You are on page 1of 8

Construction and Building Materials 112 (2016) 925–932

Contents lists available at ScienceDirect

Construction and Building Materials


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

Distance-associated chloride binding capacity of cement paste subjected


to natural diffusion
Zijian Song ⇑, Linhua Jiang, Ziming Zhang, Chuansheng Xiong
College of Mechanics and Materials, Hohai University, Nanjing 210098, PR China

h i g h l i g h t s

 Specimens undergoing natural diffusion are used for measuring the chloride binding capacity.
 Difference between conventional CBIs and natural-diffusion-determined CBIs is found.
 CBPs are proposed to quantify the hydroxyl-dependent chloride binding capacity.
 A hydroxyl-dependent chloride binding model is established.

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

Article history: Chloride binding capacity is usually quantified by chloride binding isotherms (CBIs). However, an impor-
Received 4 January 2016 tant factor influencing chloride binding capacity, namely the internal chemical environment (ICE), cannot
Received in revised form 21 February 2016 be represented in the conventional CBIs. Knowing the drawbacks of the conventional CBIs, this paper
Accepted 5 March 2016
straightly put the emphasis on the distance-associated chloride binding capacity. Specimens undergoing
Available online 9 March 2016
natural diffusion were used for measuring the chloride binding capacity at different distance. CBIs,
chloride binding profiles (CBPs), chemical profiles and microscopy result were performed to analyze
Keywords:
the rules and mechanisms. From the results of CBIs and CBPs, it is found that the conventional
Cement paste
Chloride binding
laboratory-determined CBIs would overestimate the chloride binding capacity beyond a certain distance
Distance-associated from the surface, and underestimate the chloride binding capacity near the surface. Based on the
Diffusion chemical profiles and the microscopy results, it is found that ICE changes greatly along with natural dif-
Model fusion, and meanwhile the chloride binding capacity at different distance develops discriminatively with
the ICE change. Plus, a hydroxyl-dependent chloride binding model is established to quantify the
distance-associated chloride binding capacity. The model explains the difference between the conven-
tional laboratory-determined CBIs and the natural-diffusion-determined CBIs, and it fits well with the
test results.
Ó 2016 Published by Elsevier Ltd.

1. Introduction chlorides generally exist in two main forms in concrete, including


free chlorides and bound chlorides. It is a widespread view that
Chloride-induced corrosion is one of the major causes of only free chlorides are immediately detrimental with regard to
degradation of reinforced concrete structures [1–4]. It is known the reinforcement corrosion risk though some bound chlorides
that a sufficient amount of chloride ions in concrete are able to may be released at later ages [7–11]. As a result of binding, the
break the passive layers (c-Fe2O3) of reinforcements, thereby sub- time needed for free chlorides’ reaching a threshold level has a
stantially degrading the durability and service life of reinforced considerable growth [12,13]. Knowing this, chloride binding capac-
concrete structures [5]. According to the two-stage durability ity is commonly regarded as one of the most important factors
model [6], the induction stage, also known as the longest and the deciding the length of service life of reinforced concrete structures.
critical stage, is defined as a duration from the time chlorides Chloride binding includes physical binding and chemical bind-
firstly penetrated into concrete to the time a threshold level of ing. Neither of the two binding processes is a result of single reac-
chlorides reached the steel bar. At this stage, the penetrated tion, but both contain various physical or chemical mechanisms
[7,14]. Chemical binding in an early aged cement hydrate system
⇑ Corresponding author. mainly refers to the reaction between chlorides and unhydrated
E-mail address: hhulhjiang@gmail.cn (Z. Song). C3A/C4AF to form a Friedel’s salt or its analogue [15,16]. In a mature

http://dx.doi.org/10.1016/j.conbuildmat.2016.03.020
0950-0618/Ó 2016 Published by Elsevier Ltd.
926 Z. Song et al. / Construction and Building Materials 112 (2016) 925–932

cement hydrate system, however, chemical binding is possibly due process, and thereby promoting better evaluation of chloride bind-
to the anion exchange between chloride ions and AFm (Al2O3- ing capacity in practical engineering.
Fe2O3-mono) phases [17–19]. Actually, Friedel’s salt or its analogue
is also a specific kind of AFm phases. It is known that the AFm
phases have a layered structure, which contains a positively 2. Materials and method
charged octahedral principal layer [Ca2(Al,Fe)(OH)6]+ and a nega-
2.1. Materials
tively charged interlayer [x.nH2O] [19,20]. In cement chemistry,
the common species of the interlayer x anions are usually Cl (Frie- Ordinary Portland cement (OPC) was used and the detailed chemical composi-
del’s salt), 0.5CO2
3 (Monocarboaluminate), OH (Hydroxy-AFm), tion of the cement is presented in Table 1. Neat cement pastes were prepared at
0.5SO24 (Monosulfoaluminate), 0.5Cl + 0.25SO2
4 (Kuzel’s salt) water-to-cement (w/c) ratios of 0.30, 0.40, 0.50 and 0.60.

etc. [21–24]. An AFm crystal may also contain more than one spe-
cies of X anions to form solid solution [21]. The point is that the 2.2. Specimens preparation and immersion test
hydrogen bonds between principal layer and interlayer anions
are weak and therefore the intruded chlorides can replace other 6 cylindrical specimens (50 mm in diameter and 100 mm in height) and 30 disc
species of interlayer anions of AFm phases to form Friedel’s salt specimens (50 mm in diameter and 3 mm in height) were casted for each w/c ratio.
After curing in saturated limewater at 20 ± 1 °C for 28 days, the specimens were
or analogue in the mature cement hydrated system [18]. Physical
taken out for an immersion test.
binding is neither unitary. Ramachandran pointed out three kinds The cylindrical specimens were used for measuring the chloride binding capac-
of binding with C-S-H: in a chemisorbed layer on hydrated calcium ity at different distance after the immersion test. Prior to the immersion test, the
silicates, present in the C-S-H interlayer spaces, and intimately cylindrical specimens underwent a surface treatment to implement the one-
bound in the C-S-H lattice [25]. dimensional diffusion. In this work, the side and bottom surfaces of each specimen
were sealed with epoxy resin, and only the top surface was kept unsealed. After
It is because of the various binding mechanisms that chloride that, the cylindrical specimens were exposed to a 0.5 mol/L NaCl solution for a per-
binding capacity is affected by many factors, e.g., content of alumi- iod of 180 d. The solution tanks are covered with polythene sheets to avoid evapo-
nates, C/S ratio of C-S-H gel, sulfate ion concentration, hydroxyl ion rations and carbonizations. In addition, the solutions were renewed every month
concentration and cation type of pore solution [26–30]. In other for maintaining the chloride concentrations.
The disc specimens were used to determine the CBIs in a conventional way (tak-
words, chloride binding capacity is strongly impacted by internal
ing no account of the distance). They were averagely divided into five groups and
chemical environment (abbreviated ‘‘ICE” below), which is known immersed in 0.1, 0.2, 0.3, 0.4 and 0.5 mol/L NaCl solutions, respectively. Considering
as the composition of pore solution and solid phases surrounding that the thin discs are possibly easy to reach saturation, a much shorter immersion
the location where chloride binding takes place. Likewise, some period, about one month, was set for these specimens. Polythene sheet covering
external interactions being able to influence the ICE could also was also used here to avoid evaporations and carbonizations.

impact the chloride binding, e.g. sulfate attack, leaching, carbona-


tion [7,31,32]. Even in a normal chloride diffusion process, ICE 2.3. Measurement of free chloride and other chemical concentrations
would also gradually change, from the surface to the deep, along
with chloride transport so as to keep electrical neutrality. After immersion, the 6 cylindrical specimens were cut layer-by-layer from the
Chloride binding is often quantified at thermodynamic equilib- exposed surface. The nominal thickness of each layer was 5 mm. The actual thick-
ness of each layer, excluding the cut abrasion, is controlled within 3–4 mm. Before
rium by chloride binding isotherms (CBIs), which are known as
pressing, the specimen layers were converted to a saturated, surface-dry condition.
the relationships between free and bound chloride ions over a range Then each specimen layer was crushed into four or five small pieces. One or two
of chloride concentrations at a given temperature [8]. CBIs are usu- pieces of each layer were randomly selected and temporarily stored for the follow-
ally provided with clear mathematical formulae and therefore they ing total chloride concentration measurement, and the rest pieces were used here
for pressing the pore solution.
can be easily applied into the chloride transport models. Consider-
The small pieces from the same-distance layers of the six cylinders were
ing this, CBIs have attracted considerable attention from the stacked together and put into a presser. At a pressure up to 300 MPa, the pore solu-
researchers. Up to present, four types of binding isotherms (i.e., lin- tion was then slowly extracted out. When sufficient pore solution was obtained, it
ear, Langmuir, Freundlich and BET binding isotherm) have been was divided into two parts and separately stored in two test-tubes. For one tube of
proposed to describe the chloride binding [33]. In order to deter- solution, the free chloride concentration was measured using an automatic titrime-
ter. For the other, the chemical concentration of the pore solution was measured by
mine suitable types and parameter values of CBIs, two main labora-
flame photometry. Prior to titration, the PH value was also tested by a glass PH elec-
tory methods are frequently used in the existed studies [15,27,34]. trode in the first tube of solution.
According to the two methods, researchers need not to take sam- The disposal of the 30 disc specimens was much easier because they need not
ples from natural diffusion specimens but directly use powder or be re-cut before pressing. After retaining samples for the total chloride concentra-
tion measurement, the crushed discs in the same group were straightly stacked in
thin disc specimens, which are soaked in different concentrations
the presser to extract the pore solution. Later, the free chloride concentration, the
of chloride solutions, instead. In that way, however, the chemical PH value and the other chemical concentrations were measured using the above
environment of chloride binding will not match the actual ICE of methods.
the concrete experiencing a natural diffusion. Likewise, the impact
of the gradual ICE variation on chloride binding capacity cannot be
precisely presented at different transport distance either. Therefore, 2.4. Measurement of total chloride concentrations

it is perhaps not very proper to straightly use the conventional


The restored specimen pieces were ground into powder using a mortar and pes-
laboratory-determined CBIs to predict the chloride binding capacity tle. Then an amount of powdered specimen was soaked in dilute nitric acid solution
at different distance of the concrete undergoing natural diffusion. to determine the total chloride concentration. Each test result was given as an aver-
Knowing the drawbacks of the conventional laboratory- age of three replicates.
determined CBIs, this paper straightly put the emphasis on the
distance-associated chloride binding in natural diffusion. A com-
2.5. Microscopy test
parison between the conventional laboratory-determined CBIs
and the natural-diffusion-determined CBIs is presented. A CBPs A small flake sample was typically taken from the cylinder specimen with 0.5
(Chloride binding profiles) method is also proposed to clearly ana- w/c ratio. The flake was selected nearly parallel to the diffusion direction and per-
lyze the chloride binding capacity at different distance of cement pendicular to the exposed surface. With the HITACHI S-3400N Scanning Electron
Microscope (SEM), the micro morphology of the cement hydration was analyzed
pastes undergoing chloride diffusion. The main purpose of this at different distances from the exposed surface. Besides, a linear Energy Dispersive
paper is to contribute to improving the understanding of X-ray Spectroscopy (EDS) scanning was also carried out in distance of 0–3 mm to
distance-associated chloride binding capacity in a natural diffusion analyze the chemical element profile in the surface region.
Z. Song et al. / Construction and Building Materials 112 (2016) 925–932 927

Table 1
Composition of cement (wt.%).

Composition SiO2 Al2O3 Fe2O3 CaO MgO SO3 Na2O K2O TiO2 Mn2O3 LOI
Content 21.52 5.13 5.25 61.86 1.46 2.28 0.21 0.83 0.14 0.10 0.05

3. Results and discussion chloride concentration in the range of 0–0.6 mol/L for all the paste
materials; 2) the chloride binding capacity increases with the w/c
3.1. CBIs ratio. However, the difference is also obvious. It can be intuitively
seen that the relationships between free and bound chloride ions
As previously mentioned, half the CBIs were determined here are much more similar to the linear rule comparing with
according to the conventional laboratory methods as a comparison. those in Fig. 1. Considering that, both the linear and the Freundlich
Fig. 1 presents the conventional laboratory-determined CBIs for binding isotherms are used here for fitting, which are presented in
cement pastes of four different w/c ratios. In Fig. 1, the amount Fig. 2. The values of parameters and the correlation coefficient (r2)
of bound chlorides is an average of three replicates, so the error for the four pastes are also given in Fig. 2. From the fitting results, it
bars are used to represent the measured ranges. The two ends of is found that the test plots match well with both the linear and the
each error bar represent the max and the min value of the three Freundlich binding isotherms. The correlation coefficients of the
replicates, respectively. The plots of Fig. 1 show that the amount two types of fitting curves are almost >95%. Even if the Freundlich
of bound chlorides increases with the free chloride concentration binding isotherm is adopted, it can be found that the value of
in the range of 0–0.6 mol/L for all the paste materials used in this parameter b becomes much larger than that in Fig. 1. This evidence
investigation. It also can be clearly seen that the relationship confirms that the binding rules obtained in Fig. 2 are strictly more
between free and bound chloride ions is nonlinear, which is in similar to the linear rule (b up to 1 = linear). This phenomenon
accordance with the results of most researchers [15,34,35]. As explains why a few of researchers found similar linear binding iso-
reported by Tang and Nilsson, the Freundlich binding isotherm therms in field engineering or natural diffusion test. Mohammed
(cb = acbf ) is the most frequently-used binding isotherm for chlo- and Sandberg considered that the linear rules obtained in field
rides in high concentration (free chloride ion concentration engineering and natural diffusion test are due to the leaching of
>0.01 mol/L) [34]. In the present paper, this type of binding iso- hydroxyl ions [36,37]. The specific mechanisms will be discussed
therm is also used for fitting. The fitting curves, the parameter val- in the later sections.
ues and the correlation coefficient (r2) for the four pastes are all
given in Fig. 1. It can be seen that the test results fit well with
the Freundlich binding isotherms. In all cases, the correlation coef-
ficients are >95%. As observed from the fitting curves, the slopes (a) 1.2
y = 1.4104x0.6943, R² = 0.9822
decrease markedly along x-axis. This indicates that the chloride
Bound chloride (% paste sample)

binding capacity decreases markedly with the increasing free chlo- 1


y = 1.188x0.7064, R² = 0.9920
ride concentration. Plus, it is still necessary to mention that the
y = 1.2384x0.8227, R² = 0.9925
chloride binding capacity increases with the w/c ratio though this 0.8
has been a widespread viewpoint. y = 0.9683x0.8339, R² = 0.9959
The rest half of the CBIs were determined using natural diffu- 0.6
sion specimens. As introduced in Section 2.3, the test samples were
taken from different distances of six natural diffusion specimens. 0.4
w/c0.3
Fig. 2 presents the CBIs determined by this method for the cement w/c0.4
pastes of different w/c ratios. It can be easily found that most rules w/c0.5
0.2
observed in Fig. 2 are similar to those observed in Fig. 1, namely, 1)
w/c0.6
the amount of bound chlorides increases markedly with the free
0
0 0.2 0.4 0.6 0.8 1 1.2
Free chloride (mol/L pore solution)
1 (b) 1.2
y = 0.9811x0.3274, R² = 0.9724 y = 1.8928x, R² = 0.9131
Bound chloride (% paste sample)

0.9
Bound chloride (% paste sample)

y = 1.6039x, R² = 0.9556
y = 0.9430x0.3629, R² = 0.9595 1
0.8 y = 1.4336x, R² = 0.9846
0.7 y = 0.8519x0.4642, R² = 0.9739
y = 1.1717x, R² = 0.9931
0.8
0.6 y = 0.7009x0.4536, R² = 0.9367

0.5 0.6
0.4 w/c0.3
w/c=0.3
0.4 w/c0.4
0.3 w/c=0.4
w/c0.5
0.2 w/c=0.5
0.2 w/c0.6
0.1 w/c=0.6

0 0
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2
Free chloride (mol/L pore solution) Free chloride (mol/L pore solution)

Fig. 1. Conventional laboratory-determined CBIs for cement pastes of different w/c Fig. 2. Natural-diffusion-determined CBIs for cement pastes of different w/c ratios,
ratios, and comparison with Freundlich binding isotherms. and comparison with (a) Linear, (b) Freundlich binding isotherms.
928 Z. Song et al. / Construction and Building Materials 112 (2016) 925–932

3.2. CBPs free chloride profiles. In Fig. 3(a), the ratio of bound to total chlo-
ride concentration (Eq. (1)) is used to characterize the chloride
Despite the numerous advantages, CBIs still have some draw- binding capacity. It is found that the chloride binding capacity gen-
backs. For example, they cannot intuitively present how the chlo- erally increases with the distance from the exposed surface,
ride binding capacity varies with the distance from the exposed whether in the measured or the calculated CBPs. This kind of
surface to the deep. Thus, CBPs (Chloride binding profiles) are pro- increase is generally in accordance with the previously mentioned
posed here to improve the drawbacks. In this paper, CBP is defined fact that the chloride binding capacity increases with the decreas-
as a relationship between chloride binding capacity and the dis- ing free chloride concentration, considering the free chloride con-
tance over a range from the exposed surface to the deep. Before centration also decrease with the distance. It is also noticed that
applying the CBPs, an important issue facing us is how to quantita- the increasing rates of the CBI-calculated CBPs are larger than
tively characterize the chloride binding capacity. From our point of those of the measured CBPs. In detail, the chloride binding capac-
view, there are two possible forms to settle it. One is to use the ities obtained from the CBI-calculated CBPs within a certain dis-
ratio of bound to total chloride concentration (cb/ct) and the other tance from the surface are a little larger than those obtained
is to use the ratio of bound to free chloride concentration (cb/cf). By from the measured CBIs, and beyond the distance the rule is oppo-
unifying the units, the two forms are given in Eqs. (1) and (2), site. Considering the fact that the CBI-recalculated CBPs were recal-
respectively. culated from the CBI data, it is safe to draw the conclusion that CBIs
would overestimate the chloride binding capacity beyond a dis-
cb
r cb =ct ¼  100% ð1Þ tance from the exposed surface of OPC pastes, and a little underes-
cb þ cf M Cl um =qw
timate the chloride binding capacity near the surface.
In Fig. 3(b), the ratio of bound to free chloride concentration
cb (Eq. (2)) is used. As can be seen, the difference between the mea-
r cb =cf ¼  100% ð2Þ
cf M Cl um =qw sured CBPs and the CBI-calculated CBPs at the deep distances is
much more obvious. The front cb/cf in the CBI-calculated CBPs is
where, cb is the bound chloride content (%), cf is the free chloride
about 3–8 times of that in the measured CBPs.
concentration (mol/L), Mcl is the molar mass of chloride ions
(35.45 g/mol), um is the mass porosity (%), qw is the water density
(1000 g/L). 3.3. Chemical profiles of pore solution
By using the two forms given in Eqs. (1) and (2), the CBP figure
is presented. Fig. 3 shows the measured CBPs and the ‘CBI- Fig. 4 typically shows the chemical profiles of pore solution
recalculated’ CBPs for the cement pastes with different w/c ratios. pressed from different distances of the w/c 0.5 cement paste.
The solid symbols represent the measured CBPs which were Among the data, the OH concentrations were converted from
obtained in the natural diffusion specimens, and the hollow ones the pH values while the others were directly determined using
are the CBI-recalculated CBPs which were recalculated on the basis the previously mentioned methods.
of the conventional laboratory-determined CBIs and the measured Owing to the concentration difference, the K+ and OH ions near
the surface greatly leached out. Thus, there was a sharp concentra-
tion drop of these ions near the surface. It can be seen from Fig. 4
(a) 1.05 that the K+ and OH concentrations decline gradually from the
1.00
Chloride binding capacity, cb/ct

w/c0.3 meas. deep to the surface. Likewise, it seems reasonable to suppose that
0.95 w/c0.3 calc. the Ca2+ profile also has a descendent front as observed in the K+
0.90
w/c0.4 meas.
and OH profiles. However, the evidence shows that the Ca2+ con-
0.85 centration does not decline but rather has a little growth within a
w/c0.4 calc.
0.80 certain distance. In fact, the similar phenomenon has been
w/c0.5 meas. observed by the previous researchers without giving an explana-
0.75
0.70
w/c0.5 calc. tion [38]. It is known that the leaching of the Ca2+ and OH ions
w/c0.6 meas. would lead to the dissolution/desorption of portlandite and other
0.65
calcium compounds. The concentration of original Ca2+ ions is
0.60 w/c0.6 calc.
too small (about 0.01 mol/L in this paper) to contend against to
0.55
0 10 20 30 40 50 the dissolution/desorption process [39], so the Ca2+ concentration
Distance from the exposed surface (mm)
0.45
(b) 80
0.4
Chloride binding capacity , cb/cf

70 w/c0.3 meas.
Concentraon (mol/L)

0.35
60 w/c0.3 calc.
0.3 OH-
50 w/c0.4 meas.
0.25 Ca2+
w/c0.4 calc.
40 Na+
0.2
w/c0.5 meas.
30 0.15 K+
w/c0.5 calc.
20 0.1
w/c0.6 meas.
10 w/c0.6 calc. 0.05
0 0
0 10 20 30 40 50 0 10 20 30 40 50
Distance from the exposed surface (mm) Distance (mm)

Fig. 3. Measured and ‘CBI-recalculated’ CBPs for cement pastes with different w/c Fig. 4. Chemical profiles in pore solution pressed from different distances of w/c 0.5
ratios using (a) cb/cf, (b) cb/cf to represent chloride binding capacity. cement paste.
Z. Song et al. / Construction and Building Materials 112 (2016) 925–932 929

in the certain distance becomes higher than the original value crystals (AFm, CH) and the needle-like structural crystals (AFt
under the combined action of leaching and dissolution/desorption. [19]). Moreover, the features of the crystals in this zone remain
Owing to the concentration difference, the Na+ ions in the intact and the crystal edges are also clear. Fig. 5(d) presents the
source solutions also penetrated into the cement pastes along with morphology of the transitional zone between degraded and unde-
the Cl ions. It is found that the Na+ ions penetrated much slower graded zone. Most familiar crystals can also be found here. How-
than Cl. This is much different from the rule obtained in the dif- ever, in this zone, the AFt crystals almost disappear and the
fusion process of NaCl electrolyte solution. In the electrolyte solu- edges of the crystals become somewhat unclear.
tion, Na+ ions will diffuse at the same speed as Cl ions so as to Fig. 6 presents the EDS linear scanning results for the w/c 0.5
keep the electrical neutrality [40]. In the cement hydrate system, cement paste. Owing to the non-uniformity of cement paste, the
not only the Na+ ions but also other species of ions and solid phases chemical element profiles are found a little disordered, especially
will have a response to the chloride diffusion. Thus, it is perhaps in some small-scale regions. At large enough scales, however, clear
nothing unusual to find the Na+ ions penetrated much slower than trends can be observed from the large amounts of data. As
Cl in cement hydrate systems considering the fact that the bulk observed in Fig. 6, the calcium content generally decreases from
diffusion coefficient of Na+ itself is much smaller than Cl. the deep to the surface of the specimen. In the degraded zone, this
It can be concluded that the chemical composition of pore kind of decrease is more evident. This indicates that the leaching of
solution changes discriminatively and obviously at different calcium ions in pore solution had been influencing the solid phases
distance when natural chloride diffusion takes place. The change near the surface. Further, it can be reckoned that the portlandite,
of the pore solution composition is also potential to influence the C-S-H gels, AFm phases and other calcium compounds near the
solid phases. That is to say the entire ICE will change along with surface had been largely dissolved. Due to the dissolution of C-S-
chloride diffusion. It is because of the ICE change that the chloride H gels and AFm phases, chloride ions lost their binding matrix
binding capacity also changes discriminatively at different dis- and therefore the content of bound chlorides essentially decreased.
tance. In Section 3.5, a model is built up to describe chloride bind- Thus, it is found in Fig. 6(c) that the chlorine profile has a drop near
ing capacity considering the change of an important ICE factor. the surface. It is noticed that, if only considering the impact of the
free ions (e.g. the hydroxyl ions) on chloride binding, this kind of
3.4. Microscopy test chloride drop would be unexpected. It is known that, as a result
of leaching, the hydroxyl ion concentration near the surface largely
Fig. 5 presents the morphology of cement hydrates of the w/c decreased [41,42]. The concentration decrease would also reduce
0.5 cement paste. It can be seen from Fig. 5 that the microstructure the binding competition of hydroxyl ions with chloride ions, and
of the near-surface hydrates changes a lot after natural diffusion, as thereby promoting the chloride binding [17,43]. According to this
expected. In Fig. 5(a), a narrow degraded zone (about 0–0.2 mm theory, the content of bound or total chlorides near the surface
from the surface) is found near the surface at low magnification should be the largest. Considering the opposite rule observed in
(30 times). Fig. 5(b) presents the morphology of the degraded zone EDS results, it can be assumed that the solid phases (being an
at a higher magnification (2.0 k times). It is found that the important element of ICE) will also impact the chloride binding
degraded zone is totally amorphous. None of the familiar crystals capacity.
can be observed in that zone. For comparison, Fig. 5(c) presents In this investigation, the degrading degree is negligibly small
the morphology of the undegraded zone at 2.0 k magnification. due to the short immersion duration and perhaps the non-flowing
The common crystals existed in cement hydrates can be easily solution. Therefore, the zone where the bound chloride is evidently
found in the undegraded zone, such as the hexagonal structural influenced by the solid phase degradation (i.e. the chloride drop zone)

Fig. 5. Morphology of cement hydrates of w/c 0.5 cement paste: (a) overview, (b) degraded zone, (c) transitional zone, (d) undegraded zone.
930 Z. Song et al. / Construction and Building Materials 112 (2016) 925–932

Fig. 6. EDS linear scanning for w/c 0.5 cement paste (a) scanning position, (b) oxygen profile, (c) chlorine profile, (d) calcium profile.

is also small (about 0.75 mm in depth). It is noticed that the small anion exchange mechanism [17,44], the hydroxyl ion concentra-
chloride drop zone cannot even be presented by the traditional tion is perhaps the most important factor influencing the
sampling and titration method since the coarse sampling intervals distance-associated chloride binding among the ICE factors. Thus,
(usually 3–5 mm). In fact, in most cases, the chloride drop zone is in a hydroxyl-dependent chloride binding model is established here.
small depth and can be ignored comparing with the long chloride As we know, the ratio of free chloride concentration to hydroxyl
diffusion depth. This does not mean that the impact of the solid concentration is frequently used to estimate the critical value for
phase degradation on chloride binding is small enough to be steel bar corrosion [45]. Cooperating with this type of critical value,
ignored in all cases. In some long duration tests, the chloride drop the hydroxyl-dependent chloride binding model is probably very
zone has been found much obvious [44]. Perhaps in the flowing compatible to predict the service life of reinforced structures.
or aggressive environment (e.g. the flowing seawater, NH4Cl-rich Eq. (3) shows the general reaction of chloride binding under the
underground water), the impact of the solid phase degradation on chloride-hydroxyl ion exchange mechanism.
chloride binding would be much stronger. Anyway, the small but 
R  OH þ Cl () R  Cl þ OH ð3Þ
truly existed chloride drop zone found in EDS results illustrate that
the chloride binding capacity is definitely impacted by solid phase where R-OH represents the binding matrix including CSH gels and
degradation (or change). AFm phases, R-Cl represents all the solid chloride phases. On the
basis of the mass action law and the empirical adsorption kinetics,
the positive and negative reaction rates of Eq. (3) are assumed as:
3.5. Modeling
r1 ¼ k1 cROH cm
f ; r 2 ¼ k2 c b cOH
n
ð4Þ
Chloride binding capacity is strongly impacted by ICE, which
contains many ions and solid phases in the cement hydrate system. where, cR-OH is the content of binding matrix; cOH is the hydroxyl
However, the influence of the ICE factors is not able to be repre- ion concentration; k1, k2, m and n are constants. When reaching
sented in the conventional CBIs. Thus, it is urgent to build up a equilibrium, the content of bound chlorides can be calculated by
multifactorial chloride binding model to apply it into practical letting r1 = r2. Eq. (5) shows the detailed formula and the simplify-
use. In most situations, it is difficult and unnecessary to consider ing process. The term in the box is the final model used in the pre-
all the ICE factors when describing the chloride binding capacity. sent paper, which is deduced on the basis of the understanding that
Before creating an applicable model, the preliminary work to be the content of binding matrix changes little with chloride binding.
done is to analyze the influence degrees of the multi ICE factors k1 c m
f c ROH Cm C OH const:
in line with the specific situation, and then select one or a few of cb ¼  k Cn f  ƒƒƒƒƒƒ!
ƒƒƒƒƒƒ cb ¼ acbf ð5Þ
k2 cnOH OH
the ICE factors as the determining factors for modeling. From our
point of view, it can still be a kind of progress even if only one It is found that, if the hydroxyl concentration is constant, the
ICE factor is taken into account. model in the box is the same as the Freundlich binding isotherm.
As previously mentioned, in most cases, the impact of solid This more or less reflects the rationality of the hydroxyl-
phase variation is very small and can be ignored. Considering the dependent model.
Z. Song et al. / Construction and Building Materials 112 (2016) 925–932 931

0.50 natural diffusion takes place. It is because of the change of the ICE
0.45 that the chloride binding capacity at different distance also devel-
0.40 ops discriminatively. In this study, the hydroxyl concentration is
0.35 considered to be the most important factor among the ICE factors
influencing the chloride binding capacity of the specimens under-
0.30 Measured plots:
cOH-

Fitting curve: w/c 0.5 going natural diffusion. Therefore, a hydroxyl-dependent chloride
0.25
y = 0.125x0.7087 binding model is established to quantify the distance-associated
0.20 R² = 0.9047
chloride binding capacity. The model explains the difference
0.15 between the conventional laboratory-determined CBIs and the
0.10 CBIs determined using natural diffusion specimens, and it fits well
0.05 with the test results.
0 2 4 6 8
cfm/cb
Acknowledgements
Fig. 7. Fitting curve of hydroxyl-dependent chloride binding model.

The authors wish to acknowledge the financial support by the


The final model clearly shows the impact of hydroxyl concen- National Natural Science Foundation of China under Project Nos.
tration on chloride binding capacity and explains why the 51278167 and 51479051 as well as the Natural Science Foundation
natural-diffusion-determined CBIs are different from the conven- of Jiangsu province under Project No. BK20131374. The supports of
tional CBIs. In the conventional laboratory methods, the hydroxyl the Experiment Center of Mechanics and Materials in Hohai
concentrations of the parallel disc specimens are similar to each University have been most helpful and are appreciated. The State
other after the same duration of leaching, so the Freundlich bind- Key Laboratory of Hydrology-water Resources and Hydraulic Engi-
ing isotherm is usually found by using these methods. In the natu- neering is also appreciated for providing some important test
ral diffusion method, the hydroxyl concentration increases with equipments.
the distance while the free chloride concentration decreases with
the increasing distance. Therefore, the larger the free chloride con-
References
centration is, the smaller the decreasing scope of the chloride bind-
ing capacity is. Thus, the natural-diffusion determined CBIs are [1] M. Dobrescu, M. Vasilescu, Reinforcing bar corrosion for buildings
found more linear. construction, Adv. Mater. Res. 1114 (2015) 219–223.
According to Eq. (5), the parameter m is the same as the param- [2] J. Zhu, Y. Zhang, D. Zhao, Durability assessment of an RC railway bridge pier
under a chloride-induced corrosion environment, Fifth International
eter b. Thus, the value of m can be evaluated from the conventional Conference on Transportation Engineering (2015).
laboratory-determined CBIs. Then the parameters k and n can be [3] L. Zeng, R. Song, Controlling chloride ions diffusion in concrete, Sci. Rep. 3
solved by fitting the equation to the experimental results deter- (2013) 1–7.
[4] Zijian Song, Linhua Jiang, Jianzhong Liu, Jiaping Liu, Influence of cation type on
mined using natural diffusion specimens. For the w/c 0.5 cement
diffusion behavior of chloride ions in concrete, Constr. Build. Mater. 99 (2015)
paste, the m is about 0.3629. Using cm f /cb as the independent vari- 150–158.
able and cOH as dependent variable, Eq. (5) are converted into a [5] E. Badogiannis, E. Aggeli, V.G. Papadakis, S. Tsivilis, Evaluation of chloride-
simple power function for fitting. The measured plots, the fitting penetration resistance of metakaolin concrete by means of a diffusion–Binding
model and of the k-value concept, Cement Concr. Compos. 63 (2015) 1–7.
curve, the obtained power function and the correlation coefficient [6] K. Tuutti, Corrosion of Steel in Concrete, Swedish Cement Concrete Research
(r2) are all given in Fig. 7. It can be seen that the new model fit well Institute, Stockholm, 1982.
with the test results (r2 > 90%). From the fitting result, the new [7] V. Baroghel-Bouny, X. Wang, M. Thiery, M. Saillio, F. Barberon, Prediction of
chloride binding isotherms of cementitious materials by analytical model or
model for the w/c 0.5 cement paste is: numerical inverse analysis, Cem. Concr. Res. 42 (9) (2012) 1207–1224.
[8] Q. Yuan, C. Shi, et al., Chloride binding of cement-based materials subjected to
c0:3629
f external chloride environment – a review, Constr. Build. Mater. 23 (1) (2009)
cb ¼ 0:0532 ð6Þ 1–13.
c1:4110
OH
[9] A. Neville, Chloride attack of reinforced concrete: an overview, Mater. Struct.
28 (176) (1995) 63–70.
[10] A.K. Suryavanshi, J.D. Scantlebury, S.B. Lyon, Corrosion of reinforcement steel
4. Conclusion embedded in high water–cement ratio concrete contaminated with chloride,
Cement Concr. Compos. 20 (3) (1998) 263–269.
[11] B. Reddy, G.K. Glass, P.J. Lim, N.R. Buenfeld, On the corrosion risk presented by
This study dealt with the distance-associated chloride binding chloride bound in concrete, Cem. Concr. Compos. 24 (1) (2002) 1–5.
capacity of cement pastes including four different w/c ratios from [12] P.F. McGrath, Development of test methods for predicting chloride penetration
0.30 to 0.60. Specimens undergoing natural diffusion were used for into high performance concrete (Ph.D. thesis), Department of Civil
Engineering, University of Toronto, Canada, 1996.
measuring the chloride binding capacity at different distance. CBIs, [13] B. Martin-Perez, H. Zibara, R.D. Hooton, M.D.A. Thomas, A study of the effect of
CBPs, chemical profiles and microscopy results were performed to chloride binding on service life predictions, Cem. Concr. Res. 30 (2000) 1215–
analyze the influence rules and mechanisms of distance on chlo- 1223.
[14] H.A. Berman, Determination of chloride in hardened Portland cement paste,
ride binding capacity, and then the conclusions are presented mortar, and concrete, J. Mater. 7 (3) (1972) 330–335.
below. [15] A. Delagrave, J. Marchand, J. Olliver, S. Julien, K. Hazrati, Chloride binding
The conventional laboratory-determined CBIs fit well with the capacity of various hydrated cement systems, Adv. Cem. Based Mater. 6 (1997)
28–35.
Freundlich binding isotherms. In comparison, the natural- [16] H. Justnes, A review of chloride binding in cementitious systems, Nord. Concr.
diffusion-determined CBIs are much more similar to the linear Res. 21 (1998) 1–6.
rules. By analyzing the CBPs results, it is found that the conven- [17] A.K. Suryavanshi, J.D. Scantlebury, S.B. Lyon, Mechanism of Friedel’s salt
formation in cements rich in tri-calcium aluminate, Cem. Concr. Res. 26 (5)
tional laboratory-determined CBIs would overestimate the chlo-
(1996) 717–727.
ride binding capacity beyond a certain distance from the exposed [18] Y. Chen, Z. Shui, W. Chen, G. Chen, Chloride binding of synthetic Ca–Al–NO3
surface of OPC pastes, and underestimate the chloride binding LDHs in hardened cement paste, Constr. Build. Mater. 93 (2015) 1051–1058.
capacity near the surface. [19] H.F.W. Taylor, Cement Chemistry, 2nd ed., Thomas Telford Publishing, London,
1997.
On the basis of the chemical profiles and the microscopy results, [20] L.G. Baquerizo, T. Matschei, K.L. Scrivener, M. Saeidpour, L. Wadsö, Hydration
it is safe to draw the conclusion that the entire ICE changes when states of AFm cement phases, Cem. Concr. Res. 73 (2015) 143–157.
932 Z. Song et al. / Construction and Building Materials 112 (2016) 925–932

[21] Luis G. Baquerizo a, Thomas Matschei a, Karen L. Scrivener b, Mahsa Saeidpour [34] L. Tang, L.-O. Nilsson, Chloride binding capacity and binding isotherms of OPC
c, Lars Wadsö c, Hydration states of AFm cement phases, Cem. Concr. Res. 73 pastes and mortars, Cem. Concr. Res. 23 (2) (1993) 247–253.
(2015) 143–157. [35] W. Sergi, S.W. Yu, C.L. Page, Diffusion of chloride and hydroxyl ions in
[22] R. Allmann, Refinement of the hybrid layer structure [Ca2Al(OH)6]+[1/ cementitious materials exposed to a saline environment, Mag. Concr. Res. 44
2SO43H2O], Neues Jahrbuch für Mineralogie Monatshefte 136–144 (1997). (158) (1992) 63–69.
[23] T. Matschei, B. Lothenbach, F.P. Glasser, The AFm phase in Portland cement, [36] T.U. Mohammed, H. Hamada, Relationship between free chloride and total
Cem. Concr. Res. 37 (2007) 118–130. chloride contents in concrete, Cem. Concr. Res. 33 (2003) 1487–1490.
[24] M. Francois, G. Renaudin, O. Evrard, A cementitious compound with [37] P. Sandberg, Studies of chloride binding in concrete exposed in a marine
composition 3CaOAl2O3CaCO311H2O, Acta Crystallogr., C 54 (9) (1998) environment, Cem. Concr. Res. 29 (1999) 473–477.
1214–1217. [38] E. Samson, J. Marchand, J.L. Ropert, J.P. Bournazel, Modeling ion diffusion
[25] V.S. Ramachandran, Possible states of chloride in the hydration of tricalcium mechanisms in porous media, Int. J. Numer. Meth. Eng. 46 (12) (1999) 2043–
silicate in the presence of calcium chloride, Mater. Struct. 4 (1) (1971) 3–12. 2060.
[26] A.K. Suryavanshi, J.D. Scantlebury, S.B. Lyon, The binding of chloride ions by [39] L. Jiang, Z. Song, H. Yang, Q. Pu, Q. Zhu, Modeling the chloride concentration
sulphate resistant cement, Cem. Concr. Res. 25 (3) (1995) 581–592. profile in migration test based on general Poisson Nernst Planck equations and
[27] C. Arya, N.R. Buenfeld, J.B. Newman, Factors influencing chloride binding in pore structure hypothesis, Constr. Build. Mater. 40 (1) (2013) 596–603.
concrete, Cem. Concr. Res. 20 (2) (1990) 291–300. [40] Y. Li, J. Lu, Theory of Electrolytic Solution, Tsinghua University Press, Beijing,
[28] D.F.H. Rasheeduzzafar, S.S. Al-Saadoun, A.S. Al-Fahtani, F.H. Dakhil, Effect of 2005.
tricalcium alumina content of cement on corrosion of reinforcing steel in [41] Y. Maltaisa, E. Samsona, J. Marchanda, Predicting the durability of Portland
concrete, Cem. Concr. Res. 20 (5) (1990) 723–738. cement systems in aggressive environments—laboratory validation, Cem.
[29] J.J. Beaudoin, V.S. Ramachandran, R.F. Feldman, Interaction of chloride and C– Concr. Res. 34 (2004) 1579–1589.
S–H, Cem. Concr. Res. 20 (6) (1990) 875–883. [42] K. Yokozeki, K. Watanabe, N. Sakata, N. Otsuki, Modeling of leaching from
[30] P.W. Brown, S. Badger, The distributions of bound sulfates and chlorides in cementitious materials used in underground environment, Appl. Clay Sci. 26
concrete subjected to mixed NaCl, MgSO4, NaSO4 attack, Cem. Concr. Res. 30 (1) (2004) 293–308.
(2000) 1535–1542. [43] J. Tritthart, Chloride binding in cement II. The influence of the hydroxide
[31] A.K. Suryavanshi, R. Narayanswamy, Stability of Friedel’s salt in carbonation in concentration in the pore solution of hardened cement paste on chloride
carbonated concrete structural elements, Cem. Concr. Res. 26 (5) (1996) 717– binding, Cem. Concr. Res. 19 (5) (1989) 683–691.
727. [44] B. Johannesson, K. Yamada, L.O. Nilsson, Y. Hosokawa, Multi-species ionic
[32] A. Neville, Chloride attack of reinforced concrete: an overview, Mater. Struct. diffusion in concrete with account to interaction between ions in the pore
28 (176) (1998) 63–70. solution and the cement hydrates, Mater. Struct. 40 (7) (2007) 651–665.
[33] Y. Zhang, Mechanics of chloride ions transportion in concrete (Ph.D. thesis), [45] Y. Wang, G. Fang, W. Ding, N. Han, Xing F, Self-immunity microcapsules for
College of Civil Engineering and Architecture, Zhejiang University, Hangzhou, corrosion protection of steel bar in reinforced concrete, Sci. Rep. 5 (2015).
China, 2008.

You might also like