Вы находитесь на странице: 1из 10

Construction and Building Materials 192 (2018) 28–37

Contents lists available at ScienceDirect

Construction and Building Materials


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

Effect of interfacial transition zone on the transport of sulfate ions in


concrete
Dandan Sun a,b, Kai Wu a, Huisheng Shi a, Lintao Zhang a, Lihai Zhang b,⇑
a
School of Materials Science and Engineering, Tongji University, Shanghai 201804, China
b
Department of Infrastructure Engineering, The University of Melbourne, Melbourne, VIC 3010, Australia

h i g h l i g h t s

 Reactive transport model of sulfate ions in concrete were proposed.


 The effect of interfacial transition zone (ITZ) on the transport of sulfate ions was researched.
 Experiments to obtain the content of sulfate ions and porosity of ITZ were carried out.
 The distribution of ettringite in ITZ and concrete were simulated.

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

Article history: Long-term external sulfate attack can result in significant concrete degradation. The purpose of this study
Received 19 April 2018 is to develop an experimentally validated reactive transport model to investigate the transport behaviour
Received in revised form 7 August 2018 of sulfate ions in concrete containing interfacial transition zone (ITZ). The model takes into account the
Accepted 15 October 2018
spatially dependent pore size in ITZ and the change of diffusion coefficient of sulfate ions due to the for-
Available online 22 October 2018
mation of ettringite and gypsum in concrete pores resulting from the chemical reaction between sulfate
ions and concrete components. The developed model was firstly validated by performing a series of
Keywords:
experiments involving concrete specimens and sodium sulfate solution. Then, a series of parametric stud-
Concrete
Interfacial transition zone (ITZ)
ies were carried out to identify the critical parameters that govern the transport of sulfate ions in con-
Sulfate ion crete. The backscattered electron microscopy (BSE) results show that the porosity of concrete within
Reactive transport ITZ is relatively higher than that in bulk paste with a six-fold higher porosity in inner ITZ (5 mm from
Ettringite aggregate surface) than that in outer ITZ (50 mm from aggregate surface). The model predictions sug-
gested that the concentration of sulfate ions initially increases rapidly in concrete during the first 20 days
and then the rate of increase gradually decreases due to the reduction in the diffusion coefficient of ions
as a result of the ettringite and gypsum formation in concrete pores. In addition, it demonstrates that self-
induced electric field has little influence in the transport of sulfate ions.
Ó 2018 Elsevier Ltd. All rights reserved.

1. Introduction and therefore initializes the degradation process [3]. The failure
mode of concrete resulting from sulfate attack can be characterized
The long-term performance assessment of concrete structure as the material expansion, cracking and spalling due to the forma-
under the environment of seawater or natural fresh water, are of tion of secondary gypsum and ettringite as a result of reaction
vital importance due to the failure consequences caused by ions between sulfate ions and different aluminate phases [4–7]. Thus,
attack. Different forms of ions attack, like chloride ions, sulfate ions when exposing to an environment naturally rich in sulfate ions,
and magnesium ions would be the origin of the degradation of con- the service life of a concrete structure could be significantly
crete [1,2]. One practically complicated degradation progress is shorten. Much effort has been spent in understanding the sulfate
external sulfate attack which can be described as the penetration attack on concrete in the past decades, the fundamental mecha-
of sulfate ions into concrete materials which is porous in nature, nism of the failure progress has not been fully understood [2,3,8–
12].
In order to fundamentally understand the mechanism of sulfate
⇑ Corresponding author.
attack with the aim of prolonging the service life of concrete struc-
E-mail addresses: dandan.sun@unimelb.edu.au (D. Sun), lihzhang@unimelb.edu.
tures, a series of experimental and theoretical studies have been
au (L. Zhang).

https://doi.org/10.1016/j.conbuildmat.2018.10.140
0950-0618/Ó 2018 Elsevier Ltd. All rights reserved.
D. Sun et al. / Construction and Building Materials 192 (2018) 28–37 29

conducted [3,7,9,11–16]. Fick’s second law is widely used for mod- by sulfate attack is a critical factor governing the durability of the
elling the diffusion of sulfate ions and mechanisms for the reac- concrete structures, and therefore concrete cracking due to sulfate
tions of calcium aluminates with sulfate ions are also considered attack would be incorporated in our future research. In present
in extended equations [3,11,17]. As the reactive transport of sulfate study, we developed a comprehensive reactive-transport approach
ions progresses, gypsum and ettringite gradually fill the concrete by modelling the concrete as a three-phase porous material (i.e.
pores, and ultimately cause expansion stresses. The rate of ettrin- paste, aggregate and ITZ). In addition, the model takes into account
gite formation depends on many environmental factors, like con- the effects of the electrical field on sulfate ion transport. Finally, an
centration of sulfate ions, pH of solution, temperature and so on experimental study was carried out to validate the model.
[18,19]. To understand sulfate ion induced corrosion process, sev-
eral models have been developed to quantify the expansive volume
2. Methodology
strain of concrete exposed to sulfate solution by considering
cement paste or concrete as a two phase porous network at meso-
Fig. 1 shows a schematic diagram of three-phase concrete
scale [3,20]. However, these models did not take into account the immersed in sulfate solution. In corroding area of the concrete
influence of ITZ which plays an important role in the corrosion pro-
(assuming no cracks in this zone), sulfate ions diffuse through cap-
cess. ITZ was formed owing to the ‘‘wall effect”, and appears gradi- illary pores and producing ettringite and gypsum through reacting
ent distribution of porosity, the platelike calcium hydroxide with the hydrated cement composites. For simplicity, it is assumed
crystals in it tend to form in oriented layers, which are the major
that concrete is a homogeneous porous material immersed in sat-
factors for its poor performance [21–24]. The micro-cracks urated sulfate solution, and the chemical reaction under external
occurred in ITZ are preferential path for the creation of intercon-
sulfate attack can be analysed with two successive chemical reac-
nections leading to a more permeable system [25,26]. Bourdette tions, which will be discussed in the next section.
found that the influence of ITZ in corrosion process should be
included in theoretical models in order to reproduce the experi-
2.1. Governing equations
mental observations [27], due to the fact that ITZ has a relatively
higher porosity in comparison to bulk paste, and therefore allows
The purpose of this study is to develop a theoretical model
sulfate ions to diffuse more easily. Irassar reported that the depo-
which takes into account diffusion, migration and chemical reac-
sition of ettringite and gypsum around the aggregate indicates the
tion of sulfate ions. In addition, the electrical field caused by the
vulnerability of the ITZ to chemical attack [28]. From a transport
movement of sulfate ions is modelled by using the extended
point of view, the effective diffusion coefficient in concrete is basi-
Nernst-Planck equation. Assuming the advective transport of sul-
cally dependent on the diffusivity of paste, the diffusivity of inter-
fate ions due to fluid flow is ignorable, the transport equation of
face and the volume fraction of aggregate and ITZ [21,23,29]. The
sulfate ions can be described as
presence of aggregates in concrete influences diffusion through
several effects; The dilution effect and tortuosity effect could inhi-
 
dci Fz
bit the ion transport, whereas the special microstructure of ITZ ¼ r Deff rci þ Deff ci ðr/Þ þ Si ð1Þ
dt RT
may enhance the ion transport, particularly in cement based mate-
rials [21,30]. Previous research showed that the effective diffusion where ci is the ion concentration in concrete (i.e. sulfate ion); Deff is
coefficient of chloride ions is 6–12 times greater in ITZ than that in the effective diffusion coefficient of ions in concrete; F is the Fara-
the bulk cement [31,32]. Without considering the microstructural day constant; z is the valency of ions; R is the gas constant; T is
changes that accompany the process of sulfate attack, it is impos- the environmental temperature; / is the electrical potential; and
sible to theoretically predict the damage progression [33,34]. Si is the source terms resulting from the chemical reaction between
Although the importance of ITZ on sulfate attack has been widely sulfate ions and different aluminate phases of concrete.
recognized, the research in this area is still limited [35,36].
For simplicity, this study mainly focuses on the elastic stage of 2.1.1. Chemical reaction of sulfate ions in concrete
concrete and the cracking of concrete is ignored in the model sim- Assuming the ingressing sulfates firstly react with calcium ions
ulation. However, it has been known that concrete cracking caused from portlandite (CH) or calcium silicate hydrates (C-S-H) to form

Fig. 1. Schematic diagram of three-phase concrete immersed in sulfate solution.


30 D. Sun et al. / Construction and Building Materials 192 (2018) 28–37


gypsum (C S H2 ), the dissolution of CH and the decalcification of C- diffusion in inert porous medium, the diffusivity of sulfate ions in
S-H lead to the diffusion of calcium ions towards the outside med- concrete could be dependent on ‘‘pore effect”, which is relevant to
ium. The deterioration of concrete durability is caused by the for- the hydration, chemical reaction production and cracks caused by
 expansion. While continuous hydration of cement and the filling
mation of ettringite (C6 AS3 H32 ) associated with the chemical effect of ettringite and secondary gypsum will improve the density

reactions of secondary gypsum (C S H2 ) and hydrated calcium alu- of concrete, the damage caused by expansive products may result
minates, such as tricalcium aluminate (C3A), tetracalcium alumi- in cracks, and ultimately the increase of concrete porosity [41].

Several attempts have been made to describe effects of the pore
nate (C4AH13), and monosulfate (C4 A S H12 ).
filling on the diffusion coefficients of the sulfate ions [7,42]. How-

Ca2þ þ SO2 ever, the problem becomes further complicated due to the fact that
4 þ 2H ! C S H2 ð2Þ
concrete is a composite material with three phases, and the
  microstructure of ITZ is different from that of the bulk cement
C3 A þ C S H2 þ 26H ! C6 AS3 H32 paste due to the packing constrains imposed by the aggregate sur-
  face [22].
C4 AH13 þ 3C S H2 þ 14H ! C6 AS3 H32 þ CH ð3Þ To determine the initial diffusion coefficient of sulfate ions in
concrete, ITZ with 50 lm thickness is explicitly considered. As
  
C4 A S H12 þ 2C S H2 þ 16H ! C6 AS3 H32 shown in Fig. 1, concrete could be considered as three phase and
its initial diffusion coefficient can be given as [13,43]:
 
3C4 AF þ 12C S H2 þ aH ! 4C6 AS3 H32 þ 2½ðA; FÞH3  6Db ð1  Va ÞðVa þ VI Þ þ 2VI ðDI  Db Þð1 þ 2Va þ 2VI Þ
D0eff ¼ Db ð9Þ
3Db ð2 þ Va ÞðVa þ VI Þ þ 2VI ð1  Va  VI ÞðDI  Db Þ
For simplicity, Eq. (3) describing the multiple processes of the
formation of ettringite (AFt) may be advantageously lumped into where Va and VI are volume fraction of aggregate and ITZ, respec-
one single expression as: tively; Db and DI are the diffusion coefficient of bulk cement paste
  and ITZ, respectively. It can be described as follows [39]:
CA þ nC S H2 ! C6 AS3 H32 ð4Þ
2ui
 Di ¼ D0 ui b ¼ D0 ð10Þ
where CA ¼ c1 C3 A þ c2 C4 AH13 þ c3 C4 A S H12 þ c4 C4 AF, and 3  ui
n ¼ 3c1 þ 2c2 þ 3c3 þ 4c4 represents the stoichiometric weighted where D0 is the diffusion coefficient of ions in water, /i is porosity of
coefficient of the sulfate phase and ci is the proportion of each alu- the phase in concrete, b is the tortuosity of the pore system, and can
minate phase, calculated as [3]: be written based on Miri-Tanaka method. That is,
C 2
ci ¼ P4 Ai ð5Þ b¼ ð11Þ
i¼1 CAi 3  ui
in which cA_i represents the molar concentration of each aluminate For given aggregate gradation and volume fraction, spheroidal
phase per unit volume of solid material. For the convenience of cal- aggregate particles with various sizes could be generated. The ITZ
culation, n is assumed to be 8/3, an equivalent reacting coefficient of volume fraction in concrete can be obtained using Monte Carlo
gypsum dissipation to produce ettringite [11]. method [22,44,45]. That is,
According to Eqs. (2) and (4), the chemical reaction rate of dis- 2 3
V I ¼ ð1  V a Þ  ð1  V a Þexp½ðt1 h þ t 2 h þ t 3 h Þ ð12Þ
sipation of sulfate ions and formation of gypsum can be expressed
as: where h is the thickness of ITZ, t1, t2, t3 are respectively defined in
@cSO2 terms of the volume fraction Va, mean diameter hDi, mean surface
4
¼ k1  cCa2þ  cSO2 ð6Þ area hSi and mean volume hVi of spherical aggregate particles as
@t 4
[44]:
@cgyp  n V a h Si
¼ k1  cCa2þ  cSO2  n  k2  cgyp  cCA ð7Þ t1 ¼ ð13Þ
@t 4 ð1  V a ÞhV i
where k1 is the chemical reaction rate of Eq. (2); and k2 is the chem-
ical reaction rate of take-up of sulfates. cCa2þ is the calcium ion con- 2pVa hDi V 2a hSi2
t2 ¼ þ ð14Þ
centration in pore solution, which can be assumed to be a constant ð1  V a ÞhV i 2ð1  V a Þ2 hV i2
at saturated solution, i.e. 21.25 mol/m3 at 298 K. cSO2 is the concen-
4

tration of sulfate ions, cCA represents the concentration of equiva- 4pVa 2pV 2a hDihSi
t3 ¼ þ ð15Þ
lent lumped reacting calcium aluminates (CA), and cgyp is the 3ð1  V a ÞhV i 3ð1  V a Þ2 hV i2
concentration of secondary gypsum.
Thus, The initial diffusion coefficient of sulfate ions in concrete could
be calculated through Eqs. (9)–(15) and measuring the porosity in
@cCA @cEtt  n ITZ and paste in this study, while the time-dependent diffusion
¼ ¼ k2  cgyp  cCA ð8Þ
@t @t coefficient of sulfate ions would change due to the filling effect
where cEtt is the concentration of ettringite. of reaction product (both ettringite and gypsum) resulting from
chemical reaction in early stage. In the present study, the current
2.1.2. Effective diffusion coefficient of sulfate ions diffusion coefficient of sulfate ions as the reaction progresses (Deff)
Effective diffusion coefficient of sulfate ions depends on the ion can be express as:
" !#
concentration in solution [37,38] and is also time and spatial
Deff Cp
dependent. The diffusion coefficient of sulfate ions in a porous ¼ exp a  ð16Þ
D0eff q  cmax
p
material, such as concrete and mortar, depend on the pore
structure characteristics of the material [39,40]. Unlike the simple
D. Sun et al. / Construction and Building Materials 192 (2018) 28–37 31

8
where cp ¼ cEtt þ cgyp , cEtt and cgyp represent the concentration of > cs ðx ¼ 0; tÞ ¼ uc c0 ; cs ðx ¼ L; tÞ ¼ uc c0
>
>
ettringite and gypsum, respectively. cmax is the maximum possible < c ðx ¼ 0; tÞ ¼ cini ; c ðx ¼ L; tÞ ¼ cini
p Ca Ca s Ca
ð22Þ
concentration of reaction products (both ettringite and gypsum) > cCA ðx ¼ 0; tÞ ¼ cini
> ; c ð x ¼ L; tÞ ¼ cini
>
: CA s CA
when all the pores in concrete are filled, which can be expressed
cEtt ðx ¼ 0; tÞ ¼ 0; cs ðx ¼ L; tÞ ¼ 0
in Eq. (17). The parameters a and q in Eq. (16) represent the filling
effect of reaction products on the diffusion coefficient of sulfate where L is the length of the concrete specimen, t is the immersion
2+
ions, and the values of these two parameters are obtained through time, cini ini
Ca and cCA are the initial concentration of Ca and CA, respec-
calibrating the experimental data on different sulfate-sodium con- tively, c0 is concentration of sulfate ions in solution.
centrations (i.e. 5 g/L and 50 g/L) as shown in Fig. 7.
uc 2.1.5. Initial conditions
cmax ¼ ð17Þ The initial condition for the present experiment can be
p
NA MINfV Ett; V gyp g
described as follows:
where uc is the porosity of concrete, NA is Inm constant, VEtt and 8
> cs ðx; t ¼ 0Þ ¼ 0;
Vgyp represent the unit cell volume of ettringite and gypsum, >
>
< c ðx; t ¼ 0Þ ¼ cini ;
respectively. The influence of ionic concentration, pore characteris- Ca Ca
x 2 Mp ð23Þ
tics and reaction production are considered in the current diffusion >
> cCA ðx; t ¼ 0Þ ¼ cCA ;
ini
>
:
coefficient (Eqs. (16) and (17)) which is governed by the total con- cEtt ðx; t ¼ 0Þ ¼ 0;
tent of both ettringite and gypsum in concrete pores.
where x is the section position of concrete specimen.
2.1.3. Self-induced electrical field
As shown in Fig. 2, the electrical field is created due to the dif- 2.1.6. Numerical solutions
ferent diffusion coefficients of sulfate ions and sodium ions [46]. The model was solved numerically using the finite element soft-
The fluxes (Ji) of the ions can be described by the Nernst-Planck ware package COMSOL Multiphysics and the values of parameters
equation [47]. used in this study are shown in Table 1.
As shown in Fig. 3, a 2D geometry model of random packing for
dci  
Ji ¼ ¼ r Deff rci þ um;i zi Fci ðr/Þ ð18Þ aggregate in three-phase concrete was created and numerically
dt modelled using COMSOL. The length of the square is 0.1 m based
where um,i is the mobility. It can be seen that the flux of sulfate ions on the experimental study. It is assumed that ITZ is a homogeneous
are dependent on the fluxes of other species, such as sodium ions. In thin layer surrounding the spherical aggregate with a thickness of
present study, it is assumed that the self-induced electrical field 50 lm, and ITZ layers are allowed to overlap with each other. The
only exists in a thin area closed to sulfate solution. Choosing the determination of the size of the three-phase composition sphere is
middle area of concrete as the ground condition for the electrolyte based on the mixture of concrete (i.e. cement: sand: coarse aggre-
potential, we obtain: gate = 45%: 30%: 25%) measured in Section 2.2, and this distribu-
tion is used in previous studies [33,36]. The volume fraction of
/0 ¼ 0 ð19Þ
ITZ is relatively small, and therefore is ignored in this study. A tri-
For the potential, the Poisson equation (Gauss’s law) states
r  ðer/Þ ¼ q ð20Þ Table 1
List of the parameters used in this study.
where e is the permittivity, q is the charge density. The charge den-
sity depends on the ion concentration according to Parameters Value References

q ¼ Fðcþ  c Þ ð21Þ Gas constant (R) 8.314 (JK1mol1) [20,46]


Environmental temperature 298 (K) [20,46]
(T)
Faraday constant (F) 96,480 (JK1mol1) [20,46]
2.1.4. Boundary conditions
Diffusion coefficient of sulfate 1.07  109 (m2/s) [6]
Considering the diffusion process as one-dimensional, the ions in water (D0)
boundary conditions can be expressed as: Initial concentration of Ca2+ 21.25 (mol/m3) [48]
(cini
Ca )
Initial concentration of CA 117.06 (mol/ m3) Obtained in
(cini
CA )
this
experimental
study
Porosity of ITZ (uITZ) 21.13% Obtained in
this
experimental
study
Porosity of paste (ub) 11.26% Obtained in
this
experimental
study
Chemical reaction rate of Eq. 3.05  108 [41]
(2) (k1) (m3mol1s1)
Chemical reaction rate of Eq. 1.22  109(m3mol1s1) [41]
(4) (k2)
Volume fraction of aggregate 55% Obtained in
(Va) this
experimental
study
Equivalent reacting coefficient 8/3 [20]
of gypsum dissipation to
produce ettringite (n)
Fig. 2. Schematic diagram of self-induced electrical field.
32 D. Sun et al. / Construction and Building Materials 192 (2018) 28–37

Fig. 3. 2D geometry model of random distribution of aggregate in three phase concrete.

angular mesh is used for paste modelling, while a finer size is


applied in ITZ area, as shown in Fig. 3b.

2.2. Experimental study

The purpose of this section is to conduct experiments to cali-


brate and validate the theoretical model developed in this study.

2.2.1. Materials
Portland cement (OPC 52.5R) was used in this study and its
chemical composition, as determined by X-ray fluorescence Fig. 4. Diagram of experimental setup.
(XRF), is given in Table 2. Concrete specimens were prepared with
a water/binder ratio of 0.45, the volume fraction of aggregate
0/5 mm, 5/8 mm, 8/10 mm are 30%, 5% and 20%, respectively. concentration. After being dried to a constant weight, 0.5 g of spec-
100  100  100 mm cubes were cast and all the samples were imen powder was dissolved in 100 mL of water with boric acid and
cured 24 h and then stored in concrete standard compartment hydrochloric acid in a beaker, heat the solution until boiling, and
with a constant temperature of 20 ± °C and a relatively humidity then add heated BaCl22H2O quickly into hot sample and filter
of 95 ± 5% for 56 days. the cooled solution. Filter paper and solid substance were put in
porcelain crucible, and were transferred at oven and ignited at
2.2.2. Experimental design 850 ± 20 °C. Finally, the only solid (BaSO4) was weighted. After
2.2.2.1. Sulfate ion precipitation. After being cured for 56 days, all obtaining the specimen powder, the drilling hole was carefully
specimens were sealed on 4 sides with only two opposite side left sealed. It can be seen in Fig. 5 that, after 12 weeks, the drilled sur-
open for one-dimensional diffusion (shown in Fig. 4). Specimens face of concrete is still intact and without cracks. As shown in pre-
were then immersed in 5 g/L and 50 g/L sodium sulfate solution, vious studies, a relatively homogeneous structure of the cement
respectively, 3 specimens were cast for each solution. The solution stone is obtained for specimen at the time of immersion of
was changed every month to ensure a stable erosion environment. 180 days [7]. Table 3 shows sulfate distribution in 5 g/L and 50 g/
Drilling method was used to obtain the specimen powder for L sodium sulfate solution, respectively.
analysis. Through barium sulphate gravimetric method (GB/T
22660.8–2008), the specimen powder at 0–5 mm in specimens 2.2.2.2. Quantitative analysis using backscattered electron microscopy
was obtained at a time interval of 4 weeks to analyse the sulfate (BSE). For microstructural analysis, a rectangular block sample was

Table 2
Chemical composition of Portland cement.

Chemical analysis (w/%)


SiO2 Al2O3 Fe2O3 CaO SO3 MgO Na2O K2O
20.00 4.26 3.29 65.1 2.92 0.675 0.0716 0.759
D. Sun et al. / Construction and Building Materials 192 (2018) 28–37 33

calibration process is to obtain one set of parameters which can


describe the penetration of sulfate ions in specimens exposed to
5 g/L and 50 g/L sodium-sulfate solution, respectively. The equiva-
lent concentration variation due to the variation of the position can
be expressed as,
Rm
 0
cSO2 dx
c¼ 4
ð24Þ
m

where m is the drilling depth (m = 5 mm). By using the least-


squares fitting, the values of a and q in Eq. (16) are determined as
0.1 and 0.9, respectively.
Fig. 7 shows the normalized time-dependent concentration of
sulfate ions in specimens exposed to 5 g/L and 50 g/L sodium-
sulfate solution, respectively. It shows that the normalized concen-
Fig. 5. Specimen under 50 g/L sulfate attack after 12 weeks. tration of sulfate ions in 50 g/L becomes much lower than that of
5 g/L with increase of time due to fact that the increase of the con-
Table 3 centration of sulfate ions in solution leads to the rapid accumula-
Sulfate content in concrete at depth of 5 mm (%). tion of gypsum and ettringite in pores, and therefore hinder the
0d 28 d 56 d 84 d
further transport of sulfate ions.
To investigate the influence of reaction and migration on the
5 g/L 0.168 0.179 0.184 0.185
50 g/L 0.168 0.326 0.327 0.329
transport of sulfate ions, Fig. 8 shows the normalized time-
dependent distribution of sulfate ions at d = 5 mm under diffusion,
taken from the surface of each specimen. Samples pre-dried to a diffusion + reaction, and diffusion + reaction + migration, respec-
constant weight were vacuum impregnated with low viscosity tively. It can be seen that chemical reaction has a great effect on
epoxy resin, and then cured for 9 h at room temperature. After that, the concentration of sulfate ions, whereas the self-induced electri-
the samples were ground firstly using SiC paper of 800 grit for cal field has limited influence in the transport of sulfate ions.
about 60 s, then 1200 grit for about 4 mins, and finally 2400 grit Fig. 9 shows the normalized steady-state distribution of ettrin-
for about 4 mins. The last step was to polish the samples using gite with the distance from the aggregate surface when specimens
the diamond paste of 9, 3, 0.25 lm sequentially for about 5 mins were immersed in 50 g/L Na2SO4 sodium-sulfate solution. It can be
each to remove surface scratches, and then cleaned up samples seen that the steady-state concentration of ettringite within ITZ is
with a low-relief polishing cloth. During observation process, an higher than that in bulk paste with a nearly seven-fold higher nor-
acceleration voltage of 20 kV is necessary. Around 35 images were malized concentration in inner ITZ (5 mm from aggregate surface)
collected to analyse porosity distribution of ITZ and paste by using than that in outer ITZ (50 mm from aggregate surface). The high
overflow method [33,49,50]. The average porosity as a function of accumulation of ettringite in ITZ could easily cause expansion
the distance away from the aggregate surface is given in Fig. 6b, the and further damage, and is the main reason for the cracking in
thickness of ITZ is about 50 mm and the results are consistent with ITZ when specimens immersed in sulfate solution.
other previous relevant experimental works [21,25,32,33]. Fig. 10a shows the normalized time-dependent ettringite con-
centration in specimens when exposed to different sulfate-sodium
concentrations (i.e. 5 g/L, 10 g/L, 30 g/L and 50 g/L). The results show
3. Results and discussion that, when exposed to a relatively low sulfate-sodium concentra-
tion (i.e. 5 g/L), it takes around 1800 days for ettringite concentra-
3.1. Model calibration and validation tion to reach the steady-state concentration, whereas the ten-fold
increase of sulfate-sodium concentration (i.e. 50 g/L) results in the
The parameters a and q in Eq. (16) could be calibrated by fitting significant decrease in the time for ettringite concentration to reach
the experimental data, as shown in Fig. 7. The purpose of this its steady-state (i.e. around 240 days). Fig. 10b shows normalized

Fig. 6. BSE image analysis: (a) Strips delineation, (b) Porosity distribution of ITZ in concrete after being cured 56d.
34 D. Sun et al. / Construction and Building Materials 192 (2018) 28–37

Fig. 7. Normalized time-dependent concentration of sulfate ions in specimens exposed to (a) 5 g/L sodium-sulfate solution, (b) 50 g/L sodium-sulfate solution.

depth dependent ettringite concentration in specimens exposed to


50 g/L sulfate-sodium concentration for 100 days, 300 days,
500 days and 1000 days, respectively. The gaps between the con-
centrations of sulfate-sodium represent the location of aggregates,
where the concentrations of ettringite are zero. The penetration
depth with maximum possible ettringite concentration (i.e. cEtt/c-
Ett-max = 1) at 100 days, 300 days, 500 days and 1000 days are
2 mm, 5.5 mm, 7.5 mm and 12 mm, respectively.
The concentration profiles of sulfate ions in concrete specimens
immersed in 50 g/L sodium sulfate solution for a period of 100 day
and 28 day are shown in Fig. 11, respectively. It can be seen
directly that the penetration depth of sulfate ions increases with
time. While dilution and tortuosity effect caused by spherical
aggregate change the transmission path of sulfate ions, their trans-
port could be enhanced in the microstructure of ITZ where the
porosity is relatively high. Once the ITZ is percolated due to the
high volume fraction (i.e. Va = 0.55), the formation of leaking paths
could enhance the transport of corrosive ions (shown in Fig. 3b).
Therefore, there is significant different transport behaviour in
Fig. 8. Normalized time-dependent distribution of sulfate ions under diffusion,
cement paste and ITZ.
diffusion + reaction, and diffusion + reaction + migration, respectively. Sodium-sul-
fate concentration is 50 g/L and d = 5 mm.
3.2. Parametric study

After validation, a series parametric studies were carried out to


investigate the effects of model parameters (i.e. a and q) on the
transport of sulfate ions in concrete. It can be seen from Fig. 12a
that when exposed to 5 g/L sodium-sulfate solution, a ten-fold
change in a has little impact in time-dependent concentration of
sulfate ions. In contrast, when exposed to 50 g/L sodium-sulfate
solution, a ten-fold increase of a could result in 50% decrease in
sulfate ion concentration at t = 100 days (Fig. 12b). Thus, a is a crit-
ical parameter which influences the value of Deff remarkably espe-
cially when concrete is exposed to high concentration of sulfate
ions. Similar to the parameter a, Fig. 13a shows that the maximum
filling rate q (0 < q < 1) has little influence in the concentration of
sulfate ions when sodium-sulfate concentration in solution is low
(i.e. 5 g/L), whilst under high sodium-sulfate concentration (i.e.
50 g/L), a five-fold increase of q could result in 30% decrease of
the concentration of sulfate ions.

3.3. Sensitivity analysis

Fig. 14 shows the influence of model parameters (i.e. ‘‘pore fill-


Fig. 9. Normalized steady-state distribution of ettringite in ITZ. ing factor” a, maximum filling rate q and electrolyte potential /) in
D. Sun et al. / Construction and Building Materials 192 (2018) 28–37 35

Fig. 10. (a) Normalized time-dependent ettringite concentration in specimens when exposed to different sulfate-sodium concentrations (i.e. 5 g/L, 10 g/L, 30 g/L and 50 g/L),
(b) Normalized depth dependent ettringite concentration in specimens exposed to 50 g/L sulfate-sodium concentration for 100 days, 300 days, 500 days and 1000 days,
respectively.

Fig. 11. Concentration distribution of sulfate ions at 100 day and 28 day, respectively.

Fig. 12. Parametric study of time-dependent distribution of sulfate ions (d = 5 mm). (a) 5 g/L sodium-sulfate solution, (b) 50 g/L sodium-sulfate solution.
36 D. Sun et al. / Construction and Building Materials 192 (2018) 28–37

Fig. 13. Parametric study of time-dependent distribution of sulfate ions (d = 5 mm). (a) 5 g/L sodium-sulfate solution, (b) 50 g/L sodium-sulfate solution.

Fig. 14. Sensitivity analysis of different model parameters (t = 56 days; d = 5 mm). The ‘‘Ratio” in abscissa axis represents the change in values of parameters (i.e. q, a, /)
according to their base values (i.e. q = 0.9, a = 0.1, / = 0.5), (a) 5 g/L sodium-sulfate solution, (b) 50 g/L sodium-sulfate solution.

concentration of sulfate ions in concrete at t = 56 days. The purpose time-dependent change of diffusion coefficient of sulfate ions
of the sensitivity analysis is to identify the dominant parameters due to the formation of ettringite and gypsum in concrete pores
which govern the corrosion of concrete, and provide guidance for resulting from the chemical reaction between sulfate ions and con-
future experimental studies in this area. The simulation results crete components, and the self-induced electrical field produced by
show that the model parameters exert their influence only when the ions. The following are the major findings:
sulfate ion concentration in solution is high (i.e. 50 g/L). It demon-
strates that there is a strong negative linear relationship between a  The concentration of sulfate ions initially increases rapidly in
and the change of concentration of ettringite in concrete, while the concrete during the first 20 days and then the rate of increase
parameter q is positively related to the change of concentration of gradually decreases due to the reduction in the diffusion coeffi-
ettringite. In addition, / which represents electrolyte potential, has cient of ions as a result of the formation of ettringite in concrete
little influence in the change of ettringite concentration in pores.
concrete.  The results of BSE quantitative analysis show that the porosity
of concrete within ITZ is relatively higher than that in bulk
paste. The porosity of inner ITZ (i.e. 5 mm from the aggregate
4. Conclusions surface) could be six-fold higher than that of outer ITZ (i.e.
50 mm from the aggregate surface).
The present study investigates the effect of ITZ on the transport  The steady-state concentration of ettringite within ITZ is higher
of sulfate ions in concrete by developing numerical model in than that in bulk paste with a nearly seven-fold higher concen-
conjunction with experiments. The model takes into account the tration in inner ITZ than that in outer ITZ.
spatially dependent distribution of pore size in ITZ, the
D. Sun et al. / Construction and Building Materials 192 (2018) 28–37 37

 Our parametric study and sensitivity analysis show that ‘‘pore [20] X.-B. Zuo, W. Sun, C. Yu, Numerical investigation on expansive volume strain in
filling” factor a is the most critical parameter representing the concrete subjected to sulfate attack, Constr. Build. Mater. 36 (2012) 404–410.
[21] K.L. Scrivener, A.K. Crumbie, P. Laugesen, The interfacial transition zone (ITZ)
formation of ettringite and gypsum in concrete. a is signifi- between cement paste and aggregate in concrete, Interface Sci. 12 (4) (2004)
cantly negatively correlated to the concentration of ettringite. 411–421.
 The self-induced electric field has little influence in the trans- [22] E.J. Garboczi, D.P. Bentz, Analytical formulas for interfacial transition zone
properties, Adv. Cem. Based Mater. 6 (3–4) (1997) 99–108.
port of ions at early stage of corrosion. [23] J. Ollivier, J. Maso, B. Bourdette, Interfacial transition zone in concrete, Adv.
Cem. Mater. 2 (1) (1995) 30–38.
Conflict of interest [24] W. Li, J. Xiao, Z. Sun, S. Kawashima, S.P. Shah, Interfacial transition zones in
recycled aggregate concrete with different mixing approaches, Constr. Build.
Mater. 35 (2012) 1045–1055.
We declare that no conflict of interest exits in the submission of [25] J. Xiao, W. Li, D.J. Corr, S.P. Shah, Effects of interfacial transition zones on the
this manuscript, and manuscript is approved by all authors for stress–strain behavior of modeled recycled aggregate concrete, Cem. Concr.
Res. 52 (2013) 82–99.
publication. [26] J. Xiao, W. Li, Z. Sun, D.A. Lange, S.P. Shah, Properties of interfacial transition
zones in recycled aggregate concrete tested by nanoindentation, Cem. Concr.
Acknowledgements Compos. 37 (2013) 276–292.
[27] B. Bourdette, E. Ringot, J. Ollivier, Modelling of the transition zone porosity,
Cem. Concr. Res. 25 (4) (1995) 741–751.
The authors would like to acknowledge the financial support [28] E. Irassar, V. Bonavetti, M. Gonzalez, Microstructural study of sulfate attack on
provided by the China Scholarship Council and National Natural ordinary and limestone Portland cements at ambient temperature, Cem.
Concr. Res. 33 (1) (2003) 31–41.
Science Foundation of China (Grant No. 51608382).
[29] A. Leemann, B. Münch, P. Gasser, L. Holzer, Influence of compaction on the
interfacial transition zone and the permeability of concrete, Cem. Concr. Res.
References 36 (8) (2006) 1425–1433.
[30] K.-Y. Liao, P.-K. Chang, Y.-N. Peng, C.-C. Yang, A study on characteristics of
[1] A. Neville, The confused world of sulfate attack on concrete, Cem. Concr. Res. interfacial transition zone in concrete, Cem. Concr. Res. 34 (6) (2004) 977–989.
34 (8) (2004) 1275–1296. [31] A. Asbridge, G. Chadbourn, C. Page, Effects of metakaolin and the interfacial
[2] P.K. Mehta, Mechanism of sulfate attack on portland cement concrete— transition zone on the diffusion of chloride ions through cement mortars, Cem.
Another look, Cem. Concr. Res. 13 (3) (1983) 401–406. Concr. Res. 31 (11) (2001) 1567–1572.
[3] N. Cefis, C. Comi, Chemo-mechanical modelling of the external sulfate attack in [32] J. Maso, Interfacial Transition Zone in Concrete, CRC Press, 2014.
concrete, Cem. Concr. Res. 93 (2017) 57–70. [33] K. Wu, Experimental Study on the Influence of ITZ on the Durability of
[4] M. Santhanam, M.D. Cohen, J. Olek, Mechanism of sulfate attack: a fresh look: Concrete Made with Different Kinds of Blended Materials, Ghent University,
part 1: summary of experimental results, Cem. Concr. Res. 32 (6) (2002) 915– 2014.
921. [34] Z. Liu, G. De Schutter, D. Deng, Z. Yu, Micro-analysis of the role of interfacial
[5] M. Santhanam, M.D. Cohen, J. Olek, Mechanism of sulfate attack: a fresh look: transition zone in ‘‘salt weathering” on concrete, Constr. Build. Mater. 24 (11)
Part 2. Proposed mechanisms, Cem. Concr. Res. 33 (3) (2003) 341–346. (2010) 2052–2059.
[6] P. Gospodinov, R. Kazandjiev, M. Mironova, The effect of sulfate ion diffusion [35] X. Li, Y. Xu, S. Chen, Computational homogenization of effective permeability
on the structure of cement stone, Cem. Concr. Compos. 18 (6) (1996) 401–407. in three-phase mesoscale concrete, Constr. Build. Mater. 121 (2016) 100–111.
[7] P.N. Gospodinov, R.F. Kazandjiev, T.A. Partalin, M.K. Mironova, Diffusion of [36] K. Wu, H. Shi, L. Xu, G. Ye, G. De Schutter, Microstructural characterization of
sulfate ions into cement stone regarding simultaneous chemical reactions and ITZ in blended cement concretes and its relation to transport properties, Cem.
resulting effects, Cem. Concr. Res. 29 (10) (1999) 1591–1596. Concr. Res. 79 (2016) 243–256.
[8] S. Lorente, M.-P. Yssorche-Cubaynes, J. Auger, Sulfate transfer through [37] W.-M. Ni, The mathematics of diffusion, Siam (2011).
concrete: migration and diffusion results, Cem. Concr. Compos. 33 (7) (2011) [38] G.E. Zaikov, A.L. Iordanskiĭ, V.S. Markin, Diffusion of electrolytes in polymers,
735–741. VSP1988.
[9] J.-K. Chen, C. Qian, H. Song, A new chemo-mechanical model of damage in [39] E.J. Garboczi, Permeability, diffusivity, and microstructural parameters: a
concrete under sulfate attack, Constr. Build. Mater. 115 (2016) 536–543. critical review, Cem. Concr. Res. 20 (4) (1990) 591–601.
[10] X.-B. Zuo, J.-L. Wang, W. Sun, H. Li, G.-J. Yin, Numerical investigation on [40] M.A.B. Promentilla, T. Sugiyama, T. Hitomi, N. Takeda, Quantification of
gypsum and ettringite formation in cement pastes subjected to sulfate attack, tortuosity in hardened cement pastes using synchrotron-based X-ray
Comput. Concr. 19 (1) (2017) 19–31. computed microtomography, Cem. Concr. Res. 39 (6) (2009) 548–557.
[11] R. Tixier, B. Mobasher, Modeling of damage in cement-based materials [41] C. Sun, J. Chen, J. Zhu, M. Zhang, J. Ye, A new diffusion model of sulfate ions in
subjected to external sulfate attack. I: formulation, J. Mater. Civ. Eng. 15 (4) concrete, Constr. Build. Mater. 39 (2013) 39–45.
(2003) 305–313. [42] E. Samson, J. Marchand, Modeling the transport of ions in unsaturated cement-
[12] R. Tixier, B. Mobasher, Modeling of damage in cement-based materials based materials, Comput. Struct. 85 (23–24) (2007) 1740–1756.
subjected to external sulfate attack. II: Comparison with experiments, J. [43] S. Caré, E. Hervé, Application of a n-phase model to the diffusion coefficient of
Mater. Civ. Eng. 15 (4) (2003) 314–322. chloride in mortar, Transp. Porous Media 56 (2) (2004) 119–135.
[13] A.E. Idiart, C.M. López, I. Carol, Chemo-mechanical analysis of concrete [44] B. Lu, S. Torquato, Nearest-surface distribution functions for polydispersed
cracking and degradation due to external sulfate attack: a meso-scale model, particle systems, Phys. Rev. A 45 (8) (1992) 5530.
Cem. Concr. Compos. 33 (3) (2011) 411–423. [45] J.J. Zheng, Z.Q. Guo, X.D. Pan, P. Stroeven, L.J. Sluys, ITZ volume fraction in
[14] Y. Xi, Z.P. Bažant, Modeling chloride penetration in saturated concrete, J. concrete with spheroidal aggregate particles and application: Part I. Numerical
Mater. Civ. Eng. 11 (1) (1999) 58–65. algorithm, Mag. Concr. Res. 63 (7) (2011) 473–482.
[15] W. Li, Z. Luo, Z. Sun, Y. Hu, W.H. Duan, Numerical modelling of plastic–damage [46] O. Truc, J.-P. Ollivier, L.-O. Nilsson, Numerical simulation of multi-species
response and crack propagation in RAC under uniaxial loading, Mag. Concr. diffusion, Mater. Struct. 33 (9) (2000) 566–573.
Res. 70 (9) (2017) 459–472. [47] E. Samson, J. Marchand, J.L. Robert, J.P. Bournazel, Modelling ion diffusion
[16] J. Xiao, W. Li, D.J. Corr, S.P. Shah, Simulation study on the stress distribution in mechanisms in porous media, Int. J. Numer. Meth. Eng. 46 (12) (1999) 2043–
modeled recycled aggregate concrete under uniaxial compression, J. Mater. 2060.
Civ. Eng. 25 (4) (2012) 504–518. [48] K. Nakarai, T. Ishida, K. Maekawa, Modeling of calcium leaching from cement
[17] C. Xiong, L. Jiang, Y. Zhang, H. Chu, Modeling of damage in cement paste hydrates coupled with micro-pore formation, J. Adv. Concr. Technol. 4 (3)
subject to external sulfate attack, Comput. Concr. 16 (6) (2015) 847–864. (2006) 395–407.
[18] S.L. Tracy, S.R. Boyd, J.D. Connolly, Effect of curing temperature and cement [49] Y. Gao, G. De Schutter, G. Ye, H. Huang, Z. Tan, K. Wu, Porosity characterization
chemistry on the potential for concrete expansion due to DEF, PCI J. 49 (1) of ITZ in cementitious composites: concentric expansion and overflow
(2004) 46–57. criterion, Constr. Build. Mater. 38 (2013) 1051–1057.
[19] G. Escadeillas, J.-E. Aubert, M. Segerer, W. Prince, Some factors affecting [50] J. Jiang, G. Sun, C. Wang, Numerical calculation on the porosity distribution and
delayed ettringite formation in heat-cured mortars, Cem. Concr. Res. 37 (10) diffusion coefficient of interfacial transition zone in cement-based composite
(2007) 1445–1452. materials, Constr. Build. Mater. 39 (2013) 134–138.

Вам также может понравиться