Вы находитесь на странице: 1из 8

Available online at www.sciencedirect.

com

ScienceDirect
Procedia Engineering 172 (2017) 859 – 866

Modern Building Materials, Structures and Techniques, MBMST 2016

Problems connected with use of hot-dip galvanized reinforcement in


concrete elements
Radka Pernicovaa, Daniel Dobiasa, Petr Pokornya *
a
Klokner Institute of CTU ,Solinova 7, Prague 166 08, Czech Republic

Abstract

The goal of this article is to evaluate risks arising from using hot-dip galvanized reinforcement in concrete elements. The article
provides detailed summary of current experimental activities but also earlier positive remarks about applicability of hot-dip
galvanized reinforcement, mainly from the perspective of corrosion and bond strength with concrete. Based on previously
obtained data, the article disproves barrier effect of zinc coating on base steel. The reason is the initial corrosion reaction of zinc
coating in fresh concrete producing hydrogen. Further data prove that forming hydrogen irreversibly increases porosity of cement
on reinforcement/concrete interface which can significantly reduce bond strength. Sufficient filling of pores by zinc corrosion
product could not be confirmed.

©
© 2017
2016TheTheAuthors. Published
Authors. by Elsevier
Published Ltd. Ltd.
by Elsevier This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Peer-review under responsibility of the organizing committee of MBMST 2016.
Peer-review under responsibility of the organizing committee of MBMST 2016
Keywords: Hod-dip galvanized; steel reinforcement; concrete; porosity; corosion; phase interface.

1. Introduction

Unacceptable corrosion rate of common unalloyed steel caused by carbonation or, more frequently,
contamination of concrete by chlorides significantly limits functional lifetime of steel-concrete constructions. It is in
fact the voluminous corrosion products of steel which compromise the integrity of covering layer of concrete,
eventually requiring fast and expensive countermeasures [1,2].

* Radka Pernicová. Tel.: +420 224 35 5313.


E-mail address: radka.pernicova@cvut.cz

1877-7058 © 2017 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Peer-review under responsibility of the organizing committee of MBMST 2016
doi:10.1016/j.proeng.2017.02.086
860 Radka Pernicova et al. / Procedia Engineering 172 (2017) 859 – 866

Despite that the research focused on use of alternative reinforcing materials (e.g. fabric, stainless steel, alkaline-
resistant glass, polymer and carbon fiber) has greatly progressed, their application in common constructions is still
marginal (typically only as complementary reinforcement – e.g. for strengthening of common ribbed carbon steel
reinforcement). Some alternative materials do not provide concrete with sufficient mechanical properties,
respectively, their benefit is less effective compared to common steel. Large scale application of the other materials
would greatly increase the construction [1,3].
For this reason, common rebar from unalloyed (carbon) steel with surface geometry adjusted to form
standardized ribs or imprints. Corrosion protection of such reinforcement is facilitated almost exclusively by
ensuring sufficient thickness of covering concrete and increasing the resistance of concrete against chloride and CO 2
penetration (e.g. by reduction of w/c factor, increasing prolonging the concrete treatment period and its quality,
application of more suitable cement types).
Use of other anti-corrosion measures is uncommon. Deployment of corrosion inhibitors seems to be only mildly
effective because of its inability to maintain critical concentration over long term on steel/concrete interface).
Application of cathodic protection (sacrificial anode or by DC source) cannot be used everywhere and is
unambiguously on of the more expensive anti-corrosion systems [2,3].
Evaluation of effectivity of protective coatings for common steel rebar is the focus of expert community since the
beginning of the last century. Reasons are logical – possibility of fast and easy application of corrosion protection
system (coating) without the need to maintain it while also keeping the mechanical properties of reinforcing steel.
From this point of view, epoxy coatings and coatings produced by hot-dip galvanizing are of the highest interest.
The economic studies show that application of hot-dip galvanized coatings for steel rebar protection does not
significantly increase the construction expenses. Idea of their application is further supported by the data of their
resistance to atmospheric corrosion. Use of standard-prescribed thickness, sufficiently thick coating (often a
combination of coatings: galvanized coating and organic coating) can assure required longevity of given
construction [3,4].
From the perspective of corrosion engineering, protective coatings provide only barrier protection (unless it
provides any other: either by destimulation, inhibition or electrochemical mechanism) i.e. they work as an
“insulation” separating the susceptible material from aggressive environment. They ensure longer time until
activation of steel which will then corrode at unhindered rate. In reality, transport of oxygen and humidity from the
surface to the rebar also has some effect.
This article strictly focuses on coatings produced by hot-dip galvanizing [1,4].

Nomenclature

A Surface treatment of steel by hot-dip galvanizing


B Unclear key questions about application of hot-dip -galvanized steel in concrete
C Evaluation of initial corrosion damage to zinc coating in fresh concrete
D Extent of effect of initial corrosion damage on rebar/concrete bond strength

2. Surface treatment of steel by hot-dip galvanizing

During hot-dip galvanizing (liquid zinc of 450-470 °C) alloy-like iron-zinc coating grows on coated components.
This is a result complex process of reciprocal diffusion of both metals resulting in elementary intermetallic bonds
and subsequently phase transformations. These phenomena take place in both the surface layer of zinced metal but
also at the interface of solid metal and liquid alloy. Depending on steel composition, temperature, liquid alloy
composition, wall thickness of zinced component, coating time, surface state and mode and rate of cooling, different
intermetallic Fe-Zn compounds can be formed. Appearance of typical hot-dip galvanized coating is shown in Fig.
1(a) [5,6]. Coating layers vary in composition and thickness. Iron content increases in the direction toward the steel
substrate. The layers are denoted by Greek alphabet letters: gamma (Γ), gamma1 (Γ1), delta (δ), zeta (ζ), eventually
eta (η). Individual phases significantly differ not only in composition and grain morphology but also in mechanical
properties [6].
Radka Pernicova et al. / Procedia Engineering 172 (2017) 859 – 866 861

The top layer, so-called η-phase, comprises basically from pure zinc and forms by plain solidification of liquid
zinc. From metallurgical point of view, this phase is defined as solid substitutional solution of iron in zinc (iron
content is cca 0.03 wt. %). Zinc coating relief corresponds to the local unevenness of the substrate surface, but the
final roughness is primarily affected by grains of solidifying η-phase. Zinc solidifies in hexagonal system (hcp) and
is characteristic for low hardness and high ductility at normal temperatures. Intermetallic phase zeta (ζ) can be
described as FeZn13. Iron content is 5 – 6.2 wt. %. In coating, it crystallizes in basal centred monoclinic system. In
the case of delta (δ) phase, the existence of two different crystalline structures is discussed: δ1k (or δ) and δ1p (or δ1)
which correspond to FeZn7 (δ1k), resp. FeZn10 (δ1p) (or also Fe13Zn126). New works on this topic suggest that both
delta phases exist. Both previously mentioned intermetallic compounds crystallize in hexagonal system. Iron content
in these phase is around 7 – 11.5 wt. %. Bottom layer comprises of two gamma phases (Γ1 + Γ), iron content is
about 23.5 – 28 wt. %. Phase Γ1 (e.g. Fe5Zn21) crystallizes in face centred cubic (fcc) system and Γ phase (Fe3Zn10)
in body centred cubic (bcc). Arrangement of individual phases in the coating is depicted in Fig. 1(b) [5-7].

Fig. 1. (a) Typical appearance and thickness of hot-dip galvanized coating on steel; (b) schema of individual intermetallic phase arrangement in
the coating.

3. Unclear key questions about application of hot-dip galvanized steel in concrete

Mass application of epoxy-coatings for steel rebar protection is limited by its reduced bond-strength with
concrete. This fact was unambiguously confirmed during years by data from pull-out and beam-tests. Prediction of
life-prolongation of construction considering the thickness of defect-free protection coating is in accordance with
results from corrosion tests. Thusly protected rebar is locally used for example for reinforcement of coastal
constructions – bridges and wharfs in the USA. In these constructions, lowered bond-strength was considered
during static design [3, 4].
Long-term research projects focused on effectivity of anti-corrosion protection of steel rebar by galvanized zinc
produce, compared to epoxy coatings, contradicting results. The expert public only agrees on several topics:

x during the galvanizing, mechanical properties of steel rebar are not negatively affected;
x rebar corrodes in active state in fresh concrete for certain amount of time, producing hydrogen. The surface
is quickly covered by corrosion product based on Ca[Zn(OH)3]2·2H2O [3,8,9].

Effect of initial corrosion damage of zinced rebar on future barrier effect is topic dividing the expert public.
Simultaneously, the question if hydrogen and growing corrosion products negatively affect bond-strength with
concrete was never fully answered. The same applies for the corrosion products of zinc.

4. Extent of effect of initial corrosion damage in fresh concrete

Basic literature source, focusing on this problem is the “Galvanized steel reinforcement in concrete” monography
[10] by professor Stephen R. Yeomans. However, apart from this work, older papers can be found that, based on
862 Radka Pernicova et al. / Procedia Engineering 172 (2017) 859 – 866

realized experiments suggest either decreased bond-strength of zinced rebar or unexpectedly short effectivity of
coating in standard amount of chlorides [11,12,13]. Records of smaller constructions reinforced by zinced rebar with
catastrophic results during construction stripping (load capacity reduced by decreased bond strength caused by zinc
corrosion in fresh concrete) [9,14,15]. To present day, the feasibility of application of zinced rebar in concrete is still
doubted.
Most fundamental result found by the group in 10 years can be summarized as follows: “Corrosion of zinc in
cement is closely related to pH value of pore solution, however up to pH 13.3, zinced steel corrodes in passive
state.” The reason is formation continuous and very compact coating of Ca[Zn(OH)3]2·2H2O. Above pH 13.3, the
coating corrodes in active state because its surface is no longer covered by complex zincate (corrosion products
comprise of ZnO and ε-Zn(OH)2). If the coating does not comprise of η-phase (i.e. the outer-most layer is the ζ
phase – FeZn13), its ability to transient to corrosion in passive state is greatly limited since the lowered amount of
zinc in intermetallic hinders the formation of compact passive layer of complex zincate [10, 16-20].
Our experience and experimental findings about corrosion behaviour of zinced steel in fresh and curing concrete
significantly differ from the finding in works mentioned above. We studied the effect of pH and calcium presence
(in form of Ca(OH)2) on corrosion behaviour of hot-dip galvanized steel in model environment simulating pore
solution of pH 12.6; 13.0 and 13.5. Environment of pH 12.6 (saturated solution of Ca(OH)2) models concrete pore
solution without the addition of KOH. However, since concrete contains certain amount of alkaline metal oxides
(K2O or Na2O) which with water form corresponding hydroxides with (KOH or NaOH), real pH of fresh pore
solution is rather 13.0 (for pure Portland cement up to 13.5).
During exposition of hot-dip galvanized steel (coating thickness was 90 μm ± 10 μm) in model pore solution of
pH 12.6 (all expositions took 6 days in closed containers), the surface was covered by fine crystals of
Ca[Zn(OH)3]2·2H2O. Appearance of the coating on cross-cut reveals only thin layer of corrosion products on the
surface, the coating is otherwise undamaged.
Apparent disruption of coating integrity was observed only after exposure in pore solution of pH 13.0. Sample
surface is clearly covered by zincate crystals which are much larger than in the previous test. The thickness of
corrosion products measured on the cross-cut view was about 10 μm. Corrosion damage was found in the outer η
phase. This phase dissolves in exposure environment and forms continuous layer of corrosion products on its surface
which easily descale. Small-scale disintegration in the layer structure is also apparent. Otherwise, the ζ phase,
cluster of δ phases and also cluster of Γ phases seems undamaged.
Surface of samples exposed in environment of pH 13.5 was structured, covered with large plate-like zincate
crystals. Cross-cut section of the coating shows that its integrity is strongly damaged after 6-weeks across the whole
coating thickness. It seemed that the η phase fully decomposed, i.e. part of it was dissolved and the non-adherent
remaining part was lost during the sample manipulation or during sample preparation for metallographic analysis.

Fig. 2. Surface of hot-dip galvanized steel after exposure in environment of pH (a) 12. 6 (b) 13.0 (c) 13.5.
Radka Pernicova et al. / Procedia Engineering 172 (2017) 859 – 866 863

Fig. 3. Cross-cut section of hot-dip galvanized coating in environment of pH (a) 12.6 (b) 13.0 (c) 13.5.

Results of parallel electrochemical corrosion experiments well correlate with discovered findings. During
exposition in environment of pH 12.6, the sample was corroded in active state producing hydrogen and after cca 3
days, the transient to passive state occurred. However, most of the sample exposed at pH 13.0 this transient does not
occur even after 6 days. The situation is similar for samples exposed at pH = 13.5. It is evident that surface was
completely covered by corrosion products based on Ca[Zn(OH)3]2·2H2O. This was confirmed by XRD analysis of
corrosion products. Size of produced zincate crystals increases rapidly with rising pH. Nevertheless, the compact
layer of corrosion product is not capable of preventing corrosion of zinced steel in active state. Therefore, it seems
that presence of calcium cations destabilizes ZnO and ε-Zn(OH) which normally facilitate transient to passive state
in (pH 12.6 to 13.0 in presence of KOH).
Evaluation of corrosion behaviour of hot-dip galvanized steel in fresh concrete is of great importance. Our data
from resistometric probes prepared from pure zinc foil, placed in concrete suggest that zinc is not capable of
transient to passive state in such environment. Apparent zinc passivity, i.e. detectable decrease of corrosion rate
corresponds to limited availability of water on the sample surface (after the concrete sample has dried); resp. the
corrosion rate of zinc is limited by transport of water to its surface. As it was previously stated that η-phase is in fact
pure zinc, previously published conclusions about exposition of zinc in model pore solutions can be thusly
confirmed. Corrosion products of Ca[Zn(OH)3]2·2H2O on the surface of zinced steel in concrete do not indeed
assure transient of coating to passive state [21]. This fact is partly shown on cross-section of zinc coating on steel
placed in cement paste (in situ test – use of pure Portland cement CEM I) for 28 days. For evaluation, the containers
with samples were placed in a bath – either on a grate (eq. humidity above open water surface of 95 % RH) or in the
water itself. Samples were then segmented and processed to metallographic cross-cuts. Results are shown in Fig. 4
(a) and (b). The appearance of reference sample produced on steel which was not placed in cement paste is shown in
Fig. 4 (c). It is apparent that damage to the sample placed in humid atmosphere (attack of η phase and only
peripheral attack of ζ phase) is much lesser compared to the sample placed under the water surface. The η phase is
unequivocally dissolved and the ζ phase is strongly damaged. Strong cracks reach across the whole coating
(emphasized by separation step and sample grinding), the cracks also disrupt the compactness of δ phase clusters. It
is important to remind about the results confirming the conclusions acquired by resistometric method [21].
864 Radka Pernicova et al. / Procedia Engineering 172 (2017) 859 – 866

Fig. 4. Cross-cut of hot-dip galvanized steel placed in steel for 28 days and exposed in: (a) humid atmosphere (95 % R.H.); (b) under water
surface (c) reference sample.

5. Extent of effect of initial corrosion damage on rebar/concrete bond strength

Previously mentioned work of professor Yeomans widely discusses effect of initial corrosion of zinced rebar on
bond strength with concrete. The author collaborates on this project with expert on bond- strength: O. Kayali.
Authors verified bond strength of both smooth and ribbed rebar with zinc coating and compared it to bond strength
of non-coated rebar of similar geometry. The aim of extensive experimental work was to verify if the initial
corrosion of zinc coating, evolution of hydrogen and formation of corrosion products based on zincate affects bond
strength between rebar and concrete. Authors proved by standardised tests (pull-out test and beam test) that bond
strength is indeed not affected. Authors described negative effect on evolution of porosity of cement on phase
interface (increased porosity), however, they acknowledge that the effect of hydrogen on total bond-strength is
marginal. The effect of Ca[Zn(OH)3]2·2H2O corrosion products is not discussed in detail [10].
As suggested in introduction, in the course of a century, contradicting results were also reported. For example –
authors reported slight reduction of bond strength of zinced rebar with concrete, however cases of significant
reduction of load-bearing capacity caused by reduced bond strength were also reported. Reasons for these
phenomena is, according to experts, increase of the porosity of cement, growth of corrosion products of
Ca[Zn(OH)3]2·2H2O or just reduction of rebar diameter due to corrosion, probably even in curing concrete [14,15].
Comparing the bond strength of smooth zinced bars with NSC grade concrete (normal strength concrete) to bond
strength of non-coated steel bars by pull-out test unequivocally proves that initial corrosion of zinc coating in fresh
concrete and probably not even the expected increase of corrosion rate in case of samples placed under water cannot
be neglected [22].
It is apparent that hydrogen, evolving from cathodic depolarization, accompanying zinc corrosion, negatively
affects porosity of cement. This can be confirmed by the pictures of cement collected from phase interface of zinced
rebar/cement (Portland cement CEM I) after 28 days of curing. Fig. 5 shows the cement from the phase interface,
specifically cement curing in humid environment (95 %) and in water. For comparison, Fig. 6 (a) shows appearance
of cement from phase interface of non-coated steel/cement. These results show that even after 28 days, the pores are
not filled with corrosion products, i.e. the pore structure is preserved. Pore structure formed in real concrete can be
observed even after above mentioned bond strength test of hot-dip zinc galvanized smooth specimen with concrete
Fig. 6 (b). These results are in line with results reported in other literature sources which evaluated the porosity by
mercury porosimetry [23,24].
Radka Pernicova et al. / Procedia Engineering 172 (2017) 859 – 866 865

Fig. 5. Surface of cement on hot-dip galvanized steel interface after 28 days of concrete curing in: (a) atmosphere of 95 % RH; (b) in water.

Fig. 6. Surface of cement on phase interface (a) surface of non-coated steel after 28 days of curing in atmosphere of 95 % RH; (b) surface of
smooth zinced bar after bond-strength test.

6. Conclusion

Hot-dip galvanizing of common carbon steel reinforcement can provide certain advantages. In theory, it can
significantly prolong lifetime of the construction given the concrete with carbonation effect. It is however important
to mention that hot-dip galvanized zinc only prolongs time to activation of substrate steel, i.e. until the zinc is
consumed by corrosion reaction.
Predicted problems corresponding to use of zinced rebar in concrete are also legitimate and cannot be
overlooked. One of the problems is definitely the initial corrosion of zinc surface in fresh concrete. Initial corrosion
damage results in reduced bond strength with concrete; the mechanism involves hydrogen evolution increases of
porosity of cement; expansion of corrosion products based on Ca[Zn(OH)3]2·2H2O, compromising the compactness
of cement.
Different findings based on our results based on evaluation of corrosion behaviour of zinced steel in both fresh
and curing concrete and its effect on bond strength can be explained by different pH value of tested fresh cement
mixtures. Effect of other chemical additives in modern concrete mixtures (e.g. plastificators, fluiders and aeration
additives) also cannot be neglected. Evaluation of published results can be complicated due to unclear definition of
experimental conditions, sample production, their processing and properties – composition of used cement. In some
cases, the process of bond strength testing is not described, resp. real arrangement. For these reasons, published
information cannot be always easily compared.
New experimental results confirm the fact that corrosion behaviour of zinced steel in concrete is affected by
866 Radka Pernicova et al. / Procedia Engineering 172 (2017) 859 – 866

sufficient supply of water, resp. change of humidity of pore environment. It was documented that increase of
humidity is accompanied by re-initiation of corrosion of zinc in concrete. It is also apparent that zinc surface is
unable to passivate just by corrosion products based on Ca[Zn(OH)3]2·2H2O. Sufficient supply of water/humidity
thus supports further corrosion damage of zinced steel. Bond strength of zinced reinforcement with concrete can be
further reduced by reduction of its diameter by reducing the thickness of zinc coating. Essential is the effect of water
transport to rebar surface; it can explain, in literature often mentioned, failure of this anti-corrosion system even in
environment with normal levels of chlorides.
Despite some advantages provided by hot-dip coating, it is noted that feasibility of zinced rebar is questionable,
since the conditions of fresh concrete mixture lead to damage to zinc coating and, more importantly, its bond-
strength with cement.
Designers of concrete structures with zinced reinforcement should follow construction guides for reinforcement
which consider the danger of reducing bond strength of the rebar with concrete from the point of view of ensuring
the requirements of first and second limit state.

Acknowledgements

This research has been supported by GACR Agency of the Czech Republic No. GACR 14-20856S.

References

[1] M. Collepardi, Moderní beton, 1. ed., ČKAIT v edici betonové stavitelství, Pelhřimov, 2009.
[2] P. Aitcin, Vysokohodnotný beton, 2. ed., ČKAIT v edici betonové stavitelství, Pelhřimov, 2005.
[3] P. Pokorný, Vliv koroze zinkované oceli na soudržnost s betonem, Koroze a ochrana materiálu 56(4) (2012) 119-135.
[4] D.E. Tonini D. a kol., Chloride corrosion of steel in concrete, American Society for Testing and Materials, Baltimore 1977.
[5] A. R.Marder, The metallurgy of zinc – coated steel, Progress in Materials Science 45 (2000) 191-271.
[6] V. Kuklík, J. Kudláček, Žárové zinkování, 1. ed., AČSZ, Praha 2014.
[7] P. Pokorný a kol., Description of structure of Fe-Zn intermetallic compounds present in hot-dip galvanized coatings on steel, Metalurgija
54(4), 707-710.
[8] P. Pokorný, Kritická diskuze k vlivu přídavku chemických látek do cementu na korozní chování zinkované oceli, Koroze a ochrana materiálu
58(1) (2014) 31-35.
[9] G. Rehm, A. Lämmke, Untersuchungen über Reaktionen des Zinks unter Einwirkung von Alkalien im Hinblick auf das Verhalten verzinkter
Stähle im Beton, Betonstein-Zeitung 6 (1970) 360-365.
[10] S. R.Yeomans, Galvanized steel reinforcement in concrete, 2. ed., Elsevier, Canberra, 2004.
[11] H. Arup, Galvanized steel in concrete, Materials Performance 18(4) (1979) 41-44.
[12] K. Menzel, Zur Korrosion von verzinktem Stahl in kontakt mit Beton – IWB (Mittelungen), Universität Stutgart, 1992.
[13] W.G. Hime, M. Machin, Performance variances of galvanized steel in mortar and Concrete, Corrosion 49(10) (1993) 858-860.
[14] A.H. Burggrabe, Einflußfaktoren für das Verbundverhalten glatter verzinkter Bewehrungsstäbe aus Stahl im Beton, Der Bauingenieur 46
(1971) 366-369.
[15] K.E. Robinson, The bond strength of galvanized reinforcement, Technical report: TRA/220/1956, Cement and Concrete Association 52
Grosvendor Gardens London SW1, 1956, p. 7.
[16] A. Macias, C. Andrade, Corrosion of galvanized steel in dilute Ca(OH)2 solutions (pH 11,1 - 12,6), British Corrosion Journal 22(3) (1987)
162–171.
[17] A. Macias, C. Andrade, Corrosion of galvanized steel reinforcements in alkaline solutions. (Part 1: Electrochemical results), British
Corrosion Journal 22(2) (1987) 113-118.
[18] A. Macias, C. Andrade, Corrosion of galvanized steel reinforcements in alkaline solutions. (Part 2: SEM study and identification of
corrosion products), British Corrosion Journal 22(2) (1987) 119-130.
[19] M.T. Blanco, A. Macias, C. Andrade, SEM study of the corrosion products of galvanized reinforcements immersed in solutions in the pH
range 12,6-13,6, British Corrosion Journal 19(1) (1984) 41-48.
[20] A. Macias, C. Andrade, Stability of the calcium hydroxyzincate protective layer developed on galvanized reinforcements after a further
increase of the pH value, British Corrosion Journal 36 204) (1986) 19–28.
[21] V. Kučera, P. Pokorný, M. Kouřil, Kinetika dějů řídících korozi zinkované oceli v betonu (Laboratorní projekt I), VŠCHT, 2015, s. 26.
[22] P. Pokorný a kol., Zhodnocení vlivu koroze žárově zinkované oceli na soudržnost hladkých prutů s betonem třídy „NSC“, Koroze a ochrana
materiálu 59(2) (2015) 53-65.
[23] P. Rovnaníková, P. Bayer, Mikrostruktura cementového tmelu v okolí pozinkované výztuže, 9, konference žárového zinkování, 2003, pp.
57–62.
[24] P. Rovnaníková a kol., Impact of galvanized steel corrosion on cement paste microstructure, EUROCCOR 2004-Nice, 2004, p. 9.

Вам также может понравиться