You are on page 1of 12

Construction and Building Materials 221 (2019) 351–362

Contents lists available at ScienceDirect

Construction and Building Materials


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

Natural organic antioxidants from green tea form a protective layer to


inhibit corrosion of steel reinforcing bars embedded in mortar
Ivan Pradipta a, Daniel Kong a,⇑, Joash Ban Lee Tan b
a
School of Engineering, Monash University Malaysia, Jalan Lagoon Selatan, Bandar Sunway, 47500 Subang Jaya, Selangor, Malaysia
b
School of Science, Monash University Malaysia, Jalan Lagoon Selatan, Bandar Sunway, 47500 Subang Jaya, Selangor, Malaysia

h i g h l i g h t s

 Natural antioxidants are potential ‘green’ mixed-type corrosion inhibitors.


 Green tea is rich in natural antioxidants and was chosen as the antioxidant source.
 Green tea extract formed protective layer on rebar (mixed-type corrosion inhibitor).
 Protective layer formation was confirmed visually and with microscopic methods.
 The layer reduced rebar weight loss and was enriched with calcium carbonate.

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

Article history: In our previous work, we have demonstrated that at equal volume, green tea extract (GT) exhibited a
Received 7 December 2018 higher inhibition efficiency (IE) than commercial calcium nitrite corrosion inhibitor, on corrosion of steel
Received in revised form 15 April 2019 reinforcing bar (rebar) embedded in mortar (75–80 vs. 14–24%). GT behaved as a mixed-type corrosion
Accepted 2 June 2019
inhibitor which increased rebar polarization resistance (Rp); indicating that it formed a protective layer
Available online 18 June 2019
on rebar surface. In this paper, the formation of a protective layer was confirmed with visual inspection
and microscopic examinations. In presence of this layer, rebar corrosion was reduced. This is evidenced
Keywords:
by a reduced rebar weight loss and further indicated by a similar chloride permeability between control
Protective layer
Microscopic examinations
and GT concrete, and a similar corrosion rate (CR) between control reinforced mortar and reinforced
EDX mortar incorporating solely the residual solid of GT. The similarities ruled out the plausible higher IE
XRD of GT due to an improved physical barrier of mortar/concrete against corrosion. Analyses with energy-
FTIR dispersive X-ray spectroscopy (EDX), X-ray diffraction spectroscopy (XRD), and Fourier-transform
Weight loss infrared spectroscopy (FTIR) suggested that the layer was enriched with calcium, specifically the calcium
CR carbonate polymorphs. Although the formation of protective layer and increase in Rp were influenced by
Rp the magnitude of antioxidant activity, in this study GT has shown a better IE than CI. This should
Chloride permeability
encourage more studies on the IE of sustainable ‘green’ corrosion inhibitors in concrete.
Antioxidant activity
Ó 2019 Elsevier Ltd. All rights reserved.

1. Introduction inhibitors under alkaline conditions, such as those seen in concrete


applications [1,12–19]. Thus, we suggested an investigation on the
Commercial inorganic corrosion inhibitors such as calcium corrosion inhibition efficiency of natural antioxidants at alkaline
nitrite corrosion inhibitor (CI) are easy-to-use and cost-effective concrete pH.
alternative to stainless steel reinforcing bar (rebar) and cathodic Natural antioxidants are organic compounds which are abun-
protection for preventing rebar corrosion [1–4]. However, the dant in nature [20]. These antioxidants possess multiple polar
potential health and environmental hazards of inorganic corrosion atoms and/or electron-rich bonds which may promote the adsorp-
inhibitors has fostered a reduced use of these inhibitors, and sub- tion of antioxidant molecules on rebar surface, by donating elec-
stitution with less-toxic and more environmentally-friendly trons to vacant d-orbitals of iron atoms and forming coordinate
(‘green’) organic corrosion inhibitors [1,3,5–11]. Nonetheless, there bonds [11,21]. This adsorption confers natural antioxidants a
are only limited studies on the efficiency of green corrosion potential anti-corrosion activity as mixed-type corrosion inhibi-
tors, which form protective layer to inhibit electrochemical reac-
⇑ Corresponding author. tions of rebar corrosion simultaneously: anodic reaction of iron
E-mail address: daniel.kong@monash.edu (D. Kong). oxidation and cathodic reaction of oxygen reduction [22]. Being

https://doi.org/10.1016/j.conbuildmat.2019.06.006
0950-0618/Ó 2019 Elsevier Ltd. All rights reserved.
352 I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362

one of the richest sources of natural antioxidants, green tea has a Table 1
ten-to-thirty-fold higher antioxidant activity than the berries, Volume of calcium nitrite corrosion inhibitor (CI).

which are notably the other rich sources of natural antioxidants Corrosion Chloride-to-nitrite Volume of inhibitor
[23–28]. Thus, we selected green tea as the source of natural inhibitor ratio (L/m3 concrete)
antioxidants, postulating that it should have a better opportunity CI 0.9 40.14
to prevent rebar corrosion given this extremely high antioxidant 1.2 30.11
activity. The green tea was administered as green tea extract. 1.5 24.10
1.8 20.08
Corrosion inhibition efficiency (IE) of green tea extract (GT) was
compared with the IE of CI, which was incorporated as a positive
control. To our knowledge, there has been no study which reported
Table 2
the IE of GT at alkaline pH of concrete, nor compared the IE with
Volume of green tea extract (GT).
the IE of a commercial corrosion inhibitor (e.g. CI). CI and GT were
added during mortar mixing, to inhibit the corrosion of steel rebars Corrosion inhibitor Volume of inhibitor (L/m3 concrete)
embedded in the mortar (reinforced mortar). Dosages of CI were GT 40.14
selected encompassing four chloride-to-nitrite ratios, and dosages 30.11
24.10
of GT were selected similar to the CI dosages. The similar dosages
20.08
allow a justification on selected GT dosages, an IE comparison
between GT and CI, and an indirect adjustment of GT dosages to
chloride concentration [3,9,29–31]. of 1.22 kg/m3, and concrete density of 2,350 kg/m3. The CI concen-
Our previous investigation has shown that rebars embedded in tration and density were obtained from the specification sheet of
mortar admixed with GT displayed a significantly lower corrosion product used.
rate (CR) than both rebars embedded in control mortar, and mortar
admixed with CI [32]. Therefore, GT exhibited a higher IE than CI.
2.1.3. Dosage of resuspended GT solid
The lower CR was due to an increase in polarization resistance
Residual solid of GT lacks in the active compounds (contained in
(Rp) and a reduction in anodic slope (i.e. rate of iron oxidation).
supernatant) which are hypothesized to be responsible for the
The increase in Rp indicated that GT formed a protective layer on
electron donation and anti-corrosion activity of GT as a mixed-
rebar surface [2,6,33]. Therefore, in this paper, the formation of a
type corrosion inhibitor. Therefore, CR of reinforced mortar which
layer is confirmed with visual inspection and microscopic exami-
incorporated the residual solid of GT (without the supernatant)
nations (optical microscope and scanning electron microscopes).
indicates whether GT reduces CR by improving the physical protec-
Contribution of the layer towards reduction in CR was investigated
tion of mortar against corrosion (i.e. filling up mortar pores and
by comparing rebar weight loss in absence and in presence of the
creating denser mortar matrix), in addition to increasing rebar
layer. On the other hand, the plausible reduction in CR due to an
polarization resistance.
improved physical barrier of mortar/concrete against corrosion
GT was initially prepared according to volume shown in Table 2,
was examined by comparing chloride permeability of control and
and centrifuged to separate the supernatant and residual solid. The
GT concrete. Additionally, composition of the layer was elucidated
supernatant was discarded, and the residual solid was resuspended
by analyses with energy-dispersive X-ray spectroscopy, X-ray
in ultra-pure water (UPW) to the original volume shown in Table 2.
diffraction spectroscopy, and Fourier-transform infrared spec-
troscopy. Lastly, effect of GT’s antioxidant activity on CR and Rp
2.1.4. Preparation of mortar mixture
of GT reinforced mortar was also investigated.
Mortar mixture was prepared with Portland Cement CEM II/B-S
42.5N and fine aggregate according to mix design proportion
2. Experimental studies shown in Table 3. Properties of the fine aggregate are shown in
Table 4. The water/cement ratio (w/c) of 0.54 has been widely used
2.1. Materials and sample preparations for accelerated corrosion by other researchers [40,43–50], and was selected to provide
suitable permeability for completing corrosion investigation
2.1.1. Preparation of CI and GT within the available timeframe of this study (under twelve
Dosages of CI were selected encompassing the chloride-to-
nitrite ratios recommended by manufacturer (1.2 and 1.5) and Table 3
wider range of ratios (0.9 and 1.8). On the other hand, GT was pro- Constituents of mortar mixture.
duced by ultrasonic-assisted extraction of green tea leaves with Constituent Proportion (kg/m3)
hot water at 10% (w/w) for 30 min, and removal of large particles
Cement 420.00
from the aqueous extract by filtration. Upon extraction and filtra- Water 226.51
tion, concentration of GT was determined by oven-drying method Fine aggregate 1084.28
at 100 °C [34]. The concentration ranged from 2.0 to 3.0% (w/w), Water/cement = 0.54
and was used to determine the water content of GT.
The CR of GT and CI reinforced mortar were compared at equal
inhibitor volume (Section 2.1.2) [30,35]. Additionally, CR of rein- Table 4
Properties of fine aggregate.
forced mortar which incorporated residual solid of GT was also
determined (Section 2.1.3). The accelerated corrosion employed Properties Remarks
3.5% sodium chloride (NaCl) solution to simulate seawater Effective size 1.5–3.0 mm
[9,31,36–43]. Uniformity coefficient 1.5 ± 8%
Average fineness modulus 2.19 ± 0.08
Main composition Quartz crystalline silica (approximately 94%
2.1.2. Dosage of CI and GT of total composition)
During mortar mixing, GT was added into mortar at equal Appearance Fine to coarse sand grains with brown/grey
to light grey color
volume to CI (Tables 1 and 2) [30,35]. The CI volume was calculated
Specific gravity Approximately 2.6
based on CI concentration (i.e. total solid) of 30% (w/w), CI density
I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362 353

months). Additionally, no chemical admixture (e.g. superplasti- Table 5


cizer) was added into the mortar for this w/c, since the mortar Constituents of concrete mixture.

had exhibited an acceptable workability. Thus, there was no inter- Constituent Proportion (kg/m3)
ference from other chemical admixture during the evaluation on IE Cement 420.00
of corrosion inhibitors. Water 234.59
During casting of reinforced mortar specimens, additional water Coarse aggregate 611.77
contributed to mortar mixture by the corrosion inhibitors was Fine aggregate 1,083.19
Water/cement = 0.56
taken into account, to maintain a consistent w/c in control mortar
and mortar admixed with the corrosion inhibitors. Water content
of the inhibitor was the difference between total inhibitor weight
per liter (i.e. density) and inhibitor total solid per liter.
identical to those shown in Table 4. On the other hand, the coarse
aggregates were 10-mm crushed gravels. The w/c was adopted
2.1.5. Preparation of steel rebars from the w/c used for accelerated corrosion (0.54), with an addi-
Steel rebars of 12 mm diameter and 130 mm length were cut tional adjustment for the moisture content of coarse aggregate.
from commercial carbon steel rebars [11,31,34]. Prior to use, the Similar to the preparation of mortar mixture, additional water con-
rebars were sandblasted, degreased with acetone, cleaned with tributed to concrete mixture by the inhibitors was taken into
UPW, and wiped dry [31,51]. The rebars were painted with high account during concrete mixing.
performance epoxy paint [11], except at surface areas of
1,886 mm2 (in the middle) and 943 cm2 (at the top), which
2.2.2. Preparation of CI and GT
remained unpainted for exposure to NaCl solution and electrical
Volume of CI and GT being admixed into concrete are shown in
connection [52,53]. The painted steel rebars were aligned at the
Table 6. Only CI and GT volume which showed the highest IE (i.e.
center of PVC pipe molds as illustrated in Fig. 1. The distance
lowest CR) during accelerated corrosion were selected for RCPT.
between the lower end of rebar and lower end of the mold was
30 mm, to prevent the exposure of unpainted-electrical connection
section to NaCl, and to ensure a good access to oxygen [54].
Table 6
Volume of calcium nitrite corrosion inhibitor (CI) or green tea extract (GT) being
2.2. Materials and sample preparations for rapid chloride permeability admixed into concrete.
test (RCPT) Corrosion inhibitor Volume of inhibitor (L/m3 concrete)
Control –
2.2.1. Preparation of concrete mixture
CI 30.11
Concrete specimens for RCPT were prepared according to mix GT 30.11
design shown in Table 5. Properties of the fine aggregate are

Fig. 1. Reinforced mortar specimen (Adapted and modified from Alghamdi & Ahmad [55]).
354 I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362

2.3. Methodologies with a Metrohm Autolab M204 potentiostat employing a three-


electrode setup [43]. The working electrode (WE) was the steel
2.3.1. Casting reinforced mortar specimens rebar embedded in mortar, and the reference electrode (RE) and
Procedures for mixing and casting reinforced mortar specimens counter electrode (CE) were saturated calomel electrode and AISI
were adopted and modified from Gencel et al. [56] and Karahan 304 stainless steel plate (120  100  10 mm). Since the corrosion
and Atis [57]. In brief, a laboratory-scale mixer was used for mortar measurement was performed at the end of each drying period and
mixing. Upon mixing, the fresh mortar mixture was filled into PVC prior to beginning of a new wetting period, the specimen was in
pipe molds (Fig. 1) in three-equal layers. Each layer was compacted dry condition. Thus, the electrical connection between RE and
on a vibrating table for 25 s. Three control specimens and three WE was established using a sponge wetted with diluted detergent
specimens for each volume of each corrosion inhibitor were [43,68]. Prior to LPR measurement, open circuit potential (OCP) of
prepared. After 24 h, the hardened mortar specimens were the rebar was measured. When the OCP value had been stable
de-molded and cured for 28 days in saturated limewater following (after 10 min, fluctuation less than 10 mV), OCP was determined
ASTM C 511 [57]. and the rebar was polarized with a scan potential of ±20 mV from
OCP and a scan rate of 0.166 mV/s [11,69]. CR and Rp of the rebars
2.3.2. Casting concrete specimens were obtained by analyzing the plots of log current vs. potential
Procedures for mixing and casting concrete specimens were with Nova 1.11 software.
similar to the previously described procedures for mixing and cast-
ing reinforced mortar specimens, except for the absence of steel 2.3.4.2. Weight loss method. At the end of accelerated corrosion test,
rebars. Moreover, a bigger concrete mixer (20 L capacity) and reinforced mortar specimens were broken and embedded steel
cylindrical steel molds (diameter 100 mm and height 200 mm) rebars were extracted. Rebar weight loss was determined accord-
were used in place of the smaller laboratory-scale mixer (5 L ing to ASTM G1-90.
capacity) and PVC pipe molds. Three control specimens and three
specimens for each of CI and GT concrete were prepared. 2.3.5. Ferric reducing power assay
According to Feng et al. [11], electron donation promotes the
2.3.3. Accelerated corrosion adsorption of a mixed-type corrosion inhibitor on rebar surface.
After 28-day curing, reinforced mortar specimens were sub- In our previous investigation, it was indicated that GT behaved
jected to accelerated corrosion, employing a combination of as a mixed-type corrosion inhibitor [32]. Hence, the electron dona-
impressed current and cyclic wetting-drying exposure. The tion capacity of GT was measured with ferric reducing power
selected current density was 100 mA/cm2, and the wetting-drying antioxidant assay [11,20].
cycle was four days of wetting and three days of drying Procedures for performing the ferric reducing power assay were
[43,46,58–60]. The current density has been used in other studies adapted from the study by Chan et al. [70]. Antioxidant activity of
[45,48,49,61–64]. It was originally proposed by Andrade et al. the four GT volume used for accelerated corrosion were quantified
[65] and further supported by El Maaddawy and Soudki [48]. by accounting for the dilution factor of each volume in mortar. For
Setup for the accelerated corrosion is shown in Fig. 2 and was per- each of the four diluted GT, three different volumes were pipetted
formed at laboratory temperature (26.4 ± 1.7 °C, 70.3 ± 7.5% relative into test tubes, one test tube for each volume, and made up to 1 mL
humidity). The anode was the steel rebar embedded in mortar, the with UPW. Afterwards, 2.5 mL phosphate buffer (0.2 M and pH 6.6)
cathode was AISI 304 stainless steel plate (120  100  10 mm), and 2.5 mL of 1% (w/v) potassium ferricyanide were added into
and the electrolyte was 3.5% NaCl solution [52]. The reinforced mor- each test tube. The aliquot was mixed and incubated at 50 °C for
tar specimen and stainless steel plate were immersed up to 20 mm 20 min. The reaction was stopped by adding 2.5 mL of 10% (w/v)
from the top of the stainless steel plate, to reduce the corrosion of trichloroacetic acid. The 8.5 mL mixture in each test tube was sep-
the crocodile clip used to hold the plate (Fig. 2). The specimens were arated into three test tubes (i.e. three replicates), each containing
connected in series to subject all specimens to the same applied cur- 2.5 mL solution. Each of the 2.5 mL solution was diluted with
rent [48,66]. Nonetheless, mortar pores dry out during the drying 2.5 mL UPW, and subsequently added with 500 mL of 0.1% (w/v)
period, which increases mortar resistivity to current flow. Therefore, ferric chloride. The solution was mixed and incubated in the dark
applying a current density of 100 mA/cm2 during drying period ele- for 30 min prior to absorbance measurement at 700 nm. Ferric
vated the total potential of connected specimens above the maxi- reducing power results were expressed as mg gallic acid equivalent
mum potential of the power supply. This triggered the safety (GAE). The calibration equation for gallic acid standard was
measure of the power supply to automatically shut down the y = 17.085x where y was the absorbance value at 700 nm and x
applied current. Nevertheless, given the significance of cyclic was the gallic acid concentration in mg/mL.
wetting-drying to simulate an exposure to marine environment
and to give a better representation of corrosion products formed 2.3.6. Surface examination of rebar
under natural environment [67], the combination of impressed cur- Rebar surface was initially examined visually and further exam-
rent and cyclic wetting-drying exposure was maintained. However, ined with optical microscope and scanning electron microscope
the impressed current was only applied during the wetting period, (SEM). The surface was also subjected to elemental and mineral
and the specimens were allowed to corrode naturally during the analyses with energy-dispersive X-ray spectroscopy (EDX) and
drying period. During the drying period, the specimens were X-ray diffraction spectroscopy (XRD). Additionally, presence of
removed from NaCl solution, and the corrosion reaction was sus- organic functional groups on the rebar surface was detected with
tained by oxygen ingress. A similar setup was used by Malumbela Fourier-transform infrared spectroscopy (FTIR).
et al. [47]. The accelerated corrosion study was conducted until
the 12th wetting-drying cycle [32]. 2.3.6.1. Optical microscope. Extracted rebars were observed with
ZEISS Stemi 2000 stereo microscope complemented with Zen
2.3.4. Corrosion measurement software. Magnification of the objective and camera adapter were
2.3.4.1. Electrochemical method. Electrochemical corrosion mea- 1.0x and 0.5x.
surement was performed with linear polarization resistance
(LPR) technique at laboratory temperature (26.4 ± 1.7 °C, 2.3.6.2. SEM and EDX. Rebar surface was examined and analyzed
70.3 ± 7.5% relative humidity). The measurement was conducted with Hitachi SU8010 field emission scanning electron microscope
I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362 355

Fig. 2. Accelerated corrosion employing a combination of impressed current and cyclic wetting-drying. The specimens were removed from sodium chloride solution during
drying period (Adapted and modified from Guneyisi & Gesoglu [52] and Fayala et al. [43]).

(FE-SEM) equipped with EDX. Hitachi S3400N-II variable pressure 2.3.6.4. FTIR. Rebar surface was analyzed with Nicolet iS10 FTIR
scanning electron microscope (VP-SEM) was employed for the equipped with ATR Smart iTR Diamond. The resolution was 4 cm 1
examination of the surface at lower magnification. The accelerating and the number of scans were 64 [73,74]. The wavenumber
voltage was 15 kV. ranged from 525 to 4,000 cm 1 [74]. Background spectrum was
collected prior to the collection of every rebar spectrum, and was
2.3.6.3. XRD. Homogenous (i.e. non-localized) layer was scraped subtracted from the rebar spectrum to eliminate the effect of
from rebar surface with a scalpel, and analyzed in powder form accumulated carbon dioxide and water vapor during spectra
with Bruker D8 Discover X-ray diffractometer [71]. The 2h ranged collection [73].
from 5 to 90°. Cu-Ka radiation source and Lynxeye detector were
employed for the analyses (40 kV, 40 mA, 0.15418 nm) [71,72]. 2.3.7. RCPT
The peaks were identified with data files from Joint Committee After day moist curing, concrete specimens were subjected to
on Powder Diffraction Standards- International Centre for Diffrac- RCPT according to ASTM C1202 [75]. Prior to testing, the cylindrical
tion Data (JCPDS-ICDD) and confirmed with published literatures. concrete specimens of 100 mm diameter and 200 mm height were
356 I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362

cut into discs of 100 mm diameter and 50 mm height. The discs CR was more prominent at higher GT volume (40 and 30 L/m3
were fitted into cells of PROOVE’it RCPT instrument, and potential concrete).
difference of 60 V was applied across the discs for 6 hours. Magni-
tude of electrical charge passed through the discs after 6 hours was 3.1.2. Effect of resuspended GT solid
used to determine the chloride permeability of the concrete spec- Fig. 4 shows the CR of control reinforced mortar specimens and
imens according to ASTM C1202. reinforced mortar specimens admixed with residual GT solid. It can
be observed that there was no significant difference in the CR of
2.3.8. Statistical analyses the control specimens and the specimens which incorporated
Results are expressed as mean ± standard deviation of three residual GT solid. As mentioned in Section 2.1.3, the residual solid
samples. Results on CR are presented in figures with error bars, lacks in the active compounds contained in supernatant, which are
to represent the mean and standard deviation of the three samples. hypothesized to be responsible for anti-corrosion activity of GT.
On the other hand, results on antioxidant activity, elemental anal- Therefore, the solid only plausibly improves the physical barrier
yses, and rebar weight loss are presented in tables. Significant dif- of mortar by filling up mortar pores and creating denser mortar
ferences in the results were evaluated with one-way analysis of matrix. Hence, the similar CR observed in Fig. 4 suggests a similar
variance (one-way ANOVA) followed by Tukey HSD post hoc test protective quality of mortar in control and GT mortar. Therefore,
using SPSS 20 software [76]. Within single set of data, the signifi- the active compounds contained in supernatant were responsible
cant differences are represented by superscripted letters after pre- for anti-corrosion of GT by donating electrons as a mixed-type cor-
sented values. The superscripted letters ‘a’ until ‘e’ represent rosion inhibitor.
different statistical groupings, and each group is significantly dif-
ferent based on p-value of 0.05. Therefore, values followed by dif-
3.2. RCPT
ferent superscripted letters are significantly different from each
other.
Magnitude of electrical charge passed through control concrete
and concrete admixed with CI or GT after 6 hours are presented in
3. Results and discussion Table 7. As shown in Table 7, all concrete specimens exhibited a
‘High’ permeability to chloride ion according to ASTM C1202 (elec-
The following sections present the results on CR, chloride per- trical charge >4000 Coulumb) [77]. However, the magnitude of
meability, antioxidant activity, examination of rebar surface, and electrical charge passed through GT concrete was not significantly
rebar weight loss. different from the magnitude of charge passed through control
concrete. This indicates a similar chloride permeability of the con-
3.1. Effect of corrosion inhibitors on CR trol and GT concrete. Thus, the similar chloride permeability fur-
ther supported that the lower CR of GT reinforced mortar
Effect of CI, GT, and resuspended GT solid on CR are described in observed in Fig. 3 was not due to an improved physical barrier of
Sections 3.1.1 and 3.1.2. mortar/concrete against corrosion.

3.1.1. Effect of CI and GT 3.3. Relationships among antioxidant activity, CR, and Rp
Fig. 3 shows the CR of control reinforced mortar specimens and
reinforced mortar specimens admixed with CI or GT. As shown in Effect of GT’s antioxidant activity on CR and Rp of GT reinforced
Fig. 3, GT specimens showed a lower CR than control and CI mortar are presented in Table 8. As shown in Table 8, antioxidant
specimens, thus demonstrating a higher IE of GT than CI. The lower activity of GT affects the CR and Rp of GT reinforced mortar.

500

450 Control
(14.91 ± 1.53 MPa)
400 CI- 40 L/m³ concrete
Corrosion rate (µm/year)

(19.72 ± 0.78 MPa)


350 CI- 30 L/m³ concrete
(20.27 ± 4.12 MPa)
300 CI- 24 L/m³ concrete
(21.46 ± 3.21 MPa)
250 CI- 20 L/m³ concrete
(18.52 ± 2.96 MPa)
200 GT- 40 L/m³ concrete
(23.01 ± 1.11 MPa)
150 GT- 30 L/m³ concrete
(18.05 ± 2.21 MPa)
100 GT- 24 L/m³ concrete
(19.28 ± 1.38 MPa)
50 GT- 20 L/m³ concrete
(18.99 ± 2.84 MPa)
0
1 2 3 4 5 6 7 8 9 10 11 12
Wetting-drying cycle no.

Fig. 3. Mean ± standard deviation in corrosion rate (n = 3) of control reinforced mortar and reinforced mortar admixed with calcium nitrite corrosion inhibitor (CI) or green
tea extract (GT). Values presented in the brackets are the 28-day compressive strength of the mortar.
I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362 357

500

450

400
Corrosion rate (µm/year)
350
Control
300
Resuspended GT solid- 40 L/m³
concrete
250
Resuspended GT solid- 30 L/m³
concrete
200
Resuspended GT solid- 24 L/m³
concrete
150
Resuspended GT solid- 20 L/m³
concrete
100

50

0
1 2 3 4 5 6 7 8 9 10 11 12
Wetting-drying cycle no.

Fig. 4. Mean ± standard deviation in corrosion rate (n = 3) of control reinforced mortar and reinforced mortar admixed with residual solid of green tea extract (GT) being
resuspended in water.

Table 7 Nonetheless, accurate correlation coefficients among those param-


Electrical charge passed through concrete after 6 hours. eters could not be drawn due to the larger variations in CR and Rp
Corrosion inhibitor Volume of inhibitor Electric charge than variation in antioxidant activity. This discrepancy might be
(L/m3 concrete) (Coulumb) the effect of the difference in working pH: accelerated corrosion
Control – 5114 ± 586a at alkaline pH of mortar and antioxidant assay at normal pH.
Calcium nitrite corrosion inhibitor 30 8628 ± 995b As shown in study by Muzolf et al. [78], pH affected electron-
Green tea extract 30 5177 ± 569a donating ability: at pH higher than the pKa of potent antioxidant
Values are expressed as mean ± standard deviation of three samples. Within the compounds in GT (epicatechin, epigallocatechin, epicatechin gal-
same column, values followed by different superscripted letters are significantly late, and epigallocatechin gallate), the compounds are deproto-
different from each other based on Tukey HSD test with p-value of 0.05 nated. The deprotonation increases the electron donation capacity
of the compounds, and thereby enhancing antioxidant activity with
Table 8
Effect of green tea extract (GT) antioxidant activity on polarization resistance and increasing pH. Unfortunately, to our knowledge, there is yet an ana-
corrosion rate of GT reinforced mortar. lytical method to quantify the electron donation capacity of crude
plant extract at alkaline pH of concrete (pH 12–13) [78–81]. Increas-
Ferric reducing power Corrosion rate Polarization resistance
(mg GAE) (mm/year) (X) ing working pH of ferric reducing power assay to pH 12–13 reduced
the absorbance values of gallic acid standard, indicating that the
109.51 ± 4.30a 45.18 ± 6.80a 221.18 ± 29.85a
84.27 ± 2.75b 46.37 ± 3.34a 211.14 ± 7.16a
assay was not suitable to measure antioxidant activity at such a high
75.91 ± 1.87c 54.48 ± 11.78a 214.75 ± 16.11a pH. Therefore, the ferric reducing power values presented in Table 8
66.49 ± 3.91c 60.14 ± 12.07a 190.50 ± 38.26a were measured at original working pH of the assay. Nonetheless,
25.44 ± 0.82d 134.25 ± 41.15b 76.18 ± 11.10b despite the absence of accurate correlation coefficients, it can be
21.47 ± 0.52e 195.88 ± 42.16b 64.49 ± 5.85b
observed that the magnitude of antioxidant activity influenced the
Values are expressed as mean ± standard deviation of three samples. Within the ability of GT to increase Rp and reduce CR: the higher is the electron
same column, values followed by different superscripted letters are significantly donation capacity of GT measured with ferric reducing power assay,
different from each other based on Tukey HSD test with p-value of 0.05. Abbrevi-
the higher is the Rp and the lower is the CR.
ation: GAE = gallic acid equivalent.

Fig. 5. Steel reinforcing bar extracted from control mortar and mortar admixed with 30 L/m3 concrete of calcium nitrite corrosion inhibitor (CI) or green tea extract (GT).
358 I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362

3.4. Surface examination of rebar 3.4.1. Visual inspection


Fig. 5 shows an example of three steel rebars extracted from
Visual inspection, optical microscope examination, SEM comple- each mortar group (control, CI, and GT mortar). In each mortar
mented with EDX, XRD, and FTIR analyses of steel rebars extracted group, a similar finding was observed on the three rebars. There-
after accelerated corrosion are presented in the next sections. fore, only one rebar from each mortar group is presented in

Fig. 7. Scanning electron microscope image of steel reinforcing bar extracted from
Fig. 6. Optical microscope image of steel reinforcing bar extracted from (a) control (a) control mortar, (b) mortar admixed with calcium nitrite corrosion inhibitor at
mortar, (b) mortar admixed with calcium nitrite corrosion inhibitor at 30 L/m3 30 L/m3 concrete, and (c) mortar admixed with green tea extract at 30 L/m3
concrete, and (c) mortar admixed with green tea extract at 30 L/m3 concrete. concrete.
I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362 359

Fig. 5. Moreover, only rebars extracted from mortar added with CI 3.4.3. SEM-EDX
or GT at 30 L/m3 concrete are presented, because this volume The white layer on the surface of rebars extracted from GT and
showed the highest IE during accelerated corrosion. Nonetheless, CI mortar was further examined with VP-SEM, and analyzed with
a similar result was observed on the rebars extracted from mortar FE-SEM equipped with EDX to elucidate the elemental composition
added with the other volume of CI or GT. of the layer. The VP-SEM images are shown in Fig. 7 and the ele-
As shown in Fig. 5, rebar extracted from GT mortar had the least mental analyses are presented in Table 9.
corrosion damage (i.e. largest uncorroded area), in agreement with As shown in Table 9, oxygen constitutes 40–60% of all rebar sur-
the lower CR of GT reinforced mortar presented in Fig. 3. A white face (except for untested rebar surface), suggesting the formation of
layer was observed on the uncorroded area of this rebar, and this iron oxides [72]. Moreover, Table 9 suggests that the white layer on
layer was also observed on rebar extracted from CI mortar. Further rebar extracted from GT mortar was enriched with calcium, as evi-
examination of the layer with microscopic methods and analyses denced by the significantly lower calcium content on rebar extracted
of the layer with EDX, XRD, and FTIR are presented in the next from control mortar, on which the white layer was absent. Further-
sections. more, the calcium was derived from mortar, as calcium was not
detected on the surface of untested rebars. The presence of this
calcium-enriched layer significantly reduced the iron content on
3.4.2. Optical microscope examination rebars extracted from GT mortar than on rebars extracted from con-
Optical microscope images of the white layer observed on the trol mortar and untested rebars. The white layer on rebar extracted
uncorroded area of rebars extracted from GT and CI mortar are pre- from CI mortar was similarly enriched with calcium.
sented in Fig. 6. Fig. 6 also compares the rebars with rebar
extracted from control mortar: rebar from control mortar is mostly 3.4.4. FTIR and XRD
covered with brownish-black corrosion product. More importantly, Calcium enrichment of the white layer on rebar extracted from
the rebar does not exhibit the white layer observed on rebars from GT mortar was further confirmed with XRD and FTIR analyses
GT and CI mortar. (Figs. 8 and 9). XRD spectra shown in Fig. 8 indicate that the white

Table 9
Elemental analyses on surface of steel reinforcing bars extracted from control mortar and mortar admixed with calcium nitrite corrosion inhibitor (CI) or green tea extract (GT).

Corrosion inhibitor Volume of inhibitor (L/m3 concrete) Percentage weight


Oxygen Chloride Calcium Iron
Control – 41.52 ± 5.98a 3.60 ± 2.84a 1.94 ± 1.49a 52.94 ± 4.10a
CI 30 44.60 ± 6.38a 4.49 ± 0.67a 20.38 ± 7.60b 30.54 ± 12.33bc
GT 30 59.11 ± 10.26bc 3.52 ± 0.98a 22.98 ± 5.43b 14.39 ± 11.06bd
Untested/original – 11.16 ± 3.75bd – – 88.83 ± 3.75be

Values are expressed as mean ± standard deviation of three samples. Within the same column, values followed by different superscripted letters are significantly different
from each other based on Tukey HSD test with p-value of 0.05.

Fig. 8. XRD spectra of steel reinforcing bar extracted from control mortar and mortar admixed with calcium nitrite corrosion inhibitor or green tea extract.
360 I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362

Fig. 9. FTIR spectra of steel reinforcing bar extracted from control mortar and mortar admixed with green tea extract.

layer on rebar from GT mortar was enriched with three major cal- Table 10
cium carbonate polymorphs namely calcite, aragonite, and vaterite Weight loss of steel reinforcing bars.
[82–85]. The main peak of aragonite, vaterite, and calcite are indi- Corrosion inhibitor Volume of inhibitor Weight loss
cated by peak 1–3. Although the main peak of aragonite and vater- (L/m3 concrete) (%)
ite (peak 1 and 2) [84–88] overlap with the peaks of corrosion Control – 2.94 ± 0.40a
products [38,72,89–91], main peak of calcite (peak 3) was attribu- Calcium nitrite corrosion 40 2.88 ± 0.32a
ted to calcite alone (JCPDS PDF 00-005-0586 and Ref. inhibitor (CI) 30 2.68 ± 0.38ac
[84,87,88,92,93]) and was not detected on rebar extracted from 24 2.62 ± 0.25ad
20 2.88 ± 0.23a
control mortar. Nevertheless, the presence of the three calcium
Green tea extract (GT) 40 2.09 ± 0.18bcd
carbonate polymorphs on rebar extracted from GT mortar was also 30 1.82 ± 0.16b
suggested by the FTIR spectra shown in Fig. 9. 24 2.24 ± 0.17a
The FTIR absorption bands corresponding to the three poly- 20 2.39 ± 0.20a
Resuspended residual solid 40 2.76 ± 0.23a
morphs are 1,405 and 711 cm 1 for calcite, 961 and 871 cm 1 for
of green tea extract 30 2.55 ± 0.27a
aragonite and vaterite, and 871 cm 1 for calcite and vaterite 24 2.59 ± 0.36a
[73,74,86,88,94,95]. The aforementioned bands represent different 20 2.62 ± 0.32a
vibrations of carbonate ions: asymmetrical stretching vibration
Values are expressed as mean ± standard deviation of three samples. Within the
(m3) at 1,405 cm 1, symmetrical stretching vibration (m1) at same column, values followed by different superscripted letters are significantly
961 cm 1, out-of-plane bending vibration (m2) at 871 cm 1, and different from each other based on Tukey HSD test with p-value of 0.05.
in-plane bending vibration (m4) at 711 cm 1 [73]. On the other
hand, the band at 1,641 cm 1, 2,337–2,361 cm 1, and 3,402 cm 1 with the absence of significant difference in CR as presented in
correspond to the vibration of O–H bending, CO2, and O–H stretch- Fig. 4, Table 10 shows the absence of significant difference in
ing respectively [92,96,97]. weight loss between rebars from control mortar and rebars
Overall, XRD and FTIR analyses illustrated that GT induces the extracted from mortar admixed with resuspended GT solid.
formation of protective layer on rebar, which is enriched with cal- Therefore, in overall, the presence of the calcium carbonate-
cium carbonate polymorphs (calcite, aragonite, and vaterite). The enriched layer reduces rebar corrosion as evidenced by the reduced
corrosion inhibition by this layer is illustrated in next section, by rebar weight loss.
comparing rebar weight loss in absence and in presence of this
layer. 3.6. Overall effects of GT on rebar corrosion

3.5. Rebar weight loss We have observed that rebars embedded in GT mortar had
lower CR than the rebars embedded in control and CI mortar.
Weight loss of rebars extracted from control mortar and mortar The lower CR was not due to an improved physical barrier of mor-
admixed with corrosion inhibitors (CI, GT, and resuspended GT tar/concrete against corrosion, since control and GT concrete had a
solid) are presented in Table 10. As shown in Table 10, in agree- similar chloride permeability, and reinforced mortar incorporating
ment with Fig. 3, rebars extracted from GT mortar had lower solely residual solid of GT had a similar CR to control reinforced
weight loss than the rebars extracted from control and CI mortar. mortar. Instead of improving the physical barrier of mortar/
In particular, the rebars extracted from mortar added with GT at concrete GT reduced CR by forming a layer enriched with calcium
30 L/m3 concrete showed a significantly lower weight loss than carbonate. This layer inhibited rebar corrosion, as suggested by a
the rebars from control and CI mortar. On the other hand, in line decreased rebar weight loss in presence of this layer.
I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362 361

4. Conclusion [14] C. Kamal, M.G. Sethuraman, Caulerpin—a bis-indole alkaloid as a green


inhibitor for the corrosion of mild steel in 1 M HCl solution from the marine
alga Caulerpa racemosa, Ind. Eng. Chem. Res. 51 (31) (2012) 10399–10407.
In this paper, the formation of protective layer on rebar from GT [15] P.B. Raja et al., Evaluation of green corrosion inhibition by alkaloid extracts of
mortar was confirmed with visual inspection and microscopic Ochrosia oppositifolia and isoreserpiline against mild steel in 1 M HCl medium,
Ind. Eng. Chem. Res. 52 (31) (2013) 10582–10593.
examinations (optical microscope and SEM). In presence of this
[16] G. Moretti, F. Guidi, G. Grion, Tryptamine as a green iron corrosion inhibitor in
layer, rebar weight loss was reduced, and analyses with EDX, 0.5 M deaerated sulphuric acid, Corros. Sci. 46 (2) (2004) 387–403.
XRD, and FTIR suggested that this layer was enriched with calcium, [17] F.S. de Souza, A. Spinelli, Caffeic acid as a green corrosion inhibitor for mild
steel, Corros. Sci. 51 (3) (2009) 642–649.
precisely the calcium carbonate polymorphs (calcite, aragonite,
[18] E.S. Ferreira et al., Evaluation of the inhibitor effect of L-ascorbic acid on the
and vaterite). Corrosion inhibition by this layer was further corrosion of mild steel, Mater. Chem. Phys. 83 (1) (2004) 129–134.
demonstrated by a similar chloride permeability between control [19] A.K. Satapathy et al., Corrosion inhibition by Justicia gendarussa plant extract
and GT concrete, and a similar CR between control reinforced mor- in hydrochloric acid solution, Corros. Sci. 51 (12) (2009) 2848–2856.
[20] D.J. Charles, Antioxidant assays, in: Antioxidant Properties of Spices, Herbs
tar and reinforced mortar which was added with residual solid of and Other Sources, Springer New York, New York, NY, 2013, pp. 9–38.
GT. These similarities ruled out the plausible reduction in CR due [21] I. Urquiaga, F. Leighton, Plant polyphenol antioxidants and oxidative stress,
to an improved physical protection of mortar/concrete against cor- Biol. Res. 33 (2000) 55–64.
[22] C.M. Hansson, L. Mammoliti, B.B. Hope, Corrosion inhibitors in concrete—Part
rosion. Although the formation of protective layer and increase in I: The principles, Cem. Concr. Res. 28 (12) (1998) 1775–1781.
Rp were influenced by the magnitude of antioxidant activity, in [23] A.J. Stewart, W. Mullen, A. Crozier, On-line high-performance liquid
this study GT has demonstrated a better IE than CI especially at chromatography analysis of the antioxidant activity of phenolic compounds
in green and black tea, Mol. Nutr. Food Res. 49 (1) (2005) 52–60.
the higher range of volume. This should encourage more studies [24] A.K. Atoui et al., Tea and herbal infusions: Their antioxidant activity and
on the use of sustainable ‘green’ corrosion inhibitors in concrete. phenolic profile, Food Chem. 89 (1) (2005) 27–36.
For example, a future study can investigate the IE of more [25] S.M. Henning et al., Bioavailability and antioxidant activity of tea flavanols
after consumption of green tea, black tea, or a green tea extract supplement,
cost-effective sources of antioxidants. These materials can be
Am. J. Clin. Nutr. 80 (6) (2004) 1558–1564.
derived from various agricultural by-products such as fruit peels, [26] K.W. Lee et al., Cocoa has more phenolic phytochemicals and a higher
fruit seeds, and solid wastes from cereal grain processing (for antioxidant capacity than teas and red wine, J. Agric. Food. Chem. 51 (25)
(2003) 7292–7295.
instance corncob, rice bran, and wheat husk) [98–102].
[27] G. Borges et al., Identification of flavonoid and phenolic antioxidants in black
currants, blueberries, raspberries, red currants, and cranberries, J. Agric. Food.
Chem. 58 (7) (2009) 3901–3909.
Acknowledgement [28] I.F. Benzie, Y. Szeto, Total antioxidant capacity of teas by the ferric reducing/
antioxidant power assay, J. Agric. Food. Chem. 47 (2) (1999) 633–636.
[29] A.M. Abdel-Gaber, E. Khamis, A. Hefnawy, Utilizing Arghel extract as corrosion
The authors would like to thank Advanced Engineering Platform inhibitor for reinforced steel in concrete, Mater. Corros. 62 (12) (2011) 1159–
of Monash University Malaysia for funding this research. 1162.
[30] I. Kondratova, P. Montes, T. Bremner, Natural marine exposure results for
reinforced concrete slabs with corrosion inhibitors, Cem. Concr. Compos. 25
(4) (2003) 483–490.
Declaration of interest [31] F. Bolzoni et al., Experiences on corrosion inhibitors for reinforced concrete,
Int. J. Corros. Scale Inhib. 3 (4) (2014) 254–278.
[32] I. Pradipta, D. Kong, J.B.L. Tan, Corrosion inhibition of green tea extract on
The authors do not have any conflict of interest. steel reinforcing bar embedded in mortar, Materials Science and Engineering
Conference Series, 2018.
[33] M. Ormellese et al., A study of organic substances as inhibitors for chloride-
References induced corrosion in concrete, Corros. Sci. 51 (12) (2009) 2959–2968.
[34] N. Etteyeb, X. Nóvoa, Inhibition effect of some trees cultivated in arid regions
against the corrosion of steel reinforcement in alkaline chloride solution,
[1] H. Gerengi, H.I. Sahin, Schinopsis lorentzii extract as a green corrosion inhibitor
Corros. Sci. 112 (2016) 471–482.
for low carbon steel in 1 M HCl solution, Ind. Eng. Chem. Res. 51 (2) (2011)
[35] N.S. Berke, M.C. Hicks, Predicting long-term durability of steel reinforced
780–787.
concrete with calcium nitrite corrosion inhibitor, Cem. Concr. Compos. 26 (3)
[2] V. Shubina et al., Biomolecules as a sustainable protection against corrosion of
(2004) 191–198.
reinforced carbon steel in concrete, J. Cleaner Prod, (2015).
[36] S. Jiang et al., Effects of deoxyribonucleic acid on cement paste properties and
[3] S.A. Asipita et al., Green Bambusa arundinacea leaves extract as a sustainable
chloride-induced corrosion of reinforcing steel in cement mortars, Cem.
corrosion inhibitor in steel reinforced concrete, J. Cleaner Prod. 67 (2014)
Concr. Compos. 91 (2018) 87–96.
139–146.
[37] J. Okeniyi, C. Loto, A. Popoola, Effects of Phyllanthus muellerianus leaf extract
[4] T.A. Söylev, M.G. Richardson, Corrosion inhibitors for steel in concrete: State-
on steel reinforcement corrosion in 3.5% NaCl-immersed concrete, Metals 6
of-the-art report, Constr. Build. Mater. 22 (4) (2008) 609–622.
(11) (2016) 255.
[5] A.A. Khadom, A.F. Hassan, B.M. Abod, Evaluation of environmentally friendly
[38] R. Vera et al., Corrosion products of reinforcement in concrete in marine and
inhibitor for galvanic corrosion of steel–copper couple in petroleum waste
industrial environments, Mater. Chem. Phys. 114 (1) (2009) 467–474.
water, Process Saf. Environ. Prot. 98 (2015) 93–101.
[39] Q. Jiang et al., Corrosion behavior of arc sprayed Al–Zn–Si–Re coatings on
[6] N.A. Negm et al., Gravimetric and electrochemical evaluation of
mild steel in 3.5 wt% NaCl solution, Electrochim. Acta 115 (2014) 644–656.
environmentally friendly nonionic corrosion inhibitors for carbon steel in 1
[40] S. Altoubat, M. Maalej, F. Shaikh, Laboratory simulation of corrosion damage
M HCl, Corros. Sci. 65 (2012) 94–103.
in reinforced concrete, Int. J. Concr. Struct. Mater. 10 (3) (2016) 383–391.
[7] M.A. Chidiebere et al., Corrosion inhibition and adsorption behavior of Punica
[41] S. Caré et al., Mechanical properties of the rust layer induced by impressed
granatum extract on mild steel in acidic environments: Experimental and
current method in reinforced mortar, Cem. Concr. Res. 38 (8) (2008) 1079–1091.
theoretical studies, Ind. Eng. Chem. Res. 51 (2) (2012) 668–677.
[42] T.H. Ha et al., Accelerated short-term techniques to evaluate the corrosion
[8] P.B. Raja, S. Ghoreishiamiri, M. Ismail, Natural corrosion inhibitiors for steel
performance of steel in fly ash blended concrete, Build. Environ. 42 (1) (2007)
reinforcement in concrete- A review, Surf. Rev. Lett. 22 (3) (2015) 1550040.
78–85.
[9] J.O. Okeniyi, C.A. Loto, A.P.I. Popoola, Electrochemical performance of
[43] I. Fayala et al., Effect of inhibitors on the corrosion of galvanized steel and on
Anthocleista djalonensis on steel reinforcement corrosion in concrete
mortar properties, Cem. Concr. Compos. 35 (1) (2013) 181–189.
immersed in saline/marine simulating-environment, Trans. Indian Inst. Met.
[44] L. Abosrra, A. Ashour, M. Youseffi, Corrosion of steel reinforcement in
67 (6) (2014) 959–969.
concrete of different compressive strengths, Constr. Build. Mater. 25 (10)
[10] H. Verbruggen, H. Terryn, I. De Graeve, Inhibitor evaluation in different
(2011) 3915–3925.
simulated concrete pore solution for the protection of steel rebars, Constr.
[45] S. Caré, A. Raharinaivo, Influence of impressed current on the initiation of
Build. Mater. 124 (2016) 887–896.
damage in reinforced mortar due to corrosion of embedded steel, Cem. Concr.
[11] L. Feng, H. Yang, F. Wang, Experimental and theoretical studies for corrosion
Res. 37 (12) (2007) 1598–1612.
inhibition of carbon steel by imidazoline derivative in 5% NaCl saturated Ca
[46] M. Otieno, H. Beushausen, M. Alexander, Chloride-induced corrosion of steel
(OH)2 solution, Electrochim. Acta 58 (2011) 427–436.
in cracked concrete–Part I: Experimental studies under accelerated and
[12] S. Deng, X. Li, Inhibition by Ginkgo leaves extract of the corrosion of steel in
natural marine environments, Cem. Concr. Res. 79 (2016) 373–385.
HCl and H2SO4 solutions, Corros. Sci. 55 (2012) 407–415.
[47] G. Malumbela, M. Alexander, P. Moyo, Interaction between corrosion crack
[13] P.B. Raja et al., Neolamarckia cadamba alkaloids as eco-friendly corrosion
width and steel loss in RC beams corroded under load, Cem. Concr. Res. 40 (9)
inhibitors for mild steel in 1 M HCl media, Corros. Sci. 69 (2013) 292–
(2010) 1419–1428.
301.
362 I. Pradipta et al. / Construction and Building Materials 221 (2019) 351–362

[48] T.A. El Maaddawy, K.A. Soudki, Effectiveness of impressed current technique [75] V. Achal, A. Mukerjee, M. Sudhakara Reddy, Biogenic treatment improves the
to simulate corrosion of steel reinforcement in concrete, J. Mater. Civ. Eng. 15 durability and remediates the cracks of concrete structures, Constr. Build.
(1) (2003) 41–47. Mater. 48 (2013) 1–5.
[49] C. Alonso et al., Factors controlling cracking of concrete affected by [76] M.J. Norušis, SPSS 14.0 Guide to Data Analysis, Prentice Hall Upper Saddle,
reinforcement corrosion, Mater. Struct. 31 (7) (1998) 435–441. River, NJ, 2006.
[50] D. Koleva et al., Quantitative characterisation of steel/cement paste interface [77] N. Chahal, R. Siddique, A. Rajor, Influence of bacteria on the compressive
microstructure and corrosion phenomena in mortars suffering from chloride strength, water absorption and rapid chloride permeability of fly ash
attack, Corros. Sci. 48 (12) (2006) 4001–4019. concrete, Constr. Build. Mater. 28 (1) (2012) 351–356.
[51] J. Blunt, G. Jen, C.P. Ostertag, Enhancing corrosion resistance of reinforced [78] M. Muzolf et al., pH-dependent radical scavenging capacity of green tea
concrete structures with hybrid fiber reinforced concrete, Corros. Sci. 92 catechins, J. Agric. Food. Chem. 56 (3) (2008) 816–823.
(2015) 182–191. [79] P. Janeiro, A.M.O. Brett, Catechin electrochemical oxidation mechanisms,
[52] E. Güneyisi, M. Gesoğlu, A study on durability properties of high-performance Anal. Chim. Acta 518 (1–2) (2004) 109–115.
concretes incorporating high replacement levels of slag, Mater. Struct. 41 (3) [80] B. Tyrakowska et al., TEAC antioxidant activity of 4-hydroxybenzoates, Free
(2008) 479–493. Radical Biol. Med. 27 (11–12) (1999) 1427–1436.
[53] M. Pech-Canul, P. Castro, Corrosion measurements of steel reinforcement in [81] F. Nanjo et al., Scavenging effects of tea catechins and their derivatives on 1,
concrete exposed to a tropical marine atmosphere, Cem. Concr. Res. 32 (3) 1-diphenyl-2-picrylhydrazyl radical, Free Radical Biol. Med. 21 (6) (1996)
(2002) 491–498. 895–902.
[54] N.S. Berke, D.F. Shen, K.M. Sundberg, Comparison of the polarization [82] Y. Zhang, H.X. Guo, X.H. Cheng, Role of calcium sources in the strength and
resistance technique to the macrocell corrosion technique, Corrosion rates microstructure of microbial mortar, Constr. Build. Mater. 77 (2015) 160–167.
of steel in concrete, ASTM International, 1990. [83] N.K. Dhami, M.S. Reddy, A. Mukherjee, Improvement in strength properties of
[55] S.A. Alghamdi, S. Ahmad, Service life prediction of RC structures based on ash bricks by bacterial calcite, Ecol. Eng. 39 (2012) 31–35.
correlation between electrochemical and gravimetric reinforcement [84] C.G. Kontoyannis, N.V. Vagenas, Calcium carbonate phase analysis using XRD
corrosion rates, Cem. Concr. Compos. 47 (2014) 64–68. and FT-Raman spectroscopy, Analyst 125 (2) (2000) 251–255.
[56] O. Gencel et al., Workability and mechanical performance of steel fiber- [85] E.T. Stepkowska et al., Calcite, vaterite and aragonite forming on cement
reinforced self-compacting concrete with fly ash, Compos. Interfaces 18 (2) hydration from liquid and gaseous phase, J. Therm. Anal. Calorim. 73 (1)
(2011) 169–184. (2003) 247–269.
[57] O. Karahan, C.D. Atisß, The durability properties of polypropylene fiber [86] E. Loste et al., The role of magnesium in stabilising amorphous calcium
reinforced fly ash concrete, Mater. Des. 32 (2) (2011) 1044–1049. carbonate and controlling calcite morphologies, J. Cryst. Growth 254 (1–2)
[58] C.K. Nmai, Multi-functional organic corrosion inhibitor, Cem. Concr. Compos. (2003) 206–218.
26 (3) (2004) 199–207. [87] W. De Muynck et al., Bacterial carbonate precipitation as an alternative
[59] H. Mihashi, S.F.U. Ahmed, A. Kobayakawa, Corrosion of reinforcing steel in surface treatment for concrete, Constr. Build. Mater. 22 (5) (2008) 875–
fiber reinforced cementitious composites, J. Adv. Concr. Technol. 9 (2) (2011) 885.
159–167. [88] C. Wang et al., Synthesis of nanosized calcium carbonate (aragonite) via a
[60] M. Maalej, S.F. Ahmed, P. Paramasivam, Corrosion durability and structural polyacrylamide inducing process, Powder Technol. 163 (3) (2006) 134–
response of functionally-graded concrete beams, J. Adv. Concr. Technol. 1 (3) 138.
(2003) 307–316. [89] A. Liu et al., Formation of lepidocrocite (c-FeOOH) from oxidation of
[61] C.G. Berrocal et al., Corrosion-induced cracking and bond behaviour of nanoscale zero-valent iron (nZVI) in oxygenated water, RSC Adv. 4 (101)
corroded reinforcement bars in SFRC, Composites Part B: Eng. 113 (2014) 57377–57382.
(Supplement C) (2017) 123–137. [90] J.D. Walker, R. Tannenbaum, Characterization of the sol gel formation of
[62] D. Tang et al., Influence of surface crack width on bond strength of reinforced iron(III) oxide/hydroxide nanonetworks from weak base molecules, Chem.
concrete, ACI Mater. J. 108 (1) (2011). Mater. 18 (20) (2006) 4793–4801.
[63] J.A. González et al., Comparison of rates of general corrosion and maximum [91] M. Sheydaei, S. Aber, Preparation of nano-lepidocrocite and an investigation
pitting penetration on concrete embedded steel reinforcement, Cem. Concr. of its ability to remove a metal complex dye, CLEAN–Soil, Air, Water 41 (9)
Res. 25 (2) (1995) 257–264. (2013) 890–898.
[64] K. Zandi Hanjari, D. Coronelli, Anchorage capacity of corroded reinforcement, [92] J.D. Rodriguez-Blanco, S. Shaw, L.G. Benning, The kinetics and mechanisms of
Chalmers University of Technology, 2010. amorphous calcium carbonate (ACC) crystallization to calcite via vaterite,
[65] C. Andrade, M.C. Alonso, J.A. Gonzalez, An initial effort to use the corrosion Nanoscale 3 (1) (2011) 265–271.
rate measurements for estimating rebar durability, Corrosion rates of steel in [93] L. Wang, I. Sondi, E. Matijević, Preparation of uniform needle-like aragonite
concrete, ASTM International, 1990. particles by homogeneous precipitation, J. Colloid Interface Sci. 218 (2)
[66] S. Ahmad, Techniques for inducing accelerated corrosion of steel in concrete, (1999) 545–553.
Arabian J. Sci. Eng. 34 (2) (2009) 95. [94] D. Chakrabarty, S. Mahapatra, Aragonite crystals with unconventional
[67] G. Malumbela, P. Moyo, M. Alexander, A step towards standardising morphologies, J. Mater. Chem. 9 (11) (1999) 2953–2957.
accelerated corrosion tests on laboratory reinforced concrete specimens, J. [95] B. Plav, S. Kobe, B. Orel, Identification of crystallization forms of CaCO3 with
South Afr. Instit. Civ. Eng. 54 (2) (2012) 78–85. FTIR spectroscopy, Kovine Zlitine Tehnol 33 (1999) 6.
[68] C. Andrade, I. Martínez, 14 - Techniques for measuring the corrosion rate [96] N. Nakayama, Inhibitory effects of nitrilotris(methylenephosphonic acid) on
(polarization resistance) and the corrosion potential of reinforced concrete cathodic reactions of steels in saturated Ca(OH)2 solutions, Corros. Sci. 42 (11)
structures, in: Non-Destructive Evaluation of Reinforced Concrete Structures, (2000) 1897–1920.
Woodhead Publishing, 2010, pp. 284–316. [97] S. De Angelis Curtis et al., Crystal structure and thermoanalytical study of a
[69] H. Saricimen et al., Effectiveness of concrete inhibitors in retarding rebar cadmium(II) complex with 1-allylimidazole, J. Anal. Appl. Pyrol. 87 (1) (2010)
corrosion, Cem. Concr. Compos. 24 (1) (2002) 89–100. 175–179.
[70] E. Chan et al., Effects of different drying methods on the antioxidant [98] D.P. Makris, G. Boskou, N.K. Andrikopoulos, Polyphenolic content and in vitro
properties of leaves and tea of ginger species, Food Chem. 113 (1) (2009) antioxidant characteristics of wine industry and other agri-food solid waste
166–172. extracts, J. Food Compos. Anal. 20 (2) (2007) 125–132.
[71] O. Poupard et al., Corrosion damage diagnosis of a reinforced concrete beam [99] B. Sultana et al., Antioxidant potential of extracts from different agro wastes:
after 40 years natural exposure in marine environment, Cem. Concr. Res. 36 Stabilization of corn oil, Grasas Aceites 59 (3) (2008) 205–217.
(3) (2006) 504–520. [100] Z. Khiari, D.P. Makris, P. Kefalas, An investigation on the recovery of
[72] J.K. Singh, D.D.N. Singh, The nature of rusts and corrosion characteristics of antioxidant phenolics from onion solid wastes employing water/ethanol-
low alloy and plain carbon steels in three kinds of concrete pore solution with based solvent systems, Food Bioprocess Technol. 2 (4) (2009) 337.
salinity and different pH, Corros. Sci. 56 (2012) 129–142. [101] C.S. Ku, S.P. Mun, Antioxidant activities of ethanol extracts from seeds in fresh
[73] E. Loftus, K. Rogers, J. Lee-Thorp, A simple method to establish calcite: Bokbunja (Rubus coreanus Miq.) and wine processing waste, Bioresour.
aragonite ratios in archaeological mollusc shells, J. Quat. Sci. 30 (8) (2015) Technol. 99 (10) (2008) 4503–4509.
731–735. [102] S. Okonogi et al., Comparison of antioxidant capacities and cytotoxicities of
[74] Y. Dauphin, J.P. Cuif, C.T. Williams, Soluble organic matrices of aragonitic certain fruit peels, Food Chem. 103 (3) (2007) 839–846.
skeletons of Merulinidae (Cnidaria, Anthozoa), Comp. Biochem. Physiol. B:
Biochem. Mol. Biol. 150 (1) (2008) 10–22.

You might also like