Next Article in Journal
Preference Index of Sustainable Natural Fibers in Stone Matrix Asphalt Mixture Using Waste Marble
Next Article in Special Issue
New Materials and Technologies for Durability and Conservation of Building Heritage
Previous Article in Journal
Delayed Absorption Superabsorbent Polymer for Strength Development in Concrete
Previous Article in Special Issue
Influence of Polyurea Coatings on Low-Longitudinal-Reinforcement-Ratio Reinforced Concrete Beams Subjected to Bending
 
 
materials-logo
Article Menu
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Improvement of Durability of Reinforced Concretes for Sustainable Structures: A Review on Different Approaches

by
Luigi Coppola
1,*,
Silvia Beretta
2,
Maria Chiara Bignozzi
3,
Fabio Bolzoni
2,
Andrea Brenna
2,
Marina Cabrini
1,
Sebastiano Candamano
4,
Domenico Caputo
5,
Maddalena Carsana
2,
Raffaele Cioffi
6,
Denny Coffetti
1,
Francesco Colangelo
6,
Fortunato Crea
4,
Sabino De Gisi
7,
Maria Vittoria Diamanti
2,
Claudio Ferone
6,
Patrizia Frontera
8,
Matteo Maria Gastaldi
2,
Claudia Labianca
7,
Federica Lollini
2,
Sergio Lorenzi
1,
Stefania Manzi
3,
Milena Marroccoli
9,
Michele Notarnicola
7,
Marco Ormellese
2,
Tommaso Pastore
1,
MariaPia Pedeferri
2,
Andrea Petrella
7,
Elena Redaelli
2,
Giuseppina Roviello
6,
Antonio Telesca
9 and
Francesco Todaro
7
add Show full author list remove Hide full author list
1
Department of Engineering and Applied Sciences, Università di Bergamo, Viale Marconi 5, 24044 Dalmine, Italy
2
Department of Chemistry, Materials and Chemical Engineering “G. Natta”, Politecnico di Milano, Via Mancinelli 7, 20131 Milan, Italy
3
Department of Civil, Chemical, Environmental and Materials Engineering, Università di Bologna, Via Terracini 28, 40131 Bologna, Italy
4
Department of Mechanical, Energy and Management Engineering, Università della Calabria, Via Bucci-Cubo 46C, 87036 Rende, Italy
5
Department of Chemical, Materials and Industrial Engineering, Università “Federico II” di Napoli, Piazzale Tecchio 80, 80125 Naples, Italy
6
Department of Engineering, Università Parthenope di Napoli, Via Amm. Acton 38, 80133 Naples, Italy
7
Department of Civil, Environmental, Land, Building Engineering and Chemistry, Politecnico di Bari, Via Orabona 4, 70126 Bari, Italy
8
Department of Civil Engineering, Energy, Environmental and Materials, Università Mediterranea di Reggio Calabria, Via dell’Università 25, 89122 Reggio Calabria, Italy
9
School of Engineering, Università della Basilicata, Viale dell’Ateneo Lucano 10, 85100 Potenza, Italy
*
Author to whom correspondence should be addressed.
Submission received: 4 March 2022 / Revised: 4 April 2022 / Accepted: 6 April 2022 / Published: 7 April 2022

Abstract

:
The topic of sustainability of reinforced concrete structures is strictly related with their durability in aggressive environments. In particular, at equal environmental impact, the higher the durability of construction materials, the higher the sustainability. The present review deals with the possible strategies aimed at producing sustainable and durable reinforced concrete structures in different environments. It focuses on the design methodologies as well as the use of unconventional corrosion-resistant reinforcements, alternative binders to Portland cement, and innovative or traditional solutions for reinforced concrete protection and prevention against rebars corrosion such as corrosion inhibitors, coatings, self-healing techniques, and waterproofing aggregates. Analysis of the scientific literature highlights that there is no preferential way for the production of “green” concrete but that the sustainability of the building materials can only be achieved by implementing simultaneous multiple strategies aimed at reducing environmental impact and improving both durability and performances.

1. Introduction

In the field of construction materials, it is increasingly evident that traditional environmental parameters (such as global warming potential (GWP), and gross energy requirement (GER)) as well as life cycles analyses are needed—but not sufficient—to define the sustainability of a building material. Simple parameters based on concrete composition, CO2 emissions, and compressive strength such as those proposed by Damineli et al. [1] are no longer adequate for a holistic treatment of the issue. It is essential to combine information regarding the material performances and durability with the evaluation of its environmental impact. In other words, it is not possible to define a construction material as “green” without a deep investigation of its property evolution in different environments over time.
The phenomena of early degradation, primarily those promoted by carbon dioxide or chlorides, can greatly reduce the sustainability of cementitious materials, both traditional and innovative, as widely reported in the scientific literature [2,3]. Therefore, this review aims to collect the main strategies currently available for obtaining durable and sustainable reinforced concrete structures, using both traditional and innovative materials.

2. Corrosion Mechanisms in Reinforced Concrete Structures

The protective capacity of reinforced concrete against carbon steel corrosion is one of the fundamental points that have made it the most used construction material for industrial and civil structures. Steel reinforcements give tensile strength to cementitious materials, and concrete offers protective conditions to preserve the steel from corrosion, thus making production of durable structures possible. The protective action is due to the formation of hydration products of Portland cement, which increases the alkalinity of the water inside the pores of the hardened concrete. In fact, the corrosion behavior of carbon steel is strongly influenced by the pH of the pore solution, and it is assumed that it is passive when it exceeds 11.5. In these conditions, the corrosion rate of carbon steel reinforcements becomes negligible due to the formation of a protective passive film, which slows down the anodic process of metal dissolution. Portland cement is composed by calcium silicates, which, reacting with water during the hardening process, lead to the formation of calcium hydroxide. This substance is a strong, slightly soluble hydroxide, which saturates the water of the pores. At room temperature, a simple saturated solution of this substance has a pH around 12.5. However, the pH of fresh cement paste is generally higher due to the presence of small amounts of sodium and potassium hydroxides, determining the increase in the pH up to 13.5. These alkalinity levels are reached immediately during the mixing, thus promoting a rapid passivation of the reinforcement [4,5]. The free corrosion potential of rebars rapidly increases, during the setting and hardening phase, up to potentials typical of passive conditions [6,7]. Fresh concrete is a suspension of water, solid particles of different granulometry, and cement dust, where water represents an amount of only about 20%. The solution in contact with steel reinforcements is limited to the adjacent water film, while the solid/liquid ratio increases as the degree of hydration increases. The alkali content of this water thin layer, responsible for the passivity of steel, does not depend only on the content of the above-mentioned hydroxides or on the possible presence of pozzolanic material, but also on the consumption of hydroxyl ions for the formation of the passive film itself. The protectiveness tends to increase over time and it becomes stable only after several months embedded in the cement matrix, as reported also by Andrade et al. [8].
The protective action by Portland cement concrete, however, is not only due to high pH values, but it also depends on the presence of chlorides and on the ability of the cement matrix to decrease the chloride and carbonation penetration through the concrete cover.
Chlorides break the passive film and promotes localized corrosion initiation of reinforcements. This is the main form of corrosion responsible for damaging concrete exposed in the marine environment or bridge decks and civil buildings exposed to de-icing salts. Localized corrosion initiation occurs once chloride concentration (by percentage to the weight of cement) exceeds a critical concentration threshold at the steel surface. In structures exposed to the atmosphere, where the embedded steel rebars are characterized by a high corrosion potential, this critical threshold in Portland cement concrete is usually between 0.4 and 1% [9]. Higher values are found in water-saturated concrete, in which the steel corrosion potential is lower. The alkalinity and characteristics of the concrete/reinforcement interface are the main factors influencing the critical chloride concentration threshold [10,11,12,13,14,15,16,17,18,19]. It increases with the pH and it can be described in terms of chloride to hydroxyl critical molar ratio, which is commonly considered equal to 0.6. According to this ratio, the alkalinity of Portland cement concrete therefore makes possible localized corrosion initiation only when chlorides penetrate from the environment. The value can be even higher in concrete, due to the buffering effect produced by calcium hydroxide formed during hydration of cement [12,20]. The presence of this phase on the metal surface represents a “reserve of alkalinity”, which contrasts, at the metal/cement paste interface, the pH variations involved in the initiation mechanism of localized corrosion. Only a fraction of the total chlorides already present in concrete contributes to the initiation of localized corrosion. Free chlorides, dissolved in the solution contained in the pores, are active, while a significant part is bound by the constituents of the cement [21] and does not influence the corrosion phenomenon. The two main bonding mechanisms of chlorides are by adsorption, especially on hydrated calcium silicate (C-S-H) [22], and by chemical substitution, in monosulfate calcium aluminate (phase AFm) [23,24,25], with the formation of Friedel’s salt. In addition to these phases, chlorides can also adsorb on portlandite (CH), ettringite (AFt), and other salts [25,26,27,28].
The durability of reinforced concrete structures is strictly related to the two main processes governing the corrosion of steel reinforcements, such as chlorides penetration and carbonation. Both processes affect the protective ability of concrete against steel rebar corrosion. The chloride and carbon dioxide penetration rates are mainly dependent upon the porosity of the concrete matrix, the size and distribution of the pores. It is well known, in fact, that the durability of concrete mainly depends on the mix design, placing and curing. In this view, the concrete cover thickness can be considered as the key determining factor which defines the time required for aggressive substances to reach the reinforcements.
The low corrosion rate of the reinforcements is determined mainly by passivity. Oxygen is normally present and reaches the surface of the reinforcement in amounts that promote the corrosion process. Once the passivation layer is broken, however, very different corrosion conditions can occur, in relation to the water saturation of pores. Only in water-saturated concrete, the reduced supply of oxygen, due to the slow diffusion through the pores occluded by the aqueous phase, can limit the corrosion process. This can be observed in permanently immersed concrete, in which even the possible loss of passivity would not lead to any significant corrosion [29]. However, the concrete is not typically saturated with water and the access of oxygen is such as not to constitute a limiting factor, due to the rapid diffusion through the air contained in the pores, only partially filled with water. In this case, the corrosion rate is determined by the availability of water, necessary to promote the corrosion process.
In very humid, but not saturated, concretes, the corrosion process can take place with significant rates mainly in the presence of significant chloride contamination. In these concretes, the amount of water is enough to guarantee a low electrical resistivity of the cementitious matrix, thus favoring the galvanic couple action, which controls the localized corrosion mechanism.
In carbonated concrete—without chlorides—the corrosion rate is much lower and general corrosion occurs. The corrosion rate assumes relatively low values, especially in concretes exposed to low humidity levels. The amount of electrolyte is very low, and consequently the corrosion rate is also low. In addition, the corrosion products tend to reduce the small volume of electrolyte, thus promoting the formation of patinas on reinforcements, which further decrease the anodic oxidation process of the metal. A situation of pseudo-passivity arises, with relatively high corrosion potentials, but with negligible corrosion rates. The propagation period becomes the main process in the service life of carbonated structures.
In addition, the presence of cracks and defects could represent a preferential access point for corrosive agents in concrete. However, Portland cement concrete has “smart” properties that hinder this effect, making it much less important than might be expected. The interaction between concrete and the environment leads to the precipitation of substances that tend to seal the cracks, thus making them much less critical. This is what happens, for example, in the marine environment or in contact with water that contains bicarbonate ions, calcium ions and magnesium ions, in the form of dissolved salts. In contact with the alkalinity of the concrete walls, calcium and magnesium carbonates limit the ingress of water, and they can seal relatively large cracks (below 300 μm) [30]. The barrier properties of concrete are somehow restored, thus prolonging the initiation of the corrosion phenomena. The effect is significant only for small-sized cracks and depends on the characteristics of the water and the properties of the concrete [31,32]. In the presence of major defects, however, this effect cannot be considered, and corrosion rate mechanisms are that of atmospheric corrosion rather than that of corrosion of carbon steel reinforcements in concrete.

3. Corrosion Inhibitors and Surface Treatments

Additional protection methods are necessary for reinforced concrete structures operating in severe field conditions or when very long service life is required: corrosion-resistant reinforcements, cathodic prevention, corrosion inhibitors, and surface treatments represent suitable “tools” to prevent corrosion in very aggressive environments [33].
Surface treatments to apply on the surface of reinforced concrete elements are efficient protective methods at a relatively low cost. The European Standard EN 1504-2 identifies:
(a)
Hydrophobic treatments, based on silanes, siloxanes and silicones;
(b)
Treatments able to seal the capillary pores, based on sodium silicate or magnesium fluorosilicates;
(c)
Organic coatings forming a continuous film, with a thickness between 0.1–0.3 mm, thermoplastic (acrylic, vinyl) or thermosetting (epoxy, polyurethane);
(d)
Cementitious mortars containing acrylic or vinyl polymers with polymer/cement ratio in the range of 0.3–0.6 and thickness between 1 and 5 mm.
The effect of these treatments is for two reasons: they reduce the transport of aggressive agents in concrete (oxygen, carbon dioxide, and chlorides), delaying corrosion initiation; they decrease the concrete water content, reducing the corrosion rate.
Many laboratory tests have been carried out to study their effectiveness, even if they are mainly short-term tests on water absorption, vapor permeability, adhesion, and accelerated chloride corrosion [34,35,36,37]. Hydrophobic treatment and polymer modified mortars showed the best efficiency on corrosion prevention. A long-term chloride corrosion test, lasting 17 years, showed that polymer modified coatings both delay the initiation of chloride corrosion, thanks to a strong decrease in the chloride penetration, and reduce corrosion rate [38,39]. The higher the polymer/cement ratio, the higher the coating effectiveness. However, few field-tests are available to predict the durability of surface treatments beyond a period of more than 10 years under different conditions of exposure [33].
Corrosion inhibitors can be used to both prevent and stop chloride induced corrosion and as a remedial for structures exposed to carbonation. They can be divided in two groups: admixed inhibitors (mass inhibitors), directly added as a constituent to fresh concrete, and as preventive techniques; migrating inhibitors, applied on the concrete surface which can penetrate into the hardened cement matrix, usually adopted in rehabilitation [40,41,42,43,44]. Among the mass inhibitors, inorganic ones were firstly studied since the 1950s and efficient commercial products are available. Migrating commercial corrosion inhibitors were proposed in the last 30 years, due to the growing interest in the recovery and restoration of existing buildings.
Nitrite-based inhibitors [40,45], acting as anodic passivating agents, are the most effective ones, provided a chloride/nitrite molar ratio lower than 1 is maintained. In the maximum dosage (30 L/m3) they guarantee an increase in the critical chloride content up to 3% by cement mass. They also have an effect on carbonation corrosion if dosed at 3% by cement mass. In particular, nitrites were found effective in accelerating the passivation process of active galvanized steel in fresh concrete, which is a significant aspect to consider for these types of reinforcements [46]. Concerns are with its harmfulness, solubility, and possible increase in corrosion rate in the case of low dosage.
Organic commercial inhibitors (amines, alkanolamines, and carboxylates) [40,47,48] act by adsorption on the metal surface, forming an organic monolayer. Laboratory tests, both in solution and in concrete, showed a slight increase in the critical chloride content (up to 1.2–1.5% by cement mass) for inhibitor dosages ranging from 1.5 to 10 L/m3. Few data are available on long-term efficiency; in any case, they are not as efficient as nitrite. Migrating organic inhibitors, based on similar compounds, in most cases do not reduce corrosion rate after initiation, they only delay the initiation of corrosion due to a pore blocking effect [40,49].
In the last 20 years there has been a growing interest in the study of new compounds, and to understand the mechanism of inhibition: both inorganic (zinc oxide, molybdates, borates) and organic compounds (benzoate derivatives, carboxylated ions, and amine-based substances) have been tested [13,50,51].

4. Self-Healing Strategies for High Durability Concrete

Concrete is a low-tensile strength and fragile material that is very susceptible to cracking mainly due to shrinkage, tensile stress, and freezing and thawing cycles. Generally, microcracks do not significantly jeopardize the elastomechanical performance of concrete but promote an easier penetration of external matters such as water and other chemical agents (i.e., sulfates, chloride, and acids) resulting in cement matrix degradation followed by a corrosion of steel rebars [52,53,54]. In other words, the microcrack formation is generally responsible for a reduction in a service life of concrete structures without affecting their strength [55]. For this reason, the development of techniques aiming at increasing the lifespan or reducing the maintenance costs of buildings are essential, especially in a sustainable perspective of concrete structures [56,57]. In the last years, starting from the autogenous self-healing phenomena described by Hyde and Smith [58] and Glanville [59], researchers investigated several self-healing approaches able to improve the natural capability of concrete to fill cracks.
The autogenous self-healing is defined as the natural recovery process of concretes not specifically designed for self-healing [60] and it occurs due to physical, chemical, and mechanical phenomena. The physical cause is due to swelling of hydrated cement paste next to the cracks, whereas the chemical processes are related to the continued hydration of cement and the formation of calcium carbonate crystals on the crack’s faces. Minor effects are due to mechanical causes such as the presence of fine particles that partially fill the cracks. However, the effectiveness of autogenous self-healing is rather limited and affects only the small cracks with width lower than 300 µm [61,62].
When concrete is manufactured with engineered additions able to improve the self-healing capability of mixtures, it is called autonomic self-healing or activated repairing. Several techniques have been proposed in this field, as reported in Figure 1.
The use of bacteria (also called bacterial concreting) has been shown to be effective in repairing cracks in concrete, promoting both a reduction in water penetration and chloride ion permeability with small recovery in mechanical strength [63,64]. In particular, the microbially induced calcium carbonate precipitation can occur by adding bacteria in porous aggregates [65,66,67], diatomaceous earth [68], rubber particles [69], plastic microcapsules [70,71,72], or hydrogel [73]. In any case, the effectiveness of long-term self-healing capability remains to be assessed.
A technique similar to the bacterial concreting involves the use of polymeric-based repairing agents (such as epoxy resin, methyl-methacrylate, ethyl-cyanoacrylate, or polyurethane) stored in hollow glass fibers [74], ceramic tubes [75], porous plastic fibers [76], or micro-/macrocapsules [77,78]. The healing ability is related to the microcapsule damage around the cracks that releases the healing agent. Inorganic agents (sodium and potassium silicate) were also successfully investigated in [79,80]. Nevertheless, issues related to rheology, dispersion of microcapsules, and mechanical strength loss must be solved before a widespread use of these systems [60].
Crack healing capability of concrete can be also be enhanced by adding in the mixture cross-linked polymers (also called superabsorbent polymers) that have the ability to absorb huge amount of water from the environment and to retain the liquid within their structure without dissolving. When cracks occur, these materials are exposed to the external environment and the subsequent contact with water or moisture promotes the swelling of polymers and the formation of a soft gel that prevents the ingress of external agents into concrete [81,82]. The detailed healing mechanism of superabsorbent polymers has been reported by Lee et al. [83].
The most promising technique for autonomic self-healing is the addition of expansive agents, mineral additives (also called supplementary cementitious materials), admixtures and fibers as well as their combination during the mixing [84,85,86]. Several studies evidenced that the addition of expansive agents (i.e., MgO, CaO, bentonite) and fibers both limits the shrinkage of concrete and produces compatible expansive hydrated. In this way, crack bridging capacity (strength recovery), sealing (physical closer of cracks through crystallization), and durability are improved [87,88]. On the other hand, the use of mineral additives and carboxylic acid derivatives promotes both the cement recrystallization and the salt precipitation inside cracks with an initial width up to 500–800 µm without affecting the properties of concrete at fresh and hardened state [89,90]. More details can be found in [30].

5. Corrosion Resistant Reinforcements

When the concrete cover is not able to provide the proper protection against corrosion of the traditional carbon steel reinforcement, e.g., in highly aggressive environmental conditions (especially in the presence of chlorides) or when a long service life is required, it is possible to use additional prevention/protection systems in order to guarantee the required durability [91]. The use of corrosion resistant reinforcements is one of the main additional prevention/protection systems and can be a sound choice for new structures or in repair of existing ones. The corrosion resistance of reinforcements can be obtained with coatings, both metallic (galvanized steel) or organic (epoxy coated bars), modifying the chemical composition of the steel (mainly using stainless steels) or using composite materials (FRP, Fiber Reinforced Polymers) [33,92].
The corrosion resistant reinforcement should fulfil the requirements settled for the traditional carbon steel bars, such as strength, ductility, weldability, and bond to concrete. These rebars are characterized by different corrosion behavior and costs. Their related benefits can be evaluated with performance-based approaches for the design of durability [93]. As far as the costs are concerned, although their higher initial costs, their use can lead to significant costs savings during the service life of the structure, due to a reduction in maintenance costs (direct and indirect). Moreover, a selective use in the most critical parts can be considered, thus a reduction in the initial cost can be achieved.
In carbon steel coated rebars the coating thickness and its quality (integrity) are crucial to guarantee the effectiveness of the protection [94,95]. In galvanized reinforcements, where a protective zinc-based coating is present, having typically a more or less homogeneous pure zinc η-phase on the top, the passive film, produced on the rebar surfaces, can be effective in concrete structures subjected to carbonation induced corrosion or to penetration of chlorides [96,97]. In carbonated concrete the corrosion rate is about 1–2 µm/year, thus the corrosion propagation is very slow [98]. A chloride threshold for pitting corrosion initiation in the range of 1–1.7% by weight of cement has been found, reaching also higher values than these ones, when the coating is constituted by different zinc alloys, which can be obtained from different baths in the process of hot-dip galvanizing of carbon steel reinforcements [99]. Therefore, the chlorides threshold to initiate the pitting corrosion of galvanized steel reinforcements is significantly higher than that generally considered for carbon steel rebars (0.4–1%); hence, advantages can be obtained in terms of service life extension. Furthermore, in the presence of coating discontinuities, owing to bending of rebars or welding operations, which leave uncoated substrate spots, the zinc-based coating determines a cathodic protection of the steel in correspondence of these spots [100].
In epoxy coated rebars, the epoxy resin can provide a barrier protection. This kind of resin is suitable for use in concrete (good resistance to alkaline solution, good mechanical properties, good adhesion to steel and concrete). In the presence of defects, when concrete is carbonated or in the presence of chlorides with a content higher than the chloride threshold, corrosion can occur. In carbonated concrete the attacks are, generally, limited to the area of the defects, thus also the consequences are limited [101]. In chloride contaminated concrete no advantage in pitting corrosion initiation can be achieved in the presence of defects [102]. Moreover, with these bars, the use of electrochemical techniques to assess the corrosion behavior of the reinforcements is not possible due to the presence of the electrical insulating coating. After cutting or welding, or in the presence of defects in the coating, the areas without protection have to be repaired with a paint.
In stainless steel reinforcement the corrosion resistance is given by their chemical composition (alloy elements: mainly chromium, molybdenum, nickel, nitrogen). These rebars, if properly selected, can guarantee also long service lives in harsh environmental conditions without maintenance thanks to their high resistance to corrosion. These steels do not suffer corrosion in carbonated concrete and can resist to chloride induced corrosion also in the presence of very high chloride content (also higher than 5% by cement weight) [103,104,105,106]. For this reason, their use is generally considered in chloride-rich environments with high aggressiveness. To select the most suitable type of stainless steel in terms of corrosion resistance and cost, among the different types available, the chloride threshold for pitting corrosion initiation must be known. In order to limit the costs, stainless steel reinforcement is often used in the most critical parts of the structure (or in a repaired area) and connected with the carbon steel rebars. This coupling does not lead to risk of galvanic corrosion [107].
The use of FRP reinforcement, generally GFRP (Glass Fiber Reinforced Polymers), is still to be considered in the experimental phase. Long-term data on their behavior under different exposure conditions are not available. Despite the fact they do not suffer electrochemical corrosion as steel does, they are subjected to other deterioration phenomena, e.g., due to concrete alkalinity, temperature, and humidity [108].

6. Durability Design

Worldwide, corrosion of embedded steel is the main form of premature damage of reinforced concrete structures, hence there is the need to prevent it since the design stage [33,109,110,111,112,113,114]. At this aim several approaches are available that are characterized by different levels of approximation. As introduced for the structural design in the “Model Code for Concrete Structures” issued by the International Federation for Structural Concrete (fib) in 2010, a level of approximation is a design strategy where the accuracy of the prevision can be, if necessary, progressively refined through a better estimation of the parameters related to the considered phenomenon [115]. A low level of approximation should be reserved for structures where high accuracy is not required or for a pre-design; conversely higher levels of approximation can be used in cases where higher accuracy is required and it is expected that the solution is closer to the actual behavior.
Dealing with durability, a low level of approximation can correspond to the prescriptive approach, that needs the fulfillments of minimum requirements, whilst through a performance-based approach, which consists of a real durability design, the accuracy of the prevision can be increased.
The prescriptive approach is based on the definition of an exposure class, that describes the aggressiveness of the environment to which concrete will be exposed during its service life, and the subsequent prescriptions regarding the maximum water/cement (w/c) ratio and the minimum cement content, according to the EN 206 [116]. These should be associated with minimum values of the concrete cover thickness (related to protection of rebars from corrosion), according to the Eurocode 2 [117]. These simple recommendations apply to any type of cement of the EN 197-1 standard [118] and refer to an intended service life of about 50 years. The prescribed values revealed to be inadequate in some parts of the structures, as those highly exposed to chlorides, e.g., the joints or the splash zone in marine structures [113]. Moreover, it is implicitly assumed that durability performances of concretes made with different types of cement are comparable, whilst it is well known that they behave even significantly different in relation to the resistance to aggressive agents [119,120,121,122,123,124,125]. Finally, this kind of approach does not allow to take into account the advantages of additional protections.
The performance-based approach allows to specifically design each structural element in a way that it can withstand the actual local conditions of exposure during the required service life. Among the models proposed in the recent years, the fib ‘‘Model Code for Service Life Design’’, published in 2006 [126], is one of the most used. This includes a probabilistic performance-based approach that, modelling the environmental effects on the structure, allows the evaluation of the probability that a pre-defined limit state, which corresponds to an undesired event (e.g., initiation of corrosion, cracking or spalling of concrete cover), occurs. Through these models, different design combinations, together with their reliability, can be compared, as well as the benefits connected with the use of preventative techniques [93,127]. As an example, Figure 2 shows the durability design, carried out through the fib Model Code, of a RC element exposed to the splash zone, considering a service life of 100 years and different design options, in term of types of concrete and reinforcement and concrete cover thickness. Their widespread use, however, is still limited, since indications on same input design parameters are lacking and their estimation is entrusted to the experience of the designer. Moreover, since these models are quite young compared to the length of usual service lives of RC structures and feedback data are not available yet, the reliability of their output is still under investigation [128,129].

7. Waterproofing Recycled Aggregates

Cement composites can be considered unsaturated porous materials that, when in direct contact with water, are permeated through various transport mechanisms (capillary rise, permeation, diffusion). These processes dramatically affect the durability of the concrete structures since they: (a) expose them to freeze and thaw deterioration, (b) alter cement paste composition/microstructure by dissolution and removal of its ionic compounds, (c) promote the ingress of aggressive ionic agents such as sulfates and/or chlorides.
An important characteristic of a porous material is the capillary water absorption expressed with the absorption coefficient S (kg/(m2·s0.5):
S = ( σ   c o s ϑ   r 2 μ )  
where r is the mean radius of the capillary pore, σ the surface tension of the liquid, ϑ the water contact angle. Thus S is higher as the pore size increases and as the contact angle decreases, so when dealing with porous and hydrophilic materials [130]. The cementitious matrix is made of hydrated products (mainly composed by Ca, Si, Fe) and the aggregates (65–75% of the total volume) are generally natural siliceous or limestone sand and gravel. All of these constituents contribute to the pronounced hydrophilic character of the whole cement composite which is characterized also by a peculiar porosity [131]. On the contrary, polymeric materials are intrinsically hydrophobic as they are rich in low energy groups, e.g., the -CHx ones [132,133]. For these reasons, hydrophobic cementitious materials can be easily obtained by using polymeric aggregates as a partial substitution for natural stones [134].
Replacing natural sand in cementitious mortars with grains of end-of-life tyre rubber, containing isoprene/butadiene chains, strongly reduces the penetration of water drops both in sound and cracked materials. In fact, the hydrophobic character of the aggregates is dispersed in the whole mass of the composite and exerts its effect both on the surface and in the bulk [135]. Fast water absorption, instead, has been detected when inorganic (siliceous) recycled aggregates, such as porous waste glass, have been tested [136].
However, the different surface energy of polymeric aggregates with respect to traditional concrete constituents determines a reduction in aggregate-cement bond, thus increasing both the porosity of the material and the parameter r reported in Equation (1). Nevertheless, the effectiveness of these aggregates in hindering the water ingress has been proved in terms of both water absorption rate of microliter water drops [135] and water capillary rise in partially immersed samples [137]. The first method allows a highly space resolved wetting analysis, the latter allows a quick and overall evaluation of the absorption coefficient S (Equation (1)). From Figure 3 it is evident that the absorption coefficient (calculated from the slopes of the linear fit) of rubberized mortars is less than a half of those made with natural sand [136]. Furthermore, when exposed to accelerated chloride penetration, a lower corrosion degree of steel reinforcement has been measured in rubberized concrete with respect to ordinary concrete, indirectly confirming the high water resistance of mixtures containing end-of-life tire aggregates [138].

8. Durability of Special Mixtures

8.1. Fly Ash-Based Geopolymers

A new class of materials known as geopolymers, which are part of the broad class of inorganic matrices named alkali-activated materials (AAM) has rapidly grown in interest in the last two decades in order to reduce the CO2 emissions for cement and ceramic materials productions. This new class of materials is based on alkali activation of low calcium aluminosilicate precursors able to consolidate at room or slightly higher temperatures. One of the main advantages of AAM and geopolymers is the possibility to use waste-based powders as for example coal fly ashes derived from coal fired power stations, thus promoting a circular economy approach. Many aspects of geopolymers have been studied, from the synthesis and optimization of aluminosilicate precursors to the properties of the developed products (physical, mechanical, and microstructural performances) [139,140,141,142,143,144,145,146].
As far as geopolymers durability is concerned, interesting results have been obtained about their resistance to sulfate attack [147] and alkali–silica reactions and about the high stability in the presence of fire or freeze–thaw cycles, besides a high adhesion to steel reinforcement [148,149,150,151,152,153], which suggests their use as binder in mortar and/or concrete, or for strengthening applications of reinforced concrete structures [144,154,155,156,157]. If properly designed, geopolymers perform better than ordinary Portland cement when exposed to high temperature. The rapid dehydration of the weakly bound water in the gel does not cause significant damage to the binding structure, therefore mechanical strength is retained and considerable dimensional stability at high temperature is verified [158,159,160,161,162,163]. Recycled refractory particles (RRP) have been used to develop AAM and geopolymers with enhanced performance at high temperatures [164]. RRP do not hinder the alkali activation process, and they reduce heat-induced cracking, increase the maximum temperature of dimensional stability of the composites up to 1240 °C, and improve the linear dimensional stability during heating. In addition, the room temperature curing generates a product less prone to cracking than heat curing when exposed to high temperature [38]. Considering the good performance of AAM as fire resistant materials [165,166,167], research has also focused on the use of lightweight AAM for steel protection (passive fire protection systems) [168,169,170]. Heat-induced cracking in fly ash-based alkali-activated pastes and lightweight mortars was analyzed by in situ acoustic emission detection during complete heating–cooling cycles [171]. Cracking during heating was limited and associated exclusively with the dehydration of the materials. However, samples heated to temperatures above 600 °C exhibited intense cracking on cooling.
The addition of sodium silicate to sodium hydroxide stimulated network formation in geopolymers leading to improved mechanical strength, lowering chloride ion mobility and slightly improving corrosion performances [172,173]. Studies on the corrosion behavior of steel in different room temperature cured alkali-activated fly ash mortars exposed to chloride solution showed that the most compact alkali-activated mortars have higher porosity and lower mechanical properties than a cement-based mortar, but the protectiveness afforded to the rebars is slightly higher than that obtained in traditional mortars [174].

8.2. Alkali Activated Materials

Alkali-activated materials are a recent family of binders. In the last decades, they received growing attention from academic research institutions as promising candidates in specific civil applications, such as refractory structures and concrete sewers [165,175,176,177]. Nevertheless, related marketplace shows a certain amount of concern on AAM more extensive use due to the unclear performances in terms of properties and durability, lack of extensive track record, product standards, and tailored polymer admixtures. The significant numbers of raw materials and alkaline activators, that can be used to formulate their mix design, deeply affect fundamental properties, such as shrinkage and cracking behavior, workability, development of mechanical strength, and risk of efflorescence [147,178,179,180,181]. The adopted formulations also determine their resistance to chemical, physical, and transports attack modes. Moreover, the level of confidence is also lowered by the evidence that some durability tests, tuned on Portland based products, fail in predictive ability when applied to AAM [181,182,183].
Results published by several researchers, even in the presence of a multitude of evaluation criteria and methods (length and mass changes, residual mechanical properties, ultrasonic pulse velocity, and elastic modulus) and testing conditions (salt concentration, temperature, time of immersion, periodical solution replacing, and former curing regime) clearly corroborate that AAM exhibit a higher resistance to sulfate attack than that of Portland cement (PC) materials [184,185,186,187,188,189]. The rate, severity, and mechanism of external sulfate attack on AAM concrete, that takes place in soil or marine environments, depend on the permeability of the concrete/mortar, the concentration of sulfates in the waterborne solution, on the cation accompanying the sulfate ions, on nature of the selected reactive powders and activator composition. When aluminosilicates are used as reactive powders (fly ash or metakaolin), geopolymerization leads to a N-A-S-H gel that differs from PC hydration products, characterized by the absence of high-calcium phases. This condition prevents the formation, under the sulfate attack, of gypsum or ettringite and results in a good resistance to sulfate attack [188], that is further increased if Na(OH) instead of Na2SiO3 is used as activator [185]. When Blast Furnace Slag (BFS) or other high calcium containing reactive powders are used, the formation of a hydrotalcite-type phase and of a hydration product, a calcium aluminosilicate hydrate (C-A-S-H) commonly occurs. It is less crystalline and with a lower CaO/SiO2 molar ratio than calcium silicate hydrate (C-S-H) produced by PC hydration. When exposed to sulfate attack, its durability strictly depends on cation accompanying the sulfate ions. Bakharev et al. [184], using ASTM C 1012, investigated the behavior of AAM cured for 28 days in a fog room and then immersed in a 50 g/L Na2SiO3 or 50 g/L MgSO4 solution for 12 months. No sign of deterioration was visible after sodium sulfate attack, whereas matrix degradation with formation of gypsum was observed under magnesium sulfate attack. Similar trend was found by Ye et al. [186].
When Na2SO4, KOH, and NaOH were used as activators traces of ettringite, a little expansion of the samples was visible and traces of ettringite as well as a reduction in the intensity peak of C-A-S-H gel were detected by XRD diffractometry. This behavior was related to limited gel decalcification and dealkalization, whereas, when Na2CO3 was used, the presence of CO32− suppressed ettringite formation. Under magnesium sulfate attack a more sever degradation process was observed with mechanisms that depend on the activators’ type. When NaOH or sodium silicate were used, firstly brucite was formed. It acted as a protective surface layer by delaying external ion migration, but lowered the pH of the pores solution to 10.5 thus promoting C-A-S-H gel decalcification, but not its dealumination, and gypsum production. The further migration of Mg2+ and SO42− and their reaction with decalcified C-A-S-H produced a magnesium-aluminosilicate-hydrate (M-A-S-H) and/or silica gels. The absence of ettringite was ascribed to the unfavorable pH and limited amount leached aluminum. The AAM obtained using Na2SO4 showed the weakest resistance against MgSO4 attack, due to lack of brucite protective layer due to the absence of hydroxide ions.
Ismail et al. [189] also investigated the resistance to sulfate attack of blended AAM, obtained using BFS and fly ash. Again, a key role is provided by the nature of the cation accompanying the sulfate, with negligible degradation after immersion in Na2SO4 solution and a more severe damage in MgSO4 solution, with matrix degradation and loss of samples integrity.

8.3. Calcium Sulfoaluminate Cements

Calcium sulfoaluminate (CSA) cements are special hydraulic binders generally produced from limestone, bauxite, and gypsum and they represent an important alternative to Portland cement (PC). Compared to PC, CSA cements exhibit more pronounced environmentally friendly features mainly thanks to lower synthesis temperatures (~1350 °C) and reduced limestone content (~40%) both determining a strong decrease of kiln thermal input and CO2 emissions. Furthermore, CSA clinker, which can be produced also by using industrial wastes often difficult to reuse, is more friable than PC clinker and is blended with relatively high amounts of calcium sulfates to produce CSA cements [190,191,192,193].
CSA cements contain C4A3$ (ye’elimite, the main component), calcium sulfates and a variety of calcium aluminates and silicoaluminates. The most important properties of these binders are regulated by ettringite (C6A$3H32), generated upon hydration of C4A3$ together with calcium sulfates. Depending on the conditions of C6A$3H32 formation, several technical properties can be attained (e.g., rapid-hardening, good dimensional stability, low permeability and solution alkalinity, or shrinkage compensation/self-stressing behavior) [194,195,196,197,198].
Compared to PC, there are relatively few durability studies on CSA-based cements. These mixtures have proved to be highly resistant to freeze–thaw and chemical attacks promoted by sulfates, chlorides, magnesium, and ammonium salts [193,199,200,201,202,203]. These features are mainly related to the lower porosity developed by CSA cements if compared with PC binders. In fact, porosity measurements on hydrated CSA cements, carried out with mercury intrusion porosimetry, have revealed the presence of pores with threshold radius below 25 nm [204] and only a minor content of larger pores forming an interconnected pore network [193,204], leading to low permeability [201].
As far as the carbonation is concerned, it has been found that it is due to ettringite decomposition into calcite, gypsum, and aluminum hydroxide. By now, the results obtained from carbonation tests are contradictory. In fact, a few studies state that CSA cements carbonate faster than PC [205,206,207]; other papers report that both CSA cements and PC display the same carbonation rate [208] and finally, according to other findings, CSA cements perform better than PC [209].
The behavior of CSA-based binders in relation to the durability of reinforced concrete structures is still not clear. Nevertheless, recent studies have shown that steel bars in CSA reinforced concretes, when put in chloride-free environments, seem to be protected from corrosion, despite that their alkalinity (pH = 11.5–12.0) is lower than that of PC (pH = 12.5–13.5) but sufficient to promote the passivation of embedded steel [210]. Moreover, corrosion tests on carbonated CSA concrete showed negligible corrosion rate of steel in environments up to 95% relative humidity at 20 °C temperature. Furthermore, the low alkalinity is surely favorable towards the alkali aggregate reaction [200,202,203].

9. Conclusions

This paper highlights the possible strategies for obtaining sustainable and durable concretes. In particular, it is shown how it is possible to realize durable reinforced concrete structures in different aggressive environments through an appropriate design that starts from a proper concrete composition (binders type and dosage, water content, aggregates, admixtures), passes through the choice of reinforcements (traditional carbon steel, galvanized steel, stainless steel, composite materials or coated reinforcements), and ends with the selection of additional solutions such as inhibitors, coating, self-healing techniques or waterproofing aggregates.

Author Contributions

Conceptualization and supervision L.C.; Paragraph 2: M.C. (Marina Cabrini), S.L. and T.P.; Paragraph 3: S.B., F.B., A.B., M.V.D., M.O. and M.P.; Paragraph 4: D.C. (Denny Coffetti); Paragraphs 5 and 6: M.C. (Maddalena Carsana), F.L., M.M.G. and E.R.; Paragraph 7: S.D.G., C.L., M.N., A.P. and F.T.; Paragraph 8.1: M.C.B. and S.M.; Paragraph 8.2: S.C., F.C. (Fortunato Crea) and P.F.; Paragraph 8.3: M.M. and A.T.; writing—review and editing: D.C. (Domenico Caputo), D.C. (Denny Coffetti), R.C., F.C. (Francesco Colangelo), C.F. and G.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors want to thank the members of AIMAT-CINCOMINET (Cementitious and Innovative Construction Materials Interdisciplinary Network) group for the critical revision of the paper. The AIMAT-CINCOMINET group collecting 20 Italian Universities involved in research, development, and dissemination in the field of traditional and innovative construction materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Damineli, B.L.; Kemeid, F.M.; Aguiar, P.S.; John, V.M. Measuring the eco-efficiency of cement use. Cem. Concr. Compos. 2010, 32, 555–562. [Google Scholar] [CrossRef]
  2. Coppola, L.; Bellezze, T.; Belli, A.; Bignozzi, M.C.; Bolzoni, F.; Brenna, A.; Cabrini, M.; Candamano, S.; Cappai, M.; Caputo, D.; et al. Binders alternative to Portland cement and waste management for sustainable construction—Part 1. J. Appl. Biomater. Funct. Mater. 2018, 16, 186–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Coppola, L.; Bellezze, T.; Belli, A.; Bignozzi, M.C.; Bolzoni, F.; Brenna, A.; Cabrini, M.; Candamano, S.; Cappai, M.; Caputo, D.; et al. Binders alternative to Portland cement and waste management for sustainable construction–Part 2. J. Appl. Biomater. Funct. Mater. 2018, 16, 207–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Sagüés, A.A.; Moreno, E.I.; Andrade, C. Evolution of pH during in-situ leaching in small concrete cavities. Cem. Concr. Res. 1997, 27, 1747–1759. [Google Scholar] [CrossRef]
  5. Räsänen, V.; Penttala, V. The pH measurement of concrete and smoothing mortar using a concrete powder suspension. Cem. Concr. Res. 2004, 34, 813–820. [Google Scholar] [CrossRef]
  6. Cabrini, M.; Lorenzi, S.; Pastore, T. Studio elettrochimico della formazione del film di passività sulle armature del calcestruzzo. In Giornate Naz; Della Corros.: Naples, Italy, 2013. [Google Scholar]
  7. Poursaee, A.; Hansson, C.M. Reinforcing steel passivation in mortar and pore solution. Cem. Concr. Res. 2007, 37, 1127–1133. [Google Scholar] [CrossRef]
  8. Andrade, C.; Keddam, M.; Nóvoa, X.R.; Pérez, M.C.; Rangel, C.M.; Takenouti, H. Electrochemical behaviour of steel rebars in concrete: Influence of environmental factors and cement chemistry. Electrochim. Acta 2001, 46, 3905–3912. [Google Scholar] [CrossRef]
  9. Alonso, M.C.; Sanchez, M. Analysis of the variability of chloride threshold values in the literature. Mater. Corros. 2009, 60, 631–637. [Google Scholar] [CrossRef]
  10. Angst, U.; Elsener, B.; Larsen, C.; Vennesland, O. Critical chloride content in reinforced concrete—A review. Cem. Concr. Res. 2009, 39, 1122–1138. [Google Scholar] [CrossRef]
  11. Angst, U.; Vennesland, Ø. Critical chloride content in reinforced concrete. In Concrete Repair, Rehabilitation and Retrofitting II2; Taylor & Francis Group: London, UK, 2009; pp. 311–317. [Google Scholar]
  12. PAGE, C.L. Mechanism of corrosion protection in reinforced concrete marine structures. Nature 1975, 258, 514–515. [Google Scholar] [CrossRef]
  13. Cabrini, M.; Lorenzi, S.; Pastore, T. Cyclic voltammetry evaluation of inhibitors for localised corrosion in alkaline solutions. Electrochim. Acta 2014, 124, 156–164. [Google Scholar] [CrossRef]
  14. Cabrini, M.; Lorenzi, S.; Pastore, T. Study of localized corrosion of steel reinforcement in inhibited alkaline solutions [Studio della corrosione localizzata degli acciai per armature in soluzioni alcaline inibite]. Metall. Ital. 2013, 105, 21–31. [Google Scholar]
  15. Hausmann, D.A. Steel Corrosion in Concrete—How Does It Occur? Mater. Prot. 1967, 6, 19–23. [Google Scholar]
  16. Gouda, V.K. Corrosion and corrosion inhibition of reinforcing steel I. immersed in alkaline solutions. Br. Corros. J. 1970, 5, 198–203. [Google Scholar] [CrossRef]
  17. Goni, S.; Andrade, C. Synthetic concrete pore solution chemistry and rebar corrosion rate in the presence of chlorides. Cem. Concr. Res. 1990, 20, 525–539. [Google Scholar] [CrossRef]
  18. Diamond, S. Chloride concentrations in concrete pore solutions resulting from calcium and sodium chloride admixtures. Cem. Concr. Aggreg. 1988, 44, 489–499. [Google Scholar]
  19. Yonezawa, T.; Ashworth, V.; Procter, R.P.M. Pore Solution Composition and Chloride Effects on the Corrosion of Steel in Concrete. Corrosion 1988, 44, 489–499. [Google Scholar] [CrossRef]
  20. Glass, G.K.; Buenfeld, N.R. Theoretical basis for designing reinforced concrete cathodic protection systems. Br. Corros. J. 1997, 32, 179–184. [Google Scholar] [CrossRef]
  21. Arya, C.; Buenfeld, N.R.; Newman, J.B. Factors influencing chloride-binding in concrete. Cem. Concr. Res. 1990, 20, 291–300. [Google Scholar] [CrossRef]
  22. Luping, T.; Nilsson, L.-O. Chloride binding capacity and binding isotherms of OPC pastes and mortars. Cem. Concr. Res. 1993, 23, 247–253. [Google Scholar] [CrossRef]
  23. Suryavanshi, A.K.; Scantlebury, J.D.; Lyon, S.B. Mechanism of Friedel’s salt formation in cements rich in tri-calcium aluminate. Cem. Concr. Res. 1996, 26, 717–727. [Google Scholar] [CrossRef]
  24. Matschei, T.; Lothenbach, B.; Glasser, F.P. The AFm phase in Portland cement. Cem. Concr. Res. 2007, 37, 118–130. [Google Scholar] [CrossRef]
  25. Ekolu, S.O.; Thomas, M.D.A.; Hooton, R.D. Pessimum effect of externally applied chlorides on expansion due to delayed ettringite formation: Proposed mechanism. Cem. Concr. Res. 2006, 36, 688–696. [Google Scholar] [CrossRef]
  26. Birnin-Yauri, U.A.; Glasser, F.P. Friedel’s salt, Ca2Al(OH)6(Cl,OH)·2H2O: Its solid solutions and their role in chloride binding. Cem. Concr. Res. 1998, 28, 1713–1723. [Google Scholar] [CrossRef]
  27. Elakneswaran, Y.; Nawa, T.; Kurumisawa, K. Electrokinetic potential of hydrated cement in relation to adsorption of chlorides. Cem. Concr. Res. 2009, 39, 340–344. [Google Scholar] [CrossRef]
  28. Hirao, H.; Yamada, K.; Takahashi, H.; Zibara, H. Chloride binding of cement estimated by binding isotherms of hydrates. J. Adv. Concr. Technol. 2005, 3, 77–84. [Google Scholar] [CrossRef] [Green Version]
  29. Tittarelli, F. Oxygen diffusion through hydrophobic cement-based materials. Cem. Concr. Res. 2009, 39, 924–928. [Google Scholar] [CrossRef]
  30. De Belie, N.; Gruyaert, E.; Al-Tabbaa, A.; Antonaci, P.; Baera, C.; Bajare, D.; Darquennes, A.; Davies, R.; Ferrara, L.; Jefferson, T.; et al. A Review of Self-Healing Concrete for Damage Management of Structures. Adv. Mater. Interfaces 2018, 5, 1800074. [Google Scholar] [CrossRef]
  31. Berra, M.; Colombo, A.; Pastore, T. Normal and total-lightweight reinforced concretes under cyclic immersion in 3.5% chloride solution. In Concrete under Severe Conditions: Environment and Loading; E&FN Spon: London, UK, 1995; pp. 595–604. [Google Scholar]
  32. Berra, M.; Melchiorri, G.; Pastore, T. Durability tests on normal and total-lightweight reinforced concretes in marine environment. In Proceedings of the Fourth International Symposium on Utilization of High Strength/High Performance Concrete, Paris, France, 29–31 May 1996. [Google Scholar]
  33. Bertolini, L.; Elsener, B.; Pedeferri, P.; Redaelli, E.; Polder, R.B. Corrosion of Steel in Concrete: Prevention, Diagnosis, Repair; Wiley WCH: Hoboken, NJ, USA, 2013. [Google Scholar]
  34. Basheer, P.A.M.; Basheer, L.; Cleland, D.J.; Long, A.E. Surface treatments for concrete: Assessment methods and reported performance. Constr. Build. Mater. 1997, 11, 413–429. [Google Scholar] [CrossRef]
  35. Swamy, R.N.; Tanikawa, S. An external surface coating to protect concrete and steel from aggressive environments. Mater. Struct. 1993, 26, 465–478. [Google Scholar] [CrossRef]
  36. Tittarelli, F.; Moriconi, G. The effect of silane-based hydrophobic admixture on corrosion of reinforcing steel in concrete. Cem. Concr. Res. 2008, 38, 1354–1357. [Google Scholar] [CrossRef]
  37. Coffetti, D.; Crotti, E.; Gazzaniga, G.; Gottardo, R.; Pastore, T.; Coppola, L. Protection of Concrete Structures: Performance Analysis of Different Commercial Products and Systems. Materials 2021, 14, 3719. [Google Scholar] [CrossRef] [PubMed]
  38. Diamanti, M.V.; Brenna, A.; Bolzoni, F.; Berra, M.; Pastore, T.; Ormellese, M. Effect of polymer modified cementitious coatings on water and chloride permeability in concrete. Constr. Build. Mater. 2013, 49, 720–728. [Google Scholar] [CrossRef]
  39. Brenna, A.; Beretta, S.; Berra, M.; Diamanti, M.V.; Ormellese, M.; Pastore, T.; Pedeferri, M.P.; Bolzoni, F. Effect of polymer modified cementitious coatings on chloride-induced corrosion of steel in concrete. Struct. Concr. 2019, 21, suco.201900255. [Google Scholar] [CrossRef]
  40. Elsener, B. Corrosion Inhibitors for Steel in Concrete: State of the Art Report. Constr. Build. Mater. 2001, 22, 609–622. [Google Scholar]
  41. Pastore, T.; Cabrini, M.; Coppola, L.; Lorenzi, S.; Marcassoli, P.; Buoso, A. Evaluation of the corrosion inhibition of salts of organic acids in alkaline solutions and chloride contaminated concrete. Mater. Corros. 2011, 62, 187–195. [Google Scholar] [CrossRef]
  42. Bolzoni, F.; Coppola, L.; Goidanich, S.; Lazzari, L.; Ormellese, M.; Pedeferri, M.P. Corrosion inhibitors in reinforced concrete structures Part 1: Preventative technique. Corros. Eng. Sci. Technol. 2004, 39, 219–228. [Google Scholar] [CrossRef]
  43. Coppola, L.; Coffetti, D.; Crotti, E.; Gazzaniga, G.; Pastore, T. Chloride Diffusion in Concrete Protected with a Silane-Based Corrosion Inhibitor. Materials 2020, 13, 2001. [Google Scholar] [CrossRef]
  44. Cabrini, M.; Lorenzi, S.; Coppola, L.; Coffetti, D.; Pastore, T. Inhibition effect of tartrate ions on the localized corrosion of steel in pore solution at different chloride concentration. Buildings 2020, 10, 105. [Google Scholar] [CrossRef]
  45. Berke, N.S.; Hicks, M.C. Predicting long-term durability of steel reinforced concrete with calcium nitrite corrosion inhibitor. Cem. Concr. Compos. 2004, 26, 191–198. [Google Scholar] [CrossRef]
  46. Bellezze, T.; Timofeeva, D.; Giuliani, G.; Roventi, G. Effect of soluble inhibitors on the corrosion behaviour of galvanized steel in fresh concrete. Cem. Concr. Res. 2018, 107, 1–10. [Google Scholar] [CrossRef]
  47. Maeder, U. A New Class of Corrosion Inhibitors for Reinforced Concrete. Spec. Publ. 1996, 163, 215–232. [Google Scholar] [CrossRef]
  48. Cabrini, M.; Fontana, F.; Lorenzi, S.; Pastore, T.; Pellegrini, S. Effect of Organic Inhibitors on Chloride Corrosion of Steel Rebars in Alkaline Pore Solution. J. Chem. 2015, 2015, 521507. [Google Scholar] [CrossRef] [Green Version]
  49. Ormellese, M.; Bolzoni, F.; Goidanich, S.; Pedeferri, M.; Brenna, A. Corrosion inhibitors in reinforced concrete structures Part 3–migration of inhibitors into concrete. Corros. Eng. Sci. Technol. 2011, 46, 334–339. [Google Scholar] [CrossRef]
  50. Bolzoni, F.; Brenna, A.; Fumagalli, G.; Goidanich, S.; Lazzari, L.; Ormellese, M.; Pedeferri, M. Experiences on corrosion inhibitors for reinforced concrete. Int. J. Corros. Scale Inhib. 2014, 3, 254–278. [Google Scholar] [CrossRef]
  51. Monticelli, C.; Frignani, A.; Balbo, A.; Zucchi, F. Influence of two specific inhibitors on steel corrosion in a synthetic solution simulating a carbonated concrete with chlorides. Mater. Corros. 2011, 62, 178–186. [Google Scholar] [CrossRef]
  52. Djerbi, A.; Bonnet, S.; Khelidj, A.; Baroghel-bouny, V. Influence of traversing crack on chloride diffusion into concrete. Cem. Concr. Res. 2008, 38, 877–883. [Google Scholar] [CrossRef] [Green Version]
  53. Marsavina, L.; Audenaert, K.; de Schutter, G.; Faur, N.; Marsavina, D. Experimental and numerical determination of the chloride penetration in cracked concrete. Constr. Build. Mater. 2009, 23, 264–274. [Google Scholar] [CrossRef]
  54. Kwon, S.J.; Na, U.J.; Park, S.S.; Jung, S.H. Service life prediction of concrete wharves with early-aged crack: Probabilistic approach for chloride diffusion. Struct. Saf. 2009, 31, 75–83. [Google Scholar] [CrossRef]
  55. Zhang, W.; Zheng, Q.; Ashour, A.; Han, B. Self-healing cement concrete composites for resilient infrastructures: A review. Compos. Part B Eng. 2020, 189, 107892. [Google Scholar] [CrossRef]
  56. Gartner, E.; Hirao, H. A review of alternative approaches to the reduction of CO2 emissions associated with the manufacture of the binder phase in concrete. Cem. Concr. Res. 2015, 78, 126–142. [Google Scholar] [CrossRef] [Green Version]
  57. Coffetti, D.; Crotti, E.; Gazzaniga, G.; Carrara, M.; Pastore, T.; Coppola, L. Pathways towards sustainable concrete. Cem. Concr. Res. 2022, 154, 106718. [Google Scholar] [CrossRef]
  58. Hyde, G.W.; Smith, W.J. Results of experiments made to determine the permeability of cements and cement mortars. Mater. Sci. 1889, 128, 199–207. [Google Scholar] [CrossRef]
  59. Glanville, W. The Permeability of Portland Cement Concrete; Harrison and Sons: Ventura, CA, USA, 1931; Available online: http://www.worldcat.org/title/permeability-of-portland-cement-concrete/oclc/29871375 (accessed on 9 March 2018).
  60. Rooij, M.; van Tittelboom, K.; Belie, N.; Schlangen, E. Self-Healing Phenomena in Cement-Based Materials: State-of-the-Art Report of RILEM Technical Committee; Springer: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  61. Edvardsen, C. Water permeability and autogenous healing of cracks in concrete. In Innovation in Concrete Structures: Design and Construction; ICE: London, UK, 2015; pp. 473–487. [Google Scholar] [CrossRef]
  62. Hearn, N. Self-sealing, autogenous healing and continued hydration: What is the difference? Mater. Struct. 1998, 31, 563–567. [Google Scholar] [CrossRef]
  63. Vijay, K.; Murmu, M.; Deo, S.V. Bacteria based self healing concrete—A review. Constr. Build. Mater. 2017, 152, 1008–1014. [Google Scholar] [CrossRef]
  64. Nguyen, T.H.; Ghorbel, E.; Fares, H.; Cousture, A. Bacterial self-healing of concrete and durability assessment. Cem. Concr. Compos. 2019, 104, 103340. [Google Scholar] [CrossRef]
  65. Xu, J.; Wang, X.; Zuo, J.; Liu, X. Self-Healing of Concrete Cracks by Ceramsite-Loaded Microorganisms. Adv. Mater. Sci. Eng. 2018, 2018, 5153041. [Google Scholar] [CrossRef] [Green Version]
  66. Zhang, J.; Liu, Y.; Feng, T.; Zhou, M.; Zhao, L.; Zhou, A.; Li, Z. Immobilizing bacteria in expanded perlite for the crack self-healing in concrete. Constr. Build. Mater. 2017, 148, 610–617. [Google Scholar] [CrossRef]
  67. Wiktor, V.; Jonkers, H.M. Quantification of crack-healing in novel bacteria-based self-healing concrete. Cem. Concr. Compos. 2011, 33, 763–770. [Google Scholar] [CrossRef]
  68. Wang, J.Y.; de Belie, N.; Verstraete, W. Diatomaceous earth as a protective vehicle for bacteria applied for self-healing concrete. J. Ind. Microbiol. Biotechnol. 2012, 39, 567–577. [Google Scholar] [CrossRef]
  69. Xu, H.; Lian, J.; Gao, M.; Fu, D.; Yan, Y. Self-healing concrete using rubber particles to immobilize bacterial spores. Materials 2019, 12, 2313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Araújo, M.; Chatrabhuti, S.; Gurdebeke, S.; Alderete, N.; van Tittelboom, K.; Raquez, J.M.; Cnudde, V.; van Vlierberghe, S.; de Belie, N.; Gruyaert, E. Poly(methyl methacrylate) capsules as an alternative to the proof-of-concept’’ glass capsules used in self-healing concrete. Cem. Concr. Compos. 2018, 89, 260–271. [Google Scholar] [CrossRef]
  71. Wang, J.Y.; Soens, H.; Verstraete, W.; de Belie, N. Self-healing concrete by use of microencapsulated bacterial spores. Cem. Concr. Res. 2014, 56, 139–152. [Google Scholar] [CrossRef]
  72. Dong, B.; Fang, G.; Ding, W.; Liu, Y.; Zhang, J.; Han, N.; Xing, F. Self-healing features in cementitious material with urea-formaldehyde/epoxy microcapsules. Constr. Build. Mater. 2016, 106, 608–617. [Google Scholar] [CrossRef]
  73. Wang, J.Y.; Snoeck, D.; van Vlierberghe, S.; Verstraete, W.; de Belie, N. Application of hydrogel encapsulated carbonate precipitating bacteria for approaching a realistic self-healing in concrete. Constr. Build. Mater. 2014, 68, 110–119. [Google Scholar] [CrossRef]
  74. Thao, T.D.P.; Johnson, T.J.S.; Tong, Q.S.; Dai, P.S. Implementation of self-healing in concrete–Proof of concept. IES J. Part A Civ. Struct. Eng. 2009, 2, 116–125. [Google Scholar] [CrossRef]
  75. Van Tittelboom, K.; de Belie, N.; van Loo, D.; Jacobs, P. Self-healing efficiency of cementitious materials containing tubular capsules filled with healing agent. Cem. Concr. Compos. 2011, 33, 497–505. [Google Scholar] [CrossRef]
  76. Dry, C.M. Three designs for the internal release of sealants, adhesives, and waterproofing chemicals into concrete to reduce permeability. Cem. Concr. Res. 2000, 30, 1969–1977. [Google Scholar] [CrossRef]
  77. Yang, Z.; Hollar, J.; He, X.; Shi, X. A self-healing cementitious composite using oil core/silica gel shell microcapsules. Cem. Concr. Compos. 2011, 33, 506–512. [Google Scholar] [CrossRef]
  78. Anglani, G.; Tulliani, J.M.; Antonaci, P. Behaviour of pre-cracked self-healing cementitious materials under static and cyclic loading. Materials 2020, 13, 1149. [Google Scholar] [CrossRef] [Green Version]
  79. Formia, A.; Terranova, S.; Antonaci, P.; Pugno, N.M.; Tulliani, J.M. Setup of extruded cementitious hollow tubes as containing/releasing devices in self-healing systems. Materials 2015, 8, 1897–1923. [Google Scholar] [CrossRef] [PubMed]
  80. Formia, A.; Irico, S.; Bertola, F.; Canonico, F.; Antonaci, P.; Pugno, N.M.; Tulliani, J.-M. Experimental analysis of self-healing cement-based materials incorporating extruded cementitious hollow tubes. J. Intell. Mater. Syst. Struct. 2016, 27, 2633–2652. [Google Scholar] [CrossRef]
  81. Li, D.; Chen, B.; Chen, X.; Fu, B.; Wei, H.; Xiang, X. Synergetic effect of superabsorbent polymer (SAP) and crystalline admixture (CA) on mortar macro-crack healing. Constr. Build. Mater. 2020, 247, 118521. [Google Scholar] [CrossRef]
  82. Deng, H.; Liao, G. Assessment of influence of self-healing behavior on water permeability and mechanical performance of ECC incorporating superabsorbent polymer (SAP) particles. Constr. Build. Mater. 2018, 170, 455–465. [Google Scholar] [CrossRef]
  83. Lee, H.X.D.; Wong, H.S.; Buenfeld, N.R. Self-sealing of cracks in concrete using superabsorbent polymers. Cem. Concr. Res. 2016, 79, 194–208. [Google Scholar] [CrossRef] [Green Version]
  84. Wang, X.; Fang, C.; Li, D.; Han, N.; Xing, F. A self-healing cementitious composite with mineral admixtures and built-in carbonate. Cem. Concr. Compos. 2018, 92, 216–229. [Google Scholar] [CrossRef]
  85. Wang, X.F.; Yang, Z.H.; Fang, C.; Wang, W.; Liu, J.; Xing, F. Effect of carbonate-containing self-healing system on properties of a cementitious composite: Fresh, mechanical, and durability properties. Constr. Build. Mater. 2020, 235, 117442. [Google Scholar] [CrossRef]
  86. Wang, L.; Song, X.; Yang, H.; Wang, L.; Tang, S.; Wu, B.; Mao, W. Pore Structural and Fractal Analysis of the Effects of MgO Reactivity and Dosage on Permeability and F–T Resistance of Concrete. Fractal Fract. 2022, 6, 113. [Google Scholar] [CrossRef]
  87. Qureshi, T.; Kanellopoulos, A.; Al-Tabbaa, A. Autogenous self-healing of cement with expansive minerals-I: Impact in early age crack healing. Constr. Build. Mater. 2018, 192, 768–784. [Google Scholar] [CrossRef] [Green Version]
  88. Qureshi, T.; Kanellopoulos, A.; Al-Tabbaa, A. Autogenous self-healing of cement with expansive minerals-II: Impact of age and the role of optimised expansive minerals in healing performance. Constr. Build. Mater. 2019, 194, 266–275. [Google Scholar] [CrossRef] [Green Version]
  89. Borg, R.P.; Cuenca, E.; Brac, E.M.G.; Ferrara, L. Crack sealing capacity in chloride-rich environments of mortars containing different cement substitutes and crystalline admixtures. J. Sustain. Cem.-Based Mater. 2018, 7, 141–159. [Google Scholar] [CrossRef]
  90. Ferrara, L.; Krelani, V.; Moretti, F. On the use of crystalline admixtures in cement based construction materials: From porosity reducers to promoters of self healing. Smart Mater. Struct. 2016, 25, 084002. [Google Scholar] [CrossRef]
  91. De Weerdt, K.; Colombo, A.; Coppola, L.; Justnes, H.; Geiker, M.R. Impact of the associated cation on chloride binding of Portland cement paste. Cem. Concr. Res. 2015, 68, 196–202. [Google Scholar] [CrossRef]
  92. Donnini, J.; Lancioni, G.; Bellezze, T.; Corinaldesi, V. Bond behavior of FRCM carbon yarns embedded in a cementitious matrix: Experimental and numerical results. In Key Engineering Materials; Trans Tech Publications Ltd.: Bachs, Switzerland, 2017; pp. 305–312. [Google Scholar] [CrossRef]
  93. Lollini, F.; Carsana, M.; Gastaldi, M.; Redaelli, E.; Bertolini, L. The challenge of the performance-based approach for the design of reinforced concrete structures in chloride bearing environment. Constr. Build. Mater. 2015, 79, 245–254. [Google Scholar] [CrossRef] [Green Version]
  94. ISO 14564; Epoxy-Coated Steel for the Reinforcement of Concrete. ISO: London, UK, 1999.
  95. Yeomans, S.R. Galvanized Steel Reinforcement in Concrete; Elsevier: Amsterdam, The Netherlands, 2004. [Google Scholar] [CrossRef]
  96. Roventi, G.; Bellezze, T.; Barbaresi, E.; Fratesi, R. Effect of carbonation process on the passivating products of zinc in Ca(OH)2 saturated solution. Mater. Corros. 2013, 64, 1007–1014. [Google Scholar] [CrossRef]
  97. Roventi, G.; Bellezze, T.; Giuliani, G.; Conti, C. Corrosion resistance of galvanized steel reinforcements in carbonated concrete: Effect of wet–dry cycles in tap water and in chloride solution on the passivating layer. Cem. Concr. Res. 2014, 65, 76–84. [Google Scholar] [CrossRef]
  98. Maahn, E.; Sorensen, B. Influence of Microstructure on the Corrosion Properties of Hot-Dip Galvanized Reinforcement in Concrete. Corrosion 1986, 42, 187–196. [Google Scholar] [CrossRef]
  99. Bellezze, T.; Malavolta, M.; Quaranta, A.; Ruffini, N.; Roventi, G. Corrosion behaviour in concrete of three differently galvanized steel bars. Cem. Concr. Compos. 2006, 28, 246–255. [Google Scholar] [CrossRef]
  100. Bellezze, T.; Fratesi, R.; Tittarelli, F. Corrosion behaviour of galvanized steel rebars in the presence of coating discontinuities. In Corrosion in Reinforced Concrete Structures; Raupach, M., Ed.; Woodhead Publishing: Cambridge, UK, 2007; pp. 27–37. [Google Scholar] [CrossRef]
  101. Schiessl, P.; Reuter, C. Coated reinforcing steels development and application in Europe. In Proceedings of the International Conference on Corrosion, Houston, TX, USA, 4–8 March 1991. [Google Scholar]
  102. Sagues, A.; Powers, R.G. Corrosion and corrosione control of concrete structures in Florida. What can be learned? In Proceedings of the International Conference on Repair of Concrete Structures: From Theory to Practice in a Marine Environment, Svolvear, Norway, 28–30 May 1997; pp. 49–58. [Google Scholar]
  103. Lollini, F.; Carsana, M.; Gastaldi, M.; Redaelli, E. Corrosion behaviour of stainless steel reinforcement in concrete. Corros. Rev. 2019, 37, 3–19. [Google Scholar] [CrossRef] [Green Version]
  104. The Concrete Society. Guidance on the Use of Stainless Steel Reinforcement; Technical Report No. 51; The Concrete Society: Surrey, UK, 1998. [Google Scholar]
  105. Bertolini, L.; Gastaldi, M. Corrosion resistance of low-nickel duplex stainless steel rebars. Mater. Corros. 2011, 62, 120–129. [Google Scholar] [CrossRef]
  106. Gastaldi, M.; Bertolini, L. Effect of temperature on the corrosion behaviour of low-nickel duplex stainless steel bars in concrete. Cem. Concr. Res. 2014, 56, 52–60. [Google Scholar] [CrossRef]
  107. Bertolini, L.; Gastaldi, M.; Pastore, T.; Pedeferri, M.; Pedeferri, P. Effects of Galvanic coupling between carbon steel and stainless steel reinforcement in concrete. In Proceedings of the International Conference on Corrosion and Rehabilitation of Reinforced Concrete Structures, Orlando, FL, USA, 7–11 December 1998. [Google Scholar]
  108. Trejo, D.; Gardoni, P.; Kim, J. Long-Term Performance of Glass Fiber-Reinforced Polymer Reinforcement Embedded in Concrete. ACI Mater. J. 2011, 108, 605–613. [Google Scholar]
  109. Tuutti, K. Corrosion of Steel in Concrete; Swedish Foundation for Concrete Research: Stockholm, Sweden, 1982. [Google Scholar]
  110. Page, C.L.; Treadaway, K.W.J. Aspects of the electrochemistry of steel in concrete. Nature 1982, 297, 109–115. [Google Scholar] [CrossRef]
  111. Arup, H. The Mechanism of the Protection of Steel in Concrete, Corrosion of Reinforcment in Concrete, Construction; Crane, A.P., Ed.; Ellis Horwood Ltd.: Chichester, UK, 1983; pp. 151–157. [Google Scholar]
  112. Schiessl, P. Corrosion of Steel in Concrete; Wiley: London, UK, 1988. [Google Scholar]
  113. Bertolini, L. Steel corrosion and service life of reinforced concrete structures. Struct. Infrastruct. Eng. 2008, 4, 123–137. [Google Scholar] [CrossRef]
  114. Pacheco-torgal, F.; Melchers, R.; de Belie, N.; Shi, X.; van Tittelboom, K.; Perez, A.S. Eco-Efficient Repair and Rehabilitation of Concrete Infrastructures; Woodhead Publishing: Cambridge, UK, 2018. [Google Scholar]
  115. International Federation for Structural Concrete, fib. Model Code 2010; Bulletin no. 65; fib: Lausanne, Switzerland, 2010. [Google Scholar]
  116. EN 206:2013+A1:2016; Concrete-Specification, Performance, Production and Conformity. European Union: Maastricht, The Netherlands, 2021.
  117. EN 1992-1-1:2004; Eurocodice 2: Design of Concrete Structures-Part 1-1: General Rules and Rules for Buildings. European Union: Maastricht, The Netherlands, 2004.
  118. EN 197-1:2011; Cement-Part 1: Composition, Specifications and Conformity Criteria for Common Cements. European Union: Maastricht, The Netherlands, 2011.
  119. Carsana, M.; Frassoni, M.; Bertolini, L. Comparison of ground waste glass with other supplementary cementitious materials. Cem. Concr. Compos. 2014, 45, 39–45. [Google Scholar] [CrossRef]
  120. Lollini, F.; Redaelli, E.; Bertolini, L. Effects of portland cement replacement with limestone on the properties of hardened concrete. Cem. Concr. Compos. 2014, 46, 32–40. [Google Scholar] [CrossRef]
  121. FIsfahani, T.; Redaelli, E.; Lollini, F.; Li, W.; Bertolini, L. Effects of Nanosilica on Compressive Strength and Durability Properties of Concrete with Different Water to Binder Ratios. Adv. Mater. Sci. Eng. 2016, 2016, 8453567. [Google Scholar] [CrossRef] [Green Version]
  122. Carsana, M.; Gastaldi, M.; Lollini, F.; Redaelli, E.; Bertolini, L. Improving durability of reinforced concrete structures by recycling wet-ground MSWI bottom ash. Mater. Corros. 2016, 67, 573–582. [Google Scholar] [CrossRef]
  123. Lollini, F.; Redaelli, E.; Bertolini, L. A study on the applicability of the efficiency factor of supplementary cementitious materials to durability properties. Constr. Build. Mater. 2016, 120, 284–292. [Google Scholar] [CrossRef] [Green Version]
  124. Lollini, F.; Redaelli, E.; Bertolini, L. Investigation on the effect of supplementary cementitious materials on the critical chloride threshold of steel in concrete. Mater. Struct. Constr. 2016, 49, 4147–4165. [Google Scholar] [CrossRef] [Green Version]
  125. Carsana, M.; Bertolini, L. Durability of a LWC concrete with expanded glass and silica fume. ACI Mater. J. 2017, 114, 207–213. [Google Scholar]
  126. International Federation for Structural Concrete, fib. Model Code for Service Life Design; Bullettin No. 34; fib: Lausanne, Switzerland, 2006. [Google Scholar]
  127. Lollini, F.; Gastaldi, M.; Bertolini, L. Performance parameters for the durability design of reinforced concrete structures with stainless steel reinforcement. Struct. Infrastruct. Eng. 2018, 14, 833–842. [Google Scholar] [CrossRef]
  128. Bertolini, L.; Lollini, F.; Redaelli, E. Durability design of reinforced concrete structures. Proc. Inst. Civ. Eng. 2011, 164, 73–82. [Google Scholar] [CrossRef]
  129. Lollini, F.; Redaelli, E.; Bertolini, L. Analysis of the parameters affecting probabilistic predictions of initiation time for carbonation-induced corrosion of reinforced concrete structures. Mater. Corros. 2012, 63, 1059–1068. [Google Scholar] [CrossRef]
  130. Bertolini, L. Materiali da Costruzione; Città Studi Edizioni: Milan, Italy, 2012; Volume II. [Google Scholar]
  131. Mehta, P.K. Concrete. Structure, Properties and Materials; McGraw-Hill: New York, NY, USA, 1986. [Google Scholar]
  132. Bittrich, E.; Cometa, S.; de Giglio, E.; di Mundo, R.; Itaranto, N.D.; Eichhorn, K.; Keller, B.; Lednicky, F.; Mangolini, F.; Palumbo, F.; et al. Polymer Surface Characterization; GmbH&Co KG: München, Germany, 2014. [Google Scholar]
  133. Chalangaran, N.; Farzampour, A.; Palsar, N.; Fatemi, H. Experimental investigation of sound transmission loss in concrete containing recycled rubber crumbs. Adv. Concr. Constr. 2021, 11, 447–454. [Google Scholar] [CrossRef]
  134. Chalangaran, N.; Farzampour, A.; Paslar, N. Nano Silica and Metakaolin Effects on the Behavior of Concrete Containing Rubber Crumbs. CivilEng 2020, 1, 264–274. [Google Scholar] [CrossRef]
  135. Di Mundo, R.; Petrella, A.; Notarnicola, M. Surface and bulk hydrophobic cement composites by tyre rubber addition. Constr. Build. Mater. 2018, 172, 176–184. [Google Scholar] [CrossRef]
  136. Petrella, A.; di Mundo, R.; de Gisi, S.; Todaro, F.; Labianca, C.; Notarnicola, M. Environmentally Sustainable Cement Composites Based on End-of-Life Tyre Rubber and Recycled Waste Porous Glass. Materials 2019, 12, 3289. [Google Scholar] [CrossRef] [Green Version]
  137. Di Mundo, R.; Dilonardo, E.; Nacucchi, M.; Carbone, G.; Notarnicola, M. Water absorption in rubber-cement composites: 3D structure investigation by X-ray computed-tomography. Constr. Build. Mater. 2019, 228, 116602. [Google Scholar] [CrossRef]
  138. Zhu, H.; Liang, J.; Xu, J.; Bo, M.; Li, J.; Tang, B. Research on anti-chloride ion penetration property of crumb rubber concrete at different ambient temperatures. Constr. Build. Mater. 2018, 189, 42–53. [Google Scholar] [CrossRef]
  139. Bignozzi, M.C.; Manzi, S.; Natali, M.E.; Rickard, W.D.A.; van Riessen, A. Room temperature alkali activation of fly ash: The effect of Na 2O/SiO2 ratio. Constr. Build. Mater. 2014, 69, 262–270. [Google Scholar] [CrossRef]
  140. Provis, J.L.; Bernal, S.A. Geopolymers and Related Alkali-Activated Materials. Annu. Rev. Mater. Res. 2014, 44, 299–327. [Google Scholar] [CrossRef]
  141. Singh, B.; Ishwarya, G.; Gupta, M.; Bhattacharyya, S.K. Geopolymer concrete: A review of some recent developments. Constr. Build. Mater. 2015, 85, 78–90. [Google Scholar] [CrossRef]
  142. Carabba, L.; Manzi, S.; Bignozzi, M.C. Superplasticizer addition to carbon fly ash geopolymers activated at room temperature. Materials 2016, 9, 586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Zhuang, X.Y.; Chen, L.; Komarneni, S.; Zhou, C.H.; Tong, D.S.; Yang, H.M.; Yu, W.H.; Wang, H. Fly ash-based geopolymer: Clean production, properties and applications. J. Clean. Prod. 2016, 125, 253–267. [Google Scholar] [CrossRef]
  144. Carabba, L.; Santandrea, M.; Carloni, C.; Manzi, S.; Bignozzi, M.C. Steel fiber reinforced geopolymer matrix (S-FRGM) composites applied to reinforced concrete structures for strengthening applications: A preliminary study. Compos. Part B Eng. 2017, 128, 83–90. [Google Scholar] [CrossRef]
  145. Ma, C.-K.; Awang, A.Z.; Omar, W. Structural and material performance of geopolymer concrete: A review. Constr. Build. Mater. 2018, 186, 90–102. [Google Scholar] [CrossRef]
  146. Coppola, L.; Coffetti, D.; Crotti, E.; Dell’Aversano, R.; Gazzaniga, G. The influence of heat and steam curing on the properties of one-part fly ash/slag alkali activated materials: Preliminary results. AIP Conf. Proc. 2019, 2196, 020038. [Google Scholar] [CrossRef]
  147. Mobili, A.; Belli, A.; Giosuè, C.; Bellezze, T.; Tittarelli, F. Metakaolin and fly ash alkali-activated mortars compared with cementitious mortars at the same strength class. Cem. Concr. Res. 2016, 88, 198–210. [Google Scholar] [CrossRef]
  148. Pacheco-Torgal, F.; Abdollahnejad, Z.; Camões, A.F.; Jamshidi, M.; Ding, Y. Durability of alkali-activated binders: A clear advantage over Portland cement or an unproven issue? Constr. Build. Mater. 2012, 30, 400–405. [Google Scholar] [CrossRef] [Green Version]
  149. Hossain, M.M.; Karim, M.R.; Hossain, M.K.; Islam, M.N.; Zain, M.F.M. Durability of mortar and concrete containing alkali-activated binder with pozzolans: A review. Constr. Build. Mater. 2015, 93, 95–109. [Google Scholar] [CrossRef]
  150. Pasupathy, K.; Berndt, M.; Sanjayan, J.; Rajeev, P.; Cheema, D.S. Durability of low-calcium fly ash based geopolymer concrete culvert in a saline environment. Cem. Concr. Res. 2017, 100, 297–310. [Google Scholar] [CrossRef]
  151. Wardhono, A.; Gunasekara, C.; Law, D.W.; Setunge, S. Comparison of long term performance between alkali activated slag and fly ash geopolymer concretes. Constr. Build. Mater. 2017, 143, 272–279. [Google Scholar] [CrossRef]
  152. Ameri, F.; Shoaei, P.; Zareei, S.A.; Behforouz, B. Geopolymers vs. alkali-activated materials (AAMs): A comparative study on durability, microstructure, and resistance to elevated temperatures of lightweight mortars. Constr. Build. Mater. 2019, 222, 49–63. [Google Scholar] [CrossRef]
  153. Saccani, A.; Manzi, S.; Lancellotti, I.; Barbieri, L. Manufacturing and durability of alkali activated mortars containing different types of glass waste as aggregates valorization. Constr. Build. Mater. 2020, 237, 117733. [Google Scholar] [CrossRef]
  154. Provis, J.L. Geopolymers and other alkali activated materials: Why, how, and what? Mater. Struct. 2014, 47, 11–25. [Google Scholar] [CrossRef]
  155. Monticelli, C.; Natali, M.E.; Balbo, A.; Chiavari, C.; Zanotto, F.; Manzi, S.; Bignozzi, M.C. Corrosion behavior of steel in alkali-activated fly ash mortars in the light of their microstructural, mechanical and chemical characterization. Cem. Concr. Res. 2016, 80, 60–68. [Google Scholar] [CrossRef]
  156. Monticelli, C.; Natali, M.E.; Balbo, A.; Chiavari, C.; Zanotto, F.; Manzi, S.; Bignozzi, M.C. A study on the corrosion of reinforcing bars in alkali-activated fly ash mortars under wet and dry exposures to chloride solutions. Cem. Concr. Res. 2016, 87, 53–63. [Google Scholar] [CrossRef]
  157. Mo, K.H.; Alengaram, U.J.; Jumaat, M.Z. Structural performance of reinforced geopolymer concrete members: A review. Constr. Build. Mater. 2016, 120, 251–264. [Google Scholar] [CrossRef]
  158. Barbosa, V.F.F.; MacKenzie, K.J.D. Thermal behaviour of inorganic geopolymers and composites derived from sodium polysialate. Mater. Res. Bull. 2003, 38, 319–331. [Google Scholar] [CrossRef]
  159. Masi, G.; Rickard, W.D.A.; Bignozzi, M.C.; van Riessen, A. The effect of organic and inorganic fibres on the mechanical and thermal properties of aluminate activated geopolymers. Compos. Part B Eng. 2015, 76, 218–228. [Google Scholar] [CrossRef]
  160. Zhang, H.Y.; Kodur, V.; Qi, S.L.; Wu, B. Characterizing the bond strength of geopolymers at ambient and elevated temperatures. Cem. Concr. Compos. 2015, 58, 40–49. [Google Scholar] [CrossRef]
  161. Rickard, W.D.A.; Gluth, G.J.G.; Pistol, K. In-situ thermo-mechanical testing of fly ash geopolymer concretes made with quartz and expanded clay aggregates. Cem. Concr. Res. 2016, 80, 33–43. [Google Scholar] [CrossRef]
  162. Part, W.K.; Ramli, M.; Cheah, C.B. An overview on the influence of various factors on the properties of geopolymer concrete derived from industrial by-products. Constr. Build. Mater. 2015, 77, 370–395. [Google Scholar] [CrossRef]
  163. Carabba, L.; Manzi, S.; Rambaldi, E.; Ridolfi, G.; Bignozzi, M.C. High-temperature behaviour of alkali-activated composites based on fly ash and recycled refractory particles. J. Ceram. Sci. Technol. 2017, 8, 4416. [Google Scholar]
  164. Giosuè, C.; Mobili, A.; di Perna, C.; Tittarelli, F. Performance of lightweight cement-based and alkali-activated mortars exposed to high-temperature. Constr. Build. Mater. 2019, 220, 565–576. [Google Scholar] [CrossRef]
  165. Sakkas, K.; Nomikos, P.; Sofianos, A.; Panias, D. Inorganic polymeric materials for passive fire protection of underground constructions. Fire Mater. 2013, 37, 140–150. [Google Scholar] [CrossRef]
  166. Sotiriadis, K.; Guzii, S.; Kumpová, I.; Mácová, P.; Viani, A. The Effect of Firing Temperature on the Composition and Microstructure of a Geocement-Based Binder of Sodium Water-Glass. Solid State Phenom. 2017, 267, 58–62. [Google Scholar] [CrossRef]
  167. Lahoti, M.; Tan, K.H.; Yang, E.-H. A critical review of geopolymer properties for structural fire-resistance applications. Constr. Build. Mater. 2019, 221, 514–526. [Google Scholar] [CrossRef]
  168. Watolla, M.B.; Gluth, G.J.G.; Sturm, P.; Rickard, W.D.A.; Krüger, S.; Schartel, B. Intumescent geopolymer-bound coatings for fire protection of steel. J. Ceram. Sci. Technol. 2017, 8, 351–364. [Google Scholar] [CrossRef]
  169. Carabba, L.; Moricone, R.; Scarponi, G.E.; Tugnoli, A.; Bignozzi, M.C. Alkali activated lightweight mortars for passive fire protection: A preliminary study. Constr. Build. Mater. 2019, 195, 75–84. [Google Scholar] [CrossRef]
  170. Pozzo, A.D.; Carabba, L.; Bignozzi, M.C.; Tugnoli, A. Life cycle assessment of a geopolymer mixture for fireproofing applications. Int. J. Life Cycle Assess. 2019, 24, 1743–1757. [Google Scholar] [CrossRef]
  171. Carabba, L.; Pirskawetz, S.; Krüger, S.; Gluth, G.J.G.; Bignozzi, M.C. Acoustic emission study of heat-induced cracking in fly ash-based alkali-activated pastes and lightweight mortars. Cem. Concr. Compos. 2019, 102, 145–156. [Google Scholar] [CrossRef]
  172. Criado, M.; Martínez-Ramirez, S.; Fajardo, S.; Gómez, P.P.; Bastidas, J.M. Corrosion rate and corrosion product characterisation using Raman spectroscopy for steel embedded in chloride polluted fly ash mortar. Mater. Corros. 2013, 64, 372–380. [Google Scholar] [CrossRef] [Green Version]
  173. Law, D.W.; Adam, A.A.; Molyneaux, T.K.; Patnaikuni, I.; Wardhono, A. Long term durability properties of class F fly ash geopolymer concrete. Mater. Struct. 2015, 48, 721–731. [Google Scholar] [CrossRef]
  174. Tittarelli, F.; Mobili, A.; Giosuè, C.; Belli, A.; Bellezze, T. Corrosion behaviour of bare and galvanized steel in geopolymer and Ordinary Portland Cement based mortars with the same strength class exposed to chlorides. Corros. Sci. 2018, 134, 64–77. [Google Scholar] [CrossRef]
  175. Iorfida, A.; Verre, S.; Candamano, S.; Ombres, L. Tensile and Direct Shear Responses of Basalt-Fibre Reinforced Mortar Based Materials. In Strain-Hardening Cement-Based Composites: SHCC4; Mechtcherine, V., Slowik, V., Kabele, P., Eds.; Springer: Dordrecht, The Netherlands, 2018; pp. 544–552. [Google Scholar]
  176. Candamano, S.; Crea, F.; Iorfida, A. Mechanical characterization of basalt fabric-reinforced alkali-activated matrix composites: A preliminary investigation. Appl. Sci. 2020, 10, 2865. [Google Scholar] [CrossRef] [Green Version]
  177. Coppola, B.; Palmero, P.; Montanaro, L.; Tulliani, J.M. Alkali-activation of marble sludge: Influence of curing conditions and waste glass addition. J. Eur. Ceram. Soc. 2020, 40, 3776–3787. [Google Scholar] [CrossRef]
  178. Ye, H.; Radlińska, A. Shrinkage mechanisms of alkali-activated slag. Cem. Concr. Res. 2016, 88, 126–135. [Google Scholar] [CrossRef] [Green Version]
  179. Candamano, S.; de Luca, P.; Frontera, P.; Crea, F. Production of Geopolymeric Mortars Containing Forest Biomass Ash as Partial Replacement of Metakaolin. Environments 2017, 4, 74. [Google Scholar] [CrossRef] [Green Version]
  180. Sgambitterra, E.; Lamuta, C.; Candamano, S.; Pagnotta, L. Brazilian disk test and digital image correlation: A methodology for the mechanical characterization of brittle materials. Mater. Struct. Constr. 2018, 51, 19. [Google Scholar] [CrossRef]
  181. Pacheco-Torgal, F.; Labrincha, J.; Leonelli, C.; Palomo, A.; Chindaprasirt, P. Handbook of Alkali-Activated Cements, Mortars and Concretes; Woodhead Publishing: Cambridge, UK, 2014. [Google Scholar]
  182. Provis, J.L.; van Deventer, J.S.J. Alkali Activated Materials-State-of-the-Art Report-Rilem-Tc 224-AAM; Springer: Berlin/Heidelberg, Germany, 2014. [Google Scholar] [CrossRef]
  183. Coppola, L.; Coffetti, D.; Crotti, E.; Gazzaniga, G.; Pastore, T. The durability of one-part alkali activated slag-based mortars in different environments. Sustainability 2020, 12, 3561. [Google Scholar] [CrossRef]
  184. Bakharev, T.; Sanjayan, J.G.; Cheng, Y.B. Sulfate attack on alkali-activated slag concrete. Cem. Concr. Res. 2002, 32, 211–216. [Google Scholar] [CrossRef]
  185. Bakharev, T. Durability of geopolymer materials in sodium and magnesium sulfate solutions. Cem. Concr. Res. 2005, 35, 1233–1246. [Google Scholar] [CrossRef]
  186. Ye, H.; Chen, Z.; Huang, L. Mechanism of sulfate attack on alkali-activated slag: The role of activator composition. Cem. Concr. Res. 2019, 125, 105868. [Google Scholar] [CrossRef]
  187. Gong, K.; White, C.E. Nanoscale Chemical Degradation Mechanisms of Sulfate Attack in Alkali-activated Slag. J. Phys. Chem. C 2018, 122, 5992–6004. [Google Scholar] [CrossRef]
  188. Bhutta, M.A.R.; Hussin, W.M.; Azreen, M.; Tahir, M.M. Sulphate resistance of geopolymer concrete prepared from blended waste fuel ash. J. Mater. Civ. Eng. 2014, 26, 04014080. [Google Scholar] [CrossRef]
  189. Ismail, I.; Bernal, S.A.; Provis, J.L.; Hamdan, S.; van Deventer, J.S.J. Microstructural changes in alkali activated fly ash/slag geopolymers with sulfate exposure. Mater. Struct. Constr. 2013, 46, 361–373. [Google Scholar] [CrossRef]
  190. Telesca, A.; Marroccoli, M.; Winnefeld, F. Synthesis and characterisation of calcium sulfoaluminate cements produced by different chemical gypsums. Adv. Cem. Res. 2019, 31, 113–123. [Google Scholar] [CrossRef]
  191. Telesca, A.; Marroccoli, M.; Tomasulo, M.; Valenti, G.L.; Dieter, H.; Montagnaro, F. Low-CO2 Cements from Fluidized Bed Process Wastes and Other Industrial By-Products. Combust. Sci. Technol. 2016, 188, 492–503. [Google Scholar] [CrossRef]
  192. Winnefeld, F.; Lothenbach, B. Hydration of calcium sulfoaluminate cements—Experimental findings and thermodynamic modelling. Cem. Concr. Res. 2010, 40, 1239–1247. [Google Scholar] [CrossRef]
  193. Glasser, F.P.; Zhang, L. High-performance cement matrices based on calcium sulfoaluminate–belite compositions. Cem. Concr. Res. 2001, 31, 1881–1886. [Google Scholar] [CrossRef]
  194. Xu, L.; Wu, K.; Li, N.; Zhou, X.; Wang, P. Utilization of flue gas desulfurization gypsum for producing calcium sulfoaluminate cement. J. Clean. Prod. 2017, 161, 803–811. [Google Scholar] [CrossRef]
  195. Telesca, A.; Marroccoli, M.; Tomasulo, M.; Valenti, G.L.; Dieter, H.; Montagnaro, F. Calcium looping spent sorbent as a limestone replacement in the manufacture of portland and calcium sulfoaluminate cements. Environ. Sci. Technol. 2015, 49, 6865–6871. [Google Scholar] [CrossRef] [PubMed]
  196. Telesca, A.; Marroccoli, M.; Tomasulo, M.; Valenti, G.L. Hydration Properties and Technical Behavior of Calcium Sulfoaluminate Cements. Spec. Publ. 2015, 303, 237–254. [Google Scholar]
  197. Marroccoli, M.; Pace, M.L.; Telesca, A.; Valenti, G.L. Synthesis of calcium sulfoaluminate cement from Al2O3-rich by-products from aluminium manufacture. In Proceedings of the Second International Conference on Sustainable Construction Materials and Technologies, Ancona, Italy, 28–30 June 2010; pp. 615–623. [Google Scholar]
  198. Telesca, A.; Marroccoli, M.; Coffetti, D.; Coppola, L.; Candamano, S. Tartaric acid effects on hydration development and physico-mechanical properties of blended calcium sulphoaluminate cements. Cem. Concr. Compos. 2021, 124, 104275. [Google Scholar] [CrossRef]
  199. Juenger, M.C.G.; Winnefeld, F.; Provis, J.L.; Ideker, J.H. Advances in alternative cementitious binders. Cem. Concr. Res. 2011, 41, 1232–1243. [Google Scholar] [CrossRef]
  200. Su, M.; Wang, Y.; Sorrentino, F. Development in non-Portland cements. In Proceedings of the 9th International Congress on the Chemistry of Cement, New Delhi, India, 7 June 1992. [Google Scholar]
  201. Su, M.; Wang, Y.; Zhang, L.; Li, D. Preliminary study on the durability of sulfo/ferro-aluminate cements. In Proceedings of the 10th International Congress on the Chemistry of Cement, Gothenburg, Sweden, 2–6 June 1997; pp. 1–12. [Google Scholar]
  202. Zhang, L.; Su, M.; Wang, Y. Development of the use of sulfo- and ferroaluminate cements in China. Adv. Cem. Res. 1999, 11, 15–21. [Google Scholar] [CrossRef]
  203. Zhang, L. Microstructure and Performance of Calcium Sulfoaluminate Cements. Ph.D. Thesis, University of Aberdeen, Aberdeen, UK, 2000. [Google Scholar]
  204. Bernardo, G.; Telesca, A.; Valenti, G.L. A porosimetric study of calcium sulfoaluminate cement pastes cured at early ages. Cem. Concr. Res. 2006, 36, 1042–1047. [Google Scholar] [CrossRef]
  205. Quillin, K. Performance of belite–sulfoaluminate cements. Cem. Concr. Res. 2001, 31, 1341–1349. [Google Scholar] [CrossRef]
  206. Gastaldi, D.; Bertola, F.; Canonico, F.; Buzzi, L.; Mutke, S.; Irico, S.; Paul, G.; Marchese, L.; Boccaleri, E. A chemical/mineralogical investigation of the behavior of sulfoaluminate binders submitted to accelerated carbonation. Cem. Concr. Res. 2018, 109, 30–41. [Google Scholar] [CrossRef]
  207. Coppola, L.; Coffetti, D.; Crotti, E.; Dell’Aversano, R.; Gazzaniga, G.; Pastore, T. Influence of Lithium Carbonate and Sodium Carbonate on Physical and Elastic Properties and on Carbonation Resistance of Calcium Sulphoaluminate-Based Mortars. Appl. Sci. 2019, 10, 176. [Google Scholar] [CrossRef] [Green Version]
  208. Zhang, L.; Glasser, F.P. Investigation of the microstructure and carbonation of CS¯A-based concretes removed from service. Cem. Concr. Res. 2005, 35, 2252–2260. [Google Scholar] [CrossRef]
  209. Duan, P.; Chen, W.; Ma, J.; Shui, Z. Influence of layered double hydroxides on microstructure and carbonation resistance of sulphoaluminate cement concrete. Constr. Build. Mater. 2013, 48, 601–609. [Google Scholar] [CrossRef]
  210. Carsana, M.; Canonico, F.; Bertolini, L. Corrosion resistance of steel embedded in sulfoaluminate-based binders. Cem. Concr. Compos. 2018, 88, 211–219. [Google Scholar] [CrossRef]
Figure 1. Different approaches to self-healing.
Figure 1. Different approaches to self-healing.
Materials 15 02728 g001
Figure 2. Average value of the concrete cover thickness as a function of the type of concrete (OPC = Portland cement; GGBS = ground granulated blast furnace slag; water/binder = 0.45) and the type of bar that guarantees a service life of 100 years in the splash zone, assuming a target probability of failure of 10%.
Figure 2. Average value of the concrete cover thickness as a function of the type of concrete (OPC = Portland cement; GGBS = ground granulated blast furnace slag; water/binder = 0.45) and the type of bar that guarantees a service life of 100 years in the splash zone, assuming a target probability of failure of 10%.
Materials 15 02728 g002
Figure 3. Water uptake as a function of the square root of time for mortars containing siliceous sand (Sand) and rubber grains from end-of-life tires (Rubber). Larger (L) and smaller (S) granulometric fractions of rubber have been separately use.
Figure 3. Water uptake as a function of the square root of time for mortars containing siliceous sand (Sand) and rubber grains from end-of-life tires (Rubber). Larger (L) and smaller (S) granulometric fractions of rubber have been separately use.
Materials 15 02728 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Coppola, L.; Beretta, S.; Bignozzi, M.C.; Bolzoni, F.; Brenna, A.; Cabrini, M.; Candamano, S.; Caputo, D.; Carsana, M.; Cioffi, R.; et al. The Improvement of Durability of Reinforced Concretes for Sustainable Structures: A Review on Different Approaches. Materials 2022, 15, 2728. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15082728

AMA Style

Coppola L, Beretta S, Bignozzi MC, Bolzoni F, Brenna A, Cabrini M, Candamano S, Caputo D, Carsana M, Cioffi R, et al. The Improvement of Durability of Reinforced Concretes for Sustainable Structures: A Review on Different Approaches. Materials. 2022; 15(8):2728. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15082728

Chicago/Turabian Style

Coppola, Luigi, Silvia Beretta, Maria Chiara Bignozzi, Fabio Bolzoni, Andrea Brenna, Marina Cabrini, Sebastiano Candamano, Domenico Caputo, Maddalena Carsana, Raffaele Cioffi, and et al. 2022. "The Improvement of Durability of Reinforced Concretes for Sustainable Structures: A Review on Different Approaches" Materials 15, no. 8: 2728. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15082728

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop