You are on page 1of 41

Troubleshooting Concrete Problems through Petrography

Part I: Petrographic Techniques, Diagnosing Improper Quality, Workmanship,


Atmospheric Interactions, and Freeze-Thaw Deteriorations in Concrete
Dipayan Jana and Shondeep L Sarkar1

INTRODUCTION
Concrete is a continuously changing system - starting from its mixing to its placement. Even after
hardening, concrete interacts with the environment to which it is exposed. Strength, durability or
resistance to adverse environment, and dimensional stability have long been recognized as the three
fundamental requirements for making a good quality concrete, which will provide satisfactory
performance for many years, even in a hostile environment. Less attention, however, is paid to the
microstructure, which is the basic anatomy of a durable concrete, which not only controls the strength
but also the durability and dimensional stability. Understanding the performance characteristics of a
concrete can be seriously hampered by inadequate attention to its detailed microstructure [1].
Petrography, the 150-year old science of rock-microscopy, is a powerful tool for studying
microstructural characteristics of concrete. From the selection of cementitious materials and aggregates
to be used in concrete to the quality control and quality assurance during concrete placement, from the
condition survey of an existing concrete structure to the failure investigation of deteriorated concrete,
and finally to the long-term performance recommendations – the importance of petrography in every
facet has been fully acknowledged and appreciated by the cement and concrete community [2-7].

The owner of a concrete structure often wants to know why the concrete has developed symptoms of
deterioration, how bad is it, is it premature, can he live with it, and how can it be prevented and repaired.
The ability to obtain a quick, and yet definitive and cost-effective answer to these questions has
increasingly been making petrography a dominant investigative tool for civil and structural engineers,
and architects [8,9]. A review of the basic petrographic techniques used in concrete failure investigation
are presented first, followed by the applications of petrography in the diagnosis of various physical and
chemical deleterious processes that are common in a deteriorated concrete.

MICROSCOPICAL AND MICROANALYTICAL TECHNIQUES


In a routine concrete failure investigation by petrography, light optical microscopes are the most
commonly used instruments, which are sometimes supplemented by scanning electron microscope
(Figure 1). Both optical and electron microscopes offer a wealth of information about the materials
characteristics, composition, mineralogy, and microstructure of concrete and various important
advantages over other conventional methods such as: (a) rapid evaluation of the quality of basic concrete
ingredients and their impact on the overall condition and performance of concrete; (b) identification of
any signs of improper construction practices (improper proportioning, mixing, placement, finishing, and
curing practices), which could possibly be responsible for unsatisfactory performance of concrete during
its service; (c) in situ observation of microstructural characteristics and diagnosing evidences of
deleterious physical and chemical reactions in a deteriorated concrete at high magnification and
resolution, where all concrete constituents preserve their composition, texture and distribution; (d)
identification, location, abundance and chemical composition of deleterious constituents in concrete
responsible for concrete deterioration; and, (e) rapidity and cost effectiveness in obtaining reports of

1
Dedicated to late Dr. Shondeep L Sarkar for his significant contributions to understanding concrete microstructure.
failure investigation. In failure investigation, field investigation is an important practice, which
determines the nature and extent of deteriorations and selects appropriate locations for sampling for
petrographic examinations. The examination is done by using the procedures outlined in ASTM C 295
for aggregates, ASTM C 856 for hardened concrete, and ASTM C 1324 for masonry mortars.

OPTICAL MICROSCOPY
The use of optical microscopes dates back as early as 1850 when Sorby discovered preparation of thin
sections, transparent to light, and 1897 when Tomebohm examined various clinker particles under the
optical microscopes. 'Reflected-light' and 'transmitted-light' microscopes are the two standard optical
microscopes routinely used in petrography [10-12].

In 'reflected-light microscopy' (e.g. stereomicroscopes and metallographic microscopes), white or


polarized light reflected from a polished or fresh fractured surface of a concrete is allowed to pass
through a series of lenses to magnify the surface features to as high as 100 times in stereomicroscope and
1000 times in metallurgical microscopes. Visual and stereomicroscopic examinations at low
magnifications are the very first steps of petrographic examination. Information obtainable from this step
includes quality and composition of the aggregates; in place deterioration of aggregate particles;
tightness of aggregate-paste bonds; cement-aggregate reactions; color, hardness and signs of
disintegration in the paste; depth and extent of carbonation; air void characteristics (amount, spacing,
size ranges, and distribution according to ASTM C 457 [6]; extent and severity of cracking; segregation
of coarse aggregates or bleeding; location of secondary deposits; tightness of bonding between concrete
and any topping or substrate layers; overall homogeneity of paste and concrete; identity and location of
exudations from the concrete; efflorescent salt deposits; presence and identity of coating materials
applied to the concrete, etc [13, 14]. Examinations of polished, etched, and stained surfaces by using a
metallographic microscope can identify unhydrated cement particles in the paste [13, 14].

In 'transmitted-light microscopy', white light is first filtered through a polarizer so that it can vibrate only
in one direction (perpendicular to the direction of propagation, [15]), and then transmitted through the
thin section (0.2 to 0.3 mm thick) of concrete (Figure 1). Interaction of polarized light with concrete
helps identification of various crystalline phases in aggregates, paste, air voids, and cracks by their
characteristic optical properties. A standard polarized light microscope can magnify fine details of
concrete up to 1000 times. Most modem 'petrographic microscopes' are equipped with both reflected and
transmitted light modes. Transmitted-light microscopes can be very helpful to identify the presence of
chemically (e.g., alkali-reactive) and volumetrically unstable minerals (e.g. clays) in aggregates; the
presence of corrosive contaminants in aggregates or paste; the degree of cement hydration and water-
cement ratio; microcracks in the paste; the presence of fly ash; and the identification, location and
amount of deleterious secondary deposits (e.g., alkali-silica gel, ettringite, carbonates, brucite), whose
formation can cause significant expansion and cracking in the concrete. A blue dye impregnation during
thin section preparation can highlight microcracks, air voids, capillary pores in the paste and helps to
estimate the water-cement ratio of concrete from the degree of blue coloration. Small granular mounts of
specific materials (scratched from a concrete surface) in immersion oil can determine unhydrated cement
particles in the paste, the kind of cement (i.e., Portland cement, Portland-blast furnace slag cement, or
Portland-pozzolan cement), mineral admixtures, secondary deposits, degree of hydration, carbonation,
leaching or other alterations in the paste [13,14].
Petrographic microscopes can be modified to obtain 'fluorescent-light' mode by inserting a deep blue
(UV) filter in the light path (instead of the polarizer) before it enters the thin section and impregnating
the thin section with a fluorescent dye. The resultant UV light from filter causes the dye to fluoresce,
which distinctly highlights all original voids (air voids [16], capillary pores in paste [16], cracks [17],
and aggregate porosities [17]) in concrete. Any variation in paste porosity can be determined from
variation in the degree of fluorescence. The porous areas in the paste appear light grey, whereas, the
denser areas with a low water-cement ratio appear dark grey. Fluorescent microscopy is useful in
diagnosis of microcracking, alkali-aggregate gel, and semi-quantitative determination of water-cement
ratio in concrete.

Both reflected and transmitted light microscopes can be attached to a high resolution CCD micro video
camera, 35mm camera or modern high-resolution digital cameras for capturing images of small areas of
polished or thin sections of concrete at regular intervals and storing them in a computer (equipped with
an image analyzing software) for image analysis. Air void amount and distribution can be determined
quite rapidly from image analysis of a polished or large area thin section of concrete [18]. Proportion of
various phases in concrete can be obtained from statistical analysis of multiple images taken from a large
representative area.

The wealth and accuracy of information obtained from optical microscopes, however, depends
predominantly on careful sample preparation with minimum disturbance to the original features of the
defective concrete. Selection of a representative specimen, drying, epoxy impregnation to seal all
original pores and cracks, slicing, grinding and polishing the surface to be examined, and thin sectioning
of the polished surface are the common steps in sample preparation [19].

ELECTRON MICROSCOPY AND MICROANALYSIS


Unlike optical microscopes, the application of electron microscopes to failure investigation is relatively
new, although its use in the study of other inorganic materials has long been in practice. Ability to obtain
high-resolution images of concrete surface at extremely high magnifications (up to 105 times), qualitative
and quantitative phase identification, and the ability to simultaneously analyze the chemical composition
of deleterious constituents are some of the important advantages of using electron microscopy. In
electron microscopy, polished and epoxy-impregnated concrete specimen, thin sections, or fracture
specimen are excited with a focused primary electron beam from an electron-emitting gun (Figure 1).
The accelerated primary electrons strike the sample, lose their energy, and generate a number of signals
or effects that are possible to detect by various function-dependent detectors. Descriptions of various
signals useful for concrete failure investigation are presented in the following paragraphs.

A portion of incident primary electrons undergoes violent collisions with atoms of the sample and are
scattered back out of its surface as backscattered electrons. The higher the average atomic weight of a
concrete phase, the higher the proportion of electrons backscattered from its surface, and the brighter the
appearance of that phase in the image collected by a backscatter detector [20]. As a result, backscattered
electron (BSE) imaging is used extensively in cement and concrete research to determine the abundance
and distribution of phases of different brightness (due to their different chemical compositions). BSE
imaging has been used to study the porosity of hardened cement paste [21, 22], proportion of unhydrated
cement particles in paste [23, 24], effect of silica fume [43 to 46] and superplastisizer [47] additions on
the hardened paste microstructure, etc.
If the bombarding electrons are sufficiently energetic, they can dislodge electrons from an inner orbital in
the target sample. The excited atom rapidly rearranges its electronic structure and emits characteristic X-
rays. The intensities of characteristic X-rays of different elements from a phase are detected by suitable
spectrometers attached to an electron probe microanalyzer (EPMA), which converts the elemental
intensities into their respective concentrations in the phase (see Figure 1). The elemental concentration
can be recorded either for a particular phase in concrete (phase analysis), or for spatial distribution of a
particular element in the whole image (X-ray imaging), or for variation in concentration along a traverse
(line scan) [20, 25, 26]. Figure 2 shows backscatter electron imaging, corresponding X-ray elemental
imaging, and phase analyses of two concrete samples to illustrate how readily alkali-silica reaction and
sulfate attack- the two common deleterious processes can be detected [26].

If the bombarding electrons are of low energy, they will ionize the atoms of the target sample by
dislodging loosely bound electrons as low-energy secondary electrons. Secondary electrons are detected
by scanning electron microscope (SEM), which images the topography of concrete surfaces by scanning
across it. Since secondary electrons emit from a small-localized region around the area of impact, their
signal in SEM provides a much higher image resolution than that obtainable from backscatter electrons,
which originate from larger sample volume. SEM imaging has been used extensively in determining the
location, distribution and morphologies of various cement hydration products (e.g. calcium silicate
hydrate gel, calcium hydroxide, [27]) and secondary deposits (e.g., ettringite needles, [28]) in concrete.
Most modem SEM is fully equipped with a characteristic X-ray detector (energy-dispersive
spectrometer, EDS), and backscatter detector, so that high-resolution topographic imaging, compositional
analysis, and phase proportioning from contrasting brightness can be done with this powerful instrument
(SEM-EDS).

Environmental scanning electron microscope (ESEM) [29-31] represents the latest advancement in SEM.
It allows water vapor to be retained in the sample chamber (instead of high vacuum), which helps in situ
observation of cement hydration processes and microstructural development as a function of time. It has
yet to be established as a concrete failure investigative tool, nonetheless, it holds potential, particularly in
the study of plastic shrinkage crack propagation.

A portion of primary electrons can transmit through the ultra thin concrete slice and detected by
transmission electron microscope (TEM). Due to lesser lateral spread of incident electrons, TEM can
reproduce still higher resolution and magnification than SEM, and can also provide elemental analysis
with an X-ray detector. However, its application is limited in concrete research (e.g., dissolution rate of
silica fume, [32]) due to ultra thin (< l micron) sample requirement.

In spite of numerous functional advantages over conventional optical microscopes, the higher
magnification of electron microscopes does significantly reduce the total area of an investigated sample,
which necessitates the examination of an adequate total sampling area in the specimen to obtain any
reasonably representative data or information.

TROUBLESHOOTING CONCRETE PROBLEMS


Strength and durability are the two most important concerns of concrete engineers. Microscopic
examination can provide sufficient evidence to accurately identify the reasons behind low strength or
poor durability. For example, high water-cement ratio, excessive air voids, inferior aggregates, improper
proportioning-mixing-curing, carbonation, various forms of chemical-attack, alkali-aggregate reaction,
inadequate air entrainment for freeze-thaw resistance - all possible culprits can be readily determined
from microscopic observation. Before entering into a discussion of their identifications, an overview of
the typical microstructure of concrete must be presented [1, 33, 34]. Figure 3 shows basic microstructure
of sound portland cement concrete as seen by using optical microscopes. Properties shown in the figure
are (a) well graded crushed stone coarse aggregates with tight aggregate-paste bonding, (b) closely-
spaced uniformly distributed adequately entrained air voids in the paste, (c) well graded natural sand fine
aggregates, and, (d) residual cement particles (alite with hydration rim) in the paste. Figure 4 shows high
magnification scanning electron images of calcium silicate hydrate (C-S-H) gel, calcium hydroxide (CH)
crystals, and ettringite crystals in the hardened cement paste.

(1) Failure in Quality Assurance & Improper Proportioning


Microscopical examinations shed definite light on whether an observed failure in a concrete structure is
due to the poor quality of the ingredients used in the mix, or to a good quality but improper proportioning
of ingredients.

Optical microscopy can easily identify whether a potentially alkali-reactive aggregate has been
incorporated in a concrete mixture. Elongated, poorly graded, dirty (dull-lustered, clay-coated) coarse
aggregates increase water demand, make consolidation difficult, reduce the tightness of paste-aggregate
bonding, and, as a result, reduce the strength of concrete [35, 36] (Figure 5 a, b). Contamination of an
innocuous aggregate with a trace amount of reactive constituents can make it deleterious. Optical
identification of porous aggregates (pumice, scoria, weathered chert, etc.) or water absorptive minerals
(clays), or iron sulfide minerals (pyrite) in aggregate can be directly ascribed to the aggregate popout in
concrete surface [35,37]. The presence of iron oxide or iron sulfide particles in aggregate (easily
detectable under microscope) is a plausible cause of the yellow, or yellow-red to deep rust-brown
staining and discoloration of exposed concrete surface [8, 35] (Figure 5 c). Excess free lime or magnesia
contaminants in carbonate coarse aggregates or Portland cement (especially when cement clinker cooled
slowly) can strongly react with water and substantially increase in volume by formation of hydroxides [8,
38]. Coarse-ground Portland cement can be identified by characteristic black mottling (salt-and-pepper)
appearance of the paste under stereomicroscope [8].

Distinctly higher microscopically estimated W/C than the design specified limit can be the prime cause
of a concrete's low strength. High W/C is estimated from trace amounts of unhydrated cement particles in
paste; high porosity of paste; distinct soft, weak, white to pale yellow chalky-textured mottling aspect of
paste; abundant colorless to white hexagonal platelets of calcium hydroxide as secondary deposits in air
voids or bleeding channels; weak paste-aggregate bonds (with interfaces filled with calcium hydroxide
crystals); extensive paste carbonation, fine ramifying drying shrinkage microcracks in paste due to
excessive loss of water, etc.

Any significant variation in the proportions of coarse and fine aggregates, cement content, and air
content (which are all possible to measure by microscopical point count or linear traverse analysis of
hardened concrete as described in ASTM C 457 [6]) from the design specifications can also contribute to
various undesirable results including strength reduction (e.g. by high air content, high water content, low
cement content, or high sand/paste ratio in mortar, Figure 5 d-f).

Excessive or uncontrolled drying shrinkage of concrete after hardening can cause random cracking of
concrete surface and slab curling. Shrinkage, though a common phenomenon of a normal portland
cement concrete can be high if: (a) the water content in the concrete is high, (b) the paste volume (and
cement-to-aggregate ratio) of concrete is high, (c) aggregates contain a high proportion of clay-like or
organic materials which have a high shrinkage potential, or (d) aggregate is weak and compressible (e.g.,
highly porous, micaceous and clayey sandstone) which cannot restrain the normal drying shrinkage of
the cement paste [13]. Most of these evidences are possible to diagnose by a detailed petrographic
examination. Absence of control joints in a slab, shallow cut joints, wide-spaced joints, or delayed
placement of joints in a slab can cause early shrinkage and cracking, which are possible to investigate
during field reconnaissance prior to petrographic examinations.

(2) Improper Mixing, Placement, Finishing & Inadequate Curing


Despite use of good quality cement and aggregates in a well-proportioned concrete mixture, improper
mixing, placement, finishing and curing practices can destroy the materials quality and cause a premature
failure, or development of defects.

Inadequate mixing of ingredients can reduce the strength considerably. Inadequate mixing can be
suspected from clustering of coarse aggregates, air voids or unhydrated cement particles (Figure 6 a, b);
significant variation in paste density; or, presence of sand streaks in the exposed concrete slab or in its
representative polished section (Figure 6 c [8]).

Inadequate compaction, on the other hand, can be determined from honeycombing (Figure 6 d, e) or
presence of large irregularly shaped entrapped air voids in between the coarse aggregates [8].

Bleeding, or upward migration of water in concrete in its plastic state can be determined from narrow
bleeding channels in paste around the coarse aggregates at right angles to the finished surfaces (Figure 6
f); weakening of the paste-aggregate bonds (especially towards the exposed top surface. Figure 6 g), or,
segregation of large voids (with large calcium hydroxide crystals) towards the top in an excessively wet
or non air-entrained concrete [8].

Segregation of coarse aggregates, due to either over-vibration or to an excessively wet mix, can be
readily recognized in a polished section (Figure 6 h). Segregation from over-vibration commonly exhibits
a decreasing air content and increasing fine-grained mortar towards the top.

Retempering, or addition of excess water during placement to increase the workability of the mix in hot
weather, can be detected from the presence of darker patches of paste of low water/cement ratio in the
lighter matrix of higher water-cement ratio (Figure 6 i) ; coating of dark gray paste around coarse
aggregates; air void clustering; or compositional and physical heterogeneity in the hardened cement paste
[8, 13].

Improper curing (e.g. rapid drying) transforms the normal paste into a soft, chalky -textured, and highly
carbonated (dusted) white colored paste (Figure 6 j, [8]). The proportion of unhydrated cement particles
and highly birefringent calcium carbonate crystals in the paste show an increase from interior to the
exposed surface of concrete. Plastic shrinkage cracks (Figure 6 k) in the paste also indicate improper
curing of concrete due to an increased rate of evaporation of surface water from plastic concrete than the
rate of bleeding (especially in high temperature, low humidity and high wind velocity conditions).
(3) Efflorescence and other discolorations
Efflorescence, or precipitation of white powdery salt deposits (anhydrous or hydrous, carbonate and/or
sulfate salts of calcium, sodium, potassium) at the concrete surface can be identified from field (Figure
7a) and optical microscopic studies [39, 40]. Oil immersion mount or SEM studies can also successfully
diagnose various efflorescence deposits. For example, Figure 7 b illustrates SEM image of glauberite - an
efflorescent salt in a deteriorated mortar, and, SEM morphology and corresponding X-ray elemental
spectrum (Fig 7 c, d) of a salt deposit (thenardite) at the paste-aggregate interface of a concrete, which
had been seasonally exposed to deicer salt for a number of years. The same concrete also showed the
presence of calcium chloroaluminate crystals in the paste, formed from the breakdown of the deicer salt.

(4) Carbonation
An inevitable consequence of concrete-atmosphere interaction is the reaction between the cement
hydration product (calcium hydroxide, Ca(OH)2) and atmospheric carbon dioxide (C02) to produce
calcium carbonate (CaC03). In a well-cured concrete, carbonation produces a thin (< 5 to 10 mm), hard,
resistant to abrasion and freezing-thawing, relatively impermeable coating at the finished surface
(laitance) which in a way acts as a sealant to prevent chloride or other aggressive ion penetration. But in
the case of poorly cured plastic concrete (e.g., where unvented heating during curing produced a lot of
COz), or poorly compacted plastic concrete of high W/C, severe carbonation (> 10 mm) can occur and
creates a soft, friable paste ^'1;3' 14, 40 ^g depth of carbonation increases with increasing W/C.
Carbonation can be easily detected in a fresh fracture or polished concrete section by treatment with a
phenolphthalein solution, where the solution turns colorless in the carbonated areas but becomes pink in
the noncarbonated areas (Fig 8 a-c). A thin section of carbonated concrete shows a light colored paste,
filled with minute highly birefringent crystals of calcium carbonate 8, 13 SEM images of some of the
commonly observed morphologies of CaC03 crystals1 produced during carbonation are shown in Fig 8
d-g . The long-term exposure of concrete to atmospheric C02 often results in shrinkage and cracking of
the concrete. Moreover, consumption of calcium hydroxide during carbonation can reduce the pH of
concrete to a point which facilitates rebar corrosion. Carbonation, however, does not occur aggressively
at very low or high relative humidities ^.

(5) Freeze-thaw deterioration

Common surface problems like scaling, spalling, delamination, and cracking are usually due to either
early freezing of a fresh, plastic concrete, or, to inadequate air entrainment of hardened concrete, where
frozen water (i.e. ice lenses) expands and induces cracking in the concrete. Fig 9 a, b shows that the
severity of scaling increases with decreasing air voids, due to lesser open spaces to accommodate the
volume expansion during freezing. Microscopic evidence in favor of freeze-thaw related deterioration
include, (a) extensive cracking subparallel to the exposed surface (Fig 9 c-e), (b) inadequate air
entrainment (Fig 9f), (c) large air void spacing (Fig 9g), (d) ice crystal impressions (small subradial
clusters of linear fractures, Fig 9 i, j) throughout the paste, or in aggregate-sockets indicating early
freezing of a plastic concrete ^' ^, and, (e) presence of abundant water-absorptive minerals (e.g., zeolites,
clays) in coarse aggregates, or some finely porous aggregates (cherts, argillaceous limestones,
ferruginous concretions) (Fig 9 h) in concrete which can cause expansion leading to aggregate popouts or
even surface scaling. Compared to the conventional 46x27 mm thin sections, the large-area sectioning
(50x75 mm or 100x70 mm) of concrete 41 with blue or fluorescence dye epoxy impregnation 17 is
desirable for tracing the network of microcracks in concrete. Additional evidence, like low freeze-thaw
durability (as per ASTM C666 test ^), and the presence of deicing salts on the surface, or their
derivatives in the interior of the concrete often suggest freeze-thaw related attack.

CONCLUSION

Durability of concrete has long been recognized as an important issue in construction technology.
Compare to various nondestructive, chemical, or physical testings, microscopical examination very often
offers quick, yet definite answers to concrete problems. Inferior quality of aggregates threatens concrete's
performance. Improper proportioning, compaction, mixing, placement and curing operations can destroy
the inherent quality of ingredients and develop signs of deterioration like scaling, spalling, delamination,
cracking, reductions in strength, etc. Improper curing, or high W/C ratio can cause severe carbonation,
softening of the paste and strength loss. Inadequate air entrainment shows an adverse effect on the
freeze-thaw durability of concrete. Microscopic examination is the most promising investigative method
to identify these problems, determine the extent of damage, and shed light on the underlying cause(s) of
failure. In the second part of this series, we will focus on problems associated with alkali-aggregate
reaction, external and internal sulfate attacks, seawater attack, rebar corrosion, and fire attack. In
addition, the effects of addition of various admixtures on the microstructure and their bearing on the
mechanical properties of concrete will be discussed with numerous optical and scanning electron
micrographs. These two articles will hopefully convey an important message to the construction industry
that, petrography, a long used science of rock microscopy, can offer much more than many other existing
methods in failure investigation.

ACKNOWLEDGMENTS

Sincere thanks are extended to the following individuals and organizations for providing various
photographs included in this article: Paul Stutzman (National Institute of Science and Technology), Patsy
Harms (Portland Cement Association), Ronald Sturm (Construction Technology Laboratories), the
Aberdeep Group, and Kathe Hooper (American Society for Testing and Materials).
References
1. Sarkar, S.L., "The Importance of Microstructure in Evaluating Concrete", Advances in Concrete
Technology, V.M. Malhotra ed., 2nd Ed., CANMET, Ottawa, 1994, pp 125-160.
2. Eriin, B., and Stark, D., Petrography Applied to Concrete and Concrete Aggregates, ASTM STP
1061, American Society for Testing and Materials, Philadelphia, PA, 1990,155 pp.
3. DeHays, S., and Stark, D., Petrography of cementitious materials, ASTM STP 1215, American
Society for Testing and Materials, Philadelphia, PA, 1994,208 pp.
4. Proceedings of International Conferences on Cement Microscopy; International Cement
Microscopy Association (ICMA); Duncanville, Texas, USA.
5. ASTM C 295 "Standard Practice for Petrographic Examination of Aggregates for Concrete", in
Annual Book of ASTM Standards, Section 4 - Construction, Vol 04.02-Concrete and Aggregates,
American Society for Testing and Materials, Philadelphia, PA, 1993.
6. ASTM C 457 "Microscopical Determination of Air-Void Content and Parameters of the Air-Void
System in Hardened Concrete", in Annual Book of ASTM Standards, Section 4 - Construction, Vol
04.02-Concrete and Aggregates, American Society for Testing and Materials, Philadelphia, PA,
1993.
7. ASTM C 856 "Petrographic Examination of Hardened Concrete", in Annual Book of ASTM
Standards, Section 4 - Construction, Vol 04.02-Concrete and Aggregates, American Society for
Testing and Materials, Philadelphia, PA, 1993.
8. Ray, J.A., "Things Petrographic Examination Can and Cannot Do With Concrete", Proceedings of
the 5th International Conference on Cement Microscopy, Nashville, ICMA, 1983, pp. 66-84.
9. Jana, D., "Petrography : A powerful tool for solving common concrete problems". Civil
Engineering NEWS, March 1997, pp. 40-44.
10. Campbell, D.H., "Application of the Microscope in the Concrete Industry", Proceedings of the
Third International Conference on Cement Microscopy, G. Gouder ed., 1981, pp. 286-297.
11. Erlin, B., "Methods used in Petrographic Studies of Concrete", In Analytical Techniques for
Hydraulic Cement and Concrete, ASTM STP 395, American Society for Testing and Materials,
Philadelphia, 1966, pp. 3-17.
12. Mather, K., "Applications of light microscopy in concrete research". Am. Soc. Testing Materials,
Spec. Tech. Pub, No 143,1953, pp. 51-69.
13. Mielenz, R.C., "Petrography Applied to Portland-Cement Concrete", Reviews in Engineering
Geology, Geological Society of America, Vol 1,1962, pp. 1-38.
14. Mielenz, R.C., "Diagnosing Concrete Failures", Stanton Walker Lecture Series on the Material
Sciences, Lecture No 2, University of Maryland, College Park, MD, 1964, pp. 3-47.
15. Bloss, F. D. An introduction to the methods of optical crystalligraphy; Holt, Rinehart and Winston,
New York, 1961,294pp; and Potts, P.J., A Handbook of Silicate Rock Analysis, Blackie, Glasgow,
1987, 844pp.
16. Wirgot, S., and Van Cauwelaert, F., "The Influence of Cement Type and Degree of Hydration on
the Measurement of W/C Ratio on Concrete Fluorescent Thin Sections", Petrography of
cementitious materials, S. DeHays, and D. Stark, D, eds, ASTM STP 1215, American Society for
Testing and Materials, Philadelphia, PA, 1994, pp. 91-110.
17. Anderson, K.T., and Thaulow, N., "The study of alkali-silica reactions in concrete by the use of
fluorescence thin-sections". Petrography Applied to Concrete and Concrete Aggregates, B. Eriin,
and D. Stark eds., ASTM STP 1061, American Society for Testing and Materials, Philadelphia, PA,
1990, pp.71-89.
18. Cahill, J., Dolan, J.C., and Inward, P.W., "The Identification and measurement of Entrained Air in
Concrete Using Image Analysis", Petrography of cementitious materials, S. DeHays, and D. Stark
eds., ASTM STP 1215, American Society for Testing and Materials, Philadelphia, PA, 1994, pp.
111-124.
19. Ahmed, W.U., "Petrographic Methods for Analysis of Cement Clinker and Concrete
Microstructure", Petrography of cementitious materials, S. DeHays, and D. Stark eds., ASTM STP
1215, American Society for Testing and Materials, Philadelphia, PA, 1994, pp. 1-12.
20. Goldstein, J.I., Newbury, D.E., Echlin, P., Joy, D.C., Piori, C., and Lifshin, E., Scanning Electron
Microscopy and X-ray Microanalysis, Plunum Press, New York, 1984, 673pp; and Reed, S.J.B.,
Electron Microprobe Analysis, 2nd Ed, Cambridge University Press, Cambridge, 1993,264pp.
21. Scrivener, K.L,, "The use of Backscattered Electron Microscopy and Image Analysis to Study the
Porosity of Cement Paste", Mat. Res. Soc. Symp, Proc. 137,1989, pp. 129.
22. Schrivener, K.L., "The Microstructure of Concrete", Materials Science of Concrete I, J.P. Skalny
ed., The American Ceramic Society, Westerville, 1989, pp. 127-162.
23. Scrivener, K.L., "The microstructure of anhydrous cement and its effect on hydration",
Microstructural development during hydration of cement, L.J. Struble and P.W. Brown eds..
Materials Research Society Symposia Proceedings, 1987, pp. 39-46.
24. Bentz, D.P., and Stutzman, P.E., "SEM analysis and computer modelling of hydration of Portland
cement particles". Petrography of cementitious materials, S. DeHays, and D. Stark eds., ASTM STP
1215, American Society for Testing and Materials, Philadelphia, PA, 1994, pp. 60-73.
25. Marusin, S.L., "The role of SEM-X-ray elemental analyses and dot map studies in evaluating
building materials failures". Proceedings of 18th International Conference on Cement Microscopy,
L. Jany, A. Nisperos, and J. Bayles eds., 1996, pp. 33-45.
26. Stutzman, P.E., "Application of Scanning Electron Microscopy in Cement and Concrete
Petrography", Petrography of cementitious materials, S. DeHays, and D. Stark eds., ASTM STP
1215, American Society for Testing and Materials, Philadelphia, PA, 1994, pp. 74-90.
27. Bames, B.D., Diamond, S., and Dolch, W.L., "Micromorphology of the Interfacial Zone around
aggregates in Portland Cement Mortar", Journal of the American Chemical Society, Vol 62,1979,
pp. 21-24.
28. Mehta, P.K., "Scanning electron micrographic studies ofettringite formation", Cem. Conor. Res.,
vol 6, No 2,1976,pp.169-182.
29. Kjellsen, K.O. and Jennings, H.M., "Observations of Microcracking in Cement Paste Upon Drying
and Rewetting by Environmental Scanning Electron Microscopy", Adv. Cem, B. Mater. 3,1996.,
pp. 42.
30. Caveny, B., "Cement hydration Study Using the Environmental Scanning Microscopy",
Proceedings of the 14th International Conference on Cement Microscopy, Costa Mesa,
International Cement Microscopy Association, Duncanville, TX, 1992, pp. 29.
31. Mehta, S., Jones, R., Caveny, B., Chatterji, J. and McPherson, G., "Environmental Scanning
Electron Microscope (ESEM). Examination of Individually Hydrated Portland Cement Phases",
Proceedings of the 16th International Conference on Cement Microscopy, Richmond, International
Cement Microscopy Association, Duncanville, TX, 1994, pp. 239.
32. Sarkar, S.L,, and Aitcin, P.C., "Dissolution rate of silica fume in very high strength concrete", Cem.
Concr. Res., Vol 17,1987, pp. 591-601.
33. Diamond, S., "Cement Paste Microstructure - An Overview at Several Levels", Proceedings of the
Conference on Hydraulic Cement Pastes: Their Structure and Properties, Cement and Concrete
Association, 1976, pp. 2-30.
34. Mehta, P.K., Concrete - Structure, Properties and Materials, Prentice-Hall Inc., New Jersey, 1986,
440pp.
35. Mielenz, R.C., "Petrographic Evaluation of Concrete Aggregates", Significance of Tests and
Properties of Concrete and Concrete-Making Materials, P. Klieger and J.F. Lamond eds., ASTM
STP 169C, American Society for Testing and Materials, Philadelphia, PA, 1994, pp. 341-364.
36. Leeman, M.E., and Sarkar, S.L., "Petrographic Identification of quality control failure in a
concrete", Proceedings of 18th International Conference on Cement Microscopy, L. Jany, A.
Nisperos, and J. Bayles eds., 1996, pp. 480-492.
37. Shayan, A., "Deterioration of a concrete surface due to the oxidation of pyrite contained in pyritic
aggregate", Cem. Concr. Res., Vol 18,1988, pp. 723-730.
38. Cady, P.D., and Weyers, R.E., "Petrographic examination aid in establishing the causes of
deterioration of precast concrete sewer manhole sections". Petrography Applied to Concrete and
Concrete Aggregates, B. Eriin, and D. Stark eds., ASTM STP 1061, American Society for Testing
and Materials, Philadelphia, PA, 1990, pp. 182-193.
39. Depuy, G.W., "Chemical resistance of concrete", Significance of Tests and Properties of Concrete
and Concrete-Making Materials, P. Klieger and J.F. Lamond eds., ASTM STP 169C, American
Society for Testing and Materials, Philadelphia, PA, 1994, pp. 263-281.
40. Neville, A.M., Properties of Concrete, John Wiley and Sons, me., 4th Ed, 1996, 844pp.
41. St. John, D.A., "The use of large-area thin sectioning in the petrographic examination of concrete",
Petrography Applied to Concrete and Concrete Aggregates, B. Eriin, and D. Stark eds., ASTM STP
1061, American Society for Testing and Materials, Philadelphia, PA, 1990, pp. 55-70.
42. ASTM C 666 "Test Method for Resistance of Concrete to Rapid Freezing and Thawing", in Annual
Book of ASTM Standards, Section 4 - Construction, Vol 04.02-Concrete and Aggregates, American
Society for Testing and Materials, Philadelphia, PA, 1993.
43. Scali, M.J., Chin, D., and Berke, N.S., "Effect of Microsilica and Fly Ash upon the microstructure
and permeability of concrete", Proceedings of the IXth International Conference on Cement
Microscopy, Reno, ICMA, 1987, pp. 375-397.
43. Scrivener, K.L., Bentur, A., and Pratt, P.L., "Quantitative characterization of the Transition zone in
High Strength Concretes", Adv. Cem. Res., Vol 1,1988, pp. 230-237.
44. Bentz, D.P., Stutman, P.E., and Garboczi, E,J., "Experimental and simulation studies of the
interfacial zone in concrete", Cem. Concr. Res., Vol 22,1992, pp. 891-902.
45. Marusin, S.L., "Microstructure and Pore Characteristics of Concrete containing condensed silica
fume with a superplastisizer". Proceedings of the VIII th International Conference on Cement
Microscopy, Oriando, ICMA, 1986, pp. 327-335.
47. Diamond, S., and Wang, Y., "A quantitative image analysis study of the influence of
superplastisizer on cement paste microstructure". Proceedings of 18th International Conference on
Cement Microscopy, L. Jany, A. Nisperos, and J. Bayles eds., 1996, pp. 465-479.
Figure 1: Basic configurations of polarized-light, fluorescent-light, and scanning electron microscopes
(Ref. 15).
Figure 2: Backscatter electron image (top left photo), X-ray elemental images of sodium and silicon (top right and
bottom left photos), and X-ray spectral analysis (with oxide compositions, bottom right photo) of alkali-silica reaction
gel in a crack (arrows) in concrete. In the elemental images, the sodium-silica composition of the gel is identified by
preferential locations of these elements in gel in the crack. Photo curtsey of Paul E. Stutzman, NIST.
Figure 3: Basic microstructure of portland cement concrete as seen through a stereomicroscope and a petrographic
microscope. In the top left photo, a lapped section of concrete shows dark-colored crushed limestone coarse aggregate,
light-colored natural siliceous sand fine aggregate, intimate bonds between aggregate particles and paste, and dense,
crack-free, sound nature of concrete. In top right, thin-section photomicrograph of a portland cement concrete shows
natural siliceous (quartz-quartzite-feldspar) sand fine aggregate particles. The bottom left photo shows spherical
entrained air voids in a lapped section of concrete; the voids are all less than 1 mm in diameter. The bottom-right photo
shows typical microstructure of portland cement paste containing residual alite grains with hydration rims, dark residual
ferrite phases and an overwhelming matrix phase consisting of calcium silicate hydrate and calcium hydroxide
components of cement hydration.
Figure 4: Basic microstructure of portland cement concrete as seen through a scanning electron
microscope. The photos show secondary electron images of portland cement hydration products
consisting of: (a) fibrous calcium-silicate hydrate gel (top left); (b) preponderance of platy calcium
hydroxide crystals at the paste-aggregate interface (top right); (c) massive tabular calcium hydroxide
crystals in the paste having a water-cement ratio of 0.55 (bottom left); and (d) ettringite needles and platy
calcium hydroxide crystals at the paste-aggregate interface (bottom right).
Figure 8: Secondary electron images of various morphologies of calcium carbonate crystals in concrete.
Figure 9: Freeze-thaw deterioration (continued) - (g) Aggregate popout due to freezing of the absorbed water in pore
spaces of aggregates; (h) large spacing of entrained air voids threatens freeze-thaw durability; (i) ice crystal impressions
on the exposed concrete surface; (j) subparallel ice-lens impressions in thin section.
Troubleshooting Concrete Problems through Petrography
Part II: Diagnosing Aggregate Reactions, Sulfate, Seawater, Fire Attacks, and Rebar
Corrosion in Concrete
Dipayan Jana and Shondeep L Sarkar1

INTRODUCTION
Strength, durability, and dimensional stability have long been recognized as three fundamental
parameters for making a concrete of excellent performance. Failure in achieving the desired strength, or
durability, and development of signs of distress, or any doubt in the performance record of a concrete
often calls for detailed investigation. Understanding the cause of failure, however, requires a thorough
understanding of the basic microstructure of the deteriorated concrete. Microscopical examination
provides an in-depth view of the existing microstructure, search for positive signs of deterioration to
confirm the problem, estimates the nature and extent of damage, and provides a report on possible future
performance of the questionable concrete.

The first part of this two-part article described various microscopic techniques routinely used in concrete
failure investigation, and their applications in troubleshooting various problems. Addressed were
problems associated with inadequate quality assurance of aggregates, improper proportioning, improper
mixing, placement, finishing, curing operations, carbonation, and freeze-thaw deteriorations. This part
will discuss the problems like alkali-silica reaction (ASR), alkali-carbonate reaction (ACR), sulfate
attack, corrosion of reinforcing steel, seawater attack, and fire attack. Finally, the effect of the addition of
various mineral and chemical admixtures into the concrete microstructure, their 'secrets' behind making a
concrete more durable, and the possible effects of overdoses will be discussed briefly with the help of
numerous optical and electron photomicrographs.

TROUBLESHOOTING CONCRETE PROBLEMS


(1) Alkali-Silica Reaction (ASR)
If the alkali-content of Portland cement is higher than the optimum limit (-0.6 Na2O equivalent), under
favorable moisture conditions (e.g., greater than 80 relative humidity at 23°C), certain aggregates in
concrete may undergo reaction with cement alkalis and form an alkali-silica reaction gel. Absorption of
moisture by the gel causes expansion, which may lead to cracking in concrete (Fig la) 1. Aggregates
which have potentially deleterious effects on cement alkalis are opaline or chalcedonic silica in different
rocks like chert, flint, siliceous shale, vesicular basalt, or as microscopic veinlets in innocuous carbonate
rocks; strained quartz and other silica polymorphs (tridymite and cristobalite); volcanic glasses like
obsidian and pumice; and glassy to cryptocrystalline matrix of fine grained silicic volcanic rocks like
rhyolite, dacite, andesite, etc 2. Microscopic detection of ASR entails1'^ : (a) detection of alkali-reactive
coarse aggregates, (b) internal radial cracking in coarse aggregates (Fig Ib, d), with wider cracks
emanating from the center of the aggregate becoming narrower towards the margins and extending
towards mortar, (c) presence of reaction rims (abruptly truncated against intersecting air voids) around
the aggregates (Fig Ic), with or without discontinuous circumferential cracks, (d) opaque, or translucent,
or glassy clear to white; massive (dense) optically amorphous (isotropic), spongy (textured, grainy) or
finely crystalline (lamellar or rosette-like, needle or rod-like, or, blade-like crystals 7, Fig 1m, n);
rubbery, waxy, viscous or watery (when fresh), or, hard and brittle (when dried) alkali-silica gels as

1
Dedicated to late Dr. Shondeep L Sarkar for his significant contributions to understanding concrete microstructure.
fracture fillings in aggregates, inside the cracks in paste, as linings or fillings of air voids, as reaction
rims around the aggregates, or as extrusion as gel balls on surfaces^, (e) strong signal of high alkalies in
the reaction rims or in alkali-silica gel deposits in cracks or voids under SEM-EDS 6 (Fig 1 o), (f)
weakening of the paste-aggregate bond due to overall expansion of aggregates (Fig 1 g), (g) map or
pattern cracking of hardened concrete surfaces due to volume increase by direct in-place enlargement of
the affected aggregate by gel formations (Fig 1 a), (h) presence of reactive silica in otherwise nonreactive
carbonate rocks (Fig 1 h), etc. The characteristic strong greenish yellow fluorescence of ASR gel when
treated with uranyl acetate solution and viewed in the dark with short-wave UV light, helps rapid in situ
determination of ASR in the concrete structure (Fig 1 i-1)1. Partial immersion of a polished ASR
affected concrete in water (keeping polished surface up in the air) may help slow exudations of soft gel
onto the dry surface as white microscopic dimples 9. Identification of ASR gel is an unequivocal
evidence of an alkali-silica reaction, but is not necessarily a cause of abnormal expansion or cracking2.
Petrographic examination of hardened concrete is an invaluable tool not only to rapidly identify ASR,
but also to investigate whether the cracking and deterioration in the concrete structure is due mainly to
ASR or, instead, to some other mechanism like freezing and thawing, corrosion of embedded steel,
plastic shrinkage, or chemical attack, etc.

(2) Alkali-Carbonate Reaction (ACR)


Certain dolomitic limestone or calcic dolomite coarse aggregates (containing substantial amounts of both
dolomite and calcite crystals, and clay-rich acid insoluble residue), in the presence of high-alkali cement
can create expansion and cracking in concrete by alkali-carbonate reaction. The typical ACR mechanism
involves initial 'dedolomitization' reaction between the dolomite crystals in carbonate coarse aggregate
with alkali hydroxide in pore solution forming brucite, calcite and alkali-carbonate. Subsequent
expansion, either due to growth and rearrangement of solid high volume reaction products (especially
brucite), or, due to absorption of water and alkali ions into the reaction products and into the exposed
acid insoluble clay residues, can induce cracking in concrete. The typical ACR texture shows relatively
large, rhombic crystals of dolomite [Ca, Mg(CO3)2] set in a finer-grained matrix of calcite [CaCO3], clay
and fine-grained quartz (Figure 2). The dolomite rhombs are usually sharp-edged (less commonly of
rounded corners, or less pronounced rhombic shape), typically 25-50um in size, and are set in a 2-6um
sized fine matrix. The dolomite rhombs may be sparsely distributed and 'float' in the finer matrix, or,
may be crowded together with rhombs adjacent or touching. Reactive carbonates which show very slow
expansion after a considerable period of time have relatively coarser-grained matrices of interlocking
dolomite, calcite, clay and silica minerals than the ones which expand early [11]. Thin sectioning or acid
etching of the highly polished surface can clearly display the characteristic ACR texture.

(3) External Sulfate-attack


Sulfates of sodium, calcium, and magnesium commonly present in soil or groundwater can react with
hydrated phases of exposed hardened concrete (calcium hydroxide, calcium aluminate hydrate, and
calcium silicate hydrate) and form gypsum, ettringite, and sodium or magnesium hydroxides 12'13.
Formation of these solid reaction products result in a significant increase in volume, which may lead to
disruptive expansion and cracking in the concrete. Cracking due to expansion mainly results from
occurrence of gypsum (Fig 3 e) and ettringite (Fig 3 h) in confined spaces in the hardened cement paste
as opposed to open spaces like air voids or preexisting cracks. Moreover, concrete looses its strength due
to loss of cohesion and inherent cementitious properties of the hydrated cement paste. The affected
concrete shows (a) characteristic whitish appearance with soft, powdery cement paste (chalky-texture)
and, (b) abundant tufts of white, acicular or silky crystals of ettringite filling the air voids, cracks and
pores in the paste (Fig 3 a-d). The damage usually starts at the edges and comers and is followed by
progressive cracking and spalling, reducing the concrete to a friable or even soft state. SEM investigation
found many different morphologies of ettringite, from thin needle-like crystals to thick columner crystals
(Fig 3 f-h), to large lath-like crystals14. The needle-like crystals can agglomerate in a confined space,
absorb water, expand, and induce disruption in the paste. To prove the occurrence of external sulfate-
attack, microscopic evaluation, however, should be supplemented by chemical analysis to determine
whether acid soluble sulfate content in the concrete is higher than that present in the original cement. In
addition to the chemical form of sulfate attack, the crystallization pressure resulting from precipitation of
sulfate salts and their subsequent growth in concrete exposed to alternate drying and wetting periods can
cause distress. Carbonation of ettringite (calcium aluminate sulfate hydrate) can form thaumasite
(calcium silicate sulfate carbonate) with further softening, expansion and cracking in paste lj).

(4) Internal Sulfate Attack


Contrary to the external attack of sulfate ions from water or soil, which destroys the inherent
cementitious properties of concrete by attacking the hydrated cement paste and forms ettringite, sulfates
present in cement (gypsum) to control the rate of hydration of C3A can form internal secondary ettringite
in hardened concrete (without decomposing the hydrated cement paste) at late stage of cement hydration.
Internal secondary ettringite can form either by (a) slow dissolution and reprecipitation of primary
ettringite, formed during the early stages of cement hydration in plastic concrete (mostly harmless,
common in concrete which is moist for considerable periods of time, occupies air voids or cracks formed
by other mechanisms) 16, or, by (b) slow leaching of sulfates from the hydrated cement gel bounded in it
during the heat-treatment of a precast concrete (sometimes called "delayed ettringite" - can be destructive
due to formation in hardened paste, or in non air-entrained concrete. Fig 3 f) 15. Controversy remains
whether the observed cracking in a concrete is due to expansion associated with secondary ettringite
formation in paste, or due to some other mechanism (e.g. freeze-thaw, ASR, etc) which has opened
spaces for ettringite to precipitate16. Sufficient moisture, inadequate air voids, high sulfate content in
cement, and high permeability of concrete enhances secondary ettringite related distress. Unless the
external environment is fully appraised for sulfate ions, microscopical observation of secondary ettringite
crystals in an isolated concrete specimen cannot distinguish whether it is due to external or internal
sulfate attack. Moreover, unless some direct ettringite-related distress in the paste is present, ettringite
merely lining the air voids or cracks should be considered harmless.

(5) Seawater-attack
Concrete exposed to marine environment shows various types of chemical deterioration which can be
readily observed under microscopes 12' ^' i7"2^. Following are the examples of four common types of
seawater attack: (a) Dissolved magnesium ion, or CO^ in seawater react with calcium hydroxides in
paste and precipitate brucite, or aragonite, respectively, in the pores at the surface of the concrete.
Continuous dissolution of paste makes it more and more porous (Fig 4 a). Leaching and erosion of
concrete is more severe in seawater as calcium hydroxide is more soluble in seawater than in ordinary
water, (b) Dissolved sodium chloride lowers the pH of concrete, and, as a result increases solubility of
calcium hydroxide, gypsum, monosulfates and ettringite in paste. High permeability of concrete
enhances chloride ions penetration and induces rebar corrosion, (c) During repeated wetting and drying
cycles, especially in the splash zone, dissolved salts (mostly sulfates) in seawater precipitated in the pore
spaces of paste and aggregates during drying, re-hydrate during the wetting period, grow, and exert
expansive forces on the surrounding hardened cement paste (salt weathering. Fig 4 b). (d) Dissolved
magnesium sulfate in seawater can induce sulfate attack by converting calcium hydroxide, calcium
aluminate hydrate, and calcium silicate hydrate into gypsum, brucite, and ettringite (Fig 4 a) but their
formation is not as expansive as expected from the ordinary sulfate attack, because gypsum, as well as
ettringite, are soluble in the presence of chlorides and can be leached out by the seawater.
(6) Corrosion of reinforcing steel in concrete
Penetration of chloride ions into a reinforced concrete of high permeability and absorptivity, or, severe
carbonation of paste from atmospheric C02 (reducing concrete's pH) can cause corrosion of reinforcing
steel 21. Corrosion is an electrochemical process where iron metal goes into solution and forms ferrous
or ferric iron at anode. The released electrons move to the nearby cathode and react with water and
oxygen to form hydroxyl ions. The electrical circuit is completed by movement of OH", Cl" or alkali
ions through the concrete from cathode to anode. Reactions of oxidized iron with OH", Cl" ions form
various types of corrosion products (rust. Fig 5 b), most of which are usually crystalline ferric oxide
(magnetite), and, amorphous ferrous and ferric oxides, hydroxides, chlorides and hydrates, and their
complexes. Formation of solid crystalline or amorphous corrosion products cause a significant increase
in volume eventually leading to scaling, spalling or cracking of the concrete cover around the reinforcing
steel (Fig 5 a). SEM observation of corrosion products shows various morphologies of iron oxides,
hydroxides or choloride complexes, and suggests the exfoliated lamellar product to be responsible for
expansion 22'23 (Fig 5 c, d). SEM investigtaion can also determine the effectiveness of corrosion
inhibitors (e.g. calcium nitrite, zinc oxide in admixture). The compositional and morphological
alterations at the rebar-concrete interface can help clarify the role of an inhibitor in reducing corrosion.

(7) Exposure to Fire


Concrete made with sedimentary or metamorphic coarse aggregates shows permanent change in color
when exposed to fire (goes from normal —> pink -> red —> gray —> black —> buff, as the
temperature increases from ambient to > 900°C) 10. The paste expands under high temperature and then
shrinks, cracks and becomes soft, light colored, and powdery (dehydrated) when cooled. Silica minerals
in siliceous coarse aggregates or natural sand fine aggregates undergo transformation from low to high
temperature form (at 573°C) and induce cracking or aggregate popout by subsequent increase in volume.
Microscopic investigation can accurately determine color changes, paste softness and cracking, and
proportions of high temperature silica forms to estimate the extent and severity of fire-damage. Surface
crazing, deep cracking, popouts over siliceous coarse aggregates, weakening of paste-aggregate bonds
(due to differential thermal expansion), concrete spalling and exposing of reinforcing bars, severe paste
softening (due to dehydration), significant strength reduction are some common physical consequences
of fire-attack.

EFFECTS OF MINERAL & CHEMICAL ADMIXTURES IN CONCRETE MICROSTRUCTURE


Microscopic examination can determine the type and effect of addition of various mineral admixtures on
the properties of concrete. Their influence on hydration kinetics and hydration products can also be
studied fairly accurately, with the possibility of correlating the microstructure with mechanical
properties. Fly ash in concrete can be detected by its distinctly spherical shape, and glassy or amorphous
nature (isotropic) in transmitted-light microscope (Fig 6 a), or in SEM/BSE image (Fig 6 b-d). Electron
microscopy has long been an important tool in studing hydration and pozzolanic activity of fly ash, and
its effect on paste microstructure, permeability and engineering properties of concrete (Fig 6 e-h) 24-35.
Granulated blast furnace slag appears as shard-like particles of glass, usually larger than unhydrated
cement particles. The total glass content determines its reactivity, and can be estimated from optical
microscopy or SEM. In newly hardened concrete, slag particles display characteristic bright orange to
orange-red fluorescence when viewed in dark under short-wave UV light. Concrete containing
substantial amounts of slag show characteristic dark green or blue-green colors and emit a typical odor of
hydrogen sulfide when freshly broken or wetted. These characteristic fluorescence, color, and odor,
however, disappear as the concrete ages 9. SEM or BSE imaging can be used to successfully determine
the amount and distribution of slag in hardened cement paste of any age 2^' ^' 34 (Fig 7 a-d). Under
SEM, a distinct reaction rim appears around the slag particles which become more vivid as hydration
progresses (Fig 7 b). The high magnification required to detect silica fume makes electron microscopes
ideal for quantity and distribution analysis in hardened concrete. Fig 8 a-d). BSE and SEM imaging
showed that increasing amounts of silica fume reduce the amounts of calcium hydroxide, the overall
microstructure becomes much more uniform, and the aggregate-paste interfaces become significantly
tighter (Fig 8c & 9 a,b) - all due to pozzolanic reaction between calcium hydroxide and silica fume and
an improvement in packing density due to the fine particle size of silica fume (all effects are beneficial to
concrete) ^9' ^' ^' 34 . Large overdoses of set retarding chemical admixtures can be identified from
unusual abundances of unhydrated cement particles (UPC's) in thin section. Overdoses of air-entraining
admixtures can be easily seen from excessive air-voids under optical microscopes. Large inadvertent
overdose of superplastisizers can be identified from distinct microstructural changes in BSE-images
(increased area fraction of UPCs, decreased area fraction and size of capillary pores in paste, and
increased size and area fraction of calcium hydroxide grains 36, Fig 9 c, d). Forensic investigations of the
affected concrete using microscopy, proves invaluable in ascertaining admixture overdose-related
problems like retardation or strength loss.

CONCLUSION
The above two-part discussion on the importance of microscopy in troubleshooting concrete problems
enumerates several physical and chemical deterioration processes which can be detected, identified, and
even quantified on occasion, by microscopic investigation. High permeability of concrete enhances
various chemical attacks from the external environment (e.g. sulfate or seawater attack); inferior quality
of ingredients threaten concrete's performance; improper proportioning, compaction, mixing, placement
and curing operations reduce concrete's strength and durability; reactive aggregates induce expansion and
cracking. The authors have demonstrated how microscopical investigation can provide valuable insight
into the inherent microstructure of concrete which has significant control on concrete failure. It is also
possible to study problems in specialized concretes, like fiber-reinforced concrete, high strength
concrete, or concrete made with lightweight aggregates, however, they are beyond the scope of this
article. As technology advances, more and more innovative investigative tools will appear in the market,
but, microscopic examination still remains at the nucleus of available failure investigative methods for
providing quick, cost-effective, and definitive answers to concrete problems for the engineering
community. Engineers, let's shake hands and build concrete structures for ultimate performance !

ACKNOWLEDGMENTS
Sincere thanks are extended to the following individuals and organizations for providing various
photographs or permission to use photographs included in this article: Michael A. Ozol, David Stark,
James A. Ray, Paul Stutzman, Patsy Harms (Portland Cement Association), Ronald Sturm (Construction
Technology Laboratories), Walter W. Rowe (International Cement Microscopy Association), and Kathe
Hooper (American Society for Testing and Materials).

REFERENCES
1. Stark, D., Handbook for the Identification of Alkali-Silica Reactivity in Highway Structures,
SHRP-C/FR-91-101, Strategic Highway Research Program, National Reseacrh Council,
Washington, DC, 1991,45pp.
2. Stark, D., "Alkali-Silica Reactions in Concrete", Significance of Tests and Properties of Concrete
and Concrete-Making Materials, P. Klieger and J.F. Lamond eds., ASTM STP 169C, American
Society for Testing and Materials, Philadelphia, PA, 1994, pp. 365-371.
3. Regourd, M., and Homain, H., "Microstructure of Reaction Products", Concrete Alkali-
Aggregate Reactions, P.E. Grattan-Bellew ed., Noyes Publications, New Jersey, 1987, pp. 509.
4. Proceedings of the 8th International Conference on Alkali-Aggregate Reaction, Kyoto, Japan,
1989.
5. Shayan, A., "Reexamination ofAARin an old concrete", Cem. Concr. Res., Vol 19, 1989, pp. 434-
442.
6. Sarkar, S.L., and Vogt, W.L., "An incidence of alkali-silica reaction under highly unusual
circumstances", Proceedings of 18th International Conference on Cement Microscopy, L. Jany,
A. Nisperos, and J. Bayles eds., 1996, pp. 404-415.
7. Helmuth, R., Alkali-Silica Reactivity: An overview of Research, Strategic Highway Research
Program, National Research Council, 1993,105pp.
8. Mielenz, R.C., "Petrography Applied to Portland-Cement Concrete", Reviews in Engineering
Geology, Geological Society of America, Vol 1, 1962, pp. 1-38.
9. Ray, J.A., "Things Petrographic Examination Can and Cannot Do With Concrete", Proceedings of
the 5th International Conference on Cement Microscopy, Nashville, ICMA, 1983, pp. 66-84.
10. ASTM C 856 "Petrographic Examination of Hardened Concrete", in Annual Book of ASTM
Standards, Section 4 - Construction, Vol 04.02-Concrete and Aggregates, American Society for
Testing and Materials, Philadelphia, PA, 1993.
11. Ozol, M.A., "Alkali-Carbonate Reaction", Significance of Tests and Properties of Concrete and
Concrete-Making Materials, P. Klieger and J.P. Lamond eds., ASTM STP 169C, American
Society for Testing and Materials, Philadelphia, PA, 1994, pp. 372-387.
12. Depuy, G.W., "Chemical resistance of concrete". Significance of Tests and Properties of Concrete
and Concrete-Making Materials, P. Klieger and J.F. Lamond eds., ASTM STP 169C, American
Society for Testing and Materials, Philadelphia, PA, 1994, pp. 263-281.
13. Neville, A.M., Properties of Concrete, John Wiley and Sons, Inc., 4th Ed, 1996, 844pp.
14. Mehta, P.K., "Scanning electron micrographic studies ofettringite formation", Cem. Concr. Res.,
vol 6, No 2,1976,pp.169-182.
15. Day, R.L., The effect of Secondary Ettringite formation on the Durability of Concrete : A
Literature Analysis, Research and Development Bulletin RD108T, Portland Cement Association,
Skokie, Illinois, USA, 1992, 115pp.
16. Kosmatka, S., "Ettringite and Concrete Durability", Concrete Technology Today, Vol 17, No 3,
Portland Cement Association, 1996, pp. 7-8
17. Mehta, P.K., "Durability of Concrete in Marine Environment - A Review", Performance of
Concrete in Marine Environment, V.M. Malhotra ed., ACI SP-65,1980, pp. 1-19.
18. Regourd, M., "Physico-chemical studies of cement pastes, mortars and concretes exposed to
seawater", Performance of Concrete in Marine Environment, V.M. Malhotra ed., ACI SP-
65,1980, pp. 63-81.
19. Sarkar, S.L., "The Importance of Microstructure in Evaluating Concrete", Concrete Technology :
Past, Present and Future, V.M. Malhotra ed., 1994, pp 125-160.
20. Sarkar, S.L., and Aitcin, P.C., "Phenomenological Investigation of Concrete Deterioration in a
Median Barrier", Cem. Concr. Res., Vol 21,1991, pp. 917-927.
21. Perenchio, W.F., "Corrosion of reinforcing steel", Significance of Tests and Properties of
Concrete and Concrete-Making Materials, P. Klieger and J.F. Lamond eds., ASTM STP 169C,
American Society for Testing and Materials, Philadelphia, PA, 1994, pp. 164-172.
22. Glasser, P.P., and Sagoe-Crentsil, K..K., "Steel in Concrete: Part II. Electron Microscopy
Results", Magazine of Concrete Research, Vol 41,1989, pp. 213-220.
23. Marchese, B., "Microstructural alterations of cement paste - steel interface exposed to seawater",
K. and B. Mother International Conference on Concrete Durability, J.M. Scalon ed., ACI SP100-
81,1987, pp. 1611-1633.
24. Diamond, S., "The microstructure of cement paste in concrete". Proceedings of the VIII th
International Congress on the Chemistry of Cement, Rio de Janeiro, 1986, pp. 122-147.
25. Uchikawa, H., "Effect of Blending component on hydration and structure formation",
Proceedings of the VIII th International Congress on the Chemistry of Cement, Rio de Janeiro,
1986, pp. 250-280.
26. Marusin, S., "Microstructure of Fly Ash Concrete - Part I. Effects of Calcium Sulfite and Calcium
Sulfate", Proceedings of the VIII th International Conference on Cement Microscopy, Orlando,
ICMA, 1986, pp. 52-70.
27. Scali, M.J., Chin, D., and Berke, N.S., "Effect of Microsilica and Fly Ash upon the microstructure
and permeability of concrete". Proceedings of the IXth International Conference on Cement
Microscopy, Reno, ICMA, 1987,pp.375-397.
28. Stutzman, P.E., "Application of Scanning Electron Microscopy in Cement and Concrete
Petrography", Petrography of cementitious materials, S. DeHays, and D. Stark eds., ASTM STP
1215, American Society for Testing and Materials, Philadelphia, PA, 1994, pp. 74-90.
29. Scrivener, K.L., Bentur, A., and Pratt, P.L., "Quantitative characterization of the Transition zone
in High Strength Concretes", Adv. Cem. Res., Vol 1,1988, pp. 230-237.
30. Bentz, D.P., Stutman, P.E., and Garboczi, E.J., "Experimental and simulation studies of the
interfacial zone in concrete", Cem. Concr. Res., Vol 22,1992, pp. 891-902.
31. Marusin, S.L., "Microstructure and Pore Characteristics of Concrete containing condensed silica
fume with a superplastisizer", Proceedings of the VIII th International Conference on Cement
Microscopy, Orlando, ICMA, 1986,pp.327-335.
32. Grutzeck, M.W., Roy, D.M., and Wolfe-Confer, D., "Mechanism of hydration of Portland
Cement composites containing ferrosilicon dust", Proceedings of the IV th International
Conference on Cement Microscopy, Nevada, ICMA, 1982, pp. 193-202.
33. Sarkar, S.L., "The role of Silica Fume, Slag, and Fly ash in the microstructural development of
High Performance Concrete Microstructure", Proceedings of the ACI International Conference on
High Strength Concrete, Singapore, ACI SP 149, American Concrete Institute, Detroit, 1994, pp
449-460.
34. Baalbaki, M., Sarkar, S.L., and Aitcin, P.C., "Properties and micrstructure of High Performance
Concrete containing Silica Fume, Slag, and Fly ash". Proceedings of the 4th CANMET/ACI
International Conference on Fly Ash, Slag, Silica Fume and other Natural Pozzolans in Concrete,
Istanbul, ACI SP 139, V.M. Malhotra ed., American Concrete Institute, Detroit, 1992, pp. 921-
942.
35. Sarkar, S.L., "High Strength Concrete and its Microstructure", Cement and Concrete Science and
Technology, Vol 1, Part I, S.N. Ghosh ed., ABI Books Pvt. Ltd., New Delhi, 1991, pp. 380-417.
36. Diamond, S., and Wang, Y., "A quantitative image analysis study of the influence of
superplastisizer on cement paste microstructure". Proceedings of 18th International Conference
on Cement Microscopy, L. Jany, A. Nisperos, and J. Bayles eds., 1996, pp. 465-479.
Figure Captions : Troubleshooting Concrete Problems through Microscopy : Part II
Fig 1 : Alkali-Silica Reaction in Concrete : (a) A well-defined map or pattern cracking associated with
the development of ASR in highway pavement; (b) Cracking in coarse aggregate, extending to the paste
due to ASR; (c) Reaction rim around coarse aggregate, white secondary deposits on the surface are ASR
products; (d) Subparallel cracking in paste due to ASR and drying shrinkage; (e) Cracking in chalcedonic
limestone coarse aggregate due to ASR; (f) an expanded view of the box showing chalcedonic silica - a
potentially reactive mineral to cement alkalies; (g) Weak paste aggregate bonding due to expansion of
reactive aggregate, also shown are two small ASR gel balls; (h) chalcedonic fingers in limestone - a
potentially reactive aggregate; (i) Fractured surface of concrete pavement core showing location of
reactive aggregate (granite gneiss), as photographed in ordinary light. There is no positive indication of
ASR gel on this surface in ordinary light; (j) Same surface after uranyl acetate treatment. Green and
bright yellow areas display ASR gel. Note band of ASR gel along periphery of aggregate (arrow) while
interior is free of reaction product; (k) Badly cracked coarse aggregates in concrete affected by ASR; (1)
same sample after treatment with uranyl acetate solution, positively identifying peripheral green film of
ASR gel around coarse aggregates. (m) SEM image of rosette-structured crystalline ASR product; (n)
SEM image of cracked ASR gel; (o) SEM-EDS spectrum (with corresponding oxide compositions) of
ASR gel showing peaks of alkalis (Na, K) and silica (Si).

Fig 2 : Alkali-Carbonate Reaction in Concrete : (a) A nonreactive carbonate aggregate showing a mosaic
of equigranular dolomite cryatals; (b, c) Reactive carbonate showing large rhoms of dolomite floating in
a fine-grained matrix ofcalcite, clay and quartz; (d) Alternating bands of reactive and nonreactive
carbonate rocks.

Fig 3 : Sulfate-attack in Concrete : (a, b) Ettringite needles filling air-voids in concrete; (c) Ettringite
needles lining the margins of cracks; (d) an X-ray elemental image of the same photo in 'c' showing
distribution of sulfur in ettringite; (e) SEM image of massive gypsum formation due to external sulfate-
attack in a concrete; (f) Massive delayed ettringite formation in a precast concrete pier; (g) SEM image
of ettringite needles at the paste-aggregate interface (A = aggregate, B = CH crystals, C = ettringite); (h)
SEM image of ettringite needles in hardered cement paste.

Fig 4 : Seawater -attack in Concrete : (a) SEM image showing increased porosity in paste followed by
ettringite formation in a concrete exposed to marine environment; (b) Salt weathering - SEM image of
chloroaluminate crystal (+) formation due to precipitation of dissolved salts from seawater.

Fig 5 : Rebar Corrosion : (a) Spalling of concrete bridge deck due to corrosion of underlying reinforcing
steel; (b) severe rusting of rebar and deposition of iron oxide and hydroxide products along the concrete-
rebar interface; (c) SEM image of globular and whiskery corrosion products that have burst open to
expose a floral arrangement; (d) SEM image of exfoliated lamellaes of Fe-oxide or hydroxide at the
rebar-paste interface.

Fig 6 : Fly ash in Concrete : (a) Thin section photomicrograph showing spherical residual fly ash
particles in paste (the black spheres are completely filled with carbon); (b) SEM image of fly ash
particles in a polished section of concrete; (c) Dense paste microstructure at 28-days in a low W/C high
performance concrete showing unreacted spherical fly ash particles in the paste; (d) Liesengang ring
structure (C-S-H located radically around the hydrated fly ash). The concrete microstructure represents
alternate layers ofCH and C-S-H. SEM images of successive stages ofhydration of fly ash particles - (e)
fly ash just began to hydrate; (f) at the intermediate stage of hydration; (g) floral fly ash at the advanced
stage of hydration; (h) Relict structure of a completely hydrated fly ash.

Fig 7 : Slag in Concrete : (a) SEM images of slag particles in concrete (1 = glassy phase; 2 = crystalline
phase); (b) A reaction rim (•) around a slag particle (A); (c) BSE image of angular slag particles in paste,
larger than the UPCs; (d) Slag particle in a state of hydration (surrounded by hydration rim).

Fig 8 : Silica Fume in Concrete : (a) SEM image of fresh silica fame particles; (b) TEM image of fresh
silica fame particles; (c) Strong paste-aggregate bonding in silica fame concrete; (d) TEM image of
hydration of silica fame.

Fig 9 : Comparative microstructure of normal and admixture doped concrete/mortar - (a) and (b) showing
normal and 20 silica fame doped hardened cement mortars (both 24 hours old, W/C = 0.45), respectively,
showing dense, compact, much more uniform microstructure with few CH crystals in the silica fame-
bearing mortar than those occurring in the normal one. (c) and (d) showing marked effect of
superplastisizer on paste microstructure. Compared to plain paste in (c), addition of superplastisizer
showed increased area fraction of UPCs, and decreased porosity of paste in (d).
Figure 1: Alkali-Silica Reaction in Concrete: (a) A well-defined map or pattern cracking associated with the
development of ASR in highway pavement; (b) Cracking in coarse aggregate, extending to the paste due to ASR; (c)
Reaction rim around coarse aggregate, white secondary deposits on the surface are ASR products; (d) Cracking in
concrete due to ASR.
Figure 1: Alkali-Silica Reaction in Concrete (continued): (i) Fractured surface of concrete pavement core showing
location of reactive aggregate (granite gneiss), as photographed in ordinary light. There is no positive indication of ASR
gel on this surface in ordinary light; (j) Same surface after uranyl acetate treatment. Green and bright yellow areas
display ASR gel. Note band of ASR gel along periphery of aggregate (arrow) while interior is free of reaction product;
(k) Badly cracked coarse aggregates in concrete affected by ASR; (1) same sample after treatment with uranyl acetate
solution, positively identifying peripheral green film of ASR gel around coarse aggregates [Photo curtsey of David
Stark].
Figure 1: Alkali-Silica Reaction in Concrete (continued): (m) SEM image of rosette-structured
crystalline ASR product; (n) SEM image of cracked ASR gel; (o) SEM-EDS spectrum (with
corresponding oxide compositions) of ASR gel showing peaks of alkalies (Na, K) and silica (Si).
Figure 2: Thin section photomicrographs of alkali-carbonate reactive aggregates: (a) Thin-section photomicrograph of a
non-reactive carbonate aggregate showing the typical dolomitic texture consisting of a mosaic of equigranular dolomite
crystals (top left); (b, c) Reactive carbonate aggregate showing large rhombic crystals of dolomite set in a finer-grained
matrix of calcite, clay and quartz; (d) Alternating bands of reactive and non-reactive carbonate rocks in a quarry
dolomite (bottom right). Field widths of all photos are 1 mm. [Photo curtsey of Michael Ozol].
Figure 3: The top photos are thin section photomicrographs of a normal concrete exposed to moisture for prolonged
periods, which has caused precipitation of acicular secondary ettringite crystals in air voids. Such a precipitation of
secondary ettringite in voids does not indicate any sulfate attack. The bottom photos [curtsey of Paul Stutzman] are
back scatter electron image (left) and sulfur-elemental mapping of the left image (right) in an SEM-EDS showing the
presence of ettringite in a crack, which has formed due to external sulfate attack in the concrete.
Figure 3: Sulfate-attack in Concrete (continued) - (e) SEM secondary electron image of massive gypsum
formation due to external sulfate-attack in a concrete; (f) Massive delayed ettringite formation in a
precast concrete pier; (g) SEM image of ettringite needles at the paste-aggregate interface (A =
aggregate, B = CH crystals, C = ettringite); (h) SEM image of ettringite needles in hardened cement
paste.
Figure 4: Seawater -attack in Concrete: (a) SEM image showing increased porosity in paste followed by
ettringite formation in a concrete exposed to marine environment; (b) Salt weathering - SEM image of
chloroaluminate crystal (+) formation due to precipitation of dissolved salts from seawater.
Figure 5: Corrosion of reinforcing steel in concrete: (a) Spalling of concrete bridge deck due to corrosion of underlying
reinforcing steel; (b) severe rusting of rebar and deposition of iron oxide and hydroxide products along the concrete-rebar
interface [curtsey of Ronald Strum]; (c) SEM image of globular and whiskery corrosion products that have burst open to
expose a floral arrangement; (d) SEM image of exfoliated lamellas of Fe-oxide or hydroxide at the rebar-paste interface.
Figure 6: Fly ash in Concrete: (a) Thin section photomicrograph showing spherical residual fly ash
particles in paste (the black spheres are completely filled with carbon); (b) SEM image of fly ash
particles in a polished section of concrete; (c) Dense paste microstructure at 28-days in a low W/C high
performance concrete showing unreacted spherical fly ash particles in the paste; (d) Liesengang ring
structure (C-S-H located radically around the hydrated fly ash).
Figure 6: Fly ash in Concrete (continued): (e) fly ash particles just began to hydrate; (f) at the
intermediate stage of hydration; (g) floral fly ash at the advanced stage of hydration; (h) Relict structure
of a completely hydrated fly ash.
Figure 7: Slag in Concrete: (a) SEM images of slag particles in concrete (1 = glassy phase; 2 =
crystalline phase); (b) A reaction rim (•) around a slag particle (A); (c) BSE image of angular slag
particles in paste, larger than the UPCs; (d) Slag particle in a state of hydration (surrounded by hydration
rim).
Figure 8: Silica Fume in Concrete: (a) SEM image of fresh silica fame particles; (b) TEM image of fresh
silica fame particles; (c) Strong paste-aggregate bonding in silica fame concrete; (d) TEM image of
hydration of silica fame.
Figure 9: Comparative microstructure of normal and admixture doped concrete/mortar - (a) and (b)
showing normal and 20 silica fame doped hardened cement mortars (both 24 hours old, W/C = 0.45),
respectively, showing dense, compact, much more uniform microstructure with few CH crystals in the
silica fame-bearing mortar than those occurring in the normal one. (c) and (d) showing marked effect of
superplasticizers on paste microstructure. Compared to plain paste in (c), addition of superplasticizers
showed increased area fraction of UPCs, and decreased porosity of paste in (d).

You might also like