Open access
Technical Papers
Aug 25, 2020

Accelerated Laboratory Assessment of Discrete Sacrificial Anodes for Rehabilitation of Salt-Contaminated Reinforced Concrete

Publication: Journal of Materials in Civil Engineering
Volume 32, Issue 11

Abstract

For patch repairs of chloride contaminated RC structures, the use of discrete sacrificial anodes (DSAs) is indispensable. Without DSAs embedded, due to the ring effect, the failure of the RC around the patch area is accelerated. DSAs are increasingly being used, but few studies have evaluated the effectiveness of different DSAs and effects of the surrounding environment on the performance of different DSAs. This study employed four electrical parameters and electrochemical impedance spectroscopy (EIS) to evaluate three types of DSAs embedded in chloride-contaminated concrete through wet–dry and freeze–thaw cycles. The corrosion of the reinforcements is a stochastic process and the bound chloride ions play an important role in determining the corrosion state of the reinforcement. Most of the DSAs provided effective cathodic protection, but the effectiveness of DSAs could be influenced by the corrosion state of the reinforcement. The wet–dry and freeze–thaw cycles had significant influence on the performance of two types of the DSAs. Current DSA designs do not fully utilize the embedded Zn alloy.

Introduction

The last two decades have seen increased reliance on the use of discrete sacrificial anodes (DSAs), also called point anodes, for rehabilitation of RC structures, bridge decks, and pavements (Dugarte et al. 2007; Sergi 2011). Moreover, DSAs are much needed in patch repairs of concrete contaminated with a significant amount of chloride (either from marine environments or deicer applications). Because of the difference of chloride content between the patch repair and the surrounding area, patch repair without embedded DSAs treats only the symptoms of the concrete failure (e.g., spalled concrete) and induces the so-called ring effect, which is the accelerated failure of RC around the patch area (Sergi 2011). Typically, an anode consists of a sacrificial metal (e.g., Zn alloy) encased in a cementitious mortar that is highly alkaline and somewhat porous to sustain the function of the anode over the years in service.
DSAs increasingly have been employed since the late 1990s, but their effectiveness has been evaluated by a limited number of laboratory and field studies (Dugarte et al. 2007; Sagüés et al. 2005; Sergi 2011; Trocónis de Rincón et al. 2008). One of the main challenges is that these anodes are designed to serve at least 1020  years and their evaluation takes quite a long time without acceleration. Another main challenge is that the empiric methods used for testing the efficiency of the DSAs are not fully reliable, and only indirectly indicate the state of the metal of the DSAs (Martínez and Andrade 2008). Furthermore, environmental factors of the RC structures such as moisture level, chloride content, and temperature may influence the effectiveness of DSA. The presence of chloride ions increases the risk of corrosion and the corrosion rate of the reinforcement (Ahmad 2003), and the permeation properties of concrete significantly influence the chloride diffusion (Basheer et al. 2001). Freeze–thaw actions, which are one of the most aggressive environmental attacks in cold regions, can compromise the integrity of concrete and thereby increase the permeation properties of RC structures, which decreases the corrosion initiation time and accelerates the corrosion rate (Wang et al. 2014).
The objective of this exploratory study was to preliminarily investigate and assess the relative performance of various DSAs in steel-reinforced concrete under two typical salt-contaminated levels. To accelerate the consumption rate of the sacrificial anodes, wet–dry cycles and freeze–thaw attacks were conducted on the specimens. Electrochemical parameters such as half-cell potential, corrosion current density, depolarized potential, and on potential were investigated. Half-cell potential and corrosion current density both are needed to confirm the corrosion state of the reinforcements. Depolarized potential and on potential are significant parameters for confirming the effectiveness and evaluating the protecting performance of DSAs. The relationship between the corrosion state of the rebar and the effectiveness of the sacrificial anodes should be investigated in the future, which will significantly help the prediction of service life of DSAs.

Experimental

A three-electrode system was employed in this study (Fig. 1). Two steel rebars were embedded in the mortar specimen; the top rebar served as the working electrode, and the bottom rebar served as the counter electrode. A graphite bar also was embedded in the specimen near the top rebar, which was used as the reference electrode. To investigate the throwing power of the DSAs, the top rebar was cut into three sections and reconnected using acrylic rods and epoxy, so that the top rebar was mechanically connected but electrically disconnected. There were four point-locations on the top rebar tied with a steel wire, which enabled electrical connections to reach outside the reinforced mortar specimen and facilitated the measurement of point location (node) potentials of the top rebar and the DSA. This specimen setup was employed to monitor the corrosion state of the top rebar. With the aid of electrochemical impedance spectroscopy (EIS), the corrosion current (Icorr) of Section 1 of the top rebar was obtained through the Stern–Geary equation (Moreno et al. 2004), Icorr=B/Rp, where the value of B was set at 26  mV and Rp was the polarization resistance, and in this study we used the charge transfer resistance Rct from EIS results. The corresponding corrosion current density (icorr) was calculated via the equation
icorr=Icorr/A
(1)
where A = surface area of Section 1 of the top rebar. Other electrical parameters such as instant-off potential, off4h potential, on potential, depolarized potential, and half-cell potential also periodically were measured. Half-cell potential of the top rebars was measured after the specimens were cured for 28 days, and all the electrical parameters except half-cell potential were measured after the initiation of corrosion of the top rebar was confirmed and the DSAs were connected to the top rebar.
Fig. 1. Three-electrode system and specimen geometry.

Performance Parameters

Half-Cell Potential

The half-cell potential of the top rebar was measured versus the reference electrode every weekday before the initiation of the corrosion of the top rebar, and it was measured on Tuesday and Friday of every week after the initiation of corrosion of the top rebar.

Instant-Off Potential and On Potential

The instant-off potential of the top rebar was measured versus the reference electrode 5 s after disconnecting the DSA from the top rebar every weekday.
The on potential of the top rebar was measured versus the reference electrode 5 s after reconnecting the DSA to the top rebar following the 4-h depolarization period every weekday.

Off4h Potential

The off4h potential of the top rebar was measured versus the reference electrode 4 h after disconnecting the DSA to the top rebar every weekday.

Depolarized Potential

Depolarized potential of the top rebar was obtained by subtracting off4h potential from instant-off potential.

Electrochemical Impedance Spectroscopy

Using a PARSTAT MC multichannel Potentiostat (Ametek Scientific Instruments, Berwyn, Pennsylvania), the EIS measurements were conducted on the three-electrode system in Fig. 1 and the results were analyzed using VersaStudio version 2.00 software. The EIS measurements were conducted in a frequency range of 100  kHz0.005  Hz with an applied alternating current (AC) signal of ±10  mV versus the open circuit potential (OCP) between the working electrode and reference electrode. Before conducting the EIS test, about 50  mL tap water was poured gently on the top surface of each mortar specimen. All the EIS results were interpreted using ZSimpWin version 3.60 software.

Materials

Three types of commercial DSAs were investigated, denoted S, B, and X, respectively, and each contained a sacrificial anode encased in a repair patch. The size and chemical composition of the three types of DSAs are given in Fig. 2 and Table 1. No. 3 rebar (diameter 9.525 mm), steel wire, fast-curing epoxy, and acrylic pipe were purchased from Home Depot (Lewiston, Idaho). Due to the good chemical stability, graphite bar (from Fanty_store, Shenzhen, Guangdong, China) was selected as the reference electrode. An ASTM specification C150 (ASTM 2016) Type I/II portland cement was used in this study as received. The chemical composition of cement was reported by Du et al. (2017). Analytical grade NaCl was purchased from Sigma-Aldrich (St. Louis, Missouri) and used without further purification. The fine aggregate used was multipurpose siliceous sand; before proportioning and admixing, the fine aggregate was prewetted and taken to a saturated surface dry (SSD) condition. A high-range water-reducing admixture (7920) from BASF (San Diego, California) was used to achieve a reasonable workability of the mortar mixture, at a dosage of 0.4% by mass of cement.
Fig. 2. Sizes of the three commercial DSAs.
Table 1. Chemical composition of three types of DSAs (%)
ElementSBX
Zn95.396.592.9
O1.31.02.8
C2.21.92.6
Si1.00.60.9
Others0.20.8

Specimen Fabrication

The top rebar was cut into three parts to the length of 12.2 cm (4.8 in.), 12.2 cm (4.8 in.), and 36.6 cm (14.4 in.) for Sections 1, 2, and 3, respectively. Steel wires were soldered to one end of the three parts, and then the three parts were glued with epoxy and fixed in a short length of acrylic pipe (Fig. 1). After the preparation, the top rebar was mechanically connected but electrically disconnected. The top rebar was placed in the middle of the width of the specimen, and the distance between the center of the top rebar and the top surface of the specimen was 9  cm. The bottom rebar also was placed in the middle of the width of the specimen, and the distance between the center of the bottom rebar and the bottom surface of the specimen was 3  cm. The graphite electrode was placed close to the top rebar (approximately 0.5  cm).
To simulate aged concrete of bridge decks, the specimens consisted of two layers: a chloride-contaminated layer and a chloride-free layer. Two typical levels of chloride contamination, 1.0% and 2.0% by mass of cement, were simulated. The water-to-cement mass ratio was 0.45 and the sand-to-cement mass ratio was 3.0. Triplicate specimens were prepared for each type of DSA with each level of chloride contamination. The specimens were named with a letter and two numbers, denoting type of DSA, chloride contamination level (e.g., 1 denotes 1% chloride contamination), and specimen number, respectively (Table 2).
Table 2. Sample group coding
Commercial DSA product typeSample number
Group 1S1-1
S1-3
S1-2
S2-1
S2-2
S2-3
Group 2B1-1
B1-2
B1-3
B2-1
B2-2
B2-3
Group 3X1-1
X1-2
X1-3
X2-1
X2-2
X2-3
The SSD sand first was mixed with cement in a 25-L mixer for 60 s, then water was added and mixed for 120 s. Afterward, the water-reducing admixture was added into the mortar mixture and mixed continuously for about 5 min. The fresh chloride free mixture then was cast into the mold, which already had the rebars, graphite electrode, and DSA embedded, and was carefully compacted to minimize the amount of entrapped air. The 8-cm-thick chloride-free layer was cured for 1 h, and then the fresh chloride contaminated mixture was prepared by following the same steps (except using water with a given amount of dissolved NaCl) and cast into the mold and carefully compacted. All the specimens were demolded after being cured at room temperature (RT) and relative humidity (RH) of 20%±2% for 24  h. After demolding, all the mortar specimens were cured under the same conditions for another 27 days before measurement. To enable ponding of aqueous solutions on the top surface, silicone was sprayed along the four sides of the top surface.

Testing Environment

All specimens were tested under the same conditions. The period of a wet–dry cycle was 2 weeks (1 week for the wet cycle and 1 week for the dry cycle). About 100  mL tap water was added on the top surface of the specimens every weekday during the wet cycle of the first 12 weeks. After that, to accelerate the initiation of corrosion of the rebar, 100  mL 23% by weight NaCl solution was used instead of tap water every Monday of the wet cycle. The 24-h freeze–thaw exposure of specimens was conducted in accordance with ASTM C666 (ASTM 2015b) after every six wet–dry cycles. The EIS measurement was carried out before each freeze–thaw exposure.

Scanning Electron Microscopy with Energy-Dispersive X-Ray Spectroscopy

For morphology and elemental analysis, cross-sectional samples were cut from the representative DSAs after service and from the corresponding new DSAs before service. Three strips of cross-sectional samples representing the changes in microstructural properties were cut from the cement mortar specimens incorporating three different representative DSAs. After being coated with carbon powder and vacuumed, the sliced samples were subjected to scanning electron microscopy (SEM) to examine the morphology at the microscopic level. The beam conditions used were as follows: 15  kV accelerating voltage, 50 nA beam current, and focused beam diameter. Energy-dispersive X-ray spectroscopy (EDX) semiquantified the percentages of different elements of specific sections by picking up different spots on the samples.

X-Ray Diffraction Analysis

To further confirm the phase composition of corrosion products of DSAs and find the difference between the DSAs before and after service, X-ray diffraction analysis (XRD) was conducted on the slice samples of DSAs. The XRD test was performed on a Siemens (Munich, Germany) D5000 diffractometer using Cu Kα radiation in the 2θ region between 2° and 90° with a scanning rate of 0.05°/min.

Results and Discussion

Potential Evolution of Graphite Electrode

As a reference electrode, the graphite embedded in the mortar specimen should have a stable electrode potential, which was partially validated in this work. Ag/AgCl saturated KCl aqueous electrode (Ag/AgCl) and saturated calomel electrode (SCE) are two commonly used reference electrodes in concrete (Duffó et al. 2009), which have a potential of +197 and +241  mV with respect to a standard hydrogen electrode, respectively. However, because their long-term stability could be significantly compromised by extreme service conditions, and their costs are high, an alternative reference electrode should be developed for concrete (or mortar). Muralidharan et al. (2006) investigated the feasibility of using graphite as a reference electrode in cement mortar with a chloride contamination level of 0% and 3% by weight of cement. All the mortar specimens were placed in a yard for testing. They found that the potential evolution of embedded graphite electrode versus SCE was about 75  mV for 12 months, which was limited and acceptable. Similarly, the present study also investigated the potential of embedded graphite electrode against Ag/AgCl.
The potential evolution of most of the graphite electrodes embedded in mortar specimens was about 100  mV, which was acceptable and consistent with the results of Muralidharan et al. (2006) (Fig. 3). However, the potential evolution of Graphite electrode S1-1 had two abnormal values (about 700  mV) around the 250th day, and the fluctuation of the potential evolution of Graphite electrodes B2-3 and X2-1 reached almost 200  mV. Moreover, the potential evolution of Graphite electrodes B22,X13, and X2-2 suddenly became unstable after about 400 days, 200 days, and 250 days, respectively. Then the difference of the potentials of corresponding graphite reference electrodes between wet–dry cycles could be as large as 250  mV. The possible factor that contributed to this phenomenon is that a graphite electrode is sensitive to the environmental changes in concrete, such as the fluctuations of moisture level (Muralidharan et al. 2006). Therefore, it is viable to use a graphite electrode as an embedded reference electrode to monitor the half-cell potential of reinforcements in concrete. However, periodical calibration using an external reference electrode (e.g., Ag/AgCl) is necessary.
Fig. 3. Potential evolution of embedded graphite electrodes with time: (a) in specimens with Type S DSA; (b) in specimens with Type B DSA; and (c) in specimens with Type X DSA.

Half-Cell Potential and Corrosion Current Density

All the electrical parameters were measured versus the embedded graphite electrode. However, to make the results more comprehensible, all the parameters were reported against Ag/AgCl.
Half-cell potential is one of the key parameters that provides information regarding the state of reinforcement corrosion (Ahmad 2003). It is used as a thermodynamic parameter for assessing whether or not the corrosion of reinforcement is likely to occur. According to ASTM C876 (ASTM 2015a), the probability of active corrosion of reinforcement exceeds 90% when the half-cell potential of the rebar against Ag/AgCl drops below 233  mV.
The half-cell potential of Section 1 of the top rebars in all specimens decreased to an extremum value of about 300  mV in the first 50 days, and then increased to a level above 233  mV and kept fluctuating, and after a certain amount of time (212419  days) dropped below 233  mV again (Fig. 4). Due to the chloride contamination in the top portion, the top rebar suffered from severe corrosion during the first 50 days, and the effect of passivation seemed to be counteracted, which corresponds to the decreasing process of half-cell potential of the top rebar. Then the passivation of the top rebar dominated, which protected the top rebar from corrosion and hence increased the half-cell potential. Afterward, the half-cell potentials decreased during the wet cycle and increased during the dry cycle. This finding was consistent with that of Angst et al. (2011), because the wet–dry cycle was necessary for depassivation and the rebar had a tendency to repassivate without further chloride ingress.
Fig. 4. Half-cell potential of Section 1 of top rebar of specimens embedded with DSAs: (a) Type S; (b) Type B; and (c) Type X.
The half-cell potentials of most specimens decreased one or two wet–dry cycles after every freeze–thaw action. The freeze–thaw actions generated a network of well distributed and uniform cracks, which could increase both the initial sorptivity and total water absorption and promote the transport of chloride ions (Yang et al. 2006). Therefore, the depassivation of the top rebars was facilitated and the half-cell potentials decreased. In addition, the amplitudes of the fluctuations of the half-cell potentials of specimens with a chloride contamination level of 2% by weight were larger than those of specimens with a chloride contamination level of 1% by weight. This was due mainly to the bound chloride ions, because the wet–dry cycles resulted in the lowest value of diffusion coefficient and surface chloride content (Song et al. 2008). The higher chloride contamination level resulted in a larger amount of bound chloride (Yuan et al. 2009), which increased the reservoir of chloride ions that were locally available at the steel–concrete interface (Glass and Buenfeld 1997, 2000).
Compared with previous studies (Angst et al. 2011; Bouteiller et al. 2012), the trends of half-cell potential of the embedded rebars before the initiation of corrosion were similar, increasing during the first 50100  days and then becoming stable until the initiation of corrosion. However, Bouteiller et al. (2012) found that the half-cell potential of the rebars embedded in the concrete specimens with a water-to-cement ratio of 0.45 kept fluctuating between 100 and 300  mV (against SCE) for over 600 days, whereas Angst et al. (2011) found that some rebars embedded in the concrete specimens with a water-to-cement ratio of 0.40 repassivated after a sudden decrease of half-cell potential on the 126th day. The reasons for these phenomena were that the specimens were chloride free, which made the pH value well sustained and facilitated the spontaneous repassivation (Glass and Buenfeld 2000), and that the chloride concentration of the solution used during the wet cycle was too low, which made the available chloride content at the interface between the rebar and concrete matrix much lower than the critical chloride threshold (Alonso et al. 2000; Glass and Buenfeld 1997).
When the half-cell potential of top rebar dropped below 233  mV again, it implies that the passive film was compromised and localized corrosion reinitiated. However, half-cell potential alone is not sufficient to confirm the corrosion state of rebar because it can be influenced by a number of factors, including polarization by limited diffusion of oxygen, concrete porosity, and the presence of highly resistive layers (Song and Saraswathy 2007).
To confirm the state of reinforcement corrosion, the value of corrosion current density (icorr) is a better indicator. It is a kinetic parameter that represents an overall estimate of the rate of corrosion attack of reinforcement (Ahmad 2003). The rebar is considered in a passive state when icorr is smaller than 0.1μA/cm2, whereas the rebar is in an active state when icorr is larger than 1  μA/cm2 (Ahmad 2003). The rebar is in metastable state if icorr is between 0.1 and 1  μA/cm2. The corrosion current densities obtained from electrochemical measurements are listed in Table 3. The values of corrosion current density were used to confirm that the top rebars were in a metastable state, and the sacrificial anodes were subsequently connected to the top rebar to provide cathodic protection. Based on the half-cell potential evolution and the change of corrosion current density, the time-to-corrosion (Ti) was estimated (Table 4).
Table 3. Corrosion current density of Section 1 of top rebar after wet–dry and freeze–thaw cycles
icorr(μA/cm2)After 6 wet–dry cyclesAfter 12 wet–dry cyclesAfter 18 wet–dry cyclesAfter 24 wet–dry cyclesAfter 30 wet–dry cycles
S1-10.2980.1730.6250.6910.504
S1-20.0260.0070.5600.2580.061
S1-39.94×1050.0550.3822.6990.603
S2-10.2490.1380.9792.7987.109
S2-20.0810.0122.7081.9631.519
S2-30.0090.0314.66×1040.8820.289
B1-11.95×1050.0140.2251.7470.002
B1-20.0080.1094.7944.3415.777
B1-32.30×1080.0310.0310.0150.410
B2-18.56×1080.0260.0228.04×1060.112
B2-23.12×1040.9000.8760.5140.712
B2-37.75×10110.0501.06×1091.78×1040.076
X1-10.0420.0080.1900.7950.001
X1-20.0110.0020.1330.5020.539
X1-30.0860.0010.6320.2880.240
X2-10.0100.0130.0700.0150.037
X2-21.10×1040.0513.57×1041.23×1042.328
X2-30.0130.0320.3060.0590.180
Table 4. Time-to-corrosion of Section 1 of top rebar after wet–dry and freeze–thaw cycles (days)
SpecimenTime-to-corrosion
S1-1212
S1-2245
S1-3224
B1-1168
B1-2224
B1-3377
X1-1283
X1-2276
X1-3213
S2-1220
S2-2253
S2-3283
B2-1425
B2-2255
B2-3479
X2-1
X2-2409
X2-3
Section 1 of the top rebar of each specimen reached the metastable state (0.1  μA/cm2<icorr<1  μA/cm2) after different numbers of wet–dry cycles based on the aforementioned criterion. Section 1 of B1-1 reached the metastable state after only 12 wet–dry cycles, and 10 other specimens reached the metastable state after 18 wet–dry cycles, including 4 specimens that had already initiated corrosion. Section 1 of the top rebar of S2-3 reached the metastable state after 24 wet–dry cycles, whereas Section 1 of the top rebars of B1-3, B2-1, B2-3, and X2-2 reached the metastable state after 30 wet–dry cycles. Moreover, the results of X2-1 and X2-3 showed that Section 1 of their top rebars was still in passive state.
Before the top rebars reached the metastable state, the corrosion current density of most top rebars fluctuated below 0.1  μA/cm2, which was consistent with the half-cell potential evolution. The top rebars of all specimens with the chloride contamination level of 1% by weight reached the metastable state (0.1  μA/cm2<icorr<1  μA/cm2) between 212 and 283 days, except B1-1 and B1-3. Conversely, the differences between the time-to-corrosion of specimens with the chloride contamination level of 2% by weight were much larger. This indicates that the type of embedded sacrificial anode had limited influence on the time-to-corrosion of rebar. This is consistent with the results of half-cell potential evolution. However, the total chloride content still played an important role. When the top rebar was in a metastable state, the corresponding corrosion current densities of the top rebars of most specimens with higher chloride contamination levels were higher than those of specimens with lower chloride contamination levels. This is consistent with the results of Alonso et al. (2000).
Because rebar corrosion in concrete is complicated and stochastic, statistically speaking, there could be a few uncommonly large or small values of corrosion current density and time-to-corrosion. Alonso et al. (2000) found that, based on the corrosion current density, rebars of specimens with the total chloride content of 0.48% by weight were in an active state at the beginning and entered a passive state after 36 days. Yu et al. (2010) found that the rebar in one self-compacting concrete specimen with a water-to-cement ratio of 0.5 was not corroded after 600 days, and Angst et al. (2011) found that rebars in concrete specimens with a water-to-cement ratio of 0.4 had stable pitting only during wet cycles and immediately stopped corroding during dry cycles.

Depolarized Potential and on Potential

Depolarized potential (also known as potential decay) of the rebar in concrete often is considered to be a significant parameter that verifies the effectiveness of the cathodic protection applied by sacrificial anodes (Bertolini et al. 2002). When the depolarized potential is lower than 100  mV (i.e., potential decay is higher than 100  mV), the cathodic protection is effective and a near-passive state has been induced in the protected rebar (Bertolini et al. 2002). The depolarized potentials of Sections 2 and 3 of the top rebars of representative specimens were about 0 mV (Fig. 5). This indicates that the discrete sacrificial anodes could not effectively protect Sections 2 and 3 of the top rebars because they were not electrically connected to Sections 2 and 3 of the top rebar, which means that the throwing power of discrete sacrificial anodes was limited and negligible. However, the results of the depolarized potential of Section 1 of the top rebars of all specimens, except uncorroded rebars, were lower than 100  mV (Fig. 6). This indicates that discrete sacrificial anodes could guarantee a sufficient cathodic polarization and effectively protect the connected rebar section (Bertolini et al. 2002). Because each section of the top rebar was electrically disconnected, the onset of macrocouples caused by different conditions of polarization of each section was avoided and the results of depolarized potential were reliable (Bertolini et al. 2002).
Fig. 5. Depolarized potential of sections of top rebar of representative specimens: (a) Section 2; and (b) Section 3.
Fig. 6. Depolarized potential of Section 1 of top rebar of specimens embedded with DSAs: (a) Type S; (b) Type B; and (c) Type X.
The depolarized potentials of Section 1 of the top rebars of specimens incorporating S-type DSAs were between 50 and 200  mV at the beginning of cathodic protection [Fig. 6(a)]. Section 1 of the top rebars of all specimens with the chloride contamination level of 1% by weight tended to decrease over time, whereas only that of S2-3 had a suppress trend. The rebar trend of depolarized potential indicated that the anodes needed to suppress the rebar potential to continuously and effectively provide electrons to the protected section of the top rebars. This is consistent with the decreasing trend of corrosion current density (icorr) of Section 1 of the top rebars after the connection of anodes. When the rebar was protected or repassivated, the resistance of the rebar would be kept in a range of high values or increased. In other words, the depolarized potential of DSAs would be kept in a range of low values or decreased, so as to maintain a sufficient level of current density of cathodic protection. Conversely, the depolarized potentials of S2-1 and S2-2 kept fluctuating between 100 and 200  mV after the connection and tended to be stable over time. Such results were consistent with the corresponding results of icorr, which reveals that the anode of S2-1 could not effectively protect the rebar and the anode of S2-2 could only imperfectly repassivate the rebar. Bertolini et al. (2002) found that sacrificial anodes had poor effectiveness in protecting rebars, and initiated pitting corrosion when placed a few meters above sea level. Because the testing condition was similar to sea level and Section 1 of the top rebar of S2-1 and S2-2 were actively corroded when the anodes were connected, the S-type sacrificial anode could probably prevent only new pitting from being initiated, but could not prevent existing pitting from propagating. The amplitudes of the fluctuation of depolarized potential of S2-1, S2-2, and S2-3 were on average 120% larger than those of S1-1, S1-2, and S1-3. The effect of chloride contamination level on the depolarized potential of DSAs probably could be enhanced by wet–dry cycles because the moisture at the interface between anode and concrete matrix was vital to anode performance (Liu and Shi 2009).
Similarly, the depolarized potential of B1-3, B2-1, and B2-2 was about 50  mV at the beginning of cathodic protection and tended to decrease over time, whereas the depolarized potential of B1-1 and B1-2 was about 200  mV at the beginning of cathodic protection, and then kept fluctuating, tended to be stable, and slightly increased over time, respectively [Fig. 6(b)]. The corrosion current density (icorr) of B1-1 and B2-2 effectively decreased, whereas that of B1-2 increased after the connection of the anodes, which was consistent with the corresponding results of depolarized potential. The amplitude of fluctuation B1-2 (200  mV) was larger than that of B1-1 and B1-3 (30–50 mV). A similar phenomenon was found in specimens incorporating S-type and B-type anodes, which means that cathodic protection by sacrificial anodes may not be able to effectively protect the actively corroded rebars. The chloride contamination level seemed to have a limited effect on the stability of depolarized potential of Type B DSAs; however, the effect of wet–dry and freeze–thaw cycles on the amplitude of fluctuation of depolarized potential increased from 50 to 100  mV after 150 days, which is attributable to the increased number of cracks in the cement matrix.
For specimens incorporating with X-type anodes, the depolarized potential of Section 1 of the top rebar of X1-1, X1-2, and X1-3 initially was between 10 and 60  mV, and that of X2-2 was about 200  mV [Fig. 6(c)]. All tended to decrease over time except that of X1-1. The tendency of corrosion current density of X1-3 was consistent with the results of depolarized potential. Oddly, based on the aforementioned criterion, the depolarized potential of X1-1 demonstrated a useless anode but the corrosion current density of X1-1 significantly decreased, which indicated that the corrosion current density was a more reliable parameter to confirm the effectiveness of a DSA. Comparing the amplitude of the fluctuation of depolarized potential of specimens with different chloride contamination levels (1% by weight Cl, 60  mV versus 2% by weight Cl, 250  mV) showed that the stability of the Type X DSA was greater at a lower chloride contamination level.
In addition to depolarized potential, on potential can be regarded as a driving power that guarantees the transport of electrons from DSA to rebar. It also determines the polarized potential of the rebar, which indicates whether or not the rebar is completely protected (Rajani and Kleiner 2003). There is a polarized potential shift between the on potential and polarized potential, which is dependent on how long the rebar is connected to the DSA. It often takes weeks or even months to achieve full polarization, in which rebars are polarized to 733  mV against Ag/AgCl (Rajani and Kleiner 2003). In this study, because the polarization period was only 20 or 68  h, some of the rebars might not have been completely polarized.
Fig. 7 depicts the on potential of Section 1 of the top rebars of specimens over time. The on potential of Section 1 of the top rebars of all specimens incorporating a Type S anode was about 500  mV over time, except S1-1 [Fig. 7(a)]. The on potential of S1-1 decreased from 500 to 800  mV and then increased to 600  mV over time. If the on potential of the top rebar was only 500  mV, the corresponding polarized potential could hardly be lower than 733  mV because the polarization period was only 20 or 68  h and the typical polarized potential shift is about 160  mV (Rajani and Kleiner 2003). This indicates that the S-type anode could not completely polarize the top rebar and perfectly protect it, which means that new pitting would not initiate but existing pitting would propagate (Bertolini et al. 1998). This was consistent with the results of corrosion current density after the connection of anodes. For S1-1, after the connection, the anode needed to provide more electrons to the top rebar than the amount of consumed electrons, which corresponded to the decreasing trend of the on potential. Because further corrosion of top rebar was prevented, the anode needed only to provide a lesser number of electrons to offset the consumed electrons, which corresponded to the increasing trend of the on potential. In addition, the amplitude of the fluctuation of the on potential was kept stable at about 100  mV, and the wet–dry cycles, freeze–thaw actions, and different chloride contamination levels seemed to have limited influence on the on potential.
Fig. 7. On potential of Section 1 of top rebar of specimens embedded with DSAs: (a) Type S; (b) Type B; and (c) Type X.
The on potential of Section 1 of the top rebars of all specimens incorporating Type B anode also was about 500  mV over time, except B2-1 and B1-2 [Fig. 7(b)]. However, the amplitude of the fluctuation of the on potential of specimens incorporating B type anodes was 100  mV, which was greater than that of specimens incorporating S-type anodes. This indicates that the influences of wet–dry cycles and freeze–thaw actions on the performance of Type B anodes were almost twice those on Type S anodes. For B1-2, the on potential was about 600  mV and then tended to increase over time, which indicates that this anode was losing its effectiveness of protecting the top rebar. This was consistent with the results of depolarized potential and corrosion current density.
The on potential of Section 1 of the top rebars of all specimens incorporating Type X anodes was about 600  mV over time, except X1-1 [Fig. 7(c)]. This indicates that Type X anodes could effectively transport electrons to the top rebars and fully protect them, which was consistent with the corresponding results of depolarized potential and corrosion current density. Similarly, the influences of wet–dry cycles and freeze–thaw actions on the amplitude of the fluctuation of on potential of Type X anodes were about 200  mV, which was almost 4 times that of Type S anodes.

EIS Results

Because the internal cracks of concrete may increase with the number of wet–dry cycles and freeze–thaw actions and the corrosion product of the rebar also may change with time and condition, using only one equivalent circuit could not accurately explain the complexity. Therefore, based on the simplified Randles cell, five equivalent electrical circuit models [Figs. 8(a–e)] were employed to interpret the impedance spectra of different conditions (Ribeiro and Abrantes 2016). In Fig. 8, Ro represents the offset resistance, the value of which was minor and negligible and which had no physical meaning; Rc and Qc represent the resistance and constant phase element (CPE) of the concrete that covered the rebar, respectively; and Rct and Qdl represent the charge transfer resistance and the CPE of the double layer at the rebar–concrete interface, which were composed of passive film and corrosion product, respectively. The CPE often is used instead of a perfect capacitor when the semicircle of a Nyquist plot is displaced or depressed (Ribeiro and Abrantes 2016), which is caused by the roughness of the corrosion product at the surface of rebar or the heterogeneity of the specimen (Nazari et al. 2017b). The impedance of CPE is expressed as follows (Nazari et al. 2017b):
ZCPE=1/Yo(jω)n
(2)
where Yo = magnitude of CPE; and generally 0.9n1. The existence of Warburg impedance (W) in the low frequency indicated that the corrosion of the top rebar was controlled by a diffusive process (Nazari et al. 2017a). The main differences between the equivalent circuits were how the parallel connection of Qc and Rc was added to the simplified Randles cell and to which cell the Warburg impedance should be added.
Fig. 8. Equivalent circuit models for interpreting the EIS results: (a) Ro(QcRc)(QdlRct); (b) Ro(QcRcW)(QdlRct); (c) Ro(QcRc)(QdlRctW); (d) Ro{Qc[Rc(QdlRct)]}; and (e) Ro{Qc[Rc(QdlRctW)]}.
Because the corrosion process of rebars in concrete was stochastic and complex, results of representative specimens incorporating each type of DSA and with each level of chloride contamination were selected. The impedance Nyquist plots and the corresponding Bode diagrams of Section 1 of the top rebars of representative specimens after different numbers of wet–dry cycles are shown in Figs. 9 and 10, respectively. The measured data of the impedance (real, imagined, and absolute) or the phase angle were plotted in scatters and the fitting curves were calculated from the equivalent circuit models. For S1-1 and S2-1, because there were two semicircles in the Nyquist plots and the slope of the Bode magnitude plots in low frequency increased after 12 wet–dry cycles, Equivalent circuits Figs. 8(a and d) were used to interpret the Nyquist plots of S1-1 and S2-1 before and after 12 wet–dry cycles, respectively. According to the findings of Montemor et al. (2000), the increase of the slope indicates the decrease of charge transfer resistance, which is consistent with the values in Table 5. The change of the concrete–steel interface, which was due to the growth of the rust layer at the interface, led to the use of different equivalent circuits. For B1-1, because the low-frequency part of the Nyquist plot was a line and the slope of the Bode magnitude plots in low frequency after 30 wet–dry cycles was greater than that of preceding plots, Equivalent circuits Figs. 8(a and b) were needed to interpret the Nyquist plots of B1-1 before and after 30 wet–dry cycles, respectively. Because the amount of microcracks increased with wet–dry and freeze–thaw cycles, which facilitated the diffusion of O2 and Cl in the cement matrix, the values of Warburg impedance decreased. When the top rebar was directly in contact with the external substances, the Warburg impedance element in the equivalent circuit was eliminated. Similarly, according to the change of the slope of the Bode magnitude plots, Equivalent circuits Figs. 8(c and e) were used to interpret the Nyquist plots of B2-2 after 6 and 12 wet–dry cycles, respectively. Then Equivalent circuit Fig. 8(d) was used to interpret the condition of B2-2 after 18 wet–dry cycles due to the increase of microcracks. Equivalent circuit Fig. 8(b) finally was used to interpret the Nyquist plot of B2-2 after 30 wet–dry cycles. Moreover, for X1-3, Equivalent circuit Fig. 8(b) was used during the first 6 wet–dry cycles and Equivalent circuit Fig. 8(d) was used after that to investigate the corresponding Nyquist plots. For X2-2, Equivalent circuit Fig. 8(d) was used during the first 12 wet–dry cycles, whereas Equivalent circuits Figs. 8(c and e), respectively, were used from 18 to 24 wet–dry cycles and after 30 wet–dry cycles to interpret the corresponding Nyquist plots.
Fig. 9. Nyquist diagrams of Section 1 of the top rebar: (a) S1-1; (b) S2-1; (c) B1-1; (d) B2-2; (e) X1-3; and (f) X2-2.
Fig. 10. Bode diagrams of Section 1 of the top rebar: (a) S1-1; (b) S2-1; (c) B1-1; (d) B2-2; (e) X1-3; and (f) X2-2.
Table 5. Electrochemical impedance parameters obtained by fitting EIS data to equivalent circuits
ExposureRo(Ωcm2)Rc(Ωcm2)Qc(Yo)(Ω1  cm2snf)Qc(n)Rct(Ωcm2)Qdl(Yo)(Ω1  cm2sndl)Qdl(n)Cdl(Fcm2)W(Ωcm2)
S1-1
 6 cycles2.54×1042.08×1015.68×1095.10×1016.35×1017.84×1056.70×1012.02×104
 12 cycles1.11×1092.51×1016.86×1094.90×1011.09×1028.22×1056.80×1012.68×104
 18 cycles6.76×1097.84×1018.62×1056.00×1013.03×1012.04×1095.80×1012.29×1012
 24 cycles1.65×1043.78×1011.31×1045.90×1012.76×1012.73×1095.70×1012.68×1012
 30 cycles4.32×1054.86×1019.97×1055.80×1013.78×1013.32×1095.40×1012.07×1012
S2-1
 6 cycles7.49×1055.99×1004.85×10111.00×1009.34×1011.44×1046.50×1016.17×104
 12 cycles1.41×1057.49×1006.26×10111.00×1001.69×1021.24×1046.40×1017.27×104
 18 cycles3.59×1076.59×1004.25×10111.00×1002.38×1012.73×1046.10×1019.67×104
 24 cycles9.88×1054.19×1006.71×10111.00×1008.32×1004.97×1046.40×1011.17×103
 30 cycles2.99×1001.17×1007.07×1041.00×1003.17×1001.77×1037.30×1013.49×103
B1-1
 6 cycles3.34×1006.00×1038.28×1057.80×1018.32×1052.10×1065.84×1016.05×1043.29×102
 12 cycles4.54×1004.32×1011.20×1069.33×1011.12×1031.39×1048.81×1012.93×1049.20×105
 18 cycles2.06×1035.89×1002.66×10111.00×1007.24×1016.03×1049.16×1018.89×1047.50×105
 24 cycles6.91×1063.94×1056.06×1055.53×1019.30×1004.38×10111.00×1004.38×10111.89×109
 30 cycles1.22×1041.19×1013.53×10111.00×1008.47×1037.32×1055.56×1011.82×102
B2-2
 6 cycles1.99×1057.55×1008.09×10111.00×1007.73×1042.21×1045.84×1012.44×1013.77×1022
 12 cycles2.96×1058.03×1006.95×10111.00×1001.88×1012.04×1046.30×1014.70×1041.83×1010
 18 cycles2.65×1044.27×1008.54×1061.20×1012.32×1014.60×1046.40×1011.82×103
 24 cycles3.19×1041.68×1008.95×1046.11×1014.70×1011.19×10131.33×1012.78×1067
 30 cycles1.07×1034.25×1007.30×1072.76×1012.97×1012.81×1031.00×1002.81×1037.80×104
X1-3
 6 cycles4.66×1068.16×1001.95×10108.50×1012.71×1022.38×1048.78×1014.31×1041.23×104
 12 cycles7.48×1051.02×1015.63×10107.69×1011.98×1047.73×1056.46×1014.55×103
 18 cycles9.16×1097.28×1008.64×1084.20×1013.69×1011.46×1046.29×1014.21×104
 24 cycles9.28×1078.68×1003.02×1096.80×1018.08×1011.57×1045.20×1011.80×103
 30 cycles8.13×1015.95×1001.86×1097.09×1019.13×1015.91×1054.79×1014.16×104
X2-2
 6 cycles5.69×1002.93×1031.93×1046.80×1012.11×1053.86×1061.00×1003.86×106
 12 cycles7.78×1076.59×1002.01×10101.00×1004.55×1022.07×1047.60×1018.98×104
 18 cycles3.76×1065.28×1002.72×10101.00×1006.53×1041.48×1046.50×1012.19×1021.29×1012
 24 cycles3.89×1048.68×1012.64×1046.90×1011.89×1054.01×1056.30×1018.12×1032.00×109
 30 cycles1.86×1021.41×1022.13×1046.50×1011.00×1018.62×1061.10×1015.11×10142.06×1016
 36 cycles4.00×1047.14×1003.38×1096.53×1012.91×1018.08×1044.20×1017.37×102
To correctly analyze the Nyquist plots, the semicircles should be isolated and should related to different phenomena, and a localized analysis should be performed (Ribeiro and Abrantes 2016). The diameters of the semicircles in the Nyquist plots and the slope of the Bode magnitude plots in low frequency of all the representative specimens decreased with the increase in number of wet–dry cycles before the connection of DSAs (Figs. 9 and 10). This indicates that the charge transfer resistance, Rct, decreased with the numbers of wet–dry cycles and freeze–thaw actions, which was consistent with the data in Table 5 and the results of Ribeiro and Abrantes (2016). According to the well-known Stern–Geary equation (Feliu et al. 1998), the corrosion current density increased with the decrease of charge transfer resistance before the connection of DSAs, which suggests that the top rebars of all the representative specimens were about to corrode or had been suffering from corrosion already after a certain number of wet–dry cycles and freeze–thaw actions. The decrease of charge transfer resistance consequently resulted in the decrease of the phase angle in the Bode phase plots (Montemor et al. 2000). After the connection of DSAs, the charge transfer resistance of all the representative specimens except S2-1 increased. This was consistent with the trend of the corrosion current density, indicating that all the representative DSAs except S2-1 could more or less provide some protection to the top rebars of all the representative specimens.
The values of the electrochemical parameters of each specimen are listed in Table 5. The values of Ro were mostly very small and negligible, which coincides with the definition of the corresponding part of the equivalent circuit. The p-values of Ro for each representative specimen were between 0.18 and 0.76, which are larger than 0.05 and suggest that the values of Ro of each representative specimen were not significantly different. Most of the values of Rc were on the order of 100103  Ωcm2 and the p-values of Rc for each representative specimen were around 0.37, also larger than 0.05, indicating that the influences of wet–dry cycles and freeze–thaw actions on the resistance of concrete/mortar with different chloride contamination levels were not significantly different. Compared with Rc, the range of the p values of Qc(Yo) and Qc(n) was from 0.07 to 0.94 and from 0.02 to 0.92, respectively, which suggests that the porosities of the representative specimens were slightly different, likely influenced by wet–dry cycles and freeze–thaw actions. All the values of Rct had a decreasing trend before the connection of DSAs, whereas all the values of Rct except that of S2-1 were increased after the connection of DSAs, which was consistent with the results of depolarized potential and on potential. The p-values of Rct were between 0.18 and 0.99, which statistically validates the effectiveness of all representative anodes except S2-1 and the different performance of cathodic protection of each type of anode. The values of Cdl were calculated as follows (Nazari et al. 2017b):
Cdl=(YoRct1n)1/n
(3)
The value of capacitance depends on the size of the plates, the distance between the plates, and the properties of the dielectric, and the relationship can be expressed as (Nazari et al. 2017b)
Cdl=(εεoA)/δ
(4)
where ε = dielectric constant; εo = vacuum permittivity; A = area of plate surface; and δ = distance between two plates. For S1-1, B1-1, and B2-2, the values of Cdl decreased with the increase of the values of Rct after the connection of DSA, which was consistent with the findings of Nazari et al. (2017b). The decrease of Cdl corresponds to the decrease of the area of plate surface (A) or the increase of the distance between two plates (δ) caused by the repassivation of the rebar. However, there seemed to be a time lag between the connection of DSA and the decrease of Cdl, which was mainly the result of the competition between the corrosion of the rebar and the protection or repassivation by the DSAs. For S2-2, the values of Cdl increased with the decrease of the values of Rct after the connection of DSA, which was consistent with the results of depolarized potential and on potential. The increase of Cdl suggests that the area of the plate surface (A) increased with the process of corrosion due to the absence of effective protection of DSA. The values of Cdl of X1-3 and X2-2 increased with the increase of the values of Rct. A possible explanation for the increasing trend of Cdl of X1-3 and X2-2 is that the time lag of this type of anode is longer than that of the other two types and the value of Cdl decreases when the rebar is completely repassivated or perfectly protected. Further investigations are necessary to validate the hypothesis.

SEM Observations

In Fig. 11, a series of SEM pictures demonstrates the morphology of the matrix of the representative cement mortar specimens after the embedded DSAs lost their effectiveness. There were a significant number of microcracks between the fine aggregates and along the surface of the fine aggregates. This suggests that the cement matrix had seriously deteriorated due to the wet–dry cycles and freeze–thaw cycles. With the increase of the microcracks in the cement matrix, the rebar could get access to more water, oxygen, and deleterious substances such as Cl, which accelerated the corrosion rate. This confirms the hypotheses proposed to explain the change of equivalent circuits for interpreting the EIS results. It also explains why some of the DSAs lost their effectiveness in only 2 years. In addition, the microscopic morphology of the representative DSA samples before and after service are shown in Fig. 12. The DSA samples before service consisted of the matrix and the zinc core, whereas there was a dim layer with a thickness of 100±50  μm between the matrix and the zinc core in the DSA samples after service. Unlike the matrix of DSA samples before service that were tightly attached to the zinc core, there was a crack between the layer and the matrix. This specific microcrack probably was caused by the formation of the dim layer, which was the oxidation product of zinc core.
Fig. 11. SEM observations of the morphology of matrix of representative cement mortar specimens: (a) with Type S DSA; (b) with Type B DSA; and (c) with Type X DSA.
Fig. 12. SEM observations of the morphology of DSAs: (a) before service; and (b) after service.
To further investigate the chemical composition of the dim layer and validate the hypothesis, EDX mapping was conducted on the DSA samples after service. In Fig. 13, the hot spots (i.e., spots with more signals) in the second row represent the distribution of oxygen while the hot spots in the third row represent the distribution of zinc. Obviously, the hot spots of zinc element are intensive in the zinc core and dim layer, whereas the hot spots of oxygen element are intensive in the dim layer and mortar matrix. The overlapping area of hot spots of oxygen and zinc provides convincing evidence that the dim layer is mainly consisted of the oxidation products of zinc. Moreover, Fig. 14 illustrates the EDX spectrum of randomly selected spots in the dim layer, zinc core, and matrix. The major portions of the dim layers of the representative Type S, B, and X DSAs were 81.1%, 64.4%, and 72.9% by weight zinc and 15.7%, 23.2%, and 17.4% by weight oxygen, respectively. Conversely, the major portions of the zinc cores of the representative Type S, B and X DSAs are 94.7%, 96.4%, and 92.4% by weight zinc, and 1.1%, 0.5%, and 1.7% by weight oxygen, respectively. Interestingly, there was some zinc (15% by weight) diffused in the matrix. Generally, the zinc core is fully constrained in the matrix of DSAs, which leaves no space for the oxidation product of zinc core during the service life. This significantly limits the DSA’s potential to continuously protect the connected rebar. Because the resistance of zinc oxide is much greater than that of pure zinc, the potential that is necessary for pumping electrons from the zinc core to the connected rebar will become increasingly large with the growth of the zinc oxide layer, which finally leads to the failure of DSA. Based on the SEM observation, the thickness of the zinc oxide layer was only 100±50  μm, which means that over 90% of the zinc core was wasted. Therefore, to extend the service life and improve the utilization rate of DSAs, some soft and fluid material could be embedded between the zinc core and the matrix that provides flexible room for the oxidation products. In addition, an external electrical current could be connected to the DSA for some period to reduce and rejuvenate partial oxidized zinc and decrease the resistance of the zinc oxide layer. Further studies are needed to validate the feasibility of these methods.
Fig. 13. EDX mapping of the representative DSAs after service: (a) Type S; (b) Type B; and (c) Type X.
Fig. 14. EDX spectrum of the representative DSAs after service: (a) Type S; (b) Type B; and (c) Type X.

XRD Analysis

The XRD spectra of representative samples of DSAs before and after service are depicted in Fig. 15. All types of DSAs have almost the same spectra in Fig. 15(a), indicating that the zinc core in all types of DSAs had high purity. This is consistent with the results of Table 1 and the results of previous studies (Boshkov et al. 2005; Salinas et al. 1999). The small peaks of ZnO of Type X DSAs mean that a tiny fraction of the zinc core of Type X DSAs may be oxidized during the cutting process. The peaks in the remaining figures [Figs. 15(b–d)] indicate the phase composition of DSAs after service, which was Zn(OH)2,Zn5(OH)8Cl2·H2O (ZHC), and ZnO. This is consistent with the results of previous studies (Boshkov et al. 2005; Seré et al. 1998). Compared with the spectra of DSAs before service, all the peaks of Zn of the DSAs after service exhibited a lower intensity. This indicates the consumption of zinc and confirms the effectiveness of all the DSAs that protected the connected section of top rebar.
Fig. 15. XRD spectrum of the representative DSAs before and after service: (a) representative DSAs before service; (b) Type S after service; (c) Type B after service; and (d) Type X after service.

Conclusions

This work evaluated the long-term cathodic protection performance of different types of sacrificial anodes on reinforcements of chloride contaminated mortars and investigated the protecting behavior of the sacrificial anodes through electrochemical impedance spectroscopy. Wet–dry cycles and freeze–thaw actions were conducted to accelerate the corrosion process of reinforcements. Moreover, the feasibility of using graphite electrodes as long-term reference electrodes in concrete was investigated. Based on the experimental results, the following conclusions can be drawn:
Graphite electrodes can be used as reference electrodes in cement matrix as long as the potential evolution is smaller than 100  mV. However, because graphite electrodes are sensitive to environmental changes such as moisture level, the fluctuation of the potential of a graphite electrode could be as large as 250  mV, which makes periodical calibration essential.
The corrosion process of rebar in concrete is a stochastic process, and the corrosion state of rebar should be confirmed by both half-cell potential and corrosion current density. In addition, bound chloride ions play an important role in determining the initiation of corrosion of reinforcements.
To guarantee the effectiveness of a DSA, the depolarized potential by a DSA should be lower than 150  mV. The rebar connected to a DSA could be partially protected or completely protected when the on potential is between 500 and 600  mV or when it is below 600  mV. Compared with the 1% by weight chloride-contamination level, the 2% by weight chloride-contamination level could double the fluctuation level of depolarized potential by Types S and X DSAs. For environmental changes such as wet–dry cycles and freeze–thaw actions, only the fluctuation level of depolarized potential by Type B DSAs increased from 50 to 150  mV after 150 days.
Because the corrosion of rebar in concrete is complex and stochastic, at least two different equivalent circuits are needed to interpret the EIS results before and after the initiation of corrosion of rebar, respectively. The decrease of the slope of Bode magnitude plots in a low frequency region indicates the decrease of charge transfer resistance (Rct), which helps to determine whether to use another equivalent circuit and the corrosion state of the rebar. The decrease and increase of the charge transfer resistance reveal the increase and decrease of the corrosion rate of the rebar, respectively, which identifies the effectiveness of DSAs.
Based on the SEM observation, over 90% of the zinc core in the DSAs is wasted by the end of their service life. The failure of the effectiveness of DSAs is due mainly to the limited designed space (about 100-μm thickness) for oxidization products of the zinc core and the extremely high electrical resistance of zinc oxide, which blocks the transport of electrons from zinc core to rebar. In addition, XRD analysis confirmed the composition of the crystal phase of oxidization products of zinc core, which was Zn(OH)2,Zn5(OH)8Cl2·H2O, and ZnO.

Data Availability Statement

All data, models, and code generated or used during the study appear in the published article.

Acknowledgments

This work was financially supported by a gift fund from Simpson Strong-Tie Company. The authors extend their appreciation to Dr. Mehdi Honarvar Nazari for help with the EIS and data interpretation and to Junliang Wu for help with the measurements of half-cell potential, depolarized potential, and instant-on potential. BASF–United States is acknowledged for donating the water-reducing admixture used in this work. The ideas of improving DSA design presented in this work have been protected by a patent application.

References

Ahmad, S. 2003. “Reinforcement corrosion in concrete structures, its monitoring and service life prediction: A review.” Cem. Concr. Compos. 25 (4): 459–471. https://doi.org/10.1016/S0958-9465(02)00086-0.
Alonso, C., C. Andrade, M. Castellote, and P. Castro. 2000. “Chloride threshold values to depassivate reinforcing bars embedded in a standardized OPC mortar.” Cem. Concr. Res. 30 (7): 1047–1055. https://doi.org/10.1016/S0008-8846(00)00265-9.
Angst, U. M., B. Elsener, C. K. Larsen, and Ø. Vennesland. 2011. “Chloride induced reinforcement corrosion: Electrochemical monitoring of initiation stage and chloride threshold values.” Corros. Sci. 53 (4): 1451–1464. https://doi.org/10.1016/j.corsci.2011.01.025.
ASTM. 2015a. Standard test method for corrosion potentials of uncoated reinforcing steel in concrete. C876. West Conshohocken, PA: ASTM.
ASTM. 2015b. Standard test method for resistance of concrete to rapid freezing and thawing. C666/C666. West Conshohocken, PA: ASTM.
ASTM. 2016. Standard specification for Portland cement. C150/C150M-16e1. West Conshohocken, PA: ASTM.
Basheer, L., J. Kropp, and D. J. Cleland. 2001. “Assessment of the durability of concrete from its permeation properties: A review.” Constr. Build. Mater. 15 (2–3): 93–103. https://doi.org/10.1016/S0950-0618(00)00058-1.
Bertolini, L., F. Bolzoni, P. Pedeferri, L. Lazzari, and T. Pastore. 1998. “Cathodic protection and cathodic preventionin concrete: Principles and applications.” J. Appl. Electrochem. 28 (12): 1321–1331. https://doi.org/10.1023/A:1003404428827.
Bertolini, L., M. Gastaldi, M. Pedeferri, and E. Redaelli. 2002. “Prevention of steel corrosion in concrete exposed to seawater with submerged sacrificial anodes.” Corros. Sci. 44 (7): 1497–1513. https://doi.org/10.1016/S0010-938X(01)00168-8.
Boshkov, N., K. Petrov, D. Kovacheva, S. Vitkova, and S. Nemska. 2005. “Influence of the alloying component on the protective ability of some zinc galvanic coatings.” Electrochim. Acta 51 (1): 77–84. https://doi.org/10.1016/j.electacta.2005.03.049.
Bouteiller, V., C. Cremona, V. Baroghel-Bouny, and A. Maloula. 2012. “Corrosion initiation of reinforced concretes based on Portland or GGBS cements: Chloride contents and electrochemical characterizations versus time.” Cem. Concr. Res. 42 (11): 1456–1467. https://doi.org/10.1016/j.cemconres.2012.07.004.
Du, S., H. Zha, Y. Ge, Z. Yang, and X. Shi. 2017. “Laboratory investigation into the modification of transport properties of high-volume fly ash mortar by chemical admixtures.” J. Materials Civ. Eng. 29 (10): 04017184.
Duffó, G. S., S. B. Farina, and C. M. Giordano. 2009. “Characterization of solid embeddable reference electrodes for corrosion monitoring in reinforced concrete structures.” Electrochim. Acta 54 (3): 1010–1020. https://doi.org/10.1016/j.electacta.2008.08.025.
Dugarte, M., A. A. Sagüés, R. G. Powers, and I. R. Lasa. 2007. Evaluation of point anodes for corrosion prevention in reinforced concrete. Houston: NACE International.
Feliu, V., J. A. González, C. Andrade, and S. Feliu. 1998. “Equivalent circuit for modelling the steel-concrete interface. Part I: Experimental evidence and theoretical predictions.” Corros. Sci. 40 (6): 975–993. https://doi.org/10.1016/S0010-938X(98)00036-5.
Glass, G. K., and N. R. Buenfeld. 1997. “The presentation of the chloride threshold level for corrosion of steel in concrete.” Corros. Sci. 39 (5): 1001–1013. https://doi.org/10.1016/S0010-938X(97)00009-7.
Glass, G. K., and N. R. Buenfeld. 2000. “The influence of chloride binding on the chloride induced corrosion risk in reinforced concrete.” Corros. Sci. 42 (2): 329–344. https://doi.org/10.1016/S0010-938X(99)00083-9.
Liu, Y., and X. Shi. 2009. “Cathodic protection technologies for reinforced concrete: Introduction and recent developments.” Rev. Chem. Eng. 25 (5–6): 339–388. https://doi.org/10.1515/REVCE.2009.25.5-6.339.
Martínez, I., and C. Andrade. 2008. “Application of EIS to cathodically protected steel: Tests in sodium chloride solution and in chloride contaminated concrete.” Corros. Sci. 50 (10): 2948–2958. https://doi.org/10.1016/j.corsci.2008.07.012.
Montemor, M. F., A. M. P. Simões, and M. M. Salta. 2000. “Effect of fly ash on concrete reinforcement corrosion studied by EIS.” Cem. Concr. Compos. 22 (3): 175–185. https://doi.org/10.1016/S0958-9465(00)00003-2.
Moreno, M., W. Morris, M. G. Alvarez, and G. S. Duffó. 2004. “Corrosion of reinforcing steel in simulated concrete pore solutions: Effect of carbonation and chloride content.” Corros. Sci. 46 (11): 2681–2699. https://doi.org/10.1016/j.corsci.2004.03.013.
Muralidharan, S., T.-H. Ha, J.-H. Bae, Y.-C. Ha, H.-G. Lee, K.-W. Park, and D.-K. Kim. 2006. “Electrochemical studies on the solid embeddable reference sensors for corrosion monitoring in concrete structure.” Mater. Lett. 60 (5): 651–655. https://doi.org/10.1016/j.matlet.2005.09.058.
Nazari, M. H., X. Shi, E. Jackson, Y. Zhang, and Y. Li. 2017a. “Laboratory investigation of washing practices and bio-based additive for mitigating metallic corrosion by magnesium chloride deicer.” J. Mater. Civ. Eng. 29 (1): 04016187. https://doi.org/10.1061/(ASCE)MT.1943-5533.0001727.
Nazari, M. H., M. S. Shihab, L. Cao, E. A. Havens, and X. Shi. 2017b. “A peony-leaves-derived liquid corrosion inhibitor: Protecting carbon steel from NaCl.” Green Chem. Lett. Rev. 10 (4): 359–379. https://doi.org/10.1080/17518253.2017.1388446.
Rajani, B., and Y. Kleiner. 2003. “Protecting ductile-iron Water Mains: What protection method works best for what soil condition?” J. Am. Water Works Assoc. 95 (11): 110–125. https://doi.org/10.1002/j.1551-8833.2003.tb10497.x.
Ribeiro, D. V., and J. C. C. Abrantes. 2016. “Application of electrochemical impedance spectroscopy (EIS) to monitor the corrosion of reinforced concrete: A new approach.” Constr. Build. Mater. 111 (May): 98–104. https://doi.org/10.1016/j.conbuildmat.2016.02.047.
Sagüés, A. A., V. Balakrishna, and R. G. Powers. 2005. “An approach for the evaluation of performance of point anodes for corrosion prevention of reinforcing steel in concrete repairs.” In Proc., fib Symp. Structural Concrete and Time, 35–41. Symposium Secretariat: fib Symposium Argentina.
Salinas, D. R., S. G. García, and J. B. Bessone. 1999. “Influence of alloying elements and microstructure on aluminium sacrificial anode performance: Case of Al–Zn.” J. Appl. Electrochem. 29 (9): 1063–1071. https://doi.org/10.1023/A:1003684219989.
Seré, P. R., M. Zapponi, C. I. Elsner, and A. R. Di Sarli. 1998. “Comparative corrosion behaviour of 55Aluminium–zinc alloy and zinc hot-dip coatings deposited on low carbon steel substrates.” Corros. Sci. 40 (10): 1711–1723. https://doi.org/10.1016/S0010-938X(98)00073-0.
Sergi, G. 2011. “Ten-year results of galvanic sacrificial anodes in steel reinforced concrete.” Mater. Corros. 62 (2): 98–104. https://doi.org/10.1002/maco.201005707.
Song, H.-W., C.-H. Lee, and K. Y. Ann. 2008. “Factors influencing chloride transport in concrete structures exposed to marine environments.” Cem. Concr. Compos. 30 (2): 113–121. https://doi.org/10.1016/j.cemconcomp.2007.09.005.
Song, H.-W., and V. Saraswathy. 2007. “Corrosion monitoring of reinforced concrete structures: A review.” Int. J. Electrochem. Sci. 2 (Jan): 1–28.
Trocónis de Rincón, O., Y. Hernández-López, A. de Valle-Moreno, A. A. Torres-Acosta, F. Barrios, P. Montero, P. Oidor-Salinas, and J. R. Montero. 2008. “Environmental influence on point anodes performance in reinforced concrete.” Constr. Build. Mater. 22 (4): 494–503. https://doi.org/10.1016/j.conbuildmat.2006.11.014.
Wang, Z., Q. Zeng, L. Wang, Y. Yao, and K. Li. 2014. “Corrosion of rebar in concrete under cyclic freeze–thaw and Chloride salt action.” Constr. Build. Mater. 53 (Feb): 40–47. https://doi.org/10.1016/j.conbuildmat.2013.11.063.
Yang, Z., W. J. Weiss, and J. Olek. 2006. “Water transport in concrete damaged by tensile loading and freeze–thaw cycling.” J. Mater. Civ. Eng. 18 (3): 424–434. https://doi.org/10.1061/(ASCE)0899-1561(2006)18:3(424).
Yu, H., X. Shi, W. H. Hartt, and B. Lu. 2010. “Laboratory investigation of reinforcement corrosion initiation and chloride threshold content for self-compacting concrete.” Cem. Concr. Res. 40 (10): 1507–1516. https://doi.org/10.1016/j.cemconres.2010.06.004.
Yuan, Q., C. Shi, G. De Schutter, K. Audenaert, and D. Deng. 2009. “Chloride binding of cement-based materials subjected to external chloride environment: A review.” Constr. Build. Mater. 23 (1): 1–13. https://doi.org/10.1016/j.conbuildmat.2008.02.004.

Information & Authors

Information

Published In

Go to Journal of Materials in Civil Engineering
Journal of Materials in Civil Engineering
Volume 32Issue 11November 2020

History

Received: Mar 15, 2019
Accepted: Mar 24, 2020
Published online: Aug 25, 2020
Published in print: Nov 1, 2020
Discussion open until: Jan 25, 2021

Authors

Affiliations

Jialuo He, S.M.ASCE [email protected]
Graduate Research Assistant, Laboratory of Corrosion Science and Electrochemical Engineering, Dept. of Civil and Environmental Engineering, Washington State Univ., P.O. Box 642910, Pullman, WA 99164-2910. Email: [email protected]
Associate Professor, Laboratory of Corrosion Science and Electrochemical Engineering, Dept. of Civil and Environmental Engineering, Washington State Univ., P.O. Box 642910, Pullman, WA 99164-2910 (corresponding author). ORCID: https://orcid.org/0000-0003-3576-8952. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share