You are on page 1of 11

Cement & Concrete Composites 62 (2015) 33–43

Contents lists available at ScienceDirect

Cement & Concrete Composites


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

Relation between carbonation resistance, mix design and exposure


of mortar and concrete
Andreas Leemann ⇑, Peter Nygaard, Josef Kaufmann, Roman Loser
Empa, Swiss Federal Laboratories for Materials Science and Technology, Überlandstr. 129, 8600 Dübendorf, Switzerland

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

Article history: When cement with mineral additions is employed, the carbonation resistance of mortar and concrete
Received 21 July 2014 may be decreased. In this study, mortars containing mineral additions are exposed both to accelerated
Received in revised form 8 April 2015 carbonation (1% and 4% CO2) and to natural carbonation. Additionally, concrete mixtures produced with
Accepted 23 April 2015
different cements, water-to-cement ratios and paste volumes are exposed to natural carbonation. The
Available online 14 June 2015
comparison of the carbonation coefficients determined in the different exposure conditions indicates that
mortar and concrete containing slag and microsilica underperform in the accelerated carbonation test
Keywords:
compared to field conditions. The carbonation resistance in mortar and concrete is mainly governed by
Carbonation
Accelerated testing
the CO2 buffer capacity per volume of cement paste. It can be expressed by the ratio between water added
Natural exposure during production and the amount of reactive CaO present in the binder (w/CaOreactive) resulting in a
Mix design novel parameter to assess carbonation resistance of mortar and concrete containing mineral additions.
Mortar Ó 2015 Elsevier Ltd. All rights reserved.
Concrete

1. Introduction binder-specific effects of curing on carbonation. Additionally, it


has to be taken into account that the effect of curing is dependent
The increasing use of minerals additions, e.g., ground granu- on the water-to-cement-ratio (w/c), with a stronger effect on mor-
lated blast-furnace slag (GGBS), fly ash or limestone powder, to les- tar and concrete produced with high w/c [23,24]. Such effects may
sen the CO2 emission of cementitious binders affects the durability be the reason why no material parameter governing the carbona-
of mortar and concrete. It can both improve and worsen their per- tion resistance of systems with different mineral additions has
formance depending on the type of interaction [e.g. 1–5]]. In regard been identified yet.
to carbonation resistance, a general decrease is observed when This study aims at empirically identifying a material parameter
cement clinker is replaced by mineral additions [6–11]. An (or a number of parameters) that defines the carbonation resis-
increased risk for rebar corrosion in reinforced concrete structures tance of mortar and concrete produced with cements containing
due to carbonation of the concrete cover may result. mineral additions. To ensure optimal curing, all samples were
Accelerated tests can be used to assess the carbonation resis- moist cured for 28 days before starting exposure to CO2. In a first
tance of mortar or concrete produced with cements containing series of experiments, mortars were produced. Their buffer capac-
mineral additions. However, a wide range of CO2 concentrations, ity and diffusivity were systematically changed by addition of
different relative humidity conditions and temperatures are used limestone powder, microsilica, portlandite and the use of GGBS
in such accelerated tests [8,10,12–14]. Therefore it is not clear, cement. The carbonation resistance was determined under acceler-
whether the results obtained with one specific accelerated test ated (1% and 4% CO2) and natural conditions (sheltered and unshel-
are transferable to another one and, even more important, to nat- tered). Additionally, compressive strength and oxygen diffusion
ural carbonation [15–20]. Another important point is the curing of were determined. Selected samples were further investigated with
the samples before they are exposed to carbonation. Carbonation mercury intrusion porosimetry (MIP) and scanning electron micro-
resistance increases with increased time of curing [6,15,21,22], as scopy (SEM), combined with energy dispersive X-ray spectroscopy
longer curing allows a higher degree of hydration of the binder. (EDX). In a second series of experiments, concrete mixtures pro-
However, the reactivity of cement clinker and mineral additions duced with different w/c, cement type and paste volume were
like GGBS or fly ash can be quite different, leading to exposed to natural carbonation for 2.5 years. Furthermore, com-
pressive strength and oxygen diffusion were measured.
⇑ Corresponding author. Tel.: +41 58 765 44 89.
E-mail address: andreas.leemann@empa.ch (A. Leemann).

http://dx.doi.org/10.1016/j.cemconcomp.2015.04.020
0958-9465/Ó 2015 Elsevier Ltd. All rights reserved.
34 A. Leemann et al. / Cement & Concrete Composites 62 (2015) 33–43

2. Materials and methods The same mortars were used for the MIP analysis. Samples were
extracted with pincers from carbonated and non-carbonated areas
2.1. Materials of mortar bars after an exposure in the carbonation chambers at 4%
CO2 for 35 or 91 days depending on the progress of carbonation.
2.1.1. Mortars Afterwards, the pieces were stored in isopropanol for 7 days to
The composition of the ordinary Portland cement (OPC/CEM I remove the water and then dried at 50 °C for another 7 days.
52.5 R), the GGBS cement (CEM III/B 42.5, GGBS content of 66– The calculation of the reactive calcium oxide (CaOreactive) and
80 mass-%) and the microsilica used for the mortars are given in the reacted CaO (CaOreacted) in the mortars displayed in Table 2 is
Table 1. The limestone powder used contains 92 mass-% of calcite, explained in Section 4.
5.5 mass-% of dolomite and minor amounts of quartz and mica. A
laboratory grade portlandite was used. The aggregate was standard 2.1.2. Concrete
quartz sand with a grain size of 0–2 mm. Two ordinary Portland cements (CEM I 32.5 R, CEM I 42.5 N HS),
The mix design of the mortars (Table 2) was systematically var- a cement containing 15 mass-% limestone powder (CEM II/A-LL), a
ied to study the effect of changing diffusivity and buffer capacity cement containing approximately 15 mass-% of limestone powder
on the carbonation resistance. By increasing the w/c of the mix- and 20 mass-% low-calcium fly ash (CEM II/B-M (V-LL) 32.5 R) and
tures produced solely with OPC and by replacing cement with a GGBS cement (CEM III/B 32.5 N) were used to produce the con-
limestone powder, the diffusivity was increased and the buffer crete. The composition of the five cement types used is given in
capacity decreased. Cement was replaced with microsilica to Table 3. The mix design of the concrete is given in Table 4. The
decrease both diffusivity and buffer capacity. The replacement of cubes with a side length of 150 mm were demolded after 24 h
cement with portlandite was expected to slightly increase the buf- and subsequently moist cured (>95% relative humidity) until the
fer capacity and the diffusivity. The volume of paste was kept con- age of 28 days.
stant for all mixtures and the water-to-binder ratio by mass (w/b) As the concrete mixtures were produced for another project
was the same for all mortars containing mineral additions. The dealing with sulfate resistance [25], only one cube per mixture
names of the mortars (Table 2) and the concrete mixtures was available for testing the carbonation resistance. At the age of
(Tables 3 and 4) refer to the mineral additions used and their 28 days, every cube was cut in the direction of casting in two
approximate content in mass-%. halves of identical dimensions. All sides except the cut surface
The mortars were produced according to EN 197-1 using a and the opposite cast surface were coated with epoxy. One half
Hobart mixer. From each mixture 15 mortar prisms of the cube was stored unsheltered on a roof and the other half
(40  40  160 mm3) and one cube (150  150  150 mm3) were sheltered under a roof (identical storing conditions to the mortar
produced to measure compressive strength (3 prisms), carbonation prisms).
resistance (12 prisms) and oxygen diffusion (cube). All samples At 28 days, cores (diameter of 100 mm, height of 50 mm) were
were demolded after 24 h and subsequently moist cured (>95% rel- taken from one cube to determine oxygen diffusion. Conditioning
ative humidity) until the age of 28 days. was identical to the mortar samples. The mass change during con-
After the moist curing, the mortar prisms used to test the car- ditioning was recorded to obtain information about the drying
bonation resistance were conditioned at 20 ± 1 °C and 57 ± 2% rel- behavior of the different concrete mixtures.
ative humidity until the age of 56 days. Then, the upper and lower The calculation of the reactive calcium oxide (CaOreactive) and
sides of the prisms were coated with epoxy to allow CO2 ingress the reacted CaO (CaOreacted) in the concrete mixtures shown in
only from the lateral cast surfaces. Subsequently, each of the four Table 4 is explained in Section 4.
sets consisting of three prisms were moved to one of the following
exposure conditions: carbonation chamber with 1% CO2 (20 °C, 57% 2.2. Methods
RH), carbonation chamber with 4% CO2 (20 ± 1 °C, 57 ± 2% RH),
sheltered outdoor exposure and unsheltered outdoor exposure. Compressive strength of the mortar prisms and the concrete
The samples in the sheltered outdoor exposure were stored in a cubes were determined according to DIN EN 1015-11 [26] and
shelter having one side entirely open. The unsheltered exposure EN 12390-3 [27] respectively.
was on a roof where the samples were placed on steel profiles to The procedure for the determination of carbonation coefficient
prevent capillary suction of water accumulated on the roof top. is based on SN 505 262/1 [28]. Before transferring the 12 mortar
Cores (diameter of 100 mm, height of 30 mm) were taken from prisms per mixture to the four different exposure conditions, the
the cube after 28 days to measure oxygen diffusion. The cores were initial carbonation depth was determined. An approximately
conditioned at 20 ± 1 °C and 35 ± 2% RH for seven days and then 2.5 cm thick slice was split of the prisms. The freshly broken sur-
dried in an oven at 50 °C for 24 h before the measurement started. face was sprayed with phenolphthalein and photographed.
The mass change during conditioning was recorded to obtain infor- Carbonation depth was analyzed with image analysis on 10 points
mation on the drying behavior of the mortars. per lateral cast surface. As three prisms were used for each of the
The samples for the investigation with the SEM were taken from four different exposure conditions, the total number of measure-
the mortar bars in the carbonation chamber after an exposure at 4% ments per exposure condition, age and mixture was 60. The mea-
CO2 for 35 days (mortars OPC-0.48, MS-15, P-15 and S-65). The surements on the concrete were performed the same way. After
samples were dried in an oven at 50 °C for three days, epoxy the initial measurement, the carbonation depth of the samples in
impregnated and polished. carbonation chambers was measured after an exposure of 7, 35

Table 1
Composition of the two cements and the microsilica used for the mortars.

Material CaO SiO2 Al2O3 Fe2O3 Cr2O3 MnO TiO2 P2O5 MgO K2O Na2O SO3 L.O.I.
CEM I 52.5 R 62.7 20.2 5.29 3.25 0.01 0.06 0.29 0.22 1.65 0.90 <0.03 3.82 1.34
CEM III/B 42.5 L-LH HS 48.8 30.6 9.73 1.18 n.a. n.a. n.a. n.a. 5.19 0.59 0.29 1.43 n.a.
Microsilica 0.36 92.8 0.94 0.13 n.a. 0.03 0.01 0.05 0.74 1.09 0.39 0.54 2.66

L.O.I. = loss on ignition.


n.a. = not analyzed.
A. Leemann et al. / Cement & Concrete Composites 62 (2015) 33–43 35

Table 2
Mix design of the mortars.

Mortar Cement Mineral additions w/b Sand [kg/m3] CaOreactive [kg/m3] CaOreacted [kg/m3]
3
Type Mass [kg/m ] Type Mass
[kg/m3] [mass-%]
OPC-48 CEM I 518 – – – 0.48 1448 316 246
OPC-40 CEM I 575 – – – 0.40 1449 351 250
OPC-55 CEM I 471 – – – 0.56 1448 287 235
L-7.5 CEM I 480 Limestone 36.0 7.5 0.48 1447 293 228
L-15 CEM I 446 Limestone 66.9 15.0 0.48 1450 272 212
MS-7.5 CEM I 479 Microsilica 35.9 7.5 0.48 1448 292 228
MS-15 CEM I 445 Microsilica 66.8 15.0 0.48 1449 271 211
P-7.5 CEM I 476 Portlandite 35.7 7.5 0.48 1449 317 253
P-15 CEM I 441 Portlandite 66.1 15.0 0.48 1448 319 260
S-65 CEM III/B 506 – – – 0.48 1448 238 144

Table 3
Composition of the cements used for the concrete mixtures.

Cement Abbreviation used for concrete CaO SiO2 Al2O3 Fe2O3 MgO K2O Na2O SO3 L.O.I.a
CEM I 32.5 R OPC 64.1 20.7 4.7 3 2.2 0.9 0.2 2.4 0.9
CEM I 42.5 R HS OPC-HS 59.3 19.2 3.4 5.4 4 0.7 0.3 3.2 3.8
CEM II/A-LL 42.5 N L-15 61 17.8 4.3 2.7 1.7 1 0.2 2.8 7.8
CEM II/B-M 32.5 R L-FA-35 52.3 23.6 6.9 3.8 1.7 1.2 0.3 2.5 6.9
CEM III/B 32.5 N S-65 46.4 30.1 10 1.3 5.2 0.7 0.4 4.1 1.1
a
L.O.I. = loss on ignition.

Table 4
Mix design of the concrete.

Concrete /cement [–] Aggregate [kg/m3] ms/ga [–] Cement [kg/m3] Water [kg/m3] w/c [–] Addition [kg/m3] Paste volume [l/m3] CaOreactive [kg/m3] CaOreacted [kg/m3]
OPC 1929 0.54 335 150 0.45 1.3SPb 257 212 178
OPC 1929 0.54 315 157 0.50 0.3SP 257 199 175
OPC 1931 0.54 280 168 0.60 – 257 177 165
OPC 1799 0.54 400 180 0.45 – 307 253 212
OPC 1680 1.00 450 200 0.45 1.8SP + 1.1VMAc 343 284 239
OPC-HS 1927 0.54 335 150 0.45 1.3SP 257 189 159
OPC-HS 1927 0.54 315 157 0.50 0.6SP 257 178 157
OPC-HS 1929 0.54 280 168 0.60 – 257 158 147
OPC-HS 1678 1.00 450 200 0.45 2.3SP + 1.1VMA 343 254 214
L-15 1930 0.54 330 148 0.45 0.7SP 257 174 146
L-15 1931 0.54 310 155 0.50 0.3SP 257 163 143
L-15 1923 0.54 280 168 0.60 – 257 147 137
L-15 1679 1.00 445 197 0.45 1.8SP + 1.1VMA 343 234 197
L-FA-35 1924 0.54 325 146 0.45 1.0SP 257 143 120
L-FA-35 1927 0.54 305 152 0.50 0.3SP 257 134 118
L-FA-35 1922 0.54 275 165 0.60 – 257 212 112
L-FA-35 1677 1.00 435 193 0.45 2.2SP + 1.1VMA 343 191 160
S-65 1928 0.54 325 146 0.45 1.3SP 257 151 107
S-65 1930 0.54 305 152 0.50 0.6SP 257 142 106
S-65 1925 0.54 275 165 0.60 – 257 128 101
S-65 1794 0.54 390 174 0.45 – 307 181 129
S-65 1672 1.00 440 195 0.45 SP + VMA 343 204 145
a
ms/g = mass ratio between sand and gravel.
b
SP = polycarboxylate-based superplaticizer.
c
VMA = viscosity modifying agent.

and 91 days. The carbonation depth of the mortars exposed out- where K is the carbonation coefficient, dK the carbonation depth in
doors was determined after an exposure of one year using the mm, A the initial carbonation depth in mm after curing and t the
same technique as for the samples stored in the carbonation cham- time in years.
bers. The measurements on the concrete were performed after The conditions during the entire outdoor exposure of the mor-
2.5 years (10 measurements per concrete). The carbonation coeffi- tars and part of the exposure of the concrete samples were
cient K was determined by calculating the regression of the car- recorded in a meteorological station 50 m from the exposure site
bonation depth as a function of the square root of time (Eq. (1)): (Fig. 1). The relative humidity was the highest during the winter
month, as there were several days per month with fog. During
ðdK  AÞ
K¼ p ð1Þ these days the recorded relative humidity reached 100%. The low-
t
est amount of precipitation was recorded during the winter month.
36 A. Leemann et al. / Cement & Concrete Composites 62 (2015) 33–43

25 Oxygen diffusion of the concrete shows the same


Temperature 160
cement-specific relation to w/c as compressive strength
Relative humidity
140 (Table 6). Concrete S-65 displays significantly lower values com-

Relative humidity [%],


20 Precipitation
Temperature [°C]

pared to the other mixtures at equal w/c.

precipitation [mm]
120
15 100
3.3. Carbonation coefficients of mortar and concrete
80
10
60 Decreasing the w/b in the OPC mortars decreases the carbona-
40
tion coefficient K determined in accelerated and natural conditions.
5 The cement replacement with limestone powder, microsilica and
20 GGBS generally leads to more pronounced increases of the carbon-
0 0 ation coefficient K than the use of portlandite.
Concrete OPC and OPC-HS show the lowest carbonation coeffi-
cients. The values increase with increasing clinker substitution by
mineral additions. Consequently, the highest carbonation coeffi-
Fig. 1. Monthly means of temperature, relative humidity and precipitation during
the outdoor exposure of the mortar and partly of the concrete samples (exposed cients KN are recorded for concrete S-65. The carbonation coeffi-
2.5 years). cients KN,S in the sheltered conditions are higher than KN,US in the
unsheltered conditions for the cast and the cut surfaces of the con-
crete mixtures. The values for the cut surface are lower than for the
cast ones. As the carbonation coefficients are based on only 10 mea-
The oxygen diffusion coefficient DO was measured as described surements per concrete mixture and condition, a certain variability
in [29–31] on two cores (diameter of 100 mm, height of 30 mm) for of the values has to be expected. The mean of the coefficient of vari-
the mortar mixtures and on three cores (diameter of 100 mm, ation for the carbonation coefficients is 14.7% (cast and cut surface
height of 50 mm) for the concrete mixtures. An oxygen flow was in sheltered and unsheltered exposure of all concrete mixtures).
applied on one side of the cores and a nitrogen flow on the other
side. The gas pressure on both sides of the cores was identical. 3.4. Porosity of mortar
The oxygen content in the nitrogen flow was measured until equi-
librium was reached. Afterwards, the oxygen diffusion coefficient The total porosity and the median pore size of the mortar with
was calculated according to [28]. portlandite are the highest, while the use of microsilica and GGBS
MIP was performed with a maximum pressure of lead to a decrease of both parameters (Table 7). Carbonation causes
pmax = 395 MPa. The pore structure derived from the first intrusion a decrease in the total porosity of all mortars except mortar S-65,
was used to characterize the mortars [32]. where an increase is observed (Fig. 2). The median pore size
The analysis of the hydrates in the non-carbonated and the car- increases significantly with carbonation in mortars MS and S,
bonated areas of the mortars were conducted with an environmen- whereas only slight changes are evident in the other mortars.
tal scanning electron microscope (ESEM-FEG XL30). The samples However, when not only the median Md but the entire pore size
were studied in the high vacuum mode (2.0–6.0 106 Torr) with distribution is considered, a marked increase in the volume of
an accelerating voltage of 12 kV and a beam current of 270– pores >100 nm is observed in all mortars (Table 7). Mortars
280 lA. The composition of the hydrates was determined with OPC-48 and S-65 clearly illustrate these effects of carbonation on
energy dispersive X-ray spectroscopy (EDX). An EDAX 194 UTW the pore structure (Fig. 2).
detector, a Philips digital controller and Genesis Spectrum
Software (Version 4.6.1) with ZAF corrections was used. A total
3.5. Microstructure of the mortars
of 160 points were analyzed with EDX in the non-carbonated
and carbonated area of each studied mortar.
When the microstructure of the mortars is assessed qualitatively
with SEM, a densification is observed, with decreased porosity due
3. Results to carbonation. An exception is mortar S-65, where an increase of
porosity is visible (Fig. 3). The EDX point analysis of the cement
3.1. Compressive strength of mortar and concrete paste shows the same trend in regard to the Si/Ca-ratio in all stud-
ied mortars. In the non-carbonated paste, C–S–H displays a molar
The compressive strength increases with decreasing w/b of the Si/Ca-ratio that ranges from 0.67 (mortar MS-15) to 0.37 (mortar
OPC mortars (Table 5). The replacement of cement with limestone P-15), depending on the mineral addition used. Due to the size of
powder, portlandite and GGBS leads to decrease of compressive the interaction volume of the EDX point analyses (1–3 lm) at the
strength at 28 days, while the replacement with microsilica chosen acceleration voltage, there is some intermixing of the C–
increases it. After 91 days, mortar S-65 reaches the same compres- S–H with the mineral additions. However, it is still possible to show
sive strength as mortar OPC-48. the effect of carbonation on the mineral assemblage. In the carbon-
The concrete mixtures show the usual trend of increasing com- ated areas of the mortars, a higher variation in the Si/Ca-ratio is
pressing strength with decreasing w/c (Table 6). The compressive observed, with both a trend away from the initial C–S–H composi-
strength at a given w/c decreases a few MPa when cements con- tion towards calcite and a trend towards a silicon-rich phase. This
taining mineral additions are used. can be illustrated by the example of mortar MS-15 (Fig. 4).

3.2. Oxygen diffusion of mortar and concrete 4. Discussion

The decrease of w/b of the mortars OPC leads to a decrease of 4.1. Accelerated and natural carbonation
the oxygen diffusion coefficient DO2 (Table 5). The use of limestone
powder and portlandite increases the values, while the use of The carbonation coefficients KACC of the mortars determined in
microsilica and GGBS decreases them. the carbonation chambers with 1% and 4% CO2 show a linear
A. Leemann et al. / Cement & Concrete Composites 62 (2015) 33–43 37

Table 5
Results of the mortar mixtures.

Mortar OPC-48 OPC-40 OPC-55 L-7.5 L-15 MS-7.5 MS-15 P-7.5 P-15 S-65
fc,7 [MPa] 37.8 51.8 26.3 33.3 30.7 36.6 35 33.5 31.3 14
fc,28 62.2 74.7 50.9 57.2 51.6 67.6 69.2 53.8 49 52.9
fc,91 69.7 78.8 54.9 62.6 57.2 77.2 79.0 60.6 54 69.1
DO2 [108 m/s2] 1.92 1.24 2.02 2.39 3.00 0.72 0.50 2.08 2.45 0.19
Dm [mass-%] 0.97 0.70 1.25 0.98 1.21 0.70 0.60 0.94 1.23 0.80
p
KACC,1% [mm/ y 6.44 4.29 10.43 8.37 9.37 9.03 10.29 7.39 7.85 15.71
KACC,4% 10.34 5.25 19.34 13.17 18.46 13.54 16.05 10.52 12.29 28.67
KN,S 0.87 0.49 1.63 1.42 1.69 1.24 1.35 0.91 1.22 1.46
KN,US 0.32 0.15 0.90 0.98 1.28 0.42 0.50 0.46 0.70 0.55

fc,X = compressive strength after X days.


DO2 = oxygen diffusion coefficient.
Dm = mass loss during conditioning for 7 days at 20 °C and 35% RH
KACC,X% = coefficient of carbonation in carbonation chamber with X% CO2
KN = coefficient of carbonation in sheltered (S) and unsheltered (US) natural conditions

Table 6
Results of the concrete mixtures.

Concrete/cement w/c [–] Paste volume flow fc,28 DO2 Dm KN,S-CA KN,S-CUT KN,US-CA KN,US-CUT
p
[l/m3] [cm] [MPa] [108 m/s2] [mass-%] [mm/ y]
OPC 0.45 257 59 56.3 0.87 0.6 2.1 1.1 0.5 0.5
OPC 0.50 257 51 45.8 1.22 0.79 2.9 1.8 0.5 0.5
OPC 0.60 257 52 38.9 1.76 0.83 2.1 1.6 0.5 0.5
OPC 0.45 307 62 46.6 1.47 0.69 2.4 1.6 0.5 0.5
OPC 0.45 343 70 55.5 0.84 0.67 4.3 3.4 1.6 0.9
OPC-HS 0.45 257 42 58.7 1.24 0.62 4.4 3.0 1.0 0.6
OPC-HS 0.50 257 46 47.2 2.26 0.79 3.7 3.2 2.2 1.1
OPC-HS 0.60 257 49 34.0 4.03 1.10 6.2 4.8 3.6 2.6
OPC-HS 0.45 343 50 56.7 1.51 0.76 3.1 1.9 0.5 0.5
L-15 0.45 257 42 49.7 0.97 0.67 3.7 3.5 2.0 0.9
L-15 0.50 257 50 42.4 1.38 0.69 4.1 3.1 2.5 1.7
L-15 0.60 257 56 33.2 2.37 0.99 4.7 3.4 2.2 1.6
L-15 0.45 343 55 51.7 1.17 0.89 8.4 5.2 3.9 3.0
L-FA-35 0.45 257 49 44.3 1.20 0.71 5.4 4.0 2.4 1.4
L-FA-35 0.50 257 48 38.2 1.57 0.83 5.1 3.0 1.7 1.3
L-FA-35 0.60 257 54 28.2 3.01 0.91 6.1 4.1 3.4 2.2
L-FA-35 0.45 343 66 46.3 1.69 0.90 6.6 6.5 5.4 3.3
S-65 0.45 257 45 56.2 0.21 0.3 4.8 4.1 3.0 2.1
S-65 0.50 257 43 43.3 0.62 0.50 5.0 3.7 2.0 1.1
S-65 0.60 257 46 33.0 0.78 0.79 6.0 4.6 2.6 2.0
S-65 0.45 307 43 47.2 0.47 0.51 8.1 6.0 2.3 1.5
S-65 0.45 343 53 57.9 0.32 0.47 8.7 6.7 4.7 3.4

fc,28 = compressive strength after 28 days.


DO2 = oxygen diffusion coefficient.
Dm = mass loss during conditioning for 7 days at 20 °C and 35 RH.
KN = coefficient of natural carbonation in sheltered (S) and unsheltered (US) of the cast (CA) and cut (CUT) surfaces.

Table 7 45
Porosity determined with MIP of selected non-carbonated (nc) and carbonated (c)
OPC-48, nc
samples. 40 OPC-48, c
S-65, nc
Pore volume [mm3/g]

Mortar OPC-48 L-15 MS-15 P-15 S-65 35


S-65, c
Total porosity Unc 3
[mm /g] 44.3 46.5 39.7 56.5 38.3 30
Porosity > 100 nm Unc,100 [mm3/g] 2.9 3.8 5.4 4.7 2.3 25
Median Mdnc [nm] 31.0 30.7 15.4 34.2 7.1
Total porosity Uc [mm3/g] 27.4 38.3 27.3 37.0 45.1 20
Porosity > 100 nm Unc,100 [mm3/g] 8.8 10.7 14.1 10.7 20.2 15
Median Mdc [nm] 39.7 43.3 111.0 32.3 84.8
10
5
0
1 10 100 1000
relationship between each other (Fig. 5A). As a consequence, the
Pore radius [nm]
ranking of the different mortars in order of carbonation resistance
is practically the same at these two different CO2 concentrations. Fig. 2. Pore size distribution of the mortars OPC-48 and S-65 in the non-carbonated
There is still a reasonable correlation when KACC,1% is compared to (nc) and carbonated (c) state.
the carbonation coefficient KN,S (Fig. 5B). However, there is change
in the ranking of the mortars. The carbonation coefficient KN,S of conditions at 1% CO2. However, the most pronounced change con-
the mortars MS is lower than the one of the mortars L and mortar cerns mortar S-65. Its carbonation coefficients KACC at 1% and 4%
OPC-55. The ranking was the opposite under accelerated CO2 were considerably higher than the ones of the other mortars.
38 A. Leemann et al. / Cement & Concrete Composites 62 (2015) 33–43

A 30
OPC L

Carbonation coefficient K ACC,4%


25 MS P
S
20

[mm/y1/2]
15

10

0
0 5 10 15
Carbonation coefficient K ACC,1% [mm/y1/2]

B 18
OPC L
16

Carbonation coefficient K ACC,1%


MS P
14 S
12
10

[mm/y1/2]
8
6
4
2
0
0.0 0.5 1.0 1.5 2.0
Carbonation coefficient K N,S [mm/y1/2]

Fig. 5. (A) Carbonation coefficient KACC at 4% CO2 versus carbonation coefficient KACC
at 1% CO2 (R2 = 0.96). (B) Carbonation coefficient KACC at 1% CO2 versus carbonation
Fig. 3. Non-carbonated (A) and carbonated (B) cement paste in mortar S-65.
coefficient KN,S in the sheltered outdoor exposure (R2 = 0.69).

In the sheltered outdoor exposure its carbonation coefficient KN,S is 1.5 (Table 5). More important, there is a further change in the rank-
in the same range as the one of the mortars OPC-55 and L-15. The ing of the mortars. The mortars MS now nearly reach the level of
reason for this change could be the lower CO2 content in natural mortar OPC-48. But again, the most marked change is observed
conditions or the higher relative humidity with yearly mean of for mortar S-65. Its carbonation coefficient KN,US is now lower than
76% compared to 57% in the carbonation chambers. In general, car- the values of mortars L, mortar P-15 and mortar OPC-55. It is clear
bonation is slowed down with increasing relative humidity at that the higher relative humidity slows down the rate of carbona-
levels above 60% [6,33–35]. To asses this, the data in the unshel- tion [6,33–35]. But these data clearly indicate that it is additionally
tered conditions have to be considered (Fig. 6). Due to the exposure the parameter responsible for the change in the rankings of the
to snow and rain, the moisture content of the mortars in the mortars. It has a stronger impact on the carbonation of the mortars
unsheltered exposure has to be higher compared to the one in MS and S-65 than on the other mortars. As a result of the increased
the sheltered one. As a result, the carbonation coefficient KN of moisture content, the correlation between the carbonation coeffi-
the unsheltered mortar specimens is lower by a factor of about cients KACC,1% and KN,US is worsening (Fig. 6).

0.3 18
non-carbonated OPC L
Carbonation coefficient K ACC,1%

carbonated 16
0.25 MS P
14 S
0.2 12
[mm/y1/2]
Al/Ca

10
0.15
8
0.1 6
4
0.05
2

0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Si/Ca Carbonation coefficient K N,US [mm/y1/2]

Fig. 4. Molar Al/Ca-ratio as a function of molar Si/Ca-ratio in non-carbonated area Fig. 6. Carbonation coefficient KACC at 1% CO2 versus carbonation coefficient KN,US in
and the carbonated area of mortar MS-15. the unsheltered outdoor exposure (R2 = 0.15).
A. Leemann et al. / Cement & Concrete Composites 62 (2015) 33–43 39

The lower carbonation coefficient KN of the cut compared to the comparatively high decrease of the carbonation coefficient in the
cast concrete surfaces (Table 6) can be likely attributed to optimal sheltered and unsheltered exposure could be the drying behavior
conditions for hydration of the cut surfaces whereas the cast sur- after a period of high humidity caused by fog or rain. The mortars
faces may dry to a certain degree. Moreover, the probability for S-65 and MS and concrete S-65 dry relatively slowly during the
segregation and the accumulation of bleeding water leading to a conditioning for the oxygen diffusion and additionally have very
local increase of w/c is higher on the cast surfaces. Therefore, the low oxygen diffusion coefficients DO2 (Tables 6 and 7). Therefore,
values from the cut surfaces are used in the following comparisons. it can be expected that the moisture level at the depth of the car-
The coefficients of variation calculated for the carbonation coeffi- bonation front remains high for a longer time and as a consequence
cients of the concrete mixtures as described in paragraph 3.3 seem carbonation would become slower than in a faster drying mortar or
to be in a reasonable range for the number of measurements allow- concrete.
ing such comparisons. Moreover, it has to be mentioned that the preconditioning for
Similar to the mortars there is a change in the ranking of the oxygen diffusion and the following diffusivity measurement was
concrete mixtures going from sheltered to unsheltered exposure. started at an age of 28 days. It can be expected that the diffusivity
The concrete mixtures S-65 show a more pronounced decrease of of mortar and concrete S-65 decreases more with time than the
the carbonation coefficient KN,US than the majority of the other con- one of the other mixtures. This likely increases the difference to
crete mixtures (Fig. 7). It is less pronounced than in the mortars but the other mixtures. A slower water loss due to evaporation of con-
still obvious. In the case of concrete L-FA-35 the relatively low fly crete containing GGBS and concrete containing fly ash compared to
ash content of 15 mass-% seems not to be sufficient to cause the pure OPC concrete was observed as well by Parrot [38,39]. It is
same behavior as observed in the other mixtures containing silic- expected that higher amounts of fly ash than those used in this
eous mineral additions. One possible reason for the behavior of study may cause a similar behavior to the one observed for the
mixtures containing GGBS and microsilica is their finer pore struc- mixtures S-65 and MS.
ture (Table 7) [36,37]. Capillary condensation depends on relative These results have implication for the testing of accelerated car-
humidity and pore size: the smaller the pore size, the lower the rel- bonation. Rebar corrosion due to carbonation is a problem in
ative humidity at which condensation occurs. As a result, a mate- unsheltered exposure, where there is a sufficient level of humidity
rial with a relatively high amount of fine pores may have a to support the reaction [40,41]. These are exactly the conditions
higher amount of water filled pores due to condensation than a where the correlation between carbonation coefficients KACC and
material with a coarser pore structure at identical relative humid- KN,US is poor (R2 = 0.15 for the mortars/Fig. 7), mainly due to the
ity. Transferring this to the mortar and concrete mixtures S-65 and behavior of mortars S-65 and MS. It is likely that the performance
MS, a higher amount of water filled pores could slow down carbon- of such mortar and concrete in real structures will be underesti-
ation in these mixtures. The second possible reason for the mated based on results obtained in accelerated carbonation tests
conducted at a relatively low relative humidity.

A 1.4
OPC L 4.2. Relation of carbonation with physical concrete parameters
1.2 MS P
Carbonation coefficient K N,US

S The carbonation coefficient appears to be correlated to the com-


1.0 pressive strength when the mortars containing siliceous additions
are not taken into account (Fig. 8). However, when all mortars are
0.8
[mm/y1/2]

considered, there is no clear relation. The same applies to the con-


0.6 crete mixtures where a cement-specific relation to compressive
strength and w/c is observed (Fig. 9). On the contrary, no correla-
0.4 tion is evident when all mixtures are considered.
Oxygen diffusion shows a comparable relation to carbonation as
0.2
compressive strength. Although there is a cement-specific relation,
0.0 the carbonation coefficient KN and the oxygen diffusion coefficient
0.0 0.5 1.0 1.5 DO2 of mortar and concrete (Fig. 10) as a whole do not correlate.
Carbonation coefficient K [mm/y1/2] Concrete S-65 for example exhibits the lowest oxygen diffusion
N,S
coefficient DO2, but its carbonation coefficients KN,S are the highest.
B 4 On the other side, concrete OPC and OPC-HS have the highest oxy-
OPC gen diffusion coefficients DO2 for a given carbonation coefficient.
Carbonation coefficient K N,US-CUT

OPC HS Unfortunately, no published data are available on the relation-


3 L-15 ship between concrete mix design and CO2-diffusion. It is only
L-FA-35 clear that CO2 diffusion in cement-based material is slower than
oxygen diffusion [42]. But it is not clear, whether O2 and CO2 diffu-
[mm/y1/2]

S-65
2 sion coefficients show the same relative changes as a function of
concrete mix design.
In the carbonated layer, a coarsening of the pore structure and a
1 transformation from portlandite to calcite accompanied by decalci-
fication of C–S–H is observed in the studied mortar samples. This
process goes together with a decrease of total porosity in all mor-
0
0 2 4 6 tars except for mortar S-65, where an increase is observed. The
coarsening of the pore structure seems to dominate the effect on
Carbonation coefficient K N,S-CUT [mm/y1/2]
diffusivity: the oxygen diffusion coefficient DO2 is increased by car-
Fig. 7. Carbonation coefficient KN of the mortar (A) and the concrete mixtures (B) in
bonation in cement pastes produced with pure OPC, with OPC
the sheltered and unsheltered outdoor exposure (A: R2 = 0.65, B: 0.83). The line combined with fly ash and OPC combined with GGBS [43]. Such
indicates a ratio between x and y of 1:1. an acceleration in transport due to carbonation is observed as well
40 A. Leemann et al. / Cement & Concrete Composites 62 (2015) 33–43

18 The CaOreactive of the binders used for the mortar and concrete
OPC L
Carbonation coefficient KACC,1%

16 mixtures is given in Tables 2 and 3. In the case of CEM I 52.5 used


MS P for the mortars, 1.7 mass-% CaO were subtracted, taking into
14 S account a limestone powder content of 3 mass-% (based on infor-
12 mation from cement producer). Concerning the cements used for
[mm/y1/2]

10 the concrete mixtures, 0.6 mass-% of CaO was subtracted for a


limestone powder content of 1.0 mass-% in the case of CEM I
8
32.5 and 2.8 mass-% of CaO for a limestone content of 5.0 mass
6 in the case of CEM I 42.5 HS (information from the cement pro-
4 ducer). In the concrete mixtures produced with CEM II/A-LL and
CEM II/B-M (V-LL) corrections were made for the 15 mass-% of
2
limestone powder they contain.
0 Now, the w/CaOreactive can be compared with the determined
50 60 70 80
carbonation coefficients. There is a clear correlation between the
Compressive strength 91 Tage [MPa] w/CaOreactive and the carbonation coefficients KACC determined at
Fig. 8. Carbonation coefficient KACC at 1% CO2 as a function of compressive strength
1% and 4% CO2 (R2 = 0.89 and 0.91). Both, the carbonation coeffi-
of the mortar at 91 days. The ellipsoid marks the mortars with siliceous mineral cient KN,S of the mortar and the concrete mixtures exposed to nat-
additions. ural carbonation in the sheltered exposure correlate with the
w/CaOreactive (R2 = 0.83 and 0.87/Fig. 11). In the unsheltered expo-
sure with an increased moisture content in the material the corre-
in the water permeability of pastes produced with blended lation is getting worse for the mortars (R2 = 0.29) and stays about
cements [44]. the same for the concrete mixtures (R2 = 0.84).
Service life modelling in regard to carbonation has to take into So far, the degree of hydration has not been taken into account.
account various parameters [45–48], including concrete quality. It The w/CaOreactive implies that clinker and mineral additions have
is obvious from the results of this study, that compressive strength hydrated completely, with no anhydrous material left. Of course,
and w/c alone are inadequate to characterize the behavior of con- this is not the case for the studied mortar and concrete mixtures
crete containing mineral additions. even if they have been cured for 28 days and the hydration has
likely proceeded further afterwards. As shown by the effect of cur-
4.3. Relation of carbonation with the CO2 buffer capacity ing [6,15,21,22], the degree of hydration plays a major role in the
carbonation resistance. However, it is difficult to assess it for mor-
The carbonation of mortar and concrete can be regarded as an tar and concrete exposed to carbonation. First, there are data in lit-
attack by CO2. The intruding CO2 reacts with the CaO of the erature for the degree of hydration of cement pastes but only little
hydrates to form calcite and silica-rich gels [49]. The higher the of mortar and especially of concrete [50–54]. Secondly, the degree
reactive CaO content (CaOreactive) per volume of paste is, the more of hydration is likely changing during exposure [20], since suffi-
CO2 per volume of paste can be captured and the slower the car- cient moisture is provided especially in the unsheltered exposure
bonation front should progress into the material. The CaOreactive conditions. Assessing the degree of hydration is especially difficult,
per volume of cement paste can be calculated from the composi- when two or more components like cement clinker, GGBS and fly
tion of the binder and the w/c. The CaO content of cement and min- ash with specific hydration kinetics are present. Thirdly, when con-
eral additions is measured by X-ray fluorescence, a standard crete is cured for a short time and then exposed to drying, a pro-
procedure in the characterization of these materials. If cement con- nounced gradient of the degree of hydration can be expected
tains limestone powder, the amount of CaO present in the lime- with increasing depth. This makes it impossible to accurately
stone powder has to be subtracted as it is not reactive. assess the w/CaOreacted at the carbonation front.
Consequently, the CO2 buffer capacity per volume of paste and In spite of these uncertainties, the degree hydration is esti-
with it the carbonation resistance of mortar and concrete can be mated here based on data from literature and applying some sim-
expressed by the ratio between water and CaOreactive plifications. This at least allows judging its principle effects. Data
(w/CaOreactive). Of course, the CO2 can only react with CaO that is from Parrot and Killoh [50] are used for the cement clinker. For
present in hydrates and not the one present in anhydrous cement the mortar used in the accelerated carbonation tests, a degree of
or mineral additions. This effect of the degree of hydration will be hydration of 75% (w/c = 0.40), 82% (w/b = 0.48) and 86%
discussed later in this section.
7
7
Carbonation coefficient K N,S-CUT
Carbonation coefficient K N,S-CUT

6
6
5
5
[mm/y1/2]

4
[mm/y1/2]

4
3
3 OPC
OPC
OPC HS 2 OPC HS
2
L-15 L-15
1 L-FA-35 1 L-FA-35
S-65 S-65
0 0
0.40 0.45 0.50 0.55 0.60 0.65 0 1 2 3 4
w/c Oxygen diffusion coefficient D O [10-8 m2/s]

Fig. 9. Carbonation coefficient KN of the cut concrete surfaces in the sheltered Fig. 10. Carbonation coefficient KN of the cut concrete surfaces in the sheltered
exposure as a function of w/c. exposure as a function of the oxygen diffusion coefficient DO2.
A. Leemann et al. / Cement & Concrete Composites 62 (2015) 33–43 41

A 1.6 (R2 = 0.87 and 0.85), it worsens in the sheltered (Fig. 12A) and more
OPC L so in the unsheltered exposure (R2 = 0.46 and 0.06). The concrete
1.4
mixtures show a better correlation in the sheltered exposure
Carbonation coefficient KN,S

MS P
1.2 S (Fig. 12B) than in the unsheltered one (R2 = 0.79 and 0.61).
Compared to the w/CaOreactive, the correlation of the w/CaOreacted
1.0
with the carbonation coefficients K is the same in the accelerated
[mm/y1/2]

0.8 testing, but it is worse in the sheltered and unsheltered outdoor


exposure. Assuming complete hydration, the w/CaOreactive in the
0.6
mixtures containing GGBS deviates the strongest from reality.
0.4 However, the good correlation indicates that this inaccuracy is
0.2
partly compensated by the effect of the increased moisture level
in the outdoor exposure. The influence of the CO2 buffer capacity
0.0 on carbonation is possibly lessened with increasing moisture con-
0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2
tent when mixtures with GGBS and microsilica are used.
w/CaOreactive [-] Nevertheless, the CO2 buffer capacity appears to be the govern-
ing parameter for the carbonation resistance of mortar and con-
B 8
crete considering the entire data set. Because of the uncertainties
OPC
related to the assessment of the degree of hydration, it seems to
Carbonation coefficient KN,S-CUT

OPC HS
L-15
be preferable to express the buffer capacity by the w/CaOreactive.
6
L-FA-35 This parameter can be a useful tool to assess the carbonation resis-
S-65 tance of different mix designs with variations in cement type, paste
[mm/y1/2]

volume and w/c. Its reliability can be worsened at a high moisture


4
level of mortar or concrete stored in unsheltered exposure, because
capillary condensation and drying behavior as physical parameters
are of increased importance under these conditions like discussed
2
in Section 4.1. Additionally, it has to be pointed out that the mortar
and concrete samples in this study have been moist cured for
28 days. The correlations shown above may become less accurate
0
0.6 0.8 1.0 1.2 1.4 in the case of shorter curing.
w/CaOreactive [-] The generally good correlation of both the w/CaOreactive and the
w/CaOreacted with the carbonation coefficients KACC and KN,S gives
Fig. 11. Carbonation coefficient KN,S in the sheltered outdoor exposure as a function
of the w/CaOreactive of the mortar (A: R2 = 0.79) and of the concrete mixtures (B: A 1.6
R2 = 0.87).
1.4
Carbonation coefficient KN,S

(w/c = 0.55) is assumed. For the mortars with mineral additions 1.2
(w/b = 0.48) the same degree of clinker hydration is assumed as
1.0
for the pure OPC sample. This is a simplification as the degree of
[mm/y1/2]

clinker hydration can increase with the addition of 0.8


slower-reacting mineral additions [50]. Because the CaO content
0.6
of the microsilica is very low, it was not considered. The degree
of hydration of GGBS is difficult to assess, as it depends on the ratio 0.4 OPC L
of cement clinker to GGBS, on the specific batch of GGBS and on MS P
0.2
age [51–54]. Here, it was assumed that the degree of reaction of S
GGBS is lower than the one of the clinker by a factor of 2.0 [51– 0.0
54]. Although the degree of hydration in the mortars exposed out- 0.9 1.1 1.3 1.5 1.7
doors can be expected to be higher than in the accelerated test, the w/CaOreacted [-]
same values were assumed to simplify.
B 8
In the case of concrete, average degrees of hydration for the
OPC
cement clinker during the natural exposure of 2.5 years of 84%
Carbonation coefficient KN,S-CUT

OPC HS
(w/b = 0.45), 88% (w/b = 0.50) and 93% (w/c = 0.60) were assumed,
6 L-15
independently of the amount of mineral additions. The average rel-
L-FA-35
ative humidity at the exposure site is about 76% (Fig. 1). The higher
S-65
moisture content in the unsheltered exposure likely leads to a
[mm/y1/2]

higher degree of hydration in the samples stored in these condi- 4


tions. However, due to the lack of any data the same degree of
hydration is assumed. For the fly ash a degree of hydration lower
by a factor of 2.0 than the one of the clinker was assumed 2
[51,55]. The degree of hydration of GGBS (approximate content
of 66 mass-% in the CEM III/B) was assumed to be lower than the
one of the clinker by a factor of 1.4. The higher degree of hydration 0
0.6 0.8 1.0 1.2 1.4 1.6
and smaller difference between cement clinker and GGBS com-
w/CaOreacted [-]
pared to the mortars was chosen due to longer duration of the nat-
ural exposure (2.5 years versus 1 year). Fig. 12. Carbonation coefficient KN,S in the sheltered outdoor exposure as a function
While the correlation of the w/CaOreacted with the carbonation of the w/CaOreacted of the mortar (A: R2 = 0.46) and the concrete mixtures (B:
coefficients KACC of the mortars is good at 1% and 4% CO2 R2 = 0.79).
42 A. Leemann et al. / Cement & Concrete Composites 62 (2015) 33–43

some indications about the behavior of C–S–H and portlandite in  The parameter governing the carbonation coefficient K of
regard to carbonation. Due to the use of mineral additions, their the mortar and concrete mixtures is the CO2 buffer capacity
relative amount is different in both mortar and concrete. For exam- per volume of cement paste. It can be expressed by the
ple, mortar P has a higher ratio of portlandite to C–S–H due to the w/CaOreactive or w/CaOreacted. However, both parameters
portlandite addition compared to the mortars produced with OPC. have restrictions. The first parameter assumes complete
Mortars MS and S have a lower one due to the C–S–H formed by hydration of cement and mineral additions, which is cer-
these siliceous mineral additions. In the first one, portlandite is tainly not the case for the studied mixtures. The later
consumed due to the hydration of the microsilica, resulting in parameter is limited by the uncertainty related to the
additional C–S–H formation [e.g. 56,57]]. In the second one, there assessment of the degree of hydration. The w/CaOreactive
is little portlandite due to the low clinker content and additional and w/CaOreacted are still valuable to assess the carbonation
C–S–H formation by the hydraulic GGBS reaction. However, there resistance of mortar and concrete with various mix designs.
is no trend in the carbonation coefficients indicating a slower or  The correlation of the CO2 buffer capacity with the carbon-
faster reaction with CO2 based on the different C–S–H to port- ation coefficient K varies with the relative humidity and is
landite ratios. Consequently, the contribution of C–S–H and port- becoming worse in the unsheltered outdoor exposure likely
landite to the carbonation resistance seems to be only dependent due to the increasing importance of capillary condensation
on their respective CaO-contents. It has been postulated that port- and drying behavior.
landite carbonates more slowly than C–S–H [58,59]. This seems to
In this study the buffer capacity per volume cement paste
be a physical effect: a carbonated shell can form around large port-
expressed by the w/CaOreactive or w/CaOreacted has been identified
landite agglomerates, slowing down carbonation [49,60–62].
as a novel parameter to assess the carbonation resistance of mortar
Therefore, portlandite can still be present in the zone identified
and concrete containing mineral additions. Based on these findings
as carbonated by phenolphthalein [58]. However, this seems to
further research is needed to clarify the influence of curing and
have negligible effect on the carbonation depth of the studied sam-
exposure conditions on the carbonation resistance of such systems
ples as determined with the phenolphthalein test.
in more detail. As an example, the purpose for the water storage of
all specimens during the first 28 days in this study was to mini-
mize the effect of curing on the different cements and mineral
5. Conclusions
additions used. The relation between this type of curing and the
typical curing as applied on construction sites and specified in
Mortars produced with pure OPC and additions of limestone
standards on the carbonation resistance of mortar and concrete
powder, microsilica, portlandite and GGBS were exposed to accel-
produced with various mineral additions has to be investigated.
erated carbonation with 1% and 4% CO2 and to sheltered and
Furthermore, the mechanisms leading to the decreased carbona-
unsheltered outdoor conditions. Additionally, concrete mixtures
tion coefficient KN of concrete produced with GGBS, microsilica
produced with different cement types, w/c and paste volume
and likely other siliceous mineral additions especially in the
were stored outdoors for 2.5 years in sheltered and unsheltered
unsheltered outdoor exposure should be further investigated.
outdoor exposure. All samples were moist cured for 28 days
This is of particular interest, because concrete with these types of
before being exposed to carbonation. Apart of the carbonation
mineral additions might perform significantly better in conditions
coefficient, compressive strength, oxygen diffusion and pore char-
permitting rebar corrosion than indicated by accelerated carbona-
acteristics were determined. The results allow drawing the fol-
tion tests.
lowing conclusion for the range of mix designs and the curing
used in this study:
Acknowledgements
 There is an excellent correlation between the carbonation
coefficients KACC,4% and KACC,1% of the mortars. Their ranking cemsuisse is acknowledged for the financial support of this pro-
in regard to carbonation resistance is essentially the same. ject and P. Lura for the careful review of the manuscript.
 Going from the accelerated carbonation to the carbonation
in the sheltered outdoor exposure, mortars MS and espe- References
cially mortar S-65 show a larger decrease of the carbona-
tion coefficient KN,S than the other mortars. [1] Mangat PS, Khatib JM. Influence of fly ash, silica fume, and slag on sulfate
 Going from the sheltered to the unsheltered outdoor expo- resistance of concrete. ACI Mater J 1993;92:542–52.
[2] Thomas MDA, Shehata MH, Shashiprakash SG, Hopkins DS, Cail K. Use of
sure, the decrease of the carbonation coefficient KN,US is ternary cementitious systems containing silica fume and fly ash in concrete.
again more pronounced for mortars MS and especially mor- Cem Concr Res 1999;29:1207–14.
tar S-65. The same applies to the concrete mixtures S-65. [3] Bleszynski R, Hooton RD, Thomas MDA, Rogers CA. Durability of ternary blend
concrete with silica fume and blast-furnace slag: laboratory and outdoor
There are two possible reasons for the comparatively large exposure site studies. ACI Mater J 2002;99(5):499–508.
decrease of the carbonation coefficient KN compared to the [4] Rozière E, Loukili A, El Hachem R, Grondin F. Durability of concrete exposed to
other mixtures. First, the higher amount of small pores in leaching and external sulphate attacks. Cem Concr Res 2009;39:1188–98.
[5] Leemann A, Lothenbach B, Hoffmann C. Biologically induced concrete
the mixtures MS and S-65 could lead to a higher amount deterioration in a wastewater treatment plant assessed by combining
of water filled pores due capillary condensation at a given microstructural analysis with thermodynamic modeling. Cem Concr Res
relative humidity, slowing down carbonation. Secondly, a 2010;40:1157–64.
[6] Wierig HJ. Longtime studies on the carbonation of concrete under normal
high moisture level at the carbonation front remains longer
outdoor exposure. In: Proceedings RILEM seminar on the durability of concrete
after a high humidity event, due to the low diffusivity and structures under normal outdoor exposure; 1984. p. 239–49.
slow drying of these mixtures. This would decrease the car- [7] Thomas MDA, Matthews JD. Performance of fly ash concrete in UK structures.
ACI Mat J 1993;90:586–93.
bonation rate as well.
[8] Papadakis VG. Effect of supplementary cementing materials on concrete
 There is a cement-specific correlation of the carbonation resistance against carbonation and chloride ingress. Cem Concr Res
coefficient K with w/c, compressive strength and oxygen 2000;30:291–9.
diffusion DO2 of the mortar and concrete mixtures. [9] Sisomphon K, Franke L. Carbonation rates of concretes containing high volume
of pozzolanic materials. Cem Concr Res 2007;37:1647–53.
However, if all mixtures are taken into account, no correla- [10] Rozière E, Loukili A, Cussigh F. A performance based approach for durability of
tion is evident. concrete exposed to carbonation. Constr Build Mater 2009;23:190–9.
A. Leemann et al. / Cement & Concrete Composites 62 (2015) 33–43 43

[11] Gruyaert E, Van den Heede P, De Belie N. Carbonation of slag concrete: effect of [37] Li S, Roy DM. Investigation of relations between porosity, pore structure, and
the cement replacement level and curing on the carbonation coefficient – C1 diffusion of fly ash and blended cement pastes. Cem Concr Res
effect of carbonation on the pore structure. Cem Concr Compos 1986;16:749–59.
2013;35:39–48. [38] Parrot LJ. Moisture profiles in drying concrete. Adv Cem Res 1988;1:164–70.
[12] Thomas MDA, Matthews JD. Carbonation of fly ash concrete. Mag Concr Res [39] Parrot LJ. Factors influencing relative humidity in concrete. Mag Concr Res
1992;44:217–28. 1991;43:45–52.
[13] De Ceukelaire L, Van Nieuwenburg D. Accelerated carbonation of a blast- [40] Tuutti K. Corrosion of steel in concrete. Stockholm: Svenska
furnace cement concrete. Cem Concr Res 1993;23(2):442–52. Forskningsinstitutet för cement och betong; 1982.
[14] Chen CT, Ho CW. Influence of cyclic humidity on carbonation of concrete. J [41] González JA, Andrade C. Effect of carbonation, chlorides and relative ambient
Mater Civ Eng 2013;25:1929–35. humidity on the corrosion of galvanized rebars embedded in concrete. Br
[15] Thomas MDA, Matthews JD, Haynes CA. Carbonation of fly ash concrete, vol. Corros J 1982;17:21–8.
19. ACI Spec Publ; 1992. p. 539–56. [42] Houst YF, Wittmann FH. Influence of porosity and water content on the
[16] Sanjuán MA, Andrade C, Cheyrezy M. Concrete carbonation tests in natural and diffusivity of CO2 and O2 through hydrated cement paste. Cem Concr Res
accelerated conditions. Adv Cem Res 2003;15:171–80. 1994;24:1165–76.
[17] Hyvert N, Sellier A, Duprat F, Rougeau P, Francisco P. Dependency of C–S–H [43] Ngala VT, Page CL. Effects of carbonation on the pore structure and diffusional
carbonation rate on CO2 pressure to explain transition from accelerated tests properties of hydrated cement paste. Cem Concr Res 1997;27:995–1007.
to natural carbonation. Cem Concr Res 2010;40:1582–9. [44] Auroy M, Poyet S, Le Bescop P, Torrenti JM. Impact of carbonation on water
[18] Bernal SA, Provis JL, Brice DG, Kilcullen A, Duxson P, van Deventer JSJ. transport properties of cement-based materials. In: NUWCEM, 2nd
Accelerated carbonation testing of alkali-activated binders significantly international symposium on cement-based materials for nuclear waste,
underestimates service life: the role of pore solution chemistry. Cem Concr Avignon, France; 2014.
Res 2012;42:1317–26. [45] Papadakis VG, Vayenas CG, Fardis MN. Fundamental modeling and
[19] Neves R, Branco F, de Brito J. Field assessment of the relationship between experimental investigation of concrete carbonation. ACI Mater J
natural and accelerated concrete carbonation resistance. Cem Concr Compos 1991;88:363–73.
2013;41:9–15. [46] Steffens A, Dinkler D, Ahrens H. Modeling carbonation for corrosion risk
[20] Younsi A, Turcry Ph, Aït-Mokhtar A, Staquet S. Accelerated carbonation of prediction of concrete structures. Cem Concr Res 2002;32:935–41.
concrete with high content of mineral additions: effect of interactions [47] Saetta AV, Vitaliani RV. Experimental investigation and numerical modeling of
between hydration and drying. Cem Concr Res 2013;43:25–33. carbonation process in reinforced concrete structures: part I: theoretical
[21] Osborne GJ. Carbonation and permeability of blastfurnace slag cement formulation. Cem Concr Res 2004;34:571–9.
concretes from field structures, vol. 114. ACI Spec Pub; 1992. p. 1209–38. [48] Marques PF, Chastr C, Nunes Â. Carbonation service life modelling of RC
[22] Fatthudi NI. Concrete carbonation as influenced by curing regime. Cem Concr structures for concrete with Portland and blended cements. Cem Concr
Res 1988;18:426–30. Compos 2013;37:171–84.
[23] Geiker MR, Laugesen P. On the effect of laboratory conditioning and freeze/ [49] Groves GW, Brough A, Richardson IG, Dobson CM. Progressive changes in the
thaw exposure on moisture profiles in HPC. Cem Concr Res 2001;31:1831–6. structure of hardened C3S cement pastes due to carbonation. J Am Ceram Soc
[24] Hooton RD, Geiker MR, Bentz EC. Effects of curing on chloride ingress and 1991;74:2891–6.
implications on service life. ACI Mater J 2002;99:201–6. [50] Parrot LJ, Killoh DC. Prediction of cement hydration. Proc Br Ceram Proc
[25] Loser R, Leemann A. Sulfatwiderstand von Beton: verbessertes Verfahren 1984;35:41–53.
basierend auf der Prüfung nach SIA 262/1, Anhang D, UVEK, Bundesamt für [51] Feng X, Garboczi EJ, Bentz DP, Stutzman PE, Mason TO. Estimation of the
Strassen, Bericht Nr. 1416, Bern; 2013. degree of hydration of blended cement pastes by a scanning electron
[26] DIN EN 1015-11. Methods of test for mortar for masonry Determination of microscope point-counting procedure. Cem Concr Res 2004;34:1787–93.
flexural and compressive strength of hardened mortar. 2007. [52] Lumley JS, Gollop RS, Moir GK, Taylor HFW. Degrees of reaction of the slag in
[27] EN 12390-3. Testing hardened concrete – Part 3: compressive strength of test some blends with Portland cements. Cem Concr Res 1996;26:139–51.
specimens; 2002. [53] Pane I, Hansen W. Investigation of blended cement hydration by isothermal
[28] SN 505 262/1, Betonbau – Ergänzende Festlegungen, Anhang I, calorimetry and thermal analysis. Cem Concr Res 2005;35:1155–64.
Karbonatisierungswiderstand, SIA Zürich; 2013. [54] Kocaba V, Gallucci E, Scrivener KL. Methods for determination of degree of
[29] Lawrence CD. Transport of oxygen through concrete. In: Glasser FP, editor. The reaction of slag in blended cement pastes. Cem Concr Res 2012;42:511–25.
chemistry and chemically-related properties of cement. British Ceramic [55] Ben Haha M, De Weerdt K, Lothenbach B. Quantification of the degree of
Society Proceedings, vol. 35; 1984. p. 277–93. reaction of fly ash. Cem Concr Res 2010;40:1620–9.
[30] Buenfeld NR, Okundi E. Effect of cement content on transport in concrete. Mag [56] Mitchell DRG, Hinczak I, Day RA. Interaction of silica fume with calcium
Concr Res 1998;50:339–51. hydroxide solutions and hydrated cement pastes. Cem Concr Res
[31] Villani C, Loser R, West MJ, Di Bella C, Lura P, Weiss JW. An inter lab 1998;28:1571–84.
comparison of gas transport testing procedures: oxygen permeability and [57] Lothenbach B, Scrivener K, Hooton RD. Supplementary cementitious materials.
oxygen diffusivity. Cem Concr Compos 2014;53:357–66. Cem Concr Res 2011;41:1244–56.
[32] Kaufmann J, Loser R, Leemann A. Analysis of cement-bonded materials by [58] Parrott LJ, Killoh DC. Carbonation in a 36 year old, in-situ concrete. Cem Concr
multi-cycle mercury intrusion and nitrogen sorption. J Colloid Interface Sci Res 1989;19:649–56.
2009;336:730–7. [59] Castellote M, Andrade C, Turrillas X, Campo J, Cuello GJ. Accelerated
[33] De Ceukelaire L, Van Nieuwenburg D. Accelerated carbonation of a blast- carbonation of cement pastes in situ monitored by neutron diffraction. Cem
furnace cement concrete. Cem Concr Res 1993;23:442–52. Concr Res 2008;38:1365–73.
[34] Verbeck G. Carbonation of hydrated Portland cement. In: Crawford PH, [60] Van Balen K. Carbonation reaction of lime, kinetics at ambient temperature.
editor. Cement and Concrete. Baltimore: American Society for testing Cem Concr Res 2005;35:647–57.
Material; 1958. [61] Thiery M, Villain G, Dangla P, Platret G. Investigation of the carbonation front
[35] Russell D, Basheer PAM, Rankin GIB, Long AE. Effect of relative humidity and shape on cementitious materials: effects of the chemical kinetics. Cem Concr
air permeability on prediction of the rate of carbonation of concrete. Proc Inst Res 2007;37:1047–58.
Civ Eng Struct Build 2001;146:319–26. [62] Morandeau A, Thiéry M, Dangla P. Investigation of the carbonation mechanism
[36] Mehta PK, Gjørv OE. Properties of Portland cement concrete containing fly ash of CH and C–S–H in terms of kinetics, microstructure changes and moisture
and condensed silica-fume. Cem Concr Res 1982;12:587–95. properties. Cem Concr Res 2014;56:153–70.

You might also like