You are on page 1of 10

Journal of Advanced Concrete Technology Vol. 7, No.

1, 41-50, February 2009 / Copyright 2009 Japan Concrete Institute

41

Scientific paper

Temperature Dependency of Chloride Induced Corrosion in Concrete


Nobuaki Otsuki1, Marish S. Madlangbayan1,2, Takahiro Nishida3, Tsuyoshi Saito1 and Melito A.
Baccay4
Received 11 September 2008, accepted 19 January 2009

Abstract
Maintenance of reinforced concrete is a major concern around the world. In particular, special attention is now given to
chloride induced corrosion, which is considered as one of the most serious causes of concrete deterioration. Since corrosion is an electrochemical process, the influence of temperature on the deterioration of reinforced concrete should be
considered. From these backgrounds, the objective of this study is to investigate the influence of temperature on the
deterioration process of chloride induced corrosion in reinforced concrete. The results show that the rates of diffusion of
substances and corrosion of steel bars rise with increases in temperature and these phenomena are explained by the Arrhenius theory using the concept of activation energy.

1. Introduction
The maintenance of reinforced concrete is now becoming
a major concern around the world due to limited construction budgets and diminishing supplies of
high-quality materials for construction. Consequently,
much research related to the deterioration or maintenance
of reinforced concrete structures has already been reported.
Chloride induced corrosion is one of the main causes
of deterioration in reinforced concrete. Normally, the
alkalinity of concrete around a steel bar provides a passive environment (Hausmann 1964; Bentur et al. 1997).
This passive state can be broken down by the presence of
Cl- in concrete, and thus the steel bar can easily start to
corrode with the presence of O2 and H2O. As the corrosion product (rust) occupies a greater volume than the
original steel (Mehta and Monteiro 1993), the expansive
pressure causes concrete to crack or spall. Therefore Clinduced corrosion can be a very important durability
issue especially in coastal and marine areas (Leeming
1983; Srensen and Maahn, 1982; Popovics et al. 1983).
Since the deterioration of concrete is based on diffusion
or chemical reactions, the deterioration progress can also
be considered as having temperature dependency. Generally, the rate of material diffusion and the chemical
reactions rise with increases in temperature.
In developed countries such as the USA, EU and Japan,

Department of International Development Engineering,


Tokyo Institute of Technology, Tokyo, Japan.
2
Civil Engineering Department, University of the
Philippines-Los Banos, Laguna, Philippines.
E-mail:madlangbayan.m.aa@m.titech.ac.jp
3
Center Research Institute for Electric Power Industry,
Abiko-shi, Chiba, Japan.
4
Technological University of the Philippines, Ayala Ave.,
Ermita, Manila, Philippines.

standards for construction materials or construction


methods of reinforced concrete have been established
based on research studies and vast experience under mild
temperature conditions (around 20C). These standards
are now being used and followed by some developing
countries without enough local investigation. When
performing investigations using the standards established
in another country, temperature becomes one of the most
important parameters along with the materials used.
In view of the above, it is necessary to determine the
influence of temperature on the deterioration progress of
reinforced concrete due to chloride attack. The results of
such investigation can in turn be used as grounds for the
discussion of appropriate maintenance methods under
several temperature conditions.

2. Research significance
The rates of diffusion and steel corrosion in concrete are
important factors in the study of the deterioration of
reinforced concrete structures. Temperature is also another important factor that should be taken into consideration due to the possibility of a faster rate of deterioration in reinforced concrete structure with increases in
temperature. However, research studies related to this
matter are inadequate. In addition, there are only a few
research studies that can explain the deterioration of reinforced concrete structures based on the Arrhenius theory.
This study aims to fill in this particular research gap.

3. Materials and experiments


3.1 Materials
Ordinary Portland cement was used as binder. The
physical properties and chemical components of the
cement used are listed in Table 1. River sand and crushed
stones were used as fine and coarse aggregates, respectively. The physical properties of the fine and coarse
aggregates used are listed in Table 2. Tap water was used

42

N. Otsuki, M. S. Madlangbayan, T. Nishida, T. Saito and M. A. Baccay / Journal of Advanced Concrete Technology Vol. 7, No. 1, 41-50, 2009

as mixing water. Prior to mixing the concrete, the tap


water was kept inside the experimental room in order to
keep the temperature of the water the same as the room
temperature (202C). Water-reducing and air-entraining
agents were used as chemical admixtures. The type of
steel embedded in the concrete specimen was deformed
13-mm-diameter steel bar (SD 345) conforming to the
Japanese Industrial Standard (JIS) and having a measured yield strength of 358 MPa. The chemical compositions and physical properties of the steel used are listed in
Table 3.
3.2 Environmental conditions
In this study, the influence of temperature on the deterioration of reinforced concrete was investigated. The
temperature during the experiments was controlled at
20C, 30C and 40C in an environment controlled
chamber. The relative humidity was controlled and
maintained at 60%.
3.3 Items of investigation
(1) Cl diffusivity of minute region in concrete
The test pieces for Cl- diffusivity were taken from cylindrical specimens (100 x 200 mm) prepared in the
laboratory using the mix proportions summarized in
Table 4. The bleeding ratio of concrete was measured in

accordance with JIS A1123. The bleeding ratio was taken


as the percentage of the height of visible water at the top
surface of concrete to the height of the fresh concrete.
The preparation of the test piece for minute diffusion test
is briefly described below. The preparation technique for
the test piece is shown in Fig. 1. From the central part of
the cylindrical specimen (100 x 200 mm), a bar of 15 x
15 x 50 mm was cut using a diamond cutter. The bar was
then cut crosswise into a 15 x 15 x 4 mm piece by means
of a low speed diamond saw. The 15 x 15 x 4 mm piece
was then adhered to a glass plate using electron wax. A 5
x 5 x 4 mm test piece was cut from the 15 x 15 x 4 mm
specimen, carefully selecting the portions containing
only the mortar matrix phase. The test piece from the
glass plate was then detached by applying heat. The 5 x 5
x 4 mm test piece was then placed on the plastic mould
and attached using hard epoxy. Both sides of the test
piece were then polished using a grinder and a polisher
until a 0.5 mm thickness was obtained.
The test piece was then placed inside an acrylic cylinder cell as shown in Fig. 2. A thin rubber sheet was
used as a water proof membrane. One side of the cell was
filled with NaCl solution (3.0 wt%) and the other side
was filled with a saturated Ca(OH)2 solution. The
time-dependent changes of chloride ion concentrations at
the Ca(OH)2 solution side was checked everyday. The

Table 1 Physical properties and chemical components of cement.

Density (g/cm3)

Compressive strength (MPa)


3d
7d
28d
28.4
43.0
60.2
Total alkali (%)
Cl-(%)
0.62
0.006

Specific surface area (cm2/g)

3.16
MgO (%)
1.6

3270
Loss on ignition (%)
1.4

SO3 (%)
1.9

Table 2 Physical properties of fine aggregate and coarse aggregate.

Kind of aggregate
Fine aggregate
Course aggregate
1*
4*

Density (g/cm3)
DRY
SSD1*
2.60
2.54
2.64
2.62

F.M.2*
2.59
7.00

W.A.R.3*
(%)
2.20
0.93

WV4*
(kg/l)
1.74
1.54

SSD- Surface saturated dry condition 2*F.M.- Fine modulus 3*W.A.R- Water adsorption ratio
WV- Weight per volume
Table 3 Chemical compositions and physical properties of steel.

JIS standard
Measured

Fe
(%)

98.78

C
(%)

0.21

Si
(%)

0.18

Mn
(%)

0.78

P
(%)
<0.050
0.023

S
(%)
<0.05
0.029

Yield strength
( N/mm2 )
295<
358

Table 4 Mix proportion of concrete specimens for Cl diffusivity test.

Water
Cement
Sand
Gravel
WRA
AEA
Bleeding
(kg/m3)
(kg/m3)
(kg/m3)
(kg/m3)
(g/m3)
(g/m3)
(%)
140
255
862
1042
561
17.85
0.18
0.55
210
382
733
855
840
26.74
0.51
280
509
596
734
1120
35.64
4.58
WRA and AEA refer to water-reducing admixture and air-entraining admixture, respectively.
W/C

N. Otsuki, M. S. Madlangbayan, T. Nishida, T. Saito and M. A. Baccay / Journal of Advanced Concrete Technology Vol. 7, No. 1, 41-50, 2009

J = Q.

Vcell
A

(1)

where J = flux of ions (mol/cm2/s), Q = penetration


speed (mol/cm3/s), Vcell = solution volume in the saturated Ca(OH)2 solution (cm3), A = cross sectional area of
the test piece (cm2).
The diffusion coefficient was calculated using Ficks
first law of diffusion. When the concentration gradient of
chloride ion was determined, the flux of chloride ion was
estimated. The diffusion coefficient of the chloride ion
was then calculated using Eq. 2.

Concrete Specimen
15mm
15
mm

200 mm

15
mm Cut
15mm

Mortar
Matrix

mm

Cut

4m

15

50mm
50 mm
100mm

44mm
mm

Electron Wax

Adhere

Cut

Glass Plate

mm

55mm
mm

55mm
mm

5
55mm
mm

mm
15mm
m 15

4mm
4 mm

55mm
mm
Detach from
Glass Plate

6
Exposed
Test Piece

Unified with
Epoxy Resin
55mm

35
35mm

55mm Epoxy

Resin
Matrix

Polish

54mm

0.5mm
0.5
mm

Fig. 1 Preparation technique for test piece of minute diffusion test.

Test Piece 5 mm

NaCl Solution
Solution
NaCl
(3.0
wt%)
(3.0wt%)

Ca(OH)
Ca(OH)22 Solution
Solution
(Saturated)
(Saturated)

Epoxy Resin
Matrix
Attached with
Rubber Sheet

Test
Piecesetting
set in epoxy
Test
Piece
in Epoxy

Fig. 2 Diffusion cell.

Dcl =

J cl
C cl
x

(2)

where Dcl
= chloride diffusion coefficient (cm2/s),
J cl = flux of chloride ions (mol/cm2/s), C cl = concenx
tration gradient (cm2/s).
In order to evaluate the effect of temperature on Cldiffusivity on minute regions in concrete, the minute
diffusion test (Otsuki et al. 2004) was conducted under
three levels of fixed exposure temperatures (20C, 30C
and 40C). The mixing, casting and curing for each cylindrical specimen was performed inside a controlled
environmental chamber over a temperature range of 20C,
30C and 40C, respectively, and 90% constant relative
humidity. Subsequent to a curing period of 7 days, the
specimens were transferred to another environment
controlled chamber with 75% relative humidity where
their temperatures were set to maintain the same temperatures as the specimens casting and curing.
(2) Oxygen permeability in concrete
The size of the specimen prepared for the oxygen permeability test was 10 x 10 x 40 cm using the mix proportions summarized in Table 5. A steel bar attached
with lead wire was embedded at the middle part of the
concrete specimen. The measurement set-up of the oxygen permeability test is shown in Fig. 4. As shown in the
figure, the electrochemical cell consists of a reinforcing
steel bar as the working electrode, a stainless steel plate
that acts as a counter electrode, and a Ag/AgCl reference
electrode. The limiting cathodic current density was
measured using a potentiostat and the rate of oxygen
permeability was obtained from the limiting current
(Nagataki et al. 1996) using Eq. 3.

dQ
i
= lim
dt
nF

(3)

where dQ dt = oxygen permeability (mol/cm2/s), i lim =


limiting cathodic current density (A/cm2), F = Faradays
constant (96,500 coulombs/mol), n = number of electron exchanged by oxygen gas (=4).
In order to evaluate the effect of temperature on oxy-

C (mol/l)
Cl
Cl- Concentration

concentration of chloride ions was then measured using


an ion chromatography device. The concentration change
of chloride ions at the saturated Ca(OH)2 solution side is
schematically shown in Fig. 3. In this figure, the inclination where the concentration change became constant
is called the penetration speed. The flux can be calculated
from the penetration speed, because the penetration
speed and the flux are related as shown in Eq. 1.

43

= /

Time

(sec)
T
(s)
-

Fig. 3 Variation in concentration of Cl in Ca(OH)2 cell.

N. Otsuki, M. S. Madlangbayan, T. Nishida, T. Saito and M. A. Baccay / Journal of Advanced Concrete Technology Vol. 7, No. 1, 41-50, 2009

I macro =

I i -1,i - I i ,i +1

Reference Electrode Potential


Potentiostat
Stat
(Ag/AgCl)

Stainless Board
Steel Plate
Stainless

Steel Bar Concrete


Fig. 4 Measurement set-up of oxygen permeability test.

30 cm
30cm

15
cm
15cm

Epoxy
Resin

(4)

Si

where I macro = macrocell corrosion current density


(A/cm2), S i = surface area of steel element No. i (cm2),
I i 1,i = current flowing from steel element i-1 to i
(A), I i ,i +1 = current flowing from steel element i to
i+1 (A).
On the other hand, the method of determining the microcell corrosion rate was as shown in Fig. 8. First, the
polarization resistance on the steel element was determined by AC impedance with a FRA (Frequency Response Analyzer). The amplitude of excitation voltage

Vertical Steel Bar

(3) Corrosion rate of steel bars


In this study, macrocell corrosion and microcell corrosion were separately evaluated using divided steel bars,
the preparation of which will be mentioned later. In
macrocell corrosion, the anode and the cathode exist
non-uniformly, while in the case of microcell corrosion,
the anode and the cathode exist uniformly along the
length of the steel bar.
The details of the reinforced concrete specimen are
shown in Fig. 5. The size of the specimen prepared for
the corrosion test was 150 x 30 x 15 cm. The mix proportions used to prepare these concrete specimens were
the same as those listed in Table 5 except that the W/C
used was only 0.55. The concrete specimens were admixed with 10 kg/m3 of sodium chloride in order to accelerate the corrosion in the experiment. The concrete
specimens described in this section were embedded with
divided steel bars composed of divided steel bar elements,
as shown in Fig. 6, for the purpose of making separate
measurements for the macrocell corrosion and microcell
corrosion rates. The divided steel bars were electrically
connected using lead wires soldered on each steel element. The measurement of macrocell corrosion rate was
conducted as shown in Fig. 7. The macrocell current
flowing between adjacent steel bars was measured using
zero resistance ammeters. The macrocell current density
for a steel element i was calculated using Eq. 4. The
anodic current density was denoted as positive, while the
cathodic current was denoted as negative.

was 50 mV peak to peak with a frequency range of 0.05


Hz to 5000 Hz. The polarization resistance was obtained
as diagramed in a typical Nyquist plot shown in Fig. 9.
The polarization resistance ( Rp i ) for a steel element (i)
was obtained as the difference between the total resistance ( Rt ) and the concrete resistance ( Rs ). In this

Horizontal Steel Bar

gen permeability in concrete, the oxygen permeability


test (Miyagawa et al. 1989) was conducted under three
levels of fixed temperatures (20C, 30C or 40C). The
experiment technique of controlling the temperature
during the mixing, casting and curing of each prismatic
specimen was the same as that used for the cylinder
specimens for the Cl- diffusivity test.

150
cm
150cm

44

Upper Part

Middle Part

Lower Part

Concrete
Concrete Cover: 5cm
5 cm
Steel Bars: Divided (shown in Fig.6)
Fig. 6)
D16, SD345
Upper and Lower Surface was
Coated by Epoxy Resin
Fig. 5 Outline of reinforced concrete specimen.

Table 5 Mix proportions of concrete specimen for oxygen permeability test.

W/C
0.45
0.55

Water
(kg/m3)
153
165

Cement
(kg/m3)
339
300

Sand
(kg/m3)
774
774

Gravel
(kg/m3)
1018
1018

WRA
(g/m3)
746
660

WRA and AEA refer to water-reducing admixture and air-entraining admixture, respectively.

AEA
(g/m3)
24
21

N. Otsuki, M. S. Madlangbayan, T. Nishida, T. Saito and M. A. Baccay / Journal of Advanced Concrete Technology Vol. 7, No. 1, 41-50, 2009

case, the authors calculated the microcell corrosion current density according to the famous equation of Stern
and Geary (1957) shown in Eq. 5.
I micro =

K
R pi S i

(5)

where I micro = microcell corrosion current density


(A/cm2), K = constant (= 0.0209V) (Tsuru et al. 1979),
Rp i = polarization resistance of steel element i(),

Lead Wire
33cm
mm

Steel Element

Epoxy Resin

Fig. 6 Outline of divided steel bar.

Ammeter

Ii-1,i

Ii,i+1

i-1

S i = surface area of steel element i (cm2).


In order to determine the rate of macrocell and microcell corrosion in mm/year, the conversion factor 100
A=1.16 mm/y was used (Miyazato et al. 2001). The
basis of the conversion factor is shown below:
100 x106 ( A / cm 2 ) 55.85( g / mol ) 60 sec

96500(C / mol ) 7.86( g / cm 3 ) x 2 1min


60 min 24 hr 365days 10 mm

= 1.16 mm / y
1hr 1day 1 y
1cm

The macrocell and microcell corrosion rate were


measured at the upper, middle and lower or bottom parts
of the 150 x 30 x 10 concrete specimens. Measurement of
corrosion was done at the upper, middle and lower part
since it is already established that concrete can have
varying concrete properties along its vertical direction. In
a study by Wainwright and Ait-Aider (1995), variations
in the effective water content occurred throughout the
depth of the concrete, thus producing corresponding
changes in concrete properties.
In order to evaluate the effect of temperature on the
corrosion rate of steel bars in concrete, corrosion monitoring tests (Otsuki et al. 2000) were conducted under the
same three levels of fixed temperatures (20C, 30C and
40C). The experiment technique of controlling the
temperature during the mixing, casting and curing of
each prismatic specimen was the same as that used for
the cylinder specimens for the Cl- diffusivity test.
(4) Specific concrete resistance
The specific resistance of concrete was measured using a

|Z|

i+1

Z (Im)

Steel Bar

Concrete

Fig. 7 Measuring method of current density of macrocell


corrosion.

Z (Re)
Rs

Reference Electrode
(Ag/AgCl)

i
Steel Bar

Rpi
Rt

FRA Potentiostat
Potential Stat

Stainless Board

45

where:
|Z|
Z (Im)
Z (Re)

Rs
Rt
Rpi

: Impedance vector
: Imaginary component of impedance
: Real component of the impedance
: Phase angle
: Concrete resistance
: Total resistance
: Polarization resistance (Rt-Rs)

Concrete

Fig. 8 Measurement setup of A-C Impedance with FRA.

Fig. 9 Typical Nyquist plot.

46

N. Otsuki, M. S. Madlangbayan, T. Nishida, T. Saito and M. A. Baccay / Journal of Advanced Concrete Technology Vol. 7, No. 1, 41-50, 2009

corrosion monitor. The connections between the divided


steel bars were removed. In this condition, the electric
current could not flow from one steel element to another,
and the voltage of each steel element was uniformly
distributed. The specific concrete resistance was measured at 10 kHz frequency with 50 mV of voltage. The
resistance measured using this frequency was considered
as the solution resistance in the evaluation of steel corrosion (Moshiduki 1996), and in this study this resistance
was defined as concrete resistance. The specific resistance was calculated using the value of this concrete
resistance based on Eq. 6.

R SC

A
= RC
C

(6)

where R SC = specific concrete resistance ( cm),


RC = concrete resistance (), A = area of concrete
(cm2), C = cover depth (cm).

4. Effect of temperature on material


diffusivity and steel corrosion in concrete
4.1 Theoretical background
Generally, the diffusion of substances and chemical reactions rise with increases in temperature. Since diffusion and chemical reactions are the basis by which most
cases of concrete deterioration occur, the increase in
concrete deterioration with increases in temperature has
to be considered.
In discussions of the influence of temperature on the
diffusion of substance and the chemical reaction, the
activation energy derived from the Arrhenius equation is
the most important factor. Therefore in this section, the
deterioration progress of reinforced concrete was discussed based on the activation energy calculated by the
Arrhenius equation.
In general, the influences of temperature on diffusion
or chemical reaction are theoretically illustrated using the
Arrhenius equation as shown in Eq. 7.

E
k = a exp

RT

(7)

where k = rate constant, a = frequency factor (constant),


E = activation energy (kJ/mol), R = ideal gas constant
(8.3110-3 kJ/K/mol), T = absolute temperature (K)
From Eq. 7, the increase in the diffusion rate and
chemical reactions with increases in temperature can be
theoretically confirmed. During the process of chemical
reactions, the reactants seem to shift to the products side
in an active state with high energy. The difference in
energy between the active state and the reactant is called
the activation energy (Ohki et al. 1994). In this process,
where multiple chemical reactions are sequentially generated, the activation energy follows the chemical reaction that determines the rate of reaction.
On the other hand, by taking the logarithm of Eq. 7, the
equation is transformed into Eq. 8.

E
1
log k =
log10 e + log a
R
T

(8)

From experimental results, the relationship between


the logarithm of the rate constant and reciprocal of absolute temperature is generally called the Arrhenius plot.
Therefore, the following relationship (Eq. 9) can be obtained from the Arrhenius plot.

1
log k = A + B
T

(9)

where k = rate constant based on rate of reaction, T =


absolute temperature (K), A, B = experimental constants.
The activation energy can then be calculated by using
Eqs. 8 and 9 to obtain Eq. 10.
E a =

A
R = 19.19 A
log 10 e

(10)

where A = slope of the Arrhenius plot derived from


experimental results (K)
4.2 Temperature effect on material diffusivity in
concrete
In the work of Alexander et al. (2001), it was reported
that diffusion rates are dependent on the temperature,
moisture content of the concrete, type of diffusant and
the inherent diffusibility of the material. The influence of
temperature on diffusivity of Cl- and O2 in concrete is
shown in Fig. 10. From these graphs, it can be seen that
the rate of substance diffusion such as Cl- or O2 in concrete rises with increases in temperature.
The Cl- diffusivity results in Fig. 10(a) showed that the
concrete specimens with a 4.58% bleeding ratio had a
higher tendency toward increased Cl- diffusivity as the
temperature was increased compared to the specimens
with relatively lower bleeding ratios of 0.51% and 0.18%.
Therefore, it was confirmed through the minute diffusion
test that concrete suffering from higher bleeding will
easily allow higher penetration of chloride ions as the
temperature is increased.
The O2 permeability results in Fig. 10(b) showed that
the concrete specimens with 0.55 W/C ratio had a higher
tendency toward increased O2 permeability as the temperature was increased compared to the specimens with a
0.45 W/C ratio. This observation can be attributed to the
more porous and permeable property of concrete with
higher water content. Therefore it was confirmed through
the oxygen permeability test that concrete with higher
W/C ratio will have a higher tendency toward oxygen
diffusion as the temperature is increased.
The Arrhenius plots of the diffusivity of Cl- and O2 in
concrete are shown in Fig. 11. From these figures, it can
be noted that the relationship between logarithms of the
diffusivity and reciprocal of absolute temperature are
linearly distributed. Therefore, it can be concluded that
the diffusion of Cl- and O2 in concrete follows the Ar-

N. Otsuki, M. S. Madlangbayan, T. Nishida, T. Saito and M. A. Baccay / Journal of Advanced Concrete Technology Vol. 7, No. 1, 41-50, 2009

10

Bleeding Ratio

O2Permeability x 10-12
(mol/cm
(mol/cm22/sec)
/s)

Cl- Diffusivity x 10-7


(cm
(cm22/sec)
/s)

15

0.18%
0.51%
4.58%

10
5
0
10

20

30

40

50

W/C

0.55
0.45

8
6
4
2
0
10

Temperature (C)

47

20

30

40

50

Temperature (C)

(a) Cl diffusivity
(b) O2 permeability
Fig. 10 Influence of temperature on material diffusivity in concrete.
-

Bleeding Ratio

Log
Log(O(O
2Permeability)
2 Permeability)

Log (Cl- diffusivity )

-4

0.18%
0.51%
4.58%

-6
-8
-10
3.1

3.2

3.3

3.4

-10

W/C

0.55
0.45

-11
-12
-13
3.1

3.5

3.2

3.3

3.4

3.5

1000/Temperature(1/K)
1000/Temperature (1/K)

1000/Temperature(1/K)
1000/Temperature
(1/K)

(a) Cl diffusivity
(b) O2 permeability
Fig. 11 Arrhenius plots of material diffusivity in concrete.
-

Table 6 Activation energies of material diffusivity obtained from Arrhenius plots.

Present study (for concrete)


Literature surveys (Page et al. 1981) (for cement paste)

rhenius theory. The activation energies of the material


diffusivity in concrete obtained from the Arrhenius plots
were then calculated and are shown in Table 6. The
authors made an attempt to compare these activation
energy values with those from the works of other researchers. However, an extensive survey by the authors
yielded no activation energy values for concrete in the
literature. Instead, comparison was made with activation
energy values for cement pastes in the work of Page et al.
(Page and Lambert 1987; Page et al. 1981). It was seen
that activation energies of Cl- diffusivity is 2 to 4 times
higher in concrete than in cement pastes. In the case of O2
diffusivity, activation energies are 3.4 to 6 times higher in
concrete than in cement pastes. Furthermore, Table 6
shows that in general, whether in concrete or cement
paste, the activation energy of Cl- diffusivity is higher
than that of O2 permeability, and therefore Cl- diffusivity
has higher temperature dependence than O2 permeability.

Activation energy (kJ/mol)


ClO2
70.4134.9
51.192.2
32.244.6
14.922.0

4.3 Temperature effect on corrosion of steel bar


in concrete
The influence of temperature on macrocell and microcell
corrosion rate in concrete is shown in Fig. 12. From these
graphs, it can be seen that the macrocell and microcell
corrosion rates in concrete rise with increases in temperature, which is consistent with the theoretical expectation that corrosion should increase with temperature.
According to Melchers (2002), generally the corrosion
rate doubles with every 10oC increase in temperature
when the corrosion process is kinetically controlled. Also
in these graphs, the corrosion rates at different temperatures were higher at the upper layer of the concrete
specimen compared to the lower layer. This trend may be
attributed to the quality of the concrete at different positions along the vertical direction. In the upper parts, the
oxygen permeability is higher and the concrete resistance
is lower compared to the lower parts of the concrete

N. Otsuki, M. S. Madlangbayan, T. Nishida, T. Saito and M. A. Baccay / Journal of Advanced Concrete Technology Vol. 7, No. 1, 41-50, 2009

Macrocell Corrosion Rate


(mm/year)

0.03
Position

Upper
Middle
Lower

0.02

0.01
0
0

10

20

30

40

Temperature (C)

50

Microcell Corrosion Rate


(mm/year)

corrosion have higher variation while the activation


energy of microcell corrosion are closer to each other and
are seemingly constant regardless of the specific concrete
resistance.
The difference in behavior of activation energy between macrocell and microcell corrosion induced by Clis explained as follows. In the case of macrocell corrosion, the anodic region and cathodic region exist separately. Therefore, it is considered that the rate of macrocell corrosion is easily influenced by the specific concrete resistance and the rate limiting condition also
changes depending on the specific resistance. On the
other hand, in the case of microcell corrosion, concrete
resistance has only a slight influence since the anodic and
cathodic regions of microcells are close to each other.
The relationship between activation energies of corrosion rate and O2 permeability in concrete is shown in
Fig. 15. It can be seen that the activation energy of corrosion rate is almost the same as that of O2 permeability.
This phenomenon can be explained by Fig. 16, where the
relationship between the measured corrosion rate and
calculated corrosion rate from O2 permeability is shown.
It can be seen that the corrosion rate of a steel bar in
concrete is controlled by O2 permeation because both
phenomena show almost the same values. Therefore, it

specimen. Previous work by Mohammed et al. (2002)


explained the reason behind higher corrosion rates at the
upper portions of concrete members. It was explained
that the upper layer of the concrete member is more
porous, which leads to rapid diffusion of oxygen and
moisture resulting in accelerated corrosion rate of the
steel bar.
The corresponding Arrhenius plots of the corrosion
rate in concrete are shown in Fig. 13. From these figures,
it can be seen that the relationship between the logarithms of corrosion rate and reciprocal of absolute temperature are linearly distributed. Therefore, it can be
concluded that the behavior of steel corrosion in concrete
follows the Arrhenius theory. In addition, the activation
energies of macrocell and microcell corrosion in concrete
obtained from Arrhenius plots are shown in Fig. 14.
From these figures, it can be seen that macrocell and
microcell corrosion in concrete has different temperature
dependencies. The activation energy values of macrocell
corrosion ranged from about 7 kcal/mol to about 20
kcal/mol at different values of specific concrete resistance. On the other hand, the activation energy values of
microcell corrosion ranged from 9 to 15 kcal/mol at
different values of specific concrete resistance. This
shows that the activation energy values of macrocell

0.003
Position

Upper
Middle
Lower

0.002

0.001

0
0

10

20

30

40

Temperature (C)

(a) Macrocell corrosion

(b) Microcell corrosion

Log (Macrocell Corrosion Rate)

Fig. 12 Influence of temperature on corrosion rate of steel bar in concrete.


0

-2

Position

-4

Upper
Middle

-6
3.1

3.2

3.3

3.4

1000/Temperature(1/K)
1000/Temperature (1/K)
(a) Macrocell corrosion

3.5

Log (Microcell Corrosion Rate)

48

0
Position

Upper
Middle
Lower

-2

-4

-6
3.1

3.2

3.3

3.4

1000/Temperature(1/K)
1000/Temperature
(1/K)
(b) Microcell corrosion

Fig. 13 Arrhenius plots of corrosion rate of steel bar in concrete.

3.5

50

N. Otsuki, M. S. Madlangbayan, T. Nishida, T. Saito and M. A. Baccay / Journal of Advanced Concrete Technology Vol. 7, No. 1, 41-50, 2009

can be concluded that the activation energy of corrosion


of steel bars was almost same as that of oxygen permeability. However this conclusion is limited and only valid
for oxygen diffusion controlled corrosion. In other words,
the results of activation energy of corrosion of steel bars
would depend on the experimental conditions that consider corrosion rate vs. oxygen supply.

the logarithms of these phenomena, diffusion of


substance or steel corrosion in concrete, were proportional to the reciprocal of absolute temperature.
This fact indicates that the deterioration of reinforced concrete due to steel corrosion induced by
Cl- agrees well with the Arrhenius equation.
(2) The activation energies of deterioration phenomenon in reinforced concrete were obtained by the
Arrhenius equation using the results of the study as
shown in Table 4. It was confirmed that the temperature dependency of the material diffusivity in
concrete depends on the kind of substance and that
the diffusion of Cl- has higher temperature dependency than that of O2. On the other hand, it was
confirmed that the rates of macrocell corrosion and
the microcell corrosion had different temperature
dependencies. The activation energy of macrocell

5. Conclusions

20

10

00

55
10
15
20
10
15
20
Concrete Resistance (kcm)
Concrete Resistance (k cm)

Activation Energy (kcal/mol)

Activation Energy (kcal/mol)

Based on this study, the following conclusions were


drawn.
(1) It was confirmed that the rate of diffusion of Cl- and
O2 increased as the temperature was increased. The
rate of macrocell and microcell steel corrosion in
concrete induced by Cl- also increased as the temperature was increased. It was also confirmed that

30

49

30

20

10

0
10
15
20
00 Concrete
55 Resistance
10
15
20
(kcm)
Concrete Resistance (k cm)

(a) Macrocell corrosion

(b) Microcell corrosion

Activation Energy of Steel Corrosion


(kcal/mol)

Fig. 14. Activation energies of corrosion of steel bar in concrete.

Corrosion Rate Measured


by Experiment (mm/year)

30

20

10

0
0

10

20

30

Activation Energy of Oxygen Permeation


(kcal/mol)

Fig. 15 Relationship between activation energies of corrosion rate and O2 permeability in concrete.

0.006

0.004

0.002

0
0

0.002

0.004

0.006

Corrosion Rate Calculated


by Oxygen Permeability (mm/year)

Fig. 16 Relationship between measured corrosion rate


and calculated corrosion rate from O2 permeability.

50

N. Otsuki, M. S. Madlangbayan, T. Nishida, T. Saito and M. A. Baccay / Journal of Advanced Concrete Technology Vol. 7, No. 1, 41-50, 2009

corrosion has higher variation at different specific


resistance values, while for microcell corrosion, the
activation energy seemed constant regardless of the
specific concrete resistance. Moreover it was confirmed that the activation energy of corrosion of
steel bar was almost the same with that of oxygen
permeability but only for the limited case where the
corrosion rate of steel bar in concrete was controlled
by the oxygen permeability.
Acknowledgement
The support given by the Japanese Society for the Promotion of Science (JSPS) through the Core University
Program (Environmental Engineering) is gratefully acknowledged.
References
Alexander, M. G., Mackechnie, J. R. and Ballim, Y.
(2001). Use of durability indexes to achieve durable
cover concrete in reinforced concrete structures. In:
S.Mindess and J. Skalny Eds. Materials Science of
Concrete, Vol. VI, The American Ceramic Society,
483-511.
Bentur, A., Diamond, S. and Berke, N. S. (1997). Steel
corrosion in concrete, fundamentals and civil
engineering practice. Modern concrete technology 6,
London and New York: E & FN Spon.
Hausmann, D. A. (1964). Electrochemical behavior of
steel in concrete. American Concrete Institute
Journal, 61(2), 171-188.
Leeming, M. B. (1983). Corrosion of steel
reinforcement in off-shore concrete experience from
the concrete-in-the-oceans programme. In: A.P.
Crane Ed. Corrosion of Reinforcement in Concrete
Construction, Chichester, UK: Ellis Horwood Ltd.,
59-78.
Mehta, P. K. and Monterio, P. J. M. (1993). Concrete
structure, properties and materials. Englewood Cliffs,
N. J.: Prentice Hall.
Melchers, R. E. (2002). Effect of temperature on the
marine immersion corrosion of carbon steels. NACE
International, 58(9), 768-774.
Miyagawa, T., Matsumura, T., Kobayashi, K. and Fujii,
M. (1989). Oxygen transmission through concrete
related to reinforcement corrosion. Journal of
Materials, Concrete Structures and Pavements of
JSCE, 11 (408), 111-120. (in Japanese)
Miyazato, S., Otsuki, N. and Kimura, H. (2001).
Estimation method of macrocell corrosion rate of
rebar in existing concrete structures using
non-destructive test. East Asia-Pacific Conference
(EASEC 8), 2, 531-542.
Mohammed, T. U., Otsuki, N., Hamada, H. and Yamaji, T.

(2002). Chloride-induced corrosion with presence of


gap at steel-concrete interface. ACI Materials
Journal, 99(2), 149-156.
Mochizuki, N. (1999). Corrosion monitoring of
reinforcing steel bars in concrete using A.C.
impedance spectroscopy. ZAIRYO TO KANKYO, 48,
693-696. (in Japanese)
Nagataki, N., Otsuki, N., Moriwaki, A. and Miyazato, S.
(1996). The experimental study on corrosion
mechanism of reinforced concrete at local repair
part. Journal of the Japan Society of Civil Engineers,
544(32), 109-119. (in Japanese)
Ohki, M., Ohsawa, T., Tanaka, M. and Chihara, H.
(1994). Dictionary of chemistry. Tokyo: Tokyo
Kagaku Dojin. (in Japanese)
Otsuki, N., Miyazato, S., Diola, N. B. and Suzuki, H.
(2000). Influences of bending crack and
water-cement ratio on chloride-induced corrosion of
main reinforcing bars and stirrups. ACI Materials
Journal, 97(4), 454-464.
Otsuki, N., Yodsudjai, W., Nishida, T. and Yamane, H.
(2004). New test methods for measuring strength and
chloride ion diffusion coefficient of minute region in
concrete. ACI Materials Journal, 101(2), 146-153.
Page, C. L. and Lambert, P. (1987). Kinetics of oxygen
diffusion in hardened cement pastes. Journal of
Materials Science, 22, 942-946.
Page, C. L., Short, N. R. and El Tarras, A. (1981).
Diffusion of chloride ions in hardened cement
pastes. Cement and Concrete Research, 11, 395-406.
Popovics, S., Simeonov, Y., Boshinov, G. and Barovsky,
N. (1983). Durability of reinforced concrete in sea
water. In: A.P. Crane Ed. Corrosion of Reinforcement
in Concrete Construction, Chichester, UK: Ellis
Horwood Ltd., 19-38.
Smith, W. F. (1996). Principles of material science and
engineering. 3rd Ed., McGraw-Hill.
Srensen, B. and Maahn, E. (1982). Penetration rate of
chloride in marine concrete structures. Nordic
Concrete Research, 1, 24. 1-24. 18.
Stern, M. and Geary, A. L. (1957). Electrochemical
polarization. I. A theoretical analysis of the shape of
the polarization curves. Journal of the
Electrochemical Society, 104, 56-63.
Tsuru, T., Maeda, R. and Haruyama, S. (1979).
Application of A. C. method monitor to local
corrosion. Technology of Corrosion Prevention, 28,
638-644.
Wainwright, P. J. and Ait-Aider, H. (1995). The
influence of cement source and slag additions on the
bleeding of concrete. Cement and Concrete Research,
25(7), 445-456.

You might also like