You are on page 1of 10

Construction and Building Materials 200 (2019) 255–264

Contents lists available at ScienceDirect

Construction and Building Materials


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

External sulfate attack on compressed stabilized earth blocks


Wesley V.D.C. Bezerra ⇑, Givanildo A. Azeredo
Laboratório de Ensaios de Materiais e Estruturas, Universidade Federal da Paraíba, João Pessoa CEP 58051-900, Brazil

h i g h l i g h t s

 Pioneer study with sulfate attack on compressed stabilized earth blocks.


 Wetting time and sodium sulfate concentration were highlighted.
 Failure mechanism associated with thenardite and mirabilite precipitation.
 A procedure to test CSEB against external sodium sulfate attack was recommended.

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

Article history: Cementitious materials can be damaged by external sulfate attack due to ingress of sulfate ions from
Received 24 May 2018 external environment. This process causes loss of cohesion, increase in porosity, local expansion, spalling
Received in revised form 6 December 2018 and cracking. The objectives of this work were to study the influence of the capillary absorption time and
Accepted 20 December 2018
of the sulfate ions concentration on compressed stabilized earth blocks (CSEB) exposed to sulfate attack.
Available online 26 December 2018
The CSEB were submitted to cycles of capillary absorption of sodium sulfate solution and drying. Two sets
of procedures were used to evaluate the wear of the samples. In the first set of procedures, the capillary
Keywords:
absorption time was varied. In the second one, the sodium sulfate concentration was varied instead. The
Sulfate attack
Earth blocks
CSEB used had the same dosage of earth, cement and water and for each procedure 3 samples were
Capillarity tested. The masses were measured after taking the samples out of the solution and after a 2-week drying
time, being that a cyclic process for each procedure. The room temperature and relative humidity allowed
crystallization of thenardite and mirabilite. Granulometry, XRD, XRF and SEM tests were used to identify
the materials. The results showed that capillary absorption time and sulfate concentration influences the
wear of the CSEB.
Ó 2018 Elsevier Ltd. All rights reserved.

1. Introduction and supporting walls, tunnel linings, water channels, a water reser-
voir, a mining waste water processing basin and a sewage man-
Sulfate attack encompasses a series of harmful physical and hole. Many other examples of sulfate attack in different countries
chemical reactions between sulfate ions and components of hard- were also cited by Crammond [5].
ened cementitious materials. Consequences include expansion, Sulfate attack on concrete, mortar and paste samples have been
cracking, spalling, decrease in strength and diminishing cohesive- extensively studied in the last decades [6–9], but insufficient data
ness of the cement hydration products [1]. is available when testing alternative cementitious materials sam-
Various structures made of cementitious materials, mainly con- ples such as the compressed stabilized earth blocks (CSEB), one
crete, deteriorated by sulfate attack were mentioned in previous which uses earth as the primary raw material.
studies. In Switzerland, Romer et al. [2] reports damage caused Earth in natura is a natural building material that has been used
in tunnel structures in contact with groundwater. In the United since ancient times, along with wood and stone [10,11]. Although
Kingdom, Crammond [3] cites sulfate wear on a variety of build- there is no consensus of when man began to use earth for construc-
ings, from foundations of domestic houses and highway bridges tion purposes, it’s not far from the truth to claim that earthen
to tunnel linings and harbor steps. In Germany, Bellmann et al. buildings began with the agricultural societies, between 12,000
[4] discusses 20 different cases, including river bridges foundations and 7000 BCE [12]. With the increasing population growth and
housing deficit dilemma, earth construction becomes an efficient
solution due to its low cost and local availability [13–15]. Addition-
⇑ Corresponding author. ally, at the end of a building’s life the earthen material can easily be
E-mail address: bezerrawesleyvitor@gmail.com (W.V.D.C. Bezerra).

https://doi.org/10.1016/j.conbuildmat.2018.12.115
0950-0618/Ó 2018 Elsevier Ltd. All rights reserved.
256 W.V.D.C. Bezerra, G.A. Azeredo / Construction and Building Materials 200 (2019) 255–264

reused or returned to the ground without any interference with the 100.00
environment [16].
80.00
However, most soils in their natural condition lack certain

% Passing
requirements needed for constructions, such as strength, dimen- 60.00
sional stability, durability and water resistance [10,17]. To over- 40.00
come these shortcomings, stabilization techniques that reduces
porosity, decreases permeability and increases mechanical 20.00
strength can be applied [18]. The compressed earth block was 0.00
one of the several building materials developed in order to meet 0.00 0.01 0.10 1.00 10.00
those requirements [19], being usually chemically stabilized with Grain size (mm)
cement or lime [20,21].
As the soil-structures interactions play a role on the durability Fig. 1. Grain size distribution curve of the earth.
of the materials, it is important to be aware of the external factors
that may damage the constructions, such as high concentrations of
salts like sulfates and chlorides.
Sulfate attack is usually classified in accordance with its origin,  100% of the particles smaller than 4.8 mm;
 Between 10% and 50% of the particles smaller than 0.075 mm (#200);
external or internal, and with its nature, physical or chemical.  Liquid limit 45%;
External sulfate attack (ESA) occurs when environmental sulfate  Plasticity index 18%.
present in soils, groundwater, industrial effluents and decaying
organic matter penetrates porous materials [1,22]. Internal sulfate X-ray fluorescence (XRF) and x-ray diffraction (XRD) were used to identify
attack (ISA) occurs when the sulfate source is inside the material, the minerals present on the earth. The XRF displayed predominance of SiO2,
Al2O3, and Fe2O3 (Table 1). The XRD indicated presence of quartz and kaolinite
usually coming from cement with high sulfate concentration or
(Fig. 2).
aggregates contaminated with gypsum [22].
Physical sulfate attack (PSA) occurs due to continuous growth of
2.2. Manufacture of the samples
salt crystals that generates pressure inside the pores of the mate-
rial [23,24]. When the pressure in small pores exceeds the tensile The earth used was dried in the open air to achieve hygroscopic moisture.
strength of the binder, microcracks occur [25]. In case of larger The content of Portland cement and water were 12% and 4.5% by mass,
pores, insufficient pressure is generated to cause damage. In case respectively. Dimensions of the blocks are indicated in Fig. 3. The blocks were
cured for 14 days in a covered place protected from rain and sunlight. Daily
of sodium sulfate (Na2SO4), these formed crystals are due to con-
watering was done twice to ensure adequate occurrence of hydration
tact of porous materials with high concentrations of aqueous reactions.
sodium sulfate (Na2SO4) under specific temperature and relative
humidity conditions, with phase changes between thenardite (Na2- 2.3. Experimental procedure
SO4) and mirabilite (Na2SO410H2O) taking place [24,26,27].
Other forms of sulfate compounds such as calcium sulfate Eighteen blocks were selected and submitted to two different set of procedures
(CaSO4), magnesium sulfate (MgSO4) and sulfuric acid (H2SO4) (Fig. 4). On the first set (Group A), the influence of the time in contact with aqueous
sodium sulfate solution – capillary absorption time – was studied. On the second
can also be harmful, but are usually associated to chemical sulfate
one (Group B), the influence of sodium sulfate concentration was studied using
attack (CSA) [27]. CSA occurs due to the reactions between sulfate the optimal capillary absorption time from Group A. The drying time for all proce-
ions and the main components of hydrated cement paste [23]. dures was 2 weeks.
Some examples of these reactions are the precipitation of addi- To remove excessive moisture, the blocks stayed 24 h in a thermal oven at
tional ettringite and gypsum [28]. 100 °C. After that, the blocks were separated and rested at room temperature.
The blocks were then placed in PVC containers on aluminum support bars and
The focus of this work is to evaluate the sodium sulfate attack aqueous sodium sulfate solution was gradually added to a 2 cm height (Fig. 5).
on CSEB. The samples were exposed to external sulfate attack by The solution was replaced as its level dropped due to absorption and evaporation.
cycles of capillary absorption and drying. With that, this research During the experiments, no visual signs of aluminum corrosion on the support bars
aimed at: studying the influence of the capillary absorption time was noticed.
To test the wear of compressed stabilized earth blocks exposed to sodium
and of the sodium sulfate concentration on the wear of CSEB sam-
sulfate, the RILEM TC 127-MS-A.1 [31] was used as base. This report recom-
ples. The damage was evaluated by verifying the occurrence of PSA mends capillarity and subsequent immersion of masonry blocks using sodium
and CSA. sulfate at 10% concentration by mass as an accelerated procedure. In this
study only capillarity was used, also taking as reference the NBR 9779
(1995) [32], which addresses capillarity tests on hardened concrete and mortar
samples.
2. Materials and methods Natural occurring sulfate concentration ranges from 0.015 to 1% on water bod-
ies, 0.0003–2% on groundwater and approximately 0.26% on seawater [2,4,33–37].
2.1. Materials However, at those concentrations, failure of the tested materials can take several
years to happen. Hence, 10%, 6% and 3% concentrations were chosen for the
Silty sand, Portland cement CP II-Z-32 and tap water were used to manufacture experiments.
the compressed stabilized earth blocks. The CP II-Z-32 contains 6–14% pozzolan
material and 0–10% carbonaceous material. The pozzolan is required to have at
least 75% activity at 28 days and the carbonaceous material is required to be com-
posed of at least 85% CaCO3. The remaining percentage is composed of clinker and
Table 1
calcium sulfates.
X-ray fluorescence of the earth.
Sieves from 76 mm to 0.075 mm and laser diffraction analysis were used to
obtain the grain size distribution of the earth (Fig. 1), their coarse-grained fraction Compound Percentage composition
(d > 0.075 mm) and fine-grained fraction (d < 0.075 mm), respectively. The results
SiO2 62.27
indicated that the earth is composed of: 5.15% of gravel, 77.29% of sand, 16.78%
Al2O3 23.97
of silt, and 0.78% of clay [29]. Before its usage, the earth was sifted through a
Fe2O3 9.96
4.8 mm and the retained fraction was discarded.
TiO2 3.13
The soil index properties obtained were: liquid limit of 28.64%, plastic limit of
ZrO2 0.21
18.62% and plasticity index of 10.02%. According to the NBR 10833 [30], its proper-
SO3 0.12
ties meet the requirements for the manufacture of compressed stabilized earth
Others 0.34
blocks:
W.V.D.C. Bezerra, G.A. Azeredo / Construction and Building Materials 200 (2019) 255–264 257

Fig. 2. X-ray diffraction of the earth.

Fig. 5. Capillarity test in PVC containers.

 Procedure 3: The blocks 7, 8 and 9 went through the same steps as procedure 1,
except the capillary absorption time (step 3) being 2 weeks.
 Procedure 4: The blocks 10, 11 and 12 went through the same steps as proce-
dure 1, except the blocks being initially submitted to aqueous sodium sulfate
solution at 10% concentration by mass for 24 h on the two first cycles. From
the third cycle on, tap water was used to replace the sodium sulfate solution.

Group B – Concentration study:


From Group A, the optimal procedure was 1 week in solution and 2 weeks dry-
ing (procedure 2), as discussed on the results section. All the blocks on Group B
underwent that procedure, except blocks 13, 14 and 15 which were tested with
3% aqueous sodium sulfate solution and blocks 16, 17 and 18 with 6% aqueous
Fig. 3. Dimensions of the samples. sodium sulfate solution. Group B blocks were exposed in a closed room with tem-
perature and relative humidity values of 26.0 ± 2.0 °C and 75.0 ± 8.5%, respectively.

3. Results and discussion


Group A – Capillary absorption time study:

At the end of the capillary absorption and drying phases, the sam-
 Procedure 1:
ples were weighed and photographed for analysis of wear caused by
1. The blocks 1, 2 and 3 were placed inside the container on the aluminum the sodium sulfate solution. It was intended to carry out the tests until
support bars and aqueous sodium sulfate solution at 10% concentration failure or further, but some procedures were stopped earlier because
by mass was gradually added until reaching a 2 cm height from the base
of no tendency to failure in the next cycles. Blocks were considered to
of the block. This time was registered as time zero (t = 0 h);
2. At t = 4 h, the solution was replaced to 2 cm; have failed when intense spalling and/or cracking occurred.
3. At t = 24 h, the blocks were taken out of the solution and weighted, The blocks had mean average weight of 3310 ± 100 g. Addition-
obtaining Mw1 (in grams); ally, the procedures started being cyclic only after the first capillary
4. The container was emptied and cleaned, then the blocks were placed absorption and drying cycle, therefore for calculations the initial
inside on support bars for drying. The samples remained in a closed
room with temperature and relative humidity of 27.5 ± 0.5 °C and
mass was considered to be the mass of the first drying phase (Md1).
71.5 ± 5.0%, respectively;
5. After 2 weeks, the blocks were subtly brushed until the free particles
3.1. Group A: capillary absorption time study
were removed. After that, they were weighted again, obtaining Md1;
6. The steps 1–5 were repeated until failure of the samples.
 Procedure 2: The blocks 4, 5 and 6 went through the same steps as procedure 1, The mass loss of the specimens is presented in Table A.1
except the capillary absorption time (step 3) being 1 week. (Appendices) and was calculated using the Eq. (1):

Fig. 4. Capillary absorption time and sulfate concentration for each procedure. S = sodium sulfate.
258 W.V.D.C. Bezerra, G.A. Azeredo / Construction and Building Materials 200 (2019) 255–264

Mijð%Þ ¼ ðMij  Md1Þ=Md1  100 ð1Þ The SEM samples went through a drying process at 100 °C for
24 h before the tests. At the surface portion sample, occurrence
and shape of thenardite (Then) and dehydrated mirabilite (d-
where i = w (wetting by capillary absorption) or d (drying), j = cycle Mirab) crystals are shown in Figs. 12 and 13. The thenardite iden-
1, cycle 2, cycle 3 . . . for d values or cycle 2, cycle 3, cycle 4 . . . for w tified is micron/submicron sized, in shape of anhedral crystals.
values; and Md1 = mass of the blocks weighted at the end of the dry- Similar thenardite crystals were observed by Rodriguez-Navarro
ing phase of the cycle 1. et al. [26] and Aye et al. [38] at both laboratory and field condi-
The distribution of the mass loss over time is shown in Fig. 6. In tions. Authors have reported that thenardite dissolves to the solu-
general, the procedures with samples exposed for longer periods at tion and subsequently crystallizes into thenardite and mirabilite
capillary absorption phases yielded higher mass loss, with dimin- phases upon drying [26,39]. The specimens were under tempera-
ishing increases. The 2 weeks capillary absorption time procedure ture and relative humidity conditions of 27.5 ± 0.5 °C and
(procedure 3), despite being the most aggressive, resulted in simi- 71.5 ± 5.0%, respectively. These conditions allowed precipitation
lar mass loss compared to the 1-week capillary absorption time of both thenardite and mirabilite according to Flatt & Scherer
procedure (procedure 2). Additionally, these procedures (2 and 3) [39] and Haynes & Bassuoni [24].
were the only capable of bringing samples to failure, both in At the inner section sample, predominance of dehydrated mir-
84 days. That said, procedure 2 was selected as most effective abilite occurred (Fig. 14). Even though ettringite is not stable at
because it was able to provide failure of the samples with a more temperatures above 65 °C [1], two clusters of ettringite
detailed degradation history (more cycles). Procedure 4 was the (C6AS3H32) needles (Fig. 15) were identified indicating possible
least aggressive one, because from the third cycle on, just water slight occurrence of simultaneous chemical sulfate attack. Higher
was used instead of sodium sulfate solution on the capillary solution supersaturation rate facilitates thenardite precipitation
absorption steps. The surface wear of the specimens is shown in [26]. As the absorption cycles of the blocks were done by capillar-
Fig. 7. ity, the exposed surface formed a drying front and had higher
The 2 cm solution height (starting from the base of the samples) supersaturation rate compared to inner sections. Additionally,
was enough to completely wet the blocks by capillarity. Progres- the majority of the damage occurred at outer layers of the speci-
sive efflorescence was observed both at drying and capillary mens. That could be a possible explanation for thenardite abun-
absorption stages (Fig. 8). After the end of the exposure cycles, dance at surface and scarcity at inner sections.
samples were taken from efflorescence deposits and surface & cen-
tral portions of cross sections (Fig. 9) for XRD, XRF and SEM (scan-
ning electron microscope) tests. From efflorescence samples, XRD 3.2. Group B: concentration study
(Fig. 10) and XRF (Table 2) showed that it is mainly composed of
thenardite and mirabilite. According to Bassuoni & Nehdi [7], there Similar to Group A, at each step mass measurements were
is a drying front on the surface layer of the specimens which allows made, and Eq. (1) was used for determining the mass loss. The
supersaturation and crystallization of thenardite to occur. At dry- results over time are presented in Fig. 16 and Tables A.2 and A.3
ing stages, the efflorescence and surface scaling were more intense. (Appendices). As expected, the higher the sulfate concentration,
On the central portion samples (Fig. 9), a whitish precipitate the higher the wear rate is. Chemically and physically, there is
subflorescence (sodium sulfate crystals) was observed internally more precipitation and crystallization of sulfates on micropores
at the cross sections, while the outer layers had lower concentra- associated with pressure generation. The local values of tempera-
tions mainly due to efflorescence carrying the precipitates to the ture and relative humidity – 26.0 ± 2.0 °C and 75.0 ± 8.5%, respec-
external environment. Up to 1-week capillary absorption time, tively – allowed precipitation of thenardite and mirabilite
longer exposure time to sodium sulfate solutions yielded higher [24,39]. This results in microcracking, as well as leaching of cemen-
precipitation rate of sodium sulfate crystals in the samples titious materials causing diminishing cohesiveness and aggregate
(Table 3). Higher sulfate concentration was observed on procedure loss [9]. Even though the 3% and 6% sodium sulfate procedures
2 (1-week capillary absorption time) and procedure 3 (2-week cap- yielded slower wear rates than the 10% procedure, 6% sodium sul-
illary absorption time). This ratifies the similar mass loss between fate resulted in higher total mean mass loss without failure
procedures 2 and 3, as well as both reaching failure at the same (intense surface aggregate loss occurred, but no surface cracks or
time (84 days). Additionally, it also ratifies the increase of mass spalling was identified). That can be explained by higher precipita-
loss between procedure 1 and procedures 2 & 3. XRD indicated tion and crystallization rate of sulfates on the 10% concentration
presence of thenardite, kaolinite and quartz (Fig. 11). procedure, resulting in higher pressure generation on micropores

Fig. 6. Group A mass loss over time.


W.V.D.C. Bezerra, G.A. Azeredo / Construction and Building Materials 200 (2019) 255–264 259

Fig. 7. Surface wear of the specimens. From left to right: Block 2 (90 days, procedure 1); Block 5 (84 days, procedure 2); Block 9 (84 days, procedure 3); Block 11 (75 days,
procedure 4).

Fig. 10. X-ray diffraction of the efflorescence.

Table 2
X-ray fluorescence of the efflorescence.
Fig. 8. Efflorescence at both drying and capillary absorption stages.
Efflorescence
SO3 44.206
Na2O 36.797
SiO2 8.852
Al2O3 4.456
CaO 3.252
Fe2O3 1.603
TiO2 0.372
Others 0.462

Most of the damage to the blocks was caused by physical sulfate


attack, with thenardite and mirabilite precipitating into microp-
ores and generating pressure. Traces of ettringite were also found,
indicating possible slight occurrence of chemical sulfate attack.
From this study, it was possible to conclude that:

Fig. 9. Cross sections from samples showing a whitish precipitate on inner sections.  The capillary absorption time influenced the wear on samples.
In general, longer exposure intervals caused greater damage
and leading samples to failure. The surface wear of the specimens with diminishing increases. Spalling and cracking occurred on
is shown in Fig. 17. samples tested with 1 week and 2 weeks capillary absorption
time using 10% sodium sulfate concentration.
4. Conclusions  Higher sulfate concentrations caused greater damage on sam-
ples. Even though surface wear on samples tested with lower
Compressed stabilized earth block samples were exposed to concentrations (i.e. 3% and 6%) was more subtle, greater total
external sulfate attack by cycles of capillary absorption and drying. mass loss occurred with no visible spalling or cracking.
260 W.V.D.C. Bezerra, G.A. Azeredo / Construction and Building Materials 200 (2019) 255–264

Table 3
X-ray fluorescence of the central portions.

Procedure 1 Procedure 2 Procedure 3 Procedure 4


SiO2 40.115 39.983 39.521 44.823
Al2O3 19.814 19.477 19.285 22.135
CaO 16.244 15.211 16.032 13.983
SO3 8.043 8.837 8.934 5.073
Fe2O3 7.454 7.328 7.367 7.901
Na2O 5.111 6.113 5.786 2.842
TiO2 1.778 1.772 1.787 1.989
Others 1.439 1.279 1.288 1.253

Fig. 11. X-ray diffraction of block 6 (procedure 2) at central portion.

Fig. 14. SEM image of inner section sample showing silty sand with sodium sulfate
precipitates.

Fig. 12. SEM image of surface sample showing predominance of thenardite and
dehydrated mirabilite crystals.

Fig. 15. SEM image of inner section sample showing dehydrated mirabilite and
traces of ettringite crystals.

 Among the procedures tested, the authors recommend using


procedure 2 (cycles of 1 week of capillary absorption with 10%
sodium sulfate solution and 2 weeks drying) to test CSEB
against sulfate attack because it was able to lead samples to fail-
ure with a more gradual degradation history.

Due to the fact that lower sulfate concentrations yielded higher


total mass loss without failure (spalling and/or cracking) of the
samples, monitoring additional parameters (e.g. mechanical
strength) is advised when testing materials against sulfate attack.

Conflict of interest
Fig. 13. SEM image of surface sample showing the shape of thenardite and
dehydrated mirabilite crystals. None.
W.V.D.C. Bezerra, G.A. Azeredo / Construction and Building Materials 200 (2019) 255–264 261

Fig. 16. Group B mass loss over time.

Fig. 17. Surface wear of the specimens. Left: Block 14 (294 days, procedure 5); Right: Block 16 (294 days, procedure 6).

Acknowledgements Appendices

The authors are grateful for the financial support of FINEP-INO-


VATEC, to the assistance of Prof. Gibson Rocha Meira, to the Labo-
ratório de Solidificação Rápida (LSR-UFPB) and to the Laboratório de
Ensaios de Materiais e Estruturas (LABEME-UFPB).

Table A.1
Mass loss of the specimens from the procedures 1–4. M̅ = mean average of the mass loss from the 3 specimens; r = sample standard deviation of the mass loss from the 3
specimens.

DMBlock1 (%) DMBlock2 (%) DMBlock3 (%) M (%) r (%) Cumulative Days

Md1 0.0 0.0 0.0 0.0 0.0 15


Mw2 10.8 11.7 13.5 12.0 1.4 16
Md2 1.8 2.3 1.3 1.8 0.5 30
Mw3 10.9 12.0 12.1 11.7 0.6 31
Md3 0.1 0.6 1.5 0.3 1.1 45
Mw4 10.4 12.4 11.2 11.3 1.0 46
Md4 3.2 1.8 5.1 3.4 1.6 60
Mw5 6.3 9.0 6.8 7.4 1.4 61
Md5 6.6 4.7 8.9 6.7 2.1 75
Mw6 2.7 5.4 2.1 3.4 1.7 76
Md6 7.4 4.3 8.6 6.8 2.2 90


DMBlock4 (%) DMBlock5 (%) DMBlock6 (%) M (%) r (%) Cumulative Days

Md1 0.0 0.0 0.0 0.0 0.0 21


Mw2 10.8 12.3 11.2 11.4 0.8 28
Md2 0.9 0.9 2.0 1.3 0.6 42

(continued on next page)


262 W.V.D.C. Bezerra, G.A. Azeredo / Construction and Building Materials 200 (2019) 255–264

Table A.1 (continued)



DMBlock4 (%) DMBlock5 (%) DMBlock6 (%) M (%) r (%) Cumulative Days

Mw3 10.1 9.1 9.7 9.6 0.5 49


Md3 3.6 4.2 4.5 4.1 0.5 63
Mw4 8.5 8.0 7.8 8.1 0.3 70
Md4 8.5 9.8 7.8 8.7 1.0 84
Mw5 0.7 1.0 1.7 0.5 1.3 91
Md5 10.8 11.4 9.0 10.4 1.3 105

DMBlock7 (%) DMBlock8 (%) DMBlock9 (%) M (%) r (%) Cumulative Days

Md1 0.0 0.0 0.0 0.0 0.0 28


Mw2 10.1 10.8 10.1 10.3 0.4 42
Md2 2.5 3.6 2.8 3.0 0.5 56
Mw3 9.0 9.5 10.6 9.7 0.8 70
Md3 7.3 10.2 11.3 9.6 2.1 84
Mw4 1.7 0.9 3.0 0.7 2.3 98
Md4 7.6 11.3 12.2 10.4 2.4 112

DMBlock10 (%) DMBlock11 (%) DMBlock12 (%) M (%) r (%) Cumulative Days

Md1 0.0 0.0 0.0 0.0 0.0 15


Mw2 10.0 9.9 10.2 10.0 0.1 16
Md2 1.1 0.8 0.6 0.9 0.3 30
Mw3 9.3 9.1 9.6 9.3 0.2 31
Md3 4.1 3.2 3.1 3.4 0.5 45
Mw4 6.5 7.2 7.7 7.2 0.6 46
Md4 5.5 4.2 4.0 4.5 0.8 60
Mw5 5.1 6.2 6.8 6.1 0.8 61
Md5 5.5 4.2 3.8 4.5 0.9 75

Table A.2
Mass loss of procedure 5 specimens.

DMBlock13 (%) DMBlock14 (%) DMBlock15 (%) M (%) r (%) Cumulative Days

Md1 0.0 0.0 0.0 0.0 0.0 21


Mw2 9.2 10.6 9.6 9.8 0.7 28
Md2 1.1 2.0 1.2 1.4 0.5 42
Mw3 8.9 9.6 9.4 9.3 0.3 49
Md3 3.1 5.3 3.9 4.1 1.1 63
Mw4 8.1 7.1 8.5 7.9 0.7 70
Md4 2.7 5.5 3.4 3.9 1.4 84
Mw5 7.7 5.8 7.6 7.0 1.1 91
Md5 2.8 6.5 3.7 4.4 1.9 105
Mw6 7.7 5.8 8.1 7.2 1.2 112
Md6 2.8 6.5 4.0 4.5 1.9 126
Mw7 7.1 4.9 7.2 6.4 1.3 133
Md7 3.7 7.4 4.2 5.1 2.0 147
Mw8 7.0 4.6 6.9 6.1 1.4 154
Md8 4.4 8.1 5.5 6.0 1.9 168
Mw9 7.0 4.1 6.6 5.9 1.6 175
Md9 5.7 9.1 6.7 7.2 1.8 189
Mw10 6.7 4.1 6.3 5.7 1.4 196
Md10 6.0 9.4 7.0 7.5 1.8 210
Mw11 5.4 2.7 5.4 4.5 1.5 217
Md11 6.1 9.5 6.9 7.5 1.7 231
Mw12 5.0 2.7 5.2 4.3 1.4 238
Md12 6.3 9.5 6.6 7.5 1.7 252
Mw13 4.2 1.9 4.7 3.6 1.5 259
Md13 7.4 10.5 7.6 8.5 1.7 273
Mw14 3.9 1.9 4.6 3.5 1.4 280
Md14 7.8 11.1 7.9 8.9 1.9 294

Table A.3
Mass loss of procedure 6 specimens.

DMBlock16 (%) DMBlock17 (%) DMBlock18 (%) M (%) r (%) Cumulative Days

Md1 0.0 0.0 0.0 0.0 0.0 21


Mw2 8.2 8.5 8.5 8.4 0.2 28
Md2 1.5 2.2 1.9 1.9 0.3 42
Mw3 6.9 7.6 7.6 7.4 0.4 49
W.V.D.C. Bezerra, G.A. Azeredo / Construction and Building Materials 200 (2019) 255–264 263

Table A.3 (continued)



DMBlock16 (%) DMBlock17 (%) DMBlock18 (%) M (%) r (%) Cumulative Days

Md3 4.5 7.5 6.1 6.0 1.5 63


Mw4 5.2 3.3 4.6 4.4 1.0 70
Md4 4.8 9.2 6.5 6.9 2.2 84
Mw5 4.1 0.4 2.8 2.5 1.8 91
Md5 5.2 10.1 7.3 7.5 2.5 105
Mw6 4.5 0.3 3.0 2.6 2.1 112
Md6 5.9 11.2 7.7 8.3 2.7 126
Mw7 3.4 0.9 2.2 1.6 2.2 133
Md7 6.4 11.7 8.7 8.9 2.7 147
Mw8 2.7 1.6 1.6 0.9 2.3 154
Md8 7.2 12.8 10.0 10.0 2.8 168
Mw9 2.7 2.1 0.7 0.4 2.4 175
Md9 8.6 14.3 11.3 11.4 2.9 189
Mw10 2.3 2.5 0.3 0.0 2.4 196
Md10 9.6 15.5 12.2 12.4 3.0 210
Mw11 0.0 4.5 1.2 1.9 2.3 217
Md11 9.9 16.5 12.8 13.1 3.4 231
Mw12 1.0 5.2 2.1 2.8 2.2 238
Md12 11.2 16.7 13.5 13.8 2.8 252
Mw13 1.4 6.8 3.4 3.9 2.7 259
Md13 12.4 18.6 14.9 15.3 3.1 273
Mw14 2.1 7.6 4.0 4.5 2.8 280
Md14 13.0 19.6 16.7 16.4 3.3 294

References [17] P.J. Walker, Strength, durability and shrinkage characteristics of cement
stabilised soil blocks, Cem. Concr. Compos. 17 (1995) 301–310, https://doi.org/
10.1016/0958-9465(95)00019-9.
[1] P.K. Mehta, P.J.M. Monteiro, Concrete: Microstructure, Properties, and
[18] V. Rigassi, Compressed Earth Blocks: Manual of Production, 1985.
Materials, McGraw-Hill Companies, Inc., 2006, p. 684, https://doi.org/
[19] B. Taallah, A. Guettala, The mechanical and physical properties of compressed
10.1036/0071462899.
earth block stabilized with lime and filled with untreated and alkali-treated
[2] M. Romer, L. Holzer, M. Pfiffner, Swiss tunnel structures: concrete damage by
date palm fibers, Constr. Build. Mater. 104 (2016) 52–62, https://doi.org/
formation of thaumasite, Cem. Concr. Compos. 25 (2003) 1111–1117, https://
10.1016/j.conbuildmat.2015.12.007.
doi.org/10.1016/S0958-9465(03)00141-0.
[20] A.R.A.G. Pinto, Durabilidade e resistência de matriz de solo estabilizada com
[3] N.J. Crammond, The thaumasite form of sulfate attack in the UK, Cem. Concr.
resina de mamona e fibras de pupunha para uso em construções com terra
Compos. 25 (2003) 809–818, https://doi.org/10.1016/S0958-9465(03)00106-9.
crua, Pontifícia Universidade Católica do Rio de Janeiro, 2013.
[4] F. Bellmann, W. Erfurt, H.M. Ludwig, Field performance of concrete exposed to
[21] H.B. Nagaraj, M.V. Sravan, T.G. Arun, K.S. Jagadish, Role of lime with cement in
sulphate and low pH conditions from natural and industrial sources, Cem.
long-term strength of compressed stabilized earth blocks, Int. J. Sustain. Built
Concr. Compos. 34 (2012) 86–93, https://doi.org/10.1016/j.
Environ. 3 (2014) 54–61, https://doi.org/10.1016/j.ijsbe.2014.03.001.
cemconcomp.2011.07.009.
[22] M. Collepardi, A state-of-the-art review on delayed ettringite attack on
[5] N. Crammond, The occurrence of thaumasite in modern construction – a
concrete, Cem. Concr. Compos. 25 (2003) 401–407, https://doi.org/10.1016/
review, Cem. Concr. Compos. 24 (2002) 393–402, https://doi.org/10.1016/
S0958-9465(02)00080-X.
S0958-9465(01)00092-0.
[23] M.T. Bassuoni, M.M. Rahman, Response of concrete to accelerated physical salt
[6] H. Haynes, R. O’Neill, M. Neff, R. Kumar Mehta, Salt weathering distress on
attack exposure, Cem. Concr. Res. 79 (2016) 395–408, https://doi.org/10.1016/
concrete exposed to sodium sulfate environment, ACI Mater. J. 105 (2008) 35–
j.cemconres.2015.02.006.
43, https://doi.org/10.14359/19205.
[24] H. Haynes, M.T. Bassuoni, Physical salt attack on concrete, Concr. Int. 33 (2011)
[7] M.T. Bassuoni, M.L. Nehdi, Durability of self-consolidating concrete to different
38–42.
exposure regimes of sodium sulfate attack, Mater. Struct. Constr. 42 (2009)
[25] W. Müllauer, R.E. Beddoe, D. Heinz, Sulfate attack expansion mechanisms,
1039–1057, https://doi.org/10.1617/s11527-008-9442-2.
Cem. Concr. Res. 52 (2013) 208–215, https://doi.org/10.1016/j.
[8] R. El-Hachem, E. Rozire, F. Grondin, A. Loukili, New procedure to investigate
cemconres.2013.07.005.
external sulphate attack on cementitious materials, Cem. Concr. Compos. 34
[26] C. Rodriguez-Navarro, E. Doehne, E. Sebastian, How does sodium sulfate
(2012) 357–364, https://doi.org/10.1016/j.cemconcomp.2011.11.010.
crystallize? Implications for the decay and testing of building materials, Cem.
[9] G. Massaad, E. Rozière, A. Loukili, L. Izoret, Advanced testing and performance
Concr. Res. 30 (2000) 1527–1534.
specifications for the cementitious materials under external sulfate attacks,
[27] J.P. Skalny, J. Marchand, I. Odler, Sulfat Attack on Conrete, 2003.
Constr. Build. Mater. 127 (2016) 918–931, https://doi.org/10.1016/
https://books.google.com/books?id=0ZN-z35qV1sC&pgis=1.
j.conbuildmat.2016.09.133.
[28] M. Whittaker, L. Black, Current knowledge of external sulfate attack, Adv. Cem.
[10] I. Alam, A. Naseer, A.A. Shah, Economical stabilization of clay for earth
Res. 27 (2015) 532–545, https://doi.org/10.1680/adcr.14.00089.
buildings construction in rainy and flood prone areas, Constr. Build. Mater. 77
[29] ISO, ISO 14688-1: Geotechnical investigation and testing — Identification and
(2015) 154–159, https://doi.org/10.1016/j.conbuildmat.2014.12.046.
classification of soil — Part 1: Identification and description, (2017).
[11] Q. Piattoni, E. Quagliarini, S. Lenci, Experimental analysis and modelling of the
[30] ABNT, NBR 10833: Fabricação de tijolo de solo-cimento com a utilização de
mechanical behaviour of earthen bricks, Constr. Build. Mater. 25 (2011) 2067–
prensa manual ou hidráulica - Procedimento, 2013.
2075, https://doi.org/10.1016/j.conbuildmat.2010.11.039.
[31] L. Binda, P. Bekker, G. Borchelt, N. Bright, F. Emrich, M. Forde, H. Gallegos, C.
[12] F. Pacheco-Torgal, S. Jalali, Earth construction: lessons from the past for future
Groot, E. Hedstrom, S. Lawrence, P. Maurenbrecher, C. Modena, A. Page, F.
eco-efficient construction, Constr. Build. Mater. 29 (2012) 512–519, https://
Pires, C. Republic, J. Roberts, S. Schmidt, P. Schubert, M. Schuller, J. Schwartz, T.
doi.org/10.1016/j.conbuildmat.2011.10.054.
West, RILEM TC 127-MS-A.1 Determination of the resistance of wallettes
[13] R.A. Bradley, M. Gohnert, I. Bulovic, A.M. Goliger, D.B. Surat, Steep catenary
against sulphates and chlorides, Mater. Struct. 31 (1998) 2–19, https://doi.org/
earth-brick shells as a low-cost housing solution, J. Archit. Eng. 23 (2017)
10.1007/BF02486406.
04016018, https://doi.org/10.1061/(ASCE)AE.1943-5568.0000234.
[32] ABNT, NBR 9779: Argamassa e concreto endurecidos - Determinação da
[14] O. Izemmouren, A. Guettala, S. Guettala, Mechanical properties and durability
absorção de água por capilaridade, (1995) 2.
of lime and natural Pozzolana stabilized steam-cured compressed earth block
[33] T.P.H. van den Brand, K. Roest, G.-H. Chen, D. Brdjanovic, M.C.M. van
bricks, Geotech. Geol. Eng. 33 (2015) 1321–1333, https://doi.org/10.1007/
Loosdrecht, Long-term effect of seawater on sulfate reduction in wastewater
s10706-015-9904-6.
treatment, Environ. Eng. Sci. 32 (2015) 622–630, https://doi.org/10.1089/
[15] A.O. Olotuah, Recourse to earth for low-cost housing in Nigeria, Build. Environ.
ees.2014.0306.
37 (2002) 123–129, https://doi.org/10.1016/S0360-1323(00)00081-0.
[34] D.E. Canfield, J. Farquhar, Animal evolution, bioturbation, and the sulfate
[16] J.E. Oti, J.M. Kinuthia, J. Bai, Compressive strength and microstructural analysis
concentration of the oceans, Proc. Natl. Acad. Sci. 106 (2009) 8123–8127,
of unfired clay masonry bricks, Eng. Geol. 109 (2009) 230–240, https://doi.org/
https://doi.org/10.1073/pnas.0902037106.
10.1016/j.enggeo.2009.08.010.
[35] E.G. Swenson, Concrete in sulphate environments, Can. Build. Dig. (1971).
264 W.V.D.C. Bezerra, G.A. Azeredo / Construction and Building Materials 200 (2019) 255–264

[36] F. Mittermayr, A. Baldermann, C. Kurta, T. Rinder, D. Klammer, A. Leis, J. [38] T. Aye, C.T. Oguchi, Y. Takaya, Evaluation of sulfate resistance of Portland and
Tritthart, M. Dietzel, Evaporation – a key mechanism for the thaumasite form high alumina cement mortars using hardness test, Constr. Build. Mater. 24
of sulfate attack, Cem. Concr. Res. 49 (2013) 55–64, https://doi.org/10.1016/j. (2010) 1020–1026, https://doi.org/10.1016/j.conbuildmat.2009.11.016.
cemconres.2013.03.003. [39] R.J. Flatt, G.W. Scherer, Hydration and crystallization pressure of sodium
[37] ACI Committee, ACI 318M-05: Building code requirements for structural sulfate: a critical review, in: Mat. Res. Soc. Symp. Proc., 2002, https://doi.org/
concrete and commentary, 2005. 10.1557/PROC-712-II2.2.

You might also like