You are on page 1of 14

Journal of Building Engineering 45 (2022) 103619

Contents lists available at ScienceDirect

Journal of Building Engineering


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

Numerical study on effects of the degree of water saturation on


ICCP of chloride-contaminated RC structures
Bingbing Guo a, *, Guofu Qiao b, **, Peng Han b, Qiang Fu a
a
School of Civil Engineering / Key Lab of Engineering Structural Safety and Durability / State Key Laboratory of Green Building in Western China,
Xi’an University of Architecture and Technology, Xi’an, 710055, China
b
School of Civil Engineering, Harbin Institute of Technology, Harbin, 150090, China

A R T I C L E I N F O A B S T R A C T

Keywords: The effect of the degree of concrete water saturation on impressed current cathodic protection
Reinforced concrete structures (ICCP) of chloride-contaminated reinforced concrete (RC) structures was investigated using a
Steel corrosion numerical model that includes transport of oxygen and ions, electrode reactions, and steel po­
Cathodic protection larization. The results indicate that the local current density and potential at the steel surface are
Degree of water saturation significantly affected by the saturation degree. The risk of the hydrogen embrittlement of steel
Numerical model
may rise in a voltage-controlled ICCP when the concrete saturation degree increases, and the
applied voltage needs to be reduced in this case. For RC structures having a high saturation
degree, ICCP has a stronger ability to increase the pH around the steel reinforcement and to
remove Cl− . When the concrete saturation degree is low, the stationary numerical model based on
the Laplace equation is recommended to investigate or design ICCP if neglecting its effect on the
Cl− removal and pH increase. When the concrete saturation degree is high, the time-dependent
numerical model, including transport of oxygen and ions, electrode reactions, and steel polari­
zation, is recommended.

1. Introduction
Structures made of reinforced concrete (RC) are widely employed in industry and civil architecture as well as in road and bridge
engineering and for hydraulic structures. Unfortunately, corrosion of reinforcing steel may seriously impair the long-term durability of
RC structures, especially in coastal and de-icing environments [1,2]. This poses a huge threat to the rapid development of the economy
and to the safety of human life [3]. For instance, the collapse of the Morandi Bridge in Italy in 2018 caused the death of 43 individuals,
and this disaster was attributed to the poor maintenance of this bridge, which was vulnerable to chloride- and sulfate-induced
corrosion of the cables [4]. Thus, several corrosion prevention and control methods have been developed to improve the long-term
durability of RC structures, to prolong their service life, and to ensure their safety [5,6]. Among the available methods, cathodic
protection (CP) is considered to be a direct and effective way to prevent or arrest steel corrosion of both new and existing RC structures.
With this method, an external current is introduced into the reinforcing steel bars to force them to act as a cathode, which prevents the
dissolution reaction of iron in an electrochemical corrosion cell [7–11]. In CP, a distinction can be made between impressed current
cathodic protection (ICCP) and sacrificial anode cathodic protection (SACP) [12]. The former depends on an external current source

* Corresponding author. 13 Yanta Road, Xi’an, 710055, PR China.


** Corresponding author.
E-mail addresses: guobingbing212@xauat.com (B. Guo), qgf_forever@hit.edu.cn, qgf_forever@hit.edu.cn (G. Qiao), 19b333004@stu.hit.edu.cn (P. Han),
fuqiangcsu@163.com (Q. Fu).

https://doi.org/10.1016/j.jobe.2021.103619
Received 12 July 2021; Received in revised form 24 October 2021; Accepted 8 November 2021
Available online 12 November 2021
2352-7102/© 2021 Elsevier Ltd. All rights reserved.
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

using essentially inert anodes, whereas the latter is based on the natural difference in the electrochemical potential between the
applied galvanic anode material and the embedded reinforcing steel [13]. In general, ICCP is more suitable for RC structures with
serious corrosion and long-life requirements because it is more flexible and reliable than SACP [14].
Numerical simulation has always been an economical and convenient tool to investigate CP since it can predict the protection
current density distribution and electrical potential at the steel surface. In the past decades, it has been successfully used to study the
influence of various factors on CP [7,15–21], such as the position of anode, the moisture content in concrete, the concrete resistivity,
the boundary conditions, etc. In these investigations, the numerical model based on the Laplace equations has been widely used as
follows:

∇2 φ = 0 (1)
Additionally,

i = − σ ∇φ (2)
⎧ a,c
⎨ φ = φ0 a,c
ia,c = i0 a,c (3)
⎩ a,c
i = f (φa,c )

The domain is governed by Equations (1) and (2), in which i , φ, and σ are the current density, the potential, and the concrete
conductivity, respectively. The boundary conditions at the surfaces of the anode and cathode are governed by Eq. (3), in which φa,c and
ia,c are the potential and current density at the surfaces of anode and cathode, respectively; φa,c a,c
0 and i0 are the initial potential and
current density, respectively.
The above equations can also be used to design CP [15,22]. This numerical modelling pertains to steady-state conditions and is
therefore independent of time. However, in addition to the negative potential shift of the reinforcing steel (polarization), in RC
structures also electrode reactions and transport of negatively and positively charged ions occur. Generally, this will gradually increase
the pH value in the vicinity of the steel bar and mitigate or prevent further chloride penetration or even remove the chloride ions from
the concrete. These effects are additional benefits of ICCP applied to RC structures, which has already been confirmed by several
experimental results [9,23,24]. Unfortunately, these beneficial effects have not been taken into account in the numerical models based
on the Laplace equation. Generally, for RC structures, numerical models of CP that consider these beneficial effects have been seldom
reported. In the study of Polder et al. [11], a numerical model including oxygen reduction was developed, and the results indicated that
the pH increase at the steel surface caused by CP is more significant in the early stage. Liu et al. [8] used a numerical model based on
the Nernst-Planck equation to investigate CP, and their numerical results also indicated that CP can lead to an increase of the pH value
in the vicinity of the steel, and can slow down chloride attack.
However, concrete consists of solid, liquid and gas phases, and the majority of concrete structures in practical engineering are not
completely saturated with water. In general, the degree of concrete water saturation (i.e., the ratio of pore water content to porosity) is
closely related to the environmental relative humidity (RH). When environmental RH is low/high for a long time, it may decrease/
increase the degree of concrete water saturation. The transport of ions in the liquid phase contained in the pore system determines
concrete conductivity. Thus, on the macro level, the degree of water saturation will directly affect concrete conductivity, and further
determines the magnitude of the current delivered to RC structures (including the current density between rebars) in a voltage-
controlled CP system or the output voltage in a current-controlled CP system. Moreover, the oxygen concentration in the vicinity of
the steel has a significant influence on the steel polarization and on the rate of electrode reactions. Oxygen transport in concrete
primarily occurs in the gas phase, and the gas and liquid phases are complementary to each other in the entire pore structure. Thus, the
degree of water saturation also influences the rate of oxygen transport in the concrete, and consequently, affects the steel polarization
and the rate and type of electrode reactions. Muehlenkamp et al. [21] investigated the effect of the saturation degree on the current
density and potential at the steel surface using a numerical model based on the Laplace equation, in which oxygen transport was
considered. The same numerical model was developed as one of the application libraries about ICCP of RC structures in COMSOL
Multiphysics [25]. However, beneficial effects of ICCP on Cl− migration and pH were not considered in their numerical models.
Therefore, in this study, the effect of the concrete saturation degree on ICCP was investigated using a numerical model including
transport of oxygen and ions, electrode reactions, and steel polarization. It should be noted that moisture transport in the concrete may
have an effect on CP, and this will not be discussed here because of the length and scope of this paper.

2. Theoretical equations
2.1. Oxygen transport in concrete
Oxygen plays a vital role in the electrode reactions at the steel surface when ICCP is applied to RC structures. Oxygen from the
surrounding atmosphere enters the concrete, passes through the cover concrete, and thereafter participates in the cathodic electrode
reactions. In this study, Fick’s second law was used to describe oxygen transport in concrete:
∂cO2
= ∇ ·(DO2 ∇cO2 ) (4)
∂t

where cO2 (mol/m3) is the oxygen concentration; DO2 (m2/s) is the diffusion coefficient of oxygen in concrete, which is strongly

2
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

dependent on the saturation degree apart from the material properties of concrete (i.e., pore size distribution, capillary effect, etc.).
Theoretically, DO2 can be determined using the following equation [26–28]:
k · krg ρg
DO2 = (5)
μC

where ρ (kg/m3) and μ (N⋅s/m2) are the density and dynamic viscosity of oxygen, respectively; k (m2) is the intrinsic permeability,
which is the material attribute of concrete; krg (− ) is the relative gas permeability, which depends on the concrete saturation degree; C
(1/m) is the capillary capacity function, which is the derivative of the saturation degree to the capillary pressure, and it is also a
function of the saturation degree [28]. Therefore, the oxygen diffusion coefficientDO2 , of a given concrete is dependent on its saturation
degree.

2.2. Ion transport in concrete structures with ICCP


The primary free ions in the concrete pore solution are Na+, K+, Ca2+, OH− , and SO2− 4 , and when RC structures are subjected to
chloride attack, also Cl− . Among them, the concentrations of Ca2+ and SO2−
4 are very low, and they can be discarded for simulating ion
transport in concrete. The application of ICCP causes the migration of these free (dissolved) ions in the concrete. Cations (i.e., Na+ and
K+) will move towards the steel bar (the cathode) while anions (i.e., OH− and Cl− ) will move in the direction of the anode. Thus, ICCP
can retard further chloride penetration or remove chloride that is present in concrete. In general, ion transport is jointly affected by
convection, and diffusion and migration. Water transport will not be considered in this study, which means convection is neglected,
and consequently transport of ions in concrete structures provided with ICCP can be described using the following equations:
∂cm
= − ∇ · Nm (6)
∂t
( )
zm c m F
Nm = Dm − ∇cm − ∇φ (7)
RT

where cm (mol/m3), Nm (mol/(m2⋅s)), Dm (m2/s) and zm (− ) are the concentration, flux, diffusion coefficient, and electric charge
number of the m-th free ion in the concrete, respectively; φ (V) is the electric potential; F (9.6485 × 104 C/mol), R (8.314 J/(mol⋅K)),
and T (K) denote the Faraday constant, universal gas constant, and absolute temperature, respectively. Dm mainly depends on the
tortuosity and constrictivity of the concrete pore structure, saturation degree, temperature, and diffusion coefficient of ions in the free
water [29,30]. This numerical study was conducted for a constant temperature, and consequently the effect of temperature was not
considered. It is assumed that
DCl (θ = 1)
Dm = Dm free · · f (θ) (8)
DCl_free

where Dm_free is the diffusion coefficient of ion i in an aqueous solution, the values of which are given in Ref. [31]. DCl (θ = 1) is the
chloride diffusion coefficient of the completely water saturated concrete, which can be easily measured using electrically accelerated
test methods [32]; thus, the ratioDDClCl_free
(θ=1)
can be considered to reflect the combined effect of tortuosity and constrictivity of the pore
structure. f(θ) accounts for the effect of the variation of the saturation degree (θ), which can be obtained from the following empirical
relationship [29,30]:

f (θ) = θ7/3 (9)


In addition, chloride can be chemically and physically bound by cement hydrates when penetrating into concrete from the external
environment. This significantly affects chloride transport in the concrete. In general, four types of chloride binding isotherms can
describe the chloride binding behaviors in concrete, i.e., linear, Langmuir, Freundlich, and BET binding isotherms [33]. Among them,
the Langmuir isotherm is commonly used in numerical modelling, and is expressed as follows:
αcCl
sCl = (10)
1 + βcCl

where sCl (kg/m3 of concrete) is the concentration of the bound chloride; α and β are empirical constants.
The electric current that CP supplies to the concrete is equal to the electrical charge that is transported per unit of time by free ions
dissolved in the concrete pore solution, and thus,
∑ n ∑n ( )
zm c m F
i=F zm Nm = F zm Dm − ∇cm − ∇ψ (11)
m=1 m=1
RT

where i is the current density in concrete (A/m2). According to the continuity equation, as to maintain electroneutrality, the current
density i should satisfy the following equation:
∇i = 0 (12)

3
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

Equations (3)–(12) are the basic equations used for simulating ion transport and spatial distribution of an electric field in concrete
structures with ICCP. To solve these equations, the relevant boundary conditions need to be defined.

2.3. Electrode reactions at the steel surface


Theoretically, when ICCP is applied, the electrode reactions at the cathodically polarized steel surface should be that of oxygen
reduction (O2 + 2H2O + 4e− →4OH− ), which will result into an increase of the pH in the vicinity of the reinforcing steel (the cathode).
However, when the protection current density supplied to the reinforcing steel is quite small, the steel is not completely protected,
which means that also to some limited extent iron oxidation (Fe→Fe2++ 2e− ) may occur. Normally, with ICCP the steel will not be
polarized to potential where hydrogen evolution is possible. However, when some conditions change, such as cracking of concrete
cover, and the increase in the degree of concrete water saturation, it may result in situations where the protection current density that
the steel receives becomes very high particularly for a voltage-controlled ICCP system. This will then possibly cause hydrogen evo­
lution (2H2O + 2e− →2OH− + H2), and consequently increase the risk of hydrogen embrittlement of the steel, especially in prestressed
concrete structures. Thus, three types of electrode reactions, i.e., oxygen reduction, iron oxidation and hydrogen evolution, may
simultaneously occur during the ICCP operation. The local current density of each electrode reaction can be obtained using the Tafel
equation [21]:
( )
ctO − 2.3(E − EO2 )
iO2 = 0 2 i0_O2 exp , (13)
c O2 b O2
( )
2.3(E − EFe )
iFe = i0_Fe exp , and (14)
bFe
( )
− 2.3(E − EH2 )
iH2 = i0_H2 exp . (15)
bH2
ct
Here, the ratio cO0 2 reflects the effect of the oxygen concentration on electrode kinetics; c0O2 is the initial oxygen concentration at the
O2

steel surface; ctO2 is the oxygen concentration at the steel surface at time t, which is determined by Equations (4) and (5); E is the
potential difference between concrete electrolyte and steel electrode over the steel-concrete interface. iO2, i0_O2, EO2, and bO2 are the
current density, exchange current density, equilibrium potential, and Tafel slope, respectively, for the oxygen reduction reaction; iFe,
i0_Fe, EFe, and bFe are the current density, exchange current density, equilibrium potential, and Tafel slope, respectively, for the iron
oxidation reaction; iH2, i0_H2, EH2, and bH2 are the current density, exchange current density, equilibrium potential, and Tafel slope,
respectively, for the hydrogen evolution reaction. The values of these parameters that are used in the numerical simulations are given
in Section 3.
The overall current (I) at the steel-concrete interface is the sum of these three local currents:
∫ ∫ ∫
I = iFe + iO2 + iH2 (16)
s s s

where s is the surface area of the reinforcing steel.


At the anode-concrete interface, the OH− oxidation (4OH− - 4e− →H2O + O2) is dominant, and the oxygen produced is assumed to
disappear into the surrounding atmosphere. There is also a possibility for chlorine evolution to occur, especially when ICCP with a
strong external electrical field is applied to control corrosion of RC structures exposed to a chloride-laden environment. Moreover, the
type of anode can affect the nature of electrode reactions. For example, mixed metal oxide-coated titanium anodes are capable of
facilitating only oxygen evolution reaction. These two electrode reactions can lower the pH value at the anode-concrete interface [34].
In general, depending on the anode material, the potential difference across the anode-concrete interface may be 0.2–0.5 V. The
polarization will be more significant when applying higher current densities. Considering that the products of anodic reaction does not
participate in the ion transport in concrete, the nonlinear polarization behavior at the anode-concrete interface is not taken into ac­
count in this study [17,21].

3. Benchmark example
Muehlenkamp et al. [21] and the application libraries of COMSOL [25] used the numerical model based on Laplace equations
(including Equations (1), (2), (4) and (13)~(15)) to study the effect of moisture on CP of a bridge beam. In their investigations, a
thermally sprayed zinc anode was used [21,25]. In practical engineering, some inert conductive materials, such as activated titanium,
conductive organic coatings, conductive cement-based materials, etc., are generally used as impressed current anode while thermally
sprayed zinc is used as a sacrificial (galvanic) anode. When zinc is used as the anode of ICCP, it cannot play the role of the anode of
SACP due to the intervention of external power supply, and works only as a conductive anode. In this case, the anodic reaction is
Zn→Zn2++2e− , which differs from other common inert impressed current anodes. In this study, to be in line with Muehlenkamp et al.
[21], the same CP example was used to investigate the effect of the degree of water saturation on the electrode reactions, and ability of
ICCP to remove Cl− and to increase pH. In their examples, the size of the specimen was 127 × 63.5 mm2, and two steel bars were
embedded. One quarter of this RC specimen was used to build the geometrical model because its geometry is completely symmetric in

4
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

nature. The geometrical model and its finite element mesh are shown in Fig. 1. The diameter of the embedded steel bar was 12.7 mm,
and the thickness of cover was 25.4 mm. The anode was applied on the left side of the concrete specimen. According to Refs. [21,25],
except for the anodic and cathodic boundaries, the other boundaries were assumed to be insulated. More details about this model can
be found in Ref. [21] or in the application libraries of COMSOL [25].
The values of some parameters from Ref. [21], as listed in Table 1, were used in this study. The initial K+ concentration was 71.79
mol/m3 of pore solution, and the initial pH was 13.19. The initial Na+ and Cl− concentrations were assumed to have a negative linear
relationship with their penetration depth. Their values at the concrete surface (x = 0) are 500 mol/m3 and 584 mol/m3 of pore so­
lution, respectively. Their values at the center of the steel (x = 31.75 mm) are 342 mol/m3 and 422 mol/m3 of pore solution,
respectively. Thus, the initial average ratio of Cl− to OH− at the steel-concrete interface was 2.2, which means that according to
Ref. [36] the steel was actively corroding. The value of DCl (θ = 1) was 9.2 × 10− 12 m2/s [37]. The values of α and β in Eq. (9) [38] are
given in Table 1. Although these values on the ionic concentration and the chloride diffusion coefficient were not given in Ref. [21],
they are considered reasonable because those chosen values are based on the experimental results in Refs. [37,39]. According to
Ref. [21], an external driving voltage of 1 V was applied between the external anode and the steel in the ICCP of this RC structure,
which includes the steel-concrete interface.
The oxygen concentration at the anode-concrete interface is equal to that of the current atmosphere, the value of which is 8.6 mol/
m3. When the reinforcing steel suffers from serious corrosion, there will be a gradient of oxygen concentration in the concrete due to
the oxygen consumption at the steel surface. However, this phenomenon does not exist when the steel has just become depassivated or
is still in a passive condition. Thus, in this numerical model, the initial oxygen concentration in the entire concrete cross-section is
assumed to be equal to that present in the atmosphere. According to Section 2.1, the oxygen diffusion coefficient of concrete depends
on the degree of water saturation, and the values at different saturation degrees are given in Ref. [21], as depicted in Fig. 2. In this
study, a quadratic polynomial was used to fit these data, and the resulting regression relationship showed a high correlation coefficient
(R2 = 0.998). The fitted quadratic polynomial as shown in Eq. (17) was used to express the relationship between the oxygen diffusion
coefficient and the saturation degree.
( )
DO2 = 33.023θ2 − 57.596θ + 25.630 × 10− 9 (17)

In order to investigate how the degree of water saturation affects ICCP, it was assumed that the saturation degree remains constant
over time throughout the entire cross section.

4. Results and discussion


4.1. Oxygen concentration, local current densities and potential at the steel surface under different saturation degrees
Oxygen is one of the main reactants for cathodic reactions at the steel surface. Thus, the oxygen concentration in the vicinity of the
steel directly affects the rate of the electrode reactions. Fig. 3 illustrates the predicted average oxygen concentration at the steel surface
under different saturation degrees for ICCP with applying a driving voltage of 1 V. When the saturation degree is less than 0.4, the
average oxygen concentration is almost equal to its initial value, i.e., the oxygen concentration in the surrounding atmosphere. This
indicates that oxygen transport can completely compensate for its consumption during electrode reactions, and oxygen transport is
unlikely to be a limiting factor. When the saturation degree ranges from 0.4 to 0.7, the average oxygen concentration at the steel
surface decreases, and then it becomes steady with the operation of ICCP. However, when the saturation degree is more than 0.7, the
oxygen transport cannot satisfy its consumption during electrode reactions, and according to the numerical calculations, this will lead
to an oxygen concentration of approximately zero. Therefore, three situations regarding the oxygen concentration at the steel surface
can be distinguished: (1) when the saturation degree is low, the supply easily compensates for the consumption and the oxygen
concentration at the steel surface is equal to that present in the surrounding atmosphere; (2) when the saturation degree is moderate,
the supply is approximately equal to the consumption, and the oxygen concentration is lower than that present in the atmosphere; (3)
when the saturation degree is high, the supply is less than the consumption, and the oxygen concentration is approximately zero. In
practice, there may be one or two situations among them. For instance, a significant amount of the external oxygen can easily penetrate
to the steel surface when the cover thickness is small or gas permeability of concrete is high. Then, the supply of oxygen can easily
compensate for its consumption, and the oxygen concentration at the steel surface will be equal to that present in the surrounding
atmosphere.

Fig. 1. Geometrical model and finite element mesh.

5
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

Table 1
Parameter values of the numerical model.

Symbol Unit Value


2 5
i0_Fe A/m 7.1 × 10−
i0_O2 A/m2 7.7 × 10− 5

i0_H2 A/m2 1.1 × 10− 2

EFe V vs.CSE -0.76


EO2 V vs.CSE 0.189
EH2 V vs.CSE -1.03
bFe V/decade 0.41
bO2 V/decade 0.18
bH2 V/decade 0.15
А / 0.98
В / 0.29

Note that the values related to electrode reactions are from Ref. [21]. α and β are the parameters related to chloride binding, and their values are from Ref. [38].

Fig. 2. Relationship between the oxygen diffusion coefficient and saturation degree used in the numerical model.

Fig. 3. Average oxygen concentration at the steel surface.

Each local current density reflects the corresponding electrode reaction at the steel surface under ICCP. The larger the value is, the
more obvious this electrode reaction is. Thus, we can infer the corrosion control effect from the numerical results regarding the local
current densities. Fig. 4 illustrates the average local current densities at the steel surface under different saturation degrees. The local
current density of iron oxidation in Fig. 4(a) is also called the corrosion current density. Apparently, the steel is still corroding at a very
low and negligible rate when applying CP. It can be seen from Fig. 4(a) that the steel has a tendency to corrode with a decrease in the

6
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

Fig. 4. Average local current densities at the steel surface.

saturation degree, which agrees with the findings in Refs. [21,25]. This also means that the increase in the saturation degree con­
tributes to steel polarization. Fig. 4(b) and (c) depict the average local current densities caused by oxygen reduction and hydrogen
evolution, respectively. When the saturation degree is less than 0.7, the oxygen reaction is completely dominant with the ICCP applied
in 180d. However, when the saturation degree is more than 0.7, the average local current density of the oxygen reaction decreases with
the operation of ICCP, and that of the hydrogen evolution gradually emerges. This indicates that the risk probability of the hydrogen
embrittlement of steel becomes greater.
In general, when the corrosion current density at the steel surface is less than 2 mA/m2, the steel is assumed to be passive. Hence, it
can be inferred from the results in Fig. 4 that the applied 1-V ICCP can completely suppress the corrosion process in this specimen
regardless of the concrete saturation degree. However, the risk of hydrogen embrittlement of steel may increase when the saturation
degree significantly increases, and the risk will become higher with the ICCP operation. Thus, the output voltage of ICCP should be
reduced when the degree of water saturation of the concrete shows a significant increase.
EN ISO 12696 suggests that the designed current density in CP is 2–20 mA/m2 for corroded RC structures, and it is 0.2–2 mA/m2 for

7
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

new RC structures (which is also called cathodic prevention) [35]. The suggested value is significantly lower than the numerical results
in Fig. 4, particularly when the degree of water saturation is higher than 0.4. The numerical model adopts some ideal conditions, and
neglects the effect of anodic polarization, which may produce a certain gap. In addition, applying the current density above 20 mA/m2
greatly increases the possibility of hydrogen evolution, which is not allowed in practice. Thus, according to the numerical results
presented in this study, the applied external electric field in ICCP should be reduced when the degree of water saturation increases.
The potential at the steel surface is usually measured to evaluate the corrosion control effect of CP in practical engineering, and
thus, the numerical model predicts the potential at the steel surface when applying ICCP for 180 days under different saturation
degrees, as illustrated in Fig. 5. It is found that the potential at the steel surface becomes more negative with increasing the degree of
water saturation. This again indicates that the steel is more easily polarized when the saturation degree of concrete is high especially
for a voltage-controlled ICCP system. It also means that the increase in the saturation degree may cause that in a voltage-controlled
ICCP system the possibility of hydrogen embrittlement of steel rises.
It is noted that the above-mentioned results are based on a voltage-controlled ICCP. For a current-controlled ICCP system, it can be
inferred that an increase in the degree of water saturation will cause a decrease in the system resistance, in particular that of the
concrete, and consequently result in a decrease of the required output voltage of the ICCP.

Fig. 5. The predicted potential of the steel surface after applying ICCP for 180 days under different saturation degrees (unit: V vs CSE): (a), (b), (c), (d), (e), (f) and (g)
denote the potential when the saturation degree is 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, and 0.8, respectively.

8
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

4.2. pH and Cl− concentration at the steel surface under different saturation degrees when applying ICCP
Oxygen reduction produces OH− , and this results into an increase of the pH in the vicinity of the steel. Fig. 6 illustrates the
development over time of the calculated average pH at the steel-concrete interface. The pH at the steel surface is determined by two
aspects: (1) the OH− amount produced by electrode reaction; (2) the OH− migration toward the anode. It is clear that the pH gradually
increases with the operation time of ICCP. This increasing trend is less significant with the decrease in the saturation degree especially
when the operation of ICCP is less than 60d. When the operation of ICCP is over 60d, this increasing trend gradually disappears, and
even there is a decreasing trend in the cases of the saturation degree of 0.6–0.7. This decreasing trend is attributed to that the effect of
the OH− migration on the pH is greater than that of electrode reactions. Thus, in view of the ability of ICCP to increase the pH, there
would be no urgent need to apply ICCP for a long time.
In this study, a specimen simulating a chloride-contaminated RC structure was used to investigate the ability of ICCP to remove Cl−
at the steel surface under different saturation degrees. The initial Cl− concentration at the steel surface is 422 mol/m3 of pore solution,
which has been specified in Section 3. The development over time of the distribution of the concentration of free (dissolved) Cl−
around the steel resulting from ICCP operation is illustrated in Fig. 7. It can be clearly seen that it is difficult for ICCP to remove the Cl−
at the steel surface when the degree of water saturation is quite low, and the increase in saturation degree contributes to remove Cl− for
ICCP. At the macro level, an increase of the saturation degree will result into an increase of the electrical current delivered to RC
structures for a voltage-controlled ICCP system because the concrete resistivity decreases. However, when the degree of water satu­
ration is very high, such as over 0.7 in this example, oxygen transport cannot meet the consumption requirement of the electrode
reaction over the ICCP operation. This leads to a rise in the polarization resistance across the steel-concrete interface, meaning that the
proportion of the potential difference across the interface in the entire system has an increase. Consequently, the potential difference
across concrete that is to remove Cl− decreases. This is the reason why the ability of ICCP to remove Cl− at the saturation degree over
0.7 is lower than other some situations in Fig. 7.
In the study of Christodoulou et al. [9], ICCP had been applied to a RC structure for 5 years, and after that they found that the steel
could remain a passive state although the RC structure was still subjected to chloride attack. Therefore, considering some potential
negative effects of a long-term ICCP, such as accelerating the ageing of external anodes [34] and decreasing the steel-concrete
interfacial bond strength [6], a highly efficient and long-life ICCP should be in real-time adjustments. For instance, ICCP can be
interrupted when the steel is re-passive due to the Cl− removal and pH increase; the magnitude of the applied current density should be
appropriately adjusted when the concrete saturation degree changes.

4.3. Effect of saturation degree on the current density distribution


The introduction of a fixed voltage between the external anode and the embedded steel bar resulting from the application of CP
causes a directional migration of both positively and negatively charged ions in the concrete pore solution, which produces an
electrical current in the concrete. This current is equal to that CP supplies to RC structure. The saturation degree determines the
available amount of electrolyte and the resistance against ionic migration, and thus it affects the delivered current for a voltage-
controlled ICCP system. The numerical model predicts the distribution of the current density under different saturation degrees,
and Fig. 8 illustrates the numerical results obtained at 30d and 180d after activation of ICCP. It is clearly shown that the electric current
flows to the steel from the anode, and the surface of the steel facing the anode receives more protection current than that the opposite
surface does, as reported by Hassanein et al. [17].
It can be observed from the results presented in Fig. 8 that the difference between the distributions of the current densities at 30d
and 180d becomes smaller with decreasing the saturation degree. In particular, when the saturation degree is 0.2, the numerical results
at 30d and 180d are approximately the same. This indicates that the numerical results are almost independent of time although this
numerical model is time-dependent. This can be also seen from the results of the oxygen concentration (Fig. 3) and the local current

Fig. 6. Average pH at the steel surface for different saturation degrees.

9
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

Fig. 7. Average Cl− concentration at the steel surface for different saturation degrees.

densities (Fig. 4) in the case that the saturation degree is less than 0.5. The application libraries of COMSOL Multiphysics [25] use the
stationary numerical model based on the Laplace equation of charge transport and the diffusion equation for oxygen transport to
simulate this ICCP. The numerical results1 concerning the current density are also approximately the same as those of this study when
the saturation degree is lower. This indicates that both the stationary and time-dependent numerical models can simulate ICCP of an
RC structure well in this case. The time-dependent numerical model in this study is based on the transport of oxygen and ions, electrode
reactions and steel polarization, involving multiple physical fields. The numerical calculation process is more complex, and requires a
lot of computing resources. But the numerical model based on the Laplace equation is more simple and easier to be calculated. Thus,
the stationary numerical model based on the Laplace equation is recommended to investigate ICCP of RC structures with a low
saturation degree if neglecting the beneficial effects of ICCP on the Cl− removal and pH increase. For instance, the stationary numerical
model is recommended in this study when the degree of water saturation is below 0.3.
When the saturation degree is higher, the steel will receive a higher polarization current, which means that a larger amount of
oxygen will be consumed, but the amount of oxygen that is transferred through the concrete cover to the steel surface from the external
environment is limited, and cannot meet the consumption rate. As a consequence, the steel potential will shift to more negative values,
and cathodic reactions involving hydrogen evolution will gradually develop. The electric current that a voltage-controlled CP supplies
to the RC structures shows a significant increase over time, and the numerical results are also dependent on time. Therefore, for a RC
structure having a high saturation degree, the application of a time-dependent numerical model, including transport of oxygen and
ions in concrete, electrode reactions, and steel polarization, is recommended.

4.4. Verification of the numerical model


On the macro level, the concrete resistivity affects the spatial distribution of electrical field induced by ICCP, and determines the
throwing power of ICCP. It reflects the material property, and is also an essential input parameter for the numerical model based on the
Laplace equation. In fact, the directional movement of charged ions caused by the gradient of the electrical potential produces an
electrical current, and hence, concrete resistivity is a variable in the numerical model based on the Nernst-Planck equation, rather than
an input parameter, which can be obtained by an inverse calculation of Eq. (2), i.e.,
∇φ
σ(x, y, t) = − → (18)
i

here φ and i are the electrical potential and the current density, respectively, and they are the direct variables of the numerical
modelling which are related to the time (t), and the location (x, y). Thus, the resistivity is related to the time (t), and the location (x, y).
In theory, the resistivity of concrete can vary over time because the electrical potential and the current density are dependent on the
ICCP operation time. The average concrete resistivity at the entire cross section as a function of time was calculated, as illustrated in
Fig. 9. It is clear that the degree of water saturation significantly affects the concrete resistivity. The concrete resistivity under the
saturation degree of 0.8 is two orders of magnitude smaller than the saturation degree of 0.2. During the early operation of ICCP, the
concrete resistivity demonstrates a slight decrease only when the degree of water saturation is lower than 0.4. With the operation of
ICCP, the resistivity hardly varies. This result indicates that the ICCP operation has little influence on the concrete resistivity.
To verify the numerical model, the concrete resistivity before applying ICCP under different saturation degrees was inversely
calculated, which is listed in Table 2. The calculation results regarding the concrete resistivity depends on the initial input parameters.
However, some initial parameters, such as the ionic concentration and chloride diffusion coefficient, were not given in the study of

1
The numerical results can be seen in the application libraries of COMSOL Multiphysics, which would not be presented in this paper.

10
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

Fig. 8. Current densities in RC structure with different saturation degrees at 30 d and 180 d (unit: A/m2). In this figure, 0.8, 0.7, 0.6, 0.5, 0.4, 0.3 and 0.2 are the
saturation degrees of concrete. For example, Fig. 8(a) is the distribution of the current density at 30 days when the saturation degree is 0.8.

11
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

Fig. 8. (continued).

Muehlenkamp et al. [21]. Their values are determined based on another some experimental results in Refs. [37,39], which may not
perfectly match the concrete in this study. When the saturation degree is lower, the results regarding the concrete resistivity is more
sensitive to the initial parameters, and the error between the numerical and experimental results is enlarged. With increasing the
saturation degree, there is an increasingly good agreement between the numerical and experimental results. On the whole, the results
predicted by the numerical model is in line with that in the literature [21].

5. Conclusions
In this study, the effects of the concrete saturation degree on the oxygen concentration, local current density, and potential at the
steel surface, on the ability of ICCP to increase pH and to remove Cl− , and on the electric current density in concrete were investigated
using a numerical model. The numerical model was validated against the experimental results from the literature with regard to the
concrete resistivity. The following conclusions were drawn:
The local current density and potential at the steel surface at a fixed voltage are significantly affected by the degree of water
saturation of concrete. The risk of the hydrogen embrittlement of steel may rise if the saturation degree has an increase, and the applied
current density should be reduced in this case. For RC structures with a high saturation degree, ICCP has a stronger ability to increase
pH and to remove Cl− . A highly efficient and long-life ICCP needs to work in real-time adjustments, and for instance, ICCP can be

12
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

Fig. 9. Concrete average resistivity at the entire cross section.

Table 2
Comparison of concrete resistivity before applying ICCP between experimental and numerical results under different saturation degree (Ω⋅m).

Saturation degree 0.2 0.3 0.4 0.5 0.6 0.7 0.8

Experimental results 5727 1227 500 205 142 102 64


Numerical results 3594 1540 568 262 149 94 65

Note that the experimental results are from Ref. [21].

interrupted once when the steel is re-passive.


When the concrete saturation degree is low, a stationary numerical model based on the Laplace equation is recommended to
investigate or design the ICCP if neglecting its effect on the Cl− removal and pH increase. When the concrete saturation degree is high, a
time-dependent numerical model, including the oxygen and ion transports, electrode reactions, and steel polarization, is
recommended.

Credit author statement


Bingbing Guo: Numerical modelling, Analysis, Writing-original draft preparation, Funding acquisition. Guofu Qiao: Conceptu­
alization, Methodology, Supervision, Project administration, Funding acquisition. Peng Han: Conceptualization, Writing-Review and
Editing. Qiang Fu: Methodology, Conceptualization, Writing-Review and Editing.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgements
The authors are grateful for the financial support of the Nature Science Foundation of China (NSFC) (Project No.: 51908453 and
51578190), the China Postdoctoral Science Foundation (Project No.: 2020M673607XB and 2020T130497), the Scientific Research of
Shanxi Provincial Department of Education (Project No.: 20JK0710), the Independent Research and Development project of State Key
Laboratory of Green Building in Western China (Project No.: LSZZ202113).

References
[1] C.L. Page, K.W.J. Treadaway, Aspects of the electrochemistry of steel in concrete, Nature 297 (1982) 109–1982.
[2] M. Stefanoni, U.M. Angst, B. Elsener, Kinetics of electrochemical dissolution of metals in porous media, Nat. Mater. 18 (9) (2019) 942–947, https://doi.org/
10.1038/s41563-019-0439-8.
[3] B. Hou, X. Li, X. Ma, C. Du, D. Zhang, The cost of corrosion in China, npj Materials Degradation 1 (2017) 1–10, https://doi.org/10.1038/s41529-017-0005-2.
[4] Italy’s Morandi bridge collapse − what do we know?. https://www.engineering.com/BIM/ArticleID/17517/Italys-Morandi-Bridge-CollapseWhat -DoWe-Know.
aspx#disqus_thread.
[5] A. Goyal, H.S. Pouya, E. Ganjian, P. Claisse, A review of corrosion and protection of steel in concrete, Arabian J. Sci. Eng. (2018), https://doi.org/10.1007/
s13369-018-3303-2.
[6] B. Guo, G. Qiao, D. Niu, O. Jinping, Electrochemical corrosion control for reinforced concrete structures− A review, Int J. Electrochem Sc. 15 (2020) 5723–5740.
[7] U.M. Angst, A critical review of the science and engineering of cathodic protection of steel in soil and concrete, Corrosion Journal.ORG 12 (2019) 1420–1433.
[8] Y. Liu, X. Shi, Cathodic protection technologies for reinforced concrete: introduction and recent developments, Rev. Chem. Eng. (25) (2009) 339–388.

13
B. Guo et al. Journal of Building Engineering 45 (2022) 103619

[9] C. Christodoulou, G. Glass, J. Webb, S. Austin, C. Goodier, Assessing the long term benefits of impressed current cathodic protection, Corrosion Sci. 52 (8)
(2010) 2671–2679, https://doi.org/10.1016/j.corsci.2010.04.018.
[10] B. Guo, G. Qiao, D. Li, J. Ou, Multi-species reactive transport modeling of electrochemical corrosion control in saturated concrete structures including electrode
reactions and thermodynamic equilibrium, Construct. Build. Mater. 278 (2021) 122228, https://doi.org/10.1016/j.conbuildmat.2020.122228.
[11] R.B. Polder, W.H.A. Peelen, B.T.J. Stoop, E.A.C. Neeft, Early stage beneficial effects of cathodic protection in concrete structures, Mater. Corros. 62 (2) (2011)
105–110, https://doi.org/10.1002/maco.201005803.
[12] L. Bertolini, E. Redaelli, Throwing power of cathodic prevention applied by means of sacrificial anodes to partially submerged marine reinforced concrete piles:
results of numerical simulations, Corrosion Sci. 51 (9) (2009) 2218–2230, https://doi.org/10.1016/j.corsci.2009.06.012.
[13] P. Pedeferri, Cathodic protection and cathodic prevention, Construct. Build. Mater. 10 (5) (1996) 391–402, https://doi.org/10.1016/0950-0618(95)00017-8.
[14] K. Wilson, M. Jawed, V. Ngala, The selection and use of cathodic protection systems for the repair of reinforced concrete structures, Construct. Build. Mater. 39
(2013) 19–25, https://doi.org/10.1016/j.conbuildmat.2012.05.037.
[15] G. Qiao, B. Guo, J. Ou, F. Xu, Z. Li, Numerical optimization of an impressed current cathodic protection system for reinforced concrete structures, Construct.
Build. Mater. 119 (2016) 260–267, https://doi.org/10.1016/j.conbuildmat.2016.05.012.
[16] G. Qiao, B. Guo, J. Ou, Numerical simulation to optimize impressed current cathodic protection systems for RC structures, J. Mater. Civ. Eng. 29 (6) (2017) 1,
https://doi.org/10.1061/(ASCE)MT.1943-5533.0001837, 1.
[17] A.M. Hassanein, G.K. Glass, N.R. Buenfeld, Protection current distribution in reinforced concrete cathodic protection systems, Cement Concr. Compos. 24 (2002)
159–167, https://doi.org/10.1016/S0958-9465(01)00036-1.
[18] S. Fonna, S. Huzni, A. Zaim, Simulation of cathodic protection on reinforced concrete using BEM, J. Mech. Eng. 2 (4) (2017) 111–122.
[19] M.M.S. Cheung, C. Cao, Application of cathodic protection for controlling macrocell corrosion in chloride contaminated RC structures, Construct. Build. Mater.
45 (2013) 199–207, https://doi.org/10.1016/j.conbuildmat.2013.04.010.
[20] M. Bruns, M. Raupach, Protection of the opposite reinforcement layer of RC-structures by CP - results of numerical simulations, Mater. Corros. 61 (6) (2010)
505–511, https://doi.org/10.1002/maco.200905584.
[21] E.B. Muehlenkamp, M.D. Koretsky, J.C. Westall, Effect of moisture on the spatial uniformity of cathodic protection of steel in reinforced concrete, Corrosion-US
61 (6) (2005) 519–533.
[22] Beasy, Corrosion modeling software. www.beasy.com.
[23] G.K. Glass, A.M. Hassanein, N.R. Buenfeld, Cathodic protection afforded by an intermittent current applied to reinforced concrete, Corrosion Sci. 43 (2001)
1111–1131.
[24] A.M. Hassanein, G.K. Glass, N.R. Buenfeld, Effect of intermittent cathodic protection on chloride and hydroxyl concentration profiles in reinforced concrete, Br.
Corrosion J. 34 (4) (1999).
[25] Comsol, Multiphysics. www.comsol.eu.
[26] Z.A. Kameche, F. Ghomari, M. Choinska, A. Khelidj, Assessment of liquid water and gas permeabilities of partially saturated ordinary concrete, Construct. Build.
Mater. 65 (2014) 551–565, https://doi.org/10.1016/j.conbuildmat.2014.04.137.
[27] J.P. Monlouis-Bonnaire, J. Verdier, B. Perrin, Prediction of the relative permeability to gas flow of cement-based materials, Cement Concr. Res. 34 (5) (2004)
737–744, https://doi.org/10.1016/S0008-8846(03)00071-1.
[28] C. Zhou, Predicting water permeability and relative gas permeability of unsaturated cement-based material from hydraulic diffusivity, Cement Concr. Res. 58
(2014) 143–151, https://doi.org/10.1016/j.cemconres.2014.01.016.
[29] E. Samson, J. Marchand, Modeling the effect of temperature on ionic transport in cementitious materials, Cement Concr. Res. 37 (3) (2007) 455–468, https://
doi.org/10.1016/j.cemconres.2006.11.008.
[30] F.P. Glasser, J. Marchand, E. Samson, Durability of concrete − Degradation phenomena involving detrimental chemical reactions, Cement Concr. Res. 38 (2)
(2008) 226–246, https://doi.org/10.1016/j.cemconres.2007.09.015.
[31] W.M. Haynes, D.R. Lide, T.J. Bruno, CRC Handbook of Chemistry and Physics, CRC Press, 2014.
[32] P. Spiesz, H.J.H. Brouwers, Influence of the applied voltage on the rapid chloride migration (RCM) test, Cement Concr. Res. 42 (8) (2012) 1072–1082, https://
doi.org/10.1016/j.cemconres.2012.04.007.
[33] Q. Yuan, C. Shi, G. De Schutter, K. Audenaert, D. Deng, Chloride binding of cement-based materials subjected to external chloride environment − A review,
Construct. Build. Mater. 23 (1) (2009) 1–13, https://doi.org/10.1016/j.conbuildmat.2008.02.004.
[34] W. Guo, J. Hu, Y. Ma, H. Huang, S. Yin, J. Wei, Q. Yu, The application of novel lightweight functional aggregates on the mitigation of acidification damage in the
external anode mortar during cathodic protection for reinforced concrete, Corrosion Sci. (2019) 108366, https://doi.org/10.1016/j.corsci.2019.108366.
[35] British Standard. Cathodic protection of steel in concrete, BS EN ISO 12696 (2016).
[36] U. Angst, B. Elsener, C.K. Larsen, Ø. Vennesland, Critical chloride content in reinforced concrete − A review, Cement Concr. Res. 39 (12) (2009) 1122–1138,
https://doi.org/10.1016/j.cemconres.2009.08.006.
[37] B. Guo, Erosion of Chloride in Reinforced Concrete Structures — Migration Mechanism of Multi-Field Coupling and Active Electric Corrosion Control, PhD
Thesis, Harbin Institute of Technology, China, 2018 (In Chinese).
[38] H. Zibara, Binding of External Chloride by Cement Pastes, PhD Thesis, University of Toronto, Canada, 2001.
[39] Y. Elakneswaran, A. Iwasa, T. Nawa, T. Sato, K. Kurumisawa, Ion-cement hydrate interactions govern multi-ionic transport model for cementitious materials,
Cement Concr. Res. 40 (12) (2010) 1756–1765, https://doi.org/10.1016/j.cemconres.2010.08.019.

14

You might also like