Behavior of High-Nickel Type Weathering Steel Bars in Simulated Pore Solution and Mortar Under Chloride-Containing Environment

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Advanced Concrete Technology Vol.

19, 370-381, May 2021 / Copyright © 2021 Japan Concrete Institute 370

Scientific paper

Behavior of High-Nickel Type Weathering Steel Bars in Simulated Pore


Solution and Mortar under Chloride-Containing Environment
Emel Ken D. Benito1*, Atsushi Ueno2 and Tomoko Fukuyama3

Received 6 February 2021, accepted 17 April 2021 doi:10.3151/jact.19.370

Abstract
Weathering steel (WS) is known to develop higher corrosion resistance than ordinary steel under atmospheric condition
due to the formation of a protective, dense rust layer. This aspect, however, has not been studied so far in cement-based
materials, which are characterized by high alkalinity and limited oxygen. To address the need for durable RC structure in
extreme environments, it is necessary to study the behavior of WS in concrete. Here, a basic investigation was conducted
to compare the short-term behavior of a newly developed WS with 1% Ni (NT) to conventional WS with 1% Cr (CT) and
carbon steel (PC). One set of steel bars was exposed to solutions with varying air and pH to simulate concrete condition
under chloride-containing environment. Another set was embedded in mortar under wet-dry cycle. Corrosion degree
based on mass loss, coupled with half-cell potential, and corroded area was obtained. Results indicate that alkalinity or
low oxygen appreciably reduces the corrosion rate of steels regardless of composition. These conditions make the cor-
rosion behavior of NT comparable with other conventional steels. It is recommended to explore using longer time and
wider cracks in future studies to achieve clearer difference between the steels.

1. Introduction design needs to be developed for environments with high


airborne salt.
Weathering steel (WS) is a collective term for an array of Extensive characterization studies in recent years
low-alloyed steels that exhibit resistance to corrosion due (Kimura et al. 2004, 2005; Wu et al. 2017) provided
to the development of a dense and chloride-expelling rust evidence on the improved anti-corrosive behavior of WS
layer when the material is exposed to atmosphere (Mor- with small addition of nickel (Ni). Ni was found to fa-
cillo et al. 2014). They have low carbon content of 0.2% cilitate the growth of Fe2NiO4 in the rust layer by sub-
by weight or less, and small addition of one or several stituting the position of iron (Fe). As a result, the inner
metals other than iron (e.g. Cu, Cr, Ni, P, Si, etc.), total- rust layer becomes more negatively charged than the
ing to a maximum added alloys of 3 to 5 wt.% (Cano et al. ferric oxyhydroxide (FeOOH) compounds commonly
2018). Weathering steel is a promising alternative to found in the rust of Cr-type WS, promoting the approach
carbon steel because of small addition of expensive and of Na+ ions near the steel instead of Cl− (Kimura et al.
corrosion-resistant metals, while providing better me- 2005). It is precisely this repelling of chlorides that gives
chanical properties (Cheng et al. 2014). WS was initially the newly developed WS greater resistance compared to
designed to contain chromium (Cr) content up to 1% by conventional WS under high salt concentration.
weight. However, a decade-long laboratory and field While both mentioned WS have been extensively
experiments from 1980’s to 1990’s in Japan revealed that studied under open exposure and alternating wet-dry
this amount is not enough to combat corrosion in conditions, their performance in highly alkaline and
salt-rich, coastal regions (Kihira et al. 1999; Itou et al. low-oxygen environment (i.e., concrete confinement)
2000; Usami et al. 2003). These studies have shown that remains relatively unexplored. To illustrate the purpose
at salt (NaCl) deposition rates above 0.05 mg/dm2/day, of this study further, Fig. 1 depicts an RC storm water
corrosion rate in conventional and Cr-added WS will drainage cover that was broken up in 2016, then sub-
exceed a certain limit, leading to the formation of volu- jected to natural environment of a precast-concrete plant
minous and loosely adhering rust. Hence, a new alloy in Kawasaki City, Kanagawa, Japan for two years.
High-Ni WS was used as reinforcements. Interestingly,
the reinforcements did not exhibit appreciable corrosion
1
after two years, as shown. It is thought that high-Ni WS
Assistant Professor, Department of Civil Engineering, is effective for RC that is partially damaged or cracked.
University of the Philippines-Los Baños, Laguna, However, this observation gives no information on
Philippines. high-Ni WS durability when concrete cover is still intact.
*Corresponding author, E-mail: [email protected] WS resistance is due to the development of a dense
2
Associate Professor, Department of Civil and Environ- and stable rust layer, not on the thin passive film that
mental Engineering, Tokyo Metropolitan University, forms at early stage of concrete life (Shi et al. 2018). To
Minami-Osawa, Hachioji, Tokyo, Japan. this effect, WS is generally allowed to corrode under
3
Associate Professor, Department of Architecture and chloride-rich environment. However, when it does cor-
Urban Design, Ritsumeikan University, Kyoto, Japan.
E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 371

Table 1 Chemical composition of studied steels.


Chemical composition (wt.%)
Steel
C Si Mn P S Cu Ni Cr Total
PC 0.08 0.01 0.33 0.026 0.013 - - - 0.46
CT 0.11 0.65 0.45 0.10 0.003 0.40 0.47 1.06 3.24
NT 0.10 0.21 0.88 0.01 0.010 0.74 1.15 0.02 3.11

rode, the rust that appears will act as a barrier against RC bridge decks and piers).
further chloride intrusions. Furthermore, there is a thin This paper discusses the results of an experimental
region in steel-concrete interface (SCI), existing over a program that provides preliminary data on the behavior
distance of m to cm from steel surface, that is composed of newly developed WS with 1.0% Ni by weight, relative
of air voids, macro-cracks, crevices, slips and separations to conventional WS with 1.0 % Cr by weight, and ordi-
(Angst et al. 2017). Unlike the corrosion products in nary carbon steel. First, a series of steel bars were sub-
carbon steels, which are expansive and loose, rust in WS mitted to liquid test (Benito et al. 2019) with varying
is well-adhering and dense (Kimura et al. 2005). It is amount of oxygen and pH to simulate the conditions
hypothesized that existing spaces in SCI may provide inside a concrete with chlorides. This experiment aims to
room for rust layer to develop as the rust is densified determine which environment in general will permit a
through Ni-addition. The resulting composite would reasonable degree of protection. Then, the study pro-
therefore provide better protection against further dam- ceeded to embedding the same steel types in mortar
age after corrosion initiation, compared to carbon-steel specimens.
RC. In view of the above premises, a basic study on the
newly developed WS when embedded in concrete – 2. Materials and experimental Details
characterized by an alkaline and low-oxygen condition –
is therefore necessary. Since high-Ni WS was developed 2.1 Simulated pore solution
to promote corrosion resistance against elevated NaCl The chemical compositions of steels tested are listed in
deposition rates (≥ 0.05 mg/dm2/day), this study is vital Table 1. PC is the control specimen specified in Japan
to understand its application for coastal structures (e.g. Industrial Standard (JIS G 3505) with carbon content of
0.8% or less. CT contains about 1.0% Cr by weight and
low Ni, while NT has 1.0% Ni by weight with almost no
Cr. Both are low-alloyed steels conforming to JIS G 3114,
the standard that specifies the minimum mechanical
properties for WS in Japan. Plain round bars with 5 mm
diameter were purchased. They were cut to lengths of 50
mm, and their ends polished using SiC paper No. 120 to
remove the sharp edges. One steel bar was assigned to
each exposure period lasting 14, 21, 30, 60, 90, 120, and
150 days. Solutions were prepared according to different
conditions provided in Fig. 2, and is explained more
deeply in proceeding sections.

Fig. 1 Two-year exposed high-nickel weathering steel in


2.2 Mortar test
an RC drainage cover.
Ordinary Portland cement with a density of 3.15 g/cm3

Simulated liquid test Mortar test


Oxygen Content pH Level Steel type Oxygen content Steel type
Description Mark Description Mark Description Mark Description Mark Description Mark
Aerated liquid AE Plain carbon Plain carbon
3.5% NaCl PC PC
3-days wetting, A steel Uncracked N steel
SF (pH = 6.1)
4-days drying
7-days wetting, Cr-type WS CT Cr-type WS CT
LF 3.5% NaCl
7-days drying + NaOH
B Cracked C
Complete (pH = 13.1)
IM Ni-type WS NT Ni-type WS NT
immersion

XX ‐ X ‐ XX X ‐ XX
Series Name Series Name

Fig. 2 Parameters and corresponding symbols used in the study.


E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 372

was used as the binding material. Standard sand that ble by simple electronic aerators that introduce air flow
complies with JIS R 5201 was used as fine aggregates rate of about 1.0 L/s.
having an air-dried density of 2.61 g/cm3, and absorption Acidic and alkaline solutions were also considered to
of less than 0.2%. Steel bars with the same composition compare the rust development of WS in environment
as listed in Table 1 were also acquired with dimensions similar to atmosphere and concrete, respectively. Water
of 5 mm diameter and 200 mm length. containing 3.5% NaCl by weight was prepared to simu-
Based on literature, the formation of protective rust late an environment with low pH (case A). A separate
layer happens only when steel does not retain water on its set of 3.5% NaCl solutions were added with NaOH at
surface, or undergoes periodic drying (Morcillo et al. 0.1 mol/L so that a chloride-contaminated alkaline solu-
2019). To increase the rate of supplying and draining of tion is achieved (case B). The chemical composition of
water, and to permit rapid chloride migration, the ab- the pore water will affect the composition of corrosion
sence (case N) and presence of crack (case C) were products; however, since this study is a basic investiga-
considered. The use of cracked specimens is justified tion, the focus is mainly on pH level, and NaOH was
since RC elements do sustain tensile cracks, which sub- selected because of its high solubility and ease to keep
ject the embedded steel bars to a slow process of wetting the desired high pH value. All solutions were put in 500
and drying. Hereafter, specimens will be designated mL polypropylene bottles and were replaced every 2
according to symbols presented in Fig. 2. For each con- weeks to maintain their alkalinity and salinity as rusting
dition stated, two trial samples were allotted for 60, 90 progressed. Specimens were arranged so as not to create
or 120 days of exposure time, summing to 36 mortar contact during immersion. All NaCl solutions were
specimens casted in total. measured to have an average pH value of 6.05 ± 0.59,
while it was 13.10 ± 0.21 for the remaining NaCl and
3. Experimental procedure NaOH solutions. Acidic pH was observed in NaCl solu-
tions probably as a result of stirring, which can capture
3.1 Simulated pore solution free CO2 from the surroundings. Finally, the entire test
WS is generally used in atmosphere where air is fairly was carried out in closed room of constant conditions at
present. Concrete, on the other hand, has reduced oxygen 20°C and 60% RH. Figure 3 is a photograph of labora-
access, which can change depending on its density and tory setup shown together with schematics of different
water content. Different levels of oxygen were investi- corrosion methods.
gated to compare the behavior of studied steels under After about 14, 21, 30, 60, 90, 120, and 150 days, the
these two opposing conditions. To change the oxygen assigned steel bars were retrieved and cleaned according
content, four levels of exposure were set based on the to the procedure outlined in ASTM G1 (2003). First,
amount of air that will interact with steel bars. These corrosion products were cleaned manually by a combi-
are: continuous liquid immersion (IM), cyclic wetting nation of light scraping and sanding using SiC paper No.
and drying with short (SF) and long frequency (LF), and 400 in water. It was followed by immersion in a solution
continuous aeration (AE). SF case consisted of 3 days of of 16% ferric ammonium citrate, initially boiled at 70 to
complete immersion in solutions followed by 4 days of 90°C for 20 minutes. Finally, samples were further
drying; while LF consisted of 7 days wetting, and 7 scraped to remove the remaining rust, rinsed in reagent
days drying. Continuous aeration (AE) was made possi- acetone, and weighed. The mass loss due to corrosion

7 days 7 days
wetting drying To electric aerator

Long-frequency
wet-dry (LF) Continuous
aeration

3 days 4 days
wetting drying

Short-frequency Continuous
wet-dry (SF) immersion
Fig. 3 Laboratory setup for solution test showing different conditions.
E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 373

Table 2 Mortar mix proportions and physical properties.


W/C Unit weight (kg/m3) Flow Air Density f c
S/C
(%) W C S (mm) (%) (kg/m3) (MPa)
60 3.0 290 484 1452 169 3.3 2,254 48.41

was converted to an average value through Eq. (1), re- mapping (Elsener 2002). Protruding ends of steel bars
garded as an average of mass change over the corroded were completely coated with epoxy resin. For the
area. 120-day specimens only, one end was connected to an
electric wire using copper tape, which will serve as the
mi  m f positive terminal for potential measurement. This part
C (1)
2 R 2  2 RL was similarly covered with epoxy resin.
The proportions used for mixing, together with re-
where C = corrosion degree (g/dm2), mi = mass of the sultant physical properties of mortar in fresh and hard-
rebar before exposure (g), mf = mass of the rebar after ened state, are given in Table 2. A water-cement ratio of
removing the rust products (g), R = radius of steel bars 0.60 was selected based on previous microstructural
equal to 0.025 dm, L = measured length of the steel bar studies on mortars (Stefanoni et al. 2019), showing that
(dm). this ratio produces a porosity just less than what is re-
quired for the microstructure to behave like a bulk solu-
3.2 Mortar test tion. In this way, the pore structure is not so open to
3.2.1 Specimen preparation resemble a liquid solution but not low enough to hinder
The schematic of specimens made for this study is pre- wetting and drying action.
sented in Fig. 4. Mortars were cast in cylindrical plastic Mixing followed a stepwise procedure using an au-
molds made up of rigid polyvinyl chloride pipe with tomatic rotary mixer with 20-liter capacity as outlined in
nominal inner diameter of 150 mm. Molds were cut into Fig. 5(a). Mortars were first poured into the prepared
50 mm heights and drilled with 6 mm diameter hole molds in two layers with each layer rodded and lightly
passing their diameter. The configuration was selected compacted [Fig. 5(b)]. After 24 hours, the specimens
because of two reasons. First, plastic molds could restrict were moist-cured by wrapping them in wet cloth, and
the transport of moisture to vertical direction, and thus sprinkling them with water every week for 28 days un-
fluid transport due to wetting and drying can occur der 20°C and 60% RH [Fig. 5(c)]. After curing, half of
mainly over the top and bottom faces. Secondly, the the specimens were cracked by applying a compressive
installation is simpler because the steel bar is firmly load in a manner similar to Fig. 5(d). The load was ap-
secured without the need for spacers. plied until a crack appears in the direction perpendicular
The form height was selected to permit a cover of 22 to the axis of steel, and over the entire depth of mortars
mm. This distance is considered to be within the range (Fujiwara et al. 2017). Because the cracks induced were
influenced by wetting and drying action (Chrisp et al. practically small relative to the length of exposed area
2002), and over which corrosion activity of embedded of steel, the influence of crack on corrosion degree was
steel is sufficiently detectable by half-cell potential assumed to be the same for all bars.

With or without crack


Rigid PVC pipe
[Unit: mm]
Epoxy resin
Adhesive Cu tape
22.5 steel For 120-days
5.0 Electric wire specimens only
22.5 mortar
150
20 5 5 20
Elevation

Exposure zone

150
20 5 5 20
Plan
Fig. 4 Dimension of mortar specimens.
E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 374

3.2.2 Exposure condition with saturated copper/copper sulfate electrode (CSE) as


Only wetting and drying scenario was employed in this reference half-cell. Although HCP method is not quan-
study to simulate the conditions in the field. The samples titative, various researchers (Pourbaix et al. 1980; Shi et
were submitted to exposure immediately after cracking. al. 2018) used it to compare the possibility of corrosion
Each cycle is composed of 3 days complete immersion in in WS. This method allows us to monitor the corrosion
3.5% NaCl solution, followed by 4 days drying in closed risk of the bars non-destructively. Measurements were
room of constant temperature (20°C) and relative hu- made immediately after the specimens were taken out of
midity (60%), which lasted for 120 days. Within this the solution after wetting stage, and just before they were
period, the specimens were kept slightly elevated so that immersed after drying stage. The electrode was placed
water and air could freely circulate at the bottom of on three adjacent points of uncracked mortars to measure
mortars. Note that pH of 3.5% NaCl was not recorded the potential over the entire length of bar. However, for
since this is not a variable of interest in this part of study, the cracked mortars, the electrode was placed only at two
but similar condition as in solution test can be conven- points, about 1 to 5 mm from the crack segment (see Fig.
iently assumed. Lastly, solution is replaced every month 6). Placing the electrode directly above the crack was
to maintain its salinity. avoided to prevent fluctuations of potential due to water
percolating through the crack (Dasar et al. 2017). Finally,
3.2.3 Half-cell potential the average value measured from two separate samples
Half-cell potential (HCP) measurement was conducted was obtained, and termed as corrosion potential (Ecorr) in
on 120-day specimens following ASTM C876 (2009), the following sections.

Time (sec)
0 30 60 90 120 150 180 210 240 270 300

Rotary mixing of sand & cement


Place standard sand
Rotary mixing

Manual scraping
Rotary
mixing

low high low high

(a) manual Speed of Rotary mixer

(b) (c) (d)


Fig. 5 Flow of experiment: (a) step-wise mixing procedure and the rotary mixer used; (b) casting; (c) curing; and (d)
cracking method.

Uncracked Cracked
Voltmeter Saturated
CuSO4 solution
+
Sponge
(wet at all times)

Half-cell
Alligator clip electrode

1 to 5 mm from crack segment


Fig. 6 Half-cell potential measurement assembly.
E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 375

3.2.4 Rust coverage where C = corrosion degree of steel bars inside the
The specimens were broken up using sledge hammer to mortar (g/dm2), mio = initial mass of original length (g),
recover the bars after the assigned periods. The surface Lc = actual length after cutting (dm), Lo = original length
area of the bars within the mortar was covered with before cutting (dm), mfc = final mass of the cut length (g).
transparent adhesive tape to trace localized rust and
coloration using a fine-tip marker. The tape was detached, 4. Results and discussion
and laid flat in a white paper to be captured and analyzed
in an image processing software. The extent of rust, 4.1 Simulated pore solution
corresponding to the areas marked in the tape, was ex- 4.1.1 Influence of alkalinity and oxygen level
pressed as a fraction of the original exposed area within Figure 7 plots the corrosion degree with time, arranged
the mortar by Eq. (2). according to steel type. Under the same oxygen (air)
content, alkaline pH (case B) suppressed the damage by
Acorr 30 to 90% compared to acidic pH (case A) across all
r (2)
Ao exposure conditions. This result was expected since
dissolution of iron at pH > 11.5 favors passive products
where r = rust coverage (%), Acorr = corroded areas to form, mainly as Fe2O3 (maghemite) or Fe3O4 (mag-
marked in the tape (mm2), Ao = surface area of the bars netite) (Hansson 1984). These products are stable on the
measured within the mortar (mm2). surface of steel, unlike iron ions (Fe2+ or Fe3+) that are
continuously released when pH is acidic. Continuous
3.2.5 Corrosion degree dissolution under low pH may explain why the curves
The steel bars were then cleaned following a similar present linear increase until 90 days of exposure, then
procedure applied in solution test. When the bars were became steady thereafter; whereas those in alkaline water
recovered, significant rusting was found to occur at the remained leveled down throughout the test. Gradual
end portions covered with epoxy. To consider only the stability of corrosion curves in acidic pH can be associ-
area embedded in mortar, both ends were cut to retain ated with the decrease of oxygen penetration as rust
only a length of about 140 mm. Then, the actual length thickness increases with time.
and mass of these newly cut bars were measured, and Eq. Furthermore, corrosion degree appreciably increases if
(3) were applied as the estimate of corrosion degree. The air supply increases. This result is also expected since
said expression similarly assumes that mass loss is uni- oxygen supports the cathodic reaction. It is rather worth
form over the exposed lateral surface area of newly cut mentioning that the effect of oxygen on corrosion degree
bars. is more pronounced in acidic solution. For instance,
when pH is alkaline, an increase of oxygen exposure
L  from completely immersed state (IM) to wet-dry cycle
mio  c   m fc
C  Lo  (3)
(SF or LF) hardly changed the corrosion degree; whereas
 DLc in solution of acidic pH, corrosion increased by 3 to 4
times for the same levels of oxygen. This result shows

50 20
AE - A : AE - B : SF - A : SF - B :
40 -■- PC 16 -■- PC
PC PC
30
-▲- CT CT 12 -▲- CT CT
-●- NT NT -●- NT NT
20 8
Corrosion degree (g/dm2)

10 4

0 0
0 30 60 90 120 150 0 30 60 90 120 150 0 30 60 90 120 150 0 30 60 90 120 150
(a) Aerated liquid (b) 3-days wetting, 4-days drying
20 5
LF - B : IM - A : IM - B :
16
LF - A : 4
- - PC -■- PC - - PC
-■- PC
12 - - CT 3 -▲- CT - - CT
-▲- CT
- - NT -●- NT
8 -●- NT 2 - - NT

4 1

0 0
0 30 60 90 120 150 0 30 60 90 120 150 0 30 60 90 120 150 0 30 60 90 120 150

(c) 7-days wetting, 7-days drying (d) Complete immersion


Exposed time (days)
Fig. 7 Change of corrosion degree with time as a function of alkalinity, corrosion method and steel type. See Fig. 1 for
symbols used (Benito et al. 2019).
E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 376

that appropriate combination of oxygen and pH is nec- NT did not show any difference compared to other
essary for rust development even for WS. types in environment that remains permanently wet (AE
cases), is alkaline (SF-B and LF-B) or has low oxygen
4.1.2 Influence of steel type (IM cases). The first item is explained by the fact that
Considering the result of Fig. 7 again, it is found that corrosion layer grows rapidly at continuous contact with
Cr-added steel (CT) recorded the highest resistance only chloride-containing water by imbibing large amount of
when the solution is continuously aerated (AE). Corro- moisture. This condition forces some portion of rust to
sion degree in CT decreased by an average of 60% rela- separate from the metal, making the layer difficult to
tive to other types in acidic case (AE-A), but the reduc- stabilize (Matsushima et al. 1974).
tion was only 30% in alkaline case (AE-B). Previous Moreover, in the range of pH and oxygen content
studies postulated that Cr can interact with multiple ox- usually found in hardened cement paste, the whole ca-
ygen ions (O2–) to form chromium-oxygen ion com- thodic reaction (reduction of oxygen in water) shifts to
plexes of the form CrOx3-2x. These ions can fill the spaces lower current values (Stefanoni et al. 2019). Thus, cor-
of rust in Cr-added steels that densifies its structure, rosion rate is reduced regardless of steel composition
restricting the growth of rust layer (Yamashita et al. since the process is controlled, in large part, by the ca-
2004). While this is apparent for AE, it is not so clear in thodic reaction. This inhibiting effect impedes iron, and
other cases. It can be said that CT suppresses the corro- important alloys (Ni or Cr) to oxidize and grow into
sion degree more visibly when oxygen is abundant (AE phases known to develop the protective rust, namely
condition). α-FeOOH (Kimura et al. 2003) and Fe2NiO4 (Kimura et
On the other hand, Ni-added steel (NT) slightly lev- al. 2005) for Cr- and Ni-bearing steel, respectively. Thus,
eled off the rust growth only when it was subjected to corrosion progression in WS will only approach that of
acidic drying. The material showed an average decrease plain carbon steel. This hypothesis is supported by the
of 19% in long-frequency (LF-A), and 33% in fact that the rust morphology in all rebars, under alkaline
short-frequency wet-dry (SF-A) relative to other types and low oxygen scenarios, does have striking similarity
after 150 days. The beneficial role of periodic drying to in terms of color, uniformity, and areal extent (Fig. 8).
develop the protective layer in Ni-containing WS is a The superior resistance of WS arises from the for-
prevalent concept in literature (Itou et al. 2000; Usami et mation of a suitable rust with dense structure. Cr and Ni
al. 2003). Acidic water can dissolve the metal to start the addition are known to stabilize such layer in atmospheric
corrosion process during wetting, while the rust is stabi- settings, lowering its corrosion rate than carbon steel.
lized when wetting transitions to drying (Kimura et al. The result presented here indicates that the formation of
2005). Drying also provides a washing action on rebar the layer is impeded, and, therefore, the corrosion re-
surface, giving rise to a uniform rust coating that can sistance was comparable with plain carbon steel in an
repel chlorides away from steel. Figure 8 gives the environment that is continuously wet, has high pH or has
morphologies of corroded samples after 120 days of test, very low oxygen.
from which this supposed coating on NT is observed. It is
worth noting that dense and thin rust layers, corre- 4.2 Mortar test
sponding to IM through SF, are relatively blacker; while 4.2.1 Half-cell potential
AE with thick and loose rust shows red coloration. The variation of Ecorr with time is displayed in Fig. 9.

AE SF LF IM

B-PC

B-CT

B-NT

A-PC

A-CT

A-NT

Fig. 8 Macro-morphology of corroded bars after 120 days of exposure in different pore solutions. (Benito et al. 2019).
E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 377

Note that the x-axis has been changed to logarithmic These conditions shift Ecorr to higher values compared to
scale to amplify the effect of time. Immediately after the those during wetting. Hence, repeating this cycle yields a
first wetting time, Ecorr dropped to values below -300 mV periodic drop and rise of Ecorr as mortar undergoes wet-
vs. CSE. Thereafter, the parameter took a wide range of ting and drying.
values between -200 and -600 mVCSE, which coincide It is worth noting some variations in case C that are not
with the range of uncertain and 90% corrosion risk as per present in case N. In particular, a sudden drop of poten-
the ASTM criteria. Ecorr also remained nearly constant tial to as low as -600 mVCSE is observed in cracked
with time. Crack barely influenced the absolute values of mortar with PC steel beginning at day 30, corresponding
Ecorr throughout the test. This result is apparently the case to wetting stage (Fig. 9). For the same period, CT and NT
for all steel types. Crack supposedly brings the potential stayed equally at -370 mVCSE. During the dry periods, PC
to lesser negative values as it tends to increase the oxy- still exhibited the lowest potential at about -390 mVCSE,
gen content near the steel. A negligible difference in Ecorr followed by -300 mVCSE and -240 for NT and CT, re-
has been observed between case N and C, however, in- spectively. Hence, PC can be considered slightly at risk
dicating that only small corrosion has occurred in of corrosion more than the other types in cracked case.
cracked mortar. Clearer distinction between the types might be achieva-
There is, nonetheless, a systematic pattern on the trend ble with longer exposure.
of Ecorr. It drops during wetting, followed by a rise after It was also observed that the change of potential from
drying. This behavior is associated with two processes: wetting to drying, and vice-versa, is markedly evident in
changes in moisture content inside the mortar that con- cracked mortars. This behavior may have been due to the
trols the oxygen accessing the rebar (Elsener et al. 2003), greater tendency of water to go near and away the rebar
or changes in effective concentration of dissolved metals level in cracked matrix. PC is also observed to have the
in the water-filled pores (Ribeiro et al. 2012). When greatest potential shift from wetting to drying and vice
mortar is saturated, oxygen concentration in its pores versa compared to other steels in cracked cases. This
decreases since gas does not easily dissolve in water. At increase can be likely attributed to Fe3+ on its surface.
the same time, water entering the mortar dilutes the According to previous studies, unalloyed carbon steel
concentration of metallic ions dissolved in pores. Both favors oxide products to form with richer Fe3+ than those
processes bring the steel potential field to more negative in nickel (Tian et al. 2020) or chromium-bearing (Liu et
values. When mortar undergoes drying, oxygen pene- al. 2016) steels. The concentration of this ion dictates the
trates its pores while water simultaneously leaves the rebar potential according to Nernst equation (Stefanoni et
pores, improving the concentration of metallic ions. al. 2019). Since ionic concentration changes as a result of

0
Risk of corrosion as per ASTM C876: No Crack (N)
-100
10% corrosion
-200
-300 Uncertain

-400 90%
corrosion
-500
Severe
-■- PC
Ecorr (mV vs. CSE)

-600
-▲- CT
-700 -●- NT
⊢ ⊣ 1 std. deviation.
-800
0
Risk of corrosion as per ASTM C876: Cracked (C)
-100
10% corrosion
-200
Uncertain
-300
-400 90%
corrosion
-500
Severe
-600 - - PC
- - CT
-700 - - NT
⊢ ⊣ 1 std. deviation
-800
1 10 100
Exposure time (days)
Fig. 9 Change of Ecorr with time arranged according to steel type.
E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 378

water going in and out of mortar (Ribeiro et al. 2012), a between the parameters. Such weak correlation, and the
higher shift of potential in PC may have resulted from a fact that Ecorr lies within uncertain and 90% corrosion
higher number of charged particles with higher equilib- risk as shown in Fig. 9, is in agreement with very few
rium potential, i.e., Fe3+. rusted areas observed in Fig. 10. From these results, steel
type appears to have no remarkable influence.
4.2.2 Surface condition
Figure 10(a) shows the actual state of different steel bars 4.2.3 Progress of degree of corrosion
after 90 and 120 days of wet-dry cycle. As observed, all Since Ecorr reflects only the corrosion risk, average mass
the steels present comparable surface condition, con- loss allows for the assessment of corrosion kinetics
taining only few localized rusts and some coloration due quantitatively. The plot of the parameter with time (Fig.
to rust stains. These rusts were detected to cover only less 12) suggests no observable difference between cracked
than 5% of the exposed surfaces [Fig. 10(b)]. Clearly, no and uncracked mortars, consistent with the result of
severe corrosion has occurred. Similar to IM-B case of corrosion potential. This may have been due to limited
solution test, this observation is attributed to the fact that oxygen or chloride that accessed the rebars through the
corrosion of steel is restricted in an environment of high cracks. Figure 12 also reveals that NT actually sustained
pH and low oxygen. Consequently, localized rust tends to higher mass loss in mortar by as much as 70% than other
form regardless of steel type within the period of wet-dry types at the end of the test. To expound this data further,
cycle considered. Fig. 13 was prepared to compare the result in mortars
Fig. 11 relates the values of Ecorr with the actual de- with those in chloride-contaminated alkaline solutions.
grees of damage given by the corroded area and the It is apparent that the sequence of steel type in terms
corrosion rate. There is negligible correlation found of corrosion degree is not the same across the experi-

(a)
10
Rust stain coverage (%)

90 days 120 days


8
⊢ ⊣ 1 std. deviation ⊢ ⊣ 1 std. deviation
6 PC
4 PC
PC PC CT NT
2 CT NT CT NT CT NT

0
No Crack Cracked No Crack Cracked
(b)
Fig. 10 (a) Surface macro-morphology, and (b) rust coverage of steel bars in different conditions after 90 and 120 days of
wet-dry cycle.

0
Steel Uncracked Cracked
y = 5.79x - 363.91 y = -21.27x - 319.13
-100
Ecorr (mV vs. CSE)

PC ■ □
R² = 0.09 R² = 0.08 10% corrosion
CT ▲ △
-200
NT ● ○
Uncertain
-300
⊢ ⊣ 1 std. deviation
-400
90% corrosion
-500
Severe
-600
0 3 6 9 12 0 1 2 3 4
Corrosion rate (g/dm2/yr) Rust stain coverage (%)
Fig. 11 Relationship of Ecorr with average corrosion rate and rust stain coverage.
E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 379

mental conditions considered. In particular, NT showed Results for uncracked mortar [Fig. 13(a)] and for
the highest damage in sound mortar at 2.1 g/dm2 [Fig. IM-B solution [Fig. 13(b)] were compared because both
13(a)]. However, the trend is not consistent with the cases do not permit air near the steel bars, thus allowing
result in alkaline liquid phase of similar condition, i.e., only localized pits to occur. This was indeed the case
continuous immersion [IM-B, Fig. 13(b)]. The trend is considering the observed surface appearances in Figs. 8
also not consistent for cracked mortar [Fig. 13(c)] and and 10. However, since the rusts that formed were few
its corresponding liquid test condition, i.e., cyclic and localized in the said cases, no difference between
wet-dry [SF-B, Fig. 13(d)]. The measured values were steel types could be deduced. On the other hand, the
compared with those in aerated alkaline solution [AE-B, result of cracked mortars to SF-B and AE-B solution
Fig. 13(e)], whose corrosivity is much more severe. As were compared because, under these cases, air interacts
observed, the difference between the steel types be- with rebars, although at different degrees. In AE-B solu-
comes far less apparent. In other words, the average tion [Fig. 13(e)], even though abundant air is present,
corrosion in both sound and cracked mortars was very the rebars are permanently in contact with water, which
low (≤ 2.1 g/dm2) that the distinctions between the steel accelerates the growth of rust products in all steels. The
types given in Figs. 13(a) and (c) are actually insignifi- rusts are forced to absorb water, and thus, produce thick
cant. This is because open air (oxygen), in addition to and loose layer (Matsushima et al. 1974). Meanwhile,
water, is necessary for the formation of protective rust in under frequent wet-dry cycle or SF-B solution [Fig.
weathering steels (Morcillo et al. 2014, 2019). 13(d)], water is deposited and drained periodically,

5
Corrosion degree (g/dm2)

4 No crack Cracked
⊢ ⊣ 1 standard dev. ⊢ ⊣ 1 standard dev.
3

2 NT NT
1 PC CT
CT PC
0
0 30 60 90 120 150 0 30 60 90 120 150
Exposure time (days)
Fig. 12 Change of average mass loss due to corrosion with time in mortar.

25
(a) Uncracked (b) IM-B
20 mortar Solution ■ 60 days
■ 90 days
15
■ 120 days
10
IM-B: Continuous immersion
Corrosion degree (g/dm2)

5 SF-B: Cyclic wetting and drying


2.1 1.9 2.3 2.1
0.62 0.52 AE-B: Continuous aeration
0
PC CT NT PC CT NT
25
(c) Cracked (d) SF-B (e) AE-BSolution
20 mortar Solution 18.9
15.8 16.5
15

10

5
1.8 2.6 2.4 2.1
0.80 0.62
0
PC CT NT PC CT NT PC CT NT

Steel type
Fig. 13 Comparison of corrosion degree in mortar and liquid phase from 60 to 120 days of exposure.
E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 380

which allowed the rusting process to progress gradually. 5) Besides water, air plays a role in forming the protec-
However, the rust only covered partial areas of the bars tive rust in Ni-type WS. Thus, wet-dry condition
within the period of the investigation, which then leads seems to be more favorable for NT than mortar em-
to limited protection on NT. The rust layer under SF-B bedment. Only localized rust has formed in mortar
solution has not yet fully developed after 120 days as after 120 days of wet-dry cycle, covering only 5% of
shown in Fig. 8. In cracked mortar [Fig. 13(c)], the the total surface area. NT sustained higher corrosion
corrosion resistance of NT was not evident mainly due by about 70% than other types but the values are
to limited oxygen and water. In other words, the devel- merely 2.1 g/dm2 or lower, which can be considered
opment of protective layer on NT is dependent on the insignificant given that only few products have
suitable balance of air and water in alkaline environ- formed. In summary, within the scope of this study,
ment as in the case of wet-dry state, rather than just on the behavior of NT in mortar remains practically the
the complete absence or presence of water. It is con- same as those of conventional steels.
cluded from the foregoing discussion that high-Ni type 6) In addition to possible development of rust layer, the
WS bars in mortar behaves practically the same way as result of the study can be advanced by determining the
the conventional steels, which is contrary to the authors’ effect of Ni or Cr on the critical chloride content and
hypothesis. time of corrosion initiation of WS. Future research
It must be emphasized that Ni-type WS obtains its re- may also be done focusing on advanced electro-
sistance from the appearance of a uniform and dense chemical measurements and microstructural analysis
rust with chloride-repelling property. Because localized of rust products of WS in mortar or concrete.
corrosion only formed within the scope of this study, the
decrease of corrosion damage in NT could not be estab- Acknowledgements
lished. It is possible that that rusting in NT happens The authors acknowledge Nippon Steel and Sumitomo
quickly for short period, but it will eventually stabilize Metal Corp. and Amano Myutech Corp. for providing
and cover larger areas as it forms for longer period (Wei the steel materials for the investigation. This work is
et al. 2019). Thus, in the future, the possibility of form- part of the first author’s master studies, supervised by
ing the protective rust using longer exposure and wider the preceding authors, under the grant from Japan Govt.
cracks should be further explored. through the MEXT scholarship.

5. Conclusions References
Angst, U. M., Geiker, M. R., Michel, A., Gehlen, C.,
1) High-Cr type WS (CT) showed 30 and 60% corrosion Wong, H., Isgor, O. B., Elsener, B., Hansson, C. M.,
reduction compared to other steels under basic and François, R., Hornbostel, K., Polder, R., Alonso, M. C.,
acidic case, respectively, only when it was put con- Sanchez, M., Correia, M. J., Criado, M., Sagüés, A.
tinuously in chloride-containing water with abundant and Buenfeld, N., (2017). “The steel-concrete interface.”
air. There is no difference, however, when oxygen is Materials and Structures, 50, Article No. 143.
low. ASTM C876, (2009). “Standard test method for
2) High-Ni WS (NT) reduced the corrosion degree by corrosion potentials of uncoated reinforcing steel in
19% to 30% relative to other steels under acidic and concrete.” West Conshohocken, PA, USA: American
cyclic wet-dry state, consistent with its weathering Society for Testing and Materials.
application. It appears that dense and chlo- ASTM G1, (2003). “Standard practice for preparing,
ride-repelling rust is favored by faster wet-dry cycle. cleaning, and evaluating corrosion test specimens.”
3) NT did not reduce the corrosion degree in environ- West Conshohocken, PA, USA: American Society for
ment that remains continuously wet, is highly alkaline, Testing and Materials.
or has low oxygen. The first item is explained by the Benito, E. K. D., Ueno, A. and Fukuyama, T., (2019).
fact that continuous immersion in chloride-containing “Basic study on corrosion progression in high-nickel
water accelerates the corrosion process drastically, weathering steel by simulated liquid test.” Proceedings
resulting to rust products that absorb water. Whereas, of the Japan Concrete Institute, 41(1), 689-694.
in a solution of high pH or low oxygen, corrosion rate Cano, H., Díaz, I., de la Fuente, D., Chico, B. and
is inhibited so that the desirable dense and uniform Morcillo, M., (2018). “Effect of Cu, Cr and Ni
rust film does not form. Consequently, the corrosion alloying elements on mechanical properties and
behavior of WS approaches that of carbon steel. atmospheric corrosion resistance of weathering steels
4) Although all steels in intact mortar assumed similar in marine atmospheres of different aggressivities.”
Ecorr values, plain carbon steel (PC) were slightly Materials and Corrosion, 69(1), 8-19.
more active than WS in cracked mortar. Furthermore, Cheng, X. Q., Tian, Y. W., Li, X. G. and Zhou, C., (2014).
the change of potential from wetting to drying and “Corrosion behavior of nickel-containing weathering
vice versa is highest in PC, which may have been steel in simulated marine atmospheric environment.”
caused by the difference in concentration of ferrous Materials and Corrosion, 65(10), 1033-1037.
ions on surface of each steel. Chrisp, T. M., McCarter, W. J., Starrs, G., Basheer, P. A.
E. K. D. Benito, A. Ueno and T. Fukuyama / Journal of Advanced Concrete Technology Vol. 19, 370-381, 2021 381

M. and Blewett, J., (2002). “Depth-related variation in Technical Report, 87, 17-20.
conductivity to study cover-zone concrete during Liu, M., Cheng, X., Li, X., Pan, Y. and Li, J., (2016).
wetting and drying.” Cement and Concrete Composites, “Effect of Cr on the passive film formation mechanism
24(5), 415-426. of steel rebar in saturated calcium hydroxide solution.”
Dasar, A., Hamada, H., Sagawa, Y. and Yamamoto, D., Applied Surface Science, 389, 1182-1191.
(2017). “Electrochemical behavior of steel in cracked Matsushima, I., Ishizu, Y., Ueno, T., Kanazashi, M. and
concrete - Influence of crack width, cover, exposure Horikawa, K., (1974). “Effect of structural and
conditions and supplementary cementitious materials environmental factors in the practical use of low-alloy
(SCMs).” Proceedings of the Japan Concrete Institute, weathering steel.” Corrosion Engineering, 23(4),
39(1), 1045-1050. 177-182.
Elsener, B., (2002). “Macrocell corrosion of steel in Morcillo, M., Díaz, I., Cano, H., Chico, B. and de la
concrete - Implications for corrosion monitoring.” Fuente, D., (2019). “Atmospheric corrosion of
Cement and Concrete Composites, 24(1), 65-72. weathering steels. Overview for engineers. Part I:
Elsener, B., Gulikers, J., Polder, R. and Raupach, M., Basic concepts.” Construction and Building Materials,
(2003). “Half-cell potential measurements - Potential 213, 723-737.
mapping on reinforced concrete structures.” Materials Morcillo, M., Díaz, I., Chico, B., Cano, H. and de la
and Structures, 36, 461-471. Fuente, D., (2014). “Weathering steels: From empirical
Fujiwara, H., Nakayama, A., Fujii, T. and Ayano, T., development to scientific design. A review.” Corrosion
(2017). “Effect of blast furnace slag on steel rod Science, 83, 6-31.
corrosion in concrete.” In: Proceedings of the 17th Pourbaix, M., Van Muylder, J., Porubaix, A. and Kissel,
JSMS Symposium on Concrete Structure Scenarios, J., (1980). “An electrochemical wet and dry method
Kyoto, Japan 12-13 October 2017. Tokyo: The Society for atmospheric corrosion testing.” Rapports Techniques
of Materials Science, Japan, 29-34. CEBELCOR, 139, RT 259.
Hansson, C. M., (1984). “Comments on electrochemical Ribeiro, D. V., Labrincha, J. A. and Morelli, M. R.,
measurements of the rate of corrosion of steel in (2012). “Effect of the addition of red mud on the
concrete.” Cement and Concrete Research, 14(4), corrosion parameters of reinforced concrete.” Cement
574-584. and Concrete Research, 42(1), 124-133.
Itou, M., Tanabe, K., Ito, S., Kusunoki, T., Usami, A., Shi, J., Wang, D., Ming, J. and Sun, W., (2018). “Long-term
Kihira, H., Tsuzuki, T. and Tomita, Y., (2000). electrochemical behavior of low-alloy steel in
“Performances of coastal weathering steel.” Nippon simulated concrete pore solution with chlorides.”
Steel Technical Report, 81, 79-84. Journal of Materials in Civil Engineering, 30(4), 1-11.
Kihira, H., Usami, A., Tanabe, K., Ito, M., Shigesato, G., Stefanoni, M., Angst, U. M. and Elsener, B., (2019).
Tomita, Y., Kusunoki, T., Tzuzuki, T., Ito, S. and “Kinetics of electrochemical dissolution of metals in
Murata, T., (1999). “Development of weathering steel porous media.” Nature Materials, 18, 942-947.
for coastal atmosphere.” In: P. M. Natishan, S. Ito, D. Tian, Y., Dong, C., Wang, G., Cheng, X. and Li, X.,
A. Shifler and T. Tsuru, Eds. Proceedings of the (2020). “The effect of nickel on corrosion behaviour of
International Symposium on Corrosion and Corrosion high-strength low alloy steel rebar in simulated
Control in Saltwater Environments, Honolulu, Hawaii concrete pore solution.” Construction and Building
17-22 October 1999. New Jersey, USA: The Materials, 246, Article ID 118462.
Electrochemical Society, Inc., 127-136. Usami, A., Kihira, H. and Kusunoki, T., (2003). “3%-Ni
Kimura, M., Kihira, H., Nomura, M. and Kitajima, Y., weathering steel plate for uncoated bridges at high
(2004). “Corrosion protection mechanism of the airborne salt environment.” Nippon Steel Technical
advanced weathering steel (Fe-3.0Ni-0.40Cu, mass%) Report, 87, 21-23.
in a coastal area.” In: D. A. Shifler, T. Tsuru, P. M. Wei, J., Wang, C. G., Wei, X., Mu, X., He, X. Y., Dong, J.
Natishan and S. Ito, Eds. Proceedings of the H. and Ke, W., (2019). “Corrosion evolution of steel
International Symposium on Corrosion and Corrosion reinforced concrete under simulated tidal and
Control in Saltwater Environments II, Honolulu, immersion zones of marine environment.” Acta
Hawaii 3-8 October 2004. New Jersey, USA: The Metallurgica Sinica (English Letters), 32(7), 900-912.
Electrochemical Society, Inc., 133-142. Wu, W., Zeng, Z., Cheng, X., Li, X. and Liu, B., (2017).
Kimura, M., Kihira, H., Ohta, N., Hashimoto, M. and “Atmospheric corrosion behavior and mechanism of a
Senuma, T., (2005). “Control of Fe(O,OH)6 nano-network Ni-advanced weathering steel in simulated tropical
structures of rust for high atmospheric-corrosion marine environment.” Journal of Materials Engineering
resistance.” Corrosion Science, 47(10), 2499-2509. and Performance, 26(12), 6075-6086.
Kimura, M., Shigesato, G., Tanabe, K., Suzuki, T. and Yamashita, M., Konishi, H., Mizuki, J. and Uchida, H.,
Kihira, H., (2003). “Fe(O,OH)6 network structure of (2004). “Nanostructure of protective rust layer on
rust formed on weathering steel surface and its weathering steel examined using synchrotron radiation
relationship with corrosion resistance.” Nippon Steel x-rays.” Materials Transactions, 45(6), 1920-1924.

You might also like