling, shattering, or faulting.

Q
Quality Assurance (QA): Planned and systematic actions by an pavement owner or its
representative to provide confidence that a product or facility meet applicable stan-
dards of good practice. This involves continued evaluation of design, plan, and
specification development; contract advertisement and award; construction; mainte-
nance; and the interactions of these activities.
Quality Control (QC) Specifications: Statistically based specifications regarding materials
and methods, to which the contractor must adhere in order to achieve the agency’s
desired results.

R
Reactive Aggregate: Aggregate containing certain silica or carbonate compounds that can
react with alkalis in Portland cement and sometimes cause excessive expansion of
concrete.
GL OS SA RY 20

Ready-Mixed Concrete: Concrete delivered to a purchaser in a plastic and unhardened


state, usually in a ready-mix truck.
Rebar: Abbreviation for Reinforcing Bar. S ee Reinforced Concrete.
Reconstruction: Demolishing and removing an existing pavement and replacing it with a
completely new pavement.
Recycled Concrete: Hardened concrete that is processed for reuse as an aggregate compo-
nent of new concrete.
Rehabilitation: Repairing or modifying a structure to a desired condition. S ee also Preser-
vation, Repair, and Restoration.
Reinforced Concrete: Concrete strengthened with tensile bearing materials such as steel
reinforcing bars or metal wires, and fiber reinforcement.
Reinforcement, Transverse: Reinforcing bars placed at right angles to the longitudinal
reinforcement; it may be the main or secondary reinforcement.
Rejectable Quality Level (RQL): A lower level of contractor performance, producing
concrete that is unacceptable. S ee Acceptable Quality Level.
Relative Density: The ratio of a material’s total mass to water; used to determine concrete
mixture proportions. Also called Specific Gravity.
Release Agent: A material, usually liquid and sprayed on, used to prevent bonding of
concrete to a surface such as a form.
Repair: To replace or correct deteriorated or faulting portions of a concrete slab. S ee also
Preservation, Rehabilitation, and Restoration.
Reservoir, Joint: The channel opening in a concrete joint that is meant to be filled with a
sealant. Usually, a widening saw cut above the initial saw cut.
Restoration: Reestablishing the materials, form, and appearance of a roadway to the
original condition. S ee also Preservation, Rehabilitation, and Repair.
Resurfacing: The addition of a new concrete or asphalt layer onto an existing pavement
surface to improve smoothness or texture.
Retarder: An admixture that delays the setting of cement and so slows hardening of
concrete.
Retempering: The addition of water and remixing of concrete that has begun to harden
and has lost workability.
Retrofit Dowel Bars: In Load-Transfer Restoration (LTR), the dowels installed into slots
cut into the surface of an existing concrete pavement.
Rheology: The study of the way materials flow.
Rich Mixture: A concrete mixture containing a large amount of cement.
Rigid Pavement: Pavement that strongly resists bending and distributes loads over a
comparatively large area.
Rod, Tamping: A length of solid cylindrical metal, with one end rounded; used to compact
concrete. Also called rodding.
Rodability: The susceptibility of fresh concrete or mortar to compaction with a tamping
rod.
Roller-Compacted Concrete (RCC): A zero-slump concrete mix that is consolidated by
rolling with vibratory compactors. This method allows for rapid construction of
heavy-duty pavement.
Rolling Weight Deflectometer (RWD): A laser device that measures deflections created by
truck traffic.
GL OS SA RY 21

S
Sack of Cement: S ee Bag of Cement.
Saturated Surface-Dry: A condition of aggregate in which the permeable voids are filled
with water but there is no water on the particle surfaces.
Saw Blade, Diamond: A circular concrete-sawing blade with industrial diamonds as the
abrasion element. Blades must be continuously cooled with water to protect the metal
from overheating and dislodging the diamonds.
Scaling: Surface flaking on a hardened concrete surface, usually due to poor finishing,
repeated freeze-thaw cycles, or de-icing chemicals.
SCC: Self-Compacting Concrete, also called self-leveling or self-consolidating concrete.
Screed: To roughly level and strike off concrete to the desired height. Also refers to the
tool for striking off the concrete surface, sometimes referred to as a strike-off.
Screed Guide: Firmly anchored side forms, the top of which are at the desired finish height
of the slab, on which the screed or strike-off lies while it’s being used for screeding.
Sealant: S ee Joint Sealant.
Sealant Reservoir: S ee Reservoir, Joint.
Sealing: Filling the sawed joint with sealant material to minimize intrusion into the joint
of water and incompressible materials.
Segregation: The tendency, as concrete flows laterally, for concrete to separate into
constituents according to their size and weight. Typically, coarse aggregates are left
behind, and mortar and finer aggregates move ahead. Over-vibrating can also cause
segregation, in which mortar and wetter material rise to the top.
Self-Compacting Concrete (SCC): High-performance concrete that requires little or
no vibration or other methods of consolidation; also called self-leveling or self-
consolidating concrete.
Separation: The tendency of concrete, as it travels through passages such as chutes and
conveyers, for coarse aggregate to separate from the concrete and accumulate in one
area; also, the tendency for aggregates to separate from the water by gravitational
settlement. S ee also Bleed Water and Segregation.
Separation Layer: When installing an unbonded concrete overlay, the layer of asphalt or
other medium that separates the underlying slab from the overlay.
Set: The condition reached by concrete when it has lost plasticity and gained a certain
amount of hardness, so it resists penetration or deformation. Initial Set refers to first
stiffening. Final Set refers to attainment of significant rigidity. This early period in the
hardening is called the “setting period,” although there is no well-defined break in the
hardening process.
Setting Time: The time it takes for a specimen of concrete to attain a specified degree of
rigidity.
Settlement: Sinking of solid particles in fresh concrete, after placement and prior to initial
set. S ee also Bleed Water.
Shrinkage: A decrease in volume of concrete due to changes in moisture content, temper-
ature changes, or chemical changes.
Shrinkage Crack: A crack that results from shrinkage, usually because the slab was poured
in very cold or very hot conditions; usually occurs within the first few days after
placement.
GL OS SA RY 22

Sieve: Wire mesh or a metallic plate with regularly spaced openings used to separate
aggregate material according to size.
Silica Fume: Extremely fine noncrystalline silica used as a pozzolan in concrete.
Silicone: A resin with excellent water-repellent properties used in joint-sealing com-
pounds and concrete admixtures.
Shrinkage-Reducing Concrete: Concrete containing a shrinkage-reducing admixture
(SRA) that causes expansion during hardening and so offsets drying shrinkage.
Silo: Any conical or round-shaped storage vessel designed for holding cement or aggregate.
Slab Jacking: Injecting grout, through holes drilled in the slab, under a slab to raise it and
so lessen depressions and correct faulting.
Slab Stabilization: Injecting grout, through holes drilled in the slab, under a slab to
stabilize it and lessen the possibility of cracking.
Slag Cement: Hydraulic cement consisting mostly of ground blast-furnace slag, some-
times mixed with Portland cement or hydrated lime. Slag concrete is concrete that
contains slag cement.
Slip-Form Paving: A highway paving process that extrudes concrete through a large
machine and uses a slip-form, which is placed temporarily and removed as the
concrete is placed, obviating the need to install edge forms.
Slump: A measurement of the amount that a sample of freshly mixed concrete subsides
after it is removed from a mold.
Slump Cone: A conical mold, with a base diameter of 203 mm, top diameter 102 mm, and
height 305 mm, used to conduct a slump test on freshly mixed concrete.
Slump Loss: The measurable tendency of freshly mixed concrete to lose slump (i.e.,
become firmer) after the time when a slump test has been performed.
Slump Test: The procedure for measuring slump.
Soil Cement: A mixture of soil, Portland cement, and water, sometimes used as a base
material under pavements; also called cement-stabilized soil.
Spalling: Cracking, chipping, or fraying, which may occur in defined areas on the surface
or near a joint.
Specific Gravity: The ratio of a material’s total mass to water; used to determine concrete
mixture proportions. Also called Relative Density.
Spud Vibrator: A small vibrator used for consolidating small areas of concrete, especially
in a patch.
Standard Deviation: A probability distribution of variance; the expected deviation of
individual values from their average.
Stratification: The unwanted separation of concrete into horizontal layers, with the
heavier materials at the bottom; usually caused by over-wet concrete or by
over-vibrating.
Strike Off: S ee Screed.
S tructural C ap acity : The measurable ability of a pavement to carry traffic loads; in
AASHTO design methodology, structural capacity is expressed in terms of number of
axle loads.
Subbase: A layer in a pavement system between the subgrade and the base course, or
between the subgrade and a Portland cement concrete pavement.
Subgrade: The soil prepared and compacted to support a structure or a pavement system;
also sometimes called grade.
GL OS SA RY 23

Sulfate Attack: An attack on a concrete slab by sulfates in groundwater or soil; can result
in expansion and disintegration of the concrete.
Sulfate Resistance: The ability of concrete paste and aggregate to withstand chemical attack.
Superplasticizer: A water-reducing admixture that results in a mix requiring 12% to 30%
less water. Also called a Plasticizer.
Supplementary Cementitious Material (SCM): Cementitious materials often added to a
concrete mix along with Portland cement, including fly ash and blast furnace slag.
Also called Bulk Supplements.
Surface Moisture: Water present on the surfaces of aggregates. Surface moisture is capable
of mixing with Portland cement. Also called surface water.

T
Tensile Strength: The amount of stress that a material can support without fracture when
being stretched along its axis.
Ternary Mixture: A concrete mixture using three (3) cementitious materials.
Thermal Expansion or Stress: Expansion caused by a rise in temperature.
Tie Bar: A metal bar tied at right angles to the primary reinforcement to keep it in place.
Also a deformed bar spanning across a joint to prevent separation of the adjoining slabs.
Time of Set: The amount of time needed for concrete to attain a certain degree of hard-
ness or strength.
Transit Mixing: The process of doing all or most of the concrete mixing in a delivery truck
drum. Also called truck mixing.
Transverse Crack: A crack running more or less at a right angle to the roadway.
Transverse Joint: A joint running at a right angle to the roadway.
TRB: Transportation Research Board.
Two-Lift Paving: A paving method in which two (2) layers of concrete are laid down,
rather than a single homogenous layer. The lower layer can be composed of aggregates
of lesser quality than those in the upper layer.

U
UBC: Uniform Building Code.
Ultra-Thin White-Topping: A layer of new 5- to 10-cm-thick concrete, placed over a
prepared surface of moderately distressed asphalt. The concrete is often, but not
necessarily, high-strength fiber-reinforced concrete.
Unbonded Concrete Overlay: A layer of new concrete placed onto a separation layer of
asphalt or other separating medium, which has in turn been applied over a distressed
concrete pavement.
Unit Weight: S ee Bulk Density and Specific Gravity.
Unreinforced Concrete: S ee Plain Concrete.

V
Vibrating Screed: A power tool that vibrates as it screeds the concrete surface.
Vibration: Mechanical agitation of freshly mixed concrete to remove air bubbles and
enhance consolidation and compaction.
GL OS SA RY 24

Volume Batching: Measuring of the constituent materials for concrete by volume.

W
Water-to–Cementitious Materials Ratio: The ratio of the amount of water (not including
water absorbed by aggregates) to the amount of cementing materials (not including
aggregates).
Water-Reducing Admixture: A material that allows for using less water in a concrete mix; it
either increases slump without increasing water content or else produces greater
workability with a reduced amount of water. Also called a water reducer.
Weight-Batching: Measuring the constituent materials for concrete by weight.
White-Topping: A overlay pavement made of high-strength fiber-reinforced concrete,
placed over an existing asphalt pavement. Conventional white-topping is 10 cm or
thicker; ultra-thin white-topping is 5 to 10 cm thick.
Wire-Mesh Reinforcement: A concrete reinforcement composed of wire that has been
welded at joints, creating a grid of consistently spaced wires. Also called welded wire
fabric.
Workability: The ease with which freshly mixed concrete can be mixed, placed, molded,
and finished.

XYZ
Yield: The volume per batch of concrete, given in cubic meters.
STANDARDS ORGANIZATIONS
AND PROFESSIONAL
ASSOCIATIONS

American Association American Concrete Pavement American National Standards


of State Highway & Association Institute
Transportation Officials Chicago Office: Washington, DC,
444 N. Capitol Street NW, 5420 Old Orchard Road, Headquarters:
Suite 249 Suite A-100 1819 L Street, NW, 6th Floor
Washington, DC 20001 Skokie, IL 60077-1059 Washington, DC 20036
phone 202.624.5800 phone 847.966.2272 phone 202.293.8020
fax 202.624.5806 fax 847.966.9970 fax 202.293.9287
http://www.transportation.org Washington Office: New York City
http://www.aashto.org Office—Operations:
500 New Jersey Avenue NW, 25 W. 43rd Street, 4th Floor
7th Floor New York, NY 10036
American Ceramic Society Washington, DC 20001 phone 212.642.4900
600 N. Cleveland Avenue, phone 202.638.2272 fax 212.398.0023
Suite 210 fax 202.638.2688 http://www.ansi.org
Westerville, OH 43082 http://www.pavement.com
phone 866.721.3322
fax 301.206.9789 American Petroleum
American Concrete Pipe Institute
http://www.acers.org Association 1220 L Street, NW
1303 W. Walnut Hill Lane, Suite 305 Washington, DC 20005-4070
American Coal Ash Irving, TX 75038-3008 phone 202.682.8000
Association phone 972.506.7216 http://www.api.org
15200 E. Girard Avenue, fax 972.506.7682
Suite 3050 http://www.concrete-pipe.org
Aurora, CO 80014-3955 American Shotcrete
phone 720.870.7897 Association
American Concrete Pressure 38800 Country Club Drive
fax 720.870.7889 Pipe Association
http://www.acaa-usa.org Farmington Hills, MI 48331
3900 University Drive, Suite 110 phone 248.848.3780
Fairfax, VA 22030-2513 fax 248.848.3740
American Concrete phone 703.273.7227 http://www.shotcrete.org
Institute fax 703.273.7230
38800 Country Club Drive http://www.acppa.org
Farmington Hills, MI 48331
phone 248.848.3700 American Concrete Pumping
fax 248.848.3701 Association
http://www.aci-int.org 606 Enterprise Drive
Lewis Center, OH 43035
phone 614.431.5618
fax 614.431.6944
http://www.concretepumpers.com

25
STA N D A R D S O R GANI Z AT I ONS AND P ROF E S S I ONAL AS S OCI AT I ONS 26

American Society British Cement Association Canadian Standards Association


of Civil Engineers Minerals Products Association 5060 Spectrum Way, Suite 100
ASCE World Headquarters Limited Mississauga, ON
1801 Alexander Bell Drive Riverside House, Canada L4W 5N6
Reston, VA 20191-4400 4 Meadows Business Park phone 416.747.4000
phone 800.548.2723 Station Approach, fax 416.747.2473
fax 703.295.6333 Blackwater, Camberley http://www.csa.ca
ASCE Washington Office Surrey, GU17 9AB
101 Constitution Avenue NW, United Kingdom Cast Stone Institute
Suite 375 East phone 01276.608700 813 Chestnut Street
Washington, DC 20001 fax 01276.608701 Lebanon, PA 17042
phone 800.548.2723 http://www.cementindustry.co.uk phone 717.272.3744
fax 202.789.7859 fax 717.272.5147
http://www.asce.org British Standards Institution http://www.caststone.org
389 Chiswick High Road
American Society London W4 4AL Cement Association of Canada
of Concrete Contractors United Kingdom 502-350 Sparks Street
2025 S. Brentwood Boulevard, phone 20.8996.9001 Ottawa, ON
Suite 105 fax 20.8996.7001 Canada K1R 7S8
St. Louis, MO 63144 http://www.bsi-global.com phone 613.236.9471
phone 866.788.2722 fax 613.563.4498
fax 314.968.4367 Building and Fire Research http://www.cement.ca
http://www.ascconline.org Laboratory
NIST Cement Concrete
American Society 100 Bureau Drive, Stop 8600 & Aggregates Australia
of Heating, Refrigerating, Gaithersburg, MD 20899-8600 Level 6, 504 Pacific Highway
and Air-Conditioning Engineers phone 301.975.5900 St Leonards NSW
1791 Tullie Circle NE fax 301.975.4032 Australia 2065
Atlanta, GA 30329 http://www.bfrl.nist.gov postal address: Locked Bag 2010
phone 404.636.8400 St Leonards NSW 1590
fax 404.321.5478 Building Research Establishment phone 2.9437.9711
http://www.ashrae.org Bucknalls Lane fax 2.9437.9470
Watford WD25 9XX http://www.concrete.net.au
American Society United Kingdom
for Testing and Materials phone 1923.664000 Center for Advanced
100 Barr Harbor Drive http://www.bre.co.uk Cement-Based Materials
West Conshohocken, PA Northwestern University
19428-2959 Bundesverband der Deutschen McCormick School of Engineering
phone 610.832.9500 Zementindustrie e.V. and Applied Science
http://www.astm.org Kochstraße 6-7 2145 Sheridan Road
D-10969 Berlin Evanston, IL 60208-4400
Architectural Precast Germany phone 847.491.3858
Association phone 30.280.02.0 fax 847.467.1078
6710 Winkler Road, fax 30.280.02.250 http://acbm.northwestern.edu
Suite 8 http://www.bdzement.de
Fort Myers, FL 33919 Concrete Corrosion Inhibitors
phone 239.454.6989 Canadian Society for Civil Foundation
fax 239.454.6787 Engineering phone 301.340.7368
http://www.archprecast.org 4920 de Maisonneuve Boulevard W. http://www.corrosioninhibitors.org
Suite 201
Montréal, QC Concrete Foundations Association
Canada H3Z 1N1 113 W. First Street
phone 514.933.2634 Mt. Vernon, IA 52314
fax 514.933.3504 phone 319.895.6940
http://www.csce.ca fax 320.213.5556
http://www.cfana.org
STA N D A R D S O R GANI Z AT I ONS AND P ROF E S S I ONAL AS S OCI AT I ONS 27

Concrete Reinforcing Steel European Ready Mixed Concrete Interlocking Concrete


Institute Association Pavement Institute
933 N. Plum Grove Road Boulevard du Souverain, 68 United States:
Schaumburg, IL 60173-4758 1170 Brussels 13921 Park Center Road, Suite 270
phone 847.517.1200 Belgium Herndon VA 20171
fax 847.517.1206 phone 2.6455212 phone 703.657.6900
http://www.crsi.org http://www.ermco.eu fax 703.657.6901
Canada:
Concrete Sawing and Drilling Expanded Shale, Clay and Slate P.O. Box 85040
Association Institute 561 Brant Street
11001 Danka Way North, Suite 1 230 E. Ohio Street, Suite 400 Burlington, ON
St. Petersburg, FL 33716 Chicago, IL 60611 Canada L7R 4K2
phone 727.577.5004 phone 801.272.7070 http://www.icpi.org
fax 727.577.5012 fax 312.644.8557
http://www.csda.org http://www.escsi.org International Center for
Aggregates Research
Construction Specifications Federal Highway Administration The University of Texas at Austin
Institute U.S. Department of Transportation PRC 18B
110 S. Union Street Suite 100 1200 New Jersey Ave SE 10100 Burnet Road
Alexandria, VA 22314-3351 Washington, DC 20590 Austin, TX 78758
phone 800.689.2900 phone 202.366.4000 phone 512.471.4527
fax 703.684.8436 http://www.fhwa.dot.gov http://www.icar.utexas.edu
http://www.csinet.org
Innovative Pavement Research International Code Council
Deutsches Institut für Normung e.V. Foundation 500 New Jersey Avenue, NW,
(German Standards Institute) 5420 Old Orchard Road, 6th Floor
Burggrafenstraße 6 Suite A-100 Washington, DC 20001
10787 Berlin Skokie, IL 60077-1059 phone 888.422.7233
Germany http://www.iprf.org fax 202.783.2348
phone 30.2601.0 http://www.iccsafe.org
fax 30.2601.1231 Institute of Electrical &
http://www.din.de Electronics Engineers International Concrete Repair
2001 L Street NW, Suite 700 Institute
Electric Power Research Institute Washington, DC 20036-4910 3166 S. River Road, Suite 132
3420 Hillview Avenue phone 202.785.0017 Des Plaines, IL 60018
Palo Alto, CA 94304 fax 202.785.0835 phone 847.827.0830
phone 650.855.2000 http://www.ieee.org fax 847.827.0832
http://www.epri.com http://www.icri.org
Instituto Mexicano del Cemento
European Cement Association y del Concreto, A.C. International Grooving
(CEMBUREAU) Av. Insurgentes Sur # 1846 & Grinding Association
Rue d’Arlon 55 Col. Florida, C.P. 01030 12573 Route 9W
BE-1040 Brussels México, D.F. West Coxsackie, NY 12192
Belgium phone 5322.5740 phone 518.731.7450
phone 2.234.10.11 http://www.imcyc.com fax 518.731.7490
fax 2.230.47.20 http://www.igga.net
http://www.cembureau.be Insulating Concrete Forms
Association International Organization
European Committee for 1298 Cronson Boulevard, for Standardization (ISO)
Standardization Suite 201 1, ch. de la Voie-Creuse
Avenue Marnix 17 Crofton, MD 21114 Case postale 56
B-1000 Brussels phone 888.864.4232 CH-1211 Geneva 20,
Belgium fax: 410.451.8343 Switzerland
phone 2.550.08.11 http://www.forms.org phone 22.749.01.11
fax 2.550.08.19 fax 22.733.34.30
http://www.cen.eu http://www.iso.ch
STA N D A R D S O R GANI Z AT I ONS AND P ROF E S S I ONAL AS S OCI AT I ONS 28

International Union National Concrete National Research Council Canada


of Laboratories and Experts Masonry Association 1200 Montreal Road,
in Construction Materials, 13750 Sunrise Valley Drive Building M-58
Systems and Structures Herndon, VA 20171-4662 Ottawa, ON
RILEM phone 703.713.1900 Canada K1A 0R6
157 rue des Blains fax 703.713.1910 phone 613.993.9101
F-92220 Bagneux http://www.ncma.org fax 613.952.9907
France http://www.nrc.ca
phone 1.45.36.10.20 National Concrete Pavement
fax 1.45.36.63.20 Technology Center National Slag Association
http://www.rilem.net CP Tech Center P.O. Box 1197
2711 South Loop Drive, Pleasant Grove, UT 84062
Japan Concrete Institute Suite 4700 phone 801.785.4535
Sogo Hanzomon Building 12F Ames, IA 50010 fax 801.785.4539
1-7 Kojimachi, Chiyoda-ku phone 515.294.5798 http://www.nationalslag.org
Tokyo 102-0083 fax 515.294.0467
Japan http://www.cptechcenter.org National Stone, Sand & Gravel
phone 3.3263.1571 Association
fax 3.3263.2115 National Institute 1605 King Street
http://www.jci-net.or.jp of Building Sciences Alexandria, VA 22314
1090 Vermont Avenue NW, phone 703.525.8788
Long Term Pavement Suite 700 http://www.nssga.org
Performance (USDOT) Washington, DC 20005
Turner-Fairbank Highway phone 202.289.7800 National Terrazzo & Mosaic
Research Center fax 202.289.1092 Association
6300 Georgetown Pike http://www.nibs.org P.O. Box 2605
McLean, VA 22101 Fredericksburg, TX 78624
phone 202.493.3022 National Institute of Standards phone 800.323.9736
http://www.tfhrc.gov and Technology fax 888-362-2770
100 Bureau Drive, Stop 1070 http://www.ntma.com
Materials Research Society Gaithersburg, MD 20899-1070
506 Keystone Drive phone 301.975.6478 Operative Plasterers’ and Cement
Warrendale, PA 15086-7537 http://www.nist.gov Masons’ International Association
phone 724.779.3003 11720 Beltsville Drive, Suite 700
fax 724.779.8313 National Precast Concrete Beltsville, MD 20705
http://www.mrs.org Association phone 301.623.1000
1320 City Center Drive, Suite 200 http://www.opcmia.org
National Association Carmel, IN 46032
of Corrosion Engineers phone 317.571.9500 Occupational Safety and Health
1440 South Creek Drive fax 317.571.0041 Administration
Houston, TX 77084-4906 http://www.precast.org U.S. Department of Labor – OSHA
phone 281.228.6200 200 Constitution Avenue
fax 281.228.6300 National Ready Mixed Concrete Washington, DC 20210
http://www.nace.org Association phone 800.321.6742
900 Spring Street http://www.osha.gov
National Association Silver Spring, MD 20910
of Home Builders phone 240.485.1141 Perlite Institute
1201 15th Street NW http://www.nrmca.org 4305 N. Sixth Street, Suite A
Washington, DC 20005 Harrisburg, PA 17110
phone 202.266.8200 National Research Council phone 717.238.9723
fax 202.266.8400 500 Fifth Street NW fax 717.238.9985
http://www.nahb.com Washington, DC 20001 http://www.perlite.org
phone 202.334.2138
http://www.nas.edu/nrc
STA N D A R D S O R GANI Z AT I ONS AND P ROF E S S I ONAL AS S OCI AT I ONS 29

Portland Cement Association Slag Cement Association U.S. Army Corps of Engineers
Chicago Office: 2516 Waukegan Road, Suite 349 441 G Street NW
5420 Old Orchard Road Glenview, IL 60025 Washington, DC 20314-1000
Skokie, IL 60077 phone 847.977.6920 phone 202.761.0011
phone 847.966.6200 http://www.slagcement.org http://www.usace.army.mil
fax 847.966.8389
Washington Office: Tilt-Up Concrete Association Verein Deutscher Zementwerke e.V.
500 New Jersey Avenue NW, P.O. Box 204 (German Cement Works
7th floor Mt. Vernon, IA 52314 Association)
Washington, DC 20001 phone 319.895.6911 P.O. Box 301063
phone 202.408.9494 http://www.tilt-up.org 40410 Düsseldorf
fax 202.408.0877 Germany
http://www.cement.org Transportation Association phone 211.4578.1
of Canada fax 211.4578.296
Post-Tensioning Institute 2323 St. Laurent Boulevard http://www.vdz-online.de
38800 Country Club Drive Ottawa ON
Farmington Hills, MI 48331 Canada K1G 4J8 The Vermiculite Association
phone 248.848.3180 phone 613.736.1350 Whitegate Acre, Metheringham Fen
fax 248.848.3181 fax 613.736.1395 Lincoln LN4 3AL
http://www.post-tensioning.org http://www.tac-atc.ca United Kingdom
phone 1526.323990
Transportation Research Board fax 1526.323181
Precast/Prestressed Concrete
Institute The National Academies http://www.vermiculite.org
200 W. Adams Street, #2100 500 Fifth Street NW
Chicago, IL 60606 Washington, DC 20001 Wire Reinforcement Institute
phone 312.786.0300 phone 202.334.2934 942 Main Street, Suite 300
http://www.pci.org http://www.trb.org Hartford, CT 06103
phone 800.552.4974
Silica Fume Association U.S. Bureau of Reclamation – fax 860.808.3009
38860 Sierra Lane Materials and Engineering http://www.wirereinforcementinsti-
Lovettsville, VA 20180 Research Lab tute.org
phone 540.822.9455 1849 C Street NW
fax 540.822.9456 Washington, DC 20240-0001
http://www.silicafume.org http://www.usbr.gov/pmts/materials_lab/
SECTION 1
U.S. AND EUROPEAN
CONCRETE INDUSTRY
TRANSPORTATION
INFRASTRUCTURE
PHILOSOPHY
OF CONCRETE

CONCRETE IS A MATERIAL OF SUPERLATIVE QUALITIES. It is a material with end-


less variations, which can be adapted to provide outstanding performance in almost any
combination of natural and artificial environments. Because of its unrivaled versatility,
concrete is the most widely used building material in the world. It is used to build every-
thing from low-strength drainage courses to bridges and dams demanding the highest
possible strength.
In the transportation infrastructure, concrete is omnipresent, with uses ranging from
simple driveways and sidewalks to long-life, high-performance roads. This is due in part
to concrete’s versatility, but it is also due to its quality. While the versatility of concrete is
inherent, the quality must be designed into the entire manufacturing process. Quality
pavements can be produced only if maximum cost/benefit ratio is kept in mind. The pro-
cess begins with careful attention to desired final characteristics and a logical approach to
mix design. Concrete can be made to be alkali- or acid-resistant; it can be pervious to
facilitate drainage; or it can have low permeability to resist de-icing chemicals—but only
if it is designed that way from the beginning. The mix must accommodate the needs of the
application: The cement must be appropriate. The aggregates must be clean, sound, and
well graded. During manufacture, batching must be accurate, and mixing must be done
well. Once the concrete is ready, well-thought-out transport and proper placement are
crucial. A quality-control plan that takes all of the above into account is vital if concrete is
to have the longest life cycle at the possible lowest cost.
While a haphazard approach to making and placing concrete results in a cheaper mix-
ture, it will not result in a long-lasting concrete. Quality materials and careful planning
cost more, of course, but they do not cost proportionally more. Experience and research
have both shown that well-graded, clean aggregates, appropriate cements and admixtures,
and proper curing can more than double the life span of a roadway without doubling the
cost. Moreover, producing quality concrete ends up freeing resources to produce more
pavement, and thus more income, and more quality highways.
The nature of concrete is something that has changed over time. The natural pozzola-
nic concrete that the Romans used to build the Pantheon and the aqueducts has lasted for
31
P HI L OS OP HY OF C ONC R ET E 32

millennia, but it is not the Portland cement concrete that we need today to build our dams
and roads. The demands on concrete have changed, the ingredients have changed with
them, and mix designs have become more and more sophisticated. Concrete will continue
to evolve, especially in the face of environmental concerns. We are likely to see the use of
more recycled aggregates, of more materials like ground blast-furnace slag, fly ash, and
even rice hulls to improve the quality of concrete while taking the strain off landfills. Alter-
native concretes that generate less CO2 during manufacture and placement are already
being developed.
Producing the best product, however, will always pay for itself. The smart manufactur-
ers and the efficient contractors will do more than embrace the conventional approach.
They will stay informed of the research, of the development of new materials, of new tech-
niques, and of new equipment. While it will always be cheaper in the short run to use
tried-and-true methods, those methods will not always be the best, the most efficient, or
the least costly. In the long run, the ability to make quality concrete will always result in
better-made structures. This benefits the customer, and in the long run, whatever benefits
the customer will always benefit the supplier.
HISTORY OF
CONCRETE PAVEMENT

VARIOUS FORMS OF WHAT WE NOW CALL CONCRETE have existed for centuries,
going back as far as ancient Rome. For the purposes of this report, we will concentrate on
the concrete mixture known as Portland cement and its evolution in road construction
technology in Europe and North America.
The advent of Portland cement as a building material is relatively new. In 1824, Joseph
Aspdin, a bricklayer from Leeds, England, developed and patented the first mixture that is
known as Portland cement. The process involved burning limestone down to lime, then
mixing the residual lime with clay and pulverizing the mixture wet. This wet mixture was
then dried, crushed, and calcined in a vertical kiln. (Calcination is the process of heating
the mixture to break down calcium carbonate [limestone] into calcium dioxide [lime] and
carbon dioxide.) The resulting calcine was powdered. Neither the exact proportions of
lime and clay nor the burning temperature was stipulated in the original patent. The end
product was named “Portland” cement because the set cement looked similar to the lime-
stone found in Portland, England.
At that time, natural cement was more prevalent and cheaper to produce, which hin-
dered the use of Portland cement in the building industry. Its growth as a building mate-
rial stagnated.
In 1891, in an effort to promote Portland cement as a paving material, George W. Bar-
tholomew constructed the first section of concrete pavement in the United States. This
2.4-meter-wide strip, located in Bellefontaine, Ohio, was evidence of cement’s ability to
provide a hard, durable paved surface. The success of this experiment yielded the expan-
sion of concrete paved streets throughout the entire town. In 1894, this project was com-
pleted by William T. G. Snyder, a Bellefontaine road builder who is considered America’s
first concrete paving contactor.
The introduction of the mass-produced automobile in 1908 propelled concrete paving
into a nationwide necessity due to increased demands by the public for better roads. The
first “highway” was a 23-mile (37-km) stretch of pavement near Pine Bluff, Arkansas,
called the “Dollarway” because it cost $1 per linear foot (30 cm) to build. By 1914, about
2,350 miles (3,780 km) in the United States had been paved with Portland cement.
33
HI S TORY OF C ONC RE T E PAV E M ENT 34

President Woodrow Wilson signed the first “Federal-Aid Highway Act” in 1916. This
act mandated that the U.S. federal government would finance up to 50% of each state’s
road construction costs. However, in order to receive the federal funding, each state would
be responsible for preparing plans and specifications as well as construction and mainte-
nance. In turn, the federal highway department had the right to inspect any federally
funded roadways built to ensure that they were constructed properly and made of quality
materials [Weingroff, FHWA, 2009].
During the 1930s, especially in northern states, where the freeze-thaw cycles were
extreme, roadways began to show significant crumbling on the surface, called scaling.
Northern roads were also treated with de-icing salts to remove ice and snow from pave-
ments. Osmosis caused the moisture to move toward areas with a higher salt concentra-
tion, increasing the pressure on the pores, causing fissures. And because de-icing salts
increased cooling rates, the concrete was more susceptible to surface cooling, which fur-
ther accelerated cracking and scaling.
To solve these problems, air entrainment was introduced into the cement-manufac-
turing process. This process involves creating intentional air voids in the cement that allow
migrating water to enter. These entrained air voids within the cement reduce the pressure
to the surface pores and virtually eliminate damage to the pavement surface.
It was around this time (the 1930s) that heavy truck traffic caused another problem,
called “pumping.” At this point in time, roads were normally constructed on soil from
which the topsoil had been removed. Heavy truck loads combined with excess moisture in
the soil would cause soil and water to “pump” out the undersides of the slab. This would
weaken the support structure beneath the slab and cause cracking. To remedy this, a layer
of gravel, crushed stone, or slag was used as a subbase—as opposed to merely the underly-
ing soil—which eliminated this problem.
In the 1930s and 1940s, pavement with thickened edge sections was relatively com-
mon. For instance, pavement would be 6 inches (152 mm) thick in the center and 8 inches
(203 mm) thick along the edges—referred to as an 8-6-8 design. Heavier truck traffic man-
dated that pavements be constructed to uniform thicknesses and increased to 9–10 inches
(229–254 mm) for highways. These thicker, uniform pavements are currently used to
ensure the longevity of heavily used highways [Delatte, 2008, pp. 9–10].
The biggest hindrance to expanding the road system was the fixed forms that the con-
crete was poured into. Positioning the forms required considerable manpower before the
cement could be poured. Because of this, it took construction crews an entire day to lay just
300 meters of concrete. Slip-form pavers solved the problem and revolutionized paving
technology. A slip-form paver spreads the concrete between short forms built into each
side of the machine, strikes the slab level, and finishes the surface as the machine drives
forward along the subbase. (Lower-slump concrete is normally used with slip-form pavers
because the fresh concrete is able to hold its shape once the paver has passed.)
In 1955, Iowa’s Quad City Construction Company developed an improved track-
propelled paver that could place 7.3-meter-wide slabs 250 millimeters thick. Two (2)
HI S TORY OF C ONC RE T E PAV E M ENT 35

pavers working side-by-side could lay a country road in a single pass [Delatte, 2008, p. 6].
The introduction of the slip-form paver reduced the number of workers needed on a road
crew from 100 to about 25. Today’s slip-form pavers can lay pavement up to 48 cm thick in
widths ranging from 3.6 to 4.5 meters, at a rate of a 1.6 km or more per day.
Another timesaving improvement was the introduction of central-mixed concrete.
Large central mixers were able to mix 5.8 to 8.3 cubic meters of concrete in 45 to 75 sec-
onds—drastically reducing preparation time. This central-mixed concrete could then be
hauled from the mixer to the slip-form paver in non-agitating dump trucks without affect-
ing the quality.

Concrete Pavement Durability Research


All of these advancements led to construction of the U.S. Interstate Highway System begin-
ning in 1956. Highway engineers also recognized the need to study more efficient methods
of placing and maintaining concrete roads.
Also accelerating the rate of pavement construction was the use of sawn joints to
replace hand-tooled joints. Not only speeding installation, these sawn joints also improved
pavement smoothness [Delatte, 2008, p. 6].
Between 1958 and 1960, the American Association of State Highway Officials (which
later became the American Association of State Highway and Transportation Officials, or
AASHTO) conducted tests to study the effects of varying vehicle weight loads (truck ver-
sus car weight loads) on roads of varying thicknesses. The test road, constructed in Ottawa,
Illinois, consisted of 7 miles of two-lane (2-lane) pavements closed to the public. Half of
the road was made of concrete, and the other half asphalt. The test roads incorporated a
wide range of surface, base, and subbase thicknesses plus 16 short-span bridges. Tests
involved riding cars and trucks of various specific weights over the track. Researchers
would then study the specific pavement designs for defects.
Considered by some industry experts as yielding the best information ever developed
on pavement technology, this report revealed the importance of properly graded granular
subbases [Pasko, 1998].
The study experimented with placing steel dowels across pavement joints. The testers
discovered that when the joints were placed at no more than 4.6-meter intervals, these
dowels noticeably reduced pavement cracking by transferring traffic loads from one
pavement slab to the next. These methods are still used in present day road paving [Hud-
son, 2006].
As the United States interstate system grew, myriad advances were made in concrete
paving technology. During the 1960s, electronic controls were added to slip-form pavers
to increase efficiency and uniformity. Subgrade trimmers, which level the compacted sub-
grade soil, created improved grade control and pavement stability. Additionally, tied con-
crete shoulders, in which shoulders are connected to the pavement via tie bars, reduced
stress between the outside lane and the shoulder [Pasko, 1998].
HI S TORY OF C ONC RE T E PAV E M ENT 36

Research and development of concrete paving continued throughout that time in an


effort to improve methods of laying and maintaining concrete pavement. The U.S. Inter-
state Highway System was completed in the late 1970s and early 1980s. By this time, how-
ever, many roads had reached their lifespan limits and needed repair.
In 1976, the U.S. Congress allocated federal funding for the “3R program”—Restora-
tion, Rehabilitation, and Resurfacing of paved roads. In 1981, a fourth “R,” Reconstruc-
tion, was added to the plan. Not only were innovative methods and materials needed to
restore the roads, these repairs needed to be done relatively quickly to alleviate costs due
to traffic congestion.

Resurfacing, Restoring, and Rehabilitating


The Federal-Aid Highway Act of 1976 authorized $175 million per year for 1978 and 1979,
“for resurfacing, restoring, and rehabilitating those lanes on the Interstate System which
have been in use for more than five years and which are not on toll roads” [Federal-Aid
Highway Act of 1976]. Because the federal government would contribute 90% of the
funds, it greatly reduced the financial repair burden to each state and would significantly
improve the worn existing roads.
At this time, efforts toward finding new repair technologies were at the forefront of the
industry. The introduction of “fast-track” concrete drastically cut curing times from 5 to
14 days to as little as 12 hours or less—allowing roads to be opened to traffic more quickly.
This concrete was formulated either from a higher proportion of Type I cement to water
or from high-early-strength (Type III) cement.
Besides the obvious advantages of fast-track cement’s quicker curing times, this
cement mixture creates very strong, long-lasting pavements due to the fact that they are
air-entrained and have a relatively low water content. These factors not only improve the
pavement’s strength, they reduce the chloride or salt permeability that can damage steel
reinforcements and thus lead to road deterioration. These and other cement types will be
discussed later.
Roller-compacted concrete (RCC) is a relatively new technology that enables con-
struction crews to quickly lay industrial or heavy-duty pavement. This method makes use
of high-density asphalt equipment to lay a drier-mix concrete that is subsequently com-
pacted with rollers. The resulting pavement does not contain joints, dowels, or reinforce-
ment and it does not require forms or finishing—yet it withstands heavy-duty traffic. This
economical paving construction method is commonly used when strength, durability, and
economy are most critically needed. These areas include port, intermodal, and military
facilities; parking, storage, and staging areas; intersections; and low-speed roads [Port-
land Cement Association].
Another such innovative repair technique is dowel retrofitting in existing undowelled
slab joints. This technology involves cutting slots across pavement joints, inserting the
dowel bars, then patching the existing slots with fast-track cement. The roads can then be
HI S TORY OF C ONC RE T E PAV E M ENT 37

diamond-ground to achieve a smooth surface. In fact, this particular technique, initially


used by Washington’s Department of Transportation (D.O.T.), was expected to extend the
state’s 30-year-old concrete highways by 10 to 15 years. For more on retrofitting dowels,
see the Concrete Repair chapter.
A common problem of the steel dowels is the likelihood of corrosion over time—espe-
cially in areas where snow and ice require the use of de-icing salts. The dowels can corrode
to the point of failure, which could lead to cracking in the pavement. The corrosion can
also cause the dowel to “lock” into the pavement, preventing movement during expansion
and contraction—which can also lead to slab cracking. In the 1970s, many states man-
dated the use of epoxy-coated or other materials to resist this corrosion. Since then, sev-
eral alternate materials have been proposed [Tritsch, April 2009]:

N Fiber-reinforced polymer (FRP)


N FRP composite tube with grout
N Plastic-coated dowels
N Solid stainless steel
N Stainless steel–clad
N Stainless steel tube filled with grout

Concrete overlays were also part of the restoration initiative. Bonded, unbonded, and
white-topping were less expensive than removing and rebuilding existing roadways.
Bonded overlays, or BCOs, “consist of a new concrete overlay placed directly on top of an
existing concrete pavement. This overlay bonds to the existing concrete to create a mono-
lithic slab.” [Illinois D.O.T., 2005, p. 1]
An unbonded concrete overlay, essentially, “is a new concrete pavement constructed
over an existing concrete pavement. A flexible interlayer, typically constructed of hot-mix
asphalt (HMA), separates the concrete layers, allowing the concrete layers to move inde-
pendently of each other, thereby preventing reflective cracking in the concrete overlay.”
[Illinois D.O.T., 2005, p. 1]
White-topping “consists of a concrete overlay, or inlay, placed on an existing HMA
surface pavement.” Two (2) types are available: thin white-topping, which is 10 to 18 cm
thick, and ultra-thin white-topping, which is 5 to 10 cm thick [Illinois D.O.T., 2005, p. 1].
More on white-topping and overlays can be found in the Concrete Repair chapter.
One of the most recent road tests that began in 1987 and continues today is the
Federal Highway Administration (FHWA) long-term pavement performance program
(L.T.P.P.). Part of the strategic highway research program, this test involved rigorous field
tests of more than 2,400 sections of 800-km asphalt and concrete pavement across the
United States and Canada. From this program, an extensive number of research reports as
well as design and analysis tools were created. One of the most critical was the develop-
ment and calibration of models for the A.A.S.H.T.O. Mechanistic-Empirical Pavement
Design Guide [Delatte, p. 10].
HI S TORY OF C ONC RE T E PAV E M ENT 38

The FHWA compiled all of this L.T.P.P. data into a user-friendly database of informa-
tion called DataPave Online (www.ltpp-products.com/). This online tool includes infor-
mation on inventory, materials testing, pavement performance monitoring, climatic,
traffic, maintenance, rehabilitation, and seasonal testing data. Other tools include L.T.T.P.
Pavement Online, a rigid pavement design software system that automates the pavement
design and analysis processes (www.ltpp-products.com/rigid/index.aspx) [Delatte, p. 11].
As concern for the environment grew, concrete was found to be sufficiently “green.”
Concrete roads create less resistance and reportedly provide better fuel efficiency for vehi-
cles and have two (2) times the lifespan of asphalt roads. Additionally, concrete can be
recycled and reused as either an aggregate or a subbase layer [APCA, Reconstructing,
2009]. More on concrete and the environment can be found in the chapter Concrete Eco-
logical Positioning.

Recent Developments in Europe


Several other European countries have tested the use of concrete as a pavement material.
These countries have utilized information based on U.S. research and have adopted their
own methods for concrete paving.
Germany began building concrete roads in the late 1880s. In 1934, concrete pave-
ment was used extensively in the construction of its roadways. Between 1935 and 1939,
3,500 km of roadways were built of wire-mesh-reinforced concrete. Until the early 1960s,
Germany built primarily jointed reinforced concrete pavements on unbonded base
courses. The standard jointed plain concrete pavement design was changed in 1972, elimi-
nating expansion joints in favor of transverse contraction joints on cement-treated or
asphalt-treated base. In 1982, the first slip-form paving was done on German motorways
[FHWA, long-life concrete pavements, 2007].
The Avenue de Lorraine in Brussels, originally constructed in 1925, remained in ser-
vice until 2003, when it was overlaid with concrete. The original pavement of this road was
only 15 cm thick. In fact, Brussels has several examples of concrete roads that have served
traffic for 50 years or more. Problems over the years with slab cracking because of exces-
sive length, as well as joint faulting, led to many modifications in Belgium’s concrete pave-
ment design. It wasn’t until 1960 that Belgium built its first continuously reinforced
concrete pavement. Today, this country’s roads are constructed of jointed plain concrete
pavements with dowels in the transverse joints, and a rigid base layer [FHWA, long-life
concrete pavements, 2007].
In the 1950s, concrete pavements built on the Netherlands’ motorway system were
undowelled jointed plain concrete pavement. In the 1960s, dowelling transverse joints in
the pavement became standard practice. Almost all new concrete pavements in the Neth-
erlands have been built as continuously reinforced concrete pavement [FHWA, long-life
concrete pavements, 2007].
HI S TORY OF C ONC RE T E PAV E M ENT 39

E n d n o te s
C o n c r e t e P a v em en t s : H i gh w a y s . Portland Cement Association. Skokie, Illinois, 2009. <www.
cement.org/pavements/pv_cp_highways.asp>

Delatte, Norbert. C o n c r e t e P a v em en t D e s i g n, C o ns t r u c t i o n a n d P er f o r m a n c e. New York: Taylor &


Francis Publishers, 2008.

Hudson, Dr. W. R. “Duplicating the Test,” R o a d s & B r i d ge s magazine, Vol. 44 No. 6. Arlington
Heights, Illinois, Scranton Gillette Communications, June 2006. <http://roadsbridges.com/
Duplicating-The-Test-article7063>

1 0 0 Y e a r s o f I n n o v a t i o n. ACPA – American Concrete Pavement Association. Skokie, Illinois, 2009.


<www.pavement.com/Concrete_Pavement/About_Concrete/100_Years_of_Innovation/
index.asp>

L o n g L- i f e C o n c r e t e P a v em en t s i n E u r o p e a n d C a n a d a , Ch a p t er 5 : C em en t a n d C o n c r e t e.
U.S. Dept. of Transportation – Federal Highway Administration (FHWA), 2007. <http://
international.fhwa.dot.gov/pubs/pl07027/llcp_07_05.cfm>

Pasko, Thomas J., Jr. “Concrete Pavements—Past, Present and Future,” adapted from the S ix t h
I n t er n a t i o n a l P u r d u e C o n f er en c e o n C o n c r e t e P a v em en t D e s i g n a n d M a t er i a l s f o r H i gh
P e r f o r m a n c e, N o v . 1 8 –2 1 , 1 9 9 7 . 1998. <www.tfhrc.gov/pubrds/julaug98/concrete.htm>

P a v em en t T e c h n o l o g y A d v i s o r y —U n b o n d e d C o n c r e t e O v er l a y . M a i n t en a n c e R e p a i r a n d R e h a b i l i t a -
t i o n—P T A -M 2 , E ff. 1 0 / 1 9 9 6 , R e v . 0 2 / 2 0 0 5 . Springfield, Illinois: Illinois DOT, Bureau of
Materials and Physical Research, 2006. <www.dot.il.gov/materials/research/pdf/ptam2.pdf>

P a v em en t T e c h n o l o g y A d v i s o r y —B o n d e d C o n c r e t e O v er l a y . M a i n t en a n c e R e p a i r a n d R e h a b i l i t a -
t i o n—P T A -M 3 , E ff. 1 0 / 1 9 9 6 , R e v . 0 2 / 2 0 0 5 . Springfield, Illinois: Illinois DOT, Bureau of
Materials and Physical Research, 2006. <www.dot.il.gov/materials/research/pdf/ptam3.pdf>

P a v em en t T e c h n o l o g y A d v i s o r y —W h i t e t o p p i n g . M a i n t en a n c e R e p a i r a n d R e h a b i l i t a t i o n — P T A - M 4 ,
E ff. 1 0 / 1 9 9 6 , R e v . 0 2 / 2 0 0 5 . Springfield, Illinois: Illinois DOT, Bureau of Materials and
Physical Research, 2006. <www.dot.il.gov/materials/research/pdf/ptam4.pdf>

R e c o n s t r u c t i n g a n d R e c y c l i n g . ACPA – American Concrete Pavement Association. Skokie, Illinois,


2009. <www.pavement.com/Concrete_Pavement/Technical/Fundamentals/Reconstructing_
and_Recycling.asp>

Tritsch, Steven P., W h a t ’ s N e w i n L o a d T r a ns f e r . National Concrete Consortium Meeting, San


Antonio, Texas, April 2009. <www.cptechcenter.org/t2/documents/TritschLoadTransferNC-
CApr09.pdf>

Weingroff, Richard F. F e d er a l A i d R o a d A c t o f 1 9 1 6 : B u i l d i n g t h e F o u n d a t i o n. U.S. Dept. of


Transportation – Federal Highway Administration (FHWA), 2009. <www.fhwa.dot.gov/
infrastructure/rw96a.cfm>
SECTION 2

CONCRETE
PRACTICE
AGGREGATES

AGGREGATE IS THE PARTICULATE MATTER that makes up most of the volume in


concrete, and can include sand, natural gravel, crushed rock, slag, and ashes. Together,
these constituents typically make up some 80% of mixes used for roadways, and yet they
often are considered an inert filler with little effect on the finished product’s mechanical
properties. In fact, aggregates can increase strength, durability, stability, and workability.
Additionally, recycled aggregates are increasingly used to both reduce expense and realize
substantial environmental benefits.
Not all aggregates work equally in a mix, but because they are such a major constitu-
ent, their quality is crucial. The material must be free of chemicals and coatings that could
interfere with the hydration and bond of the cement paste. Friable aggregates are apt to
break up when mixed, requiring more water and adversely affecting durability and wear
resistance. Impurities are known to interfere with hydration and hardening [KOSMATKA,
KERKHOFF, AND PANARESE, 2008, PP. 79–80].
Certain impurities and reactive minerals, such as chalcedony, flint, and dolomite, can
cause surface chipping known as pop-outs, leaving a shallow depression 25 mm to 400
mm in diameter and half as deep.
Pop-outs typically will appear within
Q u ie t e r Pa v e me n t s
a year of placing the pavement, often
after a stretch of wet weather. Impu-
rities, reactive minerals, and the Aggregates can have a role in making quieter pavements, a
move that is already well underway in Europe, according to the
damage they cause should be avoided U.S. Federal Highway Administration. By reducing the amount
in road paving at all costs [ACP of fine aggregate, so that the coarse materials are bound by
INTERNATIONAL, WHAT’S A POP- less cement, the resulting pervious concrete has voids that
reduce road noise because less air is trapped between tires
OUT? 2001–2007].
and the road surface. Also, these pavements permit water to
Aggregates are defined as either drain with relatively little of the environmental impact associ-
fine or coarse, depending on particle ated with highways, a particularly important consideration
where there are wetland setbacks [FHWA, QUIET PAVEMENTS:
size. Coarse aggregates generally
LESSONS LEARNED FROM EUROPE, 2005].
range between 9.5 mm and 37.5 mm
in diameter. Gravels constitute the
41
AGGRE GAT ES 42

majority of coarse aggregate used in concrete. Fine aggregates (or fines) consist of natural
sand or crushed stone, with most particles passing through a 9.5-mm sieve. Aggregates
also are distinguished by whether they are natural (that is, used as dug or dredged from
surface mines or the beds of rivers and lakes); processed (crushed); or manufactured or
secondary (including recycled concrete and fly ash).

Recycled Concrete
Paving or improving highways and bridges involves dealing with the old material. That
typically has meant trucking demolition waste to a landfill, but concrete can be crushed
for use as an aggregate in many applications. Recycling concrete not only spares landfills
but also reduces the need to mine or dredge natural aggregates. Further savings can be had
by using portable recycling machinery at the highway site so that there is no need to trans-
port the demolished material to a distant plant for crushing [CONCRETE NETWORK, RECY-
CLING CONCRETE, 1999–2009].
New concrete made with recycled aggregate has several desirable characteristics
beyond costing less and reducing the burden on landfills: good workability, enhanced
durability, and resistance to freeze-thaw cycles.
That said, the compressive strength of new concrete made with recycled aggregate is
variable, depending on the original strength of the material and the water/cement ratio
used in the new mix. But overall, recycled content can be expected to conform to the same
standards applied to conventional aggregates. There is a limit to how much natural crushed
coarse aggregate can be replaced by recycled aggregates, at the risk of compromising
mechanical properties; the maximum level has been put at around 30%. At higher levels,
there is a risk of increased shrinkage with drying, although strength and freeze-thaw resis-
tance aren’t significantly compromised [PORTLAND CEMENT ASSN., MATERIALS: AGGRE-
GATES, 2009].
Recycled aggregate will typically have greater absorption and lower specific gravity
than natural aggregate and can be expected to produce concrete with higher drying
shrinkage and creep. These differences become more noticeable with increasing amounts
of recycled fine aggregates. High levels of recycled fines can also make the mix difficult to
work [PORTLAND CEMENT ASSN., MATERIALS: AGGREGATES, 2009].
Because of this increased absorption, it is recommended that recycled aggregates be
prewetted and used in an almost saturated surface-dry condition (see “Practical Experi-
ence with Recycled Concrete,” below). The Federal Highway Administration recommends
that the recycled concrete in a base layer should be in a saturated state to aid in the migra-
tion of fines throughout the mix. Also, steel wheel rollers are advised for compacting the
recycled material if steel reinforcements left in the rubble might interfere with rubber-
tired equipment.
It’s necessary to increase the percentage of cement when replacing natural aggregates
with recycled material because of the crushed concrete’s mechanical properties and shape.
AGGRE GAT ES 43

Eu r o p e a n Pr a c t ic e

Around the world, relatively few consistent regula- including recycled concrete. The material is washed
tions have been drafted with the use of recycled and graded in several sizes as needed for the lay-
aggregates in mind. Consequently, these aggre- ers of two-lift (2-lift) construction. Although reactive
gates often are subjected to the same standards as aggregates are used throughout Western Europe,
natural aggregates and are apt to perform at levels potential disadvantages are avoided by making
that make them unavailable for use in concrete for- sure to use mixtures that will yield strong, dense,
mulations [THE MASTERBUILDER, RECYCLED AGGREGATES and highly durable highways. In particular, Europe-
IN CONCRETE, 2004–2009]. ans generally aim for a higher rating of flexural
strength in the top layer of two-lift construction than
In Europe, however, standards for the use of recy-
their U.S. counterparts do.
cled aggregates have been established under EN
12620, Aggregates for Concrete. The effect of these According to Long-Life Concrete Pavements in
codes has been to encourage greater use of recy- Europe and Canada, a study done for the U.S. Fed-
cled and secondary aggregates in the construction eral Highway Administration, two-lift highways are
industry. The standards include benchmarks in- now the norm in Germany, and projects there now
tended to help to ensure that these materials will often employ a geotextile layer between the lifts to
be of a quality needed for particular applications. keep the concrete slab from bonding to its cement-
The regulations for aggregates used in road con- treated base. The study points out that the layer is
struction are more stringent than those for materials thicker than the standard material favored in the
intended for building construction. Among the United States, and porous enough to allow the
guidelines is a limit on the loss of mass in freeze- freshly poured concrete to permeate it, creating a
thaw testing; a maximum of allowable lightweight dependable bond and yet maintaining separation
organic contaminants; and standards for shape from the base.
and flakiness. Requirements for polished stone
Another difference in practice is the European con-
value were set, with the value for surfaces with
cern with road-generated noise. Highways there
exposed aggregate set slightly higher than that for
are more often constructed with exposed aggregate
conventional road surfaces. Also, standards were
surfacing, using durable, high-quality materials in
established for minimizing Alkali-Silica Reaction
the top layer of the concrete. Besides creating a
[FHWA, LONG-LIFE CONCRETE PAVEMENTS IN EUROPE AND
quieter surface, this aggregate is also credited with
CANADA, 2007].
improving durability and providing good friction for
Relative to practice in the United States, European vehicles [FHWA, LONG-LIFE CONCRETE PAVEMENTS IN
countries take considerable care in choosing a high EUROPE AND CANADA, 2007].
quality of aggregates for highway foundation layers,

Alternatively, water-reducing admixtures can be used in order to yield workability, slump,


and a water/cement ratio similar to that of conventional concrete. Water-reducing admix-
tures are additives that will give the required results with less water in the mix, and they
yield greater strength without increasing the amount of cement [CONCRETE NETWORK,
WATER REDUCTION IN THE MIX, 1999–2009]. (For more information, see the chapter
Chemical Admixtures.)
Other potential concerns are the effects of road salt, asphalt, soil, glass, wood, and
other contaminants that may find their way into the reused material. And old concrete
can’t simply be ground up without removing reinforcing steel.
AGGRE GAT ES 44

Despite these precautions, 38 states now use recycled concrete as an aggregate base,
while 11 add it to new concrete. In Europe, many government specifications require that
all new concrete contain some amount of recycled concrete aggregates. It’s expected that
the use of recycled concrete will increase in coming years, as the supply of natural
aggregates dwindles and environmental restrictions on mining become tighter [BENNETT,
2002, P. 79].

Pr a c t ic a l Exp e r ie n c e w it h Re c y c le d Co n c r e t e

At least four (4) states in the U.S. are making exten- aggregate when used as a base and subbase mate-
sive use of recycled concrete. rial. This is attributed to the cementitious action
that can still occur within the compacted base,
In Texas, the state’s Department of Transportation
adding higher supporting strength for the highway
has been using recycled concrete in highways and
[FHWA, SUMMARY OF MICHIGAN RECYCLED CONCRETE AGGRE-
streets for a number of years, and it notes engi-
GATE REVIEW, 2003].
neering, economic, and environmental benefits.
Reusing this material eliminates the need to truck The Minnesota Department of Transportation uses
solid waste to landfills as well as to haul aggregate almost 100% of the concrete removed from its
from quarries. In Houston, for example, all the con- pavements as dense-graded aggregate base [FHWA,
crete rubble generated from highway construction RECYCLED CONCRETE STUDY IDENTIFIES CURRENT USES,
is being recycled for new pavements. Contractors BEST PRACTICES, 2004].
initially had problems with mix workability because
In California, Caltrans’ specifications allow the use
it was difficult to maintain a consistent, uniform
of recycled concrete in pavement supporting layers.
saturated surface-dry condition. The remedy in
The state reports that although processing costs for
Texas was to water stockpiles of the recycled con-
the material may be higher than those for new
crete and frequently test aggregate for moisture
aggregate, savings can be had when the recycling
content [FHWA, SUMMARY OF TEXAS RECYCLED CONCRETE
plants are near project areas, because of reduced
AGGREGATE REVIEW, 2003].
hauling and overhead costs. Less obviously, there is
The Michigan Department of Transportation also the advantage of less wear and tear on high-
confirms that recycled concrete can provide per- ways from loaded trucks [FHWA, SUMMARY OF CALIFORNIA
formance comparable to or better than virgin RECYCLED CONCRETE AGGREGATE REVIEW, 2003].

F ly A sh
When coal is burned in power plants, a great deal of noncombustible fly ash is left behind
as a waste product that, until recently, was mostly trucked off to landfills. That has changed
with the realization that not only can fly ash displace some of the cement needed in mixes,
but it also has the potential to improve the quality of finished concrete in a number of
ways. It acts as a pozzolan, meaning that when combined with lime (calcium hydroxide)
and water, it forms cementitious compounds. Fly ash even facilitates pumping concrete,
because it takes the form of tiny spheres—something like ball bearings—to allow the mate-
rial to flow more smoothly.
In the United States, fly ash concrete was put to the test back in 1929 during construc-
tion of the Hoover Dam. Since then, its usefulness has become more fully appreciated,
AGGRE GAT ES 45

particularly in terms of environmental impact. Because it makes it possible to use less


cement, fly ash reduces the amount of carbon dioxide released into the atmosphere both
by the kilns used to make cement and by the chemical reaction through which cement is
made. The resulting benefits are substantial, because carbon dioxide emissions from
cement production make up a considerable share of emissions from all types of fossil-fuel
combustion. Another environmental benefit is that by taking the place of some cement
needed in a mix, fly ash reduces the impact of mining for fine aggregate [HANLE, 2006].
As for the physical properties of concrete containing fly ash, it can be stronger, more
durable, less vulnerable to shock, and less subject to bleeding and segregation (in which
concrete separates into constituents according to their size and specific weight). Alkali-
Silica Reaction (ASR) is reduced and sulfate resistance is improved. Also, because of the
ball-bearing effect of fly ash and the resulting improved flow, the concrete is relatively
impermeable and less subject to cracking [GHOSH, 1992, P. 461].
Fly ash carries two basic classifications, depending on how it was generated. Type F, a
byproduct of burning anthracite and bituminous coal, is considered preferable. Type C,
made when burning sub-bituminous coal and lignite, can cause calcification when
mixed with water. (For further information, see the chapter Supplementary Cementitious
Materials/Bulk Supplements.)
Using fly ash is not without special considerations. Air-entrained concrete has micro-
scopic air cells that minimize the potential damaging effect of freeze-thaw cycles, and
these voids may not be adequately entrained in a mix that contains fly ash with particu-
larly high levels of carbon. Air-entrainment properties also may be further limited if the
material is overworked, leading to pocking and scaling. Because fly ash varies in its level
of pozzolanic activity, the proportions of a mix might have to be altered to accommodate
it. Note also that certain types of fly ash may have considerable levels of heavy metals such
as mercury, selenium, and lead, among other toxic compounds. Finally, a potential limita-
tion is that desirable sources of fly ash can be far from the construction site, so that the
expense of hauling would cancel out any economic advantage [PORTLAND CEMENT ASSN.,
MATERIALS: AGGREGATES, 2009].

Recycled Rubber
It’s recognized that tires have potential as a recycled material, with particles of rubber tak-
ing the place of a portion of solid aggregate in concrete mixes. The resulting product has
been found to be especially tough, but with reduced strength as the percentage of rubber
goes up. The challenge is that rubber particles are highly compressible and may not bond
well within the cement matrix [HUANG, LI, PANG, AND EGGERS, 2004, PP. 187–194].
Nevertheless, it’s possible that granulated tires may prove useful as an elastic aggre-
gate, one that compensates for the brittleness of concrete so that it can handle greater
stress prior to failure [GOULIAS AND ALI, 2004].
AGGRE GAT ES 46

Potentially H armful M aterials


Concrete may fail to perform up to expectations if any of several types of foreign matter
find their way into the mix:

N Organic impurities can delay setting and hardening, and eventually reduce
strength and cause deterioration.
N Impurities such as peat, humus, and organic loam may interfere with hydration
and solidifying.
N Harder materials, including coal, lignite, and wood, have been found to disinte-
grate, cause pop-outs, or lead to staining.
N Lumps of clay tend to absorb mixing water, eventually causing pop-outs and
compromising durability and wear resistance. The lumps can be difficult to detect
when dry because they tend to look a lot like dust-coated aggregate, but when
wetted they reveal themselves by beginning to disintegrate. (Limestone dust, on the
other hand, has been observed to potentially increase the strength of concrete.)
N Iron oxide and iron sulfide are known to cause stains on exposed surfaces.
N Chlorides, found on aggregate from seaside locations or recycled concrete that has
absorbed road salt, can cause steel reinforcement to corrode [SMITH, COLLIS, AND
FOOKES, 2001, P. 213].

Al k ali-S ilica Reaction


Alkali-Silica Reaction, or ASR, is a chemical process that can degrade concrete. When the
silicas in aggregates react with the alkali hydroxides in cement, a gel is formed that absorbs
water and potentially causes expansion and cracking, with serious consquences for struc-
tural integrity.
ASR is less apt to be a problem with lowered levels of both alkali metal in the cement
and also the reactive silica content in the aggregate. Volcanic aggregates tend to be espe-
cially apt to cause ASR. By adding calcium carbonate aggregates in the form of limestone,
ASR can be prevented, but they themselves may cause problems if they contain a high level
of silica. It’s also possible to neutralize the alkalinity of the cement with fine siliceous mate-
rials, such as fly ash or metakaolin.
Another potentially damaging chemical reaction, Alkali-Carbonate Reaction (or
ACR), can take place in the presence of certain dolomitic rocks, leading to expansion and
deterioration. ACR is relatively uncommon because the aggregates that lead to the prob-
lem are usually considered unsuitable for use in concrete [KOSMATKA, KERKHOFF, AND
PANARESE, 2008, P. 95].
AGGRE GAT ES 47

W ater I mpurities
Water is yet another ingredient that can introduce impurities to the mix, influencing set-
ting time, concrete strength, and durability and causing efflorescence or staining. Of par-
ticular concern are chlorides, sulfates, alkalis, and solids. (For more information, see the
chapter Water for Concrete Production.)

W aste G lass as a S and S ubstitute


Crushed and screened waste glass, or cullet, has potential as a substitute for sand in con-
crete, including types of glass that are not considered suitable for bottle recycling. But glass
aggregate can cause an Alkali-Silica Reaction between the cement and the glass, eventually
compromising strength and durability. This is especially problematic when using fine
cullet. More research is necessary to determine whether additives can arrest this reaction
and prevent the concrete from degrading over time [TEXAS D.O.T., USE OF GLASS CULLET
IN ROADWAY CONSTRUCTION, 2009].

Characteristics of A ggregates
Because aggregates can so fundamentally affect the mixing and performance of concrete,
several of their characteristics need to be considered carefully.

GRADING. Grading affects the proportions of aggregate, cement, and water needed for
maximum durability. Aggregates with a good range of particle sizes will generally produce
concrete with fewer voids between parti-
Figure 3-1
cles. This has an economic benefit, because
as voids increase in volume, more cement Id e a l C o mb in a t io n o f Ag g r e g a t e G r a d e s

is needed, and cement is more expensive 25

than aggregate. If there is not enough


20
sand to fill the voids between coarse
Percent retained

aggregate particles, the space must be


15
filled with cement paste. These under-
sanded mixes also tend to be harsh and 10
difficult to finish. On the other hand,
aggregate combinations with excessive 5

amounts of sand or extremely fine-grained


0
sand may prove uneconomical because 37.5 19.0 9.5 2.36 0.600 0.150 0.045
the finer particles have more surface area: 25.0 12.5 4.75 1.18 0.300 0.075

As surface area increases, more cement is Sieve size in mm

required. If the cement content remained Source: Kosmatka, Steven H., Beatrix Kerkhoff, and William C. Panarese.
the same, the thinner layers of cement Design and Control of Concrete Mixtures, 14th Edition (2002).
Skokie, Illinois: Portland Cement Assn., 2008.
surrounding the aggregate would make a
AGGRE GAT ES 48

stiff concrete that’s harder to use. The paste could be made more workable by adding water,
but both strength and durability would be compromised.
Surface area and economy are also affected by the maximum size of coarse aggregate
used. In most cases, increasing the maximum size of well-graded coarse aggregate causes
a decrease in the amount of paste that’s required to produce concrete of a desired consis-
tency [PORTLAND CONCRETE ASSN., AGGREGATES FOR CONCRETE, 1999].
As general rule of thumb, the grading of fine aggregate will influence the workability
of concrete more than the grading of the coarse material. If the water/cement ratio is opti-
mal, a wide range in grading will work well.
There are five (5) types of gradation [FERGUSON, 2005, P. 204]:

N DENSE GRADATION includes particles of different sizes to allow the smaller


particles to fill in the void between the bigger particles, resulting in low porosity
and low permeability of the aggregate mass.
N GAP-GRADED aggregates include a range of two (2) or more sizes while omit-
ting intermediate sizes. There is a notable difference between European concrete
specifications and those that apply in the United States: The U.S. specs allow gap-
graded aggregate concrete for many purposes even though it is inferior to a well-
graded aggregate concrete that utilizes a linear traverse gradation of fines to
coarse aggregate. The allowance for making concrete with only two main sizes of
aggregate (a fine and a coarse material) can be seen as contrary to good practice,
in that workability, pumpability, finish, and durability all suffer. Instead, the fines
should be especially well graded to ensure placement and finishing. While gap-
graded aggregates have been said to yield compressive strengths equal to or higher
than with well-graded aggregate, they require higher levels of fines, which in turn
require higher levels of cement—and often more water to bring about the desired
strength.
N OPEN-GRADED involves either a single size or a narrow range of sizes. Conse-
quently, the voids between particles tend not to be filled by smaller particles.
N COARSE-GRADED is similar to dense grading, except that the particles are
predominantly coarse.
N FINE-GRADED also can be considered a form of dense grading in which the
particles are mostly fine.

PARTICLE SHAPE AND TEXTURE. Both the shape and surface texture of an aggregate
affect the production of concrete in several ways.
Shape can be considered in terms of sphericity and roundness. Sphericity describes
how compact the particle is. A particle that approaches the shape of a sphere or a cube,
rather than being disk-like or elongated, is higher in sphericity. Roundness describes the
sharpness or angularity of the particle. Higher sphericity translates to less surface area,
which is significant in that less water and sand will be required for good workability.
AGGRE GAT ES 49

Th e Pr o b le m w it h G a p G r a d in g

In the U.S. there prevails an allowance for gap- the higher water/cement ratios or the lesser flexural
graded aggregate concrete in many instances. and elasticity of the gap-graded concrete.
Gap-graded concrete is inferior to a well-graded
In looking closely at the life cycle of gap-graded
(multi-sized) aggregate concrete, which utilizes a lin-
concrete, the cycles are much reduced as com-
ear traverse gradation of fines to coarse aggregate.
pared, where possible, to similar well-graded
This is one of the most notable differences between
concrete that includes the missing intermediate
North American concrete and European concrete
sizes of coarse material.
specifications. In the U.S. the allowance for making
concrete with only two (2) main sizes of aggregate— Recently, certain asphalt paving equipment manu-
a fine and a coarse material—seems at odds with all facturers have entered the mainstream concrete
good practice and science where gradation of mate- production market in the U.S. These manufactur-
rials for quality concrete is concerned. The workability, ers have been inquiring of many agencies and
pumpability, finish, and durability of gap-graded specification writers as to why it is allowable in con-
concrete all suffer. Such material requires that the crete production to utilize gap grading, but not so
fines be especially well graded to affect any good in quality asphalt, and not so in concrete produced
attributes with regard to placement and finishing. in other jurisdictions such as Europe and Asia. It
The argument for the allowance of gap-graded would seem a valid question that requires further
aggregates would seem to sit with the fact that com- investigation. In looking at how these gap-graded
pressive strengths can be attained that in many aggregate mixes came into allowance, the cost or
cases are higher than with well-graded aggregate. most likely the rate of aggregate production was the
Unfortunately, gap-graded aggregates require higher reason gap-gradation was first experimented with
amounts of fines, which in turn require higher in the U.S. While gap-graded mixes have been
amounts of cement and often higher amounts of shown to make strengths and perform to a point,
water to affect the desired results with regard to those who are obviously in search of the highest
strengths. Most unfortunately, the durability of such possible quality in concrete production and meth-
mix designs is not forthcoming, most likely due to odology should avoid gap grading if possible.

Round aggregates, however, lead to concrete with little aggregate interlock; these require a
cement paste with a higher tensile strength, leading to higher cement contents.
Surface texture describes the degree of roughness or irregularity of aggregate particles.
Smooth particles require less mixing water and less cementitious material [KOSMATKA,
KERKHOFF, AND PANARESE, 2008, P. 87].
Whether they take the form of sand or coarse aggregate, these rougher mixtures gen-
erally tend to yield concrete that is difficult to work. The shape and grading of sand, in par-
ticular, has a considerable effect on the water requirement and workability. Also, these
particles may break during compaction, compromising strength. In consequence, this class
of particles is either avoided or limited to about 15% by weight of the total aggregate.
Mixes that use aggregates that are coarse, angular, elongated, or lamellar have larger
voids and will require more water and fine aggregate, meaning greater expense. The ideal
aggregate shape is a multi-faceted cube (also known as polyhedral) shape, which offers
reduced surface area for coating, but ensures excellent aggregate interlock. These aggre-
gates produce concrete that is easy to pump, work, and finish, while keeping cement con-
AGGRE GAT ES 50

tent to a minimum. These aggregates are most easily manufactured using impact (hammer
mill) crushers, as opposed to cone crushers.

RESISTANCE TO FREEZING AND THAWING. A number of variables determine resistance


to freezing and thawing: an aggregate’s porosity, absorption, permeability, pore structure,
and size. A particle may absorb so much water that it cannot accommodate the expansion
and hydraulic pressure that occur during the freezing, and a single particle near the sur-
face of the concrete can cause a pop-out. If enough particles are involved, the aggregate
may expand and cause the concrete to disintegrate. Coarse aggregate with relatively high
porosity values and medium-sized pores is the most easily saturated; consequently, it is
especially likely to cause pop-outs and general deterioration. Larger pores do not usually
become saturated or cause concrete distress, and water in very fine pores may not freeze
readily [KOSMATKA, KERKHOFF, AND PANARESE, 2008, P. 93].
Freezing and thawing cycles can initiate cracking in saturated aggregate at the bot-
tom of the slab, continuing up to the visible wearing surface. At first, the cracks are
closely spaced and parallel to joints in the concrete, and they eventually spread toward
the center of the pavement [PORTLAND CEMENT ASSN., DURABILITY: FREEZE-THAW
RESISTANCE, 2009].

WETTING AND DRYING PROPERTIES. Alternating wetting and drying can affect the dura-
bility of aggregates. Their expansion and contraction coefficients vary with moisture con-
tent, and over time certain types of aggregate can permanently increase in volume, causing
pop-outs and the eventual deterioration of concrete. Especially vulnerable are clay lumps
and other friable particles when exposed to repeated wetting and drying.

ABRASION AND SKID RESISTANCE. When concrete pavement will be constantly sub-
jected to traffic, the abrasion resistance and skid resistance of an aggregate are essential con-
siderations. The highway should maintain a degree of roughness so that there is sufficient
friction with the tires of vehicles, requiring the surface aggregates to resist polishing over
time. If vehicular traffic causes these particles to become smooth and rounded, the skid
resistance value is lowered and wet conditions can lead to a corresponding increase in acci-
dents. Minerals in aggregates will polish at different rates, and harder materials can be
selected to minimize wear in highly abrasive conditions. The key factor is the polished
stone value, which measures the resistance of an aggregate to polishing. The value can be
determined by subjecting the aggregate to a polishing process and then testing it. If an
aggregate maintains a significant microtexture, it is said to have a high polished stone value
and can be considered to offer resistance to skidding [ASI, 2007, PP. 325–329].

BULK DENSITY AND VOIDS. The bulk density (or unit weight) of an aggregate is the mass
or weight of the aggregate required to fill a container of a specified unit volume—that is,
the volume occupied by both the aggregates and the voids between particles. Generally,
AGGRE GAT ES 51

bulk density ranges from roughly 1,200 to 1,750 kg/m3. Voids range from 30% to 45% for
coarse aggregates to 40% to 50% for fine aggregates [KOSMATKA, KERKHOFF, AND PANA-
RESE, 2008, P. 87].

RELATIVE DENSITY. An aggregate’s relative density (or specific gravity) is the ratio of its
mass to that of an equal absolute volume of water. This ratio is used to determine mixture
proportions. While it is not often considered an indication of aggregate quality, the low
specific gravities of some porous aggregates are correlated with accelerated deterioration
from freeze-thaw cycles.
As explained above, each aggregate particle is made up of solid matter and voids that
may contain water. Since the aggregate mass varies with its moisture content, specific grav-
ity must be determined at a fixed moisture content. There are four (4) moisture conditions
established for aggregates, with varying amounts of water held in the pores and on the sur-
face of the particles:

N In damp or wet aggregate, the pores connected to the surface are filled with water,
with free water also on the surface.
N In saturated surface-dry (SSD) aggregate, the pores connected to the surface are
filled with water but there is no free water on the surface.
N Air-dry aggregate has a dry surface but contains some water in its pores.
N Oven-dry aggregate contains no water either in the pores or on its surface.

The volume of an aggregate is generally considered to be the volume of solid matter and
internal pores. Two (2) different values of specific gravity may be arrived at, depending
upon whether the mass used is an oven-dry or a saturated surface-dry mass. Bulk specific
gravity is the oven-dry mass divided by the mass of a volume of water equal to the SSD
aggregate volume, while bulk specific gravity SSD is the saturated surface-dry mass divided
by the mass of a volume of water equal to the aggregate volume [KOSMATKA, KERKHOFF,
AND PANARESE, 2008, P. 88].
A higher moisture content in coarse aggregate will increase the bulk density; with
fine aggregate, however, moisture levels above the saturated surface-dry state will cause
the bulk density to decrease. The reason for this is that a thin film of water on sand parti-
cles will cause them to clump together so that they aren’t as readily compacted. This pro-
cess, known as bulking, leads to an increase in volume, which translates as a decrease in
bulk density. Bulking isn’t a concern if concrete aggregates are batched by mass, but volu-
metric batching must consider bulking if the moisture content varies. Otherwise, the mix
may be too rich, with possible segregation of the concrete and reduced yield [GAMBHIR,
2004, P. 57].
Bulk density is also influenced by grading, surface texture, and particle shape. Aggre-
gates with a balance of sizes will have a relatively higher bulk density. And if particles have
a higher specific gravity, there will be a higher bulk density for a particular grading. Finally,
AGGRE GAT ES 52

smooth, rounded aggregates can be expected to have a higher bulk density than rough,
angular particles of the same material and grading [GAMBHIR, 2004, P. 57].

A ggregates in T w o-L ift Paving


In two-lift highway construction, two (2) layers are laid down rather than running a single
homogenous layer. The thick lower layer can consist of aggregate of a lesser quality in
terms of durability and strength, including recycled aggregate containing concrete rubble,
asphalt, or locally sourced aggregate. The thinner top layer is made with a high-quality
aggregate that will offer good resistance to freeze-thaw cycles, noise reduction, and supe-
rior traction. This method was widely used in various parts of the United States in the
1950s and 1960s, and is again becoming popular because as quality aggregates are less
readily available. Two-lift construction also has been encouraged with field demonstra-
tions sponsored by the Federal Highway Administration.

Endnotes
Aggregates for Concrete. Portland Cement Association, Skokie, Illinois, 1999. <www.concrete.
org/general/fE1-99.PDF>

Asi, Ibrahim M. “Evaluating Skid Resistance of Different Asphalt Concrete Mixes.” Building and
Environment. Vol. 42, Issue 1, January 2007.

Bennett, D.F.H. Innovations in Concrete. London: Thomas Telford Ltd, 2002.

“Durability: Freeze-Thaw Resistance.” Portland Cement Association, Skokie, Illinois, 2009.


<http://74.125.95.132/search?q=cache:JqapE-uLRCAJ:www.cement.org/tech/cct_dur_freeze-
thaw.asp+cracking+transverse+freezing&cd=2&hl=en&ct=clnk&gl=us>

Ferguson, Bruce K. Porous Pavement. Boca Raton, Florida: CRC Press, 2005.

Gambhir, M.L. Concrete Technology. New Delhi: Tata McGraw-Hill, 2004.

Ghosh, S. Cement and Concrete Science and Technology. London: Thomas Telford, 1992.

Goulias, D.G., and A.H. Ali, “Evaluation of Rubber-Filled Concrete and Correlation between
Destructive and Nondestructive Testing Results.” Cement and Concrete Research, Volume 34,
Issue 12, Elsevier, Amsterdam, 2004.

Hanle, L. J., K. R. Jayaraman, and J. S. Smith, 2006, “CO2 Emissions Profile of the U.S. Cement
Industry,” 2006. <www.epa.gov/ttn/chief/conference/ei13/ghg/hanle.pdf>

Huang, Baoshan; Li, Guoqiang; Pang, Su-Seng; and Eggers, John. “Investigation into Waste Tire
Rubber-Filled Concrete.” Journal of Materials in Civil Engineering, Vol. 16, No. 3, 2004.

Kosmatka, Steven H., Beatrix Kerkhoff, and William C. Panarese. Design and Control of Concrete
Mixtures, 14th Edition (2002). Portland Cement Association, Skokie, Illinois, 2008.
AGGRE GAT ES 53

Long-Life Concrete Pavements in Europe and Canada, Chapter 5: Cement and Concrete. U.S. Dept.
of Transportation – Federal Highway Administration, 2007. <http://international.fhwa.dot.
gov/pubs/pl07027/llcp_07_05.cfm>

“Materials: Aggregates.” Portland Cement Association, Skokie, Illinois, 2009. <www.cement.org/


tech/cct_aggregates_recycled.asp>

“Quiet Pavements: Lessons Learned from Europe.” Focus: Accelerating Infrastructure.


U.S. Dept. of Transportation – Federal Highway Administration, 2005. <www.tfhrc.gov/
focus/apr05/04.htm>

“Recycled Aggregates in Concrete.” The Masterbuilder, Chennai, India, 2004–2009. <www.


masterbuilder.co.in/ci/135/Concrete>

“Recycled Concrete Study Identifies Current Uses, Best Practices.” Focus: Accelerating Infrastruc-
ture. U.S. Dept. of Transportation – Federal Highway Administration, 2004. <www.tfhrc.gov/
FOCUS/apr04/01.htm>

“Recycling Concrete.” Concrete Network, Calimesa, California, 1999–2009. <www.concretenet-


work.com/concrete/demolition/recycling_concrete.htm>

Smith, Mick R., L. Collis, and P. G. Fookes. “Beneficiation.” Aggregates: Sand, Gravel and Crushed
Rock Aggregates for Construction Purposes. London: Geological Society, 2001.

“Summary of California Recycled Concrete Aggregate Review.” U.S. Dept. of Transportation –


Federal Highway Administration, 2003. <www.fhwa.dot.gov/pavement/recycling/rcaca.cfm>

“Summary of Michigan Recycled Concrete Aggregate Review.” U.S. Dept. of Transportation –


Federal Highway Administration, 2003. <www.fhwa.dot.gov/pavement/recycling/rcami.cfm>

“Summary of Texas Recycled Concrete Aggregate Review.” U.S. Dept. of Transportation – Federal
Highway Administration, 2003. <www.fhwa.dot.gov/pavement/recycling/rcatx.cfm>

“Use of Glass Cullet in Roadway Construction.” Texas Department of Transportation, Austin,


2009. <www.txdot.gov/business/contractors_consultants/recycling/glass_cullet.htm>

“Water Reduction in the Mix.” Concrete Network, Calimesa, California, 1999–2009. <www.
concretenetwork.com/concrete/concrete_admixtures/water_reduction.htm>

“What’s a Pop-Out?” ACP International, Halifax, Nova Scotia, Canada. 2001–2007. <www.
exposedaggregateconcrete.com/menu.htm>
CEMENTS

IN THE FINAL CONCRETE MIX, cement and additive materials combine with water
and chemical admixtures (reagents) to form a paste (glue) that holds the aggregates
together. The makeup of cement varies both chemically and physically—and both the
chemistry and physical makeup can be controlled during manufacture. Portland cement,
a cement high in hydraulic calcium silicates, is the cement most used in highway construc-
tion, and there are several formulations for it, with North American and European agen-
cies setting various applicable standards. While the characteristics of water and aggregate
as used in concrete are also important in the makeup of concrete pavements, they are dis-
cussed in other chapters rather than here. This chapter will focus on the nature of cement,
as it is generally known in the European and North American markets.
As in any market, the manufacture of cement in Europe and North America involves
first combining carefully controlled amounts of raw materials under high heat to form
small pellets called clinker. The clinker is then ground under controlled conditions to cre-
ate a fine-powdered cement with particle size ranging from 1 to 200 µm [Popovics, 1992,
p. 38]. The powder is far from uniform, however. A study by the National Institute of Stan-
dards and Technology found that approximately 5% of the particles were 1 µm or smaller;
approximately 9% of the particles were greater than 48 µm [Ferraris, Hackley, and
Avilés, 2004, p. 5]. Thus a differentiation in clinker is constant and allows for a wide
range—normally within limits of the specifications for concrete—of mixing, placing, and
setting characteristics, which require constant quality control measures in concrete pro-
duction and placing practice. This variation in clinker also allows for aftermarket mechan-
ical and chemical practices, whereby the cement clinker variations in size are minimized
or better equalized for improved performance of the concrete.
The makeup of clinker differs between North America and Europe. European stan-
dards, set by the European Committee for Standardization (Comité Européen de Normal-
isation, or CEN), are governed by EN 19 7 , which defines clinker as:

…a hydraulic material which shall consist of at least two-thirds by mass of cal-


cium silicates (3CaO • SiO2 and 2CaO • SiO2), the remainder consisting of
54
CE M ENT S 55

aluminum- and iron-containing clinker phases and other compounds. The


ratio of CaO to SiO2 shall not be less than 2.0. The magnesium content (MgO)
shall not exceed 5.0% by mass.

North American standards are set by the American Society for Testing and Materials
(ASTM) in ASTM C 150, Standard Specification for Portland Cement. Typical makeup of
clinker used in North America is shown in Figure 4-1, Makeup of North American Port-
land Clinker. Type I cement is commonly used for applications where there is a reasonable
period between placing it and putting the road in service.
Figure 4-1
Type III is preferred when there is a short window from
Ma k e u p o f N o r t h Ame r ic a n
placement to use. Po r t la n d Clin k e r
Standards also vary for the final makeup of the cement.
By Mass
In Europe, EN 197 establishes five (5) grades of Portland
Tricalcium silicate (CaO)3•SiO2 45–75%
cement, each with several subdivisions. The five grades are
Dicalcium silicate (CaO)2•SiO2 7–32%
listed in the box North American and European Cements.
In North America, ASTM C 150 establishes 10 categories, Tricalcium aluminate
(CaO)3•Al2O3 0–13%
also listed in North American and European Cements.
Tetracalcium aluminoferrite
Four (4) of the types of ASTM cement—IA, IIA, II(MH)A, (CaO)4•Al2O3•Fe2O3 0–18%
and IIIA—are identical to the types without the A in Gypsum CaSO4•2 H2O 2–10%
their names except that they have an air-entraining admix- Source: Piepel and Redgate, 1998.
ture. Air-entraining admixtures introduce microscopic
air bubbles into the cement, improving its resistance to
freeze/thaw cycles. These will be discussed more thoroughly in the chapter Admixtures.
The (MH) designation means that the cements are for use when moderate heat of hydra-
tion and moderate sulfate resistance are desired.
When comparing European and North American Portland cements, European CEM
I 32.5 R is equivalent to ASTM Type I. Many of the other European and North American
specifications are similar but not identical to each other, and one cannot be automatically
substituted for another.
In North America and Europe, in addition to the classes of cement listed, blended
cements may offer technical or economic advantages. Blended cements are a combination
of Portland cements with a Secondary Cementitious Material (SCM). SCMs include pul-
verized fuel ash, or “Fly Ash” (PFA); ground-granulated blast-furnace slag (GGBF); and
silica fume. Naturally occurring SCMs, called pozzolans after a village in Italy where such
material was first mined, include calcined shale and calcined clay. PFA and silica fume are
often used as adjunct cementitious materials to improve the overall characteristics of the
concrete. They may be used for a variety of reasons, including but not limited to decreas-
ing the heat of hydration in concrete containing a cement with a high C3A content by
reducing the amount of cement required; decreasing permeability; and increasing long-
term strength while providing reasonable early strength.
CE M ENT S 56

No r t h Ame r ic a n a n d Eu r o p e a n Ce me n t s

NORTH AMERICA Type IV—For use when a low heat of hydration


is desired.
In North America, ASTM C 150 defines 10 different
Portland cements for use in North American. Three Type V—For use when high sulfate resistance
(3) of them, designated by the letter “A” following is desired.
the name, have air-entraining admixtures that give
them added resistance to freeze-thaw cycles, as
well as sulfates, de-icers, and alkalis. Those with EUROPE
MH after the name release only moderate amounts
CEN standard EN 197 defines five (5) types of
of heat during hydration.
cement used in Europe:
Type I—For use when the special properties
CEM I—Portland cement is composed of Portland
specified for any other type are not required.
cement and up to 5% of minor additional
Type IA—Air-entraining cement for the same constituents.
uses as Type I, where air-entrainment is desired.
CEM II—Portland-composite cement is composed
Type II—For general use, especially when moder- of Portland cement and up to 35% of other single
ate sulfate resistance is desired. constituents.

Type IIA—Air-entraining cement for the same CEM III—Blast furnace cement is composed of
uses as Type II, where air-entrainment is desired. Portland cement and higher percentages of blast
furnace slag.
Type II(MH)—For general use, especially when
moderate heat of hydration and moderate sulfate CEM IV—Pozzolanic cement is composed of
resistance are desired. Portland cement and up to 55% of pozzolanic
constituents.
Type II(MH)A—Air-entraining cement for the same
uses as Type II (MH), where air-entrainment is CEM V—Composite cement is composed of
desired. Portland cement, blast-furnace slag, and pozzo-
lana or fly ash.
Type III—For use when high early strength is
desired.
Type IIIA—Air-entraining cement for the same [Source: European Committee for Standardization
use as Type III, where air-entrainment is desired. (Comité Européen de Normalisation, or CEN) EN 197.]

Chemical Properties of Cement


Different applications of Portland cement require different properties. Some applications
require high initial strength, low heat of hydration, or high resistance to sulfates, for exam-
ple. These properties are obtained by altering both the chemistry and the fineness of the
cement powder. Fineness, of course, is controlled during the grinding of the clinker. The
chemical makeup is controlled by altering the proportions of raw materials in order to cre-
ate certain compounds in the clinker. The four (4) most important compounds and their
attributes are listed below, followed by chemical shorthand known as the Cement Chemist
Notation (CCN):
CE M ENT S 57

TRICALCIUM SILICATE, C3S, hydrates and hardens rapidly and is largely


responsible for initial set and early strength. In general, the early strength of
Portland cement concrete is higher with increased percentages of C3S.

DICALCIUM SILICATE, C2S, hydrates and hardens slowly and contributes


largely to strength increase at ages beyond one (1) week.

TRICALCIUM ALUMINATE, C3A, liberates a large amount of heat during the


first few days of hydration and hardening. It also contributes slightly to early
strength development. Cements with low percentages of C3A are more resis-
tant to soils and waters containing sulfates.

TETRACALCIUM ALUMINOFERRITE, C4AF, is the product resulting from the


use of iron and aluminum raw materials to reduce the clinkering temperature
during cement manufacture. It contributes little to strength. Most color effects
that make cement gray are due to C4AF and its hydrates.

[Source: Kosmatka, Kerkhoff, and Panarese, 2008, pp. 41–42.]

Figure 4–2

Ch e mic a l Ma k e u p o fR a w Ma t e r ia ls a n d Clin k e r *

Potential Compounds
Type of Chemical Composition of Raw Materials, % Loss on as a Result of Clinkering, % Blaine
portland ignition, fineness,
cement SiO2 Al2O3 Fe2O3 CaO MgO SO3 % Na20 eq C3S C2S C3A C4AF m2/kg

I (min-max) 19.1–22.7 4.0–6.5 0.2–4.0 61.2–67.7 0.5–4.6 2.3–4.4 0.6–2.7 0.14–1.26 42–66 9–30 6–14 1–12 310–497
I (mean) 20.7 5.1 2.6 64.0 2.1 3.0 1.3 0.58 56 17 9 8 381

II (min-max) 20.2–22.5 3.9–5.2 2.6–4.5 61.2–66.1 0.8–4.7 1.9–3.6 0.2–1.9 0.11–1.28 46–66 11–30 4–8 8-14 318–514
II (mean) 21.1 4.5 3.4 63.8 2.2 2.7 1.1 0.54 56 19 6 10 378

III (min-max) 18.9–22.7 3.9–6.6 0.3–4.1 60.7–67.5 0.7–4.9 2.6–4.3 0.3–2.6 0.10–1.21 42–65 9–28 1–14 1–12 319–672
III (mean) 20.5 4.9 2.7 63.6 2.3 3.5 1.3 0.56 56 17 9 8 547

V (min-max) 19.7–27.1 2.0–4.7 0.4–5.7 59.7–65.9 0.5–4.3 19.3–3.1 0.3–2.5 0.06–1.02 42–64 13–46 1–5 1–15 287–681
V (mean) 21.9 3.8 3.8 63.8 2.1 2.3 1.0 0.49 56 21 4 11 385

*Air-entraining cements are not included. Source: Paul D. Tennis, “Portland Cement Characteristics–1998,”
Concrete Technology Today, Vol. 20, No. 2 (August 1999).

As can be seen in Figure 4-2, Chemical Makeup of Raw Materials and Clinker, altering
the raw materials of clinker affects its composition. The raw materials for Type I Portland
cement, for example, contain 5.2% Al2O3 and they result in a clinker that contains 19%
C3A. The raw materials for Type V Portland cement, on the other hand, contain only 3.4%
Al2O3. The resulting clinker contains only 9% C3A, resulting in cement with a lower heat
of hydration that is resistant to sulfates.
CE M ENT S 58

Properties of Cement
Cement must meet certain standards to be acceptable for construction. Industry practice,
guided in North America by ASTM standards, monitors several properties of cement:

FINENESS. Fineness is the beginning point of a dense, strong mix, and is important both
to curing time and to the time elapsed from lay down of the concrete to road access. The
finer the cement, the more quickly it hydrates, and the higher the strength in its early
stages. As a result, the trend in the United States has been toward increasingly finer cements
[ACI, 2001] and most cement manufactured in the U.S. is finer than the ASTM fineness
standard, albeit varying from clinker to clinker production, as identified earlier.
ASTM has set standards for measuring fineness using the Wagner turbidimeter test
(ASTM C115), the Blaine air-permeability test (ASTM C204), or the No. 325 (45 µm) sieve
test (ASTM 430). Most manufacturers are using a laser diffraction test, for which stan-
dards are under development.

SOUNDNESS. Soundness is the ability of hardened cement to retain its volume once it has
set. Soundness is critical, as excessive expansion/contraction leads to cracking and short
useful life. Problems can usually be traced to excess MgO in the cement, or uncombined
CaO (free lime). ASTM C151, the autoclave expansion test, in which cement is heated, sets
tests for soundness and its volume monitored. Industry standards require an expansion of
less than 0.80% of the sample’s original length.

SETTING TIME. Cement must neither begin setting too early nor finish setting too late.
The proper setting time is important to mixing, placing, and curing. Gypsum added to the
cement mixture during manufacture can regulate the set time. Fineness and water/cement
ratio also affect set time. Setting time is usually monitored by repeated tests of its resis-
tance to penetration by small rods or needles of a specified size. In North America, stan-
dards for testing setting are set by ASTM C191 and ASTM C266. Admixtures are generally
used together with curing compounds to affect setting times required in concrete pave-
ments, but concrete placement temperatures and the temperature of the fresh concrete are
also important, as is protection from direct sunlight, wind, and rain—all important vari-
ables to setting time.
There are two (2) types of abnormal setting times: false set and flash set. False set
occurs when cement undergoes a significant loss of plasticity without releasing the normal
amount of heat. Plasticity can be restored from a false set by further mixing without add-
ing water. Flash set occurs when the cement sets quickly while giving off considerable
heat. Plasticity of a flash-set concrete cannot be restored.

HEAT OF HYDRATION. The heat generated when cement and water react is known as heat
of hydration. Excessive heat of hydration will lead to cracking of the concrete paving.
Heat of hydration varies by type of cement—Type III, high early-strength cement, yields up
CE M ENT S 59

to 150% of the heat of Type I, for example, whereas Type IV, low heat-of-hydration cement,
yields only 40% to 60% of the heat in Type I. Physical properties also play a role. An increase
in fineness, water/cement ratio, or ambient temperature increases the heat of hydration.

COMPRESSIVE STRENGTH. Compressive strength is the maximum ability of a cured


sample to withstand axial pressure. Different cement pastes typically have different
strengths at different time periods. Pavement design requirements are specific to the type
of road under construction, and physical requirements of the final concrete will dictate the
characteristics of the cement utilized.

CONSISTENCY. Consistency is the ability of a freshly mixed concrete to flow, and will vary
according to the physical properties of the cement used to manufacture the concrete, as
well as the other parameters—water/cement ratio, types and ratios of aggregates, and
admixtures chosen.

LOSS ON IGNITION. Loss on ignition is measured by heating a cured sample to a high


temperature (900° to 1,000°C) and measuring the loss of mass. A high loss on ignition
can indicate rehydration of the cement during storage, or the presence of unwanted
organic matter.

WEIGHT OF CEMENT. There is a direct relation between cement density, indicated by its
weight, and its durability: The denser the cement, the stronger the concrete. Unfortu-
nately, the weight of a given volume of cement varies due to its handling and storage.
Cement that has had air introduced by pumping or aeration may weigh as little as 833
kg/m3. If the same batch is consolidated by vibration, it can weigh as much as 1,650 kg/m3
[Portland Concrete Association, 1988].

SPECIFIC GRAVITY. Specific gravity is a tool for calculating proportions of a mixture. A


properly proportioned mixture of Portland cement has a specific gravity of around 3.15,
for example, while the specific gravity of blended cements is near 2.90.

Best Practices
The best cement for a given highway depends on regional conditions, but the chosen
cement should contribute to a concrete with characteristics that ensure long-term perfor-
mance of the paving surface. What those factors are, and how to deal with them, will be
discussed more thoroughly in the chapter Mix Designs, but briefly, they include:

N The expected usage, including axle loads, road speeds, and other design parameters
N Elapsed time from placing of the concrete to road use
N The annual and daily temperature range, and the prevalence and severity of
CE M ENT S 60

annual freeze/thaw cycles. Colder regions usually benefit from cements that
readily accept air-entraining admixtures or are themselves air-entrained. Care
must be taken, however, to match the cement and bulk admixtures with the work
cycle. In the far Northern hemisphere, for example, nighttime temperatures
during the summer, when most work is done, are often be considerably colder the
daytime temperatures. This affects the strength development of the pavement, and
the design of the mix must take this into consideration. Warmer regions will
generally need cements that have lower heat of hydration, which once again must
be taken into consideration in the design of the mix—mineral bulk admixtures
may be used to reduce the heat of hydration in the cement used.
N Other local environmental conditions, including ground soil/water physical
chemistry, annual rainfall, and intensity of storms, must also be considered.

While no single cement fills all requirements at all times in all places, most countries have
developed practices to meet their local needs. A U.S. Federal Highway Administration
study looked at countries with pave-
ments that had 40-year life expec-
Ce me n t s fo r Ra p id Pa v e me n t Re c o n s t r u c t i o n
tancy, and it found the following:

In the United States, high-early-strength concrete with a cur-


CANADA. Portland cement is required ing time of 12 hours or less is used so that reconstructed roads
in road construction, but SCMs can can open more quickly. The concretes used vary. One com-
be substituted for a portion of the mon concrete used is Type I cement, with a low water-to-cement
ratio. Another includes Type III cement, which is virtually iden-
cement in any of three (3) ways: The tical to Type I except that Type III cement particles are much
mix can be up to 25% granulated smaller. The smaller cement particles increase the surface
blast-furnace slag (GBFS); up to 10% area, allowing more cement contact with the water in the con-
crete mix, resulting in faster hydration.
fly ash; or a mixture of slag and fly Durability of fast-track concretes is good, because most of
ash up to 25% of the mix [FHWA, these concretes are air-entrained and have relatively low water
Long-life concrete pavements, content—factors which not only improve strength but also
decrease the chloride or salt permeability.
2007].
[Portland Cement Assn., 2009.]

GERMANY. Highways are built with


Portland cement grade CEM I 32.5
R, which is equivalent to ASTM Type I. Several other cements can be used with the
client’s consent: Portland slag cement (CEM II /A-2 or CEM II /B-S), Portland burnt shale
cement (CEM II /A-T or B-T), Portland limestone cement (CEM II /A-LL), or blast-
furnace cement (CEM II /A) (at least 42.5 strength class) are all acceptable [FHWA, Long-
life concrete pavements, 2007].
The maximum water/cement ratio is 0.50; the minimum cement content is 320 kg/m3;
and minimum air content is 4.0%. The German guideline ZTV Beton-StB 2001, Addi-
tional Guidelines for the Construction of Concrete Pavements, sets even stricter standards:
A minimum 0.45 water/cement ratio, and minimum cement content of 350 kg/m3, which
CE M ENT S 61

rises to a minimum of 420 kg/m3 for concrete in exposed aggregate layers [FHWA, Long-
life concrete pavements, 2007].
The maximum fineness allowed is 3,500 square centimeters per gram (cm2/g). There
must be at least 2 hours between setting and placement. Alkali content of type CEM I is
limited to cement to 0.80% Na2O equivalent by mass. This has eliminated cracking that
occurred in during the 1980s in pavements that were 5 to 10 years old [FHWA, Long-life
concrete pavements, 2007].

AUSTRIA. Austrian specifications require the European standard-type CEM II cement.


Initial set time can be no less than 2 hours at 20°C. Blaine fineness can be no greater than
3,500 cm2/g, and 28-day cube strength must be no less than 7 MPa (1,000 psi) [FHWA,
Long-life concrete pavements, 2007].
Portland cement with 20% to 25% slag is used in highway construction, but cement con-
tent varies by use. In two-course (2-course) construction, the minimum cement content in
the lower course is 320 kg/m3 for fixed-form paving and 350 kg/m3 for slip-form paving. The
minimum cement content for the upper course is 370 kg/m3 for fixed-form paving, 400 kg/
m3 for slip-form paving, and 28,450 kg/m3 for an exposed aggregate layer.
An air content of 3.5% to 5.5% is required for fixed-form paving and 4.0% to 6.0% for
slip-form paving.

BELGIUM. Portland cement (CEM I), or a blast-furnace slag cement (CEM III/A), and
mixes with high cement contents, and low water/cement ratios, have lead to a durable,
high-strength concrete. Air-entraining agents, which were not used in concrete pavements
until about 12 years ago, also contribute to pavement quality. Alkali contents of up to 0.9%
are allowed because local aggregates haven’t led to alkali-aggregate reactions [FHWA, Long-
life concrete pavements, 2007].

THE NETHERLANDS. Portland fly-ash cement (CEM II /B-V 32.5 R, containing 30% to
35% fly ash) or Portland cement is preferred but not required. Blended cements of up to
60% slag are allowable [FHWA, Long-life concrete pavements, 2007].

Endnotes
ASTM Standards (various): ASTM International (American Society for Testing and Materials),
Conshohocken, Pennsylvania. <www.astm.org>

Concrete Pavements: H ighw ays. Portland Cement Association, Skokie, Illinois, 2009. <http://www.
cement.org/pavements/pv_cp_highways.asp>
CE M ENT S 62

Ferraris, Chiara F., Vincent A. Hackley, and Ana Ivelisse Avilés. Measurement of Particle Siz e
Distribution in Portland Cement Pow der: Analysis of ASTM Round Robin Studies, pp. 71–81.
Gaithersburg, Maryland: National Institute of Standards and Technology, 2008.

Grube, H., and B. Kerkhoff. The N ew German Concrete Standards DIN EN 206 -1 and DIN EN
1045-2 as the Basis for the Design of Durable Constructions. Research Institute of the German
Cement Industry, German Cement Works, Concrete Technology Reports, Vol. 29,
2001–2003.

Kosmatka, Steven H., Beatrix Kerkhoff, and William C. Panarese. Design and Control of Concrete
Mixtures, 14th Ed. Skokie, Illinois: Portland Cement Assn., 2002, rev. 2008.

Long-Life Concrete Pavements in Europe and Canada, Chapter 5: Cement and Concrete. U.S. Dept.
of Transportation – Federal Highway Administration (FHWA), 2007. <international.fhwa.
dot.gov/pubs/pl07027/llcp_07_05.cfm>

Piepel, Greg, and Trish Redgate. The American Statistician, Vol. 52, No. 1, pp. 23–30. American
Statistical Association, Feb. 1998.

Popovics, Sandor. Concrete Materials, 2nd Ed.: Properties, Specifications, and Testing, p. 38. Park
Ridge, New Jersey: Noyes Publications, 1992.
SUPPLEMENTARY
CEMENTITIOUS
MATERIALS/BULK
SUPPLEMENTS

CONCRETE IS NO LONGER SIMPLY a straightforward recipe of Portland cement,


aggregate, and water. A number of factors, both economical and environmental, have led
to the increasing use of supplementary cementitious materials (SCMs), also known as bulk
supplements. These are primarily byproducts of industry, including blast-furnace slag and
fly ash from coal-fired power stations. They may also be processed natural materials, how-
ever, such as calcined clay, calcined shale, and rice-husk ash.
SCMs can be used to displace some percentage of the Portland cement in a mix because
they themselves have cementitious properties. Rather than passively adding inert bulk,
they help to make concrete more workable, more durable, and stronger. (See Figure 5-1 for
the effects of various SCMs on the characteristics of the mix.) There will be cost savings as
well, because byproducts are cheaper than cement. Energy that would otherwise be used
to fire the kilns involved in cement production is conserved. And there is a corresponding
decrease in carbon dioxide (CO2) emissions. So the increased use of SCMs can be expected
to have a beneficial impact on the atmosphere. Carbon dioxide also is released as a byprod-
uct during the calcination of limestone, in which this mineral is converted to calcium
oxide. Again, using less cement will mean a smaller environmental burden.
Another important benefit to using industrial wastes is that every kilogram incorpo-
rated into a concrete project is 1 less kilogram that has to be trucked to a landfill. Less
obvious is the fact that concrete helps sequester the potentially harmful substances found
in some industrial byproducts, such as heavy metals, so that they are less likely to find
their way into groundwater supplies [ENVIRONMENTAL BUILDING NEWS, 2009].
Today, most concrete mixtures in the United States now include bulk supplements and
SCMs. Still, the U.S. lags behind Europe in the percentage of certain recyclable materials
used in new highway construction. Some European nations are on record as using 100%
of the available blast-furnace slag, steel slag, coal fly ash, or coal bottom ash, with the
Netherlands distinguished by its high figures. This is not just a matter of enlightened pol-
icy; using recycled materials has more-obvious advantages in regions with dense popula-
tion, limited space for landfills, modest sources of natural aggregates, and a ready supply

63
SU PPL E M E N TA RY CE M E NT I T I OUS M AT E RI AL / BUL K S UP P L E M ENT S 64

of recycled materials. But beyond a country’s profile in this regard, governments can
encourage the use of recycled products by doing the research necessary to arrive at policy
decisions, and then by stating clear and consistent standards [HOLTZ AND EIGHMY, 2000].
There are potential problems with the delayed strength development of concrete con-
taining fly ash or slag in the event of low temperatures at the site. A mix with fly ash may

Figure 5-1

Effe c t so f Su p p el em n t a r y Ce me n t i t i o u s Ma t e r ia ls
o n Fr e hs Co n c r e t e Pr o p e r t ie s
KEY:
Fly ash Natural pozzolans
reduced
GGBF Silica Calcined Calcined significantly reduced
Class F Class C slag fume shale clay Metakaolin
increased

Water requirements significantly increased

no significant change
Workability
effect varies
Bleeding and segregation

Air content * *

Heat of hydration

Setting time

Finishability

Pumpability

Plastic shringage cracking

*Effect depends on properties of fly ash, including carbon content, Source: Taylor, Kosmatka, Voigt, et al., 2007, p. 37.
alkali content, fineness, and other chemical properties.

Figure 5-2

Effe c t s o f Su p p el em n t a r y Ce me n t it io u s Ma t e r ia ls
o n H a r d e n e d Co n c r e t e Pr o p e r t ie s
KEY:
Fly ash Natural pozzolans
reduced
GGBF Silica Calcined Calcined significantly reduced
Class F Class C slag fume shale clay Metakaolin
increased

Early strength significantly increased

no significant change
Long-term strength
effect varies
Permeability

Chloride ingress

ASR

Sulfate resistance

Freezing and thawing

Abrasion resistance

Drying shrinkage

Source: Taylor, Kosmatka, Voigt, et al., 2007, p. 37.


SU PPL E M E N TA RY CE M E NT I T I OUS M AT E RI AL / BUL K S UP P L E M ENT S 65

not be substantially stronger until well after a week,


Po zo la n s in Co n c r e t e
because SCMs contribute strength through a second-
ary reaction requiring a level of calcium hydroxide.
Generally, freeze-thaw cycles are particularly apt to Pozzolans are materials that have little if any
cementitious property until they react with
compromise durability with these mixes, if de-icing
the calcium hydroxide in a mix, and they
salts penetrate and cause damage. And while air- then increase the workability and hardness
entraining agents are available to make concrete less of concrete [ACI, CEMENTITIOUS MATERIALS FOR
CONCRETE, 2001]. SCMs may affect the proper-
vulnerable to these cycles, they may not be compati-
ties of fresh and hardened concrete through
ble with high-carbon fly ash. Therefore, some specifi- pozzolanic as well as hydraulic activity [TAY-
cations restrict the use of fly ash and slag to warmer LOR, KOSMATKA, VOIGT, ET AL., 2007, P. 31].

months [ACI, CEMENTITIOUS MATERIALS FOR CON-


CRETE, 2001].
Note that because mixtures containing SCMs
have more chance of unanticipated interactions, it is important to do testing in order to
ensure that they achieve the desired results, to verify optimal levels, and to check for any
unintended effects [TAYLOR, KOSMATKA, VOIGT, ET AL., 2007, P. 31].

Blend ed Cements
Blended cements are those incorporating at least one (1) SCM. There are five (5) generally
recognized types of blended cement [CEMENT AND CONCRETE BASICS, HISTORY & MANU-
FACTURE, 2010]:

1. PORTLAND BLAST-FURNACE SLAG CEMENT (Type IS) incorporates from 25% to


70% granulated blast-furnace slag by mass.

2. POZZOLAN-MODIFIED PORTLAND CEMENT, used in general construction, has


less than 15% pozzolans and 25% to 70% slag.

3. SLAG CEMENT, with a minimum of 70% water-quenched blast-furnace slag, is


slow to develop strength and is used to construct bridge piers, among other
applications.

4. SLAG-MODIFIED CEMENT contains a maximum of 25% slag. The slag can be


interground—that is, ground along with Portland cement clinker—or be ground
in a separate operation and then added to the cement. This mix is used in general
construction.

5. PORTLAND-POZZOLAN CEMENT contains from 15% to 40% pozzolans. The


pozzolans can be either interground or processed separately.

While blended cements are common in Europe and parts of Asia, their use has lagged
behind in North America, according to the Portland Cement Association, with a prefer-
SU PPL E M E N TA RY CE M E NT I T I OUS M AT E RI AL / BUL K S UP P L E M ENT S 66

ence for adding supplementary materials directly to the concrete rather than pre-blending
them. The U.S. industry has been slow to promote the advantages of blended cement over
Portland cement. And government guidelines have been formulated in a way that discour-
ages specifying them. But the U.S. recently mandated changes to ASTM and AASHO stan-
dards for blended cements that should help promote their wider use, through simpler,
clearer terminology. The percentage of SCMs in a blended cement is noted in the product
name. As examples, Type IP (15) cement contains 15% pozzolan, and Type IS (25) cement
contains 25% slag cement [TENNIS AND MELANDER, 2010].
Traditionally, fly ash, GGBF slag, calcined clay, calcined shale, and silica fume have
been used in concrete individually, but producers have been combining two (2) or more of
these materials for optimal performance. Mixtures of Portland cement and one (1) SCM
are termed binary mixes. Ternary mixtures, with two (2) added SCMs, are becoming more
common [WANG, 2003, P. 1].

F ly A sh
En v ir o n me n t a lCo n c e r n sa b o u t Fl y Ash
Fly ash is the most widely used SCM, found in
roughly 50% of ready-mixed concrete [PORT-
LAND CEMENT ASSN., SURVEY, 2000]. Coun- At the time this study was published, the U.S. Envi-
ronmental Protection Agency is considering the safety,
tries vary greatly in how much of the fly ash storage, and handling of fly ash, due to the trace pres-
they generate will find its way into concrete; ence of heavy metals. Studies over the past 10 years,
utilization percentages range from less than however, have indicated that when fly ash is encapsu-
lated by concrete, leaching of the heavy metals into
10% up to 100% [MEYER, 2005, P. 3]. In the the environment is not a problem. Fly ash remains a
U.S., the EPA has been working with state and mainstay in concrete pavements at this time, but pos-
federal agencies, as well as with industry, to sible hazardous classifications may raise the cost and
handling procedure of the material in concrete
increase the use of coal combustion products
pavement.
in concrete production from 12.4 million tons
in 2001 to an anticipated 18.6 million tons by
2011, a jump of 50% [U.S. EPA, 2009].
Fly ash is produced by coal-fired electric and steam-generating plants. Typically, coal
is pulverized and blown with air into the boiler’s combustion chamber, where it leaves a
molten mineral residue. As the boiler tubes extract heat, the flue gases are cooled and then
harden to form ash. Coarser particles, referred to as bottom ash or slag, fall to the bottom
of the combustion chamber, while the lighter fine ash particles, termed fly ash, remain sus-
pended in the flue gas, from which they are removed by emission control [FHWA, PAVE-
MENTS, 2006].
Fly ash has several effects on concrete, both during its production and over time [TAY-
LOR, KOSMATKA, VOIGT, ET AL., 2007, P. 32]. When fly ash is used:

N Typically, less water will be needed to achieve workability.


N Setting time may be delayed.
SU PPL E M E N TA RY CE M E NT I T I OUS M AT E RI AL / BUL K S UP P L E M ENT S 67

N The strength may be initially lessened but increase


over time, because fly ash reaction rates are initially
slower but continue longer.
N Heat of hydration is reduced.
N Concrete may have greater resistance to alkali-silica
reaction (ASR) and sulfate attack.
N Permeability is reduced; consequently, resistance to
chloride ion penetration is improved.
N On the negative side, incompatibility with certain
Scanning electron microscope micrograph of fly ash
cements and chemical admixtures may cause early particles at 1000X.
stiffening.

Fly ash is typically finer than Portland cement and lime, consisting of small glass spheres
that serve to improve the fluidity and workability of fresh concrete. Fineness also increases
the pozzolanic reactivity of fly ash [FHWA, PAVEMENTS, 2006].
The quality of fly ash is affected by the characteristics of the coal that produced it,
including the co-firing of bituminous and sub-bituminous coals, as well as by the pro-
cesses with which flue gases are cleaned. The four (4) most relevant characteristics of fly
ash for use in concrete are loss on ignition (LOI), fineness, chemical composition, and
uniformity [FHWA, PAVEMENTS, 2006].
LOI is an index of unburned carbon remaining in the ash, and it is a critical character-
istic, especially for concrete applications. High carbon levels can result in significant
air-entrainment problems in fresh
concrete, adversely affecting durability.
Fly Ash Cla sifi c a t io n s
While AASHTO and ASTM specify
limits for LOI, some state transporta-
tion departments specify lower levels There are two classifications of fly ash, depending on the type
of coal burned. Class F ashes are typically derived from bitu-
[FHWA, PAVEMENTS, 2006]. minous and anthracite coals and consist primarily of an
Fineness of fly ash is affected by alumino-silicate glass, with quartz, mullite, and magnetite
the condition of the coal crushers and also present. Class F, or low-calcium fly ash, has less than
10% calcium oxide. Class C ashes are generally derived from
the grinding properties of the coal
sub-bituminous coals and consist primarily of calcium alu-
itself. Limits on fineness are addressed mino-sulfate glass, as well as quartz, tricalcium aluminate,
by ASTM and state transportation and calcium oxide). Class C ash is also referred to as high-
calcium fly ash because it typically contains more than 20%
department specifications. Fly ash can
calcium oxide.
be processed by screening or air clas-
While both classes are pozzolans, some Class C ashes will
sification to improve its fineness and hydrate and harden when exposed to water, meaning that
reactivity [FHWA, PAVEMENTS, 2006]. they may also be considered a hydraulic material [TAYLOR, KOS-
Not surprisingly, the chemical MATKA, VOIGT, ET AL., 2007, P. 32.]. Class F fly ash is generally used
at levels of 15% to 25% by mass of cementitious material;
composition of fly ash has much to do
Class C fly ash is generally used at levels of 15% to 40%.
with the type of coal consumed, as
well as with any additional fuels or
SU PPL E M E N TA RY CE M E NT I T I OUS M AT E RI AL / BUL K S UP P L E M ENT S 68

additives used in the combustion or post-combustion processes. The particular pollution


control technology also may be a factor. It’s necessary to test the chemistry of the fly ash
for particular applications. Some plants burn specific coals or modify their additives to
improve ash quality and ensure uniform fly ash characteristics from one shipment to the
next. In the United States, quality specifications vary considerably: Some states require
testing samples, while others maintain lists of approved sources and accept project suppli-
ers’ certifications of fly-ash quality [FHWA, PAVEMENTS, 2006].

S ilica F um e
Silica fume, also known as microsilica, is a byproduct of the
production of various silicon and ferrosilicon alloys [ACI
COMMMITTEE 226, 1987b]. Until the 1970s, most silica fume
was discharged into the atmosphere; but it now is collected
for environmental reasons, and can be used in concrete mixes
to improve compressive strength, bond strength, and abra-
sion resistance. Silica fume has been used to produce very
high-strength, low-permeability, and chemically resistant
Scanning electron microscope micrograph of silica
concrete [WOLSIEFER, 1984, PP. 25–31]. fume particles at 20,000X.
Levels of air-entraining agent should be increased to
maintain the required air content when using silica fume, because of the greater surface
area and the presence of carbon [CARETTE AND MALHOTRA, 1983, PP. 3–13].
It’s important to note that incorporating more than 10% silica fume can cause concrete
to become sticky, and workability must be improved by increasing the initial slump. Incor-
porating silica fume into a mixture with high-range water reducer (HRWR) makes it possi-
ble to use a lower ratio of water to cementitious materials. Silica fume has been found to
significantly reduce the chloride permeability of a concrete mix, primarily because of the
increased density of the matrix. As with other pozzolans, the use of silica fume can reduce
ASR and prevent harm from the resulting expansion [FHWA, SILICA FUME, 2008].
Silica fume levels range up to 15% by weight of cement. At this upper limit, the con-
crete tends to be very strong but brittle. Levels of 7% to 10% are more typical. Water reduc-
ers usually are not required at very low levels, but they are needed at the higher range
[FHWA, SILICA FUME, 2008].
Silica fume is available in either dry or wet form. Dry silica can be stored in silos and
hoppers, while slurried products are stored in tanks [FHWA, SILICA FUME, 2008]. The loose
bulk density of silica fume is very low, and to make for easier handling, it usually is tum-
bled in an air stream, creating larger particles held together by electrostatic forces. The
agglomerations should be broken up, and the material then uniformly distributed in the
mix [TAYLOR, KOSMATKA, VOIGT, ET AL., 2007, P. 35].
Because silica fume can reduce workability and is relatively expensive, it typically is
not used in pavements except for curbs and gutters, and where concrete will be subjected
to studded tires [TAYLOR, KOSMATKA, VOIGT, ET AL., 2007, P. 35].
SU PPL E M E N TA RY CE M E NT I T I OUS M AT E RI AL / BUL K S UP P L E M ENT S 69

G round , G ranu lated Blast-F ur nace S lag


Ground, granulated blast-furnace slag (GGBF), also called slag cement, is a granular mate-
rial formed when molten iron blast-furnace slag is rapidly chilled by immersion in water.
It has long been used as a cementitious material in concrete, but with the decline of steel
production in the United States, it has become less available and now is imported from
other countries [MEYER, 2005, P. 4]. Because of its chemistry, GGBF slag behaves much
like a cement, hydrating without the presence of added calcium. However, hydration rates
can be accelerated by increasing the pH of the mix [FHWA, GGBF SLAG, 1996].
GGBF can be substituted for cement, kilogram for kilogram. In practice, up to 70% of
the Portland cement in a mix has been replaced by this material. But the substitution rate
generally should be limited to 50%, and just 25% for concretes that will be exposed to de-
icing salts. Research has shown that the scaling resistance of concrete suffers somewhat if
GGBF substitution goes above 25% [FHWA, GGBF SLAG, 1996].
GGBF can be expected to affect concrete in several ways:

N Somewhat less water is needed to achieve the expected level of workability


[TAYLOR, KOSMATKA, VOIGT, ET AL., 2007, P. 34].
N Setting times will increase as the slag content goes up. An increase of slag content
from 35% to 65% by mass can extend the setting time by as much as 60 minutes.
This delay may be beneficial, particularly in large pours and when hot weather
prevents the formation of “cold joints” in successive pours.
N Concrete will have significantly more resistance to chloride penetration. In North
America, up to 35% of the cementitious material in paving mixes is made up of
GGBF. Higher levels may be required to resist alkali-silica reactions or to reduce
the heat of hydration [TAYLOR, KOSMATKA, VOIGT, ET AL., 2007, P. 34].
N Incorporating GGBF in cement paste can help reduce the size of pores in the mix,
resulting in decreased permeability [FHWA, GGBF SLAG, 1996]. Consequently,
these concretes may require less depth of cover to protect the reinforcing steel.
N The rate and quantity of bleeding usually will less because of the relatively higher
fineness of slag. This property also increases the need for air-entraining agents,
relative to conventional concrete. However, unlike fly
ash, slag does not contain the carbon that potentially
causes instability and air loss in concrete [FHWA,
GGBF SLAG, 1996].
N Early strengths may be depressed, but this is com-
pensated for by eventual strength gains [TAYLOR,
KOSMATKA, VOIGT, ET AL., 2007, P. 34].

The compressive strength development of slag concrete


depends primarily upon the type, the fineness, the activity
Scanning electron microscope micrograph of ground,
index, and the proportions of slag used in concrete mixtures granulated blast-furnace slag particles at 2100X.
SU PPL E M E N TA RY CE M E NT I T I OUS M AT E RI AL / BUL K S UP P L E M ENT S 70

[MALHOTRA, 1987]. In general, the strength development of concrete incorporating GGBF


is slowed. Between 7 and 28 days, the strength approaches that of the control concrete, and
then it exceeds the conventional strength [ADMIXTURES AND GROUND SLAG, 1990]. If flex-
ural strength is important, it can be improved by the addition of slag cement. This increased
flexural strength is thought to be caused by stronger bonds in the cement-slag-aggregate
system because of the shape and surface texture of slag particles [FHWA, GGBF SLAG, 1996].
Problems can arise when slag concrete is placed at low temperatures. Strength is sub-
stantially lessened for up to 14 days, and it is best to reduce the percentage of slag to
between 25% and 30% [FHWA, GGBF SLAG, 1996].
As for the freeze-thaw durability of air-entrained concrete that incorporates GGBF, it
has been found to be generally as resistant as conventional concrete [MALHOTRA, 1987].
Test results indicate that regardless of the ratio of water to cement/slag, air-entrained slag-
concrete specimens performed excellently in freeze-thaw tests, with relative durability fac-
tors greater than 91% [FHWA, GGBF SLAG, 1996].
Slag helps prevent damage from ASR in various ways: the reduction of total alkalis in
the cement-slag blend; the lower permeability of the system; and the sequestration of alka-
lis in the hydration process [FHWA, GGBF SLAG, 1996].
Proper curing is essential for concretes in which GGBF substitutes for some percent-
age of Portland cement. That’s because of the reduced heat of hydration and slower rate of
initial strength gain. These concretes are susceptible to cracking caused by drying shrink-
age. Also, set retardation will become more pronounced at lower temperatures. In cold
ambient conditions, it’s necessary to maintain favorable curing temperatures until the
concrete is strong enough to resist damage from freezing temperatures and can safely be
removed from forms. Matched curing or non-destructive tests can be used to determine
in-place concrete strength and tell when cold weather protection and forms are no longer
necessary [FHWA, GGBF SLAG, 1996].

N atur al Poz z olans


Alk a li-Silc a Re a c t io n
Natural pozzolans have been used for centuries, and the term
comes from an Italian city, Pozzuoli, whose volcanic ash was
used as a cement. Many ancient Roman structures built of Alkali-silica reactions (ASR) are a
potential issue when using aggre-
pozzolan concrete can still be seen today, attesting to the gates. ASR causes the formation of a
durability of these materials [TAYLOR, KOSMATKA, VOIGT, ET moisture-absorbing gel that may
AL., 2007, P. 34]. expand to cause cracking. The effect
can be countered by using pozzolans
Calcined clay, calcined shale, and metakaolin are most
such as fly ash or GGBF slag, as well
commonly used natural pozzolans today. Their processing as low-alkali Portland cement, to
involves heat-treating in a kiln and grinding to a fine powder reduce the free alkalis that would
react with the aggregate [ACI, CEMENTI-
[TAYLOR, KOSMATKA, VOIGT, ET AL., 2007, P. 34].
TIOUS MATERIALS FOR CONCRETE, 2001].
Calcined clays are used in general-purpose concrete
mixes, typically replacing 15% to 35% of the cement. They
SU PPL E M E N TA RY CE M E NT I T I OUS M AT E RI AL / BUL K S UP P L E M ENT S 71

Scanning electron microscope micrograph of Scanning electron microscope micrograph of


calcined clay particles at 2000X. calcined shale particles at 5000X.

improve strength development and resistance to sulfate attack, help to control alkali-silica
reactivity, and reduce permeability [TAYLOR, KOSMATKA, VOIGT, ET AL., 2007, P. 35].
Because calcined shale can contain 5% to 10% calcium, it has some cementitious or
hydraulic properties [TAYLOR, KOSMATKA, VOIGT, ET AL., 2007, P. 35].
Metakaolin is a special calcined clay produced by the low-temperature calcination of
high-purity kaolin clay, followed by grinding to an average particle size of about 1 to 2 µm;
this is about 10 times finer than cement, but still 10 times coarser than silica fume. Metaka-
olin is used for applications requiring very low permeability or very high strength, serving
more as an additive to the concrete rather than as a replacement for cement. Typically,
metakolin will be used at levels of around 10% of the cement mass [tAYLOR, KOSMATKA,
VOIGT, ET AL., 2007, P. 35].

Endnotes
ACI Committee 226. 1987b. Silica Fume in Concrete: Preliminary Report. ACI Materials Journal
March–April: 158–66. American Concrete Institute, Farmington Hills, Michigan, 1987.

“Admixtures and Ground Slag for Concrete,” Transportation Research Circular No. 365 (Decem-
ber). Washington, D.C.: Transportation Research Board, National Research Council, 1990.

“Blended Cement,” Today’s Concrete Technology, 2010. <www.todaysconcretetechnology.com/


blended-cement.html>

Carette, G. G., and V. M. Malhotra. “Mechanical Properties, Durability and Drying Shrinkage of
Portland Cement Concrete Incorporating Silica Fume.” Cement, Concrete, and Aggregates,
Volume 5, Number 1, 1983. <www.astm.org/DIGITAL_LIBRARY/JOURNALS/CEMENT/
PAGES/484.htm>

Cement and Concrete Basics: History & Manufacture of Portland Cement. Portland Cement
Association, Skokie, Illinois, 2010. <www.cement.org/basics/concretebasics_history.asp>

Cement and Concrete Basics: Supplementary Cementing Materials. Portland Cement Association,
Skokie, Illinois, 2010. <www.cement.org/basics/concretebasics_supplementary.asp>
SU PPL E M E N TA RY CE M E NT I T I OUS M AT E RI AL / BUL K S UP P L E M ENT S 72

Cementitious Materials for Concrete. ACI Education Bulletin E3-01, American Concrete Institute,
Farmington Hills, Michigan, 2001. <www.concrete.org/general/fE3-01.pdf>

“Fly Ash Facts for Highway Engineers.” Pavements. U.S.D.O.T. – Federal Highway Administration,
2006. <www.fhwa.dot.gov/PAVEMENT/recycling/fach01.cfm>

Ground Granulated Blast-Furnace Slag. U.S.D.O.T. – Federal Highway Administration, 1999.


<www.fhwa.dot.gov/infrastructure/materialsgrp/ggbfs.htm>

Hanle, Lisa J. “CO2 Emissions Profile of the U.S. Cement Industry.” U.S. Environmental Protection
Agency, Washington, D.C. <www.epa.gov/ttnchie1/conference/ei13/ghg/hanle.pdf>

Holtz, Katherine, and T. Taylor Eighmy. “Scanning European Advances in the Use of Recycled
Materials in Highway Construction,” Public Roads, July/August 2000. U.S.D.O.T. – Federal
Highway Administration. <www.tfhrc.gov/pubrds/julaug00/recycscan.htm>

Malhotra, V. M. “Properties of Fresh and Hardened Concrete Incorporating Ground Granulated


Blast Furnace Slag,” Supplementary Cementing Materials for Concrete,” ed. V. M. Malhotra,
291–331. Ottawa: Canadian Government Publishing Centre, 1987.

Meyer, Christian, “Concrete as a Green Building Material.” Proceedings of ConMat ’05, Vancouver,
British Columbia, 2005. <www.civil.columbia.edu/meyer/publications/publications/87
Concrete as a Green Building Material.pdf>

Silica Fume. U.S.D.O.T. – Federal Highway Administration, 2008. <www.fhwa.dot.gov/infrastruc-


ture/materialsgrp/silica.htm>

Survey of Mineral Admixtures and Blended Cements in Ready-Mixed Concrete, Portland Cement
Association, Skokie, Illinois, 2000.

Taylor, Peter C., Steven H. Kosmatka, Gerald F. Voigt, et al., Integrated Materials and Construction
Practices for Concrete Pavement: A State-of-the-Practice Manual. National Concrete Pavement
Technology Center, Iowa State University, 2007. <www.cptechcenter.org/publications/imcp/
index.cfm#ch3>

Tennis, Paul D., and John M. Melander, “U.S. Cement Specifications,” Concrete International,
2010. <www.allbusiness.com/company-activities-management/operations-quality-control/
13706656-1.html>

“Using Fly Ash in Concrete.” Environmental Building News, February 1, 2009. <www.building-
green.com/auth/article.cfm/2009/1/29/Using-Fly-Ash-in-Concrete/>

Wang, Kejin. Properties of Blended Cements for Concrete Pavements. National Concrete Pavement
Technology Center, Iowa State University, 2003. <www.cptechcenter.org/projects/detail.
cfm?projectID=844170850>

Wastes—Partnerships—Coal Combustion Products Partnership (C2P2), U.S. Environmental


Protection Agency, Washington, D.C., 2009. <www.epa.gov/waste/partnerships/c2p2/index.
htm>

Wolsiefer, J. “Ultra High-Strength Field Placeable Concrete with Silica Fume Admixture.” Con-
crete International: Design and Construction, Volume 6, Number 4. American Concrete
Institute, April 1984. <www.concrete.org/PUBS/JOURNALS/AbstractDetails.asp?ID=9235>
WATER FOR
CONCRETE
PRODUCTION

AS CONCRETE GAINS ACCEPTANCE as a pavement medium worldwide, the perfor-


mance of concrete pavements is coming under increased scrutiny. As the performance of
concrete pavements is looked at more closely, scientific analysis is increasingly applied to
the manufacture of the product, and the ingredients used in making the product, in addi-
tion to placement, methodology, and paving operations.
To those in other industries, the idea of scientific analysis of the ingredients of all the
constituents of any product may seem perfunctory; however, concrete is a man-made
product, more often than not produced with a wide variety of raw or natural materials. To
make the highest quality concrete, it stands to reason that the highest quality control of
each of the raw material ingredients needs to be undertaken.
Indeed, many of the materials used in the manufacture of concrete are rigorously con-
trolled. Cement is a relatively scientifically produced material, manufactured under con-
trolled environments to published standards of national and international levels, and
subject to relatively stringent quality-control measures.
Chemical admixtures are produced to manufacturing specifications that include qual-
ity standards for base and combined chemicals, and are often subjected to materials analy-
sis that includes microscopic and nanotechnology within the chemists’ quality-control
laboratories. The very nature of chemical reactions with cement and water emulsifications
needs to be carefully analyzed. Some concrete admixtures are actually food or organic-
based byproducts and as such are subjected to additional standards.
Aggregates, both coarse and fine, are usually carefully processed and selected with
regard to hardness, shape, size, and gradation. The gradation of the fine aggregates is a
careful process that allows the material to become the binding matrix of the “artificial
stone,” which concrete becomes when properly manufactured. The coarse aggregate ingre-
dients are selected for hardness and shape, resulting in the mass makeup of the product,
and providing the majority of the strength of that mass, once it’s set in a hardened state.
Careful geotechnical analyses of aggregates are undertaken in selection and processing to
ensure concrete and concrete pavement quality.

73
WAT E R F OR C ONC RE T E PRODUC T ION 74

Concrete: A Water-Based Material


Concrete is actually a water-based material, and as such all materials in concrete are
affected to some degree by water or by moisture levels. Concrete requires water for making
it—as a measured mix ingredient. Concrete—a material blended from dry bulk powders,
moist raw materials, and chemical admixtures—requires water to be formed and set. In
fact, the whole of concrete is totally dependent on water. To make it, to mix it, to form it,
and to have it harden at all, water must be present.
And yet water is the least specified of the ingredients that constitute concrete mix
designs. Historically, the term “potable” was used as the only definition of usable or allow-
able water in concrete. The Portland Cement Association defines batch water as “water
discharged into the mixer from municipal water supply, reclaimed municipal water, or
water resulting from concrete production operations,” but sets no specific standards for
chemical purity. [KOSMATKA, KERKHOFF, AND PANARESE, 2008, P. 73]. The American Con-
crete Institute (ACI) notes that “potable water is usually acceptable.” While the ACI does
say that non-potable can be used if it water meets ASTM1603, the standard sets only per-
formance standards; chemical testing remains optional [National Ready Mixed Con-
crete Assn., 2010, PP. 1–2].
Such specifications have become a problem for concrete specifiers in Europe and
North America, as the move to include recycled or reclaimed water became increasingly
specified concurrently, and the definition has begun to change. If “potable” is defined as
water free of organic substances, then for the production of concrete, the term potable no
longer implies “drinkable” or “safe for cooking” in this case, but instead a usable concrete
water that takes on pH levels and particulate levels, including maximum levels of recycled
cement and admix pastes or solids.

Concrete Water
The idea of how water should be treated ahead of concrete production, however, is a basic
question that forward-thinking concrete technologists have asked for decades. In looking
at other water-based manufacturing processes, people in such industries rarely take water
for granted. The conditioning of water to optimize the emulsifications for performance of
the product, as well as the economies in a product’s manufacture, is considered a basic
premise.
Academics within various research facilities of the former Soviet Union closely inves-
tigated “concrete water” conditioning, in their search of the basic causes for hydration and
hydration control in curing of concrete. These studies undertaken in the 1970s and early
1980s were for the most part archived following the studies, and not put into practice
beyond the laboratory proof phases. The mineral hardness was found to be important,
with metals—especially iron and heavy metals such as lead, mercury, and arsenic—affect-
ing concrete elasticity and workability.
WAT E R F OR C ONC RE T E PRODUC T ION 75

Material set times were found to vary greatly, and in long-term destructive testing,
corrosion resistance was noted to vary widely, depending on the water used in the produc-
tion of the concrete. It was confirmed that the chemical admixtures were greatly affected
by the type of water used, and that the type of water especially affected the amount of time
the concrete remained workable.
In many cases fresh (soft) rainwater, river water, or lake water is used in the produc-
tion of concrete. In other cases, well water—which generally speaking contains many more
minerals—is used. In still other situations, municipally treated water is used. The varying
sodium and other mineral and heavy metal levels of these differently sourced waters cer-
tainly are suspected to affect concrete, and the effect of chlorines and fluorides and numer-
ous other treatment reagents in municipal-treated water were all investigated. But it was
in the development of primary and secondary water treatment methods that the most
interesting results were discovered.

Treatment of Water for Concrete Production


Over the centuries that concrete has evolved, various types of filtering and processing of
water for the manufacture of concrete have been undertaken. Sand and carbon filtering,
reverse osmosis, and distillation have all been completed in attempts to provide a baseline
with regard to the water ingredient. These treatments all provide for better concrete than
utilizing water that is simply taken for granted and which varies from batch to batch.
For some years, many industries have used the electrical pass-through water treat-
ment for the polarization (splitting of molecules) of water to capture heavy metals and to
decant unwanted minerals. This type of electrical treatment works by electrodynamics
and electrical impacts, disintegrating the water molecule clusters, changing the electrical
load of the constituents, and altogether improving the water’s ability to react. The pass-
through water treatment can be adjusted to the specific needs of the industrial process and
was found to be a factor in the production of water-based products of various quality.
In the 1990s, German scientists began applying this technology to the treatment of
water used in the production of concrete in practice as well as in laboratory environs. It was
discovered that the polarization of the water molecules resulted in longer workability of
concrete—in some cases up to 3 (three) times as long, and in eventual hydration values and
compressive strengths that were from 10% to 40% higher, depending on the variables.
It is important, of course, to define a correlation between laboratory tests and the pro-
duction of particular mix designs. Pass-through water equipment must be calibrated and
made suitable for the actual production requirements. With careful adjustment of the
pass-through water treatment based on testing of the actual water used in the process,
however, a defined result is obtained and the concrete mix design altered to provide the
desired effects.
In 2007 the activities of the pass-through electrical water treatment system were
refined and patented worldwide for concrete production in a system. FML Technologies
WAT E R F OR C ONC RE T E PRODUC T ION 76

of Germany holds the marketing and technological patent protection of the water treat-
ment system specifically for concrete production [REINECKE, 2010, P. 16]. This patented
approach is to treat the water utilized in concrete production as well as emulsify the pow-
ders used in concrete production with the treated water through a high-shear colloidal
mixing system and/or high-shear vortex pumping systems. (See the chapter Mix Designs
for more.) Both the stand-alone water treatment and the emulsification mixing systems
are covered under these patents.
The results are significant, as the treated water reacts with the fine grind of the cement
colloidal emulsifications, including admixtures. Further, the patented water-treatment
system consultants undertake laboratory testing of the concrete producer’s water, aggre-
gates, and bulk powders, especially cements, as its analysis then allows for predictable
results that provide reliable forecasts of the respective concrete properties and the removal
of quantities of constituents; modification of the mix designs; or advanced curing and
accelerated strengths and strip/curing times, depending on the particular objectives of the
producer.
The patented concrete water electrical-treatment system has proven to allow the
cement and other ultrafine particulate a negative attraction, due to actual disruptions in
the water molecules—splitting the water hydrogen and oxygen components, the water is
realigned by contact with solids particulate, creating a magnetic attraction that allows for
collision of the hydrating components.

Making Better Concrete Pavements


Testing has ensured that the workability window, or “dwell,” is significantly increased when
certain admix chemicals such as water-reducing agents are used. The resulting increased
workability allows for a significant reduction in the overall water/cement ratio of pave-
ment mix designs, and enhancement of placement properties in heavily reinforced con-
crete and ancillary dividers, bridge decks, etc.
Enhancing the mix water is significant insofar as enhanced mix designs with extremely
low water/cement ratios ensure higher-strength concrete with a maximum density.
Although low- and mid-range water-reducing agents can be added to water/cement mixes
with lowest water content, interestingly the reduction of water and cement is providing
new breakthroughs in workability factors in altered mix designs with low water/cement
ratios. For higher spread or slump concrete, the emulsification process has proven to pro-
vide a fantastic pumping/placing and finishing aid.
Enhancing the mix water is significant, insofar as enhanced mix designs with extremely
low water/cement ratios ensure higher-strength concrete with a maximum density.
Enhancing the mix water so that admixtures, cement, and other mix materials that consti-
tute the emulsification matrix that binds concrete mixtures when hydrated is significant
insofar as enhanced mix designs with extremely low water/cement ratios will ensure
higher strengths, with a maximum density for a given mix.
WAT E R F OR C ONC RE T E PRODUC T ION 77

The increased workability and dwell of workability both allow for enhanced place-
ment and finish. The results of electrically treated concrete water on the overall pavement
performance is yet to be studied, but it certainly stands to reason that using better-mixed
concrete as a result of treated water in concrete will yield an improved life cycle for a given
concrete pavement.

Endnotes
“Guide for Selecting Proportions for High-Strength Concrete Using Portland Cement and Other
Cementitious Materials,” American Concrete Institute, Farmington Hills, Michigan, 2008.

Kosmatka, Steven H., Beatrix Kerkhoff, and William C. Panarese. Design and Control of Concrete
Mixtures, 14th Edition. Skokie, Illinois: Portland Cement Assn., 2008.

Reinecke, Holger. “New Technology of Water Treatment and Two-Step Mixing of Concrete.”
Concrete, Vol. 43, No. 11, Dec. 2009 / Jan. 2010.

“Standards for Mixing Water in Concrete.” National Ready Mixed Concrete Association, Silver
Spring, Maryland, 2010. <www.nrmca.org/file.asp?F=BA224C37CD24413D865D4165140408
CE.pdf&N=Standards+for+Mixing+Water+in+Concrete.pdf&C=documents>
CHEMICAL
ADMIXTURES

CONCRETE ADMIXTURES IMPROVE THE BEHAVIOR of concrete in a variety of


ways and under a variety of conditions. In North America and Europe, admixtures have
proven to lessen construction costs by doing the following: enhancing workability and
allowing for rapid set times; increasing compressive and flexural strengths; providing cer-
tain tensile behaviors; making concrete more resistant to cracking; increasing the duration
of workability of fresh concrete; helping ensure the quality of concrete in transit; and,
when desired, retarding the set times during curing.
The main types of mineral admixtures—which increase pozzolanic activity and thus
add strength to the concrete—are silica fume (microsilica), fly ash, and slag. These mineral
admixtures are often referred to as cement supplements or as pozzolan enhancements and
are addressed in the next chapter, Mix Design.
This chapter addresses the main types of chemical admixtures: retarders, accelerators,
plasticizers, water reducers, air entrainment, and specialty admixtures. Each type of
admixture will be discussed generally, with an eye toward its potential uses and possible
drawbacks for concrete pavements. For a designation of the specific actual chemical and
other components of the admixtures, see Figure 7-1. Most admixtures must conform to
American Society for Testing and Materials (ASTM) and American Association of State
Highway And Transportation Officials (AASHTO) standards, and the appropriate stan-
dards are listed in Figure 7-1. Also briefly described there are some less common
admixtures.
Admixtures are meant to improve the performance of concrete that is correctly con-
stituted, following the practices discussed in earlier chapters. Admixtures will not rectify
any deficiencies in the basic concrete mix constituents: No matter which admixtures are
utilized, it is always important to use suitable aggregates, the right amount of clean and/or
treated water, and the right amount and types of cements.
Before using an admixture, one needs careful consultation with the manufacturer to
ensure that it is appropriate. Destructive testing in the laboratory, using the actual constit-
uents, is highly recommended to anticipate any particular admixture’s effect on any par-

78
CHE M I C AL ADM I X T U R ES 79

Figure 7-1

Co n c r e t e Ad m ixt u r e sb y Cla sifi c a t io n

Type of Admixture Desired Effect Material

Accelerators Accelerate setting and Calcium chloride (ASTM D98 and AASHTO M 144)
(ASTM C494 and early-strength development Triethanolamine, sodium thiocyanate, calcium formate,
AASHTO M 194, Type C) calcium nitrite, calcium nitrate

Air detrainers Decrease air content Tributyl phosphate, dibutyl phthalate, octyl alcohol, water-
insoluble esters of carbonic and boric acid, silicones

Air-entraining admixtures Improve durability in freeze- Salts of wood resins (Vinsol resin), some synthetic detergents,
(ASTM C260 and thaw, de-icer, sulfate, and salts of sulfonated lignin, salts of petroleum acids, salts of
AASHTO M 154) alkali-reactive environments, proteinaceous material, fatty and resinous acids and their salts,
improve workability alkylbenzene sulfonates, salts of sulfonated hydrocarbons

Alkali-aggregate reactivity Reduce alkali-aggregate Barium salts, lithium nitrate, lithium carbonate,
inhibitors reactivity expansion lithium hydroxide

Antiwashout admixtures Cohesive concrete for Cellulose, acrylic polymer


underwater placements

Bonding admixtures Increase bond strength Polyvinyl chloride, polyvinyl acetate, acrylics, butadiene-styrene
copolymers

Coloring admixtures Colored concrete Modified carbon black, iron oxide, phthalocyanine, umber,
(ASTM C979) chromium oxide, titanium oxide, cobalt blue

Corrosion inhibitors Reduce steel corrosion activity Calcium nitrite, sodium nitrite, sodium benzoate, certain
(ASTM C1582) in a chloride-laden environment phosphates or fluosilicates, fluoaluminates, ester amines

Dampproofing admixtures Retard moisture penetration Soaps of calcium or ammonium stearate or oleate, butyl
into dry concrete stearate, petroleum products

Foaming agents Produce lightweight, foamed Cationic and anionic surfactants, hydrolized protein
concrete with low density

Fungicides, germicides, Inhibit or control bacterial and Polyhalogenated phenols, dieldrin emulsions, copper
and insecticides fungal growth compounds

Gas formers Cause expansion before setting Aluminum powder

Grouting admixtures Adjust and grout properties for See Air-entraining admixtures, Accelerators, Retarders,
specific applications and Water reducers

Hydration control Suspend and reactivate cement Carboxylic acids, phosphorus-containing organic acid salts
admixtures hydration with stabilizer and
activator

Permeability reducers Decrease permeability Latex, calcium stearate

Pumping acids Improve pumpability Organic and synthetic polymers, organic flocculents, organic
emulsions of paraffin, coal tar, asphalt, acrylics, bentonite and
pyrogenic silicas, hydrated lime (ASTM C141)

Retarding admixtures Retard setting time Lignin, borax, sugars, tartaric acid and salts
(ASTM C494 and
AASHTO M 194, Type B)

Shrinkage reducers Reduce drying shrinkage Polyoxyalkylene alkyl ether, propylene glycol

Superplasticizers Increase flowability of concrete, Sulfonated melamine formaldehyde condensates, sulfonated


(ASTM C1017, Type 1) reduce water/cement ratio naphthalene formaldehyde condensates, lignosulfonates,
polycarboxylates
Source: Kosmatka, Kerkhoff, and Panarese, 2008
CHE M I C AL ADM I X T U R ES 80

ticular mix design. It must be understood how the addition of an admixture will perform
at the applied temperature and at the humidity at which the mixture will be used. It is vital
to use the manufacturer’s recommended amount of admixture, because under- or over-
dosing of admixtures can create problems, and can even result in achieving the opposite
of the effects desired [concrete network].

S et R etarders
Retarding admixtures slow the rate of setting in order to allow enough time for placing,
vibrating, and finishing. Retarders are commonly used (1) if the concrete will spend longer
periods of time in transport; (2) when the weather conditions are hot, and in particular hot
and dry, which greatly accelerates the setting time; or (3) if the pour is very large.
Concrete treated with a set retarder will produce more bleed water, and for a longer
period of time, than standard concrete. For this reason, if the weather is not hot and there
are sufficient pavement placing and finishing machines, adding a set retarder could result
in down time while waiting for the bleed water to evaporate. Of particular interest here is
the finding that a two-stage (2-stage) mixing process can greatly decrease bleeding. (For
more information, see the chapter on Mixing Concrete.)
Retarders generally will cause the concrete to cure more slowly with less early strength.
In some cases, a retarder may also cause shrinkage of the concrete. A test pour made seven
(7) days ahead of time is recommended to ensure a complete understanding of how the
retarding agent will affect the in situ pavement pour structure.

S et Acceleration
Accelerating admixtures are used to speed up mix hydration. In accelerated pavements,
bleed water will disappear sooner, allowing finishing work to begin more quickly. An
accelerator also causes the concrete to gain strength sooner, so it comes closer to attaining
full strength in the first few days. It must be understood, however, that there may be long-
term strength trade-offs. In most, but not all cases, the initial rapid strength gain comes at
a price of long-term strength: An accelerated mix will not achieve the long-term strength
that a given mix design would otherwise have [purdue].
Most often, set accelerators are utilized in careful doses when the weather condi-
tions are cool (but not so cold that the concrete might be subject to unwanted freezing).
Accelerators are sometimes used in highway construction to shorten the wait time
before the road can be driven on. This often applies to repair work and fast-track lane
development.
In the North American paving market, calcium chloride has historically been used to
speed up hydration. It is a known fact that chloride is an enemy of steel-reinforced con-
crete, and in higher concentrations it has adverse effects. It has been shown to greatly
increase the rate of corrosion of metal reinforcement in areas where de-icing chemicals
CHE M I C AL ADM I X T U R ES 81

and sodium are applied to pavements. It often increases shrinkage due to surface overdry-
ing, and it can cause surface scaling and spalling. It may also darken the color of the
concrete.
As a result, calcium chloride accelerating agents are used less often in North America,
and in most European jurisdictions the use of calcium chloride in fresh concrete is forbid-
den. More aggressive designs are gaining favor, including designs with lower water/cement
ratios combined with in-ground heating of pavement beds. Hydration by using Type III
or Type HE cement, by reducing the amount of water in the mix, or by using non-calcium-
chloride accelerators are other popular
alternatives. Ac c e le r a t o r s W i l l No t St o p Fr e e zi n g
Overdosing with accelerators is a com-
mon problem in the North American mar- An accelerator is often used when the weather is cold,
ketplace. Too much calcium chloride, in and sometimes when freezing is a possibility. However,
particular, not only exacerbates the prob- an accelerator is not an antifreeze: If the water in the
concrete reaches freezing temperatures, it will freeze,
lems mentioned, it can cause the concrete even if an accelerator has been added. Rather than
to set very quickly, allowing insufficient relying on an accelerator, one should take steps to pre-
time for proper finishing [rixom and vent freezing.
mailvaganam, 1999, p. 312].

Water R educers
Water-reducing admixtures lessen the amount of water needed in the mix, and so
strengthen the concrete. They also are used to reduce the cement content and to increase
slump (and flowability) in some cases. They are so commonly used that they are some-
times referred to as the “fifth ingredient” in concrete pavements.
Water reducers fall into one of three categories: Conventional (low-range) water
reducers, mid-range water reducers, and high-range water reducers. (See the box “Types
of Water Reducers” for more information.)
Water reducers can decrease the amount of needed water by 5% to 10%. Or, if the
water content is kept constant, a water reducer will significantly increase slump. A water
reducer can also reduce the amount of cement that is needed, which lowers costs.
Tests have shown that adding a water reducer and using less water can increase the 28-
day strength of concrete by as much as 25%. If water amounts stay the same, a water
reducer can double the slump loss.
Even though they decrease the need for water, water reducers can actually increase the
amount of shrinkage—but usually not significantly.
Depending on the chemical used, a water reducer will lessen bleed water, increase it,
or cause no significant change. Similarly, the type of reducers chosen in the North Ameri-
can and European markets have shown to lessen setting time in some cases.
Some water reducers have air-entraining properties, but they usually cannot take the
place of actual air-entrainment admixtures.
CHE M I C AL ADM I X T U R ES 82

Ty p e s o f W a t e r Re d u c e r s

Water reducers come as conventional (or low-range), particularly if the mixture contains mineral
mid-range, and high-range (also called superplas- admixtures like silica fume.
ticizing) types.
High-range water reducers, also called plasticizers
Conventional (low-range) water reducers typically or superplasticizers, achieve a water reduction of
achieve a 5% water reduction or a 25 to 50 mm between 12% and 40%, or a slump reduction of
slump reduction. 300 to 1,000 mm. They can greatly accelerate
early strength gain and increase the final com-
Mid-range water reducers achieve 6% to 15%
pressive strength. They produce larger air bubbles
water reduction or a 100 to 125 mm slump
than those produced by air entrainment, which
reduction. They slow the rate of setting, and
leads to good resistance to cracking during freeze-
offer similar setting rates over a wide range of
thaw cycles [CONCRETE NETWORK].
temperatures, so concrete does not set too
quickly and remains workable. They are often However, high-range reducers also greatly reduce
used in hot weather, to counteract the acceler- bleed water, and so are used only where finishing
ating effect of high temperatures. They make can be accomplished quickly. Also, they may
the mixture easier to pump, place, and finish, significantly increase shrinkage.

Plasticiz ers
The terms plasticizer and superplasticizer are used interchangeably to refer to high-range
water reducers. They allow a low water/cement ratio, which adds to the performance of
concrete. In Europe and in some select areas of the U.S., research on the plasticizers in
concrete pavements is ongoing. Generally, however, they are used to allow acceleration in
pavement mixes with ultra-low water/cement ratios, rather than for workability.
Plasticizers increase the strength of a mix and promote a higher slump. This makes it
easier to place, move, and finish concrete. It also reduces the need for vibrating, since the
more-liquid concrete leads to fewer air pockets at forms and promotes greater adhesion to
metal reinforcement.
Plasticizers can reduce handling costs. In North America and Europe, low and mid-
range water-reducing agents are deemed appropriate where metal reinforcement is congested,
or when it is necessary to pump the concrete over long distances, such as on bridge decks.
The increased slump/flowability offered by plasticizers will typically last for no more
than an hour, unless mix-water enhancements are undertaken. Once the time window
closes, the concrete will harden rapidly. If temperatures are high, the loss of flowability
may occur even sooner. Plasticizers known as Type F (water-reducing) plasticizers last
for a shorter period than those known as Type G (water-reducing and -retarding). If the
transport distance from the plant to the highway is short, a Type G plasticizer may be
added at the plant, but usually the plasticizer is added to the batch shortly prior to pouring
[FHwa, 1999].
In some cases, plasticized concrete has more bleed water than non-plasticized con-
crete. It will have less bleed water, however, than concrete that has achieved the same
CHE M I C AL ADM I X T U R ES 83

slump simply by adding more water. Some tests have shown that plasticized concrete has
a lower resistance to scaling in cold weather, but other tests have shown no significant
drop-off in cold-weather performance [Rixom and Mailvaganam, 1999, p. 315].

Air E ntrainment
Air-entrainment admixtures cause tiny bubbles to form in the concrete mix. These bub-
bles greatly improve resistance to cracking during freeze-thaw cycles. They also help the
concrete resist the surface scaling that often occurs when de-icers are applied to highway
surfaces. Air entrainment enhances workability of the concrete, and it reduces bleed water
somewhat.
Air entrainment can be added using an air-entrainment admixture, by using air-
entraining cement, or both.
Air-entrainment bubbles are microscopic in size, only 10 to 100 µm in diameter. These
bubbles relieve pressure caused by water that has infiltrated the concrete and then has fro-
zen and expanded. To successfully resist cracking during freezing, it is important that the
bubbles be of the right size, be evenly dispersed, and not be attached to each other. A good
air-entrainment admixture will provide the right bubbles in the right places. Air entrain-
ment also makes concrete easier to work with. Air-entrained concrete requires less water;
it is less sticky; it produces less bleed water; and makes it easier to get a good hard finish.
Air entrainment can be compromised by excessive vibration or by adding water to
increase the slump. Adding superplasticizers to increase slump, however, does not appear
to lessen the air-entrained concrete’s resistance to cracking during freeze-thaw cycle, prob-
ably because they reduce water content.
When mixing is done in cold temperatures using hot water, the air-entrainment
admixture may lose some effectiveness. The air-entrainment admixture should be added
after the concrete has achieved normal temperature.
Resistance to cracking during freeze-thaw cycles can also be increased by using low
water/cement ratio; aggregate of the right types and sizes; sufficient cement of the right
types; and application a good hard finish to the concrete surface [Kosmatka, Kerkhoff,
and Panarese, 2008, pp. 129–148].

S p ecialty A dmixt ures


Here are some specific problems experienced in the North American and European mar-
kets, as well as the admixtures and practices listed by the Concrete Network that are com-
monly used to help alleviate the problems:

CORROSION PROTECTION. During the winter, chlorides may be applied to a highway for
de-icing purposes. Also, in a marine environment, salt may continually attack the con-
crete. In these situations, reinforcing metal is susceptible to rust and other corrosion.
CHE M I C AL ADM I X T U R ES 84

The first solution is to apply a concrete mix that is durable and has low permeability.
All metal reinforcement is encased in the concrete by at least 50 mm. Adding silica fume
to the mix can reduce permeability, thereby protecting the metal from corrosion.
Epoxy-coated rebar resists corrosion much better than standard metal, but it is
expensive.
Some North American and almost Co ld -a n d H o t -W e a t h e r Co n c r e t e Pl a c e me n t
all European contractors use corrosion-
inhibiting admixtures (see Figure 7-1).
If the ambient temperature is lowered by 11°C, it can
Calcium chloride should be avoided double the time concrete takes to set. This is usually
as a set accelerator, even when specifi- addressed by adding an accelerator, and perhaps a water
reducer, to the mix.
cations allow it. It will attack the metal,
In hot weather, the opposite problem often occurs—
much as other salts do. the mix can set too quickly for proper finishing. The problem
is addressed with a set retarder, water reducer, or plasti-
CONTROL OF CRACKING DUE TO CUR- cizer (superplasticizer), or a combination of thereof.

ING. During the hydration process, the


cement paste shrinks as it loses water.
In some cases this can lead to cracking, depending on the types of cement and aggregate.
Pouring of test batches of the materials to be used is generally required to see whether
cracking may be a problem.
A shrinkage-reducing admixture (SRA) lessens the shrinkage and decreases the effects
of drying, thereby solving most cracking problems that are due to curing.

FINISHING ENHANCERS. If it is difficult to produce a hard finish with a given mixture, a


mid-range water-reducing admixture will make the concrete more workable. Other admix-
tures can be added to either increase or decrease bleed water and set time, as required.

Endnotes
Concrete Admixtures. Concrete Network, Calimesa, California, 1999–2009. <www.concretenet-
work.com/concrete/concrete_admixtures/>

Concrete Admixtures – Corrosion Protection. Concrete Network, Calimesa, California, 1999–2009.


<www.concretenetwork.com/concrete/concrete_admixtures/corrosion_protection.htm>

Concrete Admixtures – Finishing Enhancers. Concrete Network, Calimesa, California, 1999–2009.


<www.concretenetwork.com/concrete/concrete_admixtures/finish_enhancement.htm>

Concrete Admixtures That Control Cracking – Reducing Drying or Shrinkage Cracking. Concrete
Network, Calimesa, California, 1999–2009. <www.concretenetwork.com/concrete/concrete_
admixtures/crack_control.htm>

Kosmatka, Steven H., Beatrix Kerkhoff, and William C. Panarese. Design and Control of Concrete
Mixtures, 14th Edition (2002). Skokie, Illinois: Portland Cement Assn., 2008.
CHE M I C AL ADM I X T U R ES 85

Rapid-1 Hardening Accelerator Concrete Admixture. Purdue University, Division of Construction


Engineering and Management, West Lafayette, Indiana, 2009. <http://rebar.ecn.purdue.edu/
ECT/links/technologies/civil/rapid1.aspx>

Rixom, Roger, and Noel Mailvaganam. Chemical Admixtures for Concrete. London:
E. & F. N. Spon, 1999.

Superplasticizers. U.S. Dept. of Transportation – Federal Highway Administration, 1999. <www.


fhwa.dot.gov/infrastructure/materialsgrp/suprplz.htm>

Water Reduction in the Mix. Concrete Network, Calimesa, California, 1999–2009.


<www.concretenetwork.com/concrete/concrete_admixtures/water_reduction.htm>
MIX DESIGNS

THE INGREDIENTS USED TO MAKE CONCRETE, and the ratios in which they are
used, affect the quality of the concrete. The formula used to create a batch of concrete is
known as its mix design, and the goal of any mix design is to produce the best possible
product at the most economical price. The actual process, described later in this chapter,
is straightforward and has been made even simpler by the development of computer soft-
ware. Before looking at the process and how it varies from country to country, however, it
is important to look at the variables affected by mix design. They include strength, work-
ability, durability, and the tendency of the pavement to shrink and crack.

Strength
Strength is currently measured in compressiv e and flexural terms. While strength is the
most universal measure of concrete quality, related qualities may be equally important:
Durability, permeability, and wear resistance are especially critical when considering the
life-cycle design of a project [KOSMATKA, KERKHOFF, AND PANARESE, 2008, P. 149]. Several
factors—emulsification mixing (wet-grinding); the advent of advanced electrical water
treatment; advanced mix design development by computer-assisted technology and man-
agement; and advancements in admixtures—are allowing for more composite-like con-
crete products, which actually provide tensile strengths normally expected only in
prestressed concrete elements. Advancements in concrete mix designs are poised to
encroach on existing tensile strengths, which will allow concrete to have a far better life-
time under loads and in settlement and freeze-thaw environments. For now, however,
concrete is generally to be gauged for strength by compressive performance, and thereafter
by flexural performance.
The desired compressive strength is used to arrive at the necessary water/cement or
water/cementitious materials ratio [DELATTE, 2008, P. 114]. For fully compacted concrete
made with clean, sound aggregates, the compressive strength under given job conditions
is governed by the quantity of mixing water used per unit of cement or cementitious mate-

86
M I X DE S IG NS 87

rials. Typically, the compressive strength is inversely related to these ratios [KOSMATKA,
KERKHOFF, AND PANARESE, 2008, PP. 149–150]. If flexural strength is to be the main deter-
minant (as is practiced in some states), laboratory tests of design materials will be needed
[DELATTE, 2008, P. 115].
Strength is largely dependent on the degree to which the cement has hydrated. Over
time, concrete will become stronger to the extent that both moisture level and temperature
allow. In some cases, however, enough water may evaporate to create voids that compro-
mise both compressive and flexural qualities. Other factors in determining concrete
strength, given a certain water/cementitious materials ratio, are the following:

N Grading, texture, shape, and strength of the aggregate


N Qualities of the cementing materials
N Content of entrained air
N Selection of admixtures
N Duration of curing
[KOSMATKA, KERKHOFF, AND PANARESE, 2008, P. 150]

The cementitious materials content is usually determined by the water/cementitious mate-


rials ratio, although a minimum cement content is frequently included in specifications,
serving to enhance durability and finishability and to improve wear resistance of slabs.
These qualities are important even though strength requirements may be achieved with
smaller proportion of cementitious materials. On the other hand, high levels of cementi-
tious materials may sacrifice economy and also compromise workability [KOSMATKA,
KERKHOFF, AND PANARESE, 2008, P. 156].

Workability
Workability is the measure of how readily the mix can be placed, consolidated, and fin-
ished without segregation. It is influenced by temperatures at the site, and of course by
having the necessary workers available over the duration of the process. Workability is
typically determined by a slump test to indicate the concrete’s consistency and ability to
flow [WASHINGTON STATE D.O.T.]. As the proportion of aggregate increases, and as less
water is used, the concrete will be stiffer and less plastic. Too much water, on the other
hand, results in a thin mix that is difficult to work [KOSMATKA, KERKHOFF, AND PANARESE,
2008, P. 153].

Durability
Durability is the measure of the concrete’s performance over the desired service life [WASH-
INGTON STATE D.O.T.].
M I X DE S IG NS 88

Shrinkage and Cracking


While the cracking that occurs with shrinking is controlled to some extent by using joints
and reinforcing steel, cracking can be controlled by paying particular attention to the mix
design [WASHINGTON STATE D.O.T.]. Shrinking also can be minimized by using less water
in the mix. The need for water is generally reduced by controlling several variables:

N Lowering the water/cementing materials ratio


N Reducing the proportion of paste by specifying large, rounded coarse aggregate
N Entraining more air
N Adding water-reducing admixtures
N Adding fly ash
[KOSMATKA, KERKHOFF, AND PANARESE 2008, PP. 150–152]
[DELATTE, 2008, PP. 100, 121]

The key water/cementitious material ratio is calculated by dividing the mass of water by
the mass of cementitious material (including cement, fly ash, slag, silica fume, and natural
pozzolans). This ratio should be the lowest value needed to handle the exposure condi-
tions of the site.

Freeze-Thaw Concerns
Thermal deformation can be prevented somewhat by incorporating an aggregate rated
with a low coefficient of thermal expansion [DELATTE, 2008, P. 121]. Entrained air further
helps prevents damage from freeze-thaw cycles and from de-icing chemicals. In the short
run, it can also improve workability.

Admixtures
When using two or more admixtures, their compatibility may become an issue, and it is
best to confirm their performance in small test batches. Note that the water content of
admixtures should be considered part of the mixing water if the admixture’s water content
is sufficient to affect the water/cementitious materials ratio by 0.01 or more [KOSMATKA,
KERKHOFF, AND PANARESE, 2008, P. 157].

North American Standards


The standard ACI mix design procedure can be described as a series of several steps
[AMERICAN CONCRETE INSTITUTE, 2009, PP. 9–10].

STEP 1 • Choice of slump. Generally, the ideal is the stiffest mix that can be placed
efficiently.
M I X DE S IG NS 89

STEP 2 • Choice of aggregate size. Concrete with large nominal maximum sizes will
require less mortar because it has fewer voids. So, it is best to incorporate the largest
size appropriate to the particular project, within certain limits. The aggregates should
be no larger than one-fifth (1/5) of the narrowest dimension between the form’s sides.
It should be no more than one-third (1/3) the depth of slabs; and no more than three-
fourths (3/4) the minimum clear spacing between reinforcing bars, bundles of bars, or
pre-tensioning strands. A nominal maximum size of aggregate may need to be chosen
if reinforcing steel, post-tension ducts, or conduits might otherwise make it difficult
to place concrete without excessive segregation, pockets, or voids. If necessary, smaller
nominal maximum sizes of aggregate can be used to yield concrete with greater
strength at a given water/cement ratio.

STEP 3 • Estimating mixing water and air. The quantity of water per unit volume of con-
crete required to produce a given slump is dependent on several factors: the nominal
maximum size, particle shape, and grading of the aggregates; the concrete tempera-
ture; the amount of entrained air; and the use of chemical admixtures. Slump is not
greatly affected by the quantity of cement or cementitious materials within normal use
levels (under favorable circumstances, the use of some finely divided mineral admix-
tures may lower water requirements slightly). Depending on aggregate texture and
shape, mixing water requirements may be somewhat above or below values given in
mix-design tables, but they should serve for an initial estimate. The differences in
water demand are not necessarily reflected in strength since other compensating fac-
tors may be involved. Particle shape is not necessarily a determinant of an aggregate’s
strength-producing contribution, with roughly equivalent results from both rounded
and angular coarse aggregates of equivalent quality and grading.

STEP 4 • Choosing chemical admixtures. Admixtures are used to bring about any of sev-
eral improvements in a mix. They can make concrete more workable, durable, and
economical. They are added to either reduce or lengthen the time of set, as well as
accelerating strength gain and controlling temperature gain. And they can substan-
tially reduce the necessary proportion of water. These chemicals should be added only
after determining how they will perform under the anticipated conditions. Also, it is
necessary to include the liquid contribution of admixtures when calculating the vol-
ume of mixing water.

STEP 5 • Determining water/cement or water/cementitious materials ratio. The required


ratio is arrived at through the strength and durability requirements, taking into con-
sideration the effect of the mix’s particular blend of aggregates, cements, and cementi-
tious materials.

STEP 6 • Determining the relationship between water-cement or water-cementitious mate-


rials ratio and compressive strength of concrete. If concrete is to be subjected to severe
M I X DE S IG NS 90

conditions, these ratios should be kept low even though a higher value may satisfy
strength requirements. When using pozzolanic materials (such as fly ash, ground
granulated blast-furnace [GGBF] slag, and silica fume), it is necessary to employ a ratio
of water-to-cement, plus pozzolanic materials (or water-to-cement, plus other cementi-
tious materials) by weight, rather than going by the usual water/cement ratio.

STEP 7 • Calculation of cement content. The required cement is equal to the estimated
content of mixing water content divided by the water/cement ratio. If, however, the
specification includes a separate minimum limit on cement in addition to require-
ments for strength and durability, the mixture must be based on whichever criterion
leads to the larger amount of cement. The use of pozzolanic or chemical admixtures
will affect properties of both the fresh and hardened concrete.

STEP 8 • Estimating coarse aggregate content. Aggregates of similar nominal maximum


size and grading will yield good workability if a given volume of coarse aggregate, on
an oven-dry-rodded basis, is used per unit volume of concrete. For equal workability,
the volume of coarse aggregate in a unit volume of concrete is dependent only on its
nominal maximum size and the fineness modulus of the fine aggregate. Differences in
the amount of mortar required for workability with different aggregates, due to differ-
ences in particle shape and grading, are compensated for automatically by differences
in oven-dry-rodded void content.

STEP 9 • Estimation of fine aggregate content. The quantity of fine aggregate can be
arrived at by either the weight method or the absolute volume method. Given an esti-
mate of the weight of the concrete per unit volume, the required weight of fine aggre-
gate is the difference between the weight of fresh concrete and that of the other
ingredients combined. Approximate values are available in mix design tables, but
amounts should be refined by making trial batches. For a more precise calculation of
the fine aggregate needed, the volume displaced by the known ingredients—water, air,
cementitious materials, and coarse aggregate—is subtracted from the unit volume of
concrete. (The volume of any ingredient in concrete is equal to its weight divided by
the density of that material, which is the product of the unit weight of water and the
specific gravity of the material.)

STEP 10 • Adjusting for aggregate moisture. When weighing aggregates, their moisture
content must be taken into consideration. They typically can be assumed to contain
moisture; their dry weights therefore are increased by the percentage of their surface
and absorbed water. The mixing water for the batch is then reduced by an amount
equal to the free moisture introduced by the aggregate, which is to say the total mois-
ture minus the absorbed figure. If the absorption (normally measured by soaking 24
hours) is higher than approximately 1%, and if the pore structure within aggregate
particles is such that a significant fraction of the absorption occurs during the time
M I X DE S IG NS 91

prior to initial set, there may be a noticeable increase in the rate of slump loss due to
an effective decrease in mixing water. Also, the effective water/cement ratio would be
decreased for any water absorbed by the aggregate prior to set; this, of course, assumes
that cement particles are not carried into aggregate particle pores.

STEP 11 • Making trial batch adjustments. The calculated mixture proportions should
be checked by means of trial batches or full-sized field batches. Only sufficient water
should be used to produce the required slump, regardless of the amount assumed in
selecting the trial proportions. The concrete should be checked for unit weight and
yield and for air content. It should also be carefully observed for proper workability,
freedom from segregation, and finishing properties. Appropriate adjustments should
be made in the proportions for subsequent batches.

Eu r o p e a n St a n d a r d s

In Europe, it is recognized that the stresses gener- as yet been adopted as a harmonized standard.
ated by traffic mean the requirements for concrete Generally, there are more exacting requirements
pavement mix design differ from that for concrete for aggregates used in road construction than those
used in other applications. However, individual cli- for buildings and other structures. There is a cap
ents are given the responsibility of factoring in the placed on the loss of mass in freeze-thaw resis-
nature of the project and the environmental condi- tance testing, a limit on allowable lightweight
tions at the site. organic contaminants, requirements for shape and
flakiness index, and another for polished stone
Aggregates play an important role in concrete
values. Procedures are in place for minimizing
pavements, in addition to the structural role they
alkali-silica reactions as well.
have in any type of concrete. They must perform
well at load-transfer joints and maintain their skid- In Germany, concrete pavement contractors are
resistance properties over time. In Europe, the responsible for mix design. Mixes typically are not
maximum aggregate size is generally considered to regarded as proprietary, unlike those for cement
be one-quarter (1/4) of the thickness, or a third products. In Belgium, concrete pavements are
(1/3) of the distance between reinforcing bars. either Portland cement or blast-furnace slag cement
with reduced alkali in order to minimize alkali-
Because practice varies from one country to
aggregate reaction. In typical practice, durability
another, there is no one test for determining com-
and strength are ensured with a high proportion of
pressive, flexural, and tensile-splitting strength. In
cement, low water/cement ratios, and air-entraining
France and Spain, for example, concrete strength
agents. In Austria, contractors are responsible for
is monitored with lab samples, while in Germany,
concrete mix design, and at their discretion they
cores are taken from the constructed pavement.
hire laboratories that will design the mix. Contrac-
In Europe, only the standard for cement (EN 197) tors’ mixtures are not proprietary [FHWA, LONG-LIFE
is currently enacted internationally, while those for CONCRETE PAVEMENTS, 2007].
aggregate, admixtures, and mixing water have not
M I X DE S IG NS 92

Endnotes
ACI 211-1-9 1 Standard Practice for Selecting Proportions for N ormal, Heav yw eight, and Mass
Concrete. Farmington Hills, Michigan: American Concrete Institute, 1991, reapproved 2009.

Delatte, Norbert. Concrete Pav ement Design, Construction, and Performance. London: Taylor &
Francis, 2008.

Kosmatka, Steven H., Beatrix Kerkhoff, and William C. Panarese. Design and Control of Concrete
Mixtures, 14th Edition (2002). Skokie, Illinois: Portland Cement Association, 2008.

L ong-L ife Concrete Pav ements in Europe and Canada, Chapter 5 : Cement and Concrete. U.S. Dept.
of Transportation – Federal Highway Administration, 2007. <http://international.fhwa.dot.
gov/pubs/pl07027/llcp_07_05.cfm>

“PCC – Fundamentals,” Washington State Department of Transportation Pavement Guide.


<http://training.ce.washington.edu/wsdot/Modules/05_mix_design/05-7_body.htm>
BATCHING
TECHNOLOGY

IN CONCRETE PRODUCTION FOR PAVEMENTS, it is generally accepted in Europe


and in North America that a mixture of concrete will be batched by weight and that it will
be recorded for performance and quality control. The batching of any mixture is often
undertaken in manufacturing for the same reasons, but in concrete production it is espe-
cially prevalent as concrete is delivered most often by carrier truckload or in batches by
transit mixer truck.
Notwithstanding some movement on the part of purveyors of volumetric and continu-
ous equipment system to obtain qualification and specification allowances, or other approv-
als by regulators to allow non-batch-type machinery in the production of civil or municipal
concrete pavement concrete, it is generally accepted that all concrete for pavement should
be made in a batch-type plant to allow for quality control recording by weight.

The B est Concrete Ba tching P lant T oday


There can be no disputing that if one were to design the optimum utopist concrete
plant for concrete pavement production, it would be theoretically include:

N Individual weigh batchers for individual powders and aggregates, water, and
chemical admixtures—no volumetric or metered material measurement devices
would be used
N Layering of all aggregate constituents or metered discharge of such to high-
intensity mixing machinery for a compulsory blending under extreme force
N Individual moisture monitoring of each aggregate with automated compensation
for yield and corrected free water. Better yet, all aggregates would be dried prior
to batching to a maximum moisture content of less than 2% by weight. This would
provide ultimate quality control of the water/cement ratio.
N First-stage mixing of all powdered constituents in an emulsification slurry in a
first-stage high-shear colloidal mixing pump or chamber. The batch water would

93
BAT CHI NG T E CHNOLOG Y 94

be pretreated, filtered, and polarized for optimum hydration. Once mixed, the
slurry would be pumped to a compulsory mixer for production of the concrete
under the aforementioned extreme forces.
N Maximum temperature control of mix water, aggregates, bulk powders, and any
reagents to allow controlled shrinkage, set, and behavior of curing of the fresh
concrete. The amount of aggregates needed for daily production would be kept
treated and awaiting the batching plant either heated or cooled. Ice or ice slurry,
liquid nitrogen, or liquid carbon dioxide would be available for the cooling of
concrete in the high-intensity mixer to effect a maximum fresh concrete tempera-
ture as specified. Warm aggregates and warm (not hot) water would be available
to ensure controlled placing temperatures in cold weather.
N Dense or semi-dense phase transfer of all bulk powders into storage silos sized to
provide for several days’ production of concrete. The silos would be sheltered
from the sun and fit with oversized, aspirated filtering systems to ensure abso-
lutely no dust outcome.
N Full plant automation whereby the plant operation personnel are not capable of
manipulation and possible interfering with the designed batch/mix
N A fully weatherproofed facility that would keep all materials out of the wind, rain,
snow, cold, and heat

The fact is that there are very few such paving batch plants in operation anywhere in the
world—even though many who are involved in the development of high-performance
concrete would wish otherwise.
There has been progress: Individual weighing is specified on many government proj-
ects in North America, a result of a long-standing U.S. Army Corps of Engineers specifica-
tion to avoid cumulative weighing error; high-intensity compulsory mixers are commonly
used throughout Europe; water treatment by means of electrical polarization is now pat-
ented and gaining use worldwide; emulsification first-stage mixing is obviously the best
way to ensure the utilization of available cement and effect maximum hydration; and tem-
perature balance and maximum and minimum concrete temperature values are in place
on many high-performance concrete projects. Yet we do not find the accumulative combi-
nation of all of these types of equipment and accessories—or even most of them—in com-
mon practice on paving projects in Europe or in North America.

From Theory to P ractice in Ba tching P lant T echnology


There exists a chasm between theory and practice in concrete production today. A
more scientific approach is needed, and is in fact beginning to appear. The concrete paving
plant of the near future is going to be influenced (as is concrete pavement itself) by com-
posite and other competitive materials.
BAT CHI NG T E CHNOLOG Y 95

Many new paving plants include high-intensity mixing machinery. Many new paving
plants are opting for some of the aforementioned accessory and ancillary devices and
equipment. It’s an absolute requirement that producers make the most of the new machin-
ery, and this is going to need to be specification-driven.
If concrete pavements were required to have extended life cycles well beyond today’s
requirements, the paving contractor or concrete producer would be quick to embrace new
technologies as they become available.
In Europe today, concrete paving plants are smaller, better weatherproofed, more
mobile, and better equipped to make well-graded concrete pavements. But these plants are
nowhere near the theoretical/utopist plant outlined earlier in this chapter.

The SBM Euromix mobile plant is an advanced portable plant


that’s equipped with a twin-shaft mixer. Unlike paving plant
configurations typical in North America, the European plant is
normally equipped with a 2-, 3-, or 4-cubic-meter-output mixer.
Emphasis is on the quality of the smaller batch size. The SBM
configuration (Austrian) is a single containerized trailer with
powder silos set on precast counterweight slabs.
BAT CHI NG T E CHNOLOG Y 96

In North America today, concrete paving plants are large-scale mass production installations, as the quan-
tities of concrete pavement oftentimes demand, with one (1) or two (2) 8-cubic-meter tilt-drum mixers each,
producing up to 300 cubic meters per hour. Unfortunately, the concrete produced in these large-scale paving
plants is not as good as it could be. In fact, as is noticed worldwide, concrete pavement in the United States has
an expected life cycle of less than what is already known to be possible. We put forth in this analytical study the
observation that the outdated batching and mixing technology, much of it many decades old—if not a full
century old—has a major role in the less-than-perfect concrete in North American roads and highways. This
outdated equipment, combined with certain mix-design allowances—most notably, gap-graded aggregates
(concrete made with only two [2] or three [3] sizes of aggregate) and other identified practices—leave room for
improvement in the North American concrete pavement marketplace for higher-performing concrete.
!"#$!%!&'()%$'(*&+,'-(&.,/(' 01&234 556&78

The MCT OnSite Mobile plant takes the containerized form to a


higher level. More advanced, but less portable, the OnSite plant
is better suited to large-scale paving and infrastructure projects,
where the plant may be located for a duration of months.
MIXING
CONCRETE

A HOMOGENOUS MIXTURE OF THE INGREDIENTS in concrete is essential to the


strength of the final product. If mixing is incomplete or if the materials are allowed to seg-
regate, particles within the mix will not be distributed evenly and hydration will not be
uniform. While some parts of the batch may meet the standards of the mix design, others
will not, and taken as a whole, the batch will not meet quality control standards [RUPNOW,
SCHAEFER, WANG, AND HERMANSON, 2007, P. 1].
While this may seem obvious, the European and North American mixing processes
differ widely, as does the quality of the mixes. European concrete is generally mixed in a
single- or twin-shaft, high-intensity mixer. Less than 5% of paving concrete is mixed this
way in North America, where the low-intensity, tilting-drum-type mixer is favored instead
because it handles higher volumes and uses less energy.

B a t c h v s . Co n t in u o u s Mixe r s

In North America, as in Europe, most paving con- and repair, as well as in specialized ultra-high-
crete is produced in a “central” or “pre” mixer—no strength and polymer concrete production.
actual mixing takes place in delivery trucks, even
though the trucks agitate the mix to keep it fresh. In
both Europe and North America, the preference is
for batch mixers, which mix a fixed amount of con-
crete, and unload it before mixing the next “batch.”
They allow for greater accuracy in mixing, and
overall they provide a more homogenous mix.
Lately, however, continuous mixers have been used
with good results, albeit on an infrequent and spe-
cial-case basis for concrete pavements. Continuous
mixers produce a constant stream of concrete at
A continuous mixer combines the work of the batch plant
one end while the raw materials enter the other and the delivery truck. The raw materials are placed in
end. In North America, volumetric continuous the truck, where they are combined and then delivered
mixer trucks are utilized increasingly for overlay through a chute in the rear.

97
M I X I NG C ONC R ET E 98

High-Intensity Mixing
by Opposing-Shaft Mixers
The European-type mixer is generally a double hori-
zontal-shaft mixer. The mixer shafts force the mass of
concrete within the tank back into its own mass by
means of opposing shafts that either have spiral blades
with arms and shovels, or have paddles that force the
mix together or subdivide the mass. Forcing the mass
into itself is called compulsion, and the mixers are
known as compulsory mixers [OBER, 2003, P. 5]. The paddles on an opposing-shaft mixer stir the mixture
around each shaft, and then force the mix against itself in the
In some cases, a higher-shear-type mixer is used, area between the shafts.
such as a planetary pan or drag-pan mixer. (Hydraulic
shear is the force created at the seam between adjoining areas of a liquid traveling at dif-
ferent speeds.) The use of pan or planetary mixers for concrete pavements is not generally
favored, however, as high-shear mixing creates increased wear and results in higher main-
tenance costs. Pan and planetary-pan mixers also are not available in the sizes required for
larger paving operations.
The mixing action of the opposing shaft mixer is low-shear, but is considered high-
intensity because the ingredients are mixed in three (3) dimensions. As a result, a homo-
geneous mix can be obtained within a very short period of time [OBER, 2003, P. 5]. In
recent years manufacturers have begun to produce larger-capacity opposing-shaft mixers
that match the higher volume of the typical North American drum mixer. While the power
drawn by the larger-capacity opposing-shaft mixers is much greater than that drawn by
the low-intensity drum mixer, an opposing-shaft mixer consumes fewer kilowatts per
batch. This is due to two (2) things: the shorter mixing time required and the fact that once
compulsion is obtained, it takes very little energy to drive the material downward on the
mixing tools. Overall electrical power consumption can be lower by as much as 50% per
year [OBER, CONCRETE PRODUCTS, 1999].

Low-Intensity Drum Mixing


The type of mixer generally employed in paving-concrete plants throughout North Amer-
ica is the low-intensity, tilting-drum-type mixer. This type of mixer was originally designed
in the 19th century and remains practically unchanged since that time. It has remained
popular because it is capable of mixing a large quantity of concrete (up to 9 cubic meters)
in a single drum with very low kilowatt input.
In a typical tilting mixer, the drum rotates on its axis, and is tilted forward to discharge
the mixed concrete. (In some cases the mixer reverses rotation to discharge.) Mixing
occurs as fins or blades on the inside drum wall pick up and loft the materials. The mix
moves up the inside wall of the drum, then falls back into the bottom of the drum. This
M I X I NG C ONC R ET E 99

results in a two-dimensional (2-D), low-intensity mix—the fins mix


materials in one (1) dimension, and the falling ingredients mix them-
selves together in a second dimension. It takes very little energy to
operate the drum, but the total kilowatts required to thoroughly
mix and discharge the concrete is actually greater than in opposing
horizontal-shaft mixers, as the mixture in a drum-type is constantly
being re-lofted and the discharge of the mixture is completed under
Drum mixers use cupped fins to loft the
full load conditions until almost empty [OBER, 2003, P. 1].
mixture up the side of the drum. The
Unfortunately, the segregation of the falling materials when they contents are mixed further as they fall back
to the bottom of the drum, but segregate
hit the bottom of the drum presents a significant problem. The low- when they land. Some are fed from the
intensity drum-type mixer is so poor at providing a homogenous rear, with a second opening on the front
end; some are fed from the same end
mixture that the North American uniformity quality-control test is as that which discharges (like a bowl or
written to allow an obviously inferior type of mixer to produce pitcher).

concrete.
The generally accepted mixer uniformity test is the U.S. Army Corps of Engineers
(USACE) Mixer Uniformity Testing Procedure CRD-C 55-92. The test requires that a
mixer full of the concrete be sampled three (3) times: at the beginning of discharge, at a
point midway through the discharge, and at the end of discharge. The three (3) samples are
compared for a variety of attributes [USACE, 1992, PP. 2–4]:

N Spread or slump
N Air entrainment
N Uniform density and specific gravity
N Aggregate composition, which is tested and compared by washing out the cement,
leaving only the fine and coarse aggregates. This aggregate is weighed and sieved,
and the size distribution and gradient differentials of the aggregates are compared.

The test is extremely accurate in assessing the homogeneity of a mixer’s batch of concrete,
and the tilting-drum-type mixer cannot pass the test if the samples are taken at the very
beginning, exact middle, and very end of the batch. By the very nature of the segregation
occurring in a drum mixer, the very last of the mixture will contain inordinately more
coarse aggregate than earlier samples. The spread and slump values will also differ, as will
hardened samples of the mixtures.
To overcome this, the test allows the first sample to be taken anywhere within the first
third of the discharge rather than requiring it at the very beginning. The second sample
can be taken within the first part of the second third, and the last sample is taken at the
very first of the last third. Such molestation of the test occurs as a rule on almost every
uniformity test taken on a tilting-drum mixer.
By contrast, fewer variations are observed in a single-shaft mixer, and in a double-
shaft opposing-type high-intensity mixer, the results are much better, even when no spe-
cial care is taken to avoid the very first or the very last of a mixer load of concrete.
M I X I NG C ONC R ET E 100

Low-intensity drum-mixed concrete has been proven over and over to be inferior to
high-intensity mixed concrete. Low-intensity mixing equipment is forbidden by German
specifications, and is not even attempted in most European markets. Yet many specifica-
tions in North America not only allow the low-intensity drum-type mixer, but actually
specify it as the mixer of choice. This discrepancy is clearly a major difference between the
European and the North American methodology. Unfortunately, low-intensity mixing,
combined with gap-graded aggregate, has been responsible for millions of cubic meters of
inferior concrete poured within the North American infrastructure.

Sin g le -Sh a ft Mixe r s

In both Europe and North America, some single- While they mix far more thoroughly than drum-type
shaft mixers are utilized in concrete paving oper- mixers, they provide much lower-intensity mixing
ations. Developed in an attempt to improve on than do opposing-horizontal-shaft mixers. While
drum-type mixers, they are basically a drum-type they avoid the segregation caused by drum-type
mixer turned inside out. Blades or spiral fins rotate mixers, they do not force or “compel” the mixture
on an axle while the drum remains stationary. They to collide into its mass. With no forced mixing
mix in a gentle manner, rolling and folding the action, the single-shaft mixer provides a midpoint
mass. between drum and compulsion mixers.

One-Stage vs. Two-Stage Mixing


Batch mixing is usually a one-stage (1-stage) process. While ingredients may be added in a
specific order, mixing is completed in a single cycle. With the trend in mix design to toward
finer cementitious materials and low water-to-binder ratios, however, the cement powder
can form lumps (agglomerate), resulting in uneven hydration and reduced workability.
Two-stage (2-stage) mixing addresses this by first mixing the cement, water, and
admixtures together to form a uniform paste. The paste is then added to the fine and
coarse aggregates, and mixing continues until the aggregates are coated.
A study by the National Concrete Pavement Technology Center found significant advan-
tages to two-stage mixing [RUPNOW, SCHAEFER, WANG, AND HERMANSON, 2007, P. 77]:

N Two-stage mixing can significantly improve concrete uniformity.


N Due to increased mixing time (from the time the cement contacts water to the
end of mixing), two-stage mixed concrete generally shows a reduced slump.
N Two-stage mixing may increase concrete strength 5%–10% over conventionally
mixed concrete. Laboratory results show an 8%–10% increase, field results show a
5%–10% increase, and a literature review shows a 10%–20% increase.

The study did find, however, that for a given air-entrainment mixture, two-stage mixing
results in lower air content than concrete mixed in a single stage.
M I X I NG C ONC R ET E 101

Colloidal Mixing
The oil-and gas-drilling industry has utilized high-shear pumps and other types of colloi-
dal mixing machinery for years to finely wet-grind powdered fines and emulsify them into
a semisolid cement. The mix is used to pour down-hole drill casements and in geological
fracture repair.
Commonly referred to as “well cementing,” this type of the emulsification is being
explored for use as the first stage in a two-stage (2-stage) mixing process. While it makes
sense that emulsifying the powdered material by fine wet-grinding in a colloidal state will
produce a much-improved hydration, there is also the need to grind coarser cements and
to increase the surface area of the available cement. Together with the emulsification of the
chemical admixture in a given mix design, strengths and resiliency are greatly enhanced.
The emulsification of the powdered constituents not only changes the performance of a
given mix design, it can affect the actual development of a particular mix design.
A study by the National Concrete Pavement Technology Center found “that concretes
produced by two-stage mixing with high shear slurry mixing will provide more workable
concrete with a greater degree of hydration and improved compressive strengths, espe-
cially at early ages.” [RUPNOW, SCHAEFER, WANG, AND HERMANSON, 2007, P. 30]
In Europe and North America, ongoing practice with first-stage mixing/emulsifica-
tion of powders is gaining momentum. Combined with advanced water treatments such
as the patented electrically charging of the water inline, results are nothing short of phe-
nomenal. For concrete pavements the reduction in cement, water, and admixtures may
result in a smaller carbon footprint for concrete pavements and increased durability of the
medium.

Endnotes
Ober, Robert. Batch Plant Basics 101, Chapter 6, “A Review of Current Mixing Technology,”
San Antonio, Texas: Plant Architects, 2003.

Ober, Robert, “Double-Duty.” Concrete Products, September 1999 <http://concreteproducts.com/


mag/concrete_double_duty/index.html>

Rupnow, Tyson D., Vernon R. Schaefer, Kejin Wang, and Benjamin L. Hermanson. Improving
Portland Cement Concrete Mix Consistency and Production Rate through Two-Stage Mixing,
Final Report. Ames, Iowa: Center for Transportation Research and Education, Iowa State
University, 2007.

Test Method for Within-Batch Uniformity of Freshly Mixed Concrete. U.S. Army Corps of Engi-
neers, Washington, D.C., 1992.
CONCRETE
PLACEMENT

A WELL-FINISHED, LONG-LASTING CONCRETE PAVEMENT requires a well-com-


pacted, well-placed concrete mixture. Placement is not the only factor—a long-lasting
pavement also requires a good mix design, as well as proper production and curing—but
placement is of paramount importance.
Concrete pavement placement can be divided into four (4) sub-categories:

N Concrete delivery/hauling
N Concrete lay-down
N Concrete consolidation/vibration
N Concrete floating/finishing/slip-forming

Delivery Options
In its early years, all concrete pavement was placed in form-
work. The concrete for major roadway and interstate (motor-
way) construction, by its very nature, required concrete
production near the pour sites. Often concrete was delivered
in dump trucks, also called tipper trucks (see Figure 11-1).
These were non-agitated, truck-mounted boxes. They were
adequate for delivering stiff concrete with low water/cement
ratios, so long as the delivery trip was short enough to prevent
initial hydration and hardening and loss of workability. When Fig. 11-1. Dump trucks can be used for delivering
concrete over short distances, but they don’t provide
hot or cold weather extremes were encountered, however, the agitation needed for longer delivery trips.
delivery by non-agitated trucks could prove problematic.
Today, non-agitated dump trucks are used where specifications allow, usually where
the delivery trip is short—under a few kilometers. In other cases, however, dump trucks
are not allowed, as the non-agitated concrete can hydrate too soon, causing the exposed
surface area to dry out, and allowing premature heat buildup within the truckload of
concrete.

102
CONC RE T E P L ACE M ENT 103

A crossover between the non-agitated dump truck and a


transit-mixer (drum) truck is the dumping or tipping agitator
box, made by Maxon Industries (www.maxon.com) and known
by the trade name AGITOR. (See Figure 11-2.) This hybrid
machine utilizes hydraulics to lift a tapered box that contains a
single-shaft agitation device. While not a mixer, it agitates the
material, keeping the concrete in suspension. A tipping agitator
can be effective for the same amount of time as a transit mixer—
60 to 90 minutes, depending on the local specification. Fig. 11-2. The AGITOR tilts to empty and has
In most cases, however, transit mixers (otherwise known as a built-in single-shaft mixer, which keeps the
concrete in suspension. It keeps the concrete
drum or truck mixers) are used. (See Figure 11-3.) Certain varia- workable during 60- to 90-minute deliveries.
tions on the truck mixer, often referred to as paving mixers, uti-
lize either enlarged feed chutes or flip-up (moveable) feed chutes. The enlarged feed chutes
allow for easier filling of the transit mixer with the stiff consistency of most lower water/
cement ratio mixes. Flip-up feed chutes ease the discharge of these lower-viscosity mixes.
In most jurisdictions the total number of drum revolutions is limited to a maximum
of 300 revolutions at no more than 8 rpm (revolutions per minute). European transit mix-
ers are normally equipped with more fins at a lower profile than North American trucks,
resulting in a screw-type action that mixes more gently. In a typical North American tran-
sit mixer, the fins are much deeper and only about one-third (1/3) as many are used,
thereby creating a lofting action. This action is desirable in transit-mixed (as opposed to
plant-mixed) concrete. If the drum is rotated too quickly or for too long, however, the
North American drum design can segregate and destroy a properly homogenized paving
mix by pulling it apart. These segregated mixtures have rock tailings at the end of the load
and extremely stiff mixture at the midpoint of discharge.
In rare cases, conveyor belts are utilized for delivering pavement mixtures from the
concrete batching/mixing plant to the job site. This methodology is sometimes employed
where runways and aircraft tarmacs
are built at extra-depth, and on dam
pavement and embankments. Conveyor
delivery is rare for roadways, but it can
be effective if the conveyor is covered
for weatherproofing.

Fig. 11-3. Drum mixers turn to keep the concrete in suspension during 60- to
90-minute deliveries. Because American mixers have fewer, but larger, internal
fins, the mix is more likely to segregate in an American-style mixer.
CONC RE T E P L ACE M ENT 104

Concr ete Lay-Down


Concrete lay-down refers to the initial placement process, which can vary from dumped
by truck and moved by front-end loader to fully automated auger placement of slip-form
paving machines.
Non-automated feed-type paving
machines take windrowed fresh con-
crete deposited by dump truck, agitat-
ing body, or conveyor in front of the
machine (usually from an offset posi-
tion). This practice is discouraged
because it requires reinforcement steel
and sub-grade supports. Automated
“slip-form” paving machines, on the
other hand, generally utilize a short,
offset conveyor feed (not to be con-
fused with the conveyor belts that run Fig. 11-4. Slip-form pavers pour pavement, form the edges of the pavement,
from a batch plant). (See Figure 11-4.) and strike it level as they work. There’s no need to set forms—the mix used
is stiff enough that rails built into the machine shape the edges as the paver
The mix is stiff enough that the con- moves forward.
crete can be laid within guides on the
machine without the need for any other forms. Controlling the amount of concrete in
front of the paving machine is crucial in providing for advancement of the slip-forming
according to specification. The offset-feed conveyors allow for an exact deposit depth of
concrete mixture in front of the machine’s spreading mechanism.

Concr ete Consolidation/V ibration


Most concrete mixes are stiff in order to control the finish and so that excess moisture is
not present. Excess moisture can weaken the very topside of the finish and can lead to pre-
mature wear, shrinkage cracking, and spalling in freeze-thaw conditions. Along with these
considerations, pavements also are often sloped or curved, and as such the stiffer consis-
tency allows for this deformation without loss of dimensional control. Because the slip-
form process requires no forms, success is dependent on the stiffer consistency of the
paving mixtures for advancement and for achieving the proper tolerances on flat and
straightforward runs.
With the desired stiff consistency of paving mixtures comes the need for vibration and
consolidation, so that the mix will flow and encompass the supporting reinforcing steel.
These stiff mixes can be understood as being “Bingham Plastic” or in a “thyrotrophic”
state, with the less viscous material becoming more flowable under agitation, or as energy
is exerted by vibration. Vibration on modern European and North American concrete
paving machinery utilizes hydraulic internal vibration devices. The vibrator devices are
CONC RE T E P L ACE M ENT 105

internal to the concrete mixture and are pushed and/ Fig. 11-5. While a stiff mixture allows
a newly placed pavement to be used
or pulled through the windrowed or conveyor/chute- more quickly, the fresh concrete
placed fresh mixture. (See Figure 11-5.) The internal may be too stiff to flow around the
steel reinforcing bars. Running an
vibrators allow placement in the material at a struck internal vibration device, like this one,
level by exerting energy and creating a flow condition through a stiff mixture makes it flow
better and fill in the voids around the
of the mix. reinforcement.

Concr ete F loating and F inishing


When internal vibration is exerted on a fresh mixture of concrete, the finest fines are
“floated” to the surface, carrying with them the surrounding matrix. The float is a gelatin-
effect that allows the transient fines to be carried along with sufficient supporting fine
aggregate and some of the smaller intermediate coarser aggregate. It is this transient
floatation that provides the desired hard surface finish on stiff concrete pavement when
freshly slip-formed or finished. As previously mentioned, the overworking or creation of
too many fines on the topsides finish will result in a weaker finish. The pavement will
also expand and contract at varying rates, which can lead to problems such as cracking
and spalling.
In the late 1970s, GOMACO Corporation (www.gomaco.com) developed a finishing
drum machine that was able to eccentrically compact, float, and finish the paving mix. The
technology was transferred from a machine built in the late ’60s for finishing hard-to-
reach bridge decks. Concrete pumping was becoming more widely used at the time, and
the GOMACO technology allowed strike-off, floating, and finishing to keep pace with the
faster placement by pump. The adaptation of the roller-drum floating/finishing apparatus
was commercially successful and led to significant growth of the company and of concrete
paving contractors, and also led to acceptance of concrete pavement.
As the placement of fresh concrete pavement became more automated and as it became
possible to pave several kilometers per day, machine-placed and -finished concrete pave-
ment made important inroads on asphalt pavement, and allowed for far more economical
models of formless and form-edged non-slip-form concrete pavement utilizing drum fin-
ishing machines. In recent years, as petroleum pricing has fluctuated greatly and the sup-
ply was uncertain from time to time, slip-formed concrete pavement has become more
commonplace in markets where asphalt was previously the specified choice. Ed Sullivan,
the chief economist for the Portland Cement Association (PCA), says that “in the past,
paving cost estimates favored asphalt—resulting in 94% of all paved highways being made
of asphalt. New paving realities are now in place—reflecting changes in global oil demand
and new refining practices which reduce liquid asphalt production. The new paving reali-
ties now show that comparative life cycle and initial bid cost assessments will increasingly
favor concrete over asphalt in the foreseeable future.” [SULLIVAN, CEMENT AMERICAS,
2009] In fact, PCA estimates say that a two-lane (2-lane) concrete highway costs 25% less
than a comparable two-lane asphalt highway [SULLIVAN, THE MONITOR, 2008].
CONC RE T E P L ACE M ENT 106

G rade Control and A dvanc ements in P roduc tivity


In 1956 the construction firm of Guntert & Zimmerman of California (www.guntert.com)
developed the first track crawler–mounted slip-form paving machine. The machine also
utilized an automatic line and grade control, which revolutionized the advancement of
concrete pavement in airport, highway, and canal lining construction.
Many variations on this apparatus followed from German and American manufactur-
ers in the decades that followed, with the inclusion of advanced hydraulics for better
response and drive capabilities, as well as for advancements in grade and line control. The
modern slip-form paver, as of the time of this publication, has evolved to include onboard
laser-line and gradient computers, which download topographic data via a global posi-
tioning system (GPS). Other interesting technology includes 4-D Building Information
Management (BIM), which incorporates the profile for the concrete pavement design.
This technology combines 3-D modeling with time as a fourth dimension to better coor-
dinate highway paving and planning.

Endnotes
Sullivan, Edward. “Paving: The New Realities,” The Monitor, October 14, 2008. <www.cement.org/
Flash-Paving Report.pdf>

Sullivan, Edward. “Planning for the Big Boom.” Cement A mericas, November 1, 2009. <http://
cementamericas.com/mag/us-cement-industry-20091101/>
CURING AND
SEALING CONCRETE
IN PAVEMENTS

CURING IS THE PROCESS OF MAINTAINING the moisture and temperature condi-


tions necessary for the development of durable concrete pavements. The process begins
when water reacts with the Portland cement in the concrete pavement mix. The mixture
continues as the mix goes from a fluid to a hardened state, through the chemical process
of hydration. Under favorable moisture content and temperature conditions, the curing

Ty p ic a l C u r in g De fe c t s

Improper curing can lead to cracking, spalling, and delamination. Cracking takes three (3) forms. The most
common form, plastic shrinkage cracking, occurs before the initial setting, when concrete is still in its plastic
(pliable) form and water loss leads to shrinkage. Large, parallel cracks may appear, shallow at first and then
deepening. Alternatively, the cracks may be very fine in appearance, only 30 mm or less in depth, and while
the effects may only be cosmetic, the cracks may make concrete vulnerable to freeze-and-thawing cycles.
Although some cracking in concrete pavement is acceptable, and can be the result of improper sub-grade
preparation and settlement, cracking caused by typical curing defects will appear much sooner. Plastic
shrinkage cracking not only can lead to the aforementioned longer-term defects, but depending on climatic
conditions, can also lead to almost immediate problems with driveability and normally will cause rejection by
transportation department or third-party inspectors.
Cracking can also be caused by thermal stress. While this type of cracking is often considered a problem
with large masses of concrete, thermal gradients within pavement can also cause trouble.
A third type of cracking is caused by drying shrinkage, occurring when hardened concrete dries more than
expected. Cracks appear at close intervals rather than at the contraction joints. This cracking in turn may
lead to faster deterioration, but it does not necessarily affect the load-bearing capacity of pavement because
typically only the top 50 mm of the concrete is involved.
Another category of deterioration is erosion of the paste and spalling of the surface. Spalling is apt to take
place when the choice of materials, the weather, and construction procedures interfere with the pavement’s
strength development. If stresses exceed the strength of the roadway, delamination may occur adjacent to
cracks and joints. Traffic and climate can then aggravate these damaged sections, leading to spalling. Spall-
ing can also result from certain chemical and mineral reactions, including alkaline and carbon interactions
that can create rapid expansions in the setting finish [POOLE, 2005, PP. 1–4].

107
CURI NG AND SE AL I NG CONC RE T E I N PAV E M E N T S 108

process continues over time to increase durability, strength, resistance to abrasion, and
volume stability. Satisfactory curing also makes concrete less vulnerable to freezing and
thawing cycles, while minimizing the corrosive effect of de-icers. This chapter provides an
overview of the common curing processes employed in pavements in North America and
Europe, with background and commentary provided on the subject in general.
Curing occurs in two (2) stages. During the initial stage, which occurs before the con-
crete begins to set, chemical changes are especially pronounced. Curing is a gradual pro-
cess, however, and isn’t considered complete until after the second, final stage, when the
concrete approaches the intended design strength. Thereafter it may continue to cure for
several years, decades, or even centuries, depending on the actual type of mix and ensuing
hydration. Throughout both stages of the curing process, evaporation and thermal stresses
are primary concerns because of pavement’s exposure to the elements and its large surface
area relative to volume. A less-than-optimal curing process can lead to reduced durability,
curling pavement, and cracking, conditions that cannot be remedied. At any point in
either stage, freezing, sudden changes in temperature, or overheating can seriously weaken
the pavement. During the initial stage, the primary evaporation concern is bleed water,
which leads to cracking if it evaporates too quickly. During the final stage, the primary
concern is keeping the pavement wet so that hydration can continue at the proper rate and
the concrete can reach maximum strength [poole, 2005, PP. 1–4].

Planning
Proper curing depends on the makeup of the mix, temperature, and ambient evaporation
rates, all of which must be taken into account before the paving job begins. The goal is to
“control” the cure in all cases—whether the prevailing conditions are hot and humid, hot
and dry, cold and damp, or cold and dry.

MIX DESIGN. The most obvious consideration is the type of Portland cement used. Port-
land types I and II cure at specified rates, comparable to each other. Type III cures more
quickly; Type IV cures more slowly. Water reducers, used to good effect in mixing and
placing, can also adversely affect curing. Admixtures can be used to retard setting time;
alternatively, in cold conditions, an accelerating additive may be included in the mix. This
can cause a reaction that greatly accelerates early hydration. Free mixing water is taken up,
which then limits bleeding or causes it to stop altogether, and there is the possibility of
internal desiccation of the concrete [FHWA, 1999]. In drier climates, fine-grained cements
and certain pozzolans, both of which limit water bleed, can also be a problem. If concrete
becomes too dry while curing is underway, there is an increased chance of plastic shrink-
age cracking [ARCHITECTURAL COMPUTER SERVICES, 2005]. In the case of advanced water
treatment by means of polarization, the curing effect of concrete pavement mixtures can
be slowed by using a small amount of water-reducing agent, and thereby controlling the
initial cure over several hours (see the chapter Water for Concrete Production).
CURI NG AND SE AL I NG CONC RE T E I N PAV E M E N T S 109

TEMPERATURE. Hot weather can increase the rate of hydration and evaporation, weaken-
ing the pavement. In cold weather, the process of hydration slows as the temperature of the
concrete drops. Strength develops more gradually, with little or no increase if the ther-
mometer drops below the freezing point. A complete freeze of the pavement will result in
a greatly weakened surface. The crystallization of the matrix (glue) weakens the bond and
almost always will result in spalling of the finish. The concrete is likely to suffer from
delayed setting, rapid slump loss, plastic shrinkage cracking, and premature stiffening.
Initial precautions against hot and cold weather include cooling or heating the mix
ingredients, and are covered in the chapter Concrete Placement. Once the mix has been
determined, the effects of hot weather can be mitigated to a certain extent by placing the
mix at night, when temperatures are cooler. The addition of chilled water in place of ambi-
ent stored water, or the addition of ice in lieu of chilled water, can greatly lower the con-
crete mix temperature. Flake ice is preferred to chip or nugget/cubed ice for its quick
thermal exchange/melting characteristics.
In cases where the water/cement ratio is too low to allow the effective offset in con-
crete temperatures by water substitution with chilled water or even with 100% flaked ice,
it is often necessary to cool the coarse aggregate by submersion in chilled water or running
the aggregate on a belt conveyor with a chilled water spray (wet belt). This aggregate cool-
ing combined with flaked ice usually is sufficient for controlling concrete temperatures. In
the most extreme cases, chilled air can also be circulated through the concrete plant’s
aggregate-storage bins.
Generally, concrete cooling is calculated using a temperature-balance worksheet,
whereby the mix design information, the aggregate and water temperatures, and the ambi-
ent conditions (humidity and air temperatures) are input and the desired concrete tem-
perature is calculated, with some or all of the cooling methods discussed.
Liquefied gases such as nitrogen or carbon dioxide are also used, where supply is avail-
able and prices low enough to utilize the gas in bulk. In the case of liquefied gases, the use
of a high-intensity twin-shaft mixer can allow for a much better result than is experienced
with the tilting-drum-type mixer. In the high-intensity mixer, the gases are entrained
much faster than in the lofting action of the mixer drum—where the liquefied gases
are released to the atmosphere too quickly. The mix becoming too cold will affect
curing times.
Cold weather can be mitigated by covering the pavement with insulating blankets, or
even with electrically heated insulating blankets. The heating of mixes is done with hot
mixing water, hot air circulation through the plant’s aggregate bins, or live steam release in
the aggregate stockpiles. The mix becoming too hot will affect curing times.

EVAPORATION. Hydration is vulnerable to a loss of water through evaporation, especially


at the surface. When ambient conditions and concrete temperatures combine to produce
conditions that create a high evaporation rate, there is an increased chance of the surface
CURI NG AND SE AL I NG CONC RE T E I N PAV E M E N T S 110

drying prematurely and the formation of shrinkage cracks. So it is important to determine


the drying conditions of the site before the scheduled placement; probable evaporation
rates are calculated by considering both concrete and air temperatures, as well as wind
velocity and relative humidity. Evaporation problems, like temperature problems, can
often be mitigated by beginning work after sundown, when temperatures moderate and
the relative humidity tends to rise [poole, 2005, p. 14].

Initial Curing
As noted earlier, curing occurs in two (2) stages. The initial curing stage covers the period
from the placing of concrete until it begins to set, or solidify. At this point, the primary
concern is maintaining the proper rate at which water bleeds to the surface of the concrete.
If bleed water dries from the surface before the concrete begins to set, plastic shrinkage
cracking occurs. (Too much bleed water, on the other hand, results in a low water/cement
ratio at the surface of the pavement, and a correspondingly weak surface layer with low
abrasion resistance. Since excessive bleed is primarily controlled in the mix design stage,
however, it is not generally considered a curing issue.) Two (2) strategies can be used to
compensate for the rapid drying of bleed water. The first is to reduce the temperature
of the concrete when it is placed. The second is to reduce evaporation. It is difficult to
moderate concrete temperature over a large surface area, but one potential measure is
to cool stockpiles of aggregates that will be used in the mix by sprinkling them with water
for a modest level of evaporative cool-
Figure 12-1
ing [TIMMS, ROBINSON, ET AL., 1952].
Recently, spray-on evaporation reduc- Effe c t o f R e d u c ni g Ev a p o r a t io n b y 5 0 %
b y Usin g a n Ev a p o r a t oi n Re d u c e r
ers have come into widespread use as
5.0
an efficient means of creating a mem-
Cumulative bleeding or evaporation, kg/m2

brane that will ensure proper curing by


4.0
slowing evaporation rates. cumulative evaporation with
no evaporation retarder
Evaporation reducers are applied
soon after the initial setting, and may 3.0

need to be reapplied several times dur- cumulative


bleeding
ing the course of the initial cure. Reduc- 2.0
time of setting

ing the evaporation rate also can serve


to lower the concrete placing tempera- 1.0
cumulative evaporation
ture. These reducers are particularly with 50% reduction
in evaporation rate
efficient when treating large-scale proj-
0
ects [poole, 2005, p. 14]. Figure 12-1 0 1 2 3 4 5

shows the application of a reducer as Time, hours


cutting evaporation by 50%.
Source: Poole, 2005. p.15.
CURI NG AND SE AL I NG CONC RE T E I N PAV E M E N T S 111

Final Curing
After concrete undergoes the change to a solid state, final curing procedures come into
play. At this point, the strength of the concrete depends on keeping it moist, as shown in
Figure 12-2. Note that concrete that is moist-cured for an entire year will show a rapid gain
in compressive strength over the first 90 days, and continue to slowly gain strength. In
contrast, concrete that is kept moist for only a limited time will show no improvement in
strength after the initial 90 days [KOSTMATKA, KERKHOFF, & PANARESE, 2008, P. 219].
There are three (3) common methods of maintaining moisture during final curing:
adding water, covering with sheet materials, Figure 12-2
and applying curing compounds. The latter is
Effe c t o f Mo ist Cu r in g Time
the most practical in large projects. o n St r e n g t h G a in o f C o n c r e t e
(originally from Gonnerman and Schuman, 1928)
60
ADDED WATER. In hot weather, adding water
during the early hardening period can keep Moist-cured entire time
50
moisture at the desired level as well as con-
Compressive strength, MPa

In air after 28 days moist curing


tributing a cooling effect through evapora-
40
In air after 7 days moist curing
tion. For smaller areas, a low dike of earth or
sand can be used for ponding water over the 30 In laboratory air entire time

pavement. On a larger scale, spraying or fog-


ging the concrete with a fine mist will increase 20

the humidity above the surface. Alternatively,


saturated wet coverings can be applied to 10

large areas. A standard material is burlap,


0
including special-purpose coverings made 0 7 28 90 365
for concrete projects. Layers of wet straw or Age at test, days
hay spread over the surface are another
option. Source: Kostmatka, Kerkhoff, and Panarese, 2008. p. 219.

SHEET MATERIALS. After the initial setting, plastic sheets can be placed over the surface
to help retain water. For good coverage, sheets should extend to cover exposed edges, and
individual sheets should be overlapped. Contact with plastic may create a mottled pattern
on the surface; this can be avoided by using sheets of polyethylene laminated with burlap.
White-pigmented sheets will reflect solar energy and keep pavements cooler.

CURING COMPOUNDS. Curing compounds have become the most popular and economi-
cal means of controlling moisture loss during the final curing of pavements. A long, slow,
moist cure is virtually assured because the compound remains on the surface. On larger
areas, powered spray equipment is used to apply curing compounds over a dampened
surface. It may be necessary to make a second application in certain situations; if so,
the additional compound should be applied when the initial coating becomes tacky
[taylor, 2005, p. 4].
CURI NG AND SE AL I NG CONC RE T E I N PAV E M E N T S 112

A number of considerations come into play when choosing a curing compound.

N White-pigmented compounds have two (2) advantages over clear products: They
keep concrete cooler by absorbing less solar radiation, and the pigments are a
visual indicator of whether or not the compound has been applied evenly.
N New, water-based curing products can be used to create less environmental impact.
N Drying time of a product needs to be considered in order to avoid having precipi-
tation wash off the coating.
N The rate of water loss varies
En v ir o n me n t a ly Fr ie n d ly Cu r in g Co mp o u n d s
from compound to compound,
and should be considered.
Most concrete curing compounds used in highway construc-
N Viscosity, the measure of a
tion have been hydrocarbon-based, containing resins and
compound’s tendency to run, mineral spirits as a vehicle. The Texas Department of Trans-
should also be considered, portation has noted that a paving project might use anywhere
especially with regard to from 500 to 30,000 gallons (1,900 to 114,000 liters) of the
product. The compounds potentially are spread from the site
surface texture. A low- by spills or by unanticipated heavy rains. Also, there is con-
viscosity compound will cern about the health impact on workers applying the
tend to disperse into the relief compounds. With these factors in mind, there has been a
move to environment-friendly alternatives. One such product
of textured or grooved pave-
combines a petroleum wax and white pigment in a water
ments, making it less effective. base, a formulation that is VOC-compliant (minimizing
One remedy is to use two (2) release of volatile organic compounds). As new environmen-
tal legislation comes into effect, the shift to water-based
light coats rather than a single
curing compounds likely will continue.
application.

Sealing Compounds
Sealing compounds are applied to the surface of hardened concrete to limit the penetra-
tion of gases and liquids. While applying them is optional, sealants help protect roadways
from freeze-thaw cycles, which can open cracks that allow dissolved de-icing salt to cor-
rode reinforcing steel and deteriorate the concrete itself [WISCONSIN D.O.T., 2006].
These products typically are applied after roughly 28 days, and function best under
certain conditions. They cannot penetrate the concrete surface if surface holes are filled
with water. And the sealing of concrete placed in late fall should be delayed until spring to
prevent causing freeze-thaw problems.
Sealing compounds can be classified as film-forming, penetrating, or silane-siloxane
products.

FILM-FORMING (or membrane-forming) sealing compounds remain mostly on the sur-


face with little penetration, and they may have to be reapplied as traffic wears through
them. The longstanding formulation has been a 50:50 mix of boiled linseed oil and min-
eral spirits, but newer compounds are now available.
CURI NG AND SE AL I NG CONC RE T E I N PAV E M E N T S 113

PENETRATING compounds are chemically active, bonding with minerals in the cement
and blocking both surface moisture and moisture from below. These sealants can increase
strength and density, and they are longer-lasting than film-forming sealants. Penetrating
compounds can discolor the concrete, however, and they require periodic reapplication.

SILANE-SILOXANE products react with the chemicals released by the cement hydration
process. They work to substantially reduce the size of the pores in the concrete, and thereby
prevent dirt, oil, and salts from penetrating while allowing the concrete to breathe [KOST-
MATKA, KERKHOFF, AND PANARESE, 2008, PP. 226–227].

Endnotes
A. G. Timms, D. L. Robinson, H. J. Gilkey, Roy W. Carlson, W. R. Johnson, G. E. Burnett, and
H. C. Vollmer, “Curing of Concrete,” Journal Proceedings, American Concrete Institute, 1952.
<www.concrete.org/PUBS/JOURNALS/AbstractDetails.asp?ID=11949>

Kostmatka, Steven H., Beatrix Kerkhoff, and William C. Panarese. Design and Control of Concrete
Mixtures, 14th Edition, Portland Cement Assn., 2008.

Lab Tests Rate Sealants under Freeze-Thaw Conditions. Wisconsin Department of Transportation,
2006. <http://on.dot.wi.gov/wisdotresearch/database/briefs/03-09concretesealers-b.pdf>

Miscellaneous Cast-in-Place Concrete. Architectural Computer Services, Riverside, California,


2005. <www.arcomnet.com/masterspec/supdoc04/frame/033053f00.html>

Poole, Toy S., Guide for Curing of Portland Cement Concrete Pavements, Vol. 1. U.S. Dept. of
Transportation – Federal Highway Administration, 2005. <www.fhwa.dot.gov/pavement/
pccp/pubs/02099/02099.pdf>

Taylor, Peter T., and Celik Ozyildirim. Curing Practice for Slipformed Concrete Pavements.
Concrete Pavement Technology Program, U.S. Dept. of Transportation – Federal Highway
Administration, 2005. <www.fhwa.dot.gov/pavement/pccp/pubs/06003/06003.pdf>

Water-Reducing. Federal Highway Administration, U.S. Dept. of Transportation – Federal


Highway Administration, 1999. <www.fhwa.dot.gov/infrastructure/materialsgrp/water.htm>
QUALITY
ASSURANCE

QUALITY ASSURANCE IS AN ONGOING PROCESS that, done properly, assures a


well-done, long-lasting concrete pavement or highway. There are three (3) basic parts of
any quality assurance program: specification design, q uality control, and acceptance. Speci-
fication design is the responsibility of the funding agency, and sets standards for acceptable
materials, equipment, and construction processes. Q uality control plans are usually drafted
by the contractor and use the specification design to plot out the materials, methods,
equipment, and staff that will be used to complete the roadway project to the funding
agency’s criteria. The acceptance program, which may run concurrently with the quality
control process, defines the testing that will be performed to determine how well the con-
tractor has fulfilled the design specifications. It is the responsibility of the funding agency,
and can include pay structures based on performance [AASHTO, 1996, P. 9].

Spec ific ation D esign in th e U .S .A


Each specification usually outlines and tracks a range of quality; statistical analysis is used
to determine the acceptable limits of each specification. Due to the variability of the con-
struction process and materials, the method of statistical analysis generally accepted by
transportation engineers describes the “percent defective,” or its counterpart “percent
within limits.” This method takes into account both the desired standard and the variability
that arises in manufacturing. It is relatively simple to understand, it can be readily applied
to most construction quality characteristics, and it promotes uniform quality within speci-
fied limits that can be attributed to ultimate performance [AASHTO, 1996, P. A-2].
Because the specifications developed in this part of the quality assurance program are
crucial to the rest of the program, they must be developed carefully. Funding agencies do
not need to determine quality targets entirely on their own, however. Much information
about pavement performance can be gleaned from historical data. Conducting prelimi-
nary literature reviews, such as a review of the various states’ standards, is a practical place
to begin. Other sources include literature from:

114
QUAL I T Y AS S URANC E 115

N Transportation Research Board (TRB)


N American Association of State Highway and Transportation Officials (AASHTO)
N American Society for Testing and Materials (ASTM)
N U.S. Department of Transportation’s Federal Highway Administration (FHWA)
N American Concrete Paving Association (ACPA)
N National Stone Association (NSA)

The funding agency should also seek input from the contractor as well as from experts in
materials testing, statistics, pavement engineering, contract administration, and others.
Contacting other agencies to obtain their input can prove very valuable, as well [BURATI,
WEED, HUGHES & HILL, 1996, P. 2].

COMPUTER SOFTWARE. Two (2) software tools can be helpful in drawing up the specifi-
cation design: HIPERPAV (HIgh PERformance Concrete PAVing) and PaveSpec. HIPER-
PAV analyzes early age performance in jointed concrete pavements, continuously
reinforced concrete pavements, and bonded concrete overlays. Based on the vast array of
data available within this program, users input the factors involved in the concrete project,
including such things as mix formulation, joint design, type of subbase, and curing
method. HIPERPAV then can accurately predict the stress and strength performance dur-
ing the initial 72-hour period—which, in turn, can accurately predict the pavement’s long-
term performance [HIPERPAV, 2010].
Another software program, PaveSpec, can also be employed to simplify the specifica-
tion process. Based on stored data, this program can be used to develop performance-
related specifications and predict future performance of a pavement. Additionally, this
software allows both agencies and contractors to gather information on best practices and
materials needed to create long-lasting, high-performance pavements. Some of PaveSpec’s
specific capabilities include:

N Simulation of pavement behavior in terms of transverse cracking, transverse joint


faulting, transverse joint spalling, and pavement smoothness over time
N Application of a user-defined maintenance and rehabilitation plan to compute
life-cycle costs
N Development of pay structures for meeting, exceeding, or failing to meet the
following acceptance quality characteristics: strength, thickness, air content,
smoothness, and consolidation around dowels
N Computation of contractor pay based on test results for the five (5) acceptance
quality characteristics listed immediately above

With PaveSpec, a user can develop a specification (using the program’s Specification
Wizard), analyze an already-developed specification, and/or use a specification for a spe-
cific project [FHWA, PAVESPEC USER’S GUIDE, 2007].
QUAL I T Y AS S URANC E 116

D ev eloping a Q uality C ontrol Plan


The quality control plan should be developed by the contractor rather than by the funding
agency, in order to emphasize that quality is the contractor’s responsibility. Based on the
specification design, the contractor specifies what materials and activities will be tested in
order to produce pavement within the acceptable quality limits. In order to maintain the
quality criteria, the plan must include:

N The contractor’s policy for obtaining quality and the assignment of quality control
accountability and responsibility
N A statistically valid sampling, testing, and analysis plan with frequencies, loca-
tions, and methods
N Provisions for disposal or rework on non-specification materials or work

The plan should provide for and maintain a system that reasonably ensures that all materi-
als and work submitted for acceptance conform to the contract. The plan should address
each design specification, including the actions and considerations needed to assess pro-
duction and construction processes in order to control the quality level of the finished
pavement. The inspections and tests required to support conformance with the contract
are made available to the agency for the period required by the contract [BURATI, WEED,
HUGHES & HILL, 1996, P. 9].
The quality control plan is submitted by the contractor to the agency at an initial pre-
construction conference [AASHTO, P. 9]. Once the quality control plan is approved, the
contractor is responsible for training staff on proper construction processes, developing
appropriate manufacturing standards, as well as establishing testing procedures and
record-keeping to ensure that the construction processes and materials adhere to the qual-
ity requirements. Testing during the construction phases ensures that the agency’s quality
standards are maintained and, if not acceptable, modified in a timely manner to ensure
quality in the finished pavement.

T esting
Testing should be performed by the contractor as part of the quality control plan, as well
as by the agency as part of its acceptance plan. There are generally two (2) different
approaches that have been used by agencies to specify the quality required. In one approach,
the agency stipulates the minimum quality control requirements and properties that the
quality control plan must contain. In the alternative approach, the agency specifies all the
requirements and properties that must be tested. There are advantages and disadvantages
to both approaches. In the first approach, by stating the minimum requirements and prop-
erties, the agency conveys to the contractor the lowest acceptable quality. The disadvan-
tage is that the contractor may view this quality level as all that is needed for adequate
quality control, rather than as the minimum level of quality. On the other hand, if the plan
QUAL I T Y AS S URANC E 117

Co n st r u c t io n Ma n a g e me n t Pr a c t i c e s i n Eu r o p e a n d Ca n a d a

U.S. quality control programs may be successful at identify the ideal contractors for a specific project.
ensuring quality construction, but the system is
Current U.S. practices use unit-price structures
somewhat adversarial, with contractors on one side
that allow for different rates of pay based on quan-
and agencies on the other. In Europe and Canada,
tity variations. However, there are rarely pay
agencies are attempting to improve quality by
incentives for completing project milestones ahead
improving relationships between public agencies,
of schedule. In Scotland, the funding agency cre-
contractors, and private engineering firms.
ates a series of completion milestones for which
Approaches include risk allocation, design-build
lump-sum payments are made to the contractor.
contracts, public/private arrangements, mainte-
Ideally, this minimizes the agency’s administrative
nance and warranty requirements, and use of
duties while offering incentives to complete project
third-party consultants to carry out contract
milestones in a timely way.
management.
Another shift in traditional management methods is
Britain and the Netherlands have both established
in delegating some of the agency’s responsibilities
systems that assist project teams in identifying
to the contractor. For instance, implementing
and quantifying project risk, using probability
milestones and lump-sum payments allows the
techniques. Risk analysis and risk allocation tech-
contractor to invoice for completed work with less
niques are then used to help agencies choose
required verification than in traditional U.S. con-
delivery and contracting strategies that can man-
tracts. Also, because contractors can perform
age and alleviate these risks. In England, the
quality control more efficiently, the agencies can
Highways Agency’s Early Contractor Involvement
rely on contractor-designed quality management
(ECI) delivery method uses a qualifications-based
systems that meet the procedures submitted in the
design and contractor selection process (instead of
project proposal. Utilizing best-value procurement
a traditional bid process). The agency and contrac-
and past performance can enable the agency to
tor then work together to develop a project budget
assume a quality audit role in construction. All of
through an “open book” target pricing system. This
the countries participating in the research viewed
method distributes the risk of missing the project
the International Standards Organization (ISO) cer-
budget between both entities—instead of just the
tification as a critical factor in shifting quality
contractor.
management to the contractor.
Many European agencies use a best-value contrac-
Because of agency staff reductions, many Euro-
tor procurement system; others hire contractors
pean agencies have turned to the private sector
solely on qualifications (without a bid price). Best-
for delivery functions, including planning, design,
value practices weigh various factors in addition to
construction management, operations, and main-
price to award a contract. These factors include
tenance. As a result, the agency focus has shifted
team qualifications, past performance, and design
from contract compliance for individual services to
alternatives, as well as other items based on the
network management by integrated service con-
specific project’s needs. A new tool created by the
tracts. These changes have led to a more symbiotic,
Ministry of Transportation in Ontario, Canada, rates
long-lasting partnership and collaboration between
contractors based on past performance. The Reg-
agencies and contractors. As agencies have become
istry Appraisal and Quality System (RAQS) is used
more cognizant of how procurement and construc-
to adjust prequalification ratings for bidding pur-
tion methods affect their design and construction
poses. The British Highways Agency’s Capability
professionals and their supply chains, contractors,
Assessment Toolkit (CAT) is used to rate company
in turn, have worked to better understand the agen-
management practices and combine them with
cies’ needs.
past performance to select contractors based
purely on qualifications. This practice has the Derived from Executive Summary: Construction Management
Practices for Effective Project Delivery, Contract Compliance, and
potential of establishing long-term partnerships Quality Assurance, International Technology Exchange Program,
between the agency and contractors, plus it can May 2004
QUAL I T Y AS S URANC E 118

is all-inclusive, the contractor may view the quality control plan as the agency’s plan rather
than the contractor’s plan [BURATI, WEED, HUGHES & HILL, 1996, P. 31].
For Portland concrete pavements, typical quality characteristics that are tested for
quality control include aggregate quality, gradation of critical sieve sizes, air content, water/
cement ratio, mix temperature, and slump [BURATI, WEED, HUGHES & HILL, 1996, P. 33].
Maintaining acceptable quality levels of these factors can ensure a reliable pavement design.
If any of these factors are under acceptable limits, they must be tested at each step to ensure
that the quality level can be corrected before it is incorporated into the mixture.
Frequency of testing—which should be outlined in the contract—is a significant fac-
tor in the effectiveness of quality control. The American Concrete Institute recommends
the following:

…aggregate moisture tests are made once or twice a day. The first batch of fine
aggregate in the morning is often overly wet since moisture will migrate overnight
to the bottom of the storage bin. As fine aggregate is drawn from the bottom of the
bin and additional aggregate is added, the moisture content should stabilize at a
lower level and the first moisture test can be made.

Slump tests should be made for the first batch of concrete each day, whenever con-
sistency of concrete appears to vary, and whenever strength-test cylinders are
made at the job site.

Air-content tests should be made often enough at the point of delivery to ensure
proper air content, particularly if temperature and aggregate grading change. An
air-content test should be performed for each sample of concrete from which cyl-
inders are made; a record of the temperature of each sample of concrete should
also be kept.

In-place concrete strengths are typically documented by casting specimens that


are field-cured (as nearly as practical) in the same manner as concrete in the struc-
ture. Field-cured specimens are commonly used to decide…when traffic will be
allowed on new pavement. ASTM C31 (AASHTO T 23) contains additional
instructions regarding the handling of field-cured cylinders. Although field-cured
specimens may be tested at any age, 7-day tests are often made for comparison
with laboratory tests at the same age; these are useful to judge if curing and pro-
tection during cold weather concreting is adequate.

A c c eptanc e
The acceptance program “defines a set of rational procedures to be used by the agency to
determine the degree of compliance with contract requirements and value of product
delivered by the contractor.” The acceptance program is designed to determine how closely
the contractor complies with the contract requirements. If bonus or incentive programs
QUAL I T Y AS S URANC E 119

Pe r fo r ma n c e -R e la t e d Sp e c i fi c a t i o n s

Performance-related specifications for Portland means and standard deviations. These factors
cement concrete pavements are based on an inno- represent the quality level for which the agency
vative approach that uses measured acceptance will pay 100%. The performance-related specifi-
quality characteristics—including strength, slab cation method, based on the acceptance quality
thickness, initial smoothness, air content, and characteristics means and standard deviation
consolidation around dowels—to predict future values produced, translates into the agency
pavement performance, through mathematical awarding incentives or disincentives.
relationships.
N Measure in-situ performance of as-constructed
This evaluation method quantifies pavement per- quality. Performance-related specification is based
formance in terms of key distresses—transverse on life-cycle cost of the as-constructed pavement;
joint faulting and spalling, transverse fatigue crack- therefore direct, in-situ samples of the completed
ing, and smoothness over time (measured by the pavement are highly desired.
International Roughness Index). This method
N Provide mathematical performance prediction
enables agencies and contractors to directly link
models in terms of four (4) distress indicators:
measured acceptance quality characteristics to the
transverse joint faulting and spalling, transverse
future life-cycle costs, thereby enabling agencies
fatigue cracking, and International Roughness
to develop rational and fair contractor pay adjust-
Index.
ments that depend on the as-constructed quality
delivered for the project, when compared to the as- N Predict future life-cycle costs based on antici-
designed quality [FHWA, 1999]. pated performance. During simulated as-designed
and as-constructed performance over time, future
A performance-related specification is able to:
maintenance and rehabilitation costs can also be
N Examine multiple acceptance quality character- simulated by using an agency-selected mainte-
istics, including strength, slab thickness, air nance and repair plan. This step can predict the
content, initial smoothness, and consolidation most likely life-cycle costs for both as-designed
around dowels, or any combination of these and as-constructed pavements.
factors. Performance-related specification also
N Provide contractor pay adjustments based on
allows trade-offs to occur (in other words, superior
predicted as-designed and as-constructed life
quality measured for one acceptance quality
cycle costs. Using the predicted life-cycle cost as
characteristic can offset inferior quality for another
the single, overall quality characteristic for both
acceptance quality characteristic).
as-designed and as-constructed pavements can
N Include both acceptance quality characteris- determine final contractor pay adjustments.
tics’ mean and acceptable variability. With
PaveSpec 3.0, discussed earlier, is software used
performance-related specification, pavement
to determine the performance-related specification
performance predictions are a function of both
for a given project. Based on stored data, this pro-
acceptance quality characteristics means and
gram can be used to develop performance-related
standard deviation. Because traditional quality
specification and predict future performance of a
assurance specifications normally specify only the
pavement. It can be used to simulate performance
mean or a percentage within limits, these mea-
of as-designed and as-constructed pavement lots,
surements do not accurately indicate future
to project maintenance costs, and to develop and
performance.
administer quality-based pay adjustments.
N Define as-designed (target) quality characteris- Abbreviated from the FHWA-HIF-09-011—TechBrief: Performance-
tics. Performance-related specification requires Related Specifications for Portland Cement Concrete Pavements,
agencies to define desired quality measurements U.S. Dept. of Transportation – Federal Highway Administration, 2009

based on acceptance quality characteristics


QUAL I T Y AS S URANC E 120

are offered, the acceptance program specifies any quality-based pay adjustments to be
made [AASHTO, 1996, P. 13].
Acceptance testing is usually done by the funding agency rather than by the contractor,
although the contractor’s quality control tests can be incorporated into the agency’s results.
If so, proper statistical methods must be used to identify differences in test results.
Once statistical methods are established, there are several ways of determining
whether a batch of concrete is acceptable. One method is based on pass/fail criteria.
Initial screening tests use a pass/fail criterion to quickly catch “non-conforming” or “sub-
quality” materials before they are incorporated into the project. The advantage of a
screening test as a pass/fail test is that it normally does not require a payment relationship
since it prevents sub-quality materials from being incorporated into the project. How-
ever, if a small sample size is used, there is an increased risk of making an incorrect deci-
sion: Because there may be only a few test results for the pass/fail decision (usually one
or two [1 or 2]), they may be insufficient to allow a rigorous statistical analysis [FHWA,
2003, Chapter 6].
Another method of accepting quality characteristics is measuring the quality within a
given range. These methods are employed when the quality characteristic is used to deter-
mine payment. There are several preferred methods of quality measure, including percent
defective (PD) and percent within limits (PWL). Both of these methods (PD and PWL)
simultaneously measure both the average level and the variability in a statistically efficient
way [FHWA, 2003, Chapter 6].
Other quality measures include average absolute deviation (AAD), the moving aver-
age, and the conformal index (CI). Though some of these quality measures are more selec-
tive than others, choosing the most effective quality measure for a project can save
money—from reduced inspection or testing and/or from eliminating the amount of sub-
grade product incorporated into the project [FHWA, 2003, Chapter 6].
The acceptance range approach is considered the most practical by highway engi-
neers. While most industries will deem a lot either acceptable or unacceptable, this is sim-
ply not appropriate in Portland cement concrete paving construction due to the considerable
time and costs invested, and resources involved to remove and replace the pavement.
Therefore, engineers believe that it is more prudent to define a high level of acceptable
quality (AQL) and a lower level that would be considered clearly rejectable (RQL). Between
these two variations (AQL and RQL), the end product is not so defective as to require
removal and replacement of the pavement, nor does it warrant full payment, since the life-
cycle costs will increase due to earlier failure [FHWA, 2003, Chapter 6].
The testing sample can be either a split sample or an independent sample. A split sam-
ple is one that has been divided into two (2) or more portions representing the same mate-
rial. These are sometimes used to verify the acceptability of an operator’s test equipment
and procedure because the variability calculated from the differences is comprised solely
of testing variability. Independent samples are those taken without regard to any other
sample that may have also been taken to represent the material in question. These are
QUAL I T Y AS S URANC E 121

sometimes used to verify an acceptance decision because each sample contains indepen-
dent information regarding all sources of variability (such as materials, sampling, and test-
ing) [BURATI, WEED, HUGHES & HILL, P. 45].

PAYMENT STRUCTURE. Relating quality and performance to payment is a preferred


method of compensation. Basically, the plan is based on determining an acceptable quality
level and then adjusting the payment according to the number of defects above this level.
The pay structure may include the use of incentives/disincentives, minimum payment
provisions, remove-and-replace provisions, and retest provisions. Performance-based
payments are usually combined with a payment factor that specifies a given percentage of
payment depending on how well specifications are met. The acceptable quality level should
be set such that it yields an expected payment of 100% of the price bid. When the reject-
able quality level is used with a payment factor, the agency must decide whether to require
removal and replacement or the assignment of a minimum payment factor [BURATI, WEED,
HUGHES & HILL, 1996, P. 99].
In a typical agreement:

…if production is uniform and no more than 1 out of every 20 quality assurance
tests fails, the process can be assumed to be in control and additional process con-
trol testing (and other actions) are not indicated.

If defects rise to 10% (1 out of 10 quality assurance tests fail), that may indicate
additional process control testing and other actions are needed.

If defects rise to 15% or higher (more than 1 out of 6 quality assurance tests fail),
that is approximately equivalent (depending on sample size) to a pay factor of less
than 0.90. This indicates serious process control problems, and the agency may
require that process to be suspended while the contractor modifies the process
control procedure (including testing) to address the problem. [FHWA, CONSTRUC-
TION, 1998]

Endnotes
AASHTO, “Implementation Manual for Quality Assurance” (IMQA-1)—a Report of the AAS-
HTO Highway Subcommittee on Construction, prepared by a Joint Construction/Materials
Quality Assurance Task Force of the AASHTO Highway Subcommittee on Construction,
February 1996.

American Concrete Institute, ACI Standard 318. Farmington Hills, Michigan. <www.concrete.org>

Burati, J. L., R. M. Weed, C. S. Hughes, H. S. Hill. FH WA -RD-0 2 -0 9 5 —O ptimal Procedures for


Q uality A ssurance Specifications, Office of Research, Development, and Technology, U.S.
Dept. of Transportation – Federal Highway Administration, 1996. <www.tfhrc.gov/pave-
ment/pccp/pubs/02095/02095.pdf>
QUAL I T Y AS S URANC E 122

“Construction Program Management and Inspection Guide, Appendix B (Part 2),” Contractor
Q uality Control Plans: Contractor Guidelines and Example, U.S. Dept. of Transportation –
Federal Highway Administration, February 1998. <www.fhwa.dot.gov/construction/
cpmi04b2.cfm>

HIPERPAV (High-Performance Paving Software), 2010. <www.hiperpav.com>

“PaveSpec User’s Guide,” U.S. Dept. of Transportation – Federal Highway Administration, March
2007. <www.fhwa.dot.gov/PAVEMENT/pccp/pavespec/index.cfm>
ECOLOGICAL
POSITIONING

THE PRODUCTION OF CONCRETE has a significant effect on the environment.


Although concrete itself is a renewable and recyclable material, which has advantages that
far outnumber those of competitive pavements, the cement component represents signifi-
cant impact on the environment. Even so, concrete offers many other environmental ben-
efits and tradeoffs that are to be considered in assessing the overall ecological position of
concrete as a pavement medium.
The ingredients that go into cement must be heated to 1,450°C, and the fossil fuel con-
sumed in the process is the third largest source of greenhouse gas emissions in the United
States. As a general rule, producing 1 ton of cement means releasing 1 ton or more of CO2
(carbon dioxide), which translates to up to 5% of total CO2 emissions. (2% of the world-
wide carbon footprint is due to cement production alone.) This presumed contribution to
global warming is no longer mainly the responsibility of the United States and other West-
ern nations. Consider that while U.S. cement production peaked at 128 million metric
tons in 2006, China set an all-time world record the following year, producing an astound-
ing 1.3 billion tons of cement [BIELLO, 2008] [CONCRETE PRODUCER, 2008].
According to the Green Highways Partnership, “Current standards for concrete pro-
duction are not sustainable,” given the industry’s enormous requirements for limestone
and fossil fuels [GREEN HIGHWAYS PARTNERSHIP, 2007]. The partnership involves the Fed-
eral Highway Administration, the EPA (U.S. Environmental Protection Agency), and state
agencies, as well as industries and local contractors. It has targeted a number of strategies
to foster environmentally friendly road building, particularly the use of recycled materials
[WISCONSIN D.O.T., 2008] [FHWA, 2008].

The Greening of Concrete


Cement producers have taken on the challenge of lowering their impact on the environ-
ment. According to American Concrete Pavement Association, the industry has lowered
the amount of energy required to make a ton of cement by 33% since 1972, with the goal
of reducing that figure another 10% by 2020 [WATHNE, 2009, PP. 6, 8].
123
E COL OGI CAL P OS I T I ONING 124

One environmental advantage of concrete highway construction has long been


acknowledged: It is highly durable, lasting as long as 40 years without energy-intensive
repairs or replacement. There are other advantages of concrete paving as well.

N Industrial wastes such as fly ash and blast-furnace slag are increasingly used as
supplementary cementitious materials, strengthening concrete and reducing its
cost. At the same time, landfills are spared an enormous burden.
N Concrete itself can be recycled, further saving landfill space and also reducing the
need to extract raw aggregates.
N Trucks and cars get better fuel economy on concrete than on asphalt [TAYLOR AND
PATTEN, 2006, PP. 1–9 (trucks); SUMITSAWAN, 2009. P. 1 (cars)].
N The development of permeable concrete paving allows the infiltration of rain-
water, helping to reduce runoff in urban areas and reducing the ecological impact
of new roads in sensitive areas.
N Concrete pavements absorb less heat than dark-colored materials such as asphalt,
and thereby minimize the adverse effects of urban heat islands.

Industrial Waste as Supplementary Cementitious Materials


Various types of industrial waste have been used successfully in concrete highway mixes,
replacing traditional materials that are more expensive and also involve the environmental
disturbance of extraction. The most commonly used waste product is fly ash, a plentiful
byproduct of burning coal to generate electrical energy. Fly ash is composed of the non-
combustible mineral impurities in the coal. To date, Europe uses a greater percentage of its
fly ash in concrete than the United States does. The American Coal Ash Association reports
that in 2008, roughly 22% of the 65.7 million metric tons of fly ash produced in the U.S.
was used in the production of concrete [AMERICAN COAL ASH ASSN., 2009]; in Europe,
where 40.4 million metric tons were produced, the figure was roughly 35% [ECOBA].
The benefits of using 1 ton of ash to replace Portland cement are impressive:

N The energy saved would provide the electricity for an average American home for
24 days.
N The landfill space thus spared could handle the solid waste generated by an
average American over a period of 455 days.
N The atmosphere is spared the equivalent of two (2) months’ worth of automobile
CO2 emissions.

A significant percentage of Portland cement can also be replaced by ground, granulated


blast-furnace slag. This byproduct of steel manufacturing is created when iron ore is con-
verted to iron. Molten slag is ground into a fine powder, and in this form it can act as a
hydraulic cement. The potential of this material is well established, having been used in con-
crete mixtures for over a century. Up to 50% of the Portland cement in a concrete mix can
E COL OGI CAL P OS I T I ONING 125

be replaced by slag. When used properly, this material has the potential to increase strength,
enhance durability, and help minimize the destructive effects of de-icing compounds.
Viewed in terms of life-cycle advantages over producing conventional concrete, a cubic
meter of slag concrete can mean a up to a 59% reduction in CO2 emissions, while using
some 42% less embodied energy [PORTLAND CEMENT ASSN., DESIGNED TO DELIVER, 2009].
The Wisconsin Highway Research Program tested the performance of both 30% and
50% slag aggregates, using grade-120 slag, a high-quality grade. Curing times were slower:
While regular concrete cures in three to four (3–4) days at 21°C, 30% slag aggregate takes
four to six (4–6) days and 50% slag aggregate needs six to eight (6–8). Concrete cures more
slowly as the temperature drops, and this difference becomes more pronounced as the per-
centage of slag in the mix goes up. This may mean that slag concretes are best used in
warmer months in some regions. While there have been concerns that concrete made
with slag might not be sufficiently strong, the Wisconsin study found that its tensile-
compressive strength was equivalent to that of conventional mixes when grade-120 mate-
rial was used [WISCONSIN D.O.T., HIGHER-GRADE SLAG, 2008].
Silica fume (also known as microsilica) is a byproduct of producing silicon and ferro-
silicon alloys in electric furnaces. It takes the form of an extremely fine powder, and until
the 1970s it routinely was dispersed into the atmosphere. Because of its small particle size
and high silicon content, it has pozzolanic properties and has been found to improve con-
crete’s strength and increase resistance to abrasion. It makes concrete less permeable, and
consequently the reinforcing steel within the roadbed isn’t as vulnerable to corrosion. Up
to 15% of the cement can be replaced with silica fume, but the usual range is from 7 to 10%
[FHWA, SILICA FUME, 2008].
Unlike the above, rice-husk ash is an agricultural waste material. Worldwide, some 100
millions tons of husks are produced each year, and after combustion about 20% of that is left
as a highly siliceous ash with pozzolanic properties; fully burned rice-husk ash may have a
silica content of from 90% to 95%. The use of rice ash in concrete has yet to be fully explored,
but is more common in rice-growing regions. (9 of the top 10 producers are in Asia, with
China being the largest.) In Brazil, the world’s tenth largest producer of rice grain, research
has found that the use of rice-husk ash reduces the water needed in the mix without reduc-
ing splitting tensile strength. The modulus of elasticity declined, however, as the rice-husk
content increased. [TASHIMA, DA SILVA, AKASAKI AND BARBOSA, 1997, PP. 2–9].
After Brazil, the next 10 of the top producers include the U.S., European countries,
and African countries, where the use of rice-husk ash isn’t common [deTWILER, BHATTY,
AND BHATTACHARJA, 1996, PP. 27–28].

Recycled Concrete
As durable as they are, concrete roadways do need to be repaired or replaced over time,
and an increasing percentage of the resulting waste material is being recycled as aggregate
for new pavement. The environmental advantages are substantial, in terms of reducing the
E COL OGI CAL P OS I T I ONING 126

need for mined materials and also trucking less debris to overburdened landfills. As an
example of using old roads to build new ones, the concrete in a stretch of highway east of
Los Angeles, dating back to 1946, has been ground up and used for repaving two (2) times
over the years [WATHNE, 2009, P. 11].
Nevertheless, the U.S. Federal Highway Administration reports that the main use of
recycled concrete aggregate has been as a base material because there has been some resis-
tance to using it in concrete paving [FHWA, 2008]. But in a trial program to repair and main-
tain its roadways and parking areas, the U.S. Department of Defense (D.O.D.) has found
that it is economically feasible to require the
contractor to use recycled materials. Accord- Co n c r e t e As a n Ai r Fi l t e r
ing to D.O.D.’s report, contractors do not
have to be offered a price differential to use A somewhat surprising environmental advantage of
environmentally preferable products, in part using concrete for paving is that the material is known
because they may be willing to explore green to absorb CO2. While it’s estimated that 5% or more of
the global carbon footprint is generated by the con-
alternatives in order to maintain a competi- crete industry, minerals in concrete will combine with
tive edge as the industry becomes more envi- CO2 in the atmosphere to form calcite (or calcium car-
ronmentally attuned [EPA, 2007]. bonate, CaCO3). This chemical reaction can continue
for decades after the concrete is produced, and it
Some energy is required to process extends below the surface because of concrete’s per-
recycled aggregate concrete. There is the meability. While only an estimated 5% of the CO2
fuel consumed in trucking the demolished released by cement production may be reabsorbed,
“when summed globally [the numbers] become sig-
concrete to a recycling plant and then to the
nificant,” according to Liv Haselbach, a civil and
job site (or moving a portable crusher to the environmental engineering professor at Washington
site). It also takes energy to crush and screen State University [AGGREGATE RESEARCH, 2009].
the old concrete. But studies show that
energy is saved by using recycled aggregate
concrete even if the alternative would be to transport new extracted aggregates only a few
miles. The greater the distance the new materials have to be hauled, the greater the sav-
ings [N.Y. STATE D.O.T., 2009].

Concrete for Better Fuel Economy


Several studies have shown that vehicles get more miles or kilometers per gallon or liter on
concrete roadways than on other surfaces. The improved mileage is attributed to the rela-
tive smoothness of concrete and its resistance to flexing [PCA]. A study at the University of
Texas found that under dry surface conditions, cars traveling on concrete at 30 miles per
hour (18.6 kph) use 10% less fuel than on than asphalt. The study concluded that the
potential savings in terms of fuel consumed and CO2 emissions over the life of a project
could be substantial and should be considered in both life-cycle cost analysis and the car-
bon footprint of a roadway project. In the Dallas–Fort Worth area alone, the annual sav-
ings were estimated to be 670 million liters (177 million gallons) of fuel, with CO2 emissions
reduced by 0.62 million metric tons [SUMITSAWAN, 2009, P. 1].
E COL OGI CAL P OS I T I ONING 127

Canada’s National Resource Council reported that the fuel efficiency of trucks was
4.3% to 9.2% better on concrete pavement than on asphalt. While the improvement varied
with season and vehicle speed and weight, the major factor was the type of road surface
[TAYLOR AND PATTEN, 2006, PP. 1–9]. It has been estimated that the resulting reduction in
vehicle-generated pollutants can compensate for the CO2 generated by producing the con-
crete within nine (9) years of placing it [WATHNE, 2009, P. 20].

Minimizing Runoff with Pervious Concrete


Pervious concrete allows some of the precipitation falling on it to pass through, reducing
the amount of runoff characteristic of paved surfaces and thereby creating less environ-
mental impact along roadways and in areas adjacent to parking lots. Specifically, this sur-
face can minimize storm-water runoff to surrounding streams and lakes, while recharging
groundwater supplies. Pervious concrete has even been found to help filter out pollutants
as the water percolates [PORTLAND CEMENT ASSN., PERVIOUS CONCRETE, 2009] [ALABAMA
COOPERATIVE EXTENSION, 2004].
Because some of the water falling on pervious parking areas percolates into the ground,
there is less need for the discharge basins that have become a common sight around newer
construction. There are safety advantages as well. Because water isn’t as likely to pool on
the surface, there may be less danger of hydroplaning, less icing in winter, and reduced
highway glare [ARIZONA STATE UNIV., 2009].

Ex p e r ime n t a lCo n c r e t e s

While standard concrete mixes will reabsorb a mod- naturally by coral. The company says that each ton
est amount of CO2 over a period of years, Novacem, of their cement can be expected to sequester a
a British company, is developing a cement that half-ton of CO2 [BIELLO, 2008].
could prove to be carbon-neutral or even carbon- An Italian company, Italcementi, has devel-
negative. It is made with magnesium oxide instead oped a cement product that turns NO and NO2—the
of calcium carbonate, which doesn’t emit the CO2 main contributors to photochemical smog—into
released from calcium carbonate, and which allows harmless compounds. Their cement contains
making cement at lower kiln temperatures. The nanoparticles of titanium dioxide, a photocatalyst
company projects that each ton of its cement will that harnesses the ultraviolet part of sunlight to
be able to capture and sequester 0.75 ton of CO2 accelerate the oxidation of pollutants. The process
but full-scale production isn’t slated until 2013 forms reactive hydroxyl radicals that in turn inacti-
[NOVACEM] [GREEN TECH ENTERPRISE, 2009]. vate pollutant molecules. The product has potential
The California-based Calera Corporation is for both roadways and paralleling sound walls.
working on a process to turn power plant flue gases Despite the promise of this technology, a study by
into cement or aggregates. By bubbling the flue the Lawrence Berkeley Laboratory for the California
gases through seawater, where the CO2 reacts with Energy Commission reports that there currently is
calcium and magnesium, Calera produces a little work on it underway in the United States [CALI-
cement similar to the marine cement generated FORNIA ENERGY COMMISSION, 2008].
E COL OGI CAL P OS I T I ONING 128

The pervious quality is built into the mix by using little or no fine aggregate and only
enough cement paste to coat the aggregate. The result is a network of interconnected voids
comprising 15% to 35% of the volume, allowing water to drain quickly. In colder regions,
freeze-thaw cycling can damage pervious concrete because of the expansion of the water
contained within the pavement, and extra care must be taken to allow water to drain from
the roadbed [PORTLAND CEMENT ASSN., WEBINAR, 2009].

Cooling Urban Heat Islands


Cities have a large thermal mass that absorbs solar energy to such an extent that these
areas have been called heat islands, with temperatures considerably above those of the sur-
rounding countryside. In warm months, this effect drives peak periods of energy demand
for air conditioning, with a consequent increase in air pollution and greenhouse gas emis-
sions [EPA, reducing urban HEAT ISLANDS, 2008, P. 1].
Concrete pavement absorbs solar energy and heats the air above, but it is more reflec-
tive than asphalt and therefore the cooler alternative. The solar reflectance of conventional
concrete can reach 40%, and lighter mixes and white-topping can raise reflectance to 70%
[EPA, COOL PAVEMENTS, P. 16]. According to the estimates in one study, increasing pave-
ment reflectivity could save the city of Los Angeles roughly $90 million annually.
While concrete is cooler than asphalt, its advantage decreases over time because traf-
fic gradually darkens its surface so that more solar energy is absorbed [EPA, COOL PAVE-
MENTS, P. 8].
Another available approach to cooler pavements is to make them pervious; when wet,
these roadways lower the temperature of the surrounding air through evaporative cooling.
Research is ongoing on the effect that these pavements have when the surface is dry [EPA,
COOL PAVEMENTS, P. 8].

Endnotes
2008 Coal Combustion Product (CCP) Production and Use Survey Report. American Coal Ash
Association, Aurora, Colorado, 2009. <www.acaa-usa.org/displaycommon.
cfm?an=1&subarticlenbr=3.>

“Announcement and Call for Papers.” U.S. Dept. of Transportation – Federal Highway Adminis-
tration, 2009 <www.fhwa.dot.gov/pavement/concrete/2010acptpconf.cfm>

Biello, David. “Cement from CO2: A Concrete Cure for Global Warming?” Scientific American,
August 7, 2008. <www.scientificamerican.com/article.cfm?id=cement-from-carbon-dioxide>

Cement and Concrete Basics. Portland Cement Association, Skokie, Illinois, 2009. <www.cement.
org/basics/>

“The China (Cement) Syndrome.” Concrete Producer, Chicago, 2008. <www.theconcretepro-


ducer.com/industry-news.asp?articleID=762241&sectionID=1419>
E COL OGI CAL P OS I T I ONING 129

“Concrete Found to Absorb CO2, Making It More Environmentally Friendly Than First Thought.”
Aggregate Research, Salida, Colorado, 2009. <www.aggregateresearch.com/article.
aspx?id=16260>

“ Cool Pavements.” U.S. Environmental Protection Agency, Washington, D.C. <www.epa.gov/hiri/


resources/pdf/BasicsCompendium.pdf>

Detwiler, Rachel J., Javed I. Bhatty, and Sankar Bhattacharja. “Supplementary Cementing
Materials for Use in Blended Cements,” Research and Development Bulletin RD112. Portland
Cement Association, 1996.

“D.O.D. Paving Materials Highlight Environmental Attributes.” U.S. Environmental Protection


Agency, Washington, D.C., 2007. <www.epa.gov/epp/pubs/case/dod_pave.htm>

“Environmentally Friendly Concrete.” Green Highways Partnership, 2007. <www.greenhighways.


org/Template.cfm?FrontID=5060>

“Evaluation of Titanium Dioxide As a Photocatalyst for Removing Air Pollutants.” California


Energy Commission, January 2008. <http://74.125.93.132/search?q=cache:q9zWsTYl8pMJ:
www.energy.ca.gov/2007publications/CEC-500-2007-112/CEC-500-2007-112.PDF>

Green Highways. U.S. Dept. of Transportation – Federal Highway Administration, 2008. <www.
tfhrc.gov/focus/apr08/01.htm>

“Higher-Grade Slag Helps Wisconsin Concrete Pavements Go Green.” Wisconsin Department of


Highways, Madison, Wisconsin, 2008. <http://on.dot.wi.gov/wisdotresearch/database/briefs/
05-01slagexpanded-br.pdf>

Kaloush, Kamil, Tom Kaczmarowski, and Stew Waller. “Pervious Concrete Pavement.” Tuscon:
Arizona State University, 2009. <www.acecaz.org/RandS/2009-presentations/Friday/
Indigo/1100/Pervious Concrete Roads and Streets 2009 ASU.pdf>

“New Concretes are Designed to Deliver.” Portland Cement Association. Skokie, Illinois, 2009.
<www.cement.org/buildings/sustainable_mix.asp>

“Novacem Eyes Carbon-Neutral Cement.” Green Tech Enterprise, Cambridge, Massachusetts,


2009. <www.greentechmedia.com/articles/read/novacem-eyes-carbon-neutral-cement>

“Pervious Concrete,” PCA Webinar. Portland Cement Association, Skokie, Illinois, 2009. <www.
cement.org/tech/WB018_pervious.htm>

“Pervious Concrete.” Portland Cement Association, Skokie, Illinois, 2009. <www.cement.org/tech/


cct_con_design_pervious.asp>

“Production and Utilisation of CCPs in 2006 in Europe (EU 15),” ECOBA (European Coal
Combustion Products Association). <www.acaa-usa.org/displaycommon.
cfm?an=1&subarticlenbr=3>

“Quantify the Energy and Environmental Effects of Using Recycled Asphalt and Recycled
Concrete for Pavement Construction.” New York State Department of Transportation, 2009.
<https://www.nysdot.gov/divisions/engineering/technical-services/trans-r-and-d-repository/
C-08-02 Final Report 9-8-09.pdf>

“Recycled Concrete Aggregate,” Pavements – Recycling. U.S. Dept. of Transportation – Federal


Highway Administration, 2008. <www.fhwa.dot.gov/pavement/recycling/rca.cfm>

Reducing Urban Heat Islands: Compendium of Strategies. U.S. Environmental Protection Agency,
Washington, D.C, 2008. <www.epa.gov/hiri/resources/compendium.htm>
E COL OGI CAL P OS I T I ONING 130

“Rice Crispy” Concrete Safeguarding Many Communities from Storm Water Runoff. Alabama
Cooperative Extension, Auburn, Alabama, 2004. <www.aces.edu/dept/extcomm/newspaper/
jan22a04.html>

“Silica Fume.” U.S. Dept. of Transportation – Federal Highway Administration, 2008. <www.fhwa.
dot.gov/infrastructure/materialsgrp/silica.htm>

Sumitsawan, Palinee, Stephan Romanoschi, and Siamak A. Ardekani. “Effect of Pavement Type
on Fuel Consumption and Emissions,” Proceedings of the 2009 Mid-Continent Transporta-
tion Research Symposium, Iowa State University, Ames, Iowa, 2009. <www.intrans.iastate.
edu/pubs/midcon2009/SumitsawanEmissions.pdf>

Tashima, Mauro M., Carlos A. R. da Silva, Jorge L. Akasaki, and Michele Beniti Barbosa. “The
Possibility of Adding the Rice Husk Ash.” Congresso Brasileiro do Concreto, 2007. <http://
congress.cimne.upc.es/rilem04/admin/Files/FilePaper/p282.pdf>

Taylor, Gordon, and Jeff Patten. Effects of Pavement Structure on Vehicle Fuel Consumption,
Centre for Surface Transportation Technology, National Research Council of Canada,
Ottawa, 2006. <www.mne.psu.edu/ifrtt/ConferenceProceedings/ISHVWD_9_2006/docs/
pdfs/session 2/s2-1 91.pdf>

Use of Fly Ash in Full-Depth Reclamation. Wisconsin Department of Transportation, 2008.


<http://on.dot.wi.gov/wisdotresearch/database/tsrs/tsrflyashfulldepth.pdf>

Wathne, Leif. Concrete Pavements and Sustainability, 39th MCPA Workshop, Feb. 19, 2009. Pages 6,
8, 19, and 20. <www.durableroads.com/documents/ConcretePavementsandSustainability-
Wathne.pdf>
MAINTENANCE

THE MAJORITY OF THIS STUDY HAS SHOWN how to produce high-quality con-
crete pavements—produced with high-quality concrete ingredients in the correct ratios,
reinforced as necessary, and enhanced with appropriate admixtures. Well-built concrete
pavement will often last for 30 or more years before repairs are needed. It will not, how-
ever, last forever.
An important part of any pavement maintenance program is the periodic post-
construction evaluation of the road surface, as well as addressing problems as they occur.
Diagnosis of concrete pavement problems is covered at the beginning of this chapter. The
two (2) most basic maintenance procedures—sealing and resealing—are covered at the
end. Work that requires disturbing or replacing the road surface is classified as repair, and
is the topic of the next chapter (Highway Repairs).

D eveloping a Maintenance Plan


Approaching pavement maintenance in an organized manner saves time and money.
Often, the sooner a problem is identified and resolved, the less it will cost to repair. Main-
tenance begins with a systematic survey of the pavment—some parts of which are likely to
be sound, while others are likely to be in various stages of decay. Inspection itself ranges
from visual to monitoring with sophisticated instruments. Once the roadways have been
examined and the problems described, schedules and budgets can be developed for mak-
ing the most effective repairs. In most cases, it is not possible to make all repairs in a single
year. A ratings system helps to determine which problems need to be addressed first, and
which can be put on an ongoing maintenance schedule.
No matter how the inspection is conducted, it’s well worth the time and trouble to
conduct a thorough field inspection of a pavement. If possible, the same team of inspec-
tors should undertake the inspection of the entire roadway, for instance, so that rating
numbers will be consistent. The report should be written and dated, both to keep track of
the rate of deterioration and to document conditions to funding agencies.

131
M AI NT E NAN C E 132

For purposes of the inspection, the pavement should be divided into sections that are
in similar condition. Depending on circumstances, the section may be several kilometers
long, or it may be measured in meters where damage is localized. It’s best to give average
ratings for a fairly long section, rather than dividing it into many small sections. For
instance, a section may be described as needing spot repairs and occasional crack-filling.
Drainage issues are especially important. These are best examined soon after a rain, or
snow melt, but a visual inspection of a dry road can yield helpful information: A well-
drained road is crowned so that the middle is at least 2% higher than the sides. Shoulders
should be well sloped away from the highway. Depressions and low spots where water
puddles are signs of trouble that should be repaired as soon as possible. Culverts and
storm drains should flow freely to avoid puddling that will damage the road.

Pavement Ratings
The simplest tool for measuring pavement condition is the PAvement Surface Evaluation
and Rating System (PASER), developed by the Transportation Information Center at the
University of Wisconsin–Madison. Roads are rated visually on a scale from 1 to 10, start-
ing with ratings of 9 and 10 as conditions that require no maintenance. The scale pro-
gresses downward through the lower numbers to 1 and 2, conditions that require complete
reconstruction [WALKER, 2002, P. 15].
In cases of severe problems, heavy traffic, or poor construction, a pavement may go
directly from 9 or 10 to 3 or even 2 or 1. But even when the pavement is built well and
traffic is moderate, it is not uncommon for deterioration to proceed rapidly, and proceed
from one stage to the next in a matter of weeks. The rating system is described in Figure
15-1. Evaluation depends on signs of visible distress, described in the section “Identify-
ing Defects.”

Measuring Eq uipment
While visual examination is simple and effective, sophisticated equipment that evaluates
pavement conditions can save time and can lead to better maintenance procedures.
Although such equipment is expensive, its speed and effectiveness can save money over
the long run. No one machine monitors all conditions, and a combination of monitors
should be used for best results.
V ehicle-b ased inertial road-p rofiling systems ( IRPS) use sensors—laser, infrared,
optical, or ultrasound—to monitor irregularities in the road as the vehicle travels along it.
Results are usually recorded in terms of the International Roughness Index (IRI), a mea-
surement of variations in the road profile over a given distance. The Pennsylvania Depart-
ment of Transportation, for example, uses equipment that measures the up and down
movement of a laser as it is driven along the highway. The scale is open-ended, with a
lower IRI indicating a smoother road.
M AI NT E NAN C E 133

A falling-w eight deflectometer ( FWD ) is used to measure the deflection and struc-
tural integrity of the pavement. It is a stationary device that drops a large weight on the
road surface, and measures the deflection at the point of impact, as well as at specific inter-
vals. A rolling-w eight deflectometer ( RWD ) uses a laser to measure deflections created
by a truck traveling at highway speeds.
L ock ed wheel trailers are used to measure skid resistance of a wet road surface. The
trailer travels along the road at highway speed, and one of the wheels is stopped from turn-
ing (or “locked” in place) repeatedly while a measured amount of water is sprayed down
in front of it. A sensor in the wheel measures skid resistance as the force between the skid-

Figure 15-1

PASER Ra t in g Sy st e m

Surface
Visible Distress* General Condition/Treatment Measures
Rating

10 None New pavement. No maintenance


Excellent required.

9 Traffic wear in wheel path. Recent concrete overlay or joint


Excellent Slight map cracking or pop-outs. rehabilitation. Like-new condition.
No maintenance required.

8 Pop-outs, map cracking, or minor surface defects. Slight surface scaling. Partial More surface wear or slight defects.
Very Good loss of joint sealant. Isolated meander cracks, tight or well sealed. Isolated Little or no maintenance required.
cracks at manholes, tight or well sealed.

7 More extensive surface scaling. Some open joints. Isolated transverse or First sign of transverse cracks (all
Good longitudinal cracks, tight or well sealed. Some manhole displacement and tight); first utility patch. More
cracking. First utility patch, in good condition. First noticeable settlement or extensive surface scaling. Seal open
heave area. joints and other routine maintenance.

6 Moderate scaling in several locations. A few isolated surface spalls. Shallow First signs of shallow reinforcement
Good reinforcement causing cracks. Several corner cracks, tight or well sealed. Open or corner cracking. Needs general
(6 mm-wide) longitudinal or transverse joints and more frequent transverse joint and crack sealing. Scaled areas
cracks (some open 6 mm). could be overlaid.

5 Moderate to severe polishing or scaling over 25% of the surface. High reinforcing First signs of joint or crack spalling
Fair steel causing surface spalling. Some joints and cracks have begun spalling. First or faulting. Grind to repair surface
signs of joint or crack faulting (6 mm). Multiple corner cracks with broken defects. Some partial depth patching
pieces. Moderate settlement or frost-heave areas. Patching showing distress. or joint repairs needed.

4 Severe polishing, scaling, map cracking, or spalling over 50% of the area. Joints Needs some full-depth repairs,
Fair and cracks show moderate to severe spalling. Pumping and faulting of joints grinding, and/or asphalt overlay to
(1 cm) with fair ride. Several slabs have multiple transverse or meander cracks correct surface defects.
with moderate spalling. Spalled area borken into several pieces. Corner cracks
with missing pieces or patches. Pavement blowups.

3 Most joints and cracks are open, with multiple parallel cracks, severe spalling, Needs extensive full-depth patching
Poor or faulting. D-cracking is evident. Severe faulting (2.5 cm) giving poor ride. plus some full-slab replacement.
Extensive patching in fair to poor condition. Many transverse and meander
cracks, open and severely spalled.

2 Extensive slab cracking, severely spalled and patched. Joints failed. Patching Recycle and/or rebuild pavement.
Very Poor in very poor condition. Severe and extensive settlements or frost heaves.

1 Restricted speed. Extensive potholes. Total reconstruction.


Failed Almost total loss of pavement integrity.

*Individual pavements will not have all of the types of distress listed Source: Walker, Donald. PASER Concrete Roads Manual, Madison, Wisconsin:
for any particular rating. They may have only one or two types. Transportation Information Center, University of Wisconsin–Madison, 2002.
<www.dot.state.il.us/blr/p016.pdf >
M AI NT E NAN C E 134

ding wheel and pavement. Skid resistance typically increases during the first two (2) years
of road life, but decreases thereafter as traffic polishes the aggregate. If resistance becomes
too low, the likelihood of accidents increases, and repairs are required [WASHINGTON
STATE D.O.T., SKID RESISTANCE, 2005].

Identifying D efects
Most roadway problems are visible to the naked eye, though in some cases, further testing
will be required. Typical problems, identified by the Transportation Information Center at
the University of Wisconsin–Madison, include the following [WALKER, 2002, PP. 6–15]:

MEANDER CRACKS. Cracks that wander in various direc-


tions are usually localized, and do not indicate general
problems with the slab. They may be due to settlement of
the substrate, or an incorrectly sloped road that does not
drain water away from the area. As with any crack, the lon-
ger it is left unsealed or unrepaired, the wider and deeper it
will become.

HAIRLINE CRACKS. Also called map cracking, this pattern


of fine cracks may be caused by improper finishing or cur-
ing too quickly. By itself, this condition is not unsafe, since
the cracks are shallow. They should, however, be inspected
regularly to be sure the cracks are not growing, spalling, or
scaling. If that occurs, an overlayment is called for.

TRANSVERSE CRACKS. Cracks running across a road can


be due to several things. They may be due to control joints
that are too far apart or to a poorly compacted substrate.
Truck traffic that is too heavy for the road can also cause
transverse cracks. Often the edges on each side of a crack
will develop spalling. A single crack can be patched and
sealed. A series of parallel transverse cracks, even if they
are not wide, means that the slab is broken in that area and
should be demolished and replaced.
M AI NT E NAN C E 135

CRACKS AT MANHOLES AND INLETS. Where concrete


abuts a storm sewer inlet or other metal object, expansion
and contraction often creates cracks and faults. These prob-
lems are usually localized. Sealing of cracks and of the joint
between concrete and metal may provide a good temporary
solution. Often, however, full-depth repairs are needed.

SHOULDER PROBLEMS. In some cases, a concrete high-


way may have a shoulder made of asphalt or even gravel.
The longitudinal joint between concrete and asphalt needs
to be sealed regularly. A white-topping overlay will firm up
an asphalt shoulder. In some cases, roller-compacted con-
crete (RCC) has been used.

POP-OUTS. Pop-outs are a series of roundish indentations


spread across the slab surface. They may be caused by poor
finishing, or by a poor concrete mix that contains chert
and other absorbent materials which react to freeze-thaw
cycles. By themselves, pop-outs do not make a road unsafe.
However, if the condition leads to spalling or scaling, an
overlay is needed.

BLOWUPS. When a joint is left open for a long period,


blowups can occur. Sand or fine gravel infiltrates the joint,
and when expansion occurs due to temperature change,
one or both sides of the joint will buckle. Often, this condi-
tion occurs where joints are spaced too far apart. The
uplifted portion should be removed, and new concrete
poured. Cutting new joints helps relieve pressure.
M AI NT E NAN C E 136

OPEN JOINTS. Both longitudinal and transverse joints


should be narrow, and should be filled and sealed. If a seal
is compromised, an opening will occur, at which point
freeze-thaw cycles and heavy traffic have an opportunity to
crack the surrounding area. A vicious cycle of wider open-
ings and further damage can then lead to serious damage
to the surface, and may well compromise the entire thick-
ness of the slab. Cracks around an open joint will only grow
in time; the cracks and spalling must be repaired, and the
joint must be sealed.

FAULTED JOINTS. If the slab on one (1) side of a joint is


raised more than 13 mm above the adjoining slab, the joint
is considered “faulted.” This makes for a bumpy ride in a
car or truck, and it also presents an opportunity for further
damage due to traffic or freeze-thaw cycles. Sealing the
joint may prevent further damage. If the fault becomes
severe, slab jacking may be the solution. In some cases, the
area may need to be demolished and replaced with a new
slab that has dowelled joints.

SCALING. Here, the mortar and the fine aggregates come


loose at the surface. Often it is the result of using concrete
that is not air-entrained, in a region subject to freezing. In
some cases, poor concrete or poor finishing technique can
remove air-entrainment bubbles, leading to scaling. De-
icing chemicals can also cause scaling. If the problem is
only at the surface, repairs may not be needed. However, if
large aggregate is coming loose, or if the scaling leads to
other problems, the problem should be addressed. If the
scaling area is large, the solution may be overlayment. If it
is small, partial-depth patching is often the solution.
M AI NT E NAN C E 137

WEAR AND POLISHING. If the aggregate is clearly exposed,


the surface mortar has worn away. This actually has the
effect of making a road less skid-resistant. Water can col-
lect in ruts, which can cause hydroplaning. Often the aggre-
gate will be worn smooth, making the surface dangerously
slick. The solution may be to grind the concrete, or to apply
an overlay.

SPALLING. Spalling occurs when chunks of the surface


concrete come loose. It may occur in the middle of a slab,
or along joints or the edge of a road. If a slab has a fairly
wide crack, the edges alongside it may develop spalling.
Small spalling areas in the middle of a slab can be patched.
Spalling along a crack or a damaged joint should be repaired
at the same time as the crack or joint is repaired.

EXPOSED METAL REINFORCING. This is usually caused


by rebar or wire mesh that wasn’t placed deep enough in
the slab. Spalling occurs, exposing the metal, which often
rusts. In some cases, rust stains are visible on the surface
before the spalling occurs. Exposed metal reinforcing is
often difficult to repair. The slab may need to be demol-
ished in the damaged area and replaced with new rein-
forced concrete.

D-CRACKS. A poor concrete mix causes D-cracks—usually


because the aggregate contains stones that absorb too much
moisture. D-cracking usually starts at the bottom of the
slab and moves up to the surface. A dark discoloring and a
series of hairline cracks are telltale signs. Once the cracks
rise to the top, the entire thickness of the slab has been
compromised. Sealing the joint may help, but in most
cases the area needs to be demolished and new concrete
poured.
M AI NT E NAN C E 138

CORNER CRACKS. Typically, a corner crack occurs where


a transverse joint meets a longitudinal joint, with the crack
connecting them forming a triangle. Often, two (2) or
more of these cracks occur at the same joint. The severity
of a corner crack can be difficult to determine. Some cor-
ner cracks are only surface-deep, while others penetrate to
the bottom of the slab. Still others angle down toward the
joints, so that a chip may pop out. Depending on the extent
of the damage, a partial- or full-depth patch may be neces-
sary, and the joint may need to be replaced. In severe cases,
it is necessary to demolish and replace an entire area.

PAVEMENT HEAVE AND SETTLING. If the substrate or a


utility trench is not well compacted, the slab may generally
settle during warm weather, and heave upward during
heavy frost. In most cases, the concrete needs to be
removed, the substrate compacted, and a new slab poured.

POTHOLES. Potholes are localized. They may be caused by


a poor substrate or concrete materials, or by poor drainage
that leaves standing water in the spot. Spalling that is left
unattended can also lead to a pothole. Full-depth patching
is usually the best fix.

PATCHES THAT NEED TO BE PATCHED. All too often,


patches are done quickly and under duress. In some cases,
the damaged area is not completely removed, so some
cracked concrete remains. In other cases, a concrete patch
may not be allowed enough time to cure adequately. Some-
times a concrete highway is repaired with asphalt—which
should be considered a temporary fix at best. Asphalt and
failing patches should be replaced with partial- or full-depth
patching sooner rather than later [DELATTE, 2008, P. 307].
M AI NT E NAN C E 139

Resealing
Though it could be called a repair, resealing of joints and cracks is so frequently called for
that we will consider it a maintenance procedure. Joints almost always fail before the con-
crete pavement does. Cracks can occur at any time, and are somewhat harder to repair
because they are crooked and because they vary in depth and width.
A number of different materials are used to seal joints and cracks. These include hot-
pour sealants such as PVC-coal tars, silicone sealants, polysulfides, and polyurethanes.
Some agencies require that joints be resealed once a certain amount of the sealant—per-
haps 30%—has failed. Others use complicated rating systems based on sealant condition,
climate, traffic levels, stone intrusion into joints, and other factors [EVANS, SMITH, AND
ROMINE, 1994, P. 19].
Sealants should keep debris and moisture out of the joint. In some cases, deterioration
of the sealant reduces its height or causes gaps. In other cases, the problem is adhesion:
The sealant itself may be fairly sound, but because it does not stick to the concrete, parti-
cles and moisture can infiltrate. In either event, debris and moisture penetrate the joint,
leading to problems.
Sometimes problems with the road surface cause the sealant to fail, exacerbating an
existing problem. Faulting, D-cracks, spalling, and other problems can contribute to seal-
ant failure. In these cases, repairs to the concrete should be undertaken at the same time
as resealing.
Resealing should take place when temperatures are moderate; otherwise, expansion
and contraction will cause the joints to be too wide or too narrow. Spring and fall are usu-
ally the best times.
Sealing and resealing are similar procedures that take place as follows:

1. EXPOSING THE AREA TO BE REPAIRED. When resealing a joint, the old sealant
must first be removed. Some compression seals can be pried out. A tractor may be
used to pull out very long sections. A tool with a long, thin blade that slips under
the sealant can also be used. With other types of sealants, a diamond-bladed saw
may be needed. If sealing a crack, it must be widened and deepened with a saw
and small blade (175–200 mm) that can make turns to follow the crack.

2. RESURFACING THE JOINT. After the sealant is removed, the joint should be
resurfaced. This exposes a fresh concrete surface to ensure that the new sealant
will adhere. In some cases, the joint needs to be widened in accordance with the
requirements of the new sealant. In most cases, a wet saw is used for this purpose.
If the joint does not need to be widened, sandblasting can be an effective resurfac-
ing procedure. This step is generally not necessary when sealing a crack.

3. CLEANING THE JOINT. After wet-sawing, there will be a slurry of moisture and
concrete dust in the joint; sandblasting leaves behind both sand and concrete dust.
These must be cleaned thoroughly. An effective method is to first flush with a
M AI NT E NAN C E 140

pressure-washer. Once the joint has dried, an air blaster is used to blow out any
dust. Vacuuming machines are also sometimes used.
Some materials call for a thin coat of specified primer to be applied at
this point.

4. APPLYING BACKER ROD. Backer rod keeps the sealant from bonding to the
bottom of the joint, which would cause undue stress. The type and size of backer
rod will vary, depending on the sealant and joint.

5. APPLYING THE SEALANT. If possible, this should be done immediately after


applying the backer rod. Temperatures should be within required limits. If rain
occurs, the joints need to be cleaned before sealing.

Cold-applied sealants include polysulfides, polyurethanes, and silicones. Typically, sealant


is pumped, using compressed air, from a storage bin through a nozzle. Bubbling needs to
be minimized, and with many cold-applied sealants, it is imperative that the joint be dry
during application, and in the period soon after application [DELATTE, 2008, PP. 313–314.]
Resealing a joint or a crack is a simple procedure that can go a long way toward pre-
venting the need for extensive repairs. When compared with the cost of other repairs, it is
indeed a cost-effective approach to prolonging the serviceable life of pavement. The longer
that a crack—or any pavement problem—goes undetected and unaddressed, the greater
and more expensive the repair is going to be. The key to catching the problem is to exam-
ine the pavement regularly, and to address problems as they occur. As the next chapter will
show, sealing cracks won’t obviate all repairs. A regular evaluation of the pavement, how-
ever, combined with routine crack sealing, will minimize the extent and cost of all required
repairs.

Endnotes
Delatte, Norbert. Concrete Pavement Design, Construction, and Performance. Abingdon, Oxon,
U.K.: Taylor & Francis, 2008.

Evans, Lynn D., K. L. Smith, and A. Russell Romine, “Materials and Procedures for Repair of Joint
Seals in Portland Cement Concrete Pavements—Manual of Practice,” FHWA-RD-99-146
(updating SHRP-H-349). U.S. Dept. of Transportation – Federal Highway Administration,
1999. <www.tfhrc.gov/pavement/ltpp/pdf/99146.pdf>

“Skid Resistance,” Pavement Evaluation, Washington State Department of Transportation (Uni-


versity of Washington), 2005. <http://training.ce.washington.edu/WSDOT/Modules/09_
pavement_evaluation/09-4_body.htm - locked_wheel_tester>

Walker, Donald. PASER Concrete Roads Manual. Transportation Information Center, University
of Wisconsin–Madison, 2002. <www.dot.state.il.us/blr/p016.pdf>
HIGHWAY
REPAIRS

THIS CHAPTER FOCUSES ON THE MAJOR TYPES OF REPAIRS to highways:


partial-depth patches, full-depth patches, load-transfer restoration (dowel retrofits),
bonded and unbonded overlays, diamond-grinding, slab stabilization (undersealing),
and repairs to drains.

Partial-Depth Repairs
In a partial-depth repair, a small, shallow section of damaged concrete is removed, and the
resulting hole is filled with a patching material. Partial-depth repairs are most commonly
used to fix spalling, scaling, and D-cracking. The repair will smooth the surface for
improved ride quality and will prevent further damage. It may add a small amount of
structural stability, as well.
Partial-depth repairs are less costly and take less time than full-depth repairs, and so
are preferable for damage that does not go deep into the concrete. If damage to the con-
crete extends to one-third (1/3) or less of the slab’s thickness, a partial-depth repair may be
the solution. However, it will repair only the deteriorated concrete, and will not reinforce
opened joints or deep cracks. If damage goes deeper, a full-depth repair is required, as dis-
cussed later in this chapter.
In some situations, a partial-depth repair will accompany another repair, such as a
dowel retrofit.
For the repair to be durable, the area of damage must be correctly identified, demol-
ished, and cleaned out thoroughly. A bonding agent is applied, and a suitable repair mate-
rial is applied and finished [fhwa, 2008, P. 5.16].
The repair material should have a proven record of durability for both the climate and
the road conditions. Important factors include freeze-thaw durability, how rapidly the
material becomes hard enough to withstand traffic, and shrinkage. There is often a trade-
off: For instance, a material with rapid strength gain will allow for quick restoration of traf-
fic, but may not be as durable, or may shrink significantly, leading to future cracks. There

141
HI GH WAY R E PA IR S 142

are three (3) basic types of materials used for partial-depth repairs: cementitious,
polymeric, and bituminous.
Cementitious materials are composed of Portland cement concrete. The concrete
should be an air-entrained, low-slump mixture with aggregates not larger than half (1/2)
the repair thickness. Type I is the most common choice of cement. A rich mixture will gain
strength fairly rapidly. High-early additives are often used. While concrete mixes that
use gypsum, magnesium phosphate, and high alumina content gain strength quickly, they
are not recommended, because they will lose strength over time. When applying a Port-
land cement patch, it’s necessary to first apply a bonding agent to ensure adhesion [fhwa,
2008, P. 5.3].
Polymeric (or polymer-based) concretes add various types of resin to a concrete mix.
The resulting material sets quickly and remains strong, but the thermal expansion proper-
ties of the repair material must be similar to the surrounding concrete; otherwise, cracks
may result. Epoxy concrete has excellent adherence, but because of the heat it generates
while setting, it often must be placed in multiple pours. Methyl Methacrylate concrete
(MMA) is easier to work with and is durable, but the fumes it generates may pose health
dangers for those who work with it. Polyester-styrene concrete is a less expensive option,
but it also has a slower strength gain [fhwa, 2008, P. 5.4].
Bituminous patching materials make for quick and inexpensive repairs, but they are
not considered a permanent solution. One study found that repairs typically lasted three
(3) to four (4) years, and then deteriorated rapidly.
If a partial-depth repair spans a joint, it will almost surely crack unless an expansion
joint insert is installed. Typically, a strip of polystyrene, polyethylene, or asphalt-impreg-
nated fiberboard is used. If the patch runs up to a joint, the insert should be placed along
the joint. If the patch traverses a joint, the insert should be placed in the middle, along the
line of the joint. The insert must extend at least 25 mm below the depth of the patch, to
ensure that patching materials on either side do not contact each other. A bonding agent
may be required.
Where the patch meets the existing concrete, the joint needs to be sealed. This is usu-
ally done by brushing or troweling on a grout mixture of 1 part Portland cement to 1 part
sand. The saw-cut run-outs should be filled as well [fhwa, 2008, P. 5.5].

Full-Depth Repairs
A roadway with deep cracks or other structural defects often needs a full-depth repair,
which essentially means demolishing a small portion of a slab and pouring a new slab.
Performed correctly, a full-depth repair results in a patch that is as strong as a new road-
way. If the slab has widespread structural problems, however, a structural overlay—or
complete demolition and repouring of the slab—may be the best solution.
Defects that require full-depth repair include exposed metal reinforcing, faulted joints,
transverse cracking, D-cracking and corner cracking, blow-ups, and pop-outs. If a partial-
HI GH WAY R E PA IR S 143

depth repair has been made to an area and has failed, a full-depth repair may be needed.
If cracks—either at a joint or mid-panel—extend down more than a third (1/3) of the slab’s
thickness, then a full-depth repair is called for. In some cases, a full-depth repair is per-
formed prior to applying an overlay.
To identify the extent of the damage, core samples need to be taken. If the slab is rein-
forced, cracks that appear as structural may not extend all the way through the slab, and
so a partial-depth repair may suffice. On a plain, or non-reinforced, slab, all cracks that
appear as structural call for a full-depth repair.
In general, the minimum recommended length of a repair is 1.8 meters. If there are no
dowels in the repaired area, a minimum length of 3 meters is recommended. The perime-
ter of the repaired area should be at least 0.6 meter from any joint. If the patch comes
within 1.8 meters of an existing joint, the patch should extend all the way up to the joint.
If two patches are 2.4 meters apart, they should be joined into one large patch. If a large
area needs a number of patches, it is often more practical to demolish the entire area and
replace the slab.
Materials options are much the same as for a partial-depth repair. Because full-depth
repairs call for large amounts of material, cement concrete is the most commonly chosen
material. Most often, Type I cement is used. An accelerator can be used to enable reopen-
ing the roadway within one to three (1–3) days. To reopen within 6 hours, Type III cement
will be needed, together with accelerators. Less suitable is high-early strength concrete
made by reducing water content, and adding a water reducer and a chemical accelerator.
While it gains strength quickly, it shrinks more.
The concrete should be air-entrained, and a curing compound may be applied once
the repair is in place.
Load transfer is critical when making a full-depth repair. If the repaired area is not
fully tied to the existing slab, differential movement will cause damage to the repair or to
the existing slab. In most cases, smooth 38-mm dowels are installed as described in “Load-
Transfer Restoration.” Agencies may have specific requirements for dowel layout. Often,
three (3) or four (4) dowels per wheel path are called for.
On a continuously reinforced slab, rebar of the same thickness and type as the existing
rebar should be used. The rebar should be spliced so that it overlaps the existing rebar and
it should be tied securely with wire. If the rebar is welded, the weld should be at least 100
mm long.
If the repaired area is rough enough to hinder rideability, it may be necessary to dia-
mond-grind it, as discussed below.

Load-Transfer Restoration (LTR)


Load-transfer restoration, also called dowel retrofit, is best done by placing dowels at
transverse joints or cracks to tie slab sections together. This transfers loads across the slab
HI GH WAY R E PA IR S 144

and keeps one section from sinking below another. It may be done at a joint that has
no joints, at a joint where existing joints are not doing their job, or at a crack [fhwa,
2008, P. 8.4].
The repair is made by cutting slots in the surface for the dowels using a diamond-
bladed slot-cutting machine. Most agencies require three to four (3–4) dowels, spaced 300
mm apart, for each wheel path. Once cut, the slots are sandblasted, then airblasted to
remove all dust and ensure a good bond with the patching material. The dowels are put in
place, and the repair material is applied [fhwa, 2008, P. 8.7].
The repair material, which fills in around the dowels, should be thermally compatible
with the concrete of the slab, should bond well to the concrete, and should exhibit little
shrinkage. Materials used for partial-depth repairs, as described above, generally fit the
bill. Portland cement concrete is a common choice. Proprietary patching materials often
have better bonding and resistance to shrinking, but cost a good deal more. If desired, a
curing compound may be applied. Because the patches are narrow, the area can usually be
opened to traffic fairly soon [fhwa, 2008, P. 8.5].
In general, if one side of a joint or crack is between 3 mm and 13 mm higher than the
other, it is a good candidate for LTR. LTR is also often performed prior to installing an
overlay, to protect the overlay from cracking and spalling.
LTR is often appropriate where an otherwise sound slab shows lack of load transfer at
specific joints. It is also a good way to reinforce a slab that is too thin or under-reinforced,
where joints are too far apart, or where sizeable cracks have developed. If the patches are
rough, it may be beneficial to diamond-grind them.

Cross-Stitching
A crack or joint that does not show much in the way of deflection can be reinforced using
cross-stitching. Rather than cutting, demolishing, and filling a large area, cross-stitching
consists of drilling small-diameter holes and filling them with tie bars and grout. Each
“stitch” is actually a pair of holes, drilled at an angle of 35° to 4° to the surface. The holes
are drilled equidistant from the crack on each side, so they intersect directly below the
crack in the middle of the slab’s thickness. Typically, the pairs are spaced 500 to 750 mm
apart along the crack or joint [fhwa, 2008, Pp. 8.17–8.18].
Cross-stitching will strengthen longitudinal cracks that are on the same plane, and
will help maintain aggregate interlock. It also works to tie longitudinal joints where one
side is slightly higher than the other. Cross-stitching can also be used to better tie a road-
way to a shoulder. However, it is not recommended for transverse cracks.
Cross-stitching is typically done using 19-mm diameter deformed tie bars, together
with epoxy patching material, to secure the tie bars. Drilling is done with a hydraulic-
powered drill, an angle guide, and a long concrete bit that’s somewhat larger in diameter
than the tie bar. Once the epoxy has set, the road can be reopened to traffic.
HI GH WAY R E PA IR S 145

Diamond-Grinding
Diamond-grinding is the process of mechanically removing a thin layer of surface con-
crete to create a slightly roughed surface. The result is a smoother, quieter ride with
improved traction. Diamond-grinding also helps solve wheel-path rutting, and smoothes
bumps made by nonstructural cracks. It also increases traction on a slab that is too smooth
due to wear and polishing.
Diamond-grinding can smooth faults of 3 to 6 mm; if faults are larger, a load-transfer
restoration (LTR) and/or an overlay may be the best solution. If the concrete has a harder-
than-usual aggregate, the grinder may not be able to cut into it. Diamond-grinding also
cannot address structural problems such as deep cracks, D-cracks, and transverse joints
with significant faulting [fhwa, 2008, P. 9.2].
The grinding machine uses a cutting head with many closely spaced cutting blades.
The head is typically about 1.27 meters wide, mounted midway between the front and rear
wheels. Sensors in the rear wheels analyze the ground surface. Newer machines can pro-
duce a number of different textures to suit different conditions.
The machine runs along the road, parallel to the centerline. It should take up only one
(1) lane of traffic, so traffic in the other lane(s) can remain open while the work is being
performed. The cutting head is fairly narrow, so multiple passes are required in order to
surface an entire lane. The goal is to grind 95% of the surface; small occasional low areas
are acceptable.

Diamond-Grooving
With diamond-grooving, a series of saw blades cuts closely spaced grooves into the slab
surface. The grooves, which may be cut longitudinally or transversely, allow water to drain
away, and so help eliminate ice buildup and hydroplaning. They also provide greater
traction.
Unlike a grinding head, a grooving head has 19-mm spaces between the diamond
blades. As a result, grooving heads are wider than grinding heads—1.8 meters or wider.
Methods are usually much the same as for diamond-grinding, though fewer passes are
needed because the head is wider. Transverse grooving makes for the best drainage, but
longitudinal grooving is much more common because it proceeds far more quickly [fhwa,
2008, P. 5.16].

Concrete Overlays
There are two (2) basic types of concrete overlay: bonded and unbonded. A bonded over-
lay, which is relatively thin, is suitable for a roadway with only surface damage and per-
haps minor structural distress. An unbonded overlay, which is thicker, can solve many
structural problems. Overlays of both types can be laid on top of a concrete, asphalt, or
composite slab [NATIONAL CONCRETE PAVEMENT TECHNOLOGY CENTER, 2008, P. 2].
HI GH WAY R E PA IR S 146

Overlays are often made with rapid-strength mixtures (often Type III) that have a
high-cement content, with the addition of accelerators or water-reducers. However, the
advantages of rapid setting must be weighed against the more important considerations of
strength and resistance to shrinkage.
The overlay should expand and contract in a way that is slightly less than that of the
existing concrete below, to minimize stress caused by dissimilar movements. (Because the
overlay will be exposed to more temperature fluctuations, it will move somewhat more
than the lower layer.)
Recommended admixtures include air-entrainment, water-reducers, and Supplemen-
tary Cementitous Materials (SCMs) such as fly ash (but see “Environmental Concerns
about Fly Ash,” in the chapter Supplementary Cementitious Materials/Bulk Supplements).
Fiber reinforcement is strongly recommended, for both bonded and unbonded overlays.
Fiber-reinforced concrete has greater flexibility and resistance to cracking, which is espe-
cially important in an overlay. Polypropylene microfibers are the most common choice,
but polyester and steel fibers are also used.

UNBONDED CONCRETE OVERLAYS OVER ASPHALT. If an asphalt or composite (asphalt-


over-concrete) surface has significant surface problems and even some structural prob-
lems, an unbonded concrete overlay (also called conventional white-topping) is often the
best fix. The overlay is basically a new concrete slab that uses the asphalt surface as a sub-
base. An unbonded overlay will rehabilitate an asphalt slab with severe potholes, rutting,
alligator cracking, shoving, and pumping.
The overlay is much the same as one designed to be laid over a concrete surface: It is
made of concrete with additives, and can be from 100 to 280 mm thick [NATIONAL CON-
CRETE PAVEMENT TECHNOLOGY CENTER, 2008 P. 20].

UNBONDED OVERLAYS OVER CONCRETE. An unbonded concrete overlay is a good way


to strengthen and repair a concrete slab that has moderate structural problems. It is basi-
cally a new slab laid on top of the existing one, with the new slab structurally separated
from the underlying slab by a layer of asphalt or other material that is sandwiched between.
If an unbonded overlay will be applied, there is usually no need to make extensive repairs
to the existing slab—which can save a good deal of time and expense.
Unbonded overlays are anywhere from 100 to 280 mm thick. They can be reinforced
in various ways, much like a new slab.
The underlying slab can have some structural problems, as long as they are not severe.
Faulting is usually not a problem, as long as a separation layer that is at least 25 mm thick
will be applied first. However, if panel tenting or blow-ups are visible, indicating a void
under the slab, it is necessary first to repair and stabilize those areas, either with full-depth
repairs or slab stabilization (undersealing).
The separation layer allows the overlay to move independently of the underlying slab,
and so prevents cracks from telegraphing up to the overlay. The separation layer is most
HI GH WAY R E PA IR S 147

often made of conventional asphalt, and is 25 mm thick. If the existing slab has large irreg-
ularities, the separation layer may be somewhat thicker. If traffic will be heavy, the overlay
will be at least 200 mm thick. Joints are doweled.
An air blower or a mechanical sweeper should be used to clean the slab of debris and
dust. The asphalt is then placed using conventional methods. The surface should be dry
before the overlay is installed.
Curing is especially important if the unbonded overlay is 150 mm or thinner. Twice
the usual amount of curing compound should be used in this case.
As soon as possible, the joints should be sawn. Transverse joints are typically spaced
more closely than on a newly poured highway. If the overlay is up to 175 mm thick, 1.8- by
1.8-meter panels are recommended. For overlays greater than 1.75 mm, 4.6-meter joint
spacing is recommended.
Edge drains are needed to provide good drainage for the separation layer. Without
proper drainage, water pressure can cause the separation layer to strip, fault, or crack.
Deep edge drains, also called subdrains, are often used. In some areas the asphalt for a
separation layer is made with less sand and more aggregate, for a more porous product,
which increases drainage. In Germany and other places, a geotextile product is used
[NATIONAL CONCRETE PAVEMENT TECHNOLOGY CENTER, 2008, PP. 18–19].

BONDED OVERLAYS. A bonded concrete overlay can increase structural strength (though
not as much as an unbonded overlay) and can smooth surface problems like scaling, pol-
ishing, cracking, and spalling. Bonded overlays are thin, and it is the bonding process that
gives them strength. Bonding turns the overlay and the existing slab into a monolithic
slab—the overlay in essence becomes part of the slab—and thus, the overlay does not need
to be strong by itself. To aid in bonding, the surface is abraded by either sandblasting or
shot-blasting, to present a rough surface that will bond better than a smooth one would.
Pressure-washing and airblasting, and perhaps both, effectively remove dust and debris
that can inhibit bonding.
For the two (2) to properly bond, the overlay must have expansion and contraction
properties that are very similar to the slab’s. A bonded overlay does not use metal rein-
forcement of any type, though it may use fiber reinforcement.
The overlay’s joints should match those of the slab below to ensure that both layers
move together. The overlay’s joints should be at least as wide as the slab’s joints. It is com-
mon practice to widen joints in the existing slab with a diamond-blade saw prior to the
overlay. In some cases, additional joints, between the overlaid joints, are cut into the over-
lay to further cut down on movement.
If the existing slab shows structural defects—such as D-cracking, deep potholes, heave
or settling, deep transverse cracks, or faulted joints—then a bonded overlay is not a good
solution. Such defects must be repaired using other methods discussed in this chapter.
After that, a bonded overlay may be applied for a smoother ride.
If cracks in the existing slab widen after the bonded overlay is installed, they will tele-
HI GH WAY R E PA IR S 148

W h it e -T o p p in g

An asphalt road that is in sound structural condi- The overlay is made of the same materials as a
tion can be improved by applying a 50- to 125-mm bonded overlay for a concrete slab: conventional
bonded concrete overlay, known as ultra-thin white- concrete with additives such as accelerators, water-
topping. The overlay can solve surface problems reducers, and high-modulus structural fibers. The
typical to asphalt slabs, including ruts, shoving, procedure for placing and finishing the concrete is
and thermal cracking. The overlay will provide a the same as for a bonded overlay over a concrete
more durable solution than standard asphalt repairs slab: Either fixed-form or slip-form construction is
would. used. Curing compound should be added at twice
the usual rate, and joints should be cut as soon as
The asphalt road must be in basically sound condi-
possible.
tion. If there is structural deterioration, or if there
are potholes or depressions due to uneven base While asphalt roads do not have joints, the overlay
support, poor drainage, or delaminating of asphalt should. Typically, joints are placed 0.9 to 2.4 meters
layers, a bonded overlay will not solve the prob- apart. Longitudinal joints should be in the center of
lems. As with a bonded overlay applied to a concrete the road, where they will not be in the wheel path.
slab, the underlying asphalt slab will carry most of
If a road is composite—with an asphalt surface
the weight, so firm bonding is essential. All needed
laminated over a concrete slab—the same meth-
repairs to the asphalt surface must be made before
ods and materials as recommended for overlaying
the overlay is applied. Potholes and medium-sized
an asphalt surface are used. The bonded overlay is
cracks must be filled, and subgrade support must
typically 50 to 125 mm thick and made of the same
be provided where needed. If patches are needed,
type of reinforced concrete as recommended for a
they should be made with concrete rather than
bonded overlay onto concrete or asphalt. Joints are
asphalt, since new asphalt is oily and does not
cut in the same configuration as for an overlay
bond well with concrete. When a concrete patch is
designed for an asphalt surface.
made, joints in the overlay should be placed directly
above the joint between the patch and the asphalt However, if problems with the asphalt surface are
surface. due to underlying concrete problems, it will be nec-
essary to make repairs—partial- and full-depth
To lower high spots and roughen the surface, the
patches, slab stabilization (undersealing), load-
asphalt should be milled. The surface does not
transfer restoration, and crack cages—as required
have to be perfect or smooth, but there should be
for a concrete slab [NATIONAL CONCRETE PAVEMENT
no spots high enough to make the overlay too thin
TECHNOLOGY CENTER, 2008, PP. 14–15]
at those points.

graph up through the overlay. If this happens, full- or partial-depth repairs, or load-trans-
fer restoration, will stabilize the cracks.
If the concrete slab has been patched with asphalt, the overlay may not bond well to
the patch. The asphalt should be cut out and replaced with a concrete patch [NATIONAL
CONCRETE PAVEMENT TECHNOLOGY CENTER, 2008, PP. 12–13].

Slab Stabilization (Undersealing) and Slab-Jacking


The subbase under a slab may erode, either due to poor materials and installation or
because of water infiltration. Once this foundational support is lost, the pavement may
settle in places, causing major faults and cracking. If loss of base support has occurred but
HI GH WAY R E PA IR S 149

damage to the pavement has not begun or is small, slab stabilization is needed. If settling
and cracking have occurred, the solution is slab-jacking, a procedure in which holes are
drilled through the slab, and pressurized grout or polyurethane material is forced in, to
raise the slab.
Because the damage occurs beneath the slab, it can be difficult to be sure whether or
not a slab needs stabilization or jacking. And if material is pumped under a slab that is
actually well supported, it can place stress on the slab and so create cracking, faulting, and
other problems.
A correct diagnosis is critical and should be made with careful deliberation before
proceeding with either slab stabilization or slab-jacking. Diagnosis is often made using a
truck with devices that can measure the difference between approach and leave-corner
deflections: If the approach deflection is significantly greater than the leave deflection,
then voids are likely to exist. The most precise and reliable information can be gained
using recently developed technologies of ground-penetrating radar and infrared thermog-
raphy. These devices can detect air-filled under-slab voids as thin as 6 mm. Water-filled
voids are more difficult to detect [fhwa, 2008, Pp. 4.2–4.4].
Slab-jacking is used only for specific depressions in the road that are noticeable to
drivers. Such depressions caused by under-slab voids are usually found at approaches to
bridges, over culverts, and over embankments.
Cement-based repair materials are the most economical and common. Pozzolanic-
and limestone-cement grout are typical choices. A ratio of 1 part cement to 3 parts pozzo-
lan (Class F fly ash), with enough water to make it fluid, is common. Additives such as
superplasticizers and water-reducers are often used [fhwa, 2008, Pp. 4.7–4.8].
Polyurethane materials are also used. In one system, two (2) different liquids are
injected. They react chemically under the slab to expand and fill voids. A high-density
polyurethane foam, known as URETEK, is another option [fhwa, 2008, P. 4.8].
It’s necessary to pump under higher pressure for slab-jacking than for slab stabiliza-
tion, but at the same time it’s important to avoid raising the slab too high. Each hole
should be raised no more than 6 mm to begin with. Once all holes have been raised a small
amount, the first hole can be revisited for further raising [fhwa, 2008, pP. 4.11–4.12].

Retrofitting Edge Drains


A well-drained highway can last twice as long as a poorly drained one. If an existing high-
way has symptoms of drainage problems, edge drains can be retrofitted. Of course, new
drainage cannot be installed under an existing highway slab. Drainage can be improved,
however, by shortening the distance that draining water needs to flow horizontally [fhwa,
2008, P. 7.1].
An older highway that lacks dowels and other reinforcements needs good drainage, or
else it will develop cracks, faulting, and other problems. If a highway is doweled and well
reinforced, moisture has comparatively little effect, and it is less likely to need retrofitted
HI GH WAY R E PA IR S 150

drains. But even a new highway can need added drainage if it is subjected to extreme
moisture conditions.
The most likely candidate for retrofitted edge drains is a road that is less than 10 years
old that shows only a small amount of cracking [fhwa, 2008, P. 7.2].
Moisture-related damage to a highway does not necessarily mean that retrofitted
drains will solve the problem. In fact, if a slab already has significant damage, an edge
drain probably will not help. Drainage will also not help sections of a slab with significant
cracking. If over 10% of the surface has cracks, and many transverse joints are spalled, a
retrofitted drain will not help. If there are voids under the pavement, then slab stabiliza-
tion—not drainage—is the most likely the better solution [fhwa, 2008, P. 7.2].

Endnotes
Concrete Pavement Preservation Workshop: Reference Manual, U.S. Dept. of Transportation –
Federal Highway Administration, 2008. <www.cptechcenter.org/publications/preservation_
reference_manual.pdf>

Guide to Concrete Overlays, 2nd edition. National Concrete Pavement Technology Center, Iowa
State University, 2008. <www.cptechcenter.org/publications/overlays/index.htm>
LIFE-CYCLE
COST ANALYSIS

RESPONSIBLE USE OF RESOURCES ARGUES that in the design of any highway, an


agency should search for the most cost-effective approach. In the simplest case, when a
funding agency is deciding between a low-cost road and a more expensive road, the deci-
sion is easy. More typically, however, an agency is considering options with different price
tags, different maintenance costs, and perhaps even different life spans. Life-Cycle Cost
Analysis (LCCA) is a technique that reduces the costs of different approaches to a com-
mon denominator so that they can be compared. It allows funding agencies to quantify the
real long-term cost of a highway investment, as opposed to the short-term or immediate
cost. It helps avoid technical solutions that may seem inexpensive initially but which, for
various reasons, can end up costing far more over time.
While a wide variety of factors are involved, the concept behind LCCA is simple: The
costs of any given alternative are computed for each year of the life of the highway. Each
cost is then deflated to its present value, and the present values are totaled. When the pres-
ent value is computed for each alternative, their true relative costs can be compared.
In practice, it is a multi-part process:

1 . Establish design alternatives.


2 . Establish a schedule of activities.
3 . Establish agency and user costs.
4 . Compute life-cycle costs.
5 . Perform risk analysis.
6 . Analyze the results.

The risk analysis—which looks at a variety of outcomes and determines which are
most likely—is sometimes omitted. If so, the analysis is called a deterministic analysis. If
the risk analysis is performed, the result is called a probabilistic analysis. Probabilistic
analyses were once too complicated to perform, but computers and readily available soft-
ware have made the process much more common.

151
L I F E - CY CL E COS T ANALYS IS 152

Several software programs are available to help conduct a life-cycle cost analysis. The
most prominent in the United States is the RealCost program, available from the Federal
Highway Administration (FHWA).

Establish Design Al ternatives, Schedules, and A gency Costs


At the outset, the agency obviously must decide what alternatives it wants to consider. If the
LCCA is to be used, the benefits of each alternative must be equal, although the approaches
can be radically different. The alternatives Figure 17-1

might be asphalt vs. concrete, recycled vs. vir-


Ty p ic a l Pe n n DOT Ma in t e n a n c e Sc h e d u l e
gin base material, different mixes of concrete,
fo r Ne w Ro a d s
or even different routes. Whatever the alter-
Year Treatment
natives are, a pavement design strategy is
5 Clean and seal 25% of longitudinal joints.
developed for each. The strategy is a list of Clean and seal 5% of transverse joints.
activities required for completion and main- 0% for neoprene seals.
Seal coat shoulders if Type 1 paved shoulders.
tenance of the road. These include:
10 Same as year 5.
N Initial preliminary engineering 15 Clean and seal 25% of longitudinal joints.
N Contract administration Clean and seal 10% of transverse joints.
5% for neoprene seals.
N Construction supervision and Seal coat shoulders if Type 1 paved shoulders.
construction 20 Concrete patch 5% of pavement area.
N Future routine and preventive Spall repair 1% of transverse joints (5 sf/joint).
Slab stabilization: minimum 25% of transverse joint.
maintenance, resurfacing, and Diamond-grind 100% of pavement area.
rehabilitation Clean and seal all longitudinal joints,
including shoulders.
N All associated administrative tasks Clean and seal all transverse joints,
7% for neoprene seals.
[WALLS AND SMITH, 1998, P. 12] Seal coat shoulders, if Type I paved shoulders.
Maintenance and protection of traffic.
User delay.
Once the activities are known, a schedule is
25 Clean and seal 25% of longitudinal joints.
developed, and costs forecast on a year-by-
Clean and seal 10% of transverse joints,
year basis for the period being analyzed. The 10% for neoprene seals.
Seal coat shoulders, if Type I paved shoulders.
analysis period needs to be long enough to
30 Concrete patch 2% of pavement area.
include long-term costs of each strategy—
Clean and seal all joints with fiber asphalt membrane.
including at least one (1) scheduled rehabili- 60-#/sy leveling course.
3.5-in ID-2 or 4-in ID-3/ID-2 overlay.
tation. The FHWA’s recommends an analysis Saw and seal joints.
period of at least 35 years, a period long Type 7 paved shoulders.
Adjust all guide rail and drainage structures.
enough to include reconstructions. A typical Maintenance and protection of traffic.
schedule of activities, developed by the User delay.

Pennsylvania Department of Transportation 35 Seal coat shoulders.


(PennDOT), is shown in Figure 17-1. S ource: Walls and Smith, 1998, p. 11
Note: The CPR strategy slated for year 20 can be moved to year 15 at the
District’s discretion. However, when doing this, the overlay at year 30 must be
moved to year 25, and another overlay added at year 33.
L I F E - CY CL E COS T ANALYS IS 153

Estimate the User Costs


In addition to the costs borne by the funding agency, users of a highway under construc-
tion also pay certain costs. While the costs are not out-of-pocket, they are nonetheless real,
and need to be part of the consideration: A road that shuts down traffic for 12 months not
only inconveniences users, it is also twice as inconvenient as a project that shuts down the
road for 6 months. In an effort to determine the expense of the inconvenience, LCAA
focuses on three (3) areas: the dollar value of delays caused by construction, vehicle oper-
ating costs, and crash costs:

N The dollar value of delays is the time lost in transit multiplied by the prevailing
wage. Delays include time lost to traffic backups due to construction, time lost to
lower speeds through the work zone, the time it takes for a driver to return to
normal speed, and time lost by those taking alternate routes.
N Vehicle operating costs include extra fuel consumption, as well as wear and tear
caused by substandard road surfaces.
N Crash costs are the costs of accidents that are caused by the construction, above
and beyond what the level of accidents normally expected along the road.
[WALLS AND SMITH, 1998, P. 33]

The FHWA’s RealCost analysis generates these costs in part on information taken from
traffic studies, and in part on default assumptions. Among the things determined by the
traffic study:

ANNUAL AVERAGE DAILY TRAFFIC (AADT): This is the annual average daily traffic
in both directions as measured before the construction begins.

ANNUAL GROWTH RATE OF TRAFFIC: The percentage by which AADT increases


each year.

MAXIMUM AADT IN BOTH DIRECTIONS: The maximum amount of traffic the


new road can handle.

COMBO TRUCKS: The percentage of the AADT that is trailers and tractor-trailers.

SINGLE-UNIT TRUCKS: The percentage of the AADT comprised of trucks with


more than six (6) tires that are not combo trucks.

NORMAL OPERATIONS SPEED LIMIT: Speed limit in normal operating conditions

NORMAL NUMBER OF LANES OPEN: Number of roadway lanes available under


normal operating conditions.

QUEUE DISSIPATION CAPACITY: Capacity of each lane once traffic has cleared
the work zone.

HOURLY TRAFFIC DISTRIBUTION: Traffic flow by hour along the highway.


[FWHA, 2004, P. 28]
L I F E - CY CL E COS T ANALYS IS 154

Using these values and similar values, as well as certain default assumptions, RealCost cal-
culates the cost to the user. The costs should be calculated for the period of construction
as well as during any rehabilitation work that will have a major effect on the road. Routine
work, such as repairing potholes, is difficult to forecast and is usually excluded from the
analysis, with no apparent effect on the outcome: Where it has been possible to accurately
forecast routine costs, they have not been large enough to affect the outcome of the analy-
sis [WALLS AND SMITH, 1998, P. xiii].

Determine Life-Cycle Cost


Once all tasks have been assigned a cost and once a schedule of costs has been developed,
life-cycle cost can be computed. Life-cycle cost is defined as the initial cost of the project
plus the net present value of all subsequent costs. It isn’t a matter of simply adding up all
the costs, however. Because a dollar can be invested, a dollar today is worth more tomor-
row. As a result, all expenditures in the future must be reduced to today’s terms. For exam-
ple, if the interest rate is 4%, $1 million invested today is worth $1.04 million a year from
now, and that $1.04 million is worth $1.082 million the following year. In economic terms,
the future value of $1 million two (2) years from now is $1.082 million if invested at 4%.
Likewise the present value of $1.082 million held two (2) years from now is $1 million.
LCCA requires that all expenditures be reduced to their present value.
The standard formula for calculating present value is:

; =
where:
1
i = the interest rate
(1+i ) n
n = the year of the expenditure (year 1 of the project, year 2, year 3, etc.).

When bringing the future costs of a project back to their present value, expenses for each
year are totaled individually. One by one, each is brought back to its present value, and the
present values are totaled. To complete the picture, the initial investment is added to the
equation. As a formula, this looks like:

; =
N

3
1
PV = Initial Cost + Rehab Costk
(1+i ) nk
k=1

where:
PV = Present Value
i = the interest rate
N = the length of the project in years equals the sum of the individual present values.

; =
N

3
1
Rehab Costk equals the sum of the individual present values.
(1+i ) nk
k=1
L I F E - CY CL E COS T ANALYS IS 155

As noted, standard practice in the United States is to set N at 35 years or more. While the
value of interest rates, i, is likely to vary over time, the long-term value of N irons out many
of the fluctuations. In the United States, the rate is based on the interest of the 10-year fed-
eral Treasury Bond, which historically has run between 3% and 5%. The rate used in anal-
ysis is typically 4% [WALLS AND SMITH, 1998, P. 6].
If the analysis were to stop here, it would look like the analysis in shown in “Determin-
istic Life-Cycle Cost Analysis,” which walks through the analysis step-by-step. It’s interest-
ing to note that the user costs are higher than the agency costs in both of the alternatives
presented, and that they are the determining factor in making alternative A cheaper.
An analysis that stops at this point is called a deterministic analysis. While it has the
advantage of being relatively simple, there are inherent risks. As shown in Figure 17-2, a
good deal of what underlies the analysis is either an estimate, an assumption, or a pro-
jected look into the future. An error in any of these can throw off the value of the analysis.
Because of the obvious risks, analysts sometimes run two (2) scenarios—a best-case sce-
nario, and a worst-case scenario. In the best-case scenario, the assumptions used are those
that are both most favorable to the Figure 17-2
project and likely to happen. (While the
most favorable results would be for the Assu mp t io n s Ma d e in LC CA
road to build itself instantly at no cost,
LCCA Component Input Variable Source
it’s not very likely, and this and similar
Initial and Future Preliminary Engineering Estimate
assumptions are outside the best-case Agency Costs
Construction Management Estimate
scenario.) In a worst-case scenario, it is
Construction Estimate
assumed that anything that is likely to
Maintenance Assumption
go wrong will go wrong. Comparing
best- and worst-case scenarios gives Timing of Costs Pavement Performance Projection

planners a better idea of the range of User Costs Current Traffic Estimate

possibilities, but there are still prob- Future Traffic Projection


lems. Not only are the scenarios still Hourly Demand Estimate
based on assumptions, there is no indi- Vehicle Distributions Estimate
cation as to which scenario is more Dollar Value of Delay Time Assumption
likely—or if either is likely at all.
Work Zone Configuration Assumption

Work Zone Hours of Operation Assumption

Work Zone Duration Assumption

Work Zone Activity Years Projection

Crash Rates Estimate

Crash Cost Rates Assumption

NPV Discount Rate Assumption


S ource: Walls and Smith, 1998, p. 82
L I F E - CY CL E COS T ANALYS IS 156

De t e r min ist ic Life -C y c le Co st An a ly si

A typical deterministic Life-Cycle Cost Analysis (LCCA) examines two (2) or more alter-
natives, each of which provides the same benefits. This hypothetical example, prepared
by the Federal Highway Administration (FHWA), assumes an interest rate of 4%, and
an analysis period of 35 years. Costs that are equal between alternatives have been
removed from the analysis.

STEP 1: ESTABLISH DESIGN ALTERNATIVES


Alternative A is characterized by fewer construction and rehabilitation activities than
Alternative B is, but the activities that Alternative A requires are more extensive and the
activities are more expensive than those of Alternative B. Alternative B requires more
frequent use of work zones to maintain level of service, but these work zones don’t last
as long as those of Alternative A.

STEP 2: ESTABLISH A SCHEDULE OF ACTIVITIES


Alternative A requires a major rehabilitation every 20 years. Alternative B requires
rehabilitation every 8 years.

Year Alternative A Activities Alternative B Activities

0 Initial construction Initial construction

12 Rehabilitation 1 (8-year service life)

20 Rehavilitation 1 (20-year service life) Rehabilitation 2 (8-year service life)

28 Rehabilitation 3 (8-year service life)

35 End of analysis period—Value of salvage or End of analysis period—Value of salvage or additional


additional life of road, if any, entered into equation, life of road, if any, entered into equation, partially
partially offsetting costs. offsetting costs.

STEP 3: ESTIMATE THE AGENCY AND USER COSTS


Agency and user costs for each activity are calculated in base-year dollars, rather than
inflated dollars. User costs are based upon user vehicle operating costs and traveler
delay associated with work-zone activities. User costs increase for similar work due to
the increase in traffic over time. At year 35, the value of the remaining life of the road
is entered for each alternative.

Alternative A Activities Alternative B Activities

Year Constant Dollar Agency Cost Constant Dollar User Cost Constant Dollar Agency Cost Constant Dollar User Cost

0 $26,000,000 $11,000,000 $20,000,000 $8,000,000

12 6,000,000 10,000,000

20 $15,000,000 $30,000,000 6,000,000 16,000,000

28 6,000,000 28,000,000

35 (3,750,000) (7,500,000) (750,000) (3,500,000)

(continued)
L I F E - CY CL E COS T ANALYS IS 157

De t e r min ist ic Life -C y c le Co st An a ly si — c o n t in u e d

STEP 4: COMPUTE LIFE-CYCLE COSTS


Using the interest rate, the present value (PV) is calculated for each of the agency and
user costs.

Alternative A Alternative B

Discount Discounted Discounted Discounted Discounted


Year Factor Agency Costs User Costs Agency Costs User Costs

0 1.0000 $26,000,000 $11,000,000 $20,000,000 $8,000,000

12 0.6246 3,747,582 6,245,970

20 0.4564 6,845,804 13,691,608 2,738,322 7,302,191

28 0.3335 2,000,865 289,337,369

35 0.2534 (950,308) (1,900,616) (190,062) (886,954)

Total Cost
31,895,496 22,790,992 28,296,707 29,998,576
(Present Value)

STEP 5: ANALYZE THE RESULTS


Alternative A has the lowest combined agency and user costs, whereas Alternative B
has the lowest initial construction and total agency costs. Based on this information
alone, the decision-maker could lean toward either Alternative A (based on overall
cost) or Alternative B (due to its lower initial and total agency costs). However, more
analysis might prove beneficial. For instance, Alternative B might be revised to see
whether user costs could be reduced through improved traffic management during
construction and rehabilitation. Analysis could be performed to see whether the results
are particularly sensitive to key assumptions, and if they are, ways of reducing the
costs of those assumptions could be explored. Finally, probabilistic analysis could help
to capture the effects of uncertainty in estimates of timing or magnitude of costs devel-
oped for either alternative.
[FHWA, 2002, p. 19]
L I F E - CY CL E COS T ANALYS IS 158

Risk A nalysis
Risk analysis attempts to quantify the uncertainty. Instead of examining best- and worst-
case scenarios, it generates a huge number of scenarios based on randomly generated
numbers. It looks not only at the best- and worst-case scenarios, but at a wide variety of
possibilities, including those that are better than best-case, and worse than worst-case. A
good analysis contains as many as 10,000 scenarios [WALLS AND SMITH, 1998, P. 95].
Figure 17-3

Co mp a r iso n o f Alt e r n a t iv e s
0.6

Alternative A
0.5
Alternative B
Relative Probability

0.4

0.3

0.2

0.1

0
18 23 28 33 38

Net Present Value ($ Millions)


S ource: Walls and Smith, 1998, p. 96

Figure 17-4

Pr o b a b ilt y
1

0.8
Relative Cumulative Probability

60%
0.6

Alternative A
Alternative B
0.4

$28.27

0.2

0
18 23 28 33 38

Net Present Value ($ Millions)


S ource: Walls and Smith, 1998, p. 97
L I F E - CY CL E COS T ANALYS IS 159

Drawn as a graph, the two (2) comparisons might look as they do in Figure 17-3,
where A and B are alternate road designs. The highest point on either curve is the mean of
all the possibilities on the curve, and is also the most likely to occur. In this case, it’s intui-
tive that Alterative A is likely to cost more than Alternative B. In other cases, it might not
be, and the curve warrants a closer look.
Looking at Figure 17-4, we see that there is a 60% probability that project costs for
Alternative B will be less than $28.27 million. In other words, in the 10,000 random sce-
narios processed, 60% of the calculated values were less than $28.27 million. The slope of
Figure 17-5
the line is important, too. The flatter the
slope of the area immediately surround-
Re su lt s o f Simu la t io n s
ing a point, the greater the uncertainty
of the value of that point. In Figure 17-4, Net Present Value ($ Millions)

the range of costs with a relative proba- Basic Statistic Alternative A Alternative B B–A

bility between 40% and 60% is greater in Minimum $ 25.40 $ 13.13 $ (12.27)

the flatter Alternative A than it is for Maximum 33.04 40.35 7.31


Alternative B. In other words, while it is Mean 28.93 27.60 (1.33)
relatively certain that A will cost less Std. Deviation 1.04 3.13 2.09
than B, it is more difficult to predict a
Percentile
narrow range of costs for Alternative A
5% 27.33 22.66 (4.67)
[WALLS AND SMITH, 1998, P. 97].
10% 27.65 23.71 (3.94)
In Figure 17-5, the numbers gener-
15% 27.88 24.39 (3.49)
ated by the simulations help further clar-
20% 28.05 24.98 (3.07)
ify the costs and risks:
25% 28.22 25.50 (2.72)
N Alternative A will cost some-
30% 28.36 25.93 (2.43)
where between $25.40 million
35% 28.49 26.34 (2.15)
and $33.04 million.
40% 28.62 26.73 (1.89)
N Alternative B will cost between
45% 28.75 27.12 (1.62)
$13.13 million and $40.35
million. 50% 28.87 27.48 (1.40)

N The mean, and most likely cost 55% 29.00 27.87 (1.13)

of A, is $28.93 million; B is less 60% 29.13 28.27 (0.86)

expensive at a mean cost of 65% 29.27 28.71 (0.56)


$27.60 million. 70% 29.42 29.13 (0.29)

75% 29.58 29.63 0.06


The lower section of the chart tells
80% 29.76 30.20 0.44
us more:
85% 30.01 30.88 0.87

N There is a 5% chance that the 90% 30.31 31.67 1.35


cost of the project will be $27.33 95% 30.75 32.98 2.23
million for alternative A, and S ource: Walls and Smith, 1998, p. 98

$13.13 million for Alternative B.


L I F E - CY CL E COS T ANALYS IS 160

N There is a 10% chance that the cost of Alternative A will be $27.65 million and
that of B $22.66 million.
N In fact, the cost of Alternative B continues to be lower in up to 70% of the scenar-
ios. Eventually, as the view expands to include more and more extreme cases,
there is a chance that Alternative A is cheaper—but only in the few cases where
the values of A are less than those than B in the Figure 17-3. The most cost-
effective alternative is likely to be B in all but a few cases, and only an agency
willing to make extreme gambles would be likely to choose A in the hope of
saving money.

Endnotes
L ife-Cycle Cost A nalysis Primer. U.S. Dept. of Transportation – Federal Highway Administration,
2002. <http://isddc.dot.gov/OLPFiles/FHWA/010621.pdf>

L ife-Cycle Cost A nalysis RealCost U ser Manual. U.S. Dept. of Transportation – Federal Highway
Administration, 2004. < www.fhwa.dot.gov/infrastructure/asstmgmt/rc210704.pdf>

Walls, James, III, and Michael R. Smith, L ife-Cycle Cost A nalysis in Pavement D esign— I nterim
T echnical Bulletin. Publication FHWA-SA-98-079. U.S. Dept. of Transportation – Federal
Highway Administration, 1998. <http://isddc.dot.gov/OLPFiles/FHWA/013017.pdf>
ALTERNATIVE
CONCRETE PAVEMENT
ROAD CONSTRUCTION

THIS CHAPTER WILL FOCUS ON THE USE OF ALTERNATIVE concrete paving


technologies, including soil-cement, two-lift (2-lift) concrete paving, fast-track cement,
roller-compacted concrete (RCC), and precast concrete slabs. In many cases, these alter-
natives offer significant economic and environmental benefits plus faster curing and set-
ting times to create long-lasting roads.
Soil-cement pavements incorporate Portland cement, water, and local soil or aggre-
gate that can be mixed either at the construction site or at a central mixing plant. The soil
material can be any combination of sand, silt, clay, gravel, and crushed stone. Suitable
locally available aggregates include slag, caliche, limerock, and scoria, and/or waste matter
such as cinders, fly ash, foundry sands, and screenings from nearby quarries and gravel
pits. Even worn, unusable gravel-based roads, either with or without bituminous surfaces,
can be recycled at the construction site to create soil cement pavements [PORTLAND
CEMENT ASSOCIATION, 2009].
A standard Portland cement mix has enough paste (cement-and-water mixture) to
coat the surface area of all aggregates and fill the voids between aggregates. Soil-cement
composition differs in that the paste is insufficient to fill the voids and coat all of the parti-
cles, resulting in a cement matrix that binds together nodules of uncemented material
[PORTLAND CEMENT ASSOCIATION, 2009].
The resulting soil-cement mixture is then spread onto the road surface, compacted or
rolled with hydraulic rollers, and cured. Because less water is added to the mixture than is
normally added to standard Portland cement, the soil-cement can be compacted to a high-
density surface [KOSMATKA, KERKHOFF, AND PANARESE, 2008, P. 327]
First used in 1935, soil-cement brought a new method of paving to the industry that
provided an economical road base, and stabilized locally available soils. This stabilization
reduced the phenomenon called “pumping,” wherein wet weather combined with heavy-
load trucks caused subbase soils to eject from under the slab. This soil ejection would then
cause the slab to become unstable and crack. Utilizing local soils that were already on site
considerably reduced the transportation costs associated with hauling gravel or other base

161
A LTE R N AT I V E CONC RE T E PAV E M E NT ROA D CONS T RUC T ION 162

materials. More than 70 years later, this technology has offered long life and low
maintenance costs to more than 160,000 km of highway in the U.S. [PORTLAND CEMENT
ASSOCIATION, 2008].
Soil-cement can be classified into three (3) types: cement-modified soils, cement-
treated base, and full-depth reclamation. The types differ depending on the materials and
methods used. Cement-modified soils combine problem soils or substandard materials
with a small proportion of Portland cement that is compacted with rollers to create a sub-
base for the road. This method creates cost savings in that only a small proportion of Port-
land cement is needed to resolve the soil issues. Additionally, by using what is readily
available at the construction site, it eliminates the need to purchase and haul aggregates
and soils from other sources [PORTLAND CEMENT ASSOCIATION, 2009].
Cement-treated base combines aggregate and/or granular soils with specific amounts
of Portland cement and water. When compacted and cured, this forms a durable paving
material that is topped with a bituminous or Portland cement concrete wearing course
[PORTLAND CEMENT ASSOCIATION, 2009].
Another suggested method of constructing soil-cement pavement is by incorporating
hydraulic binders to create a pavement with a high elasticity modulus. Some experts sug-
gest that creating transverse joints in the fresh material at 2- to 3-meter intervals can be
beneficial to the finished pavement by reducing premature roadway wear and cracking.
In addition to the advantages of high strength, durability, low initial costs, and envi-
ronmental friendliness, about 90% of the materials needed for soil-cement’s formulation
and construction are readily available at the road construction site, thus reducing
handling and hauling costs. Also, because soil-cement is compacted to a dense matrix
during construction, the resulting pavement does not buckle under traffic nor develop
potholes as with unbound aggregate base materials. Compaction of soil-cement also
reinforces inadequate subgrade areas. Additionally, the resulting matrix resists wear and
tear caused by seasonal moisture changes and freeze/thaw cycles [PORTLAND CEMENT
ASSOCIATION, 2010].
Soil-cement’s savings are also realized in equipment costs. Hydraulically bound
layers can be built using standard equipment generally used for other materials (e.g.,
unbound granular layers, bituminous mixes) thereby eliminating the need for specialized
paving equipment. If used as a base for asphalt, the stability of soil cement allows for a
thinner asphalt layer, decreasing the amount of material needed to complete the
roadway.
Soil-cement is a viable alternative to standard road construction in areas with almost
any soil conditions. “The growing use of hydraulically bound bases [soil cement] and sub-
bases and the focus of sustainability issues bring increasing recognition of the benefits of
these techniques: long-life pavements, speed of project completion, reduced use of
imported aggregates, less local construction traffic, reduced project costs.” [EUROPEAN
CONCRETE PAVING ASSOCIATION, 2009]
A LTE R N AT I V E CONC RE T E PAV E M E NT ROA D CONS T RUC T ION 163

Two-Lift Concrete Paving


Another alternative method of paving design is two-lift (2-lift) concrete paving. This con-
struction method involves placing two layers of concrete on top of each other—as opposed
to a single homogenous layer that is used in standard concrete construction. In two-lift
paving, the bottom layer is thicker yet less durable than the top layer. The major benefit of
this method is that the bottom layer can incorporate recycled aggregate, which can elimi-
nate the need to bring in aggregates from elsewhere and the subsequent hauling costs. The
thinner top layer consists of high-quality aggregates that resist freeze-thaw damage,
dampen traffic noise, and improve friction [CABLE, 2004, AND VANIKAR, 2009].
In Europe, the most common two-lift pavement design incorporates 5 to 13.9 cm of
low-cost, local aggregate base materials and the same range of high-quality, imported
aggregates in the surface layer. “Two mixing plants and two pavers are used, with the two
pavers positioned 4 to 5 meters (13 to 16 feet) apart. The top lift is placed before the bot-
tom lift starts to dry, with up to 1/2 hour between the placement of the layers.” Best prac-
tices include making the top layer as thin as possible, using shallow milling to ensure a
good bond with the subbase, and immediate curing to prevent cracks [TAKE A LOOK AT
TWO-LIFT CONCRETE PAVING, AUGUST 2007].
France, Austria, and Germany make the most abundant use of two-lift concrete pav-
ing. Each of these countries cited safety issues, noise reduction, and cost savings as main
reasons for using this method [CABLE, 2004, PP3–4]
France implemented two-lift concrete paving construction on its Highway A71. The
bottom layer incorporated local aggregates. The 50 mm top layer was composed of harder
aggregates that reduced noise and increased friction, but had to be hauled from long dis-
tances at increased cost. However, this method ultimately reduced the final road costs
because it required less of the expensive concrete to achieve the desired surface.
In Austria, government regulations require that most site materials be recycled into
the new paving project. When Freeway A1 connecting Vienna and Salzburg was rebuilt,
the existing road was broken up and hauled to a crushing plant. Depending on the amount
of asphalt it contained and the size of the crushed pieces, the recycled material was used
for both subbase and aggregate. For the subbase, aggregate was secured with cement and
topped with a 50 mm layer of asphalt. Recycled aggregate was then mixed with the bottom
concrete layer to a depth of 21.5 cm, which saved on materials costs. The 40 mm top layer
incorporated higher-quality, more durable aggregate to reduce noise, and to create a harder
top layer with increased friction.
Germany also uses two-lift paving to dampen traffic noise, increase friction, and lower
material costs. Additionally, climate conditions require the use of a harder, higher-quality
aggregate to lessen damage from extreme freeze/thaw cycles. The Munich airport has a 24
cm bottom layer that used local gravel as aggregate. The 14 cm top layer incorporated
crushed granite as an aggregate.
A LTE R N AT I V E CONC RE T E PAV E M E NT ROA D CONS T RUC T ION 164

Wirtgen GmbH (www.wirtgen.de) and GOMACO Corp. (www.gomaco.com) have


developed two-lift (2-lift) paving equipment that uses a single slip-form paver. The
GOMACO paver places concrete for the lower lift in front of the paver; concrete for the
upper lift is transferred through a hopper located on the paver. The Wirtgen equipment
features two (2) separate pavers (one [1] for each lift) that can work independently of each
other and can pave a width between 5 and 15 meters up to 43 cm deep. Additionally, the
top paver of the Wirtgen equipment can incorporate a super smoother or oscillating beam
to produce the ideal surface texture for a specific application.
Because the top layer is stiffer than the bottom layer in two-lift construction, stress
from the “bimetallic” effect can result when the two layers expand and contract at different
rates. The Federal Highway Research Institute of Germany is researching this phenome-
non as well as the layers’ property differences in order to provide recommendations on the
most advantageous concrete composition, thickness, and curing process.
Several U.S. states have also experimented with the use of two-lift construction, includ-
ing Washington, Iowa, Florida, North Dakota, and Michigan. Due to the climate varia-
tions, each state adopted different methods of implementing the technology. A study by
the Center for Portland Cement Concrete Pavement Technology found that results varied,
and that “no state agency or concrete promoter has pursued the idea of regularly building
two-lift concrete pavements as an alternative to the standard design model.” [CABLE AND
FRENTRESS, 2004, P. 14]
In Washington state, one-lift (1-lift) construction methods were implemented that
created experimental road sections with flexural strengths of 38.7 kg/cm2, 45.7 kg/cm2,
and 63.3 kg/cm2.
Because the use of studded tires is prevalent due to ice and snow conditions, the 38.7-
and 45.7-kg/cm2 sections have shown significant wear, while the 63.3-kg/cm2 section has
shown relatively minimal wear. Based on this information, some experts theorize that it is
possible for Washington to implement two-lift paving by altering the flexural strength of
the top lift. This could potentially reduce overall costs by requiring the use of only one (1)
plant and would eliminate the need to stockpile different aggregates.
Iowa experimented with two-lift pavement in which the old pavement was removed,
crushed, and used as aggregate for the bottom lift. The top lift, placed immediately after
the lower lift on the same day, incorporated virgin aggregate. In 1978, Florida also experi-
mented with two-lift pavements that incorporated a lower lift with lower flexural strength
and a top lift with higher flexural strength. At this time, both states report good road
performance.
In 1976, North Dakota used two-lift pavement construction to build a stretch of road
on U.S. Highway 2. Inferior soil conditions necessitated blending locally available aggre-
gate with a one-aggregate (1-aggregate) mixture that was then mixed with cement to form
a 15-cm layer. The right-hand lane cracked longitudinally right after construction. The
project was covered in asphalt in 1997 [CABLE AND FRENTRESS, 2004, P. 9].
A LTE R N AT I V E CONC RE T E PAV E M E NT ROA D CONS T RUC T ION 165

In 1994, Michigan implemented a European two-lift (2-lift) system that incorporated


a smaller yet high-quality aggregate on the top layer. The resulting pavement created an
exposed-aggregate surface designed to reduce noise. Michigan found that the cost of the
project was twice the cost of a conventional U.S. concrete pavement, and a 1994 study of
the surface found no perceptible noise difference between the two-lift section and adja-
cent traditional concrete pavement [CABLE AND FRENTRESS, 2004, P. 9].
Nonetheless, two-lift paving has both economic and environmental advantages.
Particularly in areas where quality aggregate is becoming scarce, two-lift paving can
utilize accumulated recycled asphalt stores, thereby saving money and helping the envi-
ronment. Successful recycling can be achieved by separating the asphalt and concrete
components from existing concrete pavements and then using 15% to 40% asphalt in the
lower lift of the new concrete pavement. In addition, because reduced amounts of quality
material are needed for the upper lift, two-lift paving can actually improve wear resistance
by implementing a lesser amount of even higher quality aggregate [CABLE AND FRENT-
RESS, 2004, P. 12].
Unfortunately, as Michigan found, construction costs are doubled for two-lift pave-
ments versus standard one-lift pavement design. In addition to needing two (2) batch
plants and two (2) slip-form paving machines, the resulting labor, permits, land, equip-
ment setup, repair, and fuel costs also increase the final cost. Also, because the top lift must
be covered or capped within 30 minutes, an equipment breakdown could significantly
increase construction costs because the bottom lift would have to be removed and rein-
stalled in order to achieve the desired pavement.

Fast-Track Paving
Another alternative to standard concrete paving technology is fast-track paving. This
method incorporates high-early-strength concrete to drastically cut curing times from 5
to 14 days down to as little as 12 hours or less—allowing roads to be opened to traffic more
quickly. This concrete can be formulated either from Type I cement, with a high cement-
to-water ratio, or by using high-early-strength (Type III) cement. In fact, because there is
no specific proportioning needed for fast-track mix, there are many different material
combinations available that can achieve rapid strength. Certain proprietary blended
cements and other admixtures can also produce rapid-strength pavements.
Besides the obvious advantages of fast-track cement’s quicker curing times, this
cement mixture creates very strong, long-lasting pavements due to the fact that they are
air-entrained and have a relatively low water content. These factors not only improve the
pavement’s strength, they also reduce the chloride or salt permeability that can damage
steel reinforcement and lead to road deterioration.
There are several advantages to fast-track concrete construction that can be enhanced
with the use of additional equipment. Minimum-clearance slip-form paving machines
A LTE R N AT I V E CONC RE T E PAV E M E NT ROA D CONS T RUC T ION 166

enable single-lane reconstruction next to traffic, reducing the number of lanes closed to
traffic. Additionally, paving machines equipped with dowel inserters eliminate the need
for dowel baskets, which can free construction lanes for other construction vehicles or
leave more lanes open to car traffic. Coring equipment advancements eliminate the need
to place utility box-outs around existing or planned manholes prior to paving, which can
speed the construction process.
However, curing of fast-track concrete can be somewhat complex. In order to achieve
high-strength pavement, moisture and heat retention are critical during the strength-gain
period—which can be difficult to maintain in colder environments. To alleviate this prob-
lem, insulating blankets, which are designed to aid early strength gain in colder climates,
can be used to reduce heat loss.
Additional curing compound may also be required in order to gain the high strength
needed. Because fast-track concrete rapidly consumes mix water during the early hydra-
tion period, it may increase the likelihood of plastic shrinkage at the pavement surface.

Roller-Compacted Concrete
Roller-compacted concrete (RCC) is a relatively new paving technology that enables con-
struction crews to quickly lay industrial or heavy-duty pavement. This method utilizes
high-density asphalt equipment to lay a zero-slump concrete that is subsequently com-
pacted with steel wheel rollers. Then, in some cases it is topped with asphalt, in some cases
it is topped with concrete, and in other low-speed applications, such as parking and ports,
it is left uncapped. The resulting pavement does not contain joints, dowels, or reinforce-
ment and it does not require forms or finishing—yet it withstands heavy-duty traffic
and has a flexural strength equal to or better than conventional concrete pavement
[DELATTE, 2008, P. 43]. This economical paving construction method is commonly used
when strength, durability, and economy are most critically needed. Although several mix
formulations for constructing RCC are available, the industry has not sanctioned a stan-
dard method for this type of design.
RCC offers several advantages—the primary of these being a low initial cost. Because
no forms, dowels, or reinforcing steel are required, placement of RCC is fast and saves at
least 15% to 30% of costs compared to slip-form or fixed-form concrete pavement con-
struction. RCC’s simple construction methods allow large areas to be paved quickly with
minimal labor. Also, because RCC shrinks less than traditional methods, joints can be fur-
ther apart.
RCC also has drawbacks—it does not look like other types of concrete and has a
coarser finish than traditional concrete pavements. Traffic traveling at higher speeds would
be subjected to a rough, bumpy ride. Again, RCC is commonly used when strength, dura-
bility and economy are most important. Applications for RCC include port, intermodal,
and military facilities; parking, storage, and staging areas; intersections; and low-speed
A LTE R N AT I V E CONC RE T E PAV E M E NT ROA D CONS T RUC T ION 167

roads. Contractors in central Ohio have used RCC on city streets for several years because
its cost is comparable to asphalt’s initial costs—sometimes even less. RCC’s durability also
minimizes rutting that can appear with a full-depth asphalt base course. Light traffic and
emergency vehicles can drive upon newly created RCC pavement almost immediately—
within 3 meters behind the paver, if necessary. However, truck traffic should be prevented
from driving on the pavement for at least a few days to allow the RCC to achieve its design
strength [PALMER, 2005].

Precast Slabs
Pavement using precast concrete slabs is being researched as a possible alternative to tra-
ditional poured concrete roadways. The slabs are fabricated or assembled off-site and
delivered for installation on a prepared foundation (either existing pavement or re-graded
foundation). This allows for greater quality control, and because the slabs are at their opti-
mum strength upon installation, traffic can access the roads relatively quickly after
construction.
Several proprietary jointed precast concrete pavement systems have been developed,
including the Fort Miller Super-Slab System, the URETEK Method, the URETEK Stitch-
In-Time Process, the Kwik Slab System, and the ModieSlab System.
Though each of these systems incorporates precast concrete slabs in their construc-
tion, they differ somewhat. The Fort Miller Super-Slab System (www.super-slab.com)
incorporates precast slabs that are placed on precision-graded fine bedding material (max-
imum aggregate size of 12 mm) and then fitted with standard dowel bars to ease load
transfer.
URETEK’s two systems are both designed for repair work (www.uretekworldwide.
com/page.php?page_id=34). In the URETEK Method, the panels are delivered to the
construction site and placed into an excavated site. High-density polyurethane foam is
used to lift, realign, seal, and void-fill concrete slabs, which are resting directly on base
soils. The Stitch-In-Time Process restores load transfer to the slabs by “stitching” the
panel to the existing slab or another panel via fiberglass boards [TAYABJI, BUCH, AND
KOHLER, PP. 8–9].
The Kwik Slab System (www.kwikslab.com) essentially simulates jointed reinforced
concrete pavement sections. Using Kwik Joint steel couplers, this system interlocks precast
concrete panels, allowing two-way (2-way) rebar continuity throughout the entire pave-
ment slab. Although there is a limit to the total length of panels that can be connected, and
a need to provide expansion joints, expansion joints have not yet been incorporated into
this system [TAYABJI, BUCH, AND KOHLER, P. 11].
ModieSlab was originally developed in the Netherlands as part of the “Road to the
Future” program in 2000. This technology features full-width, precast concrete slabs that are
placed on underlying precast reinforced concrete crossbeams with prestressed anchors and
A LTE R N AT I V E CONC RE T E PAV E M E NT ROA D CONS T RUC T ION 168

sliding planes to resist subsidence. The slabs themselves feature gutters to discharge rainwa-
ter that permeates through an upper layer of porous concrete. Additionally, channels are
built into the slab to quickly rid the surface of excess rainwater. ModieSlab also contains an
underlying pipeline system to regulate the roadway’s temperature by maintaining cooler
pavement temperatures in the summer, to control dilation, and by maintaining warmer
temperatures in the winter, to eliminate ice and snow without the use of road salts. However,
the most substantial benefits of ModieSlab are the high noise reduction (approximately
6dB) and skid resistance that it provides in traffic [TAYABJI, BUCH, AND KOHLER, P. 12].

Endnotes
“Base and Sub-Base Layers.” European Concrete Paving Association Newsletter, European Concrete
Paving Association, Brussels, Belgium, 2009. <www.eupave.eu/documents/activity-areas/
base-and-sub-base-layers.xml?lang=en>

Cable, James. Reassessing Two-Lift Paving, National Concrete Pavement Technology Center, Iowa
State University, 2004. <www.intrans.iastate.edu/pubs/t2summaries/two-lift.pdf>

Cable, James, and Daniel P. Frentress, Two-Lift Portland Cement Concrete Pavements to Meet
Public Needs. Center for Portland Cement Concrete Pavement Technology, Iowa State
University, 2004. <http://publications.iowa.gov/2942/1/TwoLiftReport.pdf>

Delatte, Norbert. Concrete Pavement Design, Construction and Performance, London: Taylor &
Francis, 2008.

Fast Track Concrete Pavements. American Concrete Pavement Association, Skokie, Illinois, 1994.

Nickelson, Bob. FDR Recent Projects: Coweta County Uses Innovative Construction Techniques to
Rebuild County Roads. Portland Cement Association, Skokie, Illinois. <www.cement.org/
pavements/pv_sc_fdr_coweta.asp>

Palmer, William D., Jr., “Paving with Roller Compacted Concrete,” Public Works magazine,
January 1, 2005. <www.pwmag.com/industry-news.asp?sectionID=770&articleID=271458>

Peilert, Joseph; Lamond, James. Significance of Tests and Properties of Concrete and Concrete-
Making Materials. West Conshohocken, Pennsylvania: ASTM International, 2006.

RCC Pavement Roller-Compacted Concrete. The Cement Association of Canada. <www.cement.


ca/index.php/en/Highways/RCC_Pavement_Roller-compacted_concrete.html>

RCC Pavement: Roller-Compacted Concrete. Southeast Concrete Association. <www.secement.


org/PDFs/FINAL_RCC_Brochure.pdf>

Soil Cement Pavement. Portland Cement Association, Skokie, Illinois, 2010. < www.cement.org/
pavements/pv_sc.asp >

“Take a Look at Two-Lift Concrete Paving,” Focus, August 2007 <www.tfhrc.gov/focus/


aug07/02.htm>

Tayabji, Shiraz, Neeraj Buch, and Erwin Kohler. “Precast Concrete Pavements—Current
Technology and Future Directions.” <www.precastsolutions.org/PrecastSolutions/PDFS/
PCPS_Intro.pdf>
A LTE R N AT I V E CONC RE T E PAV E M E NT ROA D CONS T RUC T ION 169

Van Dam, Thomas, and Joep Meyer. Life-Cycle Analysis: A Powerful Tool for Improving the
Sustainability of Concrete Pavements. National Concrete Pavement Technology Center, Iowa
State University, 2009. <www.cptechcenter.org/projects/two-lift-paving/documents/Life-
CycleAnalysis--VanDam.pdf>

Vanikar, Suneel. “Two-Lift Concrete Paving & Use of Geotextiles as a Bond Breaker.” <www.
virginiadot.org/business/resources/Materials/Virginia_Concrete_Presentations/2P-Two-Lift_
ConcPaving_UsingGeotextile_as_Bondbreaker.pdf>
TRANSFER OF
TECHNOLOGY
TO THE RUSSIAN
MARKETPLACE

THE TRANSFER OF TECHNOLOGY from Europe to North America, and from North
America to Europe, is a two-way (2-way) conduit that remains fairly open to the flow of
best practices. This is not to say that information flows evenly in both directions, however;
each location has its strengths, and the transfer of technology and best practices has an
open format conducive to learning. The exchange of ideas and free-market thinking that
prevails in these two very different markets has led to a strong bond in advancements, sci-
ence, and methodology.
This report on concrete pavement methodology has several constituent components—
but one (1) of the preeminent goals of this analytical study of the ongoing contemporary
practices in the European and North American theaters is to foster a three-fold (3-fold)
conduit for experiences in Europe, in North America, and eventually in Russia. By means
of this study, it is hoped that the transfer of technology can be encouraged among these
three regions. This study itself focuses on the two (2) developed theaters: Europe and
North America as models of both good and poor practices. It is left to the Russian reader
to assess which technologies will work best in the Russian marketplace.
Russia is unique, to state the obvious. The Russian concrete sector therefore is also
unique, and the needs of the concrete industry are not necessarily those of the North
American market or the European market—but distinctly Russian. The requirements of
the growing concrete pavement sector, in particular, can be best addressed by taking the
best practices of Canadian, Scandinavian, American, German, Italian, Spanish, British,
and other markets. The concrete industry is so diverse and specialized that general
concrete practices simply cannot be disseminated entirely in any one (1) document. The
concrete pavement sector, which this study looks at both in generalities and specifically, is
a major area of growth and opportunity in Russia. This pavement sector has some similari-
ties to each market, however, and certain opportunities for transfer of technology.

170
TR A N SF E R O F T E CHNOL OGY TO T HE RUS S I AN M ARKE T P L A C E 171

Similarities to th e N orth A merican M arketplace: Siz e


The Russian market, with its vast and diverse geography, is in many ways a reflection of
North America. In square kilometers the two (2) regions are very comparable, with Can-
ada, the United States, and Mexico together measuring roughly the same size as the for-
mer Soviet Union. The diversity of topography is also somewhat comparable, with desert
regions, coastal regions, northern forests, and freshwater lakes. These similar landscapes
allow for shared approaches to construction practices in general, and to concrete produc-
tion in particular—specifically concrete batching, heating, cooling, placement, finishing,
curing, maintenance, and repair.
The advent of mobile paving/batching plants in the U.S. and the development and
refinement of slip-form pavers allowed the North American concrete paving industry to
exponentially grow during the last 30 years. It is suggested that the adaptation of these two
(2) primary concrete equipment processes be looked to for adaptation to the Russian mar-
ketplace, with a major exception: the utilization of the outdated and inferior tilt-drum
mixer. The tilt-drum mixer, as explained in the Batching Technology and the Mixing Con-
crete chapters, is a less-than-appropriate mixing device, as it segregates the materials and
is not fit for higher-performance concrete paving mix production. In most regions of
Europe, the compulsory-type mixer is required for all important and civil concrete paving
and structures, as well as for any taxpayer-funded construction. It is recommended that
the large-capacity American-style portable paving plant—paired with a European com-
pulsory-style mixer—offers the proper through-put and mobility required in approaching
the Russian paving market. The larger North American–sized plants, especially when
combined with the larger-sized 8-meter European compulsory mixer, are well suited to
both the urban and rural Russian paving marketplace. (See www.sicoma.biz and www.
bhs-sonthofen.de for examples of appropriate machinery.)
Modern slip-form paving equipment is recommended for continued adaptation to
the Russian marketplace. Proper training of operations personnel is required, as these
machines are more complex and are computer-controlled. It is further recommended that
young persons, with no previous experience in operation of slip-form paving machinery,
should be enlisted in the training for operation of this new-style equipment. Past experi-
ence in Europe and North America indicates that personnel experienced in the older-style
machinery were far less able to adapt quickly and effectively. Training young Russian con-
struction workers on the proper operation of these large-scale slip-form paving machines
is therefore strongly recommended.
Many of the manufacturers of slip-form paver machinery offer training classes on the
operation of the equipment. The manufacturers have a vested interest in the development
of properly trained and experienced operators of their respective machines.
The number of kilometers to be paved in an average concrete pavement project for
highways is comparable in the North American marketplace to that in the Russian mar-
ketplace. The freeze-thaw cycles of the northern regions and the dry, arid conditions of the
TR A N SF E R O F T E CHNOL OGY TO T HE RUS S I AN M ARKE T P L A C E 172

southern regions are remarkably alike. Long distances and large-capacity concrete pours
are very much alike. The lack of infrastructure support and logistics in the Russian coun-
tryside notwithstanding, much of the approach to the overall undertakings of these larger
pours and longer distances can be emulated by the Russian developers, as the Russian
market continues its fast-track metamorphosis. The transfer of North American method-
ology is therefore recommended in general, albeit understood that certain support mech-
anisms will need to be developed ahead of the adapted advancements.
An analogy of the petroleum industries’ transfer of technologies over the past decade
may be valid for the concrete paving industry. Similar conditions and applications would
seem to apply.

Similarities to th e E u ropean M arketplace:


Q u ality R eq uir ements
A conduit of technological exchange exists between Russia and most of Europe on a
number of fronts in construction. Interestingly, while a great deal of European construc-
tion equipment is successfully adapted to the Russian market, success is generally limited
to urban situations. In rural applications, the transfer of technology has not proven as suc-
cessful, generally due to the dissimilar distances and quantities involved.
In recent surveys of construction trade shows held in Russia, it has been noted that
over 90% of the construction machinery marketed at these Russian trade shows was from
European manufacturers, with less than 5% of the manufacturers coming from Russia,
and even less from North America. Even so, the equipment is only a part of the larger
equation, and geography can account for much of this. The materials science is another
matter, and here the European technology is becoming embraced, as well it should.
European mix designs for concrete pavement, as with much European concrete pav-
ing methodology, are quite relevant to the Russian market. The science of concrete tech-
nology in Europe generally is agreed to be as advanced as in North America, and superior
in actual practice. The gap-gradation and price-driven modifications to mix designs in the
North American market, for instance, are not something that can be recommended for
higher-quality concrete pavements in Russia. European specifications, especially those of
Germany and the countries it influences (such as Italy), are particularly applicable to the
development of high-quality Russian pavements.
The advent of 21st-century pavement techniques, and the newer methods now being
developed as outlined in other sections of this report, bode well for the Russian market-
place as carefully analyzed and adopted from the European marketplace. The recommen-
dation to the Russian market is to take the best of the North American market (the
large-scale concrete production and slip-form equipment and methodology) and the best
of the European market (the advanced mix design, mixing technologies, and the well-
graded aggregate methodology) and combine them—putting these proven approaches to
work in Russian concrete paving.
SECTION 3

COM PREH ENSIV E


L ITERATURE & DESIG N
STANDARDS REV IEW
LATEST IN
ROAD DESIGNS

RECENT RESEARCH IN ROAD DESIGN has covered a wide range of topics from
using recycled materials to fiber reinforcement. A review of recent studies on road design
identified the following significant research.

Altoubat, S.A., J.R. Roesler, D.A. Lange, and K.A. Rieder, “Simplified Method
for Concrete Pavement Design with Discrete Structural Fibers.” Construction
and B uilding Materials 22.3 (March 2008): 384–93.

The addition of discrete macro fibers to plain concrete increased the flex ural capacity of con-
crete slabs supported on ground. Flex ural strength tests ( e.g., ASTM C7 8 ) did not capture
the added toughening mechanisms that fibers imparted to concrete slabs. Based on the
results of small- and large-scale testing, the equivalent flex ural strength ratio ( Re,3 ) was
used to quantify the increased flex ural capacity of synthetic and steel-fiber reinforced con-
crete slabs over plain concrete slabs. An effective flex ural strength approach was proposed for
the design of fiber-reinforced concrete pavements. The method applied a percentage increase
to the modulus of rupture ( MO R) of plain concrete based on the measured Re,3 value for the
fiber-reinforced concrete mix ture. The effective MO R accounts for the contribution of fibers
to the added slab flex ural capacity and fits into ex isting linear elastic-based design guidelines
for concrete pavement. Large-scale test results verified the proposed design approach for
fiber-reinforced concrete pavements. The proposed method is applicable for concrete mix -
tures reinforced with relatively low volume fractions of fibers that produce Re,3 values
between 2 0 % and 5 0 % . Such fiber concrete mix tures are practically and economically attrac-
tive, as they can be mix ed, placed, and compacted with normal paving techniques.

Bank, Lawrence C., Michael Oliva, Han Bae, and Bryan V. Bindrich, “Hybrid
Concrete and Pultruded-Plank Slabs for Highway and Pedestrian Bridges.”
Construction & B uilding Materials 24.4 (April 2010): 552–558.

Studies on two novel uses of hybrid structural members consisting of commercially pro-
duced glass-reinforced pultruded ribbed fiber-reinforced polymer ( FRP) planks and

174
L AT E S T I N ROAD DE S I GNS 175

concrete are discussed in this paper. Pultruded planks are produced by all the major pul-
truders in the world and are utiliz ed primarily as decking for platforms. These highly
optimiz ed panels have the potential to be used in many other infrastructure applica-
tions, but their flex ural stiffnesses have generally been too low to be used in highway and
pedestrian bridges, due to current span requirements. H owever, when used “ compos-
itely” with concrete or cementitious materials in a hybrid form, they have the potential
to be much more widely used. Two research studies conducted on two possible hybrid
systems of different structural depths are discussed in this paper. The first study describes
the use of pultruded planks as permanent formwork in highway bridge decks where the
plank is used with concrete to produce a solid slab of 2 0 0 -mm depth that is typical of
slabs seen in highway bridge decks. The second study describes the use of pultruded
planks in pedestrian bridge decks where the pultruded plank is used with a cement-
board or a cast-in-place concrete panel to produce a hollow slab of 7 5 -mm depth that is
typical of timber decking used in FRP pedestrian bridges. Tests were conducted on
beam-type specimens of the hybrid slabs to investigate the load-transfer mechanisms
between the pultruded plank and the cementitious “ overlays” for both the 7 5 -mm and
2 0 0 -mm depths. From analysis of the load-carrying capacity and failure mechanisms of
the hybrid slabs, it was concluded that such hybrid slabs are viable systems for both
highway and pedestrian bridge decks. A bridge deck using the 2 0 0 -mm-deep hybrid slab
system was recently constructed on a highway in Wisconsin, in the United States.

Broek, Anna Vander, “Self-Healing Concrete.” Forb es 184.8


(November 2009): 46, 48–9.

Civil engineer V ictor Li has created a bendable concrete that mends its own cracks. Li’ s
self-repairing concrete is based on a material he devised in 1 9 9 0 , called “ engineered
cementitious composite,” which rarely cracks and, if it does, the cracks are rarely more
than 5 0 microns wide. Any hairline cracks absorb moisture in the air to “ grow” new
concrete, and calcium ions within the cracked concrete combine with moisture and car-
bon diox ide to restore the concrete’ s original strength.

Choi, Seongcheo, and Moon Won, “Design of Tie Bars in Portland Cement
Concrete Pavement Considering Nonlinear Temperature Variations.”
T ransportation Research Record 2095 (2009): 24–33.

Tie bar design at longitudinal construction joints in Portland cement concrete pave-
ment is based on the so-called subgrade drag theory. According to this theory, the
required max imum spacing between tie bars is inversely proportional to the widths of
the lanes that are tied together. As the number of lanes increases to accommodate
greater traffic demand, tie bar spacing decreases. In a recent project in H ouston,
Tex as, the width of pavement lanes tied together was great enough to require 3 0 -cm
L AT E S T I N ROAD DE S I GNS 176

tie bar spacing. The Tex as Department of Transportation’ s design standards for con-
tinuously reinforced concrete pavement require the same spacing for both tie bars and
transverse steel. The use of 3 0 -cm spacing for both tie bars and transverse steel results
in a substantial increase in the cost of the project. The subgrade drag theory assumes
uniform temperature distribution through the concrete slab depth and that no curling
effect is incorporated. This assumption needed reconsideration. Field testing was con-
ducted to investigate slab behavior and to develop more rational designs for tie bars
and transverse steel. Concrete displacement gauges were installed at the free edge of
the concrete slab to measure transverse horiz ontal as well as vertical displacements.
Steel strain gauges were installed at different locations in tie bars. The concrete slab
showed curling behavior; the stresses in tie bars clearly demonstrated the effects of
curling. Theoretical analysis was conducted, and the results verified the field data. A
new design method was proposed for tie bars.

Joh, S.H., Sung-Ho Joh, M.R. Cho, Mi-Ra Cho, T.H. Kang, Tae-Ho Kang, S.A. Kwon,
Soo Ahn Kwon, J.H. Nam, and Jeong Hee Nam, “Pavement Integrity Scanner to
Characterize Modulus Contrast between Near-Surface and Deeper Material in
Concrete Pavements.” T ransportation Research Record 2070 (2008): 39–48.

A major distress problem in rigid concrete pavements is early deterioration of near-


surface layers. The quality of near-surface material governs the long-term perfor-
mance of concrete pavement. If the near-surface material of the concrete layer is
cured sufficiently well at an appropriate moisture and temperature, the concrete
pavement system can be ex pected to endure fatigue loading of traffic throughout its
design life span. A field measurement device, referred to as a pavement-integrity
scanner ( PiScanner) , was developed for use in quality control and quality assurance
programs for concrete layers of pavement systems. PiScanner is based on the enhanced
resonance search ( ERS) technique, which evaluates the vertical variation of elastic
modulus or compressive strength of the concrete layer in pavement systems. This
information can be used in the construction period of a concrete pavement to evalu-
ate the curing process of near-surface materials. In the implementation of the PiScan-
ner, two ( 2 ) design schemes were incorporated: a dedicated hardware device to
perform measurements and an automated analysis algorithm to minimiz e user
ex pertise required for analysis. The resulting PiScanner allows ERS measurements to
be performed in about 2 minutes, with complete analysis of the measurements in
about 3 minutes. Two ( 2 ) field investigations were performed on a concrete test slab
and at an ex press highway to prove the reliability and feasibility of PiScanner. V erti-
cal variations of elastic modulus of concrete layer were successfully obtained by
PiScanner.
L AT E S T I N ROAD DE S I GNS 177

Madison, Adam, “The Economy: Crushed and Recycled.” Rock Products 112.10
(October 2009): 30–32.

The article discusses the role of the recycling industry in the aggregate market and cites
the increasing demand for recycled construction materials in the U.S. It mentions that
recycled concrete or asphalt pavement prepared on site is financially advantageous for
the tax payer who pays for the highway projects. It ex plains the 2 0 0 8 Construction
Aggregates Recycled Data. Also cited is the forecast of economists on the better future
for the aggregate industry.

Tsai, Chih-Ta, Lung-Sheng Li, Chien-Chih Chang, and Chao-Lung Hwang,


“Durability Design and Application of Steel Fiber–Reinforced Concrete in
Taiwan.” Arab ian J ournal for S cience and Engineering 34.1B (April 2009): 57–59.

This paper presents the way durability has been introduced to steel fiber–r einforced
concrete in Taiwan. It is generally acknowledged that steel fibers are added to improve
the toughness, abrasion resistance, and impact strength of concrete. H owever, a locally
developed mix ture design method, the densified mix ture design algorithm ( DMDA) ,
was applied to solve not only the entanglement or balling problem of steel fibers in con-
crete or to produce steel fiber–r einforced self-consolidating concrete ( SFRSCC) with
ex cellent flow-ability, but also to increase the durability by reduction in the cement
paste content. By dense packing of the aggregates and with the aid of pozz olanic mate-
rial and superplasticiz er ( SP), concrete can flow honey-like with less entanglement of
steel fibers. Such SFRSCC has already been successfully applied in several projects, such
as construction of a low-radiation waste container, bus station pavement, road deck
panel, and two ( 2 ) art statues. So it is recommended that the SFRSCC can be used for
improving the performance of ordinary steel fiber–r einforced concrete in many ways
and should be considered for increasing the life cycle of a concrete structure.

Won, Moon, Yoon Ho Cho, Shiraz Tayabji, and Jianbo Yuan, “New Technologies
in Construction and Rehabilitation of Portland Cement Concrete Pavement and
Bridge Deck Pavement.” Selected Papers from the 2009 GeoHunan International
Conference, G eotechnical S pecial Pub lication No. 19 6 (August 2009).

This proceedings paper is divided into two sections, Innovative Techniques for Bridge
Deck Pavement Design and Construction, and New Analysis Techniques and Materials
for Portland Cement Concrete Pavement System. H unan, one of the largest commercial
provinces in China, is rapidly developing into a modern epicenter of international
finance and trade. Recent construction in many parts of China has provided geotechni-
cal and pavement engineers with great opportunities for creating cutting-edge solutions
to problems involving highway and bridge pavements. This publication will be valuable
to geotechnical engineers and professionals, as well as construction professionals.
L AT E S T I N ROAD DE S I GNS 178

Zhang, Lei, Wei Huang, and Zhendong Qian, “Optimization Design of Long-
Span Steel Bridge Deck Pavement System.” New Technologies in Construction
of Portland Cement Concrete Pavement and Bridge Deck Pavement: Selected
Papers From the 2009 GeoHunan International Conference.” G eotechnical S pecial
Pub lication No. 19 6 (August 2009): 9–18.

Asphalt mix tures have been widely used for wearing surfaces on steel decks of long-span
bridges. Due to the unique mechanical and environmental conditions of steel decks,
requirements for paving materials and paving patterns are different from those for reg-
ular road pavements. Currently there are no commonly adopted design theories or pro-
cedures for steel bridge deck surfacing. Premature failures of deck pavements have often
been observed on steel bridges, particularly on the newly built long-span bridges in
China. Therefore, two ( 2 ) main objectives are identified in this paper as: 1 ) integrating
the separate design processes for the bridge structure and the deck pavement into one
interactive process; 2 ) establishing the methodology of determining pavement system
parameters. Systematic research has been conducted to develop the design theory and
procedure from many aspects, including pavement materials and structure, mechanis-
tic characteristics of the wearing surfaces on steel decks, fatigue properties, ax le-load
equivalency conversion, and optimiz ation design procedure. The empirical-mechanistic
approach is employed for the optimiz ation of design. The case study shows that it is
more feasible to consider the orthotropic deck and its pavement as a whole structural
system during bridge design.
CLIMATE-RELATED
ISSUES

THE EFFECT OF THE ENVIRONMENT ON CONCRETE and the effect of concrete on


the environment are both pressing issues. Construction in arctic or earthquake-prone
areas presents special challenges that can best be met with new design considerations.
Likewise, recycling old concrete as a means of limiting CO emissions, disposing of waste,
and lowering costs will become increasingly important as time goes on. A summary of
recent research follows.

Charbeneau, Randall J., and Michael E. Barrett, “Drainage Hydraulics of Perme-


able Friction Courses.” Water Resources Research 44.4 (May 2008).

This paper describes solutions to the hydraulic equations that govern flow in permeable
friction courses (PFC). PFC is a layer of porous asphalt approximately 50 mm thick that
is placed as an overlay on top of an existing conventional concrete or asphalt road surface
to help control splash and hydroplaning, reduce noise, and enhance quality of storm water
runoff. The primary objective of this manuscript is to present an analytical system of
equations that can be used in design and analysis of PFC systems. The primary assump-
tions used in this analysis are that the flow can be modeled as one-dimensional, steady-
state Darcy-type flow and that slopes are sufficiently small so that the Dupuit-Forchheimer
assumptions apply. Solutions are derived for cases where storm water drainage is confined
to the PFC bed and for conditions where the PFC drainage capacity is exceeded and pon-
ded sheet flow occurs across the pavement surface. The mathematical solutions provide
the drainage characteristics (depth and residence time) as a function of rainfall intensity,
PFC hydraulic conductivity, pavement slope, and maximum drainage path length.

Chen, Dar-Hao, “Investigation of a Pavement Premature Failure on a Weak and


Moisture-Susceptible Base.” Journal of Performance of Constructed Facilities 23.5
(September/ October 2009): 309–313.

An investigation was conducted to determine the root cause of the premature pavement
failure. The premature pavement failure occurred in the form of rutting and alligator

179
CL I M AT E - R E L AT E D I S S UES 180

cracking. Although the affected portion was repaired by removing and replacing the top
75-mm asphalt concrete (AC), the repaired AC experienced recurring rutting and alli-
gator cracking in a few weeks. Through extensive field and lab testing, it was found that
the weak base is the root cause of the premature failure and the brittleness of the AC is
secondary. However, both the base and AC were built according to plan and met the
current material and field density requirements. It was concluded that density alone for
construction quality control is not sufficient, as it was not able to protect premature
failures from occurring. Although there are many different ways to minimize premature
failures, an immediate action is to include proof rolling in construction quality control.
Proof rolling has been used with success to ensure proper compaction and to locate
unstable areas, as the stability is greatly influenced by the degree of densification
achieved during compaction.

Jin-Ping, Li, and Yu Sheng, “Analysis of the Thermal Stability of an Embankment


under Different Pavement Types in High-Temperature Permafrost Regions.” Cold
Regions Science and Technology 54.2 (October 2008): 120–123.

The thermal stability of the embankment, including the influence of the height and
width of the embankment in high-temperature permafrost regions, was simulated and
discussed respectively in terms of different natural geographic conditions for asphalt
concrete pavements and cement concrete pavements. The aim is how to change the
maximum thawed depth under the asphalt concrete pavements and cement concrete
pavements with increase in service life and embankment width. The simulated results
based on the observation of data from the G214 road of Zuimatan are analyzed in this
paper. It is very important to select a suitable embankment width and height in high-
temperature permafrost regions, especially under the climate warming conditions. The
conclusion that 2.0-meter embankment height for the 8.5-meter embankment width
and 2.5-meter embankment height for the 12.0-meter embankment width are the
most suitable for the cement concrete pavement construction in the experimenting
road, respectively.

Kim, Hyung Bae, and Kwang-Ho Lee, “An Innovative Rehabilitation Approach
for the Bridge Deck Pavement.” New Technologies in Construction of Portland
Cement Concrete Pavement and Bridge Deck Pavement: Selected Papers from
the 2009 GeoHunan International Conference.” Geotechnical Special Publication
No. 196 (August 2009): 19–27.

This study presents a rehabilitation method for the bridge deck pavement that has been
damaged due to a moisture-related failure of the asphalt concrete. In order to quickly
drain the water that has infiltrated into the bridge deck pavement, this rehabilitation
method adopts a new drainage system including a porous asphalt concrete layer and a
polymer concrete–based waterproofing layer so that the water cannot stay and cause
CL I M AT E - R E L AT E D I S S UES 181

moisture-related failure on the bridge deck. The most important factors for this drain-
age system are to meet satisfactory performance of the repairing material for the dam-
aged concrete bridge deck, the waterproofing layer, and to develop a fast-track
construction technique with a limited traffic-blocking time. The porous asphalt mix-
tures for the new drainage system was suggested with the maximum aggregate size of
10 mm, and was validated through various physical and mechanical laboratory tests to
confirm its performance characteristics. In this study, methyl methacrylate (MMA)–
type material was introduced for the waterproofing on the concrete bridge deck and
damaged bridge deck repairing, and the results from a series of mechanical tests for
evaluating the performance of the MMA material shows it has good capability for
waterproofing and repairing of the concrete bridge deck which is subject to a moisture-
induced neutralization of the cement concrete. In addition, to evaluate the field perfor-
mance of the new rehabilitation method, a field study was conducted on a real bridge.
Field performance observations on both the MMA and pavement materials indicated
that the new method performs much better than traditional methods in draining water
that has infiltrated into pavement layers.

Li, Jin-ping, Yu Sheng, and Jinzhao Zhang, “Study on Diseases of Cement


Concrete Pavement in Permafrost Regions.” Cold Regions Science & Technology
60.1 (January 2010): 57–62.

Because the study on cement concrete pavement in permafrost regions is not enough,
we have made an exploration on it along G214 of China. After that the problems of
pavement along the typical road of G214 in permafrost regions were firstly analyzed
based on field investigations. The results indicated that the distribution, reasons and
mechanism of problems of cement concrete pavement were different from that of the
adjacent asphalt pavement and the general regions. Moreover, the results indicated that
longitudinal fracture and broken slabs were the major types of problems of cement con-
crete pavement in permafrost regions. Finally, based on the field measurements of set-
tlement along the G214, a possible critical settlement value of fracture failure was
calculated under specific conditions. Analysis of the reasons showed that the problems
of cement concrete pavement in permafrost regions were related to the thermal stability
of embankment and thaw deformation of embankment.

Peng, Yunfeng, and Yunlong He, “Structural Characteristics of Cement-


Stabilized Soil Bases with 3D Finite Element Method.” Frontiers of
Architecture and Civil Engineering in China 3.4 (December 2009): 428.

Cement-stabilized soil bases have been widely used in expressways due to cement’s high
strength, appropriate stiffness, good water resistance, and frost resistance. So far, the
structural characteristics and mechanical behaviors of cement-stabilized soil bases were
not investigated so much. In this paper, the 3D elastic-plastic finite element method
CL I M AT E - R E L AT E D I S S UES 182

(FEM) was used to analyze the mechanical behaviors and structural characteristics of
cement-stabilized soil bases from construction to operation. The pavement filling and
the traffic loading processes were simulated, and a contact model was used to simulate
the contact behavior between each layer of the pavement. Considering the construction
process, the structural characteristics and mechanical behaviors of cement-stabilized
soil bases were studied under asphalt-concrete pavement conditions. Furthermore, the
general rules of deformations and stresses in cement-stabilized soil bases under differ-
ent conditions were discussed, and some suggestions were put forward for the design
and construction of cement-stabilized soil bases.

Proceedings of the Institution of Civil Engineers 161.2 (May 2008): 55–63.

A test road consisting of jointed plain concrete pavement divided into 22 sections was
constructed by the Korea Highway Corporation to investigate the environmental effects
on the development of joint cracks. These 22 sections were observed for approximately
three to seven (3 to 7) weeks after construction. The pavement placed in the morning
presented a higher rate of crack development at the saw-cut joints, due to a positive
built-in temperature difference. Additional joint cracking was delayed for a few days
after construction because the pavement was exposed to rainfall. For pavement sections
placed in the afternoon, drying shrinkage appeared to have a greater effect on crack ini-
tiation at the saw-cut joints than temperature did up to three or four (3 or 4) days after
placement, when tensile creep had decreased sufficiently. The effect of joint crack time
on the long-term joint behavior was also investigated in the test road two (2) years after
placement of the concrete.

Saeed, Athar, and Michael I. Hammons, “U se of Recycled Concrete as U nbound


Base Aggregate in Airfield and Highway Pavements to Enhance Sustainability.”
Airfield and Highway Pavements: Efficient Pavements Supporting Transportation’s
Future (2008): 497–508.

This research evaluated the use of recycled concrete aggregate (RCA) from airfield and
highway pavements to enhance sustainability based on engineering, economic, and
environmental criteria. The objective was to develop evaluation and construction
guidelines by defining the minimum standards for RCA as unbound base. The research
approach included contacts with industry representatives, technical assessments, site
visits, and performance review of airfield pavements with unbound RCA layers. RCA
can be used as unbound base material if produced from uncontaminated PCC. All vir-
gin aggregate tests and their limits are applicable to RCA except the sulfate soundness
test, which is waived for RCA due to the incompatibility of PCC components with the
chemical reactants used in the test. RCA should not be used where there is a potential
for sulfate exposure from subgrade soils, ground water, or other external sources. RCA
from ASR-distressed PCC can be used considering site conditions. RCA is not a hazard
CL I M AT E - R E L AT E D I S S UES 183

to the environment. An economic analysis can be conducted by considering initial


material and construction costs for both RCA and virgin aggregate.

Tao, Mingjiang, Zhongjie Zhang, and Zhong Wu, “Simple Procedure to Assess
Performance and Cost Benefits of U sing Recycled Materials in Pavement Con-
struction.” Journal of Materials in Civil Engineering 20.11 (November 2008):
718–725.

The use of recycled materials in pavement engineering has a great potential to benefit
our society in terms of reducing demands on natural pavement materials, reducing
environmental problems, and conserving energy. However, pavement design/construc-
tion practitioners often hesitate to use recycled materials due to the lack of cost benefit
and performance information. This paper presents a simple approach to evaluate eco-
nomic effects of using recycled materials in pavement construction based on in situ
pavement testing procedures. These testing procedures, including dynamic cone pene-
trometer, California bearing ratio, Dynaflect, and falling weight defelectometer, are
commonly employed by state highway agencies. A full-scale accelerated pavement test
section was built to investigate the performance of different base course materials: Loui-
siana Class II crushed limestone, foamed-asphalt-treated recycled asphalt concrete, fly-
ash-stabilized blended calcium sulfate (BCS), and BCS stabilized with the 120-grade
ground granulated blast-furnace slag (GGBFS). Among these base materials, GGBFS-
stabilized BCS was found to have the highest strength and stiffness. On the basis of these
field testing results, life-cycle cost analyses showed that the GGBFS-stabilized BCS pro-
vides a durable and cost-effective alternative to traditional pavement base materials.

X injun, Li, Mihai Marasteanu, Jason Bausano, R. Christopher Williams, and


Ben Worel, “Factors Study in Low-Temperature Fracture Resistance of Asphalt
Concrete.” Journal of Materials in Civil Engineering 22.2 (February 2010):
145–152.

An investigation of various factors that can affect the fracture resistance of asphalt
mixtures at low temperatures was conducted. The fracture resistance of 28 asphalt
mixtures—which represent a combination of factors including binder type, binder
modifier, aggregate type, asphalt content, and air voids—was evaluated by performing
semi-circular bending fracture tests at three (3) low temperatures. Two (2) parame-
ters, fracture energy and fracture toughness, were calculated from the experimental
data. The statistical analysis indicated that fracture resistance is strongly dependent
on temperature and significantly affected by the type of aggregate and air-void con-
tent. The results of the analysis also confirmed the significance of binder grade and
modifier type with relation to cracking resistance of asphalt mixtures. Asphalt content
was also found to be significant for fracture energy but not a significant factor affecting
fracture toughness.
CL I M AT E - R E L AT E D I S S UES 184

Ying-mei, Yin, and Rong-hu Zhang, “Design of Water Diversion and Drainage
in Old Concrete Pavements.” Journal of Guangdong University of Technology 26.1
(March 2009): 88–91.

According to the practice of G321 test road reconstruction, water had bad effects on the
pavement because serious water-induced distress exists widely in traditional concrete
pavements. It discusses the water diversion and water removal design of pavements to
prevent the permeability of free water infiltrating into pavements, to remove rapidly the
pavement surface water and the water infiltrating into pavements, and to strengthen
the waterproof property of pavement structures. Based on the project practice, some
suggestions on water diversion and water removal design of old concrete pavements are
put forward.

Zhanga, X iong, and Jean-Louis Briaud, “Coupled Water Content Method for
Shrink and Swell Predictions.” International Journal of Pavement Engineering
11.1 (February 2010): 13–23.

Lane/shoulder drop-off or heave due to expansive soils beneath is a common distress in


both concrete and asphalt pavements, resulting in substantial discomfort, safety haz-
ard and vehicle damage. The situation becomes worse when exposed to moisture
changes, particularly in Texas, Colorado, Arizona, California, etc., where many soils
are expansive soils. Therefore, expansive soils pose great challenges for the design of
pavements and foundations. One of the most difficult issues for designing pavements
and foundations on expansive soils is to predict the volume change of the soils. In this
study, current methods for movement predictions are summarized and their relation-
ships and shortcomings are discussed. It is found that all these methods have the same
theoretical basis. Based on the theory of unsaturated soil mechanics, it is found that the
water content and the mechanical stress can be used to determine the soil status. Hence,
a void ratio versus mechanical stress and water content surface is constructed, which
coupled both mechanical stress and suction’s influences on volume change of expansive
soils. The new surface is developed into a coupled water content method that can be
used to predict the potential vertical swell and the potential vertical shrink simultane-
ously, while all the existing methods can only predict the potential vertical rise. The
method is used to analyze the data collected from a construction site at Arlington,
Texas. The predicted movements match the measured data reasonably well. The method
is simple and overcomes the shortcomings existing in the current movement prediction
methods for expansive soils.
CRITICAL POINTS
FOR SUCCESSFUL CONCRETE
ROAD CONSTRUCTION

THE GOAL OF CREATING A HIGHWAY with a long life span can depend on critical
issues ranging from the concrete mix, to drainage, to the local environment. Recent
research on problems and solutions follows.

Abraha, Dawit G., David C. Sego, Kevin W. Biggar, and Robert Donahue, “Sulfur
Concrete for Haul Road Construction at Suncor Oil Sands Mines.” Canadian
Geotechnical Journal 44.5 (May 2007): 564–578.

The feasibility of constructing mine roads at oils sands mines (Fort McMurray, Alberta)
using concrete prepared from bitumen extraction and upgrading by-products and mine
wastes (sulfur, fly ash, coke, and tailing sand) is evaluated. An extensive laboratory test
program, including unconfined compression testing, sonic velocity measurement, and
split tensile and freeze-thaw durability tests, was carried out to characterize the physi-
cal and mechanical properties of different mix designs of sulfur concrete. A study of the
geochemical interaction of sulfur concrete with the near-surface environment included
short-term interaction of surface-exposed sulfur concrete during the construction and
operational life of the haul road and long-term interaction of sulfur concrete with
groundwater following its eventual burial with mine waste in the mined-out pits. Haul
road test sections were designed based on the critical strain and resilient modulus design
method. Stress and strain distributions in the selected haul road cross-section induced
by the truck tires were calculated using finite element analysis. Required pavement
layer thicknesses were then determined on the basis of the truck loads, along with resil-
ient modulus and strength of the sulfur concrete and subgrade material using the criti-
cal strain and resilient modulus design method.

Ahammed, M. Alauddin, and Susan L. Tighe, “Pavement Surface Mixture,


Texture, and Skid Resistance: A Factorial Analysis.” Airfield and Highway
Pavements: Efficient Pavements Supporting Transportation’s Future (2008): 370–384.

Skidding on wet pavements contributes to a substantial portion of highway crashes. The


resistance to skidding, however, depends on the microtexture and macrotexture avail-
185
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 186

able on pavement surfaces. Several past studies have focused on this aspect but with
inadequate or inconsistent conclusions. The uniqueness of this study is that surface tex-
ture performance has been evaluated controlling the variability in aggregate mineral-
ogy, environmental condition, construction, and service. Portland cement concrete
(PCC) specimens were prepared from a single mixture to evaluate the true effect of vari-
ous surface textures on friction properties. Asphalt concrete (AC) surfaces with the
same construction record were tested to examine the true effect of mix properties on
surface texture and friction. Pavement surface texture was measured using the sand
patch method and Automated Road Analyzer (ARAN ), while the skid resistance was
measured using the British Pendulum and skid trailer. Analysis has shown that a Mean
Texture Depth (MTD) of about 1.8 mm is the optimum macrotexture for maximum
surface friction on textured concrete surfaces. Exposed aggregate concrete may not be a
preferred texture because the benefit of sand microtexture is lost with washing out of
surface mortar. AC surfaces with complex macrotexture have shown to perform differ-
ently from PCC surfaces with simple macrotexture pattern. The hypothesis that British
Pendulum N umber (BPN ) is dependent only on surface microtexture and represents
low speed friction has appeared to be invalid. The skid number-speed gradient is not
something universal but varies from mix to mix. Several statistically significant models
have also been developed correlating the skid resistance with the asphalt mix grading
composition and surface macrotexture and with concrete surface macrotexture.

Alungbe, Gabriel, “Geotextiles in Flexible Pavement Construction.”


Tech D irections 64.2 (September 2004): 22–23.

The article discusses the use of geosynthetic textiles in pavement construction. There are
two major types of pavement: flexible and rigid. Rigid pavement consists of a Portland
cement concrete pavement with optional base course between the pavement and sub-
grade. W heel load distribution in rigid pavement takes place over a relatively wide area
of the subgrade because of the pavement’s rigidity and high modulus of elasticity. The
performance of some flexible pavement in some climates has proved disappointing
because transverse cracks developed within a few years of construction. Other forms of
flexible pavement distress include map or alligator cracks, ruts, frost heave, and reflec-
tive cracking. The useful life of a highway or airfield pavement depends on the ability of
an underdrain system to adequately remove free water from under the pavement. The
use of geotextile material can keep the soil particles from migrating into the gravel layer
while permitting water to pass through. Some other advantages of using geotextile for
pavement drainage include reduced trench excavation; elimination of some trench
shoring; ability to use less-expensive gravel, such as pit-run or back-run; and elimina-
tion of sand filter in dual-media backfill. Geotextile is used to separate natural soil from
pavement materials. This prevents the fine-grained soils in the subgrade from intermix-
ing with the high-quality aggregate in the base or subbase. This will in turn insure that
the aggregate maintains its design load-bearing capacity throughout the design life.
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 187

Black, Tom, “Paving the Road to the 21st Century.” American City & County 112
(November 1997): 44.

As technologies continue to develop and advance, engineers and scientists are realizing
how much effort and money can be saved in the long run by making intelligence and
foresight key factors in the design of new pavements. This realization is due largely to a
well-coordinated intense technological initiative for improving roads, the Strategic
Highway Research Program (SHRP). SHRP was established by Congress in the 19 80s as
a $ 150 million research program to improve the performance, durability, and longevity
of the country’s roads to make them safer for both motorists and highway workers. Top-
ics discussed include the Superwave program, different pavement mix recipes, concrete
pavements, and electronic hardware and software developments.

Cheng, Chin-Min, Panuwat Taerakul, Wei Tu, Behrad Zand, Tarunjit Butalia,
and William Walker Wolfe, “Surface Runoff from Full-Scale Coal Combustion
Product Pavements during Accelerated Loading.” Journal of Environmental
Engineering 134.8 (August 2008): 591–599.

In this study, the release of metals and metalloids from full-scale Portland cement
concrete pavements containing coal combustion products (CCPs) was evaluated by
laboratory leaching tests and accelerated loading of full-scale pavement sections
under well-controlled conditions. An equivalent of 20 years of highway traffic loading
was simulated at the OSU /OU Accelerated Pavement Load Facility (APLF). Three (3)
types of Portland cement concrete driving surface layers were tested, including a con-
trol section [ i.e., ordinary Portland cement (PC) concrete] containing no fly ash and
two (2) sections in which fly ash was substituted for a fraction of the cement— i.e.,
30% fly ash (FA30) and 50% fly ash (FA50). In general, the concentrations of minor
and trace elements were higher in the toxicity characteristic leaching procedure
(TCLP) leachates than in the leachates obtained from synthetic precipitation leaching
procedure and ASTM leaching procedures. Importantly, none of the leachate concen-
trations exceeded the TCLP limits or primary drinking water standards. Surface run-
off monitoring results showed the highest release rates of inorganic elements from the
FA50 concrete pavement, whereas there were little differences in release rates between
PC and FA30 concretes. The release of elements generally decreased with increasing
pavement loading. Except for Cr, elements were released as particulates (> 0.45 µ m)
rather than dissolved constituents. The incorporation of fly ash in the PC cement con-
crete pavements examined in this study resulted in little or no deleterious environ-
mental impact from the leaching of inorganic elements over the lifetime of the
pavement system.
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 188

Cho, Yoon-Ho, and Sung-Hun Yeo, “Application of Recycled Waste Aggregate to


Lean Concrete Subbase in Highway Pavement.” Canadian Journal of Civil
Engineering 31.6 (December 2004): 1101–1108.

As aggregates recycled from various types of construction waste are continuously being
produced, interest has focused on how to apply them for use in highway pavement. This
paper considers the application of waste aggregates to lean concrete, based on basic
mechanical property tests and environmental toxicity. Compared with natural aggre-
gates, waste aggregates derived mainly from recycled concrete have low specific gravity
and high water-absorption characteristics. After testing their environmental toxicity, it
was found that waste aggregates do not release any metallic ions when introduced to
alkaline conditions but do release a small but seemingly harmless amount of metallic
ions when introduced to acidic solutions. Concrete made with waste aggregates has sig-
nificant limitations in strength, particularly flexural strength, which is the main param-
eter of quality control and design for concrete pavement. It is therefore not practical to
use waste aggregates for the surface layer of concrete without using additives or special
treatments. It is possible, however, to apply concrete with waste aggregates for lean
bases. In testing, lumps of asphalt, cement paste, bricks, and glass were classified as
impurities and were observed for changes in strength based on the percentage of impu-
rities used. If the amount of impurities is greater than 25% , the 7-day compressive
strength does not meet the strength requirements specified for lean concrete.

Fischetti, Mark, “Paving the Way.” Scientific American 293.4 (October 2005):
96–97.

Focuses on the construction of roadways. Design of the modern road, using thick depos-
its of compressed crushed, angular stones covered by hot tar to bind the top layer, by
J ohn McAdam in the early 19 th century; U se of asphalt and concrete roads in current
times, both of which are built up in layers comprising increasingly finer, denser, and
harder rock particles; Recent innovations in road design, including perpetual pavement
that is built on a tough base of asphalt, stone-matrix designs in which the size and
shape of stones within surface layers is controlled, and porous pavement, which allows
water to filter through instead of running off the road.

Ghafoori, Nader M., and Matthew W. Tays, “Abrasion Resistance of Early-


Opening-to-Traffic Portland Cement Concrete Pavements.” Journal of Materials
in Civil Engineering 19.11 (November 2007): 925–935.

The research investigation presented herein is intended to study the abrasion resistance
of early-opening-to-traffic Portland cement concrete pavements, also known as fast-
track concrete. The selected matrices are examined using two (2) categories of opening-
to-traffic times (6 and 8 hours). Type III Portland cement and three (3) different cement
factors with and without an accelerating admixture were used. The trial mixtures are
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 189

examined for plastic properties (slump, air content, bleeding, setting times, and adia-
batic temperature), bulk characteristics (demolded unit weight and compressive
strength), and resistance to abrasion. Depth of wear and rate of deterioration as func-
tions of matrix proportions, opening-time categories, and testing age (up to 20 minutes
or 3 mm, whichever is reached first) are determined. The influence of other parameters
such as cement factor, curing age, and accelerating admixture on resistance to wear of
the selected matrices are discussed. Other properties, namely rate of abrasion, relative
gain of abrasion, coefficient of variation, and abrasion index, are examined. Finally,
the relationship between the abrasion resistance (depth of wear) and bulk characteristic
(compressive strength) at both opening and maturity ages are investigated. The findings
of this investigation shall benefit highway agencies and municipalities by providing
abrasion resistance data on a wide range of matrix constituents and proportions that
may be used for fast-track concrete construction.

Gorkem, Cagri, and Burak Sengoz, “Predicting Stripping and Moisture-Induced


Damage of Asphalt Concrete Prepared with Polymer-Modified Bitumen and
Hydrated Lime.” Construction and B uilding Materials 23.6 (June 2009): 2227–2236.

Many highway agencies have been experiencing premature failures that decrease the
performance and service life of pavements. One of the major causes of premature pave-
ment failure is the moisture damage of the asphalt concrete layer. Many variables affect
the amount of water damage in the asphalt concrete layer, such as the type of aggregate,
bitumen, mixture design and construction, level of traffic, environment, and the addi-
tive properties that are introduced to the bitumen, aggregate, or bitumen-aggregate
mixture. This study is aimed to determine the effect of additives such as hydrated lime
as well as elastomeric (SBS) and plastomeric (EV A) polymer-modified bitumen (PMB)
on the stripping potential and moisture-susceptibility characteristics of hot-mix asphalt
(HMA) containing different types of aggregate (basalt-limestone aggregate mixture and
limestone aggregate). The stripping properties and moisture-susceptibility characteris-
tics of the samples have been evaluated by means of captured images and the N icholson
stripping test (ASTM D 16 6 4) as well as the modified Lottman test (AASHTO T 283),
respectively. The results indicated that hydrated lime addition and polymer modifica-
tion increased the resistance of asphalt mixtures to the detrimental effect of water.
Moreover, it was found out that samples prepared with SBS PMB exhibited more resis-
tance to water damage compared to samples prepared with EV A PMB.

Hassan, Yasser, Wael Bekheet, and A. El Halim, “Modelling in Situ Shear


Strength Testing of Asphalt Concrete Pavements Using the Finite Element
Method.” Canadian Journal of Civil Engineering 28.35 (June 2001): 541–544.

Rutting is one of the well-recognized road surface distresses in asphalt concrete pave-
ments that can affect the pavement service life and traffic safety. Previous studies have
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 190

shown that the shear strength of asphalt concrete pavements is a fundamental property
in resisting rutting. Laboratory investigation has shown that improving the shear
strength of the asphalt concrete mix can reduce surface rutting by more than 30% , and
the SU PERPAV E mix design method has acknowledged the importance of the shear
resistance of asphalt mixes as a fundamental property in resisting deformation of the
pavement. An in situ shear strength testing facility was developed at Carleton U niver-
sity, and a more advanced version of this facility is currently under development in
cooperation with the Transportation Research Board and the Ontario Ministry of
Transportation. In using this facility, a circular area of the pavement surface is forced
to rotate about a normal axis by applying a torque on a circular plate bonded to the
surface. The pavement shear strength is then related to the maximum torque. This prob-
lem has been solved mathematically in the literature for a linear, homogeneous, and
isotropic material. However, the models for other material properties are mathemati-
cally complicated and are not applicable to all cases of material properties. Therefore,
developing a model that can accurately analyze the behavior of asphalt concrete pave-
ments during the in situ shear test has proven pivotal. This paper presents the develop-
ment of a three-dimensional finite element model that can simulate the forces applied
while measuring the shear strength of the asphalt concrete pavement. A comparison
between the model results and those obtained from available analytical models and
field measurements proved the accuracy of the developed model.

He, Zhao-Yi, Qiang Xie, and Li Liang, “Finite Difference Analysis on Cantilever
Slab Model with Void beneath the Cement Concrete Slab Corner.” Tumu Jianzhu
yu Huanjing Gongcheng, Journal of Civil, Architectural and Environmental
Engineering 31.2 (April 2009): 95–99.

The advantages and shortcomings of traditional method and Finite Difference Method
(FDM) for the calculation of cantilever slab model with void beneath cement concrete
slab corner and analysis on the results were carried out. The traditional method assumes
that the cement concrete slab with void below was a cantilever slab with constant sec-
tions. The influence of loading eccentricity was not considered while more than two (2)
wheel loads were exerted on the concrete pavement, which can be overcome by using
Finite Difference Method. W ith FDM, the stress and displacement distributions of any
parts of the cement concrete slab can be easily achieved. The results indicate that when
the concrete slab with void beneath the slab corner was assumed to be a cantilever slab,
regardless of calculation by traditional method or by FDM, the cement concrete pave-
ment will be destroyed even under the condition of minimum traffic grade. The fatigue
stress and road deflection obtained by FDM are bigger than that obtained by the tradi-
tional formula, so bigger safety factor should be employed in the road design.
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 191

Jung, Jong-Suk, E.B. Owusu-Antwi, and Ji-Hwan An, “Analytical Procedures for
Evaluating Factors That Affect Joint Faulting for Jointed Plain Concrete Pave-
ments Using the Long-Term Pavement Performance Database.” Canadian Journal
of Civil Engineering 33.10 (October 2006): 1279–86.

The objective of this study was to identify and quantify design and construction features
most important to joint faulting of joint plain concrete pavements. W ith data obtained
from the long-term pavement performance (LTPP) database, an analysis approach that
combined pavement engineering expertise and modern data analysis techniques was to
develop guidelines for improved design and construction of Portland cement concrete
(PCC) pavement. The approach included typical preliminary analyses, but emphasis
was placed on using a series of multivariate data analysis techniques. Discriminant
analysis was used to develop models that classify individual pavement into perfor-
mance groups developed by cluster analysis, which was used to partition the pavements
into three (3) distinct groups representing good, normal, and poor performance. These
models can be used to classify and evaluate additional or new pavements performance
throughout the pavement’s design life. To quantify the levels of the key design and con-
struction features that contribute to performance, the classification and regression tree
procedure was used to develop tree-based models for performance measure. The analy-
sis approach described was used to develop the guidelines on the key design and con-
struction features that can be used by designers to decrease joint faulting of jointed
plain concrete pavements (J PCPs).

Kayhanian, M., A. Vichare, P.G. Green, and J. Harvey, “Leachability of Dissolved


Chromium in Asphalt and Concrete Surfacing Materials.” Journal of Environ-
mental Management 90.11 (August 2009): 3574–80.

Leachate metal pollutant concentrations produced from different asphalt and concrete
pavement surfacing materials were measured under controlled laboratory conditions.
The results showed that, in general, the concentrations of most metal pollutants were
below the reporting limits. However, dissolved chromium was detected in leachate from
concrete (but not asphalt) specimens and more strongly in the early-time leachate sam-
ples. As the leaching continued, the concentration of Cr decreased to below or close to
the reporting limit. The source of the chromium in concrete pavement was found to be
cement. The concentration of total Cr produced from leachate of different cement com-
ing from different sources that was purchased from retail distributors ranged from 124
to 6 41 µ g/L. This result indicates that the potential leachability of dissolved Cr from
concrete pavement materials can be reduced through source control. The results also
showed that the leachability of dissolved Cr in hardened pavement materials was sub-
stantially reduced. For example, the concentration of dissolved Cr measured in actual
highway runoff was found to be much lower than the Cr concentration produced from
leachate of both open and dense graded concrete pavement specimens under controlled
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 192

laboratory study. It was concluded that pavement materials are not the source of pollut-
ants of concern in roadway runoff; rather most pollutants in roadway surface runoff are
generated from other road-use or land-use sources, or from (wet or dry) atmospheric
deposition.

Lee, S. Joon, “Mechanical Performance and Crack Retardation Study of a


Fiberglass-Grid-Reinforced Asphalt Concrete System.” Canadian Journal of
Civil Engineering 35.10 (October 2008): 1042–1049.

This paper presents constitutive design considerations for a semi-rigid, resin-impreg-


nated fiberglass grid that enhances the mechanical performance of an asphalt concrete
overlay and provides a pattern to retard crack propagation. An acrylic polymer resin
covers the grid-structured fiberglass strands, thus making their viscoelastic nature
compatible with the asphalt materials. The remarkable temperature susceptibility of
the resin is observed using time-temperature superposition. The grid tensile strength
of 100 kN /m was qualified using standard measurements. Three-point bending and
cyclic fatigue loading mechanical tests affirm the retardation performance of the grid-
reinforced system in terms of crack propagation. The retardation mechanism is affected
by synchronizing the cracking pattern with the flexural stress-strain profile through
image analysis. A N ational Center for Asphalt Technology full-scale road testing pro-
gram provides assessment of the grid-reinforced pavements. A visual inspection of the
grid extracted from the traffic zone showed fair condition of the pavement.

Lee, Ying-Haur, “TKUPAV: Stress Analysis and Thickness Design Program for
Rigid Pavements.” Journal of Transportation Engineering 25.4 (July/August 1999):
338–346.

An alternative procedure for the determination of critical stresses and the thickness
design of rigid pavements is presented. The effects of a finite slab size, different gear con-
figurations, widened outer lane, tied concrete shoulder, second layer, and thermal curl-
ing are investigated. The well-known ILLI-SLAB finite-element program was used for
the analysis. The program’s applicability for stress estimation is further validated by
reproducing very favorable results to the test sections of Taiwan’s second northern high-
way, the AASHO road test, and the Arlington road test. Prediction models for stress
adjustments are developed using a projection pursuit regression technique. A simplified
stress analysis procedure is proposed and implemented in a user-friendly program
(TKU PAV ) to facilitate instant stress estimations at three (3) critical locations of the
slab; namely, the edge, corner, and interior. A modified Portland Cement Association
stress analysis and thickness design procedure is also proposed and incorporated into
the TKU PAV program. This program may be utilized for various structural analyses
and designs of jointed concrete pavements in both the metric system and the U .S. cus-
tomary system.
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 193

Lian-yu, Wei, Zhang Hai-rong, and Liu Zhao-wei, “Mechanical Analysis of


Cement Concrete Overlay on Asphalt Pavement.” Journal of Hebei University
of Technology 37.6 (December 2009): 94–99.

Cement concrete overlay is a new, effective, rapid, cost-effective and simple mainte-
nance for old asphalt pavement; although in the U nited States there has been a great
deal of testing on road stress analysis, it is still limited. Large-scale finite element analy-
sis software AN SY S has been used in building a 3D structural model to make mechani-
cal analysis about the cement overlay on old asphalt pavement. Analyses: vehicle load
layer elastic modulus and thickness of the pavement overlay of the subject impact force;
temperature under the thermal stress and temperature and vehicle load coupling; and
the impact of repairing the cracks on the old road to the load stress analysis. The article
gives some suggestions for the design of cement concrete overlay.

Mucinis, Darjusas, Henrikas Sivilevicius, and Rolandas Oginskas, “Factors


Determining the Inhomogeneity of Reclaimed Asphalt Pavement and Estimation
of Its Components’ Content Variation Parameters.” B altic Journal of Road and
B ridge Engineering 4.2 (2009): 69–79.

W ith a view to saving bitumen and mineral substances and also for the purpose of pre-
venting environmental pollution, the old asphalt pavement (AP) may be milled without
increasing the thickness of pavement structure and subjected to hot recycling for reuse
in the same or any other location. However, the suitability of reclaimed asphalt pave-
ment (RAP) for reuse depends on its composition (the content and gradation of bitu-
men) and the properties of its components. Hence, the present study aims at providing
a systematic analysis of factors that determine the inhomogeneity of RAP. To this end,
RAP samples, taken from RAP being loaded from the milling machine onto the trans-
port means and from RAP stored in the territorial storage facilities of asphalt concrete
(AC) plants, were examined in a fully accredited laboratory to determine, for each sam-
ple unit, the composition (the content and gradation of aged soluble bitumen), the
content of moisture and the gradation of non-extracted granules. The statistical char-
acteristics showing the actual homogeneity of RAP sampled from different sources were
also estimated and analyzed. Further, a model was elaborated for comparing the homo-
geneity of RAP from different sample sets by the max values of standard deviations in
the percentage mass passing the sieves and, based on the experimental data, derived the
respective regression equation. The results of the performed comparison showed the
absence of any statistically significant difference in terms of homogeneity between the
three (3) different RAP sample sets, i.e., RAP-1 sampled from the road, RAP-2 taken
from the stockpiles being formed in different locations, and RAP-3 obtained from RAP
kept stockpiled in an open storage facility. The study was finalized by determining, in
line with the Specification of Technical Requirements TRA ASFALTAS 08, the max con-
tent of RAP allowed for inclusion in a hot-mix asphalt (HMA) mixture, depending on
the actual homogeneity of RAP determined.
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 194

Papayianni, I., and E. Anastasiou, “Cost-Effective Mixtures for Concrete


Pavements.” Airfield and Highway Pavement: Meeting Today’s Challenges with
Emerging Technologies (2006): 1008–1017.

In concrete pavement construction, large quantities of industrial by-products could be


absorbed for the benefit of economy and sustainability, especially when large infrastruc-
ture projects are in progress. In Greece, a great amount of high-calcium fly ash (12 million
tons per year) and steel slag (2.5 million tons per year) is produced. The current utiliza-
tion rate of the above by-products is relatively low. Before their utilization, the local indus-
trial by-products should be tested and used properly to achieve optimum concrete mixtures
for pavements. In this paper, local steel slag is used as aggregate in combination with a
cement–fly ash binding system for the production of concrete mixtures whose strength
ranges from 20 to 70 MPa. Apart from compressive strength, the flexural tensile strength,
elastic modulus of elasticity, fracture toughness, fatigue strength, and resistance to abra-
sion have been measured in a series of concrete compositions. Based on the results, it
seems feasible to design and apply low-cost concrete mixtures for pavements, since they
may contain considerable amount of by-products for which the main cost is their trans-
portation. The technical benefits related to their use in concrete are also present.

Qun, Yang, Guo Zhongyin, and Shi Chunxiang, “Study of Axle Load Conversion
of Cement Concrete Pavement in Road Tunnel with High Groundwork Modulus.”
Journal of Tongj i University 34.7 (July 2006): 891–4.

The formula of axle-load conversion provided in the specifications of cement concrete


pavement design for highway is only functional under the condition that the modulus
of groundwork does not exceed 212 MPa. It is not applicable in the case of a road tun-
nel with high groundwork modulus. A three-dimension (3D) finite method is used to
calculate the critical stress on the bottom of the slab, and the formula of axle-load con-
version in road tunnel is formulated on the basis of the theory of fatigue equivalence. It
can be applied to designing the concrete pavement and forecasting the life span of the
pavement in a road tunnel.

Roesler, Jeffery P.E., Cristian Gaedicke, David A. Lange, Salvador Villalobos,


Robert Rodden, and Zachary Grasley, “Mechanical Properties of Concrete
Pavement Mixtures with Large-Size Coarse Aggregate.” Airfield and Highway
Pavement: Meeting Today’s Challenges with Emerging Technologies (2006):
516–527.

The use of larger maximum-size aggregates (MSA) in concrete mixtures can increase
the load-transfer efficiency across aggregate interlock joints if small crack widths are
achieved. Aggregate interlock joints can minimize the number of dowels required on
airport and low-volume concrete pavements, and thus decrease their total construction
cost. This paper describes a laboratory study evaluating the strength, fracture, and
shrinkage properties of airport concrete mixtures with 25-mm and 38-mm MSA and
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 195

several total cementitious contents. Although the split tensile and flexural strength of
the larger MSA mixtures were reduced by 20% , the 28-day fracture energy (GF) was
similar between the mixtures. The 1-day GF was significantly greater for the larger
MSA mixture, indicating a greater joint shear transfer capability. The brittleness of the
concrete was also reduced as the MSA increased. Finally, the concrete free shrinkage
was similar for the same mix proportions but different MSA.

Ryu, Sung Woo, Min-Young Park, Jung-Hee Nam, Zuog An, Jong-Oh Bae, Yoon-
Ho Cho, and Seung Woo Lee, “Initial Behavior of Thin-Bonded Continuously
Reinforced Concrete Overlay (CRCO) on Aged Jointed Concrete Pavement.” New
Technologies in Construction of Portland Cement Concrete Pavement and
Bridge Deck Pavement: Selected Papers from the 2 0 0 9 GeoHunan International
Conference ( GSP 196) (2009): 101–106.

A thin-bonded CRCO (Continuously Reinforced Concrete Overlay) on a J CO (J ointed


Concrete Overlay) was constructed at Seo Hae Ahn Express Highway in Korea. This
project includes two (2) experimental sections: 1 ) a J CO with saw-cut joints over exist-
ing transverse joints, and 2) a CRCO, which employs transverse and longitudinal steel
reinforcing rods, that is placed on top of an existing J CO. After milling 5 cm off the top
of the J CO, an overlay thickness of 10 cm remains. Several V W SGs (V ibrating W ire
Strain Gauges) were installed to evaluate curling, delamination behavior, and crack
propagation of the thin-bonded CRCO. This paper describes the early behavior of this
overlay and compares the current and new rehabilitation methods available in Korea.

Songlin, Ma, Hou Xiangshen, and Wu Dayong, “Research on Structural Design


of Interlocking Concrete Block Pavement for Rural Road.” Proceedings of the 9th
International Conference of Chinese Transportation Professionals, ICCTP 2009:
Critical Issues in Transportation System Planning, D evelopment, and Management
358 (August 2009): 2535–2538.

In China, thousands of kilometers of rural roads need to be paved. In order to reduce


the construction and maintenance costs, the interlocking concrete block pavement can
be used in low-traffic rural roads. The structural design criteria of interlocking concrete
block pavement with different types of base causes were researched. The results indi-
cated that the design criteria are different for different base course. The methods of the
criteria calculating and structure design were given.

Tutumluer, Erol, Laith Tashman, and Halil Ceylan, “Soil and Material Inputs for
Mechanistic-Empirical Pavement Design.” Proceedings of Sessions of Geo-
Denver 2007: New Peaks in Geotechnics, Geotechnical Special Publication No. 169
(February 2007):

This special publication compiles and presents recent experiences dealing with soil and
pavement material laboratory and field characterizations and input requirements for
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 196

mechanistic-empirical pavement design. The emphasis has been placed on the state of
the practice in the implementation of the Mechanistic Empirical Pavement Design
Guide. Among the topics covered in the first part of the special publication are the labo-
ratory and field characterization needs in relation to modulus input requirements for
pavement foundation geomaterials, i.e., soils and base/subbase aggregates, contribution
of geosynthetic reinforcement to granular layer stiffness, resilient modulus property
input requirements for geofoam used in pavements, input property needs for the design
and construction of roller-compacted concrete pavements, and a new neural network
modeling approach for the mechanistic-empirical design of rubblized concrete pave-
ments. The second part of this special publication focuses primarily on field evaluations
and collecting input properties for forensic and rehabilitated design purposes and deals
with topics such as investigations of subgrade soil settlements of an existing highway
alerted by pavement distresses, dynamic cone penetrometer testing of emulsified asphalt-
stabilized gravel roads, airport runway evaluation using the rolling dynamic deflectom-
eter, base course use of an asphalt recycled pavement material blended with fly ash, and
performance evaluations of rigid pavement sections built on stabilized sulfate soils.

Xu, Tao, and Xiaoming Huang, “Effects of Material Design Parameters on


Mechanical States of Bridge Deck Pavement.” New Technologies in Construction
and Rehabilitation of Portland Cement Concrete Pavement and Bridge Deck
Pavement: Selected Papers from the 2 0 0 9 GeoHunan International Conference,
Geotechnical Special Publication No. 196 (August 2009): 61–67.

Bridge deck pavement suffers from more complicated stresses than ordinary pavement,
and mechanical states of the paving layers are influenced by material design parame-
ters. After mechanical computation, the effects of material design parameters on the
mechanical state and the variation law of pavement are analyzed. Analytical results
indicate that following the traditional HAM design method can not adapt to actual
mechanical states, and easily result in usual distress types on the asphalt concrete pav-
ing. Economical and reasonable paving materials should be designed or developed
according to mechanical paving characteristics, such as multilevel dense built-in modi-
fied HMA, fiber-reinforced asphalt concrete, and modified SMA, which can harmonize
paving layers as a whole, adapt to the deformations of bridge decks, improve the
mechanical states of the paving layers, increase the service years, and offer some helpful
references for similar bridge deck pavement design.

Zhang, Lei, Zhendong Qian, and Yun Liu, “Special Considerations and Analysis
for the Bascule Bridge Pavement.” New Technologies in Construction of Portland
Cement Concrete Pavement and Bridge Deck Pavement: Selected Papers From the
2 0 0 9 GeoHunan International Conference ( GSP 196) (2009): 28–34.

The bascule bridge is one of the most appropriate bridges for ports and inland rivers,
from structural and economic viewpoints. However, the difficulty in successfully paving
C R I TI C A L P OI NT S F OR C ONC RE T E ROAD C ONS T RUC T ION 197

steel-deck pavement remains to be solved for its wide applications. Serious premature
failures of the pavements are found on many bascule bridges. So far, there is little dis-
cussion about this issue. In this study, the dynamic response of bascule bridge deck
pavement was analyzed based on the different rotation velocities and different pave-
ment elastic modulus for the Haihe Bridge being built in Tianjin, China, one of the larg-
est bascule-type bridges in the world. The results indicate that the shear stress of the
pavement will increase greatly when the main girder opens wide, and simultaneously
the girder rotation angle and rotation speed will lead to the non-uniform distribution
of the shear stress. Since it is hard to control the pavement modulus due to the changing
environmental temperature in the real world, an appropriate rotation velocity is sug-
gested for reducing the shear stress between the pavement and the steel deck.
QUALITY CONTROL FOR
DESIGN AND CONSTRUCTION
OF CONCRETE ROADS

QUALITY CONTROL IS A PROCESS THAT BEGINS with design and continues through
construction. Theoretical approaches to predicting failure have helped anticipate prob-
lems, and empirical research has helped discover and correct difficulties. Recent research
includes the following.

Butalia, Tarunjit S., “Accelerated Load Testing of Full-Scale Asphalt and Con-
crete Pavements Constructed of CCPs.” Abstracts with Programs—Geological
Society of America 39.6 (October 2007): 175.

With support from the Ohio Coal Development Office, the USDOE’s Combustion
Byproducts Recycling Consortium, and the American Coal Ash Association, The Ohio
State University embarked upon the accelerated load testing of full-scale pavement sec-
tions made of CCPs. Flexible (asphalt) and rigid (concrete) pavements were studied. In
this study, three (3) concrete and three (3) asphalt sections (one [1] control and two [2]
CCP sections each) were constructed at the OSU/OU Accelerated Pavement Load Facil-
ity in Lancaster, Ohio. Class F fly ash and lime were used in the construction of the sub-
grade of all sections. Fly ash and bottom ash were incorporated into the construction of
the base and subbase of the CCP sections of both the asphalt and concrete pavements.
The CCP concrete slab included 30% and 50% replacement of cement with Class F fly
ash. The instrumented pavements were constructed in summer of 2003. Instrumenta-
tion included pavement strain and displacement gauges, soil pressure and moisture
gauges, pore water-pressure transducers, and environmental sampling ports. The con-
trol and CCP sections were subjected to accelerated load testing equivalent to 20 years
of state highway traffic. The structural and environmental instrumentation pavement
response was measured at regular intervals of loading. Overall, the full-scale CCP pave-
ment sections demonstrated similar or better performance and response to load than
the control sections. At the end of the pavement design life, none of the asphalt sections
had failed in terms of rutting or fatigue cracking. All three (3) concrete sections demon-
strated good performance, and no fatigue cracking was observed in any of the concrete

198
QUA LI TY CO N TR O L F O R D E S I GN AND CONS T RUC T I ON OF C ONC RE T E ROAD S 199

sections over the design life or even much later, i.e., at the end of the full-scale testing
program. Both CCP base/subbase mixes out-performed the control mix based on the
collected response, performance, and FWD testing data throughout the whole full-scale
testing program. During the second phase of the full-scale testing program, the two (2)
CCP sections showed better resistance to the adverse environmental conditions and
exhibited better performance than the control section.

Chen, Dar Hao, “Field and Lab Investigations of Prematurely Cracking Pave-
ments.” Journal of Performance of Constructed Facilities 21.4 (July August 2007):
293–301.

A forensic study was conducted to identify the cause of the premature cracking on three
(3) recently completed projects that were built with the same design. Nondestructive
(ground penetration radar, falling weight deflectometer [FWD], GeoGauge, and Porta-
ble FWD), nuclear density gauge, dynamic cone penetration, and extensive laboratory
tests were performed. It was found that the initial stiffness of the treated base was found
to be excessively high by FWD backcalculation. Some sections of the backcalculated
base moduli were over 20.7 GPa. This indicates that the layer is excessively brittle for a
base material, similar to lean concrete. Six (6) specimens, which were made without a
mellowing period, exhibited cracks. There was no cracking for six (6) specimens that
had 2 days of mellowing. It was concluded that the culprit of the transverse cracking in
the main lanes was the shrinkage of lime-treated base layers. The longitudinal cracks
are related to the edge-drying, and the transverse cracks are related to the insufficient
mellowing period. Based on the findings of this study, the District implemented a 2-day
mellowing period for Quicklime treated caliche base. Three (3) newly constructed pave-
ments (age 8, 5, and 2 months) were surveyed. No cracking can be observed so far, and
the District thinks the cracking problem has been mitigated by the 2-day mellowing
period. Without the mellowing period, cracking had normally occurred 1 to 2 months
after construction.

Chen, Dar Hao, “Inspection and Condition Assessment Using Ground-


Penetrating Radar.” Journal of Geotechnical & Geoenvironmental Engineering
136.1 (January 2010): 207–214.

The nondestructive mapping of anomalies and voids under roadway pavements is criti-
cal to highway authorities because of the potential loss of support that would lead to
safety hazards. 400-MHz ground-coupled penetrating radar (GCPR) was used in this
study to characterize the subsurface conditions of three (3) roadway pavements (SH359,
IH40, and U.S. 290). The extents of the anomalies in horizontal and vertical directions
were visible in GCPR images. Coring, boring, and lab testing were performed to verify
the settlement and source of the moisture on SH359. The source of the moisture was
from the leaking water pipe, as indicated by the high chloride and chlorite contents.
QUA LI TY CO N TR O L F O R D E S I GN AND CONS T RUC T I ON OF C ONC RE T E ROAD S 200

A 1.8-m deep void (3.8 m3 in volume) under IH40 and a 1.8 m×4.6 m×3.7 m (30.6 m3
in volume) void under U.S. 290’s reinforced concrete pavements were successfully iden-
tified by GCPR and verified by field boring and coring. Fortunately, the voids near the
drainpipes were detected by GCPR in time. Otherwise, the void would have increased
in size, and that could have led to a severe hazard. This study has successfully demon-
strated that the GCPR is able to identify anomalies and voids. Therefore, engineers can
utilize the information from GCPR to undertake remedial actions with confidence.

Delatte, Norbert, Aleksandar Mrkajic, and Daniel I. Miller, “Field and Labora-
tory Evaluation of Pervious Concrete Pavements.” Transportation Research
Record 2113 (2009): 132–139.

Portland cement pervious concrete (PCPC) is a material of increasing interest for


parking lots and other applications. PCPC typically consists of coarse aggregates, Port-
land cement, water, and various admixtures. Similar materials are used for cement-
stabilized drainage layers in highway and airport pavements. In this research, in-service
PCPC pavements were inspected in the field, and cores were removed in order to inves-
tigate properties in the laboratory. Field evaluation methods included visual inspection,
two (2) surface drainage measurements, and an indirect-transmission ultrasonic pulse
velocity (UPV) test. Laboratory testing methods included void ratio, unit weight,
compressive strength, splitting tensile strength, hydraulic conductivity, and direct-
transmission UPV. Because it is compacted on the surface with screeds or rollers, PCPC
generally has higher strength, lower void ratio, and lower permeability at the surface
than at the bottom. Therefore, the properties of the tops and bottoms of core samples
were compared. Generally, the PCPC installations evaluated under this research project
have performed well in freeze-thaw environments with little maintenance required. No
visual indicators of freeze-thaw damage were observed. With the exception of some
installations in which the pore structure was sealed during construction with wet mix-
tures or overcompaction, nearly all sites showed fair to good infiltration capability on
the basis of drain-time measurements.

de Solminihac, Hernan, Marcelo G. Bustos, Anibal L. Altamira, and Juan Pablo


Covarrubias, “Functional Distress Modelling in Portland Cement Concrete
Pavements.” Canadian Journal of Civil Engineering 30.4 (August 2003): 696–703.

Concrete is widely used as a construction material in pavements by public and private


agencies that administer highway networks because of its high durability and capacity
to resist large traffic loads and very rigorous climates. Nevertheless, these agencies have
to estimate the evolution of pavement performance to plan and optimize the applica-
tion of adequate maintenance activities, allowing the pavement to be maintained at an
optimum service level throughout its lifetime. Predictive distress models of the incre-
mental type—that is, models capable of predicting annual increments of different dis-
QUA LI TY CO N TR O L F O R D E S I GN AND CONS T RUC T I ON OF C ONC RE T E ROAD S 201

tress indicators in the pavement—could be very useful tools in the implementation of


maintenance plans, with minimal need for previous data, especially with regard to
information on cumulative traffic loads. This paper offers incremental models for dis-
tress prediction in jointed plain concrete pavements, related to joint problems such as
faulting and spalling, which clearly affect the pavement ride quality. The equations
obtained allow for not only the calculation of distress predictions in analyzing road
maintenance policies, but also the adjustment of the original designs of these pave-
ments, to minimize the occurrence and magnitude of distress problems.

Khazanovich, Lev, Iliya Yut, Shariq Husein, Carly Turgeon, and Tom Burnham,
“Adaptation of Mechanistic-Empirical Pavement Design Guide for Design of
Minnesota Low-Volume Portland Cement Concrete Pavements.” Transportation
Research Record 2087 (2008): 57–67.

A new Mechanistic-Empirical Pavement Design Guide (MEPDG) was recently pro-


posed in the United States. The performance prediction models in the MEPDG were
calibrated using nationwide pavement performance data. This study examined the fea-
sibility of adapting a recent version of the MEPDG for the design of low-volume con-
crete pavements in Minnesota. To achieve this, a comprehensive evaluation of the
MEPDG Version 0.910 performance prediction models was conducted. It was found
that the faulting model produced acceptable predictions, whereas the cracking model
had to be adjusted. Although performance data from the Minnesota Road Research
facility were actively used in calibration of the MEPDG, it is necessary to perform cali-
brations against a wider range of local variables to achieve a practical design procedure
for Minnesota. The cracking model was recalibrated using the design and performance
data for 65 pavement sections located in Minnesota, Iowa, Wisconsin, and Illinois. A
prototype catalog of recommended design features for low-volume Portland cement
concrete pavements in Minnesota was developed using MEPDG Version 0.910. The cat-
alog offers a variety of feasible design alternatives for a given combination of site condi-
tions. It is recognized, however, that Version 0.910 is not the final version of the MEPDG.
Therefore, the catalog should be updated after the MEPDG software is finalized.

Lee, Duk Gyoo, “Dynamic Prediction Model of As-Built Roughness in Asphaltic


Concrete Pavement Construction.” Journal of Transportation Engineering 133.2
(February 2007): 90–95.

This paper develops a dynamic prediction model of a highway pavement contractor’s


quality-based performance using a panel (longitudinal) data analysis. This panel data
modeling uses as-built roughness measurements and pavement and contractor’s char-
acteristics for reconstructed, replaced, and resurfaced pavement projects in Wisconsin
from 1998 through 2002. Several random-effects models were first developed in in-
sample specification, and their modeling performances were measured by Akaike’s
QUA LI TY CO N TR O L F O R D E S I GN AND CONS T RUC T I ON OF C ONC RE T E ROAD S 202

information criteria, which combines goodness of fit and model complexity. Out-of-
sample specifications validated the developed random-effects models by comparing out-
of-sample forecasting accuracies. The results show that the best model has approximately
a 16% mean absolute percentage error. The results finally show that asphaltic concrete
pavement quality of construction can be predicted based on the contractor’s past qual-
ity-based performance and other construction parameters. Therefore, the dynamic pre-
diction model developed in this study could be implemented in the contractor’s
prequalifications required for advanced contracting methods.

Li, Qiang, Kelvin C. P. Wang, and Yanjun Qiu, “Comparative Study of Jointed Plain
Concrete Pavement (JPCP) Design Methodologies in China and the United States.”
International Conference on Transportation Engineering 2007 (2007): 259–264.

The pavement design specifications in China are mechanistic-based with limited capa-
bilities of performance prediction. In the U.S., the various editions of the AASHTO
Guide for Design of Pavement Structures were developed based on the results from the
1960’s AASHO Road Test and a number of studies in the following decades. Since 2002,
the U.S. pavement design standard has progressed to a mechanistic-empirical method,
known as the Mechanistic-Empirical Pavement Design Guide (MEPDG), which is
expected to replace the current AASHTO Guide released in 1993. In this paper, a com-
parative study of Jointed Plain Concrete Pavement (JPCP) design methodologies in
China and the U.S. is presented. The design outputs designed based on the Chinese
JPCP pavement specification are then used in the MEPDG design software to predict
the pavement performances during their design lives. The findings of this study show
that the new MEPDG has advantages in several aspects over the Chinese design proce-
dure. It also demonstrates that the new MEPDG procedure is applicable in China to
establish a general pavement design method and may be helpful to ease the frequently
occurring premature failure problems of pavements in China.

Qian, Wei, Huang Zhendong, Zhendong Qian, and Hedong Niu, “Load Stress
Analysis of Unbonded Prestressed Concrete Pavement.” Traffic and Transporta-
tion Studies 2002 (2002): 1548–1553.

A 20-node isoparametric element is implemented for the stress analysis of prestressed


concrete pavement (PCP). The model of friction force at the slab bottom is presented
and the friction force is found to correspond much more to the working condition of
PCP than what other models have suggested. The longitudinal prestress equivalent
model is established and a stress calculation method of higher accuracy is introduced.
Numerical examples and experimental data both verify the effectiveness of the analyti-
cal model and program. Load stress of the unbonded PCP is researched in detail using
the data from the first domestic experiment road. The test results and comparison anal-
ysis can provide important reference for prospective design and construction of PCP.
QUA LI TY CO N TR O L F O R D E S I GN AND CONS T RUC T I ON OF C ONC RE T E ROAD S 203

Wang, Zhengjun, Yiq iu Tan, and Jiani Wang, “Research on Testing Strength of
Cement Concrete Pavement Slabs with Method of Ultrasonic Flatting.” Logistics:
The Emerging Frontiers of Transportation and Development in China (2009):
2579–2584.

This research applied method of mathematics analysis to judge takeoff position range of
head wave of reflection longitudinal wave so that homologous acoustical parameter
could be determined through analyzing sonic propagation characteristic. On the basis
of this study, incorporating Nen-duo highway construction and applying local sand-
stone materials, acoustic characteristic of cement concrete, which is usually used in
concrete slabs, was proceeded with preliminary study. The research found equation of
manifold mathematical model on concrete compressive strength and velocity of sound.
In view of size effect, preliminary demonstration was processed in laboratory through
specimen of full size.
EXAMPLES

CONCRETE HAS IMPROVED IMMEASURABLY over the years because of research.


While many of the improvements are the result of work in the laboratory, field investiga-
tion on actual roads is the best way to test the effectiveness of new developments. A sum-
mary of recent research follows.

Chen, Dar-Hao, and Moon Won, “Investigation of Settlement of a Jointed


Concrete Pavement.” Journal of Performance of Constructed Facilities 26.3
(December 2009): 440–446.

A section of jointed concrete pavement on U.S. 75, which was built from 1982 to 1985,
in the Paris District of the Texas Department of Transportation (TxDOT) experienced
severe pumping and settlement, even though two (2) types of treatment (full-depth
repair and polyurethane foam injection) were performed. An extensive field investiga-
tion was conducted using ground-penetrating radar, falling weight deflectometer,
dynamic cone penetrometer, and coring to identify the causes of the continued pumping
and settlement problems, and develop an optimal repair strategy. The pavement evalu-
ation included tie bar condition, load transfer efficiency (LTE) at transverse and longi-
tudinal construction joints, and base support conditions. Some of the tie bars failed in
shear due to corrosion, which resulted in substantially low LTEs (< 40%) at longitudinal
construction joints. Pumping and settlement problems were more pronounced where
the tie bars failed; the resulting large deflections exacerbated the pumping and settle-
ment problems. The results demonstrate the importance of adequate LTEs ( > 80%) pro-
vided by tie bars, base and subgrade support, in providing satisfactory JCP performance.
Inadequate design or construction of any of these critical elements could lead to perfor-
mance problems, potentially including severe settlement, which is quite difficult to
repair. To repair this pavement section, the Paris District of TxDOT is planning to ret-
rofit tie bars by the “ slot stitching” method, along with filling the voids under the slab
using grout, followed by thin overlay using latex-modified concrete to correct the differ-
ential elevation problems at longitudinal construction joints. It is expected that this
repair strategy will address the distress problems and extend the pavement life.
204
E X AM P LES 205

Fansong, Meng, Liu Jianping, and Liu Yongzhi, “Design Principles and Frost
Damage Characteristics of the Frozen Soil Roadbed along the Heihe-Bei’an
Highway.” Bingchuan Dongtu, Journal of Glaciology and Geocryology 23.3 (2001):
307–311.

The Heihe-Bei’an Highway is located in sporadic permafrost regions, containing 17 sec-


tions of permafrost, 3.165 km in length. The mean annual ground temperature is high,
and thermal stability is poor. Sporadic permafrost occurs in swampy wetlands with
peat and surface water. Permafrost distribution and development shows obvious differ-
ences in different topographic units under the effect of conversion temperature, vegeta-
tion, soil, surface water, and ground water. Permafrost mainly appears in ditch valleys.
No permafrost appears in high slopes and mountains. Seasonally frozen and thawed
depth is about 1.5 to 2.0 m, with a maximum of 2.5 m. The thickness of permafrost is
about 3 to 6 m, with a maximum of about 9 m. Ground ice occurs in massive, micro-
layered, layered cryostructures near the permafrost table. Problems with frost heaving
and thaw settlement, i.e., frost heaving in seasonally frozen areas and thaw settlement
in permafrost areas, always occur. Frost heaving is extremely strong in swamps, strong
in slope sections, and slight in hill sections. Permafrost conditions and ground ice should
be considered in the design of roadbeds in sporadic permafrost regions in order to ensure
the stability and reliability of the roadbed. Prethawing design principles may be applied
to ice-rich and saturated-ice soil, but monitoring roadbed deformation is necessary.
Temporary road surfaces are used for saturated-ice soil and ice with soil, because the
permafrost is too thick for thawing during construction. Concrete pavement is added
after the permafrost under the roadbed has thawed. Cold permafrost must be protected
to ensure a stable roadbed.

Lee, Eul-Bum, and Kunhee Choi, “Innovative Fast-Track Approach to Concrete


Pavement Rehabilitation in Urban Highway Network.” Airfield and H ighway
Pavement: M eeting Today’ s Challenges with Emerging Technologies (2006):
897–911.

This case study presents an innovative fast-track reconstruction approach as imple-


mented in 2004 on the Interstate 15 (I-15) Devore Urban Highway Reconstruction
Project in Southern California. The “ Rapid Rehab” strategy was chosen for the Devore
Project, in which concrete truck lanes were to be rebuilt along a 4.5-km stretch of I-15
using two (2) nine-day (9-day) extended closures with 24/7 operations. In this Rapid
Rehab strategy, state-of-practice technologies were adopted to accelerate construction,
to mitigate traffic disruptions, and to propagate project information. From the initial
planning and design stages, engineers used CA4PRS (Construction Analysis for Pave-
ment Rehabilitation Strategies) software, incorporated with traffic simulation models,
to arrive at an optimal and economical rehabilitation scenario. The postconstruction
data validated the CA4PRS preconstruction analysis and simulation estimates of con-
struction productivity and traffic delays. As a result of the implementation of an Auto-
E X AM P LES 206

mated Work Z one Information System (AWIS) and proactive public outreach, a 20%
reduction in traffic demand through the Construction Work Z one (CWZ ) was achieved,
thereby reducing the maximum peak-hour delay to the “ tolerable” level. Surveys on the
project Web site showed public perception of the Rapid Rehab approach changed dra-
matically from initial reluctance to future support. Advantages of using this method of
fast-track, accelerated highway reconstruction included less traffic disruption for the
traveling public, longer pavement life expectancy (30-plus years), improved safety for
motorists and workers, and reduced agency cost compared to the traditional approach
using repeated nighttime closures.

Seo, Youngguk, “Distress Evolution in Highway Flexible Pavements: A 5-Year


Study at the Korea Highway Corporation Highway Flexible Test Road.” Journal
of Testing and Evaluation 38.1 (January 2010): 32–41.

In K orea, a 7.7 km-long test road was constructed with an aim to better understand the
behavior of pavements and to develop a mechanistic-empirical pavement design guide
(MEPDG). This 2-lane highway consisted of asphalt and concrete pavements that were
further divided into diverse sections: 33 asphalt sections and 25 concrete sections. Con-
struction of the test road began in 1997 and ended in 2002. It opened to traffic in March
2004. So far, many field tests have been carried out to evaluate the response and perfor-
mance of pavements. At asphalt sections, the first attempt was made to correlate design
parameters with the performance of pavements, and this is the focus of this paper. Per-
formance was characterized by the evolution of failures, such as rutting, cracking, and
longitudinal road profile. Design parameters included surface type, base type, base
thickness, and anti-frost layer. Field surveys were made with an automatic road ana-
lyzer for the 2002–2 007 period, and trends in pavement failures were analyzed with
respect to individual design parameters. Also, a highway present-condition index
(HPCI) was adopted to assess the overall performance of individual sections. Findings
of this study are being used to validate pavement performance models incorporated
into the MEPDG that is scheduled to be released in 2010.
COMPARISON OF CODES
BY JURISDICTION

CONCRETE CONSTRUCTION TECHNIQUES VARY WIDELY from country to coun-


try—sometimes because of local conditions, and sometimes because no uniform code
exists. Studies of different approaches to construction follow.

Beatty, Tommy L., et al., Pavement Preservation Technology in France, South


Africa, and Australia. U.S.D.O .T., Federal Highway Administration, 2002.

Traditionally, highway agencies have allowed the ride quality and structural condition
of their pavements to deteriorate to fair or poor condition before taking steps to rehabil-
itate the pavements. The aim of rehabilitation is to repair structural damage and restore
measurable pavement conditions such as ride, rutting, and cracking. This is a costly and
time-consuming activity with associated traffic disruptions and inconvenience to adja-
cent businesses and residences. In recent years, an increasing number of highway agen-
cies have found that applying a series of low-cost pavement preservation treatments can
extend the service lives of their pavements. This translates into a better investment and
increased customer satisfaction and support. France, South Africa, and Australia were
identified as nations that have innovative programs as well as new treatments for pave-
ment preservation. The French government has made a conscious decision to design
and build extremely durable, long-lasting structural pavement sections (which include
surfacing, base, and subbase materials) on the Motorway System. The South Africans,
like the French, also build robust structural sections with a long service life into their
national network. In the Republic of South Africa, roadway sections are constructed of
cement-treated subbase covered by high-quality crushed-aggregate base course. Austra-
lia uses long-term maintenance contracts (10 years), turning over total control and
responsibility for roadway system maintenance, rehabilitation, and capital improve-
ments to private contractors.

207
COM PARI S ON OF CODE S BY J URI S DI C T ION 208

Beushausen, Hans-Dieter, “Cement and Concrete in South Africa.” Betonwerk


und Fertigteil-Technik/Concrete Precasting Plant and Technology, 6–11, (2002)
12–18+ 20.

International co-operation, characterized by investments and knowledge transfer, is


one of the main engines for worldwide technical progress and development. South Afri-
ca’s international links are mainly based in Europe, and this is where the country origi-
nally obtained a substantial part of its foundation for the building industry. Based on
that foundation, South Africa has developed a well-organized and internationally
acknowledged cement and concrete industry, which produces high-quality products in
the fields of cement manufacture, concrete constructions, and research.

D’Angelo, John, et al., Warm-Mix Asphalt: European Practice. U.S.D.O .T.,


Federal Highway Administration, 2008.

A number of new technologies have been developed to lower the production and place-
ment temperatures of hot-mix asphalt (HMA). Generically, these technologies to reduce
production temperatures are referred to as warm-mix asphalt (WMA). These technolo-
gies originated in Europe in response to a variety of concerns. Beginning in 2002, inter-
est in these technologies has grown in the United States. The American Association of
State Highway and Transportation Officials (AASHTO) and the Federal Highway
Administration (FHWA) organized a scanning tour to Europe, allowing U.S. experts to
meet directly with the agencies who first used these technologies, the suppliers and con-
tractors who developed them, and the contractors who build WMA pavements. The
tour also allowed the team to view the performance of some of the oldest WMA projects
in Europe.

G ibbs, David, et al., Quiet Pavement Systems in Europe. U.S.D.O .T.,


Federal Highway Administration, 2005.

Report summarizes the findings of the quiet-pavements scan team. The team visited
five (5) countries over a 17-day period. The study design was based on a comprehensive
desk scan of published research that summarized where the technology was most used,
where it was first used, and where innovation was still being explored. The objective of
the quiet-pavements scan team was to visit countries with the most experience in quiet-
pavement technology and learn from their experience.

Hall, Kathleen, et al., Long-Life Concrete Pavements in Europe and Canada.


U.S.D.O .T., Federal Highway Administration, 2007.

Long-life concrete pavements require less frequent repair and rehabilitation and contrib-
ute to highway safety and mitigation of congestion. The Federal Highway Administration,
American Association of State Highway and Transportation Officials, and National Coop-
erative Highway Research Program sponsored a scanning study to identify design philoso-
COM PARI S ON OF CODE S BY J URI S DI C T ION 209

phies, materials requirements, construction procedures, and maintenance strategies used


in Europe and Canada to build long-life concrete pavements. The scan team observed that
concrete pavements in the countries visited are designed for 30 or more years of low-main-
tenance service life. The countries are responding to pavement-tire noise issues in urban
areas by using exposed aggregate surface. Some use catalog designs for pavements and
geotextiles as a separator layer between the cement-treated base and concrete pavement.
Team recommendations for U.S. implementation include using two-lift (2-lift) construc-
tion to build pavements, developing pavement design catalogs, using better-quality mate-
rials in pavement subbases, paying greater attention to cement and concrete mixture
properties, using a geotextile interlayer to prevent concrete slabs from bonding to the
cement-treated base, and using exposed aggregate surfaces to reduce noise.

Hall, Kathleen, et al., Nonwoven Geotextile Interlayers for Separating


Cementitious Pavement Layers: German Practice and U. S. Field Trials.
U.S.D.O .T., Federal Highway Administration, 2009.

Pavement engineering is traditionally a conservative field, but successful pavement


engineers will constantly seek out proven innovative concepts with potential to improve
pavement performance while reducing costs. Many pavement structures in the United
States consist of more than one cementitious layer, requiring separation. This could be
a new concrete pavement (jointed or continuously reinforced) atop a cementitious base
or, becoming increasingly popular, an unbonded concrete overlay. In both cases, an
interlayer is often required for separation. While hot-mix asphalt is commonly used for
this purpose, associated constructability, cost, and performance issues need to be recog-
nized. The German highway community has more than 25 years of experience using an
alternative interlayer made of a nonwoven geotextile. With proper selection and place-
ment, these interlayers have resulted in excellent performance in separating new con-
crete pavement layers from the cementitious bases commonplace on the German
motorway system. While this application also exists in the United States, the possibility
of using the nonwoven geotextile as a separation interlayer in an unbonded concrete
overlay system has also generated significant interest because of the potential cost sav-
ings involved. This report documents the German experience and gleans better prac-
tices for using nonwoven geotextile interlayers between cementitious layers in the
United States. The report covers field trials in Missouri and Oklahoma, along with
other information, so that pavement engineers can make an informed decision on the
viability of this innovative but proven alternative.

Ingram, Lon S., et al., Superior Materials, Advanced Test Methods, and
Specifications in Europe. U.S.D.O .T., Federal Highway Administration, 2004.

Report of a U.S. transportation delegation scanning study of Europe (United Kingdom,


Denmark, Germany, and The Netherlands) to review their approaches to introducing,
COM PARI S ON OF CODE S BY J URI S DI C T ION 210

approving, and specifying new products for highway construction. Included in the
report is guidance on noise issues in pavement design, construction, and maintenance.

Shilstone, James M., “Worldwide Standards for Concrete Mixtures


for Safe and Durable Runways and Taxiways.” American Society of
Civil Engineers. ASCE Conf. Proc. 262–21 (2007).

The technology that led to optimized concrete paving mixtures was born of a need to
develop a method to specify and control the quality of concrete mixtures for construc-
tion of a $600 million office building in Riyadh, Saudi Arabia, in the early 1970s. How-
ever, there were no materials standards in the kingdom then. It was necessary for this
author to learn how to achieve that goal using the non-standard, locally available mate-
rials. The research was done in a laboratory in Athens, Greece. The results included the
introduction of a nomograph on which six (6) independent variables can be plotted and
used to assess concrete mix performance. That technology is being applied today for
construction of buildings, highways, and airfields all over the world. The first verifica-
tion of the technology for airfield construction was done at the Jeddah International
Airport in Saudi Arabia.

X in-bo, Ma, and Su-mei Z hang, “Comparison of Design Methods of Load-


Carrying Capacity for Circular Concrete-Filled Steel Tube Beam Columns
in Typical Codes Worldwide,” Journal of Harbin Institute of Technology 39–4
(April 2007): 536–41.

The design methods and calculated results of load-carrying capacity for circular con-
crete-filled steel tube (CFST) beam columns in seven (7) different nodes worldwide are
compared in detail. These include American Institute of Steel Construction (AISC) –
Load and Resistance Factor Design specification for structural steel buildings (LRFD);
European Committee for Standardization (CEN) Eurocode 4-Design of composite steel
and concrete structures (EC4-94); Architectural Institute of Japan (AIJ) – Recommen-
dations for design and construction of concrete-filled steel tubular structures; and the
Chinese codes JCJ01-89, CECS code, DL/T 5085-1999, and DBJ13-5-2003. The compari-
sons between the results calculated by each code and 285 experimental results collected
from worldwide references, both excluding and including the limitations prescribed by
each code, are carried out and discussed.
CURRENT
CRITERIA

WORLDWIDE, SEVERAL ORGANIZATIONS have set standards for the production of


concrete. These standards cover everything from aggregate quality to slump testing and
strength. The purpose of these standards is to assure that concrete going into a job is ade-
quate, and that it is comparable to concrete going into similar jobs. This chapter includes
concrete-related criteria from four (4) major authorities: AASHTO, ACI, ASTM, and the
European Union. Copies of the standards may be purchased from the authorities.

Ame r ic a n Asso c ia t io n o f St a t e H i g h w a y
a n d Tr a n ps o r t a t oi n Offi c ia ls (A ASH TO)
Sp e c ifi c a t io n s , Co d e s , St a n d a r d s , a n d G u id e lin e s
AASHTO M 6 Standard Specification for Fine Aggregate for Portland Cement Concrete
AASHTO M 43 Standard Specification for Sizes of Aggregate for Road and Bridge
Construction – ASTM Designation: D 448-03a
AASHTO M 80 Standard Specification for Coarse Aggregate for Portland Cement Concrete
AASHTO M 85 Standard Specification for Portland Cement – ASTM Designation: C
150-04a
AASHTO M 155 Standard Specification for Granular Material to Control Pumping under
Concrete Pavement
AASHTO M 157 Standard Specification for Ready-Mixed Concrete
AASHTO M 233 Standard Specification for Boiled Linseed Oil Mixture for Treatment of
Portland Cement Concrete
AASHTO M 240 Standard Specification for Blended Hydraulic Cement – ASTM Designation:
C 595-03
AASHTO M 241 Standard Specification for Concrete Made by Volumetric Batching and
Continuous Mixing – ASTM Designation: C 685-95a; Replaced by AASHTO
M 241M/M 241-05
AASHTO M 295 Standard Specification for Coal Fly Ash and Raw or Calcined Natural
Pozzolan for Use in Concrete – ASTM Designation: C 618-03
AASHTO M 321 Standard Specification for High-Reactivity Pozzolans for Use in Hydraulic-
Cement Concrete, Mortar, and Grout
AASHTO T 11 Standard Method of Test for Materials Finer than 75-µm (No. 200) Sieve in
Mineral Aggregates by Washing – ASTM Designation: C 117-03

211
CURRE NT CRI T ER IA 212

AASH TO Sp e c i fi c a t i o n s , Co d e s , St a n d a r d s , a n d G u i d e l i n e s— c o n t in u e d

AASHTO T 19M/T 19 Standard Method of Test for Bulk Density (Unit Weight) and Voids in
Aggregate – ASTM Designation: C 29/C 29M-97 (2003)
AASHTO T 22 Standard Method of Test for Compressive Strength of Cylindrical Concrete
Specimens – ASTM Designation: C 39-04a
AASHTO T 23 Standard Method of Test for Making and Curing Concrete Test Specimens
in the Field – ASTM Designation: C 31-03a
AASHTO T 26 Standard Method of Test for Quality of Water to Be Used in Concrete
AASHTO T 27 Standard Method of Test for Sieve Analysis of Fine and Coarse Aggregates –
ASTM Designation: C 136-05
AASHTO T 84 Standard Method of Test for Specific Gravity and Absorption of Fine
Aggregate – ASTM Designation: C 128-97
AASHTO T 85 Standard Method of Test for Specific Gravity and Absorption of Coarse
Aggregate – ASTM Designation: C 127-88 (1993)
AASHTO T 97 Standard Method of Test for Flexural Strength of Concrete (Using Simple
Beam with Third-Point Loading) – ASTM Designation: C 78-02
AASHTO T 98 Standard Method of Test for Fineness of Portland Cement by the
Turbidimeter
AASHTO T 103 Standard Method of Test for Soundness of Aggregates by Freezing and
Thawing
AASHTO T 105 Standard Method of Test for Chemical Analysis of Hydraulic Cement –
ASTM Designation: C 114-05
AASHTO T 106M/T 106 Standard Method of Test for Compressive Strength of Hydraulic Cement
Mortar (Using 50-mm or 2-in. Cube Specimens) – ASTM Designation: C
109/C 109M-02
AASHTO T 119M/T 119 Standard Method of Test for Slump of Hydraulic Cement Concrete – ASTM
Designation: C 143/C 143M-03
AASHTO T 121M/T 121 Standard Method of Test for Density (Unit Weight), Yield, and Air Content
(Gravimetric) of Concrete – ASTM Designation: C 138/C 138M-01a
AASHTO T 126 Standard Method of Test for Making and Curing Concrete Test Specimens
in the Laboratory-HM-22; Part IIA – ASTM Designation: C 192/C 192M-95
AASHTO T 127 Standard Method of Test for Sampling and Amount of Testing of Hydraulic
Cement – ASTM Designation: C 183-02
AASHTO T 129 Standard Method of Test for Normal Consistency of Hydraulic Cement –
ASTM Designation: C 187-04
AASHTO T 131 Standard Method of Test for Time of Setting of Hydraulic Cement by Vicat
Needle – ASTM Designation: C 191-04a
AASHTO T 132 Standard Method of Test for Tensile Strength of Hydraulic Cement Mortars
AASHTO T 133 Standard Specification of Test for Density of Hydraulic Cement – ASTM
Designation: C 188-95
AASHTO T 137 Standard Method of Test for Air Content of Hydraulic Cement Mortar –
ASTM Designation: C 185-02
AASHTO T 140 Standard Method of Test for Compressive Strength of Concrete Using
Portions of Beams Broken in Flexure
AASHTO T 141 Standard Method of Test for Sampling Freshly Mixed Concrete – ASTM
Designation: C 172-04
AASHTO T 148 Standard Method of Test for Measuring Length of Drilled Concrete Cores –
ASTM Designation: C 174-87
AASHTO T 152 Standard Method of Test for Air Content of Freshly Mixed Concrete by the
Pressure Method – ASTM Designation: C 231-03
AASHTO T 154 Standard Method of Test for Time of Setting of Hydraulic Cement Paste by
Gillmore Needles – ASTM Designation: C 266-04
CURRE NT CRI T ER IA 213

AASH TO Sp e c i fi c a t i o n s , Co d e s , St a n d a r d s , a n d G u i d e l i n e s— c o n t in u e d

AASHTO T 155 Standard Method of Test for Water Retention by Concrete Curing Materials
– ASTM Designation: C 156-03
AASHTO T 157 Standard Method of Test for Air-Entraining Admixtures for Concrete –
ASTM Designation: C 233-04
AASHTO T 158 Standard Method of Test for Bleeding of Concrete – ASTM Designation: C
232-99
AASHTO T 160 Standard Method of Test for Length Change of Hardened Hydraulic Cement
Mortar and Concrete – ASTM Designation: C 157-93
AASHTO T 161 Standard Method of Test for Resistance of Concrete to Rapid Freezing and
Thawing – ASTM Designation: C 666/C 666M-03
AASHTO T 177 Standard Method of Test for Flexural Strength of Concrete (Using Simple
Beam with Center-Point Loading) – ASTM Designation: C 293-02
AASHTO T 178 Standard Method of Test for Cement Content of Hardened Portland Cement
Concrete – ASTM Designation: C 1084-92
AASHTO T 179 Standard Method of Test for Effect of Heat and Air on Asphalt Materials
(Thin-Film Oven Test) – ASTM Designation: D 1754-97 (2002)
AASHTO T 185 Standard Method of Test for Early Stiffening of Hydraulic Cement (Mortar
Method) – ASTM Designation: C 359-04
AASHTO T 186 Standard Method of Test for Early Stiffening of Hydraulic Cement (Paste
Method) – ASTM Designation: C 451-04a
AASHTO T 188 Standard Method of Test for Evaluation by Freezing and Thawing of Air-
Entraining Additions to Hydraulic Cement
AASHTO T 192 Standard Method of Test for Fineness of Hydraulic Cement by the 45-µm
(No. 325) Sieve – ASTM Designation: C 430-96
AASHTO T 195 Standard Method of Test for Determining Degree of Particle Coating of
Bituminous-Aggregate Mixtures – ASTM Designation: D 2489-67(1980)
AASHTO T 197M/T 197 Standard Method of Test for Time of Setting of Concrete Mixtures by
Penetration Resistance – ASTM Designation: C 403/C 403M-99
AASHTO T 198 Standard Method of Test for Splitting Tensile Strength of Cylindrical
Concrete Specimens-HM-22; Part IIA – ASTM Designation: C 496-96
AASHTO T 231 Standard Method of Test for Capping Cylindrical Concrete Specimens –
ASTM Designation: C 617-98 (2003)
AASHTO T 259 Standard Method of Test for Resistance of Concrete to Chloride Ion
Penetration-HM-22; PART IIB
AASHTO T 260 Standard Method of Test for Sampling and Testing for Chloride Ion in
Concrete and Concrete Raw Materials
AASHTO T 277 Standard Method of Test for Electrical Indication of Concrete’s Ability to
Resist Chloride Ion Penetration – ASTM Designation: C 1202-94
AASHTO T 309M/T 309 Standard Method of Test for Temperature of Freshly Mixed Hydraulic-
Cement Concrete – ASTM Designation: C 1064/C 1064M-05
AASHTO T 318 Standard Method of Test for Water Content of Freshly Mixed Concrete
Using Microwave Oven Drying

Ame r ic a n Co n c r e t e In ts it u t e A
( CI)
Sp e c fii c a t io n s , Co d e s , St a n d a r d s , a n d G u id e lin e s
ACI 104 Preparation of Notation for Concrete
ACI 116R Cement and Concrete Terminology
ACI 117 Specifications for Tolerances for Concrete Construction and Materials and
Commentary
ACI 121R Quality Management System for Concrete Construction
CURRE NT CRI T ER IA 214

ACI Sp e c ifi c a t io n s , Co d e s , St a n d a r d s , a n d G u id e lin e s — c o n t in u e d

ACI 122R Guide to Thermal Properties of Concrete and Masonry Systems


ACI 126.3R Guide to Recommended Format for Concrete in Materials Property Database
ACI 201.1R Guide for Making a Condition Survey of Concrete in Service
ACI 201.2R Guide to Durable Concrete
ACI 207.1R Guide to Mass Concrete
ACI 207.2R Effect of Restraint, Volume Change, and Reinforcement on Cracking of Mass Concrete
ACI 207.3R Practices for Evaluation of Concrete in Existing Massive Structures
ACI 207.4R Cooling and Insulating Systems for Mass Concrete
ACI 207.5R Roller-Compacted Mass Concrete
ACI 209.1R Report on Factors Affecting Shrinkage and Creep of Hardened Concrete
ACI 209R Prediction of Creep, Shrinkage, and Temperature Effects in Concrete Structures
ACI 210.1R Compendium of Case Histories on Repair of Erosion-Damaged Concrete in Hydraulic
Structures
ACI 210R Erosion of Concrete in Hydraulic Structures
ACI 211.1 Standard Practice for Selecting Proportions for Normal, Heavyweight, and Mass
Concrete
ACI 211.2 Standard Practice for Selecting Proportions for Structural Lightweight Concrete
ACI 211.3R Guide for Selecting Proportions for No-Slump Concrete
ACI 211.4R Guide for Selecting Proportions for High-Strength Concrete with Portland Cement and
Fly Ash – Errata
ACI 211.5R Guide for Submittal of Concrete Proportions
ACI 212.3R Chemical Admixtures for Concrete
ACI 212.4R Guide for the Use of High-Range Water-Reducing Admixtures (Superplasticizers) in
Concrete
ACI 213R Guide for Structural Lightweight-Aggregate Concrete
ACI 214.3R Simplified Version of the Recommended Practice for Evaluation of Strength Test
Results of Concrete
ACI 214.4R Guide for Obtaining Cores and Interpreting Compressive Strength Results
ACI 214R Evaluation of Strength Test Results of Concrete
ACI 215R Considerations for Design of Concrete Structures Subjected to Fatigue Loading
ACI 216R Guide for Determining the Fire Endurance of Concrete Elements
ACI 221.1R State-of-the-Art Report on Alkali-Aggregate Reactivity
ACI 221R Guide for Use of Normal-Weight and Heavyweight Aggregates in Concrete
ACI 222.3R Design and Construction Practices to Mitigate Corrosion of Reinforcement in Concrete
Structures
ACI 222R Protection of Metals in Concrete against Corrosion
ACI 223 Standard Practice for the Use of Shrinkage-Compensating Concrete
ACI 224.1R Causes, Evaluation, and Repair of Cracks in Concrete Structures
ACI 224.2R Cracking of Concrete Members in Direct Tension
ACI 224.3R Joints in Concrete Construction
ACI 224R Control of Cracking in Concrete Structures
ACI 225R Guide to the Selection and Use of Hydraulic Cements
ACI 228.1R In-Place Methods to Estimate Concrete Strength
ACI 228.2R Nondestructive Test Methods for Evaluation of Concrete in Structures
ACI 229R Controlled Low-Strength Materials
ACI 232.1R Use of Raw or Processed Natural Pozzolans in Concrete
ACI 232.2R Use of Fly Ash in Concrete
CURRE NT CRI T ER IA 215

ACI Sp e c ifi c a t io n s , Co d e s , St a n d a r d s , a n d G u id e lin e s — c o n t in u e d

ACI 233R Slag Cement in Concrete and Mortar


ACI 234R Guide for the Use of Silica Fume in Concrete
ACI 237R Self-Consolidating Concrete
ACI 301 Specifications for Structural Concrete
ACI 301M Specifications for Structural Concrete
ACI 302.1R (ERTA) Guide for Concrete Floor and Slab Construction
ACI 302.2R Guide for Concrete Slabs that Receive Moisture-Sensitive Flooring Materials
ACI 303.1 Standard Specifications for Cast-in-Place Architectural Concrete
ACI 303R Guide to Cast-in-Place Architectural Concrete Practice
ACI 304.1R Guide for the Use of Preplaced Aggregate Concrete for Structural and Mass
Applications
ACI 304.2R Placing Concrete by Pumping Methods
ACI 304.3R Heavyweight Concrete: Measuring, Mixing, Transporting, and Placing
ACI 304.4R Placing Concrete with Belt Conveyors
ACI 304.6R Guide for the Use of Volumetric-Measuring and Continuous-Mixing Concrete
Equipment
ACI 304R Guide for Measuring, Mixing, Transporting, and Placing Concrete
ACI 305.1 Specifications for Hot-Weather Concreting
ACI 305R Hot-Weather Concreting – Incorporating Errata: 6/15/2006
ACI 306.1 Standard Specifications for Cold-Weather Concreting
ACI 306R Cold-Weather Concreting
ACI 308.1 Standard Specifications for Curing Concrete
ACI 308R Guide to Curing Concrete
ACI 309.1R Behavior of Fresh Concrete During Vibration
ACI 309.2R Identification and Control of Visible Effects of Consolidation on Formed Concrete
Surfaces
ACI 309.5R Compaction of Roller-Compacted Concrete
ACI 309R Guide for Consolidation of Concrete
ACI 311.4R Guide for Concrete Inspection
ACI 325.10R State-of-the-Art Report on Roller-Compacted Concrete Pavements
ACI 325.11R Accelerated Techniques for Concrete Paving
ACI 325.12R Guide for Design of Jointed Concrete Pavements for Streets and Local Roads
ACI 325.13R Concrete Overlays for Pavement Rehabilitation
ACI 325.3R Guide for Design of Foundations and Shoulders for Concrete Pavements
ACI 325.5R Design of Continuously Reinforced Concrete Pavement for Airports
ACI 325.6R Texturing Concrete Pavements
ACI 325.7R Recommendations for Designing Prestressed Concrete Pavements
ACI 325.9R Guide for Construction of Concrete Pavements and Concrete Bases
ACI 336.2R Suggested Analysis and Design Procedures for Combined Footings and Mats
ACI 336.3R Design and Construction of Drilled Piers
ACI 341.2R Seismic Analysis and Design of Concrete Bridge Systems
ACI 341.3R Seismic Evaluation and Retrofit Techniques for Concrete Bridges
ACI 343R Analysis and Design of Reinforced Concrete Bridge Structures
ACI 345.1R Guide for Maintenance of Concrete Bridge Members
ACI 345.2R Guide for Widening Highway Bridges
ACI 345R Guide for Concrete Highway Bridge Deck Construction
CURRE NT CRI T ER IA 216

ACI Sp e c ifi c a t io n s , Co d e s , St a n d a r d s , a n d G u id e lin e s — c o n t in u e d

ACI 346 Specifications for Cast-in-Place Concrete Pipe


ACI 347 Guide to Formwork for Concrete
ACI 350.2R Concrete Structures for Containment of Hazardous Materials
ACI 350.3 Seismic Design of Liquid-Containing Concrete Structures and Commentary
ACI 350.4R Design Considerations for Environmental Engineering Concrete Structures
ACI 352.1R Recommendations for Design of Slab-Column Connections in Monolithic Reinforced
Concrete Structures
ACI 358.1R Analysis and Design of Reinforced and Prestressed-Concrete Guideway Structures
ACI 360R Design of Slabs-on-Ground
ACI 363.2R Guide to Quality Control and Testing of High-Strength Concrete
ACI 363R State-of-the-Art Report on High-Strength Concrete
ACI 364.1R Guide for Evaluation of Concrete Structures Prior to Rehabilitation
ACI 365.1R Service-Life Prediction – State-of-the-Art Report
ACI 440R State-of-the-Art Report on Fiber-Reinforced Plastic Reinforcement for Concrete
Structures
ACI 522R Pervious Concrete
ACI 523.1R Guide for Cast-in-Place Low-Density Concrete
ACI 544.1R Fiber-Reinforced Concrete
ACI 544.2R Measurement of Properties of Fiber-Reinforced Concrete
ACI 544.3R Guide for Specifying, Proportioning, Mixing, Placing, and Finishing Steel–Fiber
Reinforced Concrete
ACI 544.4R Design Considerations for Steel Fiber–Reinforced Concrete
ACI 546.1R Guide for Repair of Concrete Bridge Superstructures
ACI 546.3R Guide for the Selection of Materials for the Repair of Concrete
ACI 546R Concrete Repair Guide
ACI 547R Refractory Concrete: Abstract of State-of-the-Art Report (Revised 1983)
ACI 548.1R Guide for the Use of Polymers in Concrete
ACI 548.3R Polymer-Modified Concrete
ACI 548.4 Standard Specification for Latex-Modified Concrete (LMC) Overlays
ACI 548.5R Guide for Polymer Concrete Overlays
ACI 555R Removal and Reuse of Hardened Concrete
ACI CCS-0 Concrete Fundamentals
ACI CCS-1 Slabs on Grade-Second Edition
ACI PC Pumping Concrete – Techniques and Applications

Ame r ic a n So c ei t y of r Te st in g a n d Ma t e r ia ls (A STM)
H ig h w a y St a n d a r d s

C0029_C0029M-07 Test Method for Bulk Density (“Unit Weight”) and Voids in Aggregate
C0031_C0031M-09 Practice for Making and Curing Concrete Test Specimens in the Field
C0033_C0033M-08 Specification for Concrete Aggregates
C0039_C0039M-05E02 Test Method for Compressive Strength of Cylindrical Concrete
Specimens
C0040-04 Test Method for Organic Impurities in Fine Aggregates for Concrete
C0042_C0042M-04 Test Method for Obtaining and Testing Drilled Cores and Sawed Beams
of Concrete
CURRE NT CRI T ER IA 217

ASTM H ig h w a y St a n d a r d s— c o n t in u e d

C0070-06 Test Method for Surface Moisture in Fine Aggregate


C0078-08 Test Method for Flexural Strength of Concrete (Using Simple Beam with
Third-Point Loading)
C0087-05 Test Method for Effect of Organic Impurities in Fine Aggregate on
Strength of Mortar
C0088-05 Test Method for Soundness of Aggregates by Use of Sodium Sulfate or
Magnesium Sulfate
C0094_C0094M-09 Specification for Ready-Mixed Concrete
C0117-04 Test Method for Materials Finer than 75-µm (No. 200) Sieve in Mineral
Aggregates by Washing
C0123-04 Test Method for Lightweight Particles in Aggregate
C0125-07 Terminology Relating to Concrete and Concrete Aggregates
C0127-07 Test Method for Density, Relative Density (Specific Gravity), and
Absorption of Coarse Aggregate
C0128-07A Test Method for Density, Relative Density (Specific Gravity), and
Absorption of Fine Aggregate
C0131-06 Test Method for Resistance to Degradation of Small-Size Coarse
Aggregate by Abrasion and Impact in the Los Angeles Machine
C0136-06 Test Method for Sieve Analysis of Fine and Coarse Aggregates
C0138_C0138M-09 Test Method for Density (Unit Weight), Yield, and Air Content (Gravi-
metric) of Concrete
C0142-97R04 Test Method for Clay Lumps and Friable Particles in Aggregates
C0143_C0143M-08 Test Method for Slump of Hydraulic-Cement Concrete
C0156-05 Test Method for Water Retention by Liquid Membrane-Forming Curing
Compounds for Concrete
C0157_C0157M-08 Test Method for Length Change of Hardened Hydraulic-Cement Mortar
and Concrete
C0171-07 Specification for Sheet Materials for Curing Concrete
C0172-08 Practice for Sampling Freshly Mixed Concrete
C0173_C0173M-09 Test Method for Air Content of Freshly Mixed Concrete by the Volumet-
ric Method
C0174_C0174M-06 Test Method for Measuring Thickness of Concrete Elements Using
Drilled Concrete Cores
C0192_C0192M-07 Practice for Making and Curing Concrete Test Specimens in the
Laboratory
C0215-08 Test Method for Fundamental Transverse, Longitudinal, and Torsional
Resonant Frequencies of Concrete Specimens
C0227-03 Test Method for Potential Alkali Reactivity of Cement-Aggregate
Combinations (Mortar-Bar Method)
C0231-09 Test Method for Air Content of Freshly Mixed Concrete by the Pressure
Method
C0232_C0232M-09 Test Methods for Bleeding of Concrete
C0233-07 Test Method for Air-Entraining Admixtures for Concrete
C0260-06 Specification for Air-Entraining Admixtures for Concrete
C0289-07 Test Method for Potential Alkali-Silica Reactivity of Aggregates
(Chemical Method)
C0293-08 Test Method for Flexural Strength of Concrete (Using Simple Beam
With Center-Point Loading)
C0294-05 Descriptive Nomenclature for Constituents of Concrete Aggregates
C0295-08 Guide for Petrographic Examination of Aggregates for Concrete
CURRE NT CRI T ER IA 218

ASTM H ig h w a y St a n d a r d s— c o n t in u e d

C0309-07 Specification for Liquid Membrane-Forming Compounds for Curing


Concrete
C0311-07 Test Methods for Sampling and Testing Fly Ash or Natural Pozzolans for
Use in Portland-Cement Concrete
C0330-05 Specification for Lightweight Aggregates for Structural Concrete
C0331-05 Specification for Lightweight Aggregates for Concrete Masonry Units
C0332-07 Specification for Lightweight Aggregates for Insulating Concrete
C0341_C0341M-06 Practice for Length Change of Cast, Drilled, or Sawed Specimens of
Hydraulic-Cement Mortar and Concrete
C0387_C0387M-09 Specification for Packaged, Dry, Combined Materials for Mortar and
Concrete
C0403_C0403M-08 Test Method for Time of Setting of Concrete Mixtures by Penetration
Resistance
C0418-05 Test Method for Abrasion Resistance of Concrete by Sandblasting
C0441-05 Test Method for Effectiveness of Pozzolans or Ground Blast-Furnace
Slag in Preventing Excessive Expansion of Concrete Due to the Alkali-
Silica Reaction
C0457-08D Test Method for Microscopical Determination of Parameters of the Air-
Void System in Hardened Concrete
C0469-02E01 Test Method for Static Modulus of Elasticity and Poisson’s Ratio of
Concrete in Compression
C0470_C0470M-08 Specification for Molds for Forming Concrete Test Cylinders Vertically
C0490_C0490M-08 Practice for Use of Apparatus for the Determination of Length Change
of Hardened Cement Paste, Mortar, and Concrete
C0494_C0494M-08A Specification for Chemical Admixtures for Concrete
C0495-07 Test Method for Compressive Strength of Lightweight Insulating
Concrete
C0496_C0496M-04E01 Test Method for Splitting Tensile Strength of Cylindrical Concrete
Specimens
C0511-09 Specification for Mixing Rooms, Moist Cabinets, Moist Rooms, and
Water Storage Tanks Used in the Testing of Hydraulic Cements and
Concretes
C0512-02 Test Method for Creep of Concrete in Compression
C0535-03E01 Test Method for Resistance to Degradation of Large-Size Coarse
Aggregate by Abrasion and Impact in the Los Angeles Machine
C0566-97R04 Test Method for Total Evaporable Moisture Content of Aggregate by
Drying
C0567-05A Test Method for Determining Density of Structural Lightweight Concrete
C0586-05 Test Method for Potential Alkali Reactivity of Carbonate Rocks as
Concrete Aggregates (Rock-Cylinder Method)
C0597-02 Test Method for Pulse Velocity through Concrete
C0617-09 Practice for Capping Cylindrical Concrete Specimens
C0618-08A Specification for Coal Fly Ash and Raw or Calcined Natural Pozzolan for
Use in Concrete
C0637-98AR03 Specification for Aggregates for Radiation-Shielding Concrete
C0638-92R02 Descriptive Nomenclature of Constituents of Aggregates for Radiation-
Shielding Concrete
C0641-07 Test Method for Iron Staining Materials in Lightweight Concrete
Aggregates
C0642-06 Test Method for Density, Absorption, and Voids in Hardened Concrete
CURRE NT CRI T ER IA 219

ASTM H ig h w a y St a n d a r d s— c o n t in u e d

C0666_C0666M-03R08 Test Method for Resistance of Concrete to Rapid Freezing and Thawing
C0670-03 Practice for Preparing Precision and Bias Statements for Test Methods
for Construction Materials
C0672_C0672M-03 Test Method for Scaling Resistance of Concrete Surfaces Exposed to
De-icing Chemicals
C0684-99R03 Test Method for Making, Accelerated Curing, and Testing Concrete
Compression Test Specimens
C0685_C0685M-07 Specification for Concrete Made by Volumetric Batching and Continu-
ous Mixing
C0702-98R03 Practice for Reducing Samples of Aggregate to Testing Size
C0779_C0779M-05 Test Method for Abrasion Resistance of Horizontal Concrete Surfaces
C0796-04 Test Method for Foaming Agents for Use in Producing Cellular Concrete
Using Preformed Foam
C0802-96R08E01 Practice for Conducting an Interlaboratory Test Program to Determine
the Precision of Test Methods for Construction Materials
C0803_C0803M-03 Test Method for Penetration Resistance of Hardened Concrete
C0805_C0805M-08 Test Method for Rebound Number of Hardened Concrete
C0823_C0823M-07 Practice for Examination and Sampling of Hardened Concrete in
Constructions
C0827-01AR05 Test Method for Change in Height at Early Ages of Cylindrical Speci-
mens of Cementitious Mixtures
C0856-04 Practice for Petrographic Examination of Hardened Concrete
C0869-91R06 Specification for Foaming Agents Used in Making Preformed Foam for
Cellular Concrete
C0873_C0873M-04E01 Test Method for Compressive Strength of Concrete Cylinders Cast in
Place in Cylindrical Molds
C0876-09 Test Method for Corrosion Potentials of Uncoated Reinforcing Steel in
Concrete
C0878_C0878M-09 Test Method for Restrained Expansion of Shrinkage-Compensating
Concrete
C0881_C0881M-02 Specification for Epoxy-Resin-Base Bonding Systems for Concrete
C0882_C0882M-05E01 Test Method for Bond Strength of Epoxy-Resin Systems Used With
Concrete by Slant Shear
C0884_C0884M-98R05 Test Method for Thermal Compatibility between Concrete and an Epoxy-
Resin Overlay
C0900-06 Test Method for Pullout Strength of Hardened Concrete
C0918_C0918M-07 Test Method for Measuring Early-Age Compressive Strength and
Projecting Later-Age Strength
C0928_C0928M-08 Specification for Packaged, Dry, Rapid-Hardening Cementitious
Materials for Concrete Repairs
C0937-02 Specification for Grout Fluidifier for Preplaced-Aggregate Concrete
C0938-02 Practice for Proportioning Grout Mixtures for Preplaced-Aggregate
Concrete
C0939-02 Test Method for Flow of Grout for Preplaced-Aggregate Concrete (Flow
Cone Method)
C0940-98AR03 Test Method for Expansion and Bleeding of Freshly Mixed Grouts for
Preplaced-Aggregate Concrete in the Laboratory
C0941-02 Test Method for Water Retentivity of Grout Mixtures for Preplaced-
Aggregate Concrete in the Laboratory
CURRE NT CRI T ER IA 220

ASTM H ig h w a y St a n d a r d s— c o n t in u e d

C0942-99R04 Test Method for Compressive Strength of Grouts for Preplaced-


Aggregate Concrete in the Laboratory
C0943-02 Practice for Making Test Cylinders and Prisms for Determining Strength
and Density of Preplaced-Aggregate Concrete in the Laboratory
C0944_C0944M-99R05E01 Test Method for Abrasion Resistance of Concrete or Mortar Surfaces by
the Rotating-Cutter Method
C0953-06 Test Method for Time of Setting of Grouts for Preplaced-Aggregate
Concrete in the Laboratory
C0979-05 Specification for Pigments for Integrally Colored Concrete
C0989-09 Specification for Slag Cement for Use in Concrete and Mortars
C1017_C1017M-07 Specification for Chemical Admixtures for Use in Producing Flowing
Concrete
C1040_C1040M-08 Test Methods for In-Place Density of Unhardened and Hardened
Concrete, Including Roller-Compacted Concrete, by Nuclear Methods
C1059_C1059M-99R08 Specification for Latex Agents for Bonding Fresh to Hardened Concrete
C1064_C1064M-08 Test Method for Temperature of Freshly Mixed Hydraulic-Cement
Concrete
C1067-00R07 Practice for Conducting a Ruggedness or Screening Program for Test
Methods for Construction Materials
C1073-97AR03 Test Method for Hydraulic Activity of Ground Slag by Reaction with
Alkali
C1074-04 Practice for Estimating Concrete Strength by the Maturity Method
C1077-09 Practice for Laboratories Testing Concrete and Concrete Aggregates for
Use in Construction and Criteria for Laboratory Evaluation
C1084-02 Test Method for Portland-Cement Content of Hardened Hydraulic-
Cement Concrete
C1090-01R05E01 Test Method for Measuring Changes in Height of Cylindrical Specimens
of Hydraulic-Cement Grout
C1105-08A Test Method for Length Change of Concrete Due to Alkali-Carbonate
Rock Reaction
C1107_C1107M-08 Specification for Packaged Dry, Hydraulic-Cement Grout (Nonshrink)
C1116_C1116M-09 Specification for Fiber-Reinforced Concrete
C1137-05 Test Method for Degradation of Fine Aggregate Due to Attrition
C1138M-05 Test Method for Abrasion Resistance of Concrete (Underwater Method)
C1140-03A Practice for Preparing and Testing Specimens from Shotcrete Test
Panels
C1141_C1141M-08 Specification for Admixtures for Shotcrete
C1152_C1152M-04E01 Test Method for Acid-Soluble Chloride in Mortar and Concrete
C1170_C1170M-08 Test Method for Determining Consistency and Density of Roller-
Compacted Concrete Using a Vibrating Table
C1176_C1176M-08 Practice for Making Roller-Compacted Concrete in Cylinder Molds
Using a Vibrating Table
C1202-09 Test Method for Electrical Indication of Concrete’s Ability to Resist
Chloride Ion Penetration
C1218_C1218M-99R08 Test Method for Water-Soluble Chloride in Mortar and Concrete
C1231_C1231M-09 Practice for Use of Unbonded Caps in Determination of Compressive
Strength of Hardened Concrete Cylinders
C1240-05 Specification for Silica Fume Used in Cementitious Mixtures
C1245_C1245M-06 Test Method for Determining Bond Strength between Hardened Roller-
Compacted Concrete and Other Hardened Cementitious Mixtures
(Point-Load Test)
CURRE NT CRI T ER IA 221

ASTM H ig h w a y St a n d a r d s— c o n t in u e d

C1252-06 Test Methods for Uncompacted Void Content of Fine Aggregate (as
Influenced by Particle Shape, Surface Texture, and Grading)
C1260-07 Test Method for Potential Alkali Reactivity of Aggregates (Mortar-Bar
Method)
C1293-08B Test Method for Determination of Length Change of Concrete Due to
Alkali-Silica Reaction
C1315-08 Specification for Liquid Membrane-Forming Compounds Having Special
Properties for Curing and Sealing Concrete
C1362-09 Test Method for Flow of Freshly Mixed Hydraulic-Cement Concrete
C1383-04 Test Method for Measuring the P-Wave Speed and the Thickness of
Concrete Plates Using the Impact-Echo Method
C1385_C1385M-98R04E01 Practice for Sampling Materials for Shotcrete
C1398-07 Test Method for the Laboratory Determination of the Time of Setting of
Hydraulic-Cement Mortars Containing Additives for Shotcrete by the
Use of Gillmore Needles
C1399-07A Test Method for Obtaining Average Residual-Strength of Fiber-Rein-
forced Concrete
C1404_C1404M-98R03 Test Method for Bond Strength of Adhesive Systems Used With
Concrete as Measured by Direct Tension
C1435_C1435M-08 Practice for Molding Roller-Compacted Concrete in Cylinder Molds
Using a Vibrating Hammer
C1436-08 Specification for Materials for Shotcrete
C1438-99R05E01 Specification for Latex and Powder Polymer Modifiers for Hydraulic
Cement Concrete and Mortar
C1439-08A Test Methods for Evaluating Polymer Modifiers in Mortar and Concrete
C1451-05 Practice for Determining Uniformity of Ingredients of Concrete from a
Single Source
C1480_C1480M-07 Specification for Packaged, Pre-Blended, Dry, Combined Materials for
Use in Wet or Dry Shotcrete Application
C1524-02A Test Method for Water-Extractable Chloride in Aggregate (Soxhlet
Method)
C1542_C1542M-02 Test Method for Measuring Length of Concrete Cores
C1543-02 Test Method for Determining the Penetration of Chloride Ion into
Concrete by Ponding
C1550-08 Test Method for Flexural Toughness of Fiber-Reinforced Concrete (Using
Centrally Loaded Round Panel)
C1556-04 Test Method for Determining the Apparent Chloride Diffusion Coeffi-
cient of Cementitious Mixtures by Bulk Diffusion
C1567-08 Test Method for Determining the Potential Alkali-Silica Reactivity of
Combinations of Cementitious Materials and Aggregate (Accelerated
Mortar-Bar Method)
C1579-06 Test Method for Evaluating Plastic Shrinkage Cracking of Restrained
Fiber Reinforced Concrete (Using a Steel Form Insert)
C1580-09 Test Method for Water-Soluble Sulfate in Soil
C1581_C1581M-09 Test Method for Determining Age at Cracking and Induced Tensile
Stress Characteristics of Mortar and Concrete under Restrained
Shrinkage
C1582_C1582M-04 Specification for Admixtures to Inhibit Chloride-Induced Corrosion of
Reinforcing Steel in Concrete
C1583_C1583M-04E01 Test Method for Tensile Strength of Concrete Surfaces and the Bond
Strength or Tensile Strength of Concrete Repair and Overlay Materials
by Direct Tension (Pull-Off Method)
CURRE NT CRI T ER IA 222

ASTM H ig h w a y St a n d a r d s— c o n t in u e d

C1585-04E01 Test Method for Measurement of Rate of Absorption of Water by


Hydraulic-Cement Concretes
C1602_C1602M-06 Specification for Mixing Water Used in the Production of Hydraulic
Cement Concrete
C1603-05A Test Method for Measurement of Solids in Water
C1604_C1604M-05 Test Method for Obtaining and Testing Drilled Cores of Shotcrete
C1609_C1609M-07 Test Method for Flexural Performance of Fiber-Reinforced Concrete
(Using Beam with Third-Point Loading)
C1610_C1610M-06A Test Method for Static Segregation of Self-Consolidating Concrete
Using Column Technique
C1611_C1611M-09A Test Method for Slump Flow of Self-Consolidating Concrete
C1621_C1621M-09A Test Method for Passing Ability of Self-Consolidating Concrete by
J-Ring
C1622_C1622M-06 Specification for Cold-Weather Admixture Systems
C1646_C1646M-08A Practice for Making and Curing Test Specimens for Evaluating
Resistance of Coarse Aggregate to Freezing and Thawing in
Air-Entrained Concrete
C1679-09 Practice for Measuring Hydration Kinetics of Hydraulic Cementitious
Mixtures Using Isothermal Calorimetry
C1688_C1688M-08 Test Method for Density and Void Content of Freshly Mixed Pervious
Concrete
E0329-08 Specification for Agencies Engaged in Construction Inspection
and/or Testing

Eu r o p e a n Sp e c ifi c a t io n s , Co d e s , St a n d a r d s , a n d G u id e lin e s

EN 196 Methods of Testing Cement


Part 1: Determination of Strength
Part 2: Chemical Analysis of Cement
Part 3: Determination of Setting Time and Soundness
Part 5: Pozzolanicity Test for Pozzolanic Cements
Part 6: Determination of Fineness
Part 7: Methods of Taking and Preparing Samples of Cement
Part 21: Determination of the Chloride, Carbon Dioxide, and Alkali Content of Cement
EN 450 Fly Ash for Concrete
Part 1: Definition, Specifications, and Conformity Criteria
Part 2: Conformity Evaluation
EN 934 Admixtures for Concrete, Mortar, and Grout
Part 2: Concrete admixtures – Definitions and Requirements
Part 4: Admixtures for Grout for Prestressing Tendons – Definitions, Requirements,
and Conformity
Part 6: Sampling, Conformity Control, Evaluation of Conformity, Marking, and Labeling
EN 13263-1 Silica Fume for Concrete
EN 197-1 Part 1: Composition, Specifications, and Conformity Criteria for Common Cements
Cement
EN 197-2 Part 2: Conformity Evaluation
Cement
EN 197-3 Part 3: Composition, Specifications, and Conformity Criteria for Low-Heat Common
Cement Cements
CURRE NT CRI T ER IA 223

Eu r o p e a n Sp e c ifi c a t io n s , Co d e s , St a n d a r d s , a n d G u i d e l i n e s— c o n t in u e d

EN 197-4 Part 4: Composition, Specifications, and Conformity Criteria for Low Early Strength
Cement Blast-Furnace Cements
EN 206-1 Part 1: Specification, Performance, Production
Concrete Part 2: Specification for Constituent Materials and Concrete
EN 12350 Testing Fresh Concrete
Part 1: Sampling
Part 2: Slump Test
Part 3: Vibe Test
Part 4: Degree of Compactability
Part 5: Flow Table Test
Part 6: Density
Part 7: Air Content – Pressure Methods
EN 12390 Testing Hardened Concrete
Part 1: Shape, Dimensions, and Other Requirements for Specimens and Molds
Part 2: Making and Curing Specimens for Strength Tests
Part 4: Compressive Strength – Specification for Testing Machines
Part 5: Flexural Strength of Test Specimens
Part 6: Tensile Splitting Strength of Test Specimens
Part 7: Density of Hardened Concrete
Part 8: Depth of Penetration of Water under Pressure
EN 12390: Part 3 Compressive Strength of Test Specimens has been approved at second formal
vote and will be published by BSI in due course
EN 12620 Aggregates for Concrete
EN 14216 Composition, Specifications, and Conformity Criteria for Massive Concrete
Cement Low-Heat Cements
EN 14217 Composition, Specifications, and Conformity Criteria for Low Early Strength
Cement Low-Heat Cements
Part 1: Definitions, Requirements, and Conformity Criteria
Part 2: Conformity Evaluation
EN 13055-1 Lightweight Aggregates
Part 1: Lightweight Aggregates for Concrete and Mortar
EN 13139 Aggregates for Mortar
EN 13242 Aggregates for Unbound Materials for Use in Civil Engineering Work and
Road Construction
EN 932 Tests for General Properties of Aggregates
Part 1: Methods for Sampling
Part 2: Methods for Reducing Laboratory Samples
Part 3 : Procedure and Terminology for Simplified Petrographic Description
Part 5: Common Equipment and calibration
Part 6: Definitions of Repeatability and Reproducibility
EN 933 Tests for Geometrical Properties of Aggregates
Part 1: Determination of Particle Size Distribution – Sieving Method
Part 2: Determination of Particle size distribution – Test Sieves, Nominal Size
of Apertures
Part 3: Determination of Particle Shape – Flakiness Index
Part 4: Determination of Particle Shape – Shape Index
Part 5: Determination of Percentage of Crushed and Broken Surfaces in
Coarse Aggregate Particles
Part 7: Determination of Shell Content – Percentage of Shells in Coarse Aggregates
Part 8: Assessment of Fines – Sand Equivalent Test
Part 9: Assessment of Fines – Ethylene Blue Test
Part 10: Assessment of Fines – Grading of Fillers (Air-Jet Sieving)
CURRE NT CRI T ER IA 224

Eu r o p e a n Sp e c ifi c a t io n s , Co d e s , St a n d a r d s , a n d G u i d e l i n e s— c o n t in u e d

EN 1008 Mixing Water for Concrete


EN 1097 Tests for Mechanical and Physical Properties of Aggregates
Part 1: Determination of the Resistance to Wear (Micro-Devil)
Part 2: Methods for the Determination of Resistance to Fragmentation
Part 3: Determination of Loose Bulk Density and Voids
Part 4: Determination of the Voids of Dry Compacted Filler
Part 5: Determination of the Water Content by Drying in a Ventilated Oven
Part 6: Determination of Particle Density and Water Absorption
Part 7: Determination of the Particle Density of Filler – Pyknometer Method
Part 8: Determination of the Polished Stone Value
Part 9: Determination of the Resistance to Wear by Abrasion from Studded Tires
Nordic Test
EN 1367 Tests for Thermal and Weathering Properties of Aggregates
Part 1: Determination of Resistance to Freezing and Thawing
Part 2: Magnesium Sulfate Test
Part 3: Boiling Test for Sonnenbrand Basalt
Part 4: Determination of Drying Shrinkage
EN 1744 Tests for Chemical Properties of Aggregates
Part 1: Chemical Analysis
EN 1504 Products and Systems for the Protection and Repair of Concrete Structures
EN 1504 Part 10: Site Application of Products and Systems and Quality Control of the Works
EN 12696 Cathodic Protection of Steel in Concrete
EN 13791 Assessment of Concrete Compressive Strength in Structures or in Structural Elements
EN 14038-1 Electrochemical Re-Alkalization and Chloride Extraction Treatments for Reinforced
Concrete
Part 1: Re-Alkalization
Part 2: Chloride Extraction
EN 1542 Measurement of Bond Strength by Pull-Off
EN 1543 Determination of Tensile-Strength Development for Polymers
EN 1766 Reference Concretes for Testing
EN 1770 Determination of the Coefficient of Thermal Expansion
EN 1799 Tests to Measure the Suitability of Structural Bonding Agents for Application to
Concrete Surface
EN 12636 Determination of Adhesion Concrete to Concrete
EN 12504 Testing Concrete in Structures
Part 1: Cored Specimens – Taking, Examining, and Testing in Compression
Part 2: Non-Destructive Testing – Determination of Rebound Number
In preparation:
Part 3: Determination of Pull-Out Force
Part 4: Determination of Ultrasonic Pulse Velocity

Das könnte Ihnen auch gefallen