You are on page 1of 12

Construction and Building Materials 162 (2018) 253–264

Contents lists available at ScienceDirect

Construction and Building Materials


journal homepage: www.elsevier.com/locate/conbuildmat

Characterization of concrete affected by delayed ettringite formation


using the stiffness damage test
Eric R. Giannini a,⇑, Leandro F.M. Sanchez b, Atolo Tuinukuafe c, Kevin J. Folliard d
a
RJ Lee Group, 350 Hochberg Road, Monroeville, PA 15146, USA
b
Department of Civil Engineering, University of Ottawa, 161 Louis-Pasteur, K1N 6N5 Ottawa, ON, Canada
c
Department of Civil, Construction and Environmental Engineering, The University of Alabama, Box 870205, Tuscaloosa, AL 35487-0205, USA
d
Department of Civil, Architectural and Environmental Engineering, The University of Texas at Austin, 301 E Dean Keeton St, Stop C1700, Austin, TX 78712-0273, USA

h i g h l i g h t s

 The Stiffness Damage Test (SDT) can reliably assess DEF-damaged concrete.
 SDI and PDI indices have a near-linear correlation to DEF expansions up to 0.40%.
 SDT may characterize DEF damage more effectively than ASR damage.
 The statistical significance of SDT analyses by SDI and PDI is confirmed by ANOVA.
 Future work should consider combined ASR and DEF damage, and correlating SDT to DRI.

a r t i c l e i n f o a b s t r a c t

Article history: The stiffness damage test (SDT) may provide more information than conventional mechanical tests used
Received 28 March 2017 to assess concrete affected by expansive reactions. A new approach to analyzing SDT data involving the
Received in revised form 14 October 2017 calculation of stiffness damage and plastic damage indices (SDI and PDI) has successfully characterized
Accepted 2 December 2017
the expansion of concrete affected by alkali-silica reaction (ASR). This study is the first to implement
Available online 15 December 2017
SDI and PDI to characterize concrete affected by delayed ettringite formation (DEF). The SDI and PDI
parameters have a nearly linear relationship to expansions under 0.40%. ANOVA confirms the statistical
Keywords:
significance of the SDI and PDI parameters and that they provide information not available from standard
Concrete
Stiffness damage
elastic modulus tests.
DEF Ó 2017 Elsevier Ltd. All rights reserved.
Delayed ettringite formation
ASR
Damage characterization

1. Introduction cient load-carrying capacity. Compressive strength, though an


important indicator of concrete quality and a critical input for
Extraction and testing of core samples from concrete structures structural analyses, is not affected as much as stiffness, tensile
is extremely useful in the diagnosis and evaluation of structures strength, or flexural strength in concrete suffering from internal
affected by durability-related problems, such as alkali-aggregate swelling reaction mechanisms (ISR) such as AAR and DEF [1–3].
reactions (AAR), cyclic freezing and thawing damage, and delayed Compressive strength testing is also destructive to the specimen
ettringite formation (DEF). Core sampling is akin to taking a biopsy and precludes also performing petrographic examination of the
in medicine – an invasive procedure that generally does little harm same specimen. The stiffness damage test (SDT), however, was
to the structure. Cores are typically subjected to mechanical tests found to provide useful information on the degree of damage in
and petrographic examination to determine the type and severity concrete, such as reduction in stiffness and extent of microcracking
of the distress, and to determine whether the structure has suffi- damage; it is thus considered a more complete technique than a
simple elastic modulus test [4,5]. Moreover, it has been found that
when loads of 40% or less of the compressive strength of the con-
⇑ Corresponding author. crete area used, the SDT remains a non-destructive test of the con-
E-mail addresses: EGiannini@rjleegroup.com (E.R. Giannini), Leandro.Sanchez@ crete, even at very high levels of deterioration from ISRs [5]. Hence,
uottawa.ca (L.F.M. Sanchez), atuinukuafe@crimson.ua.edu (A. Tuinukuafe), folliard@
the SDT can be coupled with further chemical or microscopic
mail.utexas.edu (K.J. Folliard).

https://doi.org/10.1016/j.conbuildmat.2017.12.012
0950-0618/Ó 2017 Elsevier Ltd. All rights reserved.
254 E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264

evaluations performed on the same specimens, to obtain maxi- at the maximum load; this was thought to account for the orienta-
mum information about the condition of the structure [4–11]. It tion of cracking in damaged concrete.
is worth noting that the non-destructive character of the SDT is Further development of the test method by Smaoui et al. [14]
an asset, although the most important aspect of this mechanical resulted in a recommended loading level of 10 MPa and identified
testing procedure is its ability to quantify the amount of inner dis- the area of the first hysteresis loop and the accumulated plastic
tress of damaged concrete, which requires the application of loads strain over all five cycles as the most important parameters. They
of at least 30% (ideally 40%) of the compressive strength of the proposed that a linear relationship between these parameters
concrete. and ASR expansion could be established using laboratory speci-
SDT attempts to quantify damage to concrete from AAR and mens or core samples extracted from larger specimens of known
other mechanisms by analyzing stress vs. strain data obtained expansion levels [14,15]. They also noted that concrete made with
when cyclic loads are applied to core samples. Several versions of different reactive aggregates will exhibit varying responses in the
the test and associated data analyses have been proposed, but stiffness damage test; that is, linear relationships must be estab-
the basic principle remains the same; compared to sound concrete, lished for multiple reactive aggregate types in order to estimate
damaged concrete will have a reduced stiffness and will accumu- the expansion of a variety of field structures. Fig. 1 shows typical
late more plastic strain and thus dissipate more energy during stress–strain data obtained using this version of the test for an
the test. While the use of SDT as a tool for characterizing the effects undamaged sample and one damaged by ASR.
of AAR on concrete structures has been widely-described and The 10 MPa version of the test has been adopted by the US Fed-
reported in the literature [5,12–17] and most extensively for eral Highway Administration [25]. Several papers describe the
alkali-silica reaction (ASR), there are only limited reports of its application of this version of the SDT to field structures damaged
use for characterizing DEF-affected concrete [4,18]. Recent by ASR [26–29], but no studies have been published regarding
research on the use of SDT for ASR-damaged concrete has resulted the use of SDT to characterize damage from DEF or a combination
in new methods of data analysis [5,8,10,11], which are now applied of ASR and DEF using the updated procedure and analysis devel-
to prior SDT data generated for DEF-affected concrete in this paper. oped by Smaoui et al.
DEF is a form of internal sulfate attack in concrete, driven by More recent works by Sanchez et al. [5,8,10,11,30] suggest that
curing temperatures in excess of 65–70 °C and unfavorable cement the maximum applied stress in the test should not be a fixed value
chemistry [19–21]. Cement hydration and formation of C-S-H is because useful data can only be obtained if it is at least 30% of the
greatly accelerated with increased curing temperatures. With sus- compressive strength. If the load is less than 30% of the compres-
tained temperatures above 65 to 70 °C, ettringite becomes thermo- sive strength, which would be greater than 10 MPa for concrete
dynamically unstable; hydration reactions are unable to form stronger than 33.3 MPa, and greater than 5.5 MPa for all but the
ettringite, and previously-formed ettringite decomposes and weakest concretes, then it is very difficult to quantify ASR damage
returns to solution [22]. The rapidly-growing ‘‘inner” C-S-H traps with SDT. As a result, Sanchez et al. [10,11] recommended conduct-
dissolved sulfates and alumina before they can react to form ettrin- ing the test at 40% of the current strength, which would be the
gite [20,21,23]. After temperatures decrease to levels more com- same load specified for the determination of the static modulus
monly experienced by concrete in service, thermodynamics again of elasticity by ASTM C469 [31].
favor the formation of ettringite. Trapped sulfates and alumina
may be released from the C-S-H and react with water and mono- 3. Scope and research significance
sulfate to form ettringite; this can lead to deleterious expansion
and cracking of the concrete [20]. Similar to AAR, the microscale This paper presents a study of the use of SDT to characterize
characteristics of DEF include extensive microcracking. In contrast two sets of concrete cylinders, made with different aggregates
to AAR (especially ASR), where microcracking extends through and conditioned to increasing levels of expansion and damage
both the paste and aggregate particles, microcracking from DEF is from DEF. The results of standard secant modulus of elasticity tests
primarily limited to the cement paste, which expands away from and compressive strength tests on the same specimens are also
the aggregate particles. The resulting gaps and microcracks are presented and analyzed for their utility in characterize the pro-
often filled with recrystallized ettringite deposits instead of ASR gress of damage. The notable improvement over earlier work to
gel. More importantly, because DEF damage development is not characterize DEF-affected concrete is the use of analytical tools
intrinsically linked to the mineralogy of the aggregates or the dis- developed by Sanchez and coworkers that seek to correlate SDT
tribution of reactive phases, as is the case in ASR [24], it is possible damage indices to petrographic damage features quantified
that SDT could be more generally useful as a tool to diagnose and through the Damage Rating Index (DRI) [5–11,32]. The SDT data
characterize the extent of damage from DEF. are also analyzed using tools previously developed by Smaoui
et al. [14]. Finally, ANOVA analysis of the data is presented to iden-
tify the most significant variables influencing the test results.
2. Background on SDT
4. Materials and methods
The stiffness damage test (SDT) was originally developed by
Crouch and Wood [12] and involved applying five cycles of 5.5 Two sets of 100  200 mm cylinder specimens were fabricated for this study.
MPa compressive load to a core specimen and measuring its The specimens made use of the same concrete mixtures used by Giannini and Fol-
stress–strain response. In ASR-damaged concrete, the elastic mod- liard [16] in a parallel study on the mechanical properties of ASR-affected concrete;
however, the conditioning regime (described in Section 4.2) was designed to limit
ulus decreases, while the stress–strain hysteresis loops increase in the development of ASR and promote the development of DEF.
size, and increasing amounts of plastic strain accumulate during
the course of the test [12,13]. Chrisp et al. [13] placed emphasis 4.1. Materials and mixture proportions
on the elastic modulus, plastic strain and size of the hysteresis
loops of the second through fifth cycles and largely discarded the Table 1 presents the aggregates used in this study, and Table 2 shows the pro-
data from the first load cycle, primarily because they believed that portions for the two concrete mixtures. Mixture 1 contained aggregates C1 and F1,
and Mixture 2 contained aggregates C2 and F2. Aggregates C1 and F2 are considered
the first cycle induced new damage in the concrete. They also pro- non-reactive with regards to ASR; the 1-year expansion in ASTM C1293 testing for
posed calculating a non-linearity index (NLI), defined as the secant these two aggregates tested together is 0.01% [20]. When tested per ASTM C1293,
modulus at half the maximum load divided by the secant modulus reactive aggregates C2 and F1 have been documented to exhibit 1-year expansions
E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264 255

12 12

10
Expansion 10
Expansion
0.009% 0.272%
8 8

Stress (MPa)
Stress (MPa)
6 6

4 4

2 2

0 0
0 200 400 600 800 1000 0 200 400 600 800 1000
Axial Strain ( m/m) Axial Strain ( m/m)
Fig. 1. Typical stiffness damage test data for undamaged (left) and ASR-damaged (right) concrete [4]. Test was conducted per the procedure recommended by Smaoui et al.
[14].

Table 1
Details of aggregates used in the study.

Aggregate Source Mineralogy and Type Oven-dry SG Absorption (%)


C1 (coarse) San Antonio, Texas Crushed limestone 2.50 2.72
C2 (coarse) Bernalillo, New Mexico Partially crushed gravel with rhyolite and mixed volcanics 2.56 1.47
F1 (fine) El Paso, Texas Natural sand with quartz, feldspars, siliceous volcanics, and chert 2.59 0.69
F2 (fine) San Antonio, Texas Manufactured limestone sand 2.42 3.77

Table 2
Mixture proportions. Mixture 1 used aggregates C1 and F1 and Mixture 2 used
aggregates C2 and F2. Aggregate quantities are in the saturated surface-dry condition.

Quantity, kg/m3
Component Mixture 1 Mixture 2
Coarse aggregate 1107 1161
Fine aggregate 626 561
Cement 420 420
Water (w/c = 0.42) 176 176
NaOH 1.90 1.90

of 0.16% and 0.59%, respectively [20]. An ASTM C150 Type I Portland cement with an
equivalent alkali content of 0.90% Na2Oeq was used for both mixtures, and the alkali Fig. 2. Curing temperature profile to promote development of DEF.
content of the mixture was boosted to 1.25% by mass of cement through the addi-
tion of 50% w/w NaOH solution to the mixing water. This enabled the fabrication of
ASR- and DEF-affected concrete from a single mixture.
The three cylinders from each mixture designated for expansion measurements
were instrumented immediately following demolding, after completing the 24-h
4.2. Specimen preparation high-temperature curing cycle. A drill press was used to drill out a small opening
on each end, and 6.4-mm- diameter stainless steel gauge studs were inserted and
Concrete was mixed per ASTM C192 and a total of 45 cylinders were cast (24 for bonded with epoxy; while the epoxy hardened, the specimens were wrapped in
Mixture 1 and 21 for Mixture 2). Three cylinders from each mixture were tested for plastic to avoid moisture loss.
28-day compressive strength, and three cylinders were instrumented as reference The change in length of the instrumented cylinders was measured in reference
specimens for expansion measurements. The remaining 18 cylinders from Mixture to a 173-mm- long Invar bar using a digital comparator with 0.002 mm resolution.
1 and 15 cylinders from Mixture 2 were not instrumented; their expansions were Expansions were measured periodically and the average expansions for each set of
inferred based on the average expansions of the reference cylinders for each mix- instrumented specimens were assumed to be representative of the average expan-
ture. Including the three reference cylinders for each mixture, three cylinders each sions of the non-instrumented specimens. When the expansions of the reference
were tested at seven levels of expansion for Mixture 1 and six levels of expansion specimens reached a target level (based on intervals given below), a set of three
for Mixture 2. non-instrumented cylinders was subjected to the SDT (followed by a standard
To promote development of DEF, the cylinder molds were sealed and the spec- secant modulus and compressive strength test). An initial set of specimens for each
imens subjected to a high-temperature curing regime developed by Kelham [19] mixture was tested at seven days of age to provide baseline (undamaged) data; sub-
and illustrated in Fig. 2. Specimens were placed in a ramping oven programed to sequent sets would be tested as close as possible to expansion intervals of 0.10% for
increase the air temperature from 23 °C to 90 °C over a period of 4 h, hold at 90 Mixture 1 (until approximately 0.50%) and 0.08% for Mixture 2 (until approximately
°C for 12 h and decrease to 23 °C over a period of 4 h. The oven was not moisturized; 0.32%), and the final, instrumented set was tested at the end of the experimental
the seal on the molds was considered to be adequate to minimize drying during program.
high-temperature curing. Although the internal temperature of the specimens In preparation for the mechanical tests (including SDT), the ends of the speci-
was not monitored during this process, upon removal from the oven, the exterior mens were ground plane and parallel using a cylinder end-grinder on the day the
surface temperature of the cylinder molds was measured to be 41 °C, which sug- tests were conducted. After grinding the ends and between each test, the specimens
gests that the specimen temperatures lagged the air temperature profile somewhat were stored in water to maintain them in a moisture condition in accordance with
during the ramping periods. Given that the air temperature was held at 90 °C for 12 ASTM C469 [31] and ASTM C39 [34].
h, it is assumed that the specimens reached and maintained that temperature for
much of that time period. Specimens were demolded and moist-cured at 23 °C
for an additional 24 h and then stored in saturated limewater at 23 °C. The limewa- 4.3. Stiffness damage test
ter promotes leaching of alkalis from the cylinders, thereby suppressing the devel-
opment of ASR [20,33]. Therefore, despite using a highly-alkaline mixture and At each expansion level, three specimens were subjected to the SDT, following
alkali-silica reactive aggregates (C2 and F1), the measured expansion should be the procedure first outlined by Smaoui et al. [14]. Specimens were placed in a com-
attributable primarily, if not almost entirely to DEF, and not ASR. pressometer as described by ASTM C469 [31] and subjected to five load cycles using
256 E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264

a closed-loop hydraulic testing machine with a 100 kN capacity as shown in Fig. 3.


Because the platens installed on the load frame were fixed, a spherically-seated pla-
ten was inserted on top of the specimen. A loading and unloading rate of 0.1 MPa/s
was used, with a peak compressive stress of 10 MPa. An LVDT mounted in the com-
pressometer provided specimen deformation data, while the MTS controller
recorded the applied load. All data were sampled four times per second. The
load-deformation data was used to calculate and plot the stress–strain curve for
each specimen.
Two general methods of SDT data analysis were employed, hereafter referred to
as the Smaoui Method and Sanchez Method. The original parameters recommended
by Chrisp et al. [13] were developed when the test was conducted using a lower
maximum load and ultimately unsupported by any correlation to the degree of
damage, and so were not used in this study.

4.3.1. Smaoui Method analysis


The Smaoui Method analysis focused on two parameters: the area of the hys- Fig. 4. Illustration of the two parameters extracted for the Smaoui Method SDT data
teresis loop of the first loading cycle (A1) and total accumulated plastic strain over analysis.
all five loading cycles (S5). These parameters were found by Smaoui et al. to corre-
late nearly linearly to expansion from ASR [14], although more recent works are less
supportive that this is always the case [5,8,10,11,16,32]. These parameters are may vary considerably throughout the structure. Sanchez et al. [5,8,10,11,30,32]
extracted by numerical analysis of the stress and strain data generated by the suggested that the most suitable yet practical approach would be a two-step proce-
SDT, and are illustrated in Fig. 4. The average A1 and S5 at each level of expansion dure presented hereafter:
are calculated and plotted against expansion.
 Step 1: Determine the compressive strength on cores extracted from undam-
aged or less-damaged locations in the structural member under investigation,
4.3.2. Sanchez Method analysis and;
Sanchez et al. [5,8,10,11,30,32], following the above mentioned works of  Step 2: Use 40% of the value obtained in Step 1 to perform the SDT.
Smaoui et al. [14,15,26], conducted extensive studies of SDT on AAR-affected con-
crete. Using the cyclic loading technique, they evaluated concrete specimens fabri-
In addition, some practical parameters were found to influence SDT outcomes,
cated in the laboratory presenting different compressive strengths (25, 35 and 45
such as the moisture content and conditioning prior to conducting the test, location
MPa) and incorporating a wide range of reactive aggregates (both coarse and fine),
and direction of the core (i.e. longitudinal vs. transverse vs. vertical), environmental
as well as concrete cores extracted from AAR-affected structures. The authors
conditions and depth from the member surface (especially in cores drilled from dis-
showed that SDT starts being diagnostic and still non-destructive (i.e. tested sample
tressed structures) [10].
may be used to other tests after SDT cycles) when loads equal or greater to 30% of
The Sanchez Method was found very promising for assessing AAR damage and
the compressive strength are applied, although the procedure should ideally be car-
progress in concrete, especially with the use of indices as output parameters of the
ried out at 40% of the compressive strength in order to be considered a diagnostic
test, namely Stiffness Damage Index (SDI) and Plastic Deformation Index (PDI),
test procedure. However, in many practical situations, the information regarding
which represent respectively the ratio of dissipated energy or plastic deformation
mix design and 28-day compressive strength of in-service infrastructure is quite
(SI or DI, respectively) to the total energy or deformation induced in the concrete
limited, and even if quality control test records are available, in-situ strengths
(SI + SII or DI + DII, respectively) over the five cycles of testing (Fig. 5). Finally, fol-
lowing the works by Chrisp and co-workers [13], Sanchez et al. [7–9,11] confirmed
that the Non-Linearity Index (NLI), calculated by the slope of the stress–strain curve
from the origin to the half of the maximum load applied, divided by the slope curve
taken from the origin to the maximum load in the first cycle, would be an interest-
ing complementary index to consider, particularly for the assessment of cores
extracted from field structures, because this index accounts for preferred crack ori-
entation often found in AAR-affected reinforced concrete structures.

4.4. Other mechanical tests

Following the stiffness damage test, the static secant modulus of elasticity and
compressive strength of the specimens were determined in accordance with C469
[31] and ASTM C39 [34]. An estimated compressive strength (based upon an
assumed slight decline in strength from tests at the previous expansion level)
was used to determine the applied load for the first elastic modulus test, and sub-
sequent tests were based on the compressive strength of the previously tested spec-
imens in the set. The goal was to apply a load of 40% of the compressive strength;
however, the actual applied load varied from 29 to 46% of the compressive strength.

Fig. 5. Determination of the Stiffness Damage Index (SDI) and Plastic Deformation
Fig. 3. Test setup for stiffness damage test [4]. Index (PDI) based on the output from the Stiffness Damage Test (SDT).
E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264 257

Secant moduli should be nearly identical for peak loads of 30 to 45% of the compres- slip across crack surfaces while the cracks close under compression
sive strength [35]; the loads used in this study are generally within that range. This
loading [38].
sequence of tests (SDT ? E ? fc) typically involves the application of increasing
loads from one test to the next, so each test is not expected to influence the one
(s) which follow. In fact, ASTM C469 specifies that no data be recorded in the first 5.3. Elastic modulus and compressive strength
loading cycle, which supports the idea that there should be no influence on the
measured elastic modulus when the SDT is conducted at up to 40% of the compres-
Tables 3 and 4 present the elastic modulus and compressive
sive strength. However, when the compressive strength is less than 25 MPa, the
loads used in for SDT will exceed those used to determine the elastic modulus when strength for Mixture 1 and Mixture 2, respectively. The tables show
the maximum SDT load is fixed at 10 MPa. To test this, the elastic moduli obtained the moduli and strength of individual specimens, as well as the
from both the SDT (average of 2nd and 3rd cycles, per suggestions in [10] and ASTM average value for all three tested at each expansion value. The
C469 testing were compared and analyzed. tables also show the average elastic modulus of each set of three
specimens as a percentage of the value that would be expected
based on the average compressive strength of the specimens (E,%
5. Results of Predicted). This value is calculated as:

5.1. Specimen expansions Emean


E; % of Predicted ¼ pffiffiffiffi ð1Þ
4730 f c
The average expansions of the reference specimens are shown
in Fig. 6. The specimens from Mixture 1 exhibited much greater where the denominator of the equation is the SI-unit formula used
expansion than those from Mixture 2 after approximately 50 days in design equations to estimate elastic modulus in the ACI 318
after casting. However, both Mixture 1 and Mixture 2 specimens building code for normalweight concrete [39].
expanded more quickly than the cylinders cast from those mix-
tures that were conditioned to expand only from ASR using storage 6. Analyses and discussion
over water in sealed container at 38 °C. Expansion data for the ASR-
affected cylinders were presented alongside the data in Fig. 6 in 6.1. SDT – Smaoui Method analysis
earlier works by Giannini and co-workers [16,18]. The difference
in expansion behavior between the two mixtures is not particu- Smaoui et al. [14,15] proposed two primary parameters to
larly surprising; prior research has noted that the aggregates have quantify SDT results for ASR: (1) the area of the first loading cycle,
influence on expansion in DEF-affected mortars, and that this which will be referred to as A1 and; (2) the accumulated plastic
appears to be most strongly correlated to aggregate shape and tex- strain after five loading cycles, which will be referred to as S5.
ture, and the resulting quality of the paste-aggregate bond, rather Fig. 9 presents A1 plotted against expansion for each mixture and
than to alkali-reactivity [20,21,36,37]. Thus, it follows that the Fig. 10 presents S5 plotted against expansion for each mixture.
manufactured limestone sand in Mixture 2 may be expected to Error bars in these figures represent the range of A1 and S5 values
suppress expansion from DEF compared to Mixture 1, which con- measured for individual specimens. A linear regression is also
tained a natural sand. shown for the two mixtures in both Figs. 9 and 10. Each data point
represents the average of three cylinders tested at that expansion
level.
5.2. Stiffness damage test The linear regression of A1 vs. expansion has an R2 of 0.86 for
Mixture 1, and 0.79 for Mixture 2. In the case of S5, the linear
Figs. 7 and 8 present representative examples of the stress– regression yields an R2 of 0.87 for Mixture 1 and 0.72 for Mixture
strain data obtained at each expansion level for Mixtures 1 and 2. A better linear fit is evident for both A1 and S5 for Mixture 1 if
2, respectively. Qualitatively speaking, as the expansion level the final expansion level is excluded, with R2 values of 0.97 and
increases, increased signs of damage can be seen in the stress– 0.96, respectively. Comparing Figs. 9 and 10, it is clear that both
strain plots for both mixtures in the form of larger hysteresis areas parameters show a very similar relationship to expansion and it
(indicating dissipated energy), greater accumulated plastic strain, appears that the expansive phenomenon has two distinct phases:
and reduced stiffness (measured by slope of the stress–strain curve a) an initial linear trend up to expansions of about 0.40% and; b)
in the loading phases of each cycle). Furthermore, at higher expan- a second phase where the rate of increase in the damage parame-
sion levels, the loading portion of the stress–strain curve for the ters is somewhat slower compared to the initial phase.
first cycle is increasingly concave in shape, while the subsequent
loading cycles become increasingly convex in shape. These charac- 6.2. SDT – Sanchez Method
teristics also found in AAR-damaged materials, and are linked to
According to Sanchez et al. [8,10,11], the SDT is more successful
in characterizing damage in concrete (i.e. more accurate and less
influenced by a number of test parameters such as test load)
through the use of indices, SDI and PDI (stiffness damage and plas-
tic deformation indices, respectively), as described in 4.3.2.
Fig. 11 shows a plot of the above indices against expansion for
Mixture 1. Linear fits are shown in Fig. 11a and logarithmic fit
curves in Fig. 11b. Because the first set of specimens were tested
at 0.000% expansion, this data was excluded when developing
the logarithmic fit. The SDI values range from 0.14 to 0.46 and
PDI values range from 0.13 to 0.42, for expansion levels varying
from 0% to 1.01%. Moreover, the variation of both indices as a func-
tion of expansion presented a ‘‘concave” (i.e. close to a logarithmic
function) trend as well as a somehow two-phase behavior: a) a
Fig. 6. Average expansions of instrumented cylinders from the two mixtures. Error roughly linear trend up to 0.4% expansion and; b) a trend for level-
bars represent the range of expansions measured. ing off at expansions greater than 0.4%. Therefore, when all the
258 E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264

Fig. 7. Example stress–strain plots for Mixture 1 specimens (1 specimen per expansion level).

expansion levels are considered together, a linear regression Fig. 13 shows a plot of the SDI and PDI against expansion for
(Fig. 11a) does not indicate a very good fit for both indices (R2 of Mixture 2. Linear fits are shown in Fig. 13a and logarithmic fit
0.72 and 0.69 for SDI and PDI, respectively). However, when the curves in Fig. 13b. SDI values range from 0.18 to 0.32 and PDI val-
last expansion level is removed, as shown in Fig. 12, the linear ues range from 0.16 to 0.33, for expansions up to 0.45%. In this
trend is more clearly seen, and the linear regression has an R2 of case, for both indices, a linear trend is also apparent, and linear
0.97 for both indices. The logarithmic curves in Fig. 11b provide regression yields an R2 of 0.90 for SDI and 0.70 for PDI. The
a much improved fit for both indices (R2 of 0.93 and 0.88 for SDI improved linear trend for Mixture 2 compared to Mixture 1 (pri-
and PDI, respectively) through the highest expansion level. How- marily with regards to SDI) is likely related to the fact that the
ever, the steep slope of the curve at 0.1% expansion suggests this maximum expansions were less for Mixture 2. The logarithmic fit
logarithmic fit may be problematic for very small expansions and curves yield an R2 of 0.92 for SDI and 0.57 for PDI. In the case of
predict negative SDI and PDI values. A practical engineering Mixture 2, there is little benefit derived from the use of a logarith-
approach may be to use a two-part linear fit, with the slope mic fit over this range of expansions; this approach may be best
decreasing significantly at 0.4% expansion. suited only when dealing with expansions well in excess of 0.40%.
E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264 259

Fig. 8. Example stress–strain plots for Mixture 2 specimens (1 specimen per expansion level).

Table 3
Elastic modulus and compressive strength results for Mixture 1.

Exp. Elastic Modulus (GPa) Compressive Strength (MPa) E,% of SDT Load
(%) Individual Specimens Mean Individual Specimens Mean Predicted (% of fc)
0.00 22.6 24.6 22.2 23.1 27.6 27.6 26.9 27.3 93.5 36.6
0.10 24.4 18.8 16.9 20.0 33.2 26.1 25.9 28.4 79.5 35.2
0.21 20.0 10.8 13.5 14.8 30.7 23.1 24.8 26.2 61.0 38.1
0.34 8.5 9.1 6.9 8.2 20.2 20.4 18.4 19.7 39.1 50.9
0.42 8.4 5.7 8.5 7.5 20.5 16.5 19.3 18.7 36.6 53.4
0.51 6.3 7.2 6.6 6.7 19.1 18.2 17.7 18.3 33.1 54.5
1.01 6.3 4.6 6.6 5.8 16.9 14.3 16.6 16.0 30.7 62.7

Table 4
Elastic modulus and compressive strength results for Mixture 2.

Exp. Elastic Modulus (GPa) Compressive Strength (MPa) E,% of SDT Load
(%) Individual Specimens Mean Individual Specimens Mean Predicted (% of fc)
0.01 23.1 23.8 23.8 23.6 32.9 33.9 32.2 33.0 86.8 30.3
0.08 20.4 21.6 18.4 20.1 31.9 33.7 33.1 32.9 74.2 30.4
0.18 16.2 17.5 20.1 17.9 33.4 30.9 33.0 32.5 66.6 30.8
0.27 13.4 15.5 19.7 16.2 28.9 30.4 28.7 29.3 63.2 34.1
0.33 19.6 20.7 15.5 18.6 32.7 35.6 30.4 32.9 68.5 30.4
0.45 8.7 9.5 10.6 9.6 27.3 25.0 23.1 25.2 40.3 39.7
260 E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264

Fig. 9. Average area of the first loading cycle (A1) vs. expansion for Mixtures 1 and
2. Area is expressed as dissipated energy per unit volume (1 J/m3 = 1 Pa). Error bars
indicate the range of values calculated for individual specimens.

Fig. 11. Indices proposed by Sanchez vs. expansion level for Mixture 1 presenting
all the expansion levels with (a) linear fit, and (b) logarithmic fit. Error bars (black =
SDI, grey = PDI) indicate the range of values calculated for individual specimens.

Fig. 10. Average accumulated plastic strain over all five loading cycles (S5) vs.
expansion for Mixtures 1 and 2. Error bars represent the range of values calculated
for individual specimens.

An examination of the scatter in the data presented in Figs. 9–


13 also reveals that the SDI and PDI parameters exhibit signifi-
cantly less scatter than A1 and S5. As a percentage of the average
values at each expansion level, the range of individual SDI values
is an average of 24%, while the range of A1 values is on average
56% of the average value at a particular expansion level. The same
calculations for PDI and S5 yields 42% and 65%, respectively. While
this is a simplistic way to analyze the scatter of SDT data and the
calculated parameters, this further supports the use of the Sanchez
parameters of SDI and PDI in the analysis of SDT data for DEF-
damaged concrete.
Fig. 12. Indices proposed by Sanchez vs. expansion level for Mixture 1 disregarding
the last expansion level. Error bars (black = SDI, grey = PDI) indicate the range of
6.3. The use of SDT as a quantitative tool for assessing damage in values calculated for individual specimens.
concrete

Although the results discussed in the previous sections seem to of the material tested (e.g. compressive strength), and interesting
be interesting and promising, the use of the SDT in assessing con- and comparable results were found for a wide range of AAR-
crete damage should be discussed, especially regarding how the affected mixtures [5,8,10,11]. Under compression, the cracks and
method detects, measures, and most importantly ‘‘considers” dis- flaws found in the damaged medium (i.e. cracked concrete) tend
tress in damaged concrete affected by different mechanisms. The to close. If the load applied is insufficient, the cracks remain
output parameters proposed by Sanchez et al. [8,10,11] are able opened after load application, whereas if the load is too high, it cre-
to measure the actual ‘‘damage” state of a concrete based on the ates new damage in the tested material. The latter seems to pre-
hysteresis area and the plastic deformation under the stress vs. vent the use of fixed loading procedures proposed by Chrisp
strain curve of a tested material. In previous works, the reliability et al. [13] and Smaoui at al. [14,15] for quantitatively assessing real
of this approach was shown to be unrelated to the characteristics structures because in most practical cases, the characteristics of
E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264 261

Sanchez procedure for assessing concrete distress caused by com-


bined ASR and DEF and the results were promising. A fairly linear
trend was found in this case, which after some microscopic analy-
sis using the DRI, may be attributed to two main reasons: 1) the
different nature of DEF distress development, which occurs almost
exclusively in the cement paste and; 2) the lack of reaction prod-
ucts completely filling cracks caused by DEF.
In this work (Section 6.2), a linear trend of SDT parameters vs.
expansion was found for Mixture 2 concrete, whereas a two-step
behavior was found for Mixture 1 (a linear first step and a second
step with a concave or logarithmic trend). In the case of Mixture 1,
at very high expansions (>0.60%), the SDT parameters leveled off
even as expansion continued to increase. Without petrography of
the specimens to support a microstructural explanation, it is
unclear why this happens; however, prior works by Sanchez
et al. [8,10] suggest that sample conditioning parameters (e.g.
moisture content) may have some contribution to this observed
effect on SDT results. In this research, the samples remained satu-
rated throughout the conditioning (limewater soak) and testing
processes, effectively eliminating moisture as a variable. Therefore,
in order to fully understand the observed trends in SDT results, a
comprehensive petrographic evaluation is still needed and should
be addressed in future works. Otherwise, regarding practical appli-
cation of the tools, the largest expansion values attained by the
Mixture 1 specimens (>1%) are unlikely to be observed in the
majority of DEF-damaged civil structures. This is particularly true
for more heavily-reinforced concrete structures, where the steel
provides a confining force that significantly restrains expansion.
Reinforcement yielding, for example, happens at a strain of approx-
imately 0.20%; if expansions are restrained such that yielding does
not occur, the linear ‘‘first-step” relationship between SDT param-
Fig. 13. Indices proposed by Sanchez vs. expansion level for Mixture 2 with (a)
eters and expansion may be valid for damaged concrete that would
linear fit, and (b) logarithmic fit. Error bars (black = SDI, grey = PDI) indicate the
range of values calculated for individual specimens. be found in service. The analysis of the data obtained in the previ-
ous sections seems to once more support the capacity of the SDT in
characterizing damage due to DEF, as well as AAR.

field concrete are unknown, especially in relatively old


infrastructure.
Sanchez et al. [8,10,11] verified that in order to be considered a 6.4. Compressive strength and elastic modulus
quantitative tool, the SDT procedure should somehow ‘‘challenge”
the sample analyzed, and thus a load around 40% of the sample Considering that it is quite simple to also test for the elastic
strength should be used. Moreover, the prior authors determined modulus and compressive strength of a specimen following SDT,
that for AAR damage up to very high expansion levels (expansion some discussion of the value of those parameters in characterizing
levels of around 0.3–0.4%), the procedure proposed would still damage from DEF is warranted. Fig. 14a presents the static elastic
remain non-destructive. modulus (E) results (ASTM C469 test) for both mixtures vs. expan-
In the case of AAR damage up to 0.3–0.4% expansion, Sanchez sion (although with the final expansion level removed for Mixture
et al. [8,10,11] verified the presence of a similar concave trend of 1), and Fig. 14b compares this data against average moduli calcu-
SDT/PDI vs. expansion levels, as is seen for Mixture 1 in this study. lated from the 2nd and 3rd SDT loading cycles. Fig. 15 presents
The authors explained this trend according to microscopic analysis the average compressive strength results vs. expansion. Error bars
made through the use of the Damage Rating Index (DRI) [7,9]. At in Figs. 14a and 15 represent the range of moduli and compressive
the onset of ASR expansion, a number of cracks are formed within strengths measured for all specimens at each expansion value. Full
the aggregates particles. As expansion increases, new cracks are data for individual samples are given in Tables 3 and 4. The elastic
again formed in the aggregate particles but more importantly, moduli of the specimens were greatly affected by damage from
the cracks already present increase in length and width, eventually DEF, decreasing as much as 75% for Mixture 1 and 59% for Mixture
reaching the cement paste. Finally, the previously-formed cracks 2, compared to baseline (zero expansion) values. Compressive
link together forming a network of cracks. The ‘‘concave” or level- strength decreased 41% for Mixture 1 and 24% for Mixture 2, com-
ing off trend of the SDI and PDI was found to coincide primarily pared to the baseline values.
with the last step of crack development, when the cracks are Although compressive strength is a significant parameter in
growing and linking, rather than forming, because the indices assessing the load-bearing capacity of a structure, it is ultimately
measured by the SDT are obtained by the number of cracks (which the least sensitive parameter investigated with respect to charac-
are not necessarily greater at the last steps of the chemical reac- terizing DEF progress in concrete. Furthermore, as the data in
tion) vs. the width of the cracks (which keeps increasing as a func- Tables 3 and 4 show, the elastic modulus is significantly under-
pffiffiffiffi
tion of ASR development). Furthermore, the behavior of ASR gel predicted by the equation E ¼ 4730 f c given in the ACI 318 Build-
within the cracks was also questioned by the authors, and probably ing Code. Measured elastic moduli were as much as 60–70% below
they play a non-negligible role on the leveling off trend obtained. values calculated based on compressive strength. Therefore, not
Martin et al. [40] investigated the use of SDT according to the only is the elastic modulus a more suitable parameter than
262 E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264

compressive strength to characterize the degree of damage from


DEF, it must be tested for directly.
Fig. 14b shows there is a close agreement between the elastic
moduli measured using the ASTM C469 method and the 2nd and
3rd SDT loading cycles. This further supports the argument that
the SDT is nondestructive in nature, provided that the applied load
is approximately 40% or less of the compressive strength. The SDT
moduli were generally slightly higher than the ASTM C469 moduli;
for SDT loadings less than 40% of the compressive strength, the SDT
moduli were, on average, 4% higher than the ASTM C469 moduli.
The deviation between these moduli values increased for higher
expansions of Mixture 1 (0.344%), where the SDT loading
exceeded 50% of the compressive strength. A reexamination of
Figs. 7 and 8 show that the stress–strain curves of the 2nd through
5th SDT loading cycles become increasingly concave at higher
expansion levels; this means that as SDT loads exceed those
applied in the ASTM C469 test, the secant moduli calculated from
the SDT data would be expected to exceed those determined by
ASTM C469. As cracks are closed under load, the stress–strain
response of DEF-damaged concrete would be expected to stiffen
slightly; this behavior becomes more pronounced as damage
increases, and is reflected in the increasingly concave shape of
the stress–strain curve, and the disparity in moduli determined
by these two methods. However, it should be noted that either
method would be preferable to estimating the elastic modulus
from compressive strength data.

6.5. Statistical analysis

Fig. 14. (a) Average elastic modulus (ASTM C469) vs. expansion for Mixtures 1 and In order to assess the statistical validity and significance of the
2. Final expansion level excluded for Mixture 1. (b) Average elastic moduli from database generated as part of this study, analyses of variance (two-
ASTM C469 and SDT 2nd and 3rd cycles compared.
variable ANOVA) were performed, taking into account both the
individual specimens and the average results obtained at each
expansion level.
First, two-variable ANOVA was performed on the individual test
results (i.e. accounting for each specimen evaluated) obtained from
Sanchez Method in order to verify the significance of the individual
outcomes obtained at each expansion level studied (with a confi-
dence level of 95%) for all the parameters analyzed in this research
(i.e. PDI, SDI, and E, respectively). The results, presented in Table 5,
indicate that all the different expansion levels provided significant
results (i.e. the result of each set of specimens, composed of three
different cylinders, showed equal results for the level of signifi-
cance specified), as all the ‘‘F values” obtained were lower than
the ‘‘Fcritic” and the ‘‘p values” were greater than 0.05 for each case.
These results thus enabled the use of the average values obtained
at each expansion level to appraise the significance of the out-
comes obtained through the use of the Sanchez Method.
Second, following the previous statistical results, two-variable
ANOVA was performed on the average values obtained for each
set of specimens at each expansion level studied (Table 6). Analyz-
ing the data obtained, one may notice that all the results obtained
Fig. 15. Average compressive strength vs. expansion for Mixtures 1 and 2.
through the use of Sanchez Method (i.e. PDI, SDI and E, respec-

Table 5
Two-variable ANOVA performed on the individual test results obtained through the use of Sanchez Method.

Mixture Variable F Fcritic F > Fcritic a p-value p<a


1 SDI 1.30 3.74 No 0.05 .30 No
PDI 1.93 3.74 No 0.05 .18 No
E 1.93 3.74 No 0.05 .18 No
2 SDI 0.91 3.89 No 0.05 .43 No
PDI 0.36 3.88 No 0.05 .70 No
E 0.36 3.88 No 0.05 .70 No
E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264 263

Table 6
Two-variable ANOVA performed on the average test results obtained for the different parameters used in Sanchez Method.

Mixture Variable F Fcritic F > Fcritic a p-value p<a


1 SDI 8.88 2.76 Yes 0.05 .0003 Yes
PDI 9.25 2.76 Yes 0.05 .0002 Yes
E 30.59 2.76 Yes 0.05 .00 Yes
2 SDI 8.95 3.00 Yes 0.05 .0007 Yes
PDI 8.85 3.00 Yes 0.05 .0008 Yes
E 45.20 3.00 Yes 0.05 .00 Yes

tively) could be considered statistically significant, as all the would also like to acknowledge the contributions of Kerry Kreit-
‘‘F values” obtained were largely greater than the ‘‘Fcritic” and the man, Zach Webb, Brian Hanson, Eric Schell, and Dennis Fillip.
‘‘p values” were much less than 0.05, confirming thus the analyses
and data interpretation made throughout this research.
References

7. Conclusions [1] ISE, Structural Effects of Alkali-Silica Reaction, Institution of Structural


Engineers, London, 1992, 45.
[2] T. Ahmed, E. Burley, S. Rigden, A.I. Abu-Tair, The effect of alkali reactivity on
In this work, the SDT was used to assess concrete mixtures the mechanical properties of concrete, Constr. Build. Mater. 17 (2003) 123–
affected by DEF and presenting a wide range of deterioration levels. 144, https://doi.org/10.1016/S0950-0618(02)00009-0.
From this study, the following conclusions can be made: [3] R.N. Swamy, Assessment and rehabilitation of AAR-affected Structures, Cem.
Concr. Compos. 19 (5-6) (1997) 427–440, https://doi.org/10.1016/S0958-9465
(97)00035-8.
 The SDT was previously found to provide useful information on [4] E.R. Giannini, Evaluation of Concrete Structures Affected by Alkali-Silica
the degree of damage of ASR-affected concrete such as its stiff- Reaction and Delayed Ettringite Formation PhD Dissertation, The University
of Texas at Austin, Austin, Texas, 2012.
ness and inner damage. Furthermore, it has been found that [5] L.F.M. Sanchez, B. Fournier, M. Jolin, J. Bastien, Evaluation of the Stiffness
when loads equal or lower than 40% of the material’s compres- Damage Test (SDT) as a tool for assessing damage in concrete due to ASR: test
sive strength are used, the SDT remains a non-destructive loading and output responses for concretes incorporating fine or coarse
reactive aggregates, Cem. Concr. Res. 56 (2014) 213–229, https://doi.org/
method up to very high levels of deterioration of the material 10.1016/j.cemconres.2013.11.003.
tested. [6] L.F.M. Sanchez, S. Multon, A. Sellier, M. Cyr, B. Fournier, M. Jolin, Comparative
 SDT is suitable as a mechanical testing tool to quantitatively study of a chemo–mechanical modeling for alkali silica reaction with
experimental evidences, Constr. Build. Mater. 72 (2014) (2014) 301–315,
characterize DEF evolution and damage in laboratory concrete.
https://doi.org/10.1016/j.conbuildmat.2014.09.007.
Moreover, it is believed that the SDT would even be more suit- [7] L.F.M. Sanchez, B. Fournier, M. Jolin, J. Duchesne, Reliable quantification of AAR
able to appraise damage from DEF than from ASR, especially damage through assessment of the damage rating index (DRI), Cem. Concr. Res.
67 (2015) 74–92, https://doi.org/10.1016/j.cemconres.2014.08.002.
because DEF microscopic damage features are almost entirely
[8] L.F.M. Sanchez, B. Fournier, M. Jolin, J. Bastien, Evaluation of the Stiffness
found in the cement paste and/or ITZ. Thus, conceptually, the Damage Test (SDT) as a tool for assessing damage in concrete due to ASR: input
SDT parameters would be less influenced by aggregate charac- parameters and variability of the test responses, Constr. Build. Mater. 77
teristics (rock type, mineralogy, etc.) than when testing ASR- (2015) 20–32, https://doi.org/10.1016/j.conbuildmat.2014.11.071.
[9] L.F.M. Sanchez, B. Fournier, M. Jolin, M.A.B. Bedoya, J. Bastien, J. Duchesne, Use
damaged concrete. Further analyses are needed to validate this of the Damage Rating Index (DRI) to quantify damage due to alkali-silica
hypothesis. reaction in concrete incorporating reactive fine and coarse aggregates, ACI
 The SDI and PDI parameters developed by Sanchez et al. and Mater. J. 113 (2016) 395–407, https://doi.org/10.14359/51688983.
[10] L.F.M. Sanchez, B. Fournier, M. Jolin, J. Bastien, D. Mitchell, Practical use of the
found to be suitable for assessing ASR damage can be directly Stiffness Damage Test (SDT) for assessing damage in concrete infrastructure
employed to characterize DEF-affected concrete even if the affected by alkali-silica reaction, Constr. Build. Mater. 125 (2016) 1178–1188,
loads applied in SDT are slightly lower than 40% of the compres- https://doi.org/10.1016/j.conbuildmat.2016.08.101.
[11] L.F.M. Sanchez, B. Fournier, M. Jolin, D. Mitchell, J. Bastien, Overall assessment
sive strength. of alkali-aggregate reaction (AAR) in concretes presenting different strengths
 The SDI and PDI parameters are more effective ways of charac- and incorporating a wide range of reactive aggregate types and natures, Cem.
terizing DEF-damaged concrete than previous parameters iden- Concr. Res. 93 (2017) 17–31, https://doi.org/10.1016/j.
cemconres.2016.12.001.
tified by Smaoui et al.
[12] R.S. Crouch, J.G.M. Wood, Damage Evolution in AAR Affected Structures ,
 Both the SDI and PDI parameters obtained from the SDT and the International Symposium on Fracture Damage of Concrete and Rock, Vienna,
static elastic modulus provide statistically significant results July 1988.
[13] T.M. Chrisp, P. Waldron, J.G.M. Wood, Development of a non-destructive test
and could be considered unique ways of assessing damage from
to quantify damage in deteriorated concrete, Mag. Concr. Res. 45 (165) (1993)
DEF. 247–256, https://doi.org/10.1680/macr.1993.45.165.247.
 Future research to advance the use of SDT to characterize DEF [14] N. Smaoui, M.-A. Bérubé, B. Fournier, B. Bissonette, B. Durand, Evaluation of the
damage in concrete should seek to use petrographic examina- expansion attained to date by concrete affected by alkali silica reaction. Part I:
experimental study, Canadian, J. Civ. Eng. 31 (5) (2004) 826–845, https://doi.
tion to correlate the SDT and expansion data to the Damage Rat- org/10.1139/l04-051.
ing Index. It is also recommended that consideration be given to [15] N. Smaoui, B. Fournier, M.-A. Bérubé, B. Bissonette, B. Durand, Evaluation of the
combined deterioration from ASR and DEF, because DEF is more expansion attained to date by concrete affected by alkali-silica reaction. Part
II: application to nonreinforced concrete specimens exposed outside,
commonly found in concert with ASR in field structures than as Canadian, J. Civ. Eng. 31 (6) (2004) 997–1011, https://doi.org/10.1139/l04-074.
the sole damage mechanism. [16] E.R. Giannini, K.J. Folliard, Stiffness Damage and Mechanical Testing of Core
Specimens for the Evaluation of Structures Affected by ASR, in Eds. T. Drimalas,
J.H. Ideker, B. Fournier, in: 14th Int. Conf. on Alkali-Aggregate Reactions in
Concrete, Austin, TX, USA, 2012, 10.
[17] A.F. Bentivegna, J.H. Ideker, Application of prognosis and diagnosis techniques
Acknowledgements of ASR for a historic structure, in Eds. H.M. Bernardes, N.P. Hasparyk, in: 15th
Int. Conf. on Alkali-Aggregate Reaction in Concrete, São Paulo, Brazil, 2016.
[18] E.R. Giannini, K.J. Folliard, J. Zhu, O. Bayrak, K. Kreitman, Z. Webb, B. Hanson.
This research was supported in part by the Texas Department of Non-Destructive Evaluation of In-Service Concrete Structures Affected by
Transportation as part of research project 0-6491. The authors Alkali-Silica Reaction (ASR) or Delayed Ettringite Formation (DEF) -Final
264 E.R. Giannini et al. / Construction and Building Materials 162 (2018) 253–264

Report, Part I, No. FHWA/TX-13/0-6491-1, Center for Transportation Research, [29] S. Tremblay, B. Fournier, M. Thomas, T. Drimalas, K. Folliard, The Lomas
Austin, Texas, 2013. Boulevard Road Test Site, Albuquerque (New Mexico) - A Case Study on the
[19] S. Kelham, The Effect of cement composition and fineness on expansion Use of Preventative Measures Against ASR in New Concrete, in Eds. T.
associated with delayed ettringite formation, Cem. Concr. Compos. 18 (1996) Drimalas, J.H. Ideker, B. Fournier, in: 14th Int. Conf. on Alkali-Aggregate
171–179, https://doi.org/10.1016/0958-9465(95)00013-5. Reactions in Concrete, Austin, TX, USA, 2012, 10.
[20] K.J. Folliard, R. Barborak, T. Drimalas, L. Du, S. Garber, J. Ideker, T. Ley, S. [30] L. Sanchez, B. Fournier, M. Jolin, Critical Parameters of the Stiffness Damage
Williams, M. Juenger, B. Fournier, M.D.A. Thomas, Preventing ASR/DEF In New Test for Assessing Concrete Damage due to Alkali-Silica Reaction, in Eds. T.
Concrete: Final Report, Center for Transportation Research, Austin, Texas, Drimalas, J.H. Ideker, B. Fournier, in: 14th Int. Conf. on Alkali-Aggregate
2006. Reactions in Concrete, Austin, TX, USA, 2012, 10.
[21] H.F.W. Taylor, C. Famy, K.L. Scrivener, Delayed Ettringite Formation, Cem. [31] ASTM C469/C469M-14, Standard Test Method for Static Modulus of Elasticity
Concr. Res. 31 (5) (2001) 683–693, https://doi.org/10.1016/S0008-8846(01) and Poisson’s Ratio of Concrete in Compression, ASTM International, West
00466-5. Conshohocken, Pennsylvania, 2014. doi: 10.1520/C0469_C0469M.
[22] T. Ramlochan, The Effects of Pozzolans and Slag on the Expansion of Mortars [32] L.F.M. Sanchez, Contribution to the Assessment of Damage in Aging Concrete
and Concrete Cured at Elevated Temperature PhD Thesis, University of Infrastructures Affected by Alkali-Aggregate Reaction Ph.D. Dissertation,
Toronto, Toronto, 2003. Université Laval, Québec, 2014.
[23] K.L. Scrivener, M.C. Lewis, Effect of Heat Curing on Expansion of Mortars and [33] M. Thomas, K. Folliard, T. Drimalas, T. Ramlochan, Diagnosing Delayed
Composition of Calcium Silicate Hydrate Gel, Ed. B. Erlin. Ettringite - The Ettringite Formation in Concrete Structures, Cem. Concr. Res. 38 (6) (2008)
Sometimes Host of Destruction, ACI SP-177, American Concrete Institute, 841–847, https://doi.org/10.1016/j.cemconres.2008.01.003.
Farmington Hills, Michigan, 1999. 93-104. [34] ASTM C39/C39M-17, Standard Test Method for Compressive Strength of
[24] R.P. Martin, C. Bazin, F. Toutlemonde, Alkali aggregate reaction and delayed Cylindrical Concrete Specimens, ASTM International, West Conshohocken,
ettringite formation: common features and differences, in Eds. T. Drimalas, J.H. Pennsylvania, 2017. doi: 10.1520/C0039_C0039M-17.
Ideker, B. Fournier, in: 14th Int. Conf. on Alkali-Aggregate Reactions in [35] S. Popovics, Strength and Related Properties of Concrete: A Quantitative
Concrete, Austin, TX, USA, 2012, 10. Approach, John Wiley & Sons, New York, 1998.
[25] B. Fournier, M.-A. Bérubé, K.J. Folliard, M. Thomas, Report on the Diagnosis, [36] P.E. Grattan-Bellew, J.J. Beaudoin, V.-G. Vallée, Effect of Aggregate Particle Size
Prognosis, and Mitigation of Alkali-Silica Reaction (ASR) in Transportation and Composition on Expansion of Mortar Bars due to Delayed Ettringite
Structures, Report No. FHWA-HIF-09-004, Federal Highway Administration, U. Formation, Cem. Concr. Res. 28 (8) (1998) 1147–1156, https://doi.org/10.1016/
S. Department of Transportation, Washington, DC, 2010. S0008-8846(98)00084-2.
[26] M.-A. Bérubé, B. Durand, B. Bissonette, B. Fournier, N. Smaoui, Evaluation of the [37] R. Yang, C.D. Lawrence, J.H. Sharp, Effect of type of aggregate on delayed
expansion attained to date by concrete affected by alkali silica reaction. Part ettringite formation, Adv. Cem. Res. 11 (3) (1999) 119–132, https://doi.org/
III: application to existing structures, Canadian, J. Civ. Eng. 32 (3) (2005) 463– 10.1680/adcr.1999.11.3.119.
479, https://doi.org/10.1139/l04-104. [38] S. Mindess, J.F. Young, D. Darwin, Concrete, second ed., Prentice Hall, Upper
[27] L. Sanchez, P. Salva, B. Fournier, M. Jolin, N. Pouliot, A. Hovington, Evaluation of Saddle River, New Jersey, 2003.
Damage in the Concrete Elements of the Viaduct ‘‘Robert-Bourassa-Charest” [39] ACI Committee 318, ACI 318-14, Building Code Requirements for Structural
After Nearly 50 Years in Service, in Eds. T. Drimalas, J.H. Ideker, B. Fournier, in: Concrete, Farmington Hills, Michigan: American Concrete Institute, 2014.
14th Int. Conf. on Alkali-Aggregate Reactions in Concrete, Austin, TX, USA, [40] R.-P. Martin, L. Sanchez, B. Fournier, F. Toutlemonde, Diagnosis of AAR and
2012, 10. DEF: Comparison of Residual Expansion, Stiffness Damage Test and Damage
[28] M.D.A. Thomas, K. J. Folliard, B. Fournier, T. Drimalas, P. Rivard, Study of Rating Index, in Eds. H.M. Bernardes, N.P. Hasparyk, 15th Int. Conf. on Alkali-
Remedial Actions on Highway Structures Affected by ASR, in Eds. T. Drimalas, J. Aggregate Reaction in Concrete, São Paulo, Brazil, 2016.
H. Ideker, B. Fournier, in: 14th Int. Conf. on Alkali-Aggregate Reactions in
Concrete, Austin, TX, USA, 2012, 10.

You might also like