13

Corrosion sensing for assessing and monitoring civil infrastructures

A. Poursaee,    Clemson University, USA

Abstract:

Concrete structures with steel-reinforced rebar are in a continuous and losing battle with corrosive elements that naturally occur from long-term exposure to an aggressive environment. Additionally, because such structures are now required to perform longer than mandated by their original design, many are now woefully obsolete. Clear understanding of the condition of such structures is crucial, from the points of view of both economy and safety. Prediction of the deterioration of steel-reinforced concrete structures due to corrosion and adapting the accurate maintenance protocol are difficult tasks, which would be enhanced by using different corrosion measurement techniques and corrosion sensors. This chapter describes some corrosion measurement methods and their applications in corrosion sensors.

Key words

corrosion measurement; corrosion sensors; concrete deterioration; service life

13.1 Introduction

The low cost of steel-reinforced concrete, and the ready availability of raw materials with which it is formed, make it the most widely used structural material available. The durability of structures made from steel-reinforced concrete is related to its ability to impede, or greatly reduce, the rate of moisture transport or the ingress of aggressive ions. However, because many existing concrete structures are under constant degradation from hostile environments, they suffer from durability issues, particularly the corrosion of the steel bars within such structures. This corrosion, in turn, has created a multi-billion-dollar infrastructure deficit, the annual direct cost of which, in the US, is estimated at $22.6 billion a year (Koch et al., 2001) and growing. The corrosion of the reinforcing steel in concrete is a serious problem, from the perspectives of both safety and economy, which directly can affect the sustainability of the infrastructure. Indeed, the Federal Highway Administration (FHWA) report on corrosion protection of concrete bridges estimated the average annual cost through the year of 2011 for just maintaining the US bridge infrastructure to be $5.2 billion (Yunovich et al., 2001)

13.2 Principles of corrosion

Concrete provides physical corrosion resistance to the steel reinforcement by acting as a barrier, and chemical corrosion resistance due to its high pH. Concrete that is not exposed to any external influences usually exhibits a pH between 12.5 and 13.5 (Hansson, 1984). As shown in the Pourbaix diagram (Fig. 13.1) that defines the range of electrochemical potential and pH for Fe-H2O system in the alkaline environment, at potentials and pH normal within the concrete a protective passive layer forms on the surface of steel. It is believed that this layer is an ultra-thin (< 10 nm), protective oxide or hydroxide film that decreases the anodic dissolution rate to negligible levels (Zakroczymski et al., 1985; Montemor et al., 1998; Carnot et al., 2002). However, reinforcing steel does corrode. The partial or complete loss of the passive layer, known as depassivation, leads to corrosion of the steel bars.

image
13.1 Pourbaix diagram for Fe-H2 O at 25 °C (Pourbaix, 1974).

As can be seen in Fig. 13.2, the corrosive products of iron are expansive, and their formation can cause cracking and further deterioration in the concrete, which ultimately reduces the service life and causes safety problems in the infrastructure.

image
13.2 Corrosion products of iron (Lide, 1999).

Formation of passive film on iron begins with dissolution of the metal, which produces electrons and the reduction of oxygen that uses those electrons. The ferrous ions from the anodic dissolution of iron are attracted to the cathodic part of the steel and combined with hydroxide ions from the cathodic reaction of oxygen to form the ferrous hydroxide. If this film is exposed to the oxygen, other passive oxide layers, such as Fe3O4 or Fe2O3, may form on the outer surface of the film. The protective nature of this layer can be reduced, and the result would be active corrosion of steel in concrete. Chloride ions, mostly from de-icing salts or seawater, and carbon dioxide, from the atmosphere, are two major factors that can break the passive film on the surface of steel and initiate corrosion, and the mechanism will be discussed in the following sections. Insufficient oxygen to preserve the passive film, galvanic cell formation from the contact of different metals, and stray currents are the other factors that may also cause active corrosion in reinforcing steel structures.

13.2.1 Chloride induced corrosion

Chloride ions can be present in the concrete due to the use of chloride-contaminated components or the use of CaCl2 as an accelerator when mixing the concrete, or by diffusion into the concrete from the outside environment (Thuresson, 1996). A localized breakdown of the passive layer occurs when sufficient amounts of chlorides reach reinforcing bars, and the corrosion process is then initiated. Chlorides in concrete can be either dissolved in the pore solution (free chlorides) or chemically and physically bound to the cement hydrates and their surfaces (bound chlorides). Only the free chlorides dissolved in the pore solution are responsible for initiating the process of corrosion (Martin-Perez et al., 2000).

There are three theories about the chloride attack (ACI Committee 222, 1996):

1. Chloride ions penetrate to the oxide film on steel through pores or defects in the film is easier than the penetration of other ions.

2. Chloride ions are adsorbed on the metal surface in competition with dissolved O2 or hydroxyl ions.

3. Chloride ions compete with hydroxyl ions for the ferrous ions produced by corrosion, and a soluble complex of iron chloride forms, which can diffuse away from the anode, destroying the protective layer of Fe(OH)2 and permitting corrosion to continue.

13.2.2 Carbonation induced corrosion

When concrete is exposed to air, the calcium hydroxide reacts with water and carbon dioxide in the air:

CaOH2+CO2CaCO3+H2O [13.1]

image [13.1]

The effect of carbonation is to reduce the pH value of the surface layer of the concrete to less than 8.3. This pH is sufficient to make the passive layer on the reinforcement rebar unstable (Allen and Forrester, 1983). The process of carbonation can be summarized in the following steps:

1. CO2 diffuses into concrete.

2. In the presence of moisture, it reacts with CH (calcium hydroxide).

3. The pH decreases from about 12.6 to 8.3.

4. At this pH, steel is not passive.

13.2.3 Mechanism of corrosion in reinforced concrete

Corrosion is an electrochemical reaction that consists of anodic and cathodic half-cell reactions. Micro-cell corrosion is the term given to the situation where active dissolution and the corresponding cathodic half-cell reaction take place at adjacent parts of the same metal part. For a steel reinforcing bar (rebar) in concrete, this process always occurs in practice. The surface of the corroding steel can act as a mixed electrode containing both anode and cathode regions that are connected by the bulk steel. Macro-cell corrosion can also form on a single bar exposed to different environments within the concrete, or where part of the bar extends outside the concrete. In both cases, concrete pore solution functions as an electrolyte.

For steel embedded in concrete, based on the pH of the concrete (electrolyte) and the presence of aggressive ions, the following would be the possible anodic reactions (Hansson, 1984; Ahmed, 2003):

3Fe+4H2OFe3O4+8H++8e [13.2]

image [13.2]

2Fe+3H2OFe3O3+6H++6e [13.3]

image [13.3]

Fe+2H2OHFeO2+3H++2e [13.4]

image [13.4]

FeFe2++2e [13.5]

image [13.5]

The possible cathodic reactions depend on the availability of O2 and on the pH near the steel surface. The most likely reactions are as follows (Hansson, 1984; Ahmed, 2003):

2H2O+O2+4e4OH [13.6]

image [13.6]

2H2O+2eH2+2OH [13.7]

image [13.7]

13.3 Corrosion evaluation techniques

Corrosion consists of electrochemical reactions at the interface between the metal and an electrolyte solution. During the anodic reaction, a metal is oxidized and releases electrons. These electrons are consumed by the cathodic reaction in which the reduction occurs. By equating these two reactions, a corrosion current, Icorr, which is the absolute value of the corrosion rate, and half-cell potential (also called corrosion potential or open circuit potential), Ecorr, which is the probability of corrosion, can be found.

Ecorr is equivalent to the voltage of a cell or battery versus a reference electrode under no-load conditions, and can be measured with a high impedance voltmeter or potentiometer (Elsener et al., 2003; Corrosion Doctors, 2005). Icorr cannot be measured directly, but it can be estimated using electrochemical techniques, while Ecorr must be determined as the potential difference between the metal surface and a reference electrode.

As mentioned, corrosion of steel in concrete occurs via electrochemical reactions. Therefore, electrochemical techniques are ideal for the study of the corrosion processes. Usually, in electrochemical measurements, a cell consists of a working electrode (the corroding metal), a counter electrode, a reference electrode, and electrolyte. All of the electrodes are connected to a potentiostat, which allows the potential of the metal to be changed in a controlled manner and the resultant current flow to be measured as a function of potential. This changing of the potential is called ‘polarization.’ When the polarization is done potentiostatically (controlled by potential), the current is measured, and when it is done galvanostatically (controlled by current), the potential is measured (Fontana, 1987; Jones, 1992; Gamry Instruments, 2005).

13.3.1 Half-cell potential technique

The half-cell potential technique is the most widely used technique to evaluate the corrosion of the steel rebars in concrete. It was introduced in the 1970s by Richard F. Stratfull in North America and by the Danish Corrosion Center in Europe (Stratfull 1968,1972; FORCE Technology, 2004). In 1980, the test was approved as a standard by the ASTM (ASTM, 2009). This technique is based on measuring the electrochemical potential of the steel rebar with respect to a standard reference electrode (copper/copper sulfate electrode (CSE) is suggested by the ASTM-C876) placed on the surface of the concrete, and can provide an indication of the corrosion risk of the steel. Figure 13.3 shows the basics of half-cell potential measurement.

image
13.3 Apparatus for half-cell potential method described in ASTM C876 to measure surface potential associated with corrosion current.

According to the ASTM Standard recommended guidelines for interpretation of the results, if the measured potential is more negative than − 350 mV versus CSE, the probability of active corrosion is more than 90%. If the measured potential is more positive than − 200 mV versus the CSE, the probability of not having active corrosion is 90%, and between − 200 and − 350 mV is the uncertainty region. The most common way of presenting the half-cell potential data is plotting the potential distribution or potential mapping contour. It should be emphasized that, since half-cell potential value is defined as the thermodynamic measure of the ease of removing electrons from the metal in steady state condition, it cannot be used as direct measurement of corrosion rate. It should be noted that half-cell potential is the probability of corrosion activity, while Icorr is the direct measurement of corrosion rate. A simple comparison of the half-cell potential data with the ASTM guidelines on steel reinforcement corrosion probability could cause mistakes in the evaluation of the structure. It has been accepted by those who work in the field that a more negative reading of potential means a higher probability of corrosion. However, this general rule may not always be correct. Some precautions are necessary in interpreting the data from half-cell potential measurements, because there are many factors that may affect the magnitude of the potentials. Details about difficulties with the half-cell potential method are discussed in Poursaee and Hansson (2009).

13.3.2 Linear polarization resistance (LPR)

Figure 13.4 shows a schematic plot of the relationship between potential and current in the region of the open circuit potential. The curve plots the applied potential versus measured current, and vice versa. As shown in Fig. 13.4, there is an approximately linear region around the open circuit potential. The LPR measurements are performed by applying a potential in the range of ± 10 mV about the Ecorr, either as a constant pulse (potentiostatic) or a potential sweep (potentiodynamic) and measuring the current response. Alternatively, a current pulse (galvanostatic) or a current sweep (galvano-dynamic) can be applied, and potential response is measured. Polarization resistance (Rp) is the resistance of the specimen to oxidation while an external potential is applied, and the corrosion rate, which is inversely related to the Rp, can be calculated from it.

image
13.4 Schematic illustration of the linear polarization curve.

Rp is determined by calculating the slope of this linear region:

Rp=ΔEΔI [13.8]

image [13.8]

where ΔE = change in potential and ΔI = change in current. The Stern-Geary equation relates corrosion current to Rp (Stern and Geary, 1957):

Icorr=BRp [13.9]

image [13.9]

B=βaβc2.3βa+βc [13.10]

image [13.10]

The corrosion current density, icorr, can be calculated by dividing the corrosion current (Icorr) by the surface area of the polarized area (A):

icorr=BRpA [13.11]

image [13.11]

B is the Stern-Geary constant, and βa and βc are anodic and cathodic Tafel constants, respectively. The value of B should be determined empirically. However, for most cases, it can be assumed to be 0.026 V for active and 0.052 V for passive corrosion of steel in concrete (Andrade and González, 1978; Andrade et al., 1990).

The resistance measured by the LPR is actually the sum of the polarization resistance, Rp, and the electrolyte resistance, RΩ. If Rp >> RΩ, the resistance that is measured by the LPR is close enough to the polarization resistance, which can be used as the actual value. However, in some environments with low conductivity, and/or high corrosion rates, the RΩ is significant and should be considered (Jones, 1992).

According to some researchers, corrosion current densities over ~ 1 μA/cm2 have been identified as the level of high corrosion risk, and corrosion current density below 0.1 μA/cm2 are characterized as passive corrosion in the system (Gonzalez et al., 1995; Alonso et al., 2000; Polder and Peelen, 2002). However, it seems that the equipment used by these researchers generally gives lower values than does other commercial equipment (Gepraegs, 2002). Therefore, applying such definitions may over- or under-estimate the corrosion rate and cause errors in evaluation and life prediction. Interpreting the corrosion current density values of embedded steel bars in concrete, obtained from the LPR technique, is difficult in large part because determining the actual corroding area of steel is almost impossible and usually causes underestimation of the actual corrosion current density in the areas of active corrosion. The LPR has some advantages over the other measurement techniques, which makes it popular in the evaluation of the corrosion rate in reinforced concrete: it is a non-destructive technique; it is a simple method and it usually needs only a few minutes for corrosion rate determination. Because of its rapidity, it can be effectively used in kinetic studies of corrosion monitoring (Jones, 1992).

13.3.3 Potentiostatic LPR

In the potentiostatic LPR technique, a constant potential signal (usually ± 10 mV) is applied for a certain period of time, which is determined by the time for the current to reach steady state, in the form of square wave between the working electrode (steel bar in concrete) and reference electrode, and the response current is measured. By using Equation [13.8], the RP. and consequently the corrosion current density and corrosion rate, can be calculated.

13.3.4 Galvanostatic pulse technique

The galvanostatic pulse technique was introduced for field application in 1988 (Newton and Sykes, 1988). This method is a rapid non-destructive polarization technique. A short-time anodic current pulse is applied galvanostatically between a counter electrode placed on the concrete surface and the rebar. The applied current is usually in the range of 10–100 μA, and the typical pulse duration is between 5 and 30 s. The reinforcement is anodically polarized and the resulting change of the electrochemical potential of the reinforcement is measured with a reference electrode, which is usually in the center of the counter electrode and recorded as a function of polarization time (Klinghoffer, 1995; Elsener et al., 1997). A typical potential response for a corroding reinforcement is shown in the Fig. 13.5.

image
13.5 Schematic illustration of galvanostatic pulse results.

When the constant current, Iapp, is applied to the rebar, the polarization of the rebar, ηt, at given time t can be expressed as (Jones and Greene, 1966):

ηt=Iapp×Rp×1etRpCdl+RΩ [13.12]

image [13.12]

where Rp = polarization resistance, Cdl = double-layer capacitance, and RΩ = ohmic resistance of the concrete cover.

By transferring Equation [13.12] to logarithmic form, the values of Rp and Cdl, can be calculated as following (Newton and Sykes, 1988):

InηmaxηtInIapp×RptRpCdl [13.13]

image [13.13]

where ηmax is the final steady state potential value. Figure 13.6 plots Equation [13.13].

image
13.6 Schematic illustration of Equation [13.13] (Newton and Sykes, 1988).

If this straight line is extrapolated to t = 0, it will give an intercept of Iapp × Rp and the slope of the line in 1/RpCdl. The remaining over-potential corresponds to Iapp × RΩ, which is the ohmic voltage drop across the concrete cover. After determining the polarization resistance (Rp) by the above method, the corrosion current Icorr can be calculated from the Stern-Geary equation (Stern and Geary, 1957; Newton and Sykes, 1988).

13.3.5 Equipment with the guard ring

An auxiliary counter electrode, referred to as a guard ring electrode, can be used to confine the polarization to a known length of reinforcing bar. The counter electrode and the guard ring are typically arranged as annular metal rings with a reference electrode in the center. The potential or current applied from the guard ring tends to repel the signals from the central counter electrode, confining them to an area of the structure located approximately under the counter electrode, as schematically shown in Plate X in the color section between pages 294 and 295.

There are two well-known equipment works based on the galvanostatic pulse technique:

1. The GECOR is considered a galvanostatic pulse technique instrument although it is often referred to as an LPR device (Broomfield, 1996; Feliu et al., 1996; Newhouse and Weyers, 1996). The electrode assembly has a total of three reference electrodes, one located in the center and two located between the counter electrode and guard ring, and they are used to adjust the guard electrode current to maintain the potential difference between the two reference electrodes constant during the polarization procedure.

2. The GalvaPulse is a galvanostatic pulse instrument developed by the FORCE Institute in Denmark. Signal confinement is over a 70 mm length of bar, and measurements can be made with or without the guard ring electrode. The measuring cell has an Ag/AgCl reference electrode at the center with a zinc counter electrode and a zinc guard ring. Based on the suggestion of the manufacturer, 10–20 μA for 5–10 s, in passive areas, should give a reasonable polarization of the reinforcement. The recommended applied current pulse in active areas is 80–100 μA for 5–10 s.

13.3.6 Electrochemical impedance spectroscopy (EIS)

The EIS studies the system response to the application of a small amplitude alternating potential or current signal at different frequencies. The popularity of the EIS methods for reinforced concrete has increased remarkably in recent years, because analysis of the system response provides information about the double-layer capacitance, interface, structure, reactions which are taking place, corrosion rate, and electrolyte (environment) resistance (Silverman, 1990; Jones, 1992; Lasia, 1999). An electrochemical process can be considered as an electrical circuit with basic elements such as resistors, capacitors, and inductors. Therefore, in interpreting the response to an alternating current (AC), the AC circuit theory can be used successfully to demonstrate a corrosion process, and it may also be used to understand the behavior of the corrosion process and predict the corrosion rates.

In direct current, Ohm’s law is as follows:

V=IR [13.14]

image [13.14]

(V = Potential, I = Direct current, R = Actual resistor)

In the AC, Ohm’s law becomes:

V=IZ [13.15]

image [13.15]

(V = Potential, I = Alternative current, Z = Impedance)

Direct current can be viewed as alternating current at zero frequency. In this case, the resistance is composed of only one or more actual resistors. When the frequency is not zero, all circuit elements that can affect the flow of current, e.g., resistors, capacitors, and inductors, cause the resistance. The resistance created by capacitors and inductors depends on the frequency, while that created by a resistor is not dependent on frequency (Silverman, 1986). A sinusoidal current or voltage can be represented as a rotating vector, as shown in Fig. 13.7. In this figure, the x-component shows the observed current so it becomes the real component of the rotating vector, while the contribution of the y-component is not observed; there- fore, it is called the imaginary component of the rotating vector.

image
13.7 Relationship between sinusoidal AC current and rotating vector representation.

The mathematical descriptions of the two components are as follows:

Realcurrent=Ix=Icosωt [13.16]

image [13.16]

Imaginarycurrent=Iy=Isinωt [13.17]

image [13.17]

where t = time and ω = frequency in radians per second = 2πf (f = frequency in Hertz).

To separate the real (x) and imaginary (y) components, the magnitude of the imaginary part should be multiplied by j=1image 1 and then the real and imaginary values can be reported separately. The equations for AC impedance become:

Etotal=Ereal+Eimaginary+E+jE [13.18]

image [13.18]

Itotal=Ireal+Iimaginary=I+jI [13.19]

image [13.19]

Ztotal=Z+Z=E+jEI+jI [13.20]

image [13.20]

The absolute amplitude of the impedance (that is the length of the vector) and the phase angle are defined by (Princeton Applied Research, 2006):

Z=Z+Z2 [13.21]

image [13.21]

tanθ=ZZ [13.22]

image [13.22]

The goal of AC impedance is to measure the impedance Z as Z′ and Z″, and then model the response by using an equivalent simple circuit (Silverman, 1986).

Data presentation

There are different ways to illustrate the response of an electrochemical system to an applied AC potential or current. The most common plots are the Nyquist plot and Bode plot. If, at each excitation frequency, the real part is plotted on the x-axis and the imaginary part is plotted on the y-axis of a chart, a ‘Nyquist plot’ is formed. A simple corroding system can be assumed as: solution resistance, in series with a combination of a resistor and a capacitor, which represent the polarization resistance and double-layer capacitance, respectively. This simple representation is called a Randles cell and is shown in Fig. 13.8.

image
13.8 Equivalent circuit for a simple electrochemical system.

Figure 13.9 schematically illustrates the Nyquist plot for a simple electro-chemical system corresponding to the analog circuit in Fig. 13.8. It should be noted that each point on the Nyquist plot is the impedance at one frequency. On the Nyquist plot, the impedance can be represented as a vector of length |Z|, and the angle between this vector and the x-axis is the phase angle ‘θ’ (Gamry Instruments, 2006; Princeton Applied Research, 2006). At high frequencies, at the leftmost end of the semicircle, where the semicircle touches the x-axis, the impedance of the Randles cell is entirely produced by the ohmic resistance, RΩ. The frequency reaches its lower limit at the rightmost end of the semicircle. At this frequency, the Randles cell also approximates a pure resistance, but now the value is (RΩ + Rp) (Princeton Applied Research).

image
13.9 Nyquist plot for a simple electrochemical system.

The Nyquist plot has some limitations (Princeton Applied Research):

1. The frequency is not clearly shown on the plot and it is not possible to determine, for a specific point, the frequency used to record that point;

2. The ohmic and polarization resistances can be directly determined from the plot but the electrode capacitance can be only calculated if the frequency information is known, using Equation [13.23]:

C=1ωmax×Rp [13 23]

image [13 23]

3. If there are high and low impedance components in the circuit, the larger impedance controls plot scaling, and distinguishing the low impedance semicircle would probably be impossible.

A Bode plot is another popular presentation method for the impedance data. In the Bode plot, the data are plotted with log of frequency on the abscissa and both the log of absolute value of the impedance (|Z|) and phase-shift (θ) on the ordinate (Gamry Instruments, 2005). Figure 13.10 schematically shows a Bode plot for the same system shown in Fig. 13.8. Since the frequency appears as one of the axes in the Bode plot, it is easy to understand the dependence of impedance on the frequency from the plot. The log |Z| versus log ω curve can be used to determine the values of Rp and RΩ. At very high and very low frequencies, |Z| becomes independent of frequency. At the highest frequencies the ohmic resistance controls the impedance and log (RΩ) can be read from the high frequency horizontal level. On the other hand, at the lowest frequencies, polarization resistance contributes, and log (Rp + RΩ) can be read from the low frequency horizontal portion.

image
13.10 Bode plot for the same system shown in Fig. 13.9.

The Bode format is advantageous when data scatter prevents satisfactory fitting of the Nyquist semicircle. In general, the Bode plot provides a more understandable description of the frequency-dependent behavior of the electrochemical system than does the Nyquist plot, in which frequency values are not clear (Princeton Applied Research, 2006).

13.3.7 Cyclic polarization

Cyclic potentiodynamic polarization technique is a relatively non-destructive measurement that can provide information about the corrosion rate, corrosion potential, susceptibility to pitting corrosion of the metal, and concentration limitation of the electrolyte in the system. The technique is built on the idea that prediction of the behavior of a metal in an environment can be made by forcing the material from its steady state condition and monitoring how it responds to the force as the force is removed at a constant rate and the system is reversed to its steady state condition. Applied potential is the force and is raised at a continuous, often slow, rate by using a potentiostat (Silverman, 1998). This rate is called the polarization scan rate, and is an experimental parameter. The potential of the specimen is changed continuously while the resulting current is monitored, and then the applied potential is plotted versus the logarithm of the resulting current density. The conductivity of the electrolyte (environment) is a very important factor that should be considered in all electrochemical experiments, especially in the cyclic polarization technique. The electrolyte resistance causes a potential drop between the working electrode and reference electrode and can cause errors. This effect has important impacts on the interpretation, and should be compensated.

13.4 Corrosion sensors for field monitoring

Application of sensors to monitor the corrosion activity of steel in concrete and to measure the corrosion rate of steel has been of great attraction. The embeddable corrosion sensors can provide early warning of conditions that damage steel reinforcement, leading to cracking and deterioration of concrete structures. In addition, lost cost, and no or minimal need for trained personnel and maintenance, make the application of sensors even more attractive.

13.4.1 Electrochemical sensors

The sensor most used for this purpose is basically an embeddable inert and stable reference electrode in high alkaline environment in concrete such as: Mn/MnO2 or Ti/TiO2 (Muralidharan et al., 2007; Duffó and Farina, 2009; Dong et al., 2011), to perform half-cell potential measurement. The other type of sensor is designed to measure the half-cell potential and the corrosion current density, and consequently corrosion rate. These sensors are based on utilizing three electrodes (reference, counter or auxiliary, and working electrodes), which are usually used in conventional electrochemical measuring systems. Most of the sensors in this category use the metal–metal oxide (MMO) as reference electrodes, such as Mn/MnO2 and Ti/TiO2, stainless steel as counter electrode, and the reinforcing bar as the working electrode (Anderade and Martinez, 2009; Duffó and Farina, 2009; Dong et al., 2011). In the sensor developed by Poursaee, graphite rods were used as both reference and counter electrodes. The stability of graphite in high pH was evaluated, and it was concluded that the potential of the graphite is stable enough during the period of measurement, and the potentiostatic LPR can be performed successfully, using this configuration (Poursaee, 2009). However, graphite cannot be used as the permanent reference electrode, unless it undergoes special treatment (Swette et al., 1999). The sensors with the three electrochemical components can be used to perform all the techniques, such as the LPR, galvanostatic pulse, cyclic polarization, and the EIS.

Electrochemical sensors: some examples

The embedded corrosion instrument (ECI)

ECI – designed and manufactured by the Virginia Technologies (Virginia Technologies Inc, 2012), monitors five key factors in corrosion of the reinforcing steel: LPR, open circuit potential, resistivity, chloride ion concentration, and temperature. The ECI is designed to monitor bridges, buildings, dams, erosion control structures, flood control channels, parking garages, piers, pylons, roadways, and spillways. The ECI consists of a stainless steel conductivity sensor, a stainless steel counter electrode, a mild steel working electrode, a Mn/MnO2 reference electrode, and an Ag/AgCl chloride sensor. The ECI corrosion monitor is combined with a NetCon-10 interface module. The NetCon-10 is a connection module that helps in organizing large networks of corrosion monitors and guarding them against voltage spikes.

900 concrete multi-depth sensor

The multi-depth sensors, designed by the Rohrback Cosasco Systems (Rohrback Cosasco Systems, 2012) may be used to assess the depth of chloride or carbonation ingress, and the instantaneous corrosion rate. The multi-depth sensor has four galvanic couples of mild steel and stainless steel. These couples are located at four depths from the concrete surface. The couple furthest from the surface should be placed above the rebar. A zero resistance ammeter (ZRA) is used to measure the current flow between the two electrodes. An increase in current flow indicates the ingress of chloride contamination and increased corrosivity at that electrode level. The multi-depth sensor can also measure the instantaneous corrosion rate of steel in concrete by using the LPR method. The electrodes of the sensor use adjacent carbon steel elements for the LPR measurement. It seems that the couple of mild steel and stainless steel can be used as the counter and reference electrodes, respectively. The sensors can be monitored frequently or continuously to track changes in corrosion rate. It is suggested by the manufacturer that the multi-depth sensors be positioned at the most susceptible locations for corrosion, adjacent to the rebar but on the side that will see chloride or moisture ingress first. This will allow precautionary measures to be taken before the onset of corrosion.

CorrWatch multi-sensor

The CorrWatch multi-sensor, which is developed by FORCE Technology (FORCE Technology, 2012) can be embedded in new concrete structures. This probe can measure most parameters necessary for assessment of the corrosion state. The CorrWatch consists of four mild steel (anode) probes and the measuring electrodes in variable positions but with a known distance to the reinforcement and a measuring electrode of noble metal (cathode). Anode heights are flexible and can be adjusted to fit the concrete covering thickness. In order to predict when the reinforcement starts corroding, the current between the individual anodes and the cathode is measured, either by a specially designed handheld ZRA or a specially developed data logger.

SensCore corrosion sensor

The SensCore corrosion sensor, designed by the The Roctest Group (The Roctest Group, 2012) measures both corrosion initiation and corrosion rate. Those two measurements are performed at four different depths, between the concrete surface and the reinforcement bars depth, which allows the evaluation of the corrosion front progression. Humidify sensor is also combined with the SensCore to provide a complete picture of the corrosion initiation and progression in a reinforced concrete structure. The SensCore corrosion sensor is composed of four mild steel rebars that are secured to a stainless steel support. The probe is then placed in the concrete and the SensCore data logger measures the corrosion current for each bar separately. A zero current indicates that corrosion is inhibited at that depth. A non-zero current indicates that the conditions for rebar corrosion are present at that depth, while the corrosion current gives an indication of the corrosion rate.

Intertek-CAPCIS probes

Intertek-CAPCIS (Oliver et al., 2009) equipment provides sensors for measuring and monitoring the corrosion condition of steel reinforcement in concrete structures. For this purpose, three sensors were developed by the Intertek-CAPCIS: M3 probe, C4 probe, and M9 probe.

M3 probe – this sensor is a multi-element sensor developed for monitoring the corrosion rate of the reinforcing bars in concrete structures. This probe is designed to be installed during concrete construction. A standard M3 probe consists of a carbon steel working electrode, a silver/silver chloride/potassium chloride (Ag/AgCl/KCl) reference electrode, an AISI316 stainless steel auxiliary electrode, and a thermistor to measure the temperature.

C4 probe – this probe is also a multi-element sensor for monitoring the corrosion rate and condition of reinforced concrete. It is mainly designed for use with tunnel elements. However, the C4 probe can also be installed into any reinforced concrete structure. The following information can be obtained, using C4 probe: corrosion potential, corrosion rate (using LPR), concrete resistivity, concrete temperature, and concrete humidity.

M9 probe – this is a multi-layered variation of the standard M3 probe. The design provides long-term service life, which makes M9 probe ideal for use in aggressive environments or where the ability to make measurements at various cover depths yielding early warning of deterioration is crucial, such as in nuclear waste storage plants.

13.4.2 Other sensors

The onset of corrosion in steel-reinforced concrete can also be detected by different methods. The inductively coupled magnetic fields, with the core typically made of a simple LC resonator, have been used by researchers to sense the initiation of corrosion (Carkhuff and Cain, 2003; Andringa et al., 2005; Bhadra et al., 2010). Change of the inductance or the capacitance of the LC circuit will change the resonant frequency that is the basis of such sensors. The resonant frequency of the whole LC circuit, schematically shown in Fig. 13.11, is given by the Thompson formula:

f=12πLC [13.24]

image [13.24]
image
13.11 Circuit diagram for a corrosion sensor, based on inductance.

where L is the total inductance of the circuit and C is the total capacitance of the circuit.

According to the Equation [13.24], the change of the inductance or the capacitance of the circuit will change the resonant frequency. As can be seen in Fig. 13.11, the resonant frequency of the circuit depends on the state of the steel wire. When the steel wire is not broken, the total capacitance of the circuit is equal to the summation of the two capacitances (C = C1 + C2). However, when the steel wire is broken due to corrosion, the total capacitance of the circuit becomes C1. The diameter of the steel wire is usually much smaller than that for the reinforcing steel, and the steel wire will corrode before considerable corrosion damage has occurred in the reinforcement.

Eddy currents have also been used to detect the corrosion on the embedded steel bars in concrete (Gaydecki and Burdekin, 1994; Miller et al., 2003; Kumar et al., 2006). Eddy currents are generated in a conductive material by a changing magnetic field. In the eddy current technique, the magnetic field in a coil induces eddy currents in the rebar. This eddy current generates a magnetic field of its own that interferes with the main magnetic field. The change in inductance of the coil is then measured using the meter.

Sensors based on the application of fiber optics to detect corrosion have also been developed. An optical fiber consists of a glass core surrounded by a glass cladding that differs in index of refraction. The glass fiber is then coated during the manufacturing process with a protective polymer layer (Merzbachery et al., 1996). The direct spectroscopy of corroded versus un-corroded materials is the basis of this method. Spectrally broadband light is coupled into an optical fiber and then illuminates the region under measurement. The presence of corrosion is determined using color modulation of the broadband input signal (Fuhr and Huston, 1998).

The principle of light reflection has also been used to develop fiber optic sensors to detect corrosion of steel in concrete. The sensor consists of an optical fiber reflection sensor, a sacrificial metallic film, joined to a steel tube. One side of the film is finely polished and is isolated from the environment, while the other side is exposed to the corrosive environment. The corrosion pits initiated at the exposed film surface slowly penetrate the sacrificial film as the exposure time increases. The corrosion pits that reach the polished surface reduce the surface reflectivity of the polished surface. This decrease in reflectivity can be detected by the optical fiber reflectivity sensor (Wang and Huang, 2011).

The fiber Bragg grating (FBG) is used for the corrosion detection purpose as well (Lo and Xiao, 1998; Yang et al., 2006; Gaon et al., 2011). Normal optical fibers are uniform along their lengths. In a simple FBG, the refractive index of the fiber core varies periodically along the length of the fiber. The FBG reflects particular wavelengths of light and transmits all others, and therefore can be used as an inline optical filter to block certain wavelengths, or as a wavelength-specific reflector. The difference between the reflection of corrosion products and that of the steel rebar can be used to detect the corrosion activity on the surface of the rebar.

Acoustic emission (AE) monitoring of concrete is the other technique that has been used to detect rebar corrosion (Zdunek et al., 1995; Yoon et al., 2000; Golaski et al., 2002; Idrissia and Limam, 2003; Reis et al., 2003; Assouli et al., 2005; Fariduddin et al., 2007). AE energy is released when a crack propagates. Since corrosion products are expansive (Lide, 1999), they cause multiple micro-fractures, which ultimately lead to macro-cracks. This activity can be detected by AE sensors. As a result, the initiation of corrosion, location of the zone containing intense corrosion products, crack formation mechanism, and the loss of bond strength due to corrosion between reinforcing steel and the surrounding concrete, can be investigated using this method. However, at this time, expensive equipment and the need for highly trained personnel limit the application of AE technique mainly to laboratory investigations.

13.5 Conclusion and future trends

Different corrosion devices and sensors have been developed to measure the corrosion activity of reinforcing steel bars in concrete. The calculations for the measurements make certain assumptions during the test. As a result, different devices may give different corrosion rates, even when the tests were performed at the same point. However, because the components of sensors are close together, and usually different sensors such as temperature are combined with corrosion sensors, such sensors are generally able to differentiate between regions of high and low corrosion rates more efficiently and provide more accurate results. It is important to recognize that a corrosion rate measurement represents the conditions at the time of the test. Changes in the factors that may affect corrosion rate, such as temperature, concrete resistivity, and oxygen availability, will change the corrosion rate. Therefore, it is difficult to extrapolate service life based on one measurement. Measurements need to be continuous and repeated under different seasons and conditions to have a clear understanding of the corrosion activity of the steel rebars. This task is only possible by using sensors. Prediction of the deterioration of steel-reinforced concrete structures due to corrosion and adapting the accurate maintenance protocol is a difficult task, which will be enhanced by using corrosion sensors.

A typical concrete structure is not equipped with sensors, especially corrosion sensors. However, it is obvious that in the future this will change, and the management of concrete structures will use sensors to monitor such structures. Application of corrosion sensors as a tool for health monitoring of civil engineering structures is a relatively new development with enormous potential. Rising concerns for safety, convenience, efficiency factors, and sustainability tied with government mandates, will increase sensor usage to an unprecedented level, including corrosion sensors, in infrastructures and research on these concepts will continue in the near future.

13.6 References

1. ACI Committee 222. 222R-96: Corrosion of Metals in Concrete. 1996.

2. Ahmed S. Reinforcement corrosion in concrete structure, its monitoring and service life prediction-a review. Cement and Concrete Composite. 2003;30(4–5):459–471.

3. Allen RTL, Forrester JA. The investigation and repair of damaged reinforced concrete structures. In: Crane AP, ed. London: Society of Chemical Industry; 1983:193–222.

4. Alonso C, Andrade C, Castellote M, Castro P. Chloride threshold values to depassivated reinforcing bars embedded in a standardised OPC mortar. Cement and Concrete Research. 2000;30:1047–1055.

5. Anderade C, Martinez I. Embedded sensors for the monitoring of corrosion parameters in concrete structures. NDTCE’09, Non-Destructive Testing in Civil Engineering 2009; Nantes, France.

6. Andrade C, González JA. Quantitative measurements of corrosion rate of reinforcing steels embedded in concrete using polarization resistance measurements. Werkstoffe und Korrosion. 1978;29:515–519.

7. Andrade C, Marcias A, Feliu S, Escudero ML, Gonzalez JA. Quantitative measurement of the corrosion rate using a small counter electrode in the boundary of passive and corroded zones of a long concrete beam. In: Philadelphia, PA: ASTM; 1990;Berke NS, Chaker V, Whiting D, eds. Corrosion Rates of Steel in Concrete, ASTM STP 1065.

8. Andringa MM, Neikirk DP, Dickerson NP, Wood SL. Unpowered wireless corrosion sensor for steel reinforced concrete. In: 4th IEEE Conferemce on Sensors, Irvine. 2005; California, USA.

9. Assouli B, Simescu F, Debicki G, Idrissi H. Detection and identification of concrete cracking during corrosion of reinforced concrete by acoustic emission coupled to the electrochemical techniques. NDT & E International. 2005;38:682–689.

10. ASTM. C876–09: Standard Test Method for Half-cell Potentials of Uncoated Reinforcing Steel in Concrete. In: 2009:446–451. 03.02.

11. Bhadra S, Bridges GE, Thomson DJ. Coupled Coil Sensor for Detecting Surface Corrosion on Steel Reinforcement. In: 14th International Symposium on Antenna Technology and Applied Electromagnetics [ANTEM] and the American Electromagnetics Conference [AMEREM]. 2010; Fairmont Chateau Laurier, Ontario, Canada.

12. Broomfield JP. Field measurement of the corrosion rate of steel in concrete using a microprocessor controlled unit with a monitored guard ring for signal confinement. In: Berke NS, Escalante E, Nmai CK, Whiting D, eds. ASTM International 1996.

13. Carkhuff B, Cain R. Corrosion sensors for concrete bridges. IEEE Instrumentation & Measurement Magazine. 2003;6(2):19–24.

14. Carnot A, Frateur I, Marcus P, Tribollet B. Corrosion mechanisms of steel concrete moulds in the presence of a demoulding agent. Journal of Applied Electrochemistry. 2002;32:865–869.

15. Corrosion Doctors. 2005; In: http://www.corrosion-doctors.org; 2005.

16. Dong S, Lin C, Hu R, Li L, Du R. Effective monitoring of corrosion in reinforcing steel in concrete constructions by a multifunctional sensor. Electrochimica Acta. 2011;56:1881–1888.

17. Duffó GS, Farina SB. Development of an embeddable sensor to monitor the corrosion process of new and existing reinforced concrete structures. Construction and Building Materials. 2009;23:2746–2751.

18. Elsener B, Gulikers J, Polder R, Raupach M. Half-cell potential measurements – Potential mapping on reinforced concrete structures. Materials and Structures. 2003;36:461–471.

19. Elsener B, Klinghoffer O, Frolund T, Rislund E, Schiegg Y, Böhni H. Assessment of reinforcement corrosion by means of galvanostatic pulse technique. In: Blankvoll A, ed. Repair of Concrete Structures. 1997:391–400. Svolvœr, Norway.

20. Fariduddin AKM, Ohtsu M, Hossain KMA, Lachemi M. Simulation of reinforcement-corrosion-induced crack propagation in concrete by acoustic emission technique and boundry element method analysis. Canadian Journal of Civil Engineering. 2007;34:1197–1207.

21. Feliu S, Gonzalez JA, Andrade C. Electrochemical methods for onsite determination of corrosion rates of rebars. In: Berke NS, Escalante E, Nmai C, Whiting D, eds. Techniques to assess corrosion activity of steel reinforced concrete structures, ASTM STP 1276. ASTM 1996:107–118.

22. Fontana MG. Corrosion Engineering. Singapore: McGraw-Hill; 1987.

23. FORCE Technology. Evolution of corrosion in reinforced concrete structures. 2004; In: www.force.dk/NR/rdonlyres/7A1681DB-71A2-4424-B948-D44EEDC-38C35/204/17861en3.pdf; 2004.

24. Force Technology. CorroWatch Multisensor. 2012; In: http://www.forcetechnology.com/en/Menu/Products/Concrete-monitoring/Concrete-monitoring-probes/corrowatchmultisensor.htm; 2012.

25. Fuhr PL, Huston DR. Corrosion detection in reinforced concrete roadways and bridges via embedded fiber optic sensors. Smart Materials and Structures. 1998;7:217–228.

26. Gamry Instruments. DC corrosion techniques. 2005; Application Notes http://www.gamry.com/App_Notes/DC_Corrosion/GettingStartedWithEchem-CorrMeasurements.htm; 2005.

27. Gamry Instruments. Electrochemical Impedance Spectroscopy Primer. 2005; Application Notes http://www.gamry.com/App_Notes/EIS_Primer/EIS_Primer.htm; 2005.

28. Instruments Gamry. Basics of Electrochemical Impedance Spectroscopy. 2006; Gamry Instruments http://www.gamry.com/App_Notes/EIS_Primer/EIS%20Primer%202006.pdf; 2006.

29. Gaon J, Wu J, Li J, Zhao X. Monitoring of corrosion in reinforced concrete structure using Bragg grating sensing. NDT & E International. 2011;44:202–205.

30. Gaydecki PA, Burdekin FM. An inductive scanning system for two-dimensional imaging of reinforcing components in concrete structures. Measurement Science and Technology. 1994;5:1272–1280.

31. Gepraegs OK. Comparison and Evaluation of Electrochemical Techniques and Monitoring Instruments to determine the Corrosion Rate of Steel in Concrete. University of Waterloo. MA.Sc 2002.

32. Golaski L, Gebski P, Ono K. Diagnostic of reinforced concrete bridges by acoustic emission. Journal of Acoustic Emission. 2002;20:83–98.

33. Gonzalez JA, Andrade C, Alonso C, Feliu S. Comparison of rates of general corrosion and maximum pitting penetration on concrete embedded steel reinforcement. Cement and Concrete Research. 1995;25:257–264.

34. Hansson CM. Comments on electrochemical measurements of the rate of corrosion of steel in concrete. Cement and Concrete Research. 1984;14:574–584.

35. Idrissia H, Limam A. Study and characterization by acoustic emission and electrochemical measurements of concrete deterioration caused by reinforcement steel corrosion. NDT & E International. 2003;36:563–569.

36. Jones DA. Principles and Prevention of Corrosion. Prentice Hall, Upper Saddle River, NJ: Macmillan Publishing Company; 1992.

37. Jones DA, Greene ND. Electrochemical measurement of low corrosion rates. Corrosion. 1966;22:198–204.

38. Klinghoffer O. In: In Situ Monitoring of the Reinforcement Corrosion by Means of Electrochemical Methods. Published in Nordic Concrete Research 1995:16.

39. Koch GH, Brongers PH, Thompson NG, Virmani YP, Payer JH. Corrosion Costs and Preventive Strategies in the United States. Washington DC: Federal Highway Administration (FHWA), Office of Infrastructure Research and Development; 2001.

40. Kumar K, Muralidharan S, Manjula T, Karthikeyan MS, Palaniswamy N. Sensor systems for corrosion monitoring in concrete structures. Sensors & Transducers Magazine. 2006;67(5):553–560.

41. Lasia A. Electrochemical impedance spectroscopy and its applications. In: Conway BE, Bockris J, White RE, eds. New York: Kluwer Academic/Plenum Publishers; 1999:143–248. Modern Aspects of Electrochemistry. 32.

42. Lide DR. CRC Handbook of Chemistry and Physics. New York, NY: CRC Press; 1999.

43. Lo Y, Xiao F. Measurement of corrosion and temperature using a single-pitch Bragg grating fiber sensor. Journal of Intelligent Material Systems and Structures. 1998;9(10):800–807.

44. Martin-Perez B, Zibara H, Hooton RD, Thomas MDA. A study of the effect of chloride binding on service life predictions. Cement and Concrete Research. 2000;30(8):1215–1223.

45. Merzbachery CI, Kersey AD, Friebele EJ. Fiber optic sensors in concrete structures: a review. Smart Materials and Structures. 1996;5:196–208.

46. Miller G, Gaydecki P, Quek S, Fernandes BT, Zaid M. Detection and imaging of surface corrosion on steel reinforcing bars using a phase-sensitive inductive sensor intended for use with concrete. NDT&E International. 2003;36(1):19–26.

47. Montemor MF, Simoes AMP, Ferreira MGS. Analytical characterization of the passive film formed on steel in solutions simulating the concrete interstitial electrolyte. Corrosion. 1998;54(5):347–353.

48. Muralidharan S, Ha TH, Bae JH, Ha YC, Lee HG, Kim DK. A promising potential embeddable sensor for corrosion monitoring application in concrete structures. Measurement. 2007;40(6):600–606.

49. Newhouse CD, Weyers RE. Modeling the measured time to corrosion cracking. In: Berke NS, Escalante E, Nmai CK, Whiting D, eds. STP 1276 – Techniques to assess the corrosion activity of steel rein-forced concrete structures. West Conshohocken, PA: ASTM International; 1996.

50. Newton CJ, Sykes JM. A galvanostatic pulse technique for investigation of steel corrosion in concrete. Corrosion Science. 1988;28(11):1051–1074.

51. Oliver KE, Gooderham T, Newton CJ. Corrosion monitoring systems and sensors to track material durability in concrete structures and down-hole applications. Intertek-CAPCIS 2009.

52. Polder RB, Peelen WHA. Characterisation of chloride transport and reinforcement corrosion in concrete under cyclic wetting and drying electrical resistivity. Cement & Concrete Composites. 2002;24:427–435.

53. Pourbaix M. Atlas of Electrochemical Equilibria in Aqueous Solutions. Houston, Texas: National Association of Corrosion Engineers; 1974.

54. Poursaee A. Automatic system for monitoring corrosion of steel in concrete. Advances in Engineering Software. 2009;40(11):1179–1182.

55. Poursaee A, Hansson CM. Potential pitfalls in assessing chloride-induced corrosion of steel in concrete. Cement and Concrete Research. 2009;39(5):391–400.

56. Princeton Applied Research. Basics of Electrochemical Impedance Spectroscopy, Application Note AC-1. 2006; In: http://new.ametek.com/content-manager/files/PAR/078.pdf; 2006.

57. Reis HLMD, Ervin B, Kuchma DA, Bemhard JT. Evaluation of corrosion damage in steel reinforced concrete. In: International Workshop on Structural Health Monitoring, Stanford CA, USA. 2003.

58. Rohrback Cosasco Systems. 900 Concrete Multi-Depth Sensor. 2012; In: http://www.cosasco.com/multi-depth-corrosion.html; 2012.

59. Silverman DC. Tutorial on Cyclic Potentiodynamic Polarization Technique. CORROSION-98, NACE International, San Diego, CA 1998; paper No. 299.

60. Silverman DC. Primer on the AC impedance technique. In: Baboian R, ed. Electrochemical techniques for corrosion engineering. NACE 1986.

61. Silverman DC. Simple Models/Practical Answers Using the Electrochemical Impedance Technique. In: Baboian R, Dean W, eds. Corrosion Testing and Evaluation. Silver Anniversary Volume, ASTM 1990.

62. Stern M, Geary AL. Electrochemical polarisation: I A theoretical analysis of the shape of polarisation curves. Journal of The Electrochemical Society. 1957;104(1):56–63.

63. Stratfull RF. Laboratory Corrosion Test of Steel in Concrete. Sacramento, California: California division of highways, State of California; 1968.

64. Stratfull RF. Half-Cell Potential and the Corrosion of Steel in Concrete. Sacramento, California: California division of highways, State of California; 1972.

65. Swette LL, Manoukian M, Hamdan M, LaConti A, Sohanghpurwala A, Scannell WT. Reference electrode for monitoring steel-in-concrete potential. 1999; U. S. Patent. USA. 5964992.

66. The Roctest Group. SensCore Corrosion Sensor. 2012; In: http://www.roctest-group.com/sites/default/files/datasheets/products/16.1010ENDS_SensCore%20Corrosion%20Current%20Sensor.pdf; 2012.

67. Thuresson TI. Electrochemical monitoring of the influence of concrete quality on the reinforcing steel corrosion in industrial effluent. Queen’s University. MA. Sc 1996.

68. Virginia Technologies Inc. The Embedded Corrosion Instrument – ECI. 2012; In: http://www.vatechnologies.com/eciIndex.htm; 2012.

69. Wang Y, Huang H. Optical fiber corrosion sensor based on laser light reflection. Smart Materials and Structures. 2011;20:1–7.

70. Yang S, Cai HW, Geng JX. Advanced fiber grating corrosion sensors for structural health monitoring. In: Ou JP, Li H, Duan ZD, eds. The Second International Conference on Structural Health Monitoring of Intelligent Infrastructure. 2006:441–443. Shenzhen, China.

71. Yoon D, Weiss WJ, Shah SP. Assessing damage in corroded reinforced concrete using acoustic emission. Journal of Engineering Mechanics. 2000;126(3):1–11.

72. Yunovich M, Thompson NG, Balvanyos T, Lave L. Corrosion Costs and Preventive Strategies in the United States, Appendix D: Highway Bridges. Washington DC: Federal Highway Administration (FHWA), Office of Infrastructure Research and Development; 2001.

73. Zakroczymski T, Fan C-J, Szklarska-Smialowska Z. Kinetics and mechanism of passive film formation on iron in 0.05 M NaOH. Journal of Electrochemical Society. 1985;132(12):2862–2867.

74. Zdunek AD, Prine D, Li Z, Landis E, Shah SP. Early detection of steel rebar corrosion by acoustic emission monitoring. Corrosion 95 Baltimore, Maryland, USA: The NACE International; 1995.


1Mathematicians use i to stand for 1image, but electrochemists use j to avoid confusion with i, the symbol for current.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset