13

Concrete, mortar and plaster using titanium dioxide nanoparticles: applications in pollution control, self-cleaning and photo sterilization

M. Vittoriadiamanti and M.P. Pedeferri,     Politecnico di Milano, Italy

Abstract:

Many advances have been made in understanding the mechanisms of TiO2 photoactivity and in developing its potential in addressing environmental issues. This chapter provides a short overview of these mechanisms, focusing on construction materials (mortar, plaster, concrete) modified through the addition of titanium dioxide nanoparticles. Examples will be given of laboratory experiments carried out in recent years, and of current applications in the built environment; attention will also be paid to the international standards that have been, and are still being, developed for this technology.

Key words

titanium dioxide

photocatalytic cement

self-cleaning

material ageing

13.1 Introduction

TiO2 has been used for centuries as white pigment in textiles and paints, well before the discovery of its photocatalytic properties. Curiously, alterations induced by TiO2 pigments in supporting materials (degradation of paints and fabrics, or bleaching of dyes) had already been observed, but not understood. It was only in twentieth century that these alterations could be related to its photoactivation. The following list summarizes advances in research in the last 100 years in understanding mechanisms of TiO2 photo-activation, and in developing and optimizing its properties:

• The first trace of scientific works on this subject is a paper published by Keidel in 1929, suggesting an active role of TiO2 in the fading of paints.

• In 1938 the photobleaching of dyes caused by UV-irradiated TiO2 was investigated by Goodeve and Kitchener, and ascribed to the presence of active oxygen species detected on TiO2 surfaces.

• In 1964 Doerffler and Hauffe first proposed a research paper whose title contained the term ‘heterogeneous photocatalysis’: zinc oxide was used as photocatalyst.

• In the same year, Kato and Masuo reported the use of a TiO2 suspension to photocatalyse the oxidation of tetralin, opening the way to similar processes proposed by McLintock and Ritchie on ethylene and propylene in 1965.

• In 1972 Fujishima and Honda for the first time reported the production of a TiO2-based electrochemical cell for water splitting, which consisted of a single crystal rutile photoanode and a platinum counter electrode. This is now known as the ‘Honda–Fujishima effect’.

• Frank and Bard (1977a, 1977b) first used TiO2 as remedy to environmental pollution issues by studying the liquid phase reduction of cyanide ions in the presence of TiO2 powders.

• This publication was followed in the same year by the development of nitrogen reducing solar cells – photocatalytic reduction of N2 to ammonia – performed on pure and Fe-doped TiO2 (Schrauzer and Guth, 1977).

• In 1978 TiO2-catalysed organic synthesis was introduced by Kreutler and Bard (photosynthesis of methane from acetic acid).

• In 1985 TiO2 was first applied successfully in biocide bacteria photokilling (Matsunaga et al., 1985); this concept was then applied to tumour cells (Fujishima et al., 1986).

• Matthews (1987) was the first to use TiO2 powders immobilized on a substrate for liquid phase organic photocatalysis, to avoid filtration and resuspension.

• In 1991 O’Regan and Gratzel proposed the use of TiO2 as anode in dye-sensitized photovoltaic cells.

• In 1995 the superhydrophilic effect was first noticed by Fujishima’s research group, leading to the development of self-cleaning and antifogging surfaces (Wang et al., 1997).

• The first photocatalytic TiO2 products were delivered commercially in the late 1990s in the form of self-cleaning tiles and glass: in 1995, the Japanese company TOTO Ltd. started the manufacture of antibacterial Cu- or Ag-containing TiO2 tiles, mainly for operating rooms (Watanabe et al., 1995).

• The Marunouchi Building (or Marubiru), in Tokyo, opened in 2002, one of the first buildings featuring self-cleaning windows (Ohtani 2010).

• Photoactivation was proved to induce a hardness modification in the surface of TiO2, due to compressive stresses that are induced by volume increases connected with the onset of the superhydrophilic state (Shibata et al., 2003).

• The church Dives in Misericordia (Rome, Italy), designed by Richard Meier for the Jubilee and finished in 2003, is one of the first examples of photocatalytic technology in Europe: it was built with TiO2-modified cement (former constructions: a school in Mortara, Italy, in 1999, and the Cité de la Musique in Chambéry, France, in 2000).

• 20,000 m2 of self-cleaning windows were installed in the terminal building of Chubu International Airport (Japan), completed in 2005 (Fujishima and Zhang, 2006).

• The application of TiO2 as cool material in energy-saving technologies for building was proposed: it exploits the latent heat flux generated by the evaporation of thin films of water on superhydrophilic surfaces (Irie et al., 2004).

It is commonly assumed that the first studies on TiO2 photoactivated properties date back to the early 1970s, when studies from Honda and Fujishima attracted the attention of the scientific community: the above list proves clearly attracted less attention.

Photoinduced processes originate from the absorption of light by TiO2, which causes an electron to be promoted to the conduction band (CB) leaving a hole in the valence band (VB). The electron-hole pair can be used to create electricity in photovoltaic solar cells, or to drive a chemical reaction: the latter action belongs to the category of heterogeneous photocatalysis. Trapping of holes at the TiO2 surface also causes a dramatic increase in the surface wettability, which is defined as photoinduced superhydrophilicity. These phenomena will be explained in more detail in the following section.

13.2 Principles of heterogeneous photocatalysis

Many transition metal oxides are known to display a photocatalytic behaviour, i.e., these substances can act as catalysts and promote oxidation or reduction reactions when activated by electromagnetic radiation. Among these materials, titanium dioxide in its polymorphic structures (anatase, rutile and brookite) is undoubtedly the most relevant and studied, although other oxides and sulphides have attracted scientists’ attention.

An ideal semiconductor for use in photocatalytic processes should have the following characteristics: chemical and biological inertia, ease of production and utilization, efficient solar light activation, high efficiency and low cost. Titanium dioxide possesses almost all the cited characteristics with the exception of the efficiency in solar light exploiting, since activation is promoted by UV light, which is less than 10% of solar light energy (Linsebigler et al., 1995). But, how does heterogeneous photocatalysis work? The mechanism must be understood in order to explore its applications in everyday life, and especially as a complementary function of construction materials, as we will describe in the second part of this chapter.

13.2.1 Semiconductor activation

The initial process for semiconductor-supported heterogeneous photocatalysis of organic and inorganic compounds is the generation of electron-hole pairs, initiated by light absorption with energy equal to or greater than the band gap (Fig. 13.1). The process under irradiation can therefore be divided in four steps, as described by Schiavello (1988):

image

13.1 Schematic representation of photoactivation mechanism: hυ = incident radiation energy, Eg = semiconductor bandgap, A = acceptor, D = donor. Phases are numbered as in the related text.

1. absorption of light, followed by the separation of the electron–hole couple,

2. adsorption of the reagents,

3. redox reaction,

4. desorption of the products.

Photoinduced charge transfer to other species (organic or inorganic molecules, water) relies on molecules migration and adsorption onto the semiconductor surface.

At its surface, the semiconductor can donate electrons to reduce an electron acceptor, given that the semiconductor CB bottom must be higher than the acceptor reduction energy, as a necessary condition. The acceptor in most cases is represented by oxygen, which forms a superoxide ion, O2•–, or hydrogen peroxide, H2O2: they both have excellent reactivity and play important roles in photocatalytic reactions.

On the other side of the band gap, holes can combine with electrons provided by donor species adsorbed on the semiconductor surface, thus oxidizing the donor itself: in this case, the necessary condition is that the top of the semiconductor valence band must be lower than the donor oxidation energy. Typical oxidation reactions take place between semiconductor and water molecules to form hydroxyl radicals OH, which are extremely active and tend to react easily due to their strong oxidizing power.

In both cases, the rate of charge transfer is strongly influenced by the respective positions of conduction and valence band edges and the redox potential levels of adsorbate species: the greater the difference between semiconductor and adsorbate energy levels, the faster the redox reactions (Linsebigler et al., 1995).

Charge transfer must compete with charge carrier recombination, which can take place in the volume of the semiconductor or at its surface and strongly decreases photocatalytic yield, i.e., the number of events occurring per absorbed photon. Charge carrier trapping can help increase the photogenerated species’ lifetime. Localized energy surface states ascribed to irregularities and surface defects can act as charge traps, and thus reduce recombination effects; adsorbed oxygen usually acts as an electron scavenger for the trapped electrons, while trapped holes can react with oxygen ions or hydroxyl groups.

13.2.2 The most common photocatalyst: titanium dioxide

Titanium dioxide (TiO2) is the most common among titanium minerals, and is extensively used in everyday life, specially as white pigment in painting, food and cosmetic industries, thanks to its high refractive index. It is found in nature in four polymorphs: anatase, which presents a distorted tetragonal crystal structure; rutile, which is also tetragonal; brookite, with orthorhombic crystal structure; and TiO2(B), with monoclinic structure. Only two of the phases, anatase and rutile, are interesting for practical applications, as they are wide band gap semiconductors.

The Eg value for anatase is 3.20 eV, which corresponds to a wavelength absorption threshold of 384 nm. This means that its activation requires an irradiating source with wavelength lower than that indicated, that is, in the near-UV region, while visible light is not sufficiently energetic to induce photoactivity in this material (Fig. 13.2).

image

13.2 Spectral irradiance of sunlight as a function of wavelength, highlighting ultraviolect (UV), visible (Vis) and infrared (IR) fractions.

The parameters which mostly concur to determine photoactivation efficiency, and specifically photocatalysis, are the following, as defined by Carp et al. (2004):

• catalyst surface area, considering also specific surface (porosity) available

• initial concentration of the compound to be degraded (saturation regime) and formation of intermediate products competing for adsorption, even deactivating TiO2

• reaction environment (oxygen, humidity, pH)

• wavelength and intensity of activating light source, which must provide enough energy to overcome the semiconductor band gap.

In Section 13.3 the main application fields of titanium dioxide will be described, as summarized in the well-known scheme of Fig. 13.3, together with the reasons for its success in such different environments. Before that, a brief look at other semiconductors that can be used as photocatalysts is required for a comprehensive treatise.

image

13.3 Main effects connected with TiO2 photoactivity.

13.2.3 Other photocatalysts

Titanium dioxide is by far the most studied photocatalyst, but several works are also dedicated to the study of other transition metal oxides and to the evaluation of their semiconducting nature and consequent photocatalytic activity. Probably the most important reason for this broadening of the sector is the fact that titanium dioxide can only be activated by UV light, which occupies a very narrow percentage of natural light: the choice of a photocatalyst which is active under visible light would allow a wider slice of sunlight to be exploited, and therefore increase the material efficiency. Moreover, UV light is also considered harmful to humans, as it can provoke skin and eye damage.

Possible alternative semiconductors that are susceptible to photoactivation are zinc oxide, ZnO (band gap 3.4 eV), mainly for the wide range of morphologies and properties achievable depending on the synthetic routes adopted, and tungsten oxide, WO3, whose band gap of 2.7 eV is promising for applications in visible light conditions. Yet, few examples can be found of their use in photocatalytic construction materials. Studies on ZnO-modified cements usually deal with the related modification of mechanical properties, and controversial results have been proposed: the material strength was proved to increase by substituting a few percent of cement with zinc oxide (Riahi and Nazari, 2011), while Lackhoff et al. (2003) observed a set-retarding effect and a correlated decrease in mechanical resistance due to water losses during the prolonged dormant phase before setting. As for WO3, tests were performed by Linkous et al. (2000) by coating a cement substrate with either WO3 or TiO2, and results indicated that the presence of WO3 alone induced a lower photoactivity compared to TiO2.

On the other hand, good photoactivation efficiencies were observed in hybrid oxides (mainly WO3-TiO2 and SiO2-TiO2): attention is now focusing on these composite systems as a means to improve the properties of photocatalytic materials. WO3-TiO2 hybrid nanoparticles were proved to exhibit photoactivity also under visible light activation (Chai et al., 2006). Yet, no information is available on the interaction between the composite WO3-TiO2 nanoparticles and mortars or concrete, especially concerning possible influences on the setting properties of the material, or on its final mechanical behaviour.

SiO2-supported TiO2 materials have been extensively used as photocatalysts for a wide variety of reactions. The higher efficiency observed in a few works compared to pure titanium dioxide has been ascribed to different physicochemical properties, which depend on the possible interactions between the two oxides and on synthesis conditions (sol-gel, grafting of TiO2 on SiO2, coprecipitation, impregnation, chemical vapour deposition). The most interesting point about silica-supported TiO2 is that silica itself is a common ingredient of cement-based materials, which makes the integration of these composite nanopowders easier (Bellardita et al., 2010).

13.3 Applications of semiconductor photocatalysis

13.3.1 Photocatalytic degradation of pollutants

TiO2 photocatalytic activity can be exploited to solve manifold purification issues arising not only from industrial uses or production of harmful substances, but also from heating and transportation. It exhibits good efficiency in the degradation of organic pollutants as well as inorganic compounds, from nitrogen oxides to complex metal salts, in gas and liquid phase.

Several works on photocatalysis focus on the removal of organic pollutants from wastewaters, specially referring to dyes, which are toxic to microorganisms and aquatic life: as reported by Konstantinou and Albanis (2004), up to 20% of dyes used in manufacturing processes are dispersed in waste-waters, therefore efficient processes that avoid their release into natural water sources are vital to preserve the ecosystem. Several reviews can be found on TiO2 effectiveness in dye degradation: we therefore suggest the reader refers to specific literature for further information (e.g., Akpan and Hameed, 2009; Han et al., 2009).

However, in recent years the aspect of removing gas phase organic compounds has also attracted attention. Extensive studies have been carried out on the removal of outdoor and indoor pollutants, and more specifically volatile organic compounds (VOCs), by TiO2 nanoparticles under UV illumination. These particles are often integrated in air purification devices, which are commercially available from several companies. VOCs are often responsible for malodorous air in buildings. They derive from several sources, such as cooking, cleaning products, furniture, etc.; in outdoor environments, they are usually part of combustion gases, industrial exhausts, cigarette smoke (Fujishima et al., 2007). A long list of experimental studies is available in the scientific literature on this topic as well.

Photocatalysts are not only used for breaking down large volumes of soilage, they are also capable of destroying it as it accumulates, e.g., to prevent cigarette smoke residue stains, or unpleasant odours due to the presence of VOCs, of the order of 10 ppb by volume. At these concentrations, TiO2 should be able to decompose such compounds even with scarce UV light (even as low as 1 μW/cm2 according to some authors). Bright UV lamps are also used in city and highway tunnels to reduce pollution released by traffic, degrading both VOCs and nitrogen oxides, NOx, as described by Demeestere et al. (2008), Toma et al. (2004), and many others.

13.3.2 Self-cleaning

Besides photocatalytic applications of TiO2, another fascinating phenomenon arises from UV irradiation, that is, the alteration of TiO2 wettability and formation of a highly hydrophilic surface state: this behaviour is defined as photoinduced superhydrophilicity (Wang et al., 1997), and involves the reduction of Ti4+ to Ti3+ by electrons and simultaneous hole trapping at lattice sites. This reduces the bond strength between reduced titanium and the closest oxygen, which is then removed when another water molecule arrives in contact with the surface and adsorbs on it. This creates a highly hydroxylated surface layer, which is responsible for hydrogen bonds with water and consequent increased hydrophilicity of the surface, as depicted in Fig. 13.4.

image

13.4 Theoretical mechanism and practical effect of photoinduced superhydrophilicity (photographs represent a plastic sheet covered on the left side with a TiO2 layer).

In this way, water can reach a contact angle close to zero on the surface of irradiated TiO2 (Drelich et al., 2011). Moreover, this surface is not solely hydrophilic: on the contrary, it presents an amphiphilic nature, with hydrophobic and hydrophilic domains of nanometre size alternating across the surface. This allows both oils and water to spread easily on the photoactivated TiO2 surface (Fujishima and Zhang, 2006).

Regardless of the in-depth study of the chemical and photochemical mechanisms involved, this effect is of extreme practical importance, as it defines one of the most renowned abilities of titanium dioxide and the reason why it has found so many applications in building materials: the self-cleaning effect.

The formation of an amphiphilic domain network is accompanied by photocatalytic activity, as both have a common origin: UV irradiation. This double photoinduced phenomenon results in the self-cleaning effect: surface contaminants are first photomineralized, at least in part, and subsequently washed away by water, which spreads below them in tight contact with the TiO2 surface. Moreover, drop formation on superhydrophilic surfaces is avoided, which in turn precludes stain formation due to slow water evaporation from the surface (Fig. 13.5). Another effect, anti-fogging, also arises from the same mechanism: no water droplets form on surfaces with a contact angle lower than 20°, as in the case of irradiated TiO2 (Fujishima and Zhang, 2006).

image

13.5 Schematic representation of self-cleaning on TiO2 containing surfaces: PM = particulate matter, MA = mineral acids.

To be precise, this property of self-cleaning should be referred to as ‘easy cleaning’: in fact, dirt and particles can adhere on the surface of titanium dioxide, even when irradiated; it is then extremely easy to remove them, as only UV light (available in natural sunlight) and water (that can be provided by rain) are required to remove stains and dirt.

Self-cleaning has become popular in a number of fields, even in clothing: some companies are producing self-cleaning cotton fabric, for an easier cleaning and deodorizing of clothes. Yet, their chief success is in the built environment, with the production of self-cleaning glass, tiles, paints, mortars and many other materials and components. In building facades, soiling is due to the adhesion of particulate matter on the surface of the materials they are made of, which are often porous (and therefore more prone to adsorbing such compounds). Particulate usually attaches to the surface through organic bonds, such as fatty acid chains and carboxylic groups. The twofold role of TiO2 is then the photocatalytic degradation of such groups and the onset of superhydrophilicity, which drives rainwater in direct contact with the surface, removing particles which are at that point loosely adherent, thus reducing the need for maintenance.

13.3.3 Antibacterial and anti-vegetative properties

Another possible way of exploiting the photocatalytic properties of TiO2 is its ability to mediate the destruction of bacteria, viruses and other biological materials. This function is often referred to as photosterilization, and it is of great interest for applications connected to air depuration – possibly of private housing, but mainly of medical-related environments, such as operating theatres, common rooms and patient rooms – due to the chance to induce the death of bacteria, viruses, as well as allergens and fungi.

The mechanism is similar to that of photocatalytic degradation: active species are once again surface hydroxyl radical species produced by photogenerated holes, and superoxide ions produced by photogenerated electrons, which damage or destroy cell walls of biological materials (Mills and Lee, 2002; Gerrity et al., 2008). A more detailed explanation is given by Hamal et al. (2009), who ascribe the photosterilization effect to the oxidation of complex proteins and to the inhibition of enzymatic functions of bacterial cells, which lead to ultimate cell death. In this respect, the employment of Ag-doped titanium dioxide is gaining much attention so as to achieve an easy decontamination and disinfection of common rooms and offices, as well as of medical equipment and operating theatres (Page et al., 2007; Wysocka-Krol et al., 2011).

Antibacterial activity was also described as a means to control biological growth on concrete surfaces and avoid unsightly stains. As described by Kurth et al. (2007), biofilm growth on concrete and mortar causes the triggering of undesirable chemical and aesthetical changes. The introduction of TiO2 in the material does not explicate a strictly bactericidal activity in this frame, but more properly an anti-vegetative effect. Polo et al. (2011) investigated the possibility of using titanium dioxide as a control technology to counteract biofilm microorganism growth, and reported the photokilling effect of immobilized TiO2 nanopowders on planktonic cells; conversely, no cell inactivation was observed on young biofilms, suggesting that possible applications of TiO2 should focus on preventative measures rather than on the disaggregation of already existing biofilms. Linkous et al. (2000), and Linkous and Robertson (2006) also reported the inhibition of algae adhesion on cement substrates and on roofing membranes modified with TiO2 (and WO3).

13.4 TiO2 in cement-based materials

Since their earliest appearance in Japan at the end of twentieth century, the diffusion of building materials modified with photocatalytic components has been constantly diffusing, spreading also to European countries. This is correlated with the increase in the generation of pollution and depletion of natural resources caused by intense and rapid industrial expansion, which pushes towards the development of sustainable materials, technologies and energy sources.

The widespread commitment inside the scientific community in the field of sustainability has focused much on purification devices, both for waste-water treatment and to improve the quality of air in industrial areas and large urban centres. This has driven great improvements in advanced oxidation processes (AOPs), specifically considering heterogeneous photocatalysis as a substitute for older, less efficient or more expensive techniques (Palmisano et al., 2007; Sievers, 2011).

In this field, a great number of scientific studies and patents deal with titanium dioxide nanopowder production and characterization (Paz, 2010; Shapovalov, 2010). Integrating TiO2 in construction materials has gained much interest (Fujishima and Zhang, 2006; Hüsken et al., 2007) in the attempt to meet air quality requirements promoted by several national and international committees, such as the quality standards introduced in early 1990s by the US Environmental Protection Agency (EPA), through the 1990 Clean Air Act Amendments which reported ozone, particulate matter, carbon monoxide, NOx, SO2 and lead as the most hazardous air pollutants causing severe concern for human health (Fig. 13.6).

image

13.6 (a) Generic sources of emission of atmospheric pollutants and related percentages; (b) specific case: typical sources of VOC release. (source: Environment Canada website) (source: Tokyo Metropolitan Government website)

As previously cited, several types of devices implementing the use of titanium dioxide have been designed and are currently commercialized. Examples of functionalized materials are photoactive paints for interior or exterior, tiles, self-cleaning fabrics for clothing; complete devices, like TiO2-containing air purifiers, are also available. These examples, which list just some of the available photocatalytic materials, make clear the interest in active principles capable of solving the cited air quality issues, or at least to mitigate them.

Construction materials represent the most easily available medium to distribute photoactive substances over the widest surface area possible, gaining the maximum efficiency thanks to a versatile support for the photocatalyst and to a limited increase in material costs. The introduction of heterogeneous photocatalysis principles in building materials also allows the exploitation of the self-cleaning attitude conferred by the simultaneous occurrence of (i) the degradation of greasy deposits accumulating on their surface and (ii) a state of photoinduced superhydrophilicity, with consequent washing away of reaction products, as reported by several authors including Mellott et al. (2006). The applications of these materials concern horizontal (cementitious tiles, pavings) and vertical structures (plasters, coatings, concrete structures) as well as galleries (cementitious paintings, concrete panels, asphalt coatings). Self-cleaning properties are mostly exploited in white concrete buildings: among the most representative examples, the Cité des Arts et de la Musique in Chambéry, France (Fig. 13.7a), completed in 2000, and the church Dives in Misericordia, built in Rome, Italy, by architect Richard Meier in 2003 (Fig. 13.7b).

image

13.7 Examples of buildings produced with TiO2-containing cement: (a) Cité des Arts et de la Musique, Chambéry, (b) church Dives in Misericordia, Rome.

Experimental works on TiO2-containing construction materials were carried out by several research groups, usually performing the addition of titanium dioxide nanopowders or suspensions, with varying particle size, to cement pastes, plasters, mortars and concretes, generally characterized by a low water-to-cement ratio. On the other hand, little information is available on the actual behaviour of the photocatalyst integrated in the material, and particularly on its evolution in time. A summary of experimental works proposed on photoactive construction materials is reported in the following sections.

13.4.1 Interaction with hydraulic and non-hydraulic binders

The introduction of TiO2 nanoparticles in building materials surely brings advantages, mainly from an economic point of view (i.e., the decrease in maintenance costs during the lifetime of the structure), and possibly an improvement of the surrounding air quality. The latter effect is actually connected with the extension of photocatalytic surfaces in relation with the volume of air to be depolluted: a limited space such as an indoor environment can benefit from this property, while a single building in a polluted city area will not be able to have a positive effect on the huge volume of air that comes into contact with it, this contact in most cases being very short due to wind.

Yet, the interaction of TiO2 with such a complex hosting environment may also have negative consequences. In fact, materials based on either hydraulic (cement, hydraulic lime) and non-hydraulic (gypsum, lime) binders consist of a mixture of calcium-based inorganic compounds, mainly calcium oxide/hydroxide, carbonate, silicate and sulphate. These constituents do not occupy the whole volume of the material, and materials are typically porous at the micro- and also nano-scale: this porosity is where TiO2 usually finds its collocation, acting as a further aggregate, or nano-filler. Therefore, all reaction products of cement hydration that remain unbounded in the material porosities can adsorb on the TiO2 surface, if small (e.g., impurities, germs of calcium hydroxide crystals, etc.), thus ‘stealing’ active sites to external polluting substances that could be degraded in a mechanism of competitive adsorption. Furthermore, a side effect of increased electron-hole couple recombination can also occur on adsorbed species (Lackhoff et al., 2003; Kwon et al., 2006).

This is the most evident influence of the alkaline hosting material on TiO2 photoactivated properties; it is then immediate to wonder whether TiO2 itself may lead to changes in the material characteristics. Concerning the fresh state, the key modification induced by TiO2 nanopowders is a decrease in workability: in fact, this is not correlated to the chemical nature of the particles, but to their nanometric size, which produces a drastic change in the rheological behaviour of the mix. This change is quite pronounced, and must be considered in order to assure a determined workability.

On the other hand, hardening properties are just slightly affected by TiO2. Attention has been focused on the observed increase in the compressive resistance of the material: this was mainly ascribed to the already cited filling effect, which was addressed by several works and summarized in the review by Sanchez and Sobolev (2010). Yet, controversies arise when a possible active behaviour of TiO2 is considered. Lackhoff et al. (2003) and Li et al. (2007) hypothesize a pozzolanic activity of TiO2; while the former justify this as an indirect consequence of an accelerated cement hydration observed with NMR relaxometry, in the latter case no experimental validation is provided, and the assumption is probably made in the wake of the pozzolanic activity of silica nanopowders proved in previous works (Ji, 2005; Jo et al., 2007). This was further supported by Nazari and Riahi in 2010, as a consequence of observed reduction in setting time and final porosity of TiO2-containing concrete, in spite of Chen and Poon’s observations (2009a) rebutting any pozzolanic nature of TiO2. Their experimental tests showed no TiO2 mass change during hydration, which suggests an inert behaviour of the nanopowders, whose maximum effect is that of providing a wide surface for nucleation or clinging of hydration products. This issue is still open for discussion.

The majority of works in this field are related to hydraulic binder-based materials, as they represent the larger volume of products, especially concerning the realization of new structures; a few works are also available on non-hydraulic binders, focusing on the conservation of historical and contemporary structures. In this specific context, Karatasios et al. (2010) investigated the ageing of lime mortars, and specifically their carbonation. They attributed to the presence of TiO2 an effect of CO2 release, which accelerated carbonation reactions. In fact, TiO2 presence would have caused the photocatalytic degradation of organic substances adsorbed on the material surface, releasing CO2 as a reaction product. The process was considered beneficial since surface carbonation of these materials decreases the risks of calcium leaching by run-off water, one of the main causes of binder degradation, due to the lower solubility of calcium carbonate compared to calcium hydroxide (Hansen et al., 2003).

13.4.2 Ageing of the material

Understanding the evolution of the material during its whole lifetime and the possible onset of negative interferences between TiO2 and its hosting environment is just as important as studying its beneficial effects.

The alkaline material undergoes carbonation in time, which decreases capillary absorption and causes the precipitation of calcium carbonate inducing a solid volume increase of the material higher than 10% (Ceukelaire and Nieuwenburg, 1993; Castellote et al., 2009). These precipitates tend to obstruct active sites, thus decreasing the photocatalytic efficiency of TiO2 mainly for a shielding effect (Chen and Poon, 2009b). This effect builds up with the accumulation of contaminants on surfaces exposed to the environment, as noted in a report of the Hong Kong Environmental Protection Department (Yu, 2003) (Fig. 13.8).

image

13.8 Scheme of possible progressive shielding of photocatalyst due to material ageing.

Although some data are already available, the influence of carbonation on the photocatalytic and self-cleaning efficiency of TiO2 embedded in construction materials has not been examined thoroughly yet. This can be a vital aspect to define the service life of such materials, as a careful design of the mix must take into account a possible decrease, and eventually the loss, of this functionality. This is true for any kind of element considered – mortar panels for cladding systems, repair plasters, and even more importantly if the whole structure is built with photocatalytic concrete, since its degradation cannot be easily overcome by substitution or re-application. A guideline dedicated to solving design issues with photocatalytic materials would therefore be a desirable target of further research.

13.5 Efficiency of TiO2 in the built environment

Experimental works concerning the use of TiO2 nanopowders in the construction field are variegated in terms of substrate material (bare cement pastes, mortars, concrete), composition (mix proportions with different binder, water/binder ratio, photocatalyst concentration, type and quantity of sand/aggregates, additives) and final application (precast panels, paving blocks, concrete pavements, cement-based tiles, indoor and outdoor walls, masonry blocks).

Characterization tests performed are also numerous, all involving a more or less intense and prolonged irradiation with UV light or simulated sunlight. Different experimental setups are applied in laboratory testing, depending on the property to be tested:

• photocatalytic activity: measurement of the extent of gas phase degradation of inorganic pollutants (NOx) or volatile organic compounds (VOCs) at the material surface

• self-cleaning: introduction of the material in a closed chamber with soiling atmosphere, or impregnation with an organic dye, and monitoring of colour loss (colour recovery); analysis of chromatic changes during extended exposure of materials to the external environment

• superhydrophilicity: measurement of contact angle of water on the material surface

• antibacterial and anti-vegetative effect: inhibition of biofilm formation, algae adhesion and proliferation, and sterilization effects.

13.5.1 Air depollution

Investigations on photocatalytic activity of cement pastes, mortars and concrete containing TiO2 nanopowders have been performed in most cases by flowthrough methods. Nitrogen monoxide, or directly a NOx (NO + NO2) mixture, are used as polluting source with typical concentration of approximately 1 ppmv (1 part per million in volume), and NO and NO2 concentrations are compared in the inlet and outlet gas flow; a chemiluminescent NOx analyser is used as measuring unit. Tests involve a first phase of gas flow in the absence of any irradiation, to reach an equilibrium of air composition in the reactor chamber and of the NO absorption in the cementitious material itself, which clearly must not be considered in the calculation of NO degradation. Afterwards, a UV light source (or solar spectrum lamp) is switched on and the outlet gas composition sampled at predefined time intervals: NO concentration usually drops immediately by some percent, and finally goes back to its initial value when the lamp is switched off and the test interrupted.

In NOx degradation tests, it is important to keep in mind that the first chemical reaction taking place at the TiO2 surface is the reduction of NO to NO2. Therefore, after a rapid decrease of NOx concentration, a slight increase in its value can be observed in the steady state of reaction, since the drop of NO concentration is accompanied by the formation of NO2. Finally, NO2 is reduced as well, by the following reactions:

image [13.1]

image [13.2]

Degradation efficiencies of laboratory specimens range from approximately 50–60% in flow conditions to the total degradation of pollutants in batch conditions, depending on gas concentration, flux and irradiation time. Humidity was observed in many cases to slightly reduce photocatalytic activity when above a certain threshold, as the adsorption of water molecules on the surface of TiO2 nanoparticles is competitive with respect to pollutant adsorption and consequent degradation (Hüsken et al., 2009).

The result of the reaction chain of NOx degradation has drawn much attention and some concern, since the final reaction step involves the dissolution of nitrate ions in rain to form nitric acid, and the consequent acidification of rainwater that reaches the sewers. Nonetheless, the concentration of NOx in air is in the order of magnitude of tens of ppb (parts per billion), and consequently the possible concentration of HNO3 that could reach sewers is extremely limited and is not expected to produce any detrimental effect.

Similar tests are performed with VOCs as polluting source, among which propanol, butanol, toluene, formaldehyde and acetone are the most diffuse model reactants. Also in this case, removal efficiencies are almost 100% in batch conditions and close to 60% in flow conditions. Experimental tests performed in flow conditions are undoubtedly more relevant than batch ones, as they are more representative of the actual working conditions of these materials in service.

13.5.2 Self-cleaning

One of the first studies on the self-cleaning attribute of photoactive materials is that proposed by Cassar in 2004, describing the impregnation of white cement disks with a yellow dye (phenanthroquinone) and the subsequent restoration of the initial white colour in specimens containing TiO2. Similar works on cement pastes and mortars were carried out with other organic dyes, such as rhodamine B. Aqueous solutions of the dye with a concentration of 0.05 g/L were spread on the surface of mortars and allowed to dry, then colour variations during irradiation were monitored with a spectrophotometer (Ruot et al., 2009).

This method allows an easy evaluation of the self-cleaning property. The spectrophotometer measures the colour of the surface, and converts it into a set of chromatic coordinates (usually the CIELab colour system, defined by the Commission Internationale de I’Éclairage, where L* is brightness, a* varies from green to red, and b* from blue to yellow). In this space, colour changes can be measured as geometric distance between two points, which correspond to two different colours to be compared (for example, A and B in Fig. 13.9). It is possible to calculate the overall colour change (Δcol), or to analyse only the changes in hue (Δhue), or to identify changes in the saturation of a single hue (Δsat). The latter parameter is exploited in dye degradation measurements: since dyes lose their colour when the molecular structure is degraded, colour saturation can be used as a representative parameter of dye concentration. In the case of rhodamine B, which exhibits a strong magenta colour, a decreased intensity of red component of colour indicates that the dye is being degraded.

image

13.9 Colour measurements in CIELab colour space.

Yet, it is important to keep in mind that colour changes do not necessarily mean complete degradation, and the organic molecule can still be partially integer even after the colour has disappeared, since it is sufficient to break the molecule chromophore groups to induce the colour loss. This is why these tests often refer to dye ‘decolonization’ rather than ‘mineralization’, since the latter can only be measured by TOC (total organic carbon) or spectrometric measurements.

Another important experimental path that can be chosen to test the self-cleaning attribute of cement-based materials is their exposure to a polluted environment and monitoring of surface colour, which is performed with the same colour measurements described for dye decolourization. The polluted environment can be created artificially or it is possible to expose some specimens directly to urban atmosphere. In the latter case, since experimental conditions (humidity, temperature, irradiation) cannot be chosen and regulated, it is fundamental to have a monitoring station close to the exposure site, from where atmospheric data can be obtained. It is then possible to correlate them with the colour measurements, the most relevant parameters being the amount of radiation reaching the surface during the day and the time of wetness, which is closely correlated to the number of rainy days. As for the interpretation of data, a lower colour change in time is expected from a self-cleaning material. Moreover, if the soiled surface of a photoactive material returns towards its original colour right after a rainy event, and the same trend is not observed in a similar ‘blank’ material (without TiO2), then this behaviour will be due to the onset of self-cleaning (Diamanti et al., 2008).

13.5.3 A case study of life cycle assessment

A life cycle assessment (LCA) of TiO2-containing coatings for concrete pavements was performed by Hassan in 2010. The author assumed that an improvement in air quality cannot be used as the only criterion for the complete evaluation of a material that should be considered sustainable, as critical environmental factors may be omitted. Therefore, a life-cycle inventory (LCI) was performed to quantify the energy, abiotic raw material inputs and emission from cradle to grave. A Building for Environmental and Economic Sustainability (BEES) impact assessment model (Lippiat, 2007) was applied to analyse the inventory.

The reduction of the environmental impact of the material was found to act on four main categories: acidification, eutrophication, air pollutants and smog formation. Parallel increases in global warming, fossil fuel depletion, water intake, ozone depletion and impacts on human health were also assessed, caused mainly by the production phase and fossil energy consumption. Yet, the total environmental performance of the product led to the conclusion that the titanium dioxide coatings tested have an overall beneficial effect on the environment.

13.6 Pilot projects and field tests

The confused mass of information described in the previous sections surely shows the usefulness of TiO2 in decreasing environmental pollution, and more broadly in improving the quality of materials where it is contained. Nonetheless, it implies at the same time great difficulties in understanding the actual behaviour of photocatalytic materials and in classifying them on the basis of their efficiencies.

These aspects are probably better investigated through large-scale experimental setups, which are fundamental to define the actual behaviour of materials modified with titanium dioxide in real practice. In this frame, Dylla et al. (2010) proposed a new laboratory setup to evaluate the influence of various parameters (humidity, pollutant flow rate, mix design) on the efficiency of photocatalytic coatings for concrete pavements, as well as on their resistance to abrasion and wear, given the critic application. A photoreactor with fluorescent lamps was built with a surrounding circuit providing the contaminant source (NO) and desired humidity. Particularly interesting are the dimensions of the photoreactor, 25 cm × 30 cm × 2.5 cm, which allows the evaluation of large samples, up to real-size paving slabs.

Besides large-scale tests, which still belong to the class of laboratory experiments, higher relevance must be given to pilot projects, as the most realistic way to measure the working efficiency directly on site. This is the scenario from which the project PICADA (Photocatalytic Innovative Coverings Applications for Depollution Assessment) took its first steps. In this project, a consortium of eight industries and research laboratories was created, aimed at developing and optimizing industrial formulations of innovative façade coatings with de-soiling and de-polluting properties including titanium dioxide, and at establishment of a local behaviour model under different exposure conditions and in a realistic urban environment.

The biggest pilot test in Europe is probably the ‘street canyon’ site that was built in France, near Guerville, in the frame work of the PICADA project (Guerrini et al., 2007). The methodology consisted in testing the effectiveness of photocatalytic properties on a 1 : 5 scale model reproducing the environmental conditions of a street located between two buildings, in a generic urban context. Two 18 m long lanes were built, and walls were covered with normal plaster in one case and photocatalytic plaster in the other; the atmosphere within the canyons was modified by flowing engine exhaust gases. Both air composition and environmental parameters were monitored, and the efficiency of photocatalytic plaster in keeping air cleaner was proved, with an up to fourfold decrease of contaminants compared to normal plaster in favourable weather conditions (wind, mainly).

Nowadays, examples of TiO2-containing building materials can be found almost anywhere: some of them are just pilot projects, like the ‘street canyon’, others are the final product of this technology. One early example of the use of photocatalytic TiO2 in the European built environment, and probably the most renowned, is the already cited church Dives in Misericordia, designed by Richard Meier and built with precast blocks of photocatalytic concrete produced by the Italcementi group (Fig. 13.7b). This building is often used as reference for the introduction and development of the use of photocatalytic concrete in Europe. It was finished in 2003, and for 7 years a constant monitoring of its colour was performed on the three ‘sails’ that form its architecture, showing no change in brightness in the areas analysed and only slight changes in the colour coordinates a* and b* on the panels facing south, which was ascribed to the deposition of African sand carried by the sirocco wind.

A demonstration project was carried out in Bergamo, Italy, where photocatalytic slabs were installed on a road and related pavement. Two periods of constant monitoring of NOx in the surrounding air, lasting 10 days each, showed that in the presence of photocatalytic paving slabs, the pollutant concentration decreased by almost 45%, being the average concentration in that area in the order of ppbv (Guerrini and Peccati, 2007). Other works on photocatalytic roads were performed in laboratory and field conditions by Ballari et al. (2010) and in large-scale application by Beeldens (2006) in Antwerp, Belgium, where a road was paved with photocatalytic building blocks, revealing again a decrease in NOx concentration.

Several other projects have been realized and are still monitored, such as tunnels, airports and schools, whose exterior walls were either built with photocatalytic materials or coated with TiO2-containing products, such as mortars, plasters, or just paints.

13.7 Existing patents and standards relating to photocatalytic cementitious materials

Although the attention of a large part of the scientific community has been devoted to titanium dioxide for decades, practical applications are more recent, and have found a worldwide diffusion only in the last few years. The market for photocatalytic construction materials is expected to grow from a volume of $800 million to $1.5 billion by 2014, as proposed in the market report ‘Photocatalysts: Technologies and Global Markets’ by BCC Research (2010). Commercial products focus mainly on the application of coatings on other materials, in spite of the problems connected to the durability of such products due to environmental factors, or just wear. TOTO Ltd was one of the first companies to notice the economic and innovation implications of the use of TiO2 in building materials, and collaborated closely with Fujishima and its research group when TiO2 photocatalysis and self-cleaning were still in their infancy.

An idea of the maturity of a product that still attracts so much basic and applied research is given by market volumes, but also the growing number of patent applications and standards promulgated is indicative of a massive passage from laboratory to real applications.

13.7.1 Overview of current patents

Nowadays, hundreds of patents exist on photocatalytic materials for air purification, self-cleaning surfaces, antibacterial surfaces, and many other sub-categories. Just as an example, in 2009 approximately 60 patent applications were submitted to the WIPO (World Intellectual Property Organization) based on a keyword search of the concept ‘photocatalysis’, among which 41% come from Japan: the major players are TOTO Ltd (Japan) and Carrier Corp. (USA), followed by Saint Gobain Glass (France) and Italcementi (Italy). Air purification is by far the most popular, and includes air purification devices and photocatalytic filters, which are beyond the scope of this chapter. A short list, surely not comprehensive, of some patents related to photocatalytic cementitious materials is reported in Table 13.1.

Table 13.1

Examples of patents on TiO2 in building materials

image

13.7.2 Standards for materials testing

Japan was undoubtedly the first country to appreciate the potential of photocatalytic materials, to invest in their application nationwide and to formulate appropriate standards to evaluate their efficiency. All standards that have been, or are being, developed are therefore based on Japanese experience, and often refer to the corresponding JIS (Japanese Industrial Standard), which include the test methods of photocatalytic materials for air purification performance (JIS R 1701), antibacterial activity (JIS R 1702), self-cleaning performance (JIS R 1703), water-purification performance (JIS R 1704) and antifungal activity (JIS R 1705).

In fact, every single outcome of the photoactivation of titanium dioxide requires a different standard, and also the type of supporting material strongly influences the test methods and the expected results: porous cementitious materials will necessarily undergo a different procedure with respect to compact ceramics, and further differences will characterize the testing of coatings. This short survey will only focus on standards referring to cement-based, and therefore intrinsically porous, photocatalytic materials.

The formulation of such standards in Europe was entrusted to an official working group of CEN (European Committee for Standardization), who was asked to define technical specifications and guidelines: the activity is not closed yet, and the group is still working on it. In the meanwhile, some documents were produced by the ISO (International Standards Organization) working group on photocatalytic materials, starting in 2007, focusing on photocatalytic fine ceramics (Table 13.2). Other standards are under evaluation and will be published probably in the next few years. These documents will surely become a reference to assess product performances for both public and private building contractors.

Table 13.2

ISO standards on photocatalytic fine ceramics

Year Code Topic
2011 ISO22197-2 Photocatalytic removal efficiency of acetaldehyde
2011 ISO22197-3 Photocatalytic removal efficiency of toluene
2011 IS010677 Definition of standard UV light source for testing photocatalytic performances
2010 IS010676 Water purification performances through the forming ability of active oxygen
2010 IS010678 Photocatalytic degradation of methylene blue
2009 IS027447 Evaluation of antibacterial activity of photocatalytic surfaces
2009 IS027448 Self-cleaning performances through the measurement of water contact angle
2007 IS022197-1 Photocatalytic removal efficiency of nitric oxide

13.8 References

Akpan, U.G., Hameed, B.H. Parameters affecting the photocatalytic degradation of dyes using TiO2-based photocatalysts: a review. J. Hazard. Mater. 2009; 170:520–529.

Ballari, M.M., Hunger, M., Hüsken, G., Brouwers, H. NOx photocatalytic degradation employing concrete pavement containing titanium dioxide. Appl. Catal. B. 2010; 95:245–254.

Beeldens, A., Environmental friendly concrete pavement blocks: air purification in the centre of Antwerp. Proceedings of 10th International Symposium on Concrete Roads. 2006. [Brussels].

Bellardita, M., Addamo, M., Di Paola, A., Marcì, G., Palmisano, L., Cassar, L., Borsa, M. Photocatalytic activity of TiO2/SiO2 systems. J. Hazard. Mater. 2010; 174:707–713.

Carp, O., Huisman, C.L., Reller, A. Photoinduced reactivity of titanium dioxide. Progr. Solid State Chem. 2004; 32:33–177.

Cassar, L. Photocatalysis of cementitious materials: clean buildings and clear air. MRS Bull. 2004; 29:328–331.

Castellote, M., Fernandez, L., Andrade, C., Alonso, C. Chemical changes and phase analysis of OPC pastes carbonated at different CO2 concentrations. Mater. Struct. 2009; 42:515–525.

Ceukelaire, L.D., Nieuwenburg, D.V. Accelerated carbonation of a blast-furnace cement concrete. Cem. Concr. Res. 1993; 23:442–452.

Chai, S.Y., Kim, Y.J., Lee, W.I. Photocatalytic WO2/TiO2 nanoparticles working under visible light. J. Electroceram. 2006; 17:909–912.

Chen, J., Poon, C.S. Photocatalytic cementitious materials: influence of the microstructure of cement paste on photocatalytic pollution degradation. Environ. Sci. Technol. 2009; 43:8948–8952.

Chen, J., Poon, C.S. Photocatalytic construction and building materials: from fundamentals to applications. Build. Environ. 2009; 44:1899–1906.

Clean Air Act Amendments of 1990 §101–131, 42 U.S.C. §7401–7431.

Demeestere, K., Dewulf, J., De Witte, B., Beeldens, A., Van Langenhove, H. Heterogeneous photocatalytic removal of toluene from air on building materials enriched with TiO2. Build. Environ. 2008; 43:406–414.

Diamanti, M.V., Ormellese, M., Pedeferri, M.P. Characterization of photocatalytic and superhydrophilic properties of mortars containing titanium dioxide. Cem. Concr. Res. 2008; 38:1349–1353.

Doerffler, W., Hauffe, K. Heterogeneous photocatalysis I. The influence of oxidizing and reducing gases on the electrical conductivity of dark and illuminated zinc oxide surfaces. J. Catal. 1964; 3:156–170.

Drelich, J., Chibowski, E., Meng, D.D., Terpilowski, K. Hydrophilic and super-hydrophilic surfaces and materials. Soft Matter. 2011; 7:9804–9828.

Dylla, H., Hassan, M.M., Mohammad, E.N., Rupnow, T., Wright, E. Evaluation of environmental effectiveness of titanium dioxide photocatalyst coating for concrete pavement. Transport. Res. Record. 2010; 2164:46–51.

Frank, S.N., Bard, A.J. Heterogeneous photocatalytic oxidation of cyanide ion in aqueous solutions at titanium dioxide powder. J. Am. Chem. Soc. 1977; 99:303–304.

Frank, S.N., Bard, A.J. Heterogeneous photocatalytic oxidation of cyanide and sulfite in aqueous solutions at semiconductor powders. J. Phys. Chem. 1977; 81:1484–1488.

Health Perspectives, 109, A174–A177.

Fujishima, A., Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature. 1972; 238:37–38.

Fujishima, A., Zhang, X.T. Titanium dioxide photocatalysis: present situation and future approaches. Cr. Chim. 2006; 9:750–760.

Fujishima, A., Ohtsuki, J., Yamashita, T., Hayakawa, S. Behavior of tumor cells on photoexcited semiconductor surface. Photomed Photobiol. 1986; 8:45–46.

Fujishima, A., Zhang, X., Tryk, D.A. Heterogeneous photocatalysis: from water photolysis to applications in environmental cleanup. Int. J Hydrogen Energy. 2007; 32:2664–2672.

Gerrity, D., Ryu, H., Crittenden, J., Abbaszadegan, M. Photocatalytic inactivation of viruses using titanium dioxide nanoparticles and low-pressure UV lamp. J. Environ. Sci Health A. 2008; 43:1261–1270.

Goodeve, C.F., Kitchener, J.A. Photosensitisation by titanium dioxide. Trans. Farad. Soc. 1938; 34:570–579.

Guerrini, G.L., Peccati, E., Photocatalytic cementitious roads for depollution. P. Baglioni, L. Cassar. RILEM Int. Symp. on Photocatalysis, Environment and Construction Materials, Italy, RILEM. 2007:179–186.

Guerrini, G.L., Plassais, A., Pepe, C., Cassar, L., Use of photocatalytic cementitious materials for self-cleaning applications. P. Baglioni, L. Cassar. RILEM Int. Symp. on Photocatalysis, Environment and Construction Materials, Italy, RILEM. 2007:219–226.

Hamal, D.B., Haggstrom, J.A., Marchin, G.L., Ikenberry, M.A., Hohn, K., Klabunde, K.J. A multifunctional biocide/sporocide and photocatalyst based on titanium dioxide (TiO2) codoped with silver, carbon, and sulfur. Langmuir. 2009; 26:2805–2810.

Han, F., Kambala, V.S.R., Srinivasan, M., Rajarathnam, D., Naidu, R. Tailored titanium dioxide photocatalysts for the degradation of organic dyes in wastewater treatment: a review. Appl. Catal. A. 2009; 359:25–40.

Hansen, E., Doehne, E., Fidler, J., Larson, J., Martin, B., Matteini, M., Rodriguez-Navarro, C., Pardo, E.S., Price, C., de Tagle, A., Teutonico, J.M., Weiss, N. A review of selected inorganic consolidants and protective treatments for porous calcareous materials. Rev. Conserv. 2003; 4:13–25.

Hassan, M.M. Quantification of the environmental benefits of ultrafine/nanotitanium dioxide photocatalyst coatings for concrete pavement using hybrid life-cycle assessment. J. Infrastruct. Sys. 2010; 16:160–166.

Hüsken, G., Hunger, M., Brouwers, H., Comparative study on cementitious products containing titanium dioxide as photo-catalyst. P. Baglioni, L. Cassar. RILEM Int. Symp. on Photocatalysis, Environment and Construction Materials, Italy, RILEM. 2007:147–154.

Hüsken, G., Hunger, M., Brouwers, H. Experimental study of photocatalytic concrete products for air purification. Build. Environ. 2009; 44:2463–2474.

Irie, H., Sunada, K., Hashimoto, K. Recent developments in TiO2 photocatalysis: novel applications to interior ecology materials and energy saving systems. Electrochemistry. 2004; 72:807–812.

Ji, T. Preliminary study on the water permeability and microstructure of concrete incorporating nano-SiO2. Cem. Concr. Res. 2005; 35:1943–1947.

Jo, B.W., Kim, C.H., Tae, G., Park, J.B. Characteristics of cement mortar with nano-SiO2 particles. Construct. Build. Mater. 2007; 21:1351–1355.

Karatasios, I., Katsiotis, M.S., Likodimos, V., Kontos, A.I., Papavassiliou, G., Falaras, P., Kilikoglou, V. Photo-induced carbonation of lime-TiO2 mortars. Appl. Catal. B. 2010; 95:78–86.

Kato, S., Masuo, F. Titanium dioxide-photocatalyzed oxidation. I. Titanium dioxide-photocatalyzed liquid phase oxidation of tetralin. Kogyo Kagaku Zasshi. 1964; 67:42–50.

Keidel, E. Die Beeinflussung der Lichtechtheit von Teerfarblacken durch Titanweiss [Influence of titanium white on the fastness to light of coal-tar days]. Farben-Zeitung. 1929; 34:1242–1243.

Konstantinou, I.K., Albanis, T.A. TiO2-assisted photocatalytic degradation of azo dyes in aqueous solution: kinetic and mechanistic investigations – a review. Appl. Catal. B: Environ. 2004; 49:1–14.

Kreutler, B., Bard, A.J. Heterogeneous photocatalytic synthesis of methane from acetic acid – new Kolbe reaction pathway. J. Am. Chem. Soc. 1978; 100:2239–2240.

Kurth, J.C., Giannantonio, D.J., Allain, F., Sobecky, P.A., Kurtis, K.E., Mitigating biofilm growth through the modification of concrete design and practiceBaglioni P., Cassar L., eds. RILEM Int. Symp. on Photocatalysis, Environment and Construction Materials, Italy, RILEM. 2007:195–202.

Kwon, J.M., Kim, Y.H., Song, B.K., Yeom, S.H., Kim, B.S., Bin, I. Novel immobilization of titanium dioxide (TiO2) on the fluidizing carrier and its application to the degradation of azo-dye. J. Hazard. Mater. 2006; 134:230–236.

Lackhoff, M., Prieto, X., Nestle, N., Dehn, F., Niessner, R. Photocatalytic activity of semiconductor-modified cement – influence of semiconductor type and cement ageing. Appl. Catal. B. 2003; 43:205–216.

Li, H., Zhang, M., Ou, J. Flexural fatigue performance of concrete containing nano-particles for pavement. Int. J. Fatigue. 2007; 29:1292–1301.

Linkous, C.A., Robertson, R.H., Effect of photocatalytic coatings on the weathering of elastomeric roofing membrane. Proceedings of the 15th Symposium on Improving Building Systems in Hot and Humid Climates, Orlando, FL. 2006. [ESL-HH-06-07-33].

Linkous, C.A., Carter, G.J., Locuson, D.B., Ouellette, A.J., Slattery, D.K., Smitha, L.A. Photocatalytic inhibition of algae growth using TiO2, WO3, and cocatalyst modifications. Environ. Sci. Technol. 2000; 34:4754–4758.

Linsebigler, A.L., Lu, G., Yates, J.T. Photocatalysis on TiO2 surfaces: principles, mechanisms, and selected results. Chem. Rev. 1995; 95:735–758.

Lippiat, B., Building for environmental and economic sustainability (BEES)Technical manual and user guide. Springfield, VA: National Institute of Standards and Technology, 2007.

Matsunaga, T., Tomato, R., Nakajima, T., Wake, H. Photoelectrochemical sterilization of microbial cells by semiconductor powders. FEMS Microbiol. Lett. 1985; 29:211–214.

Matthews, R.W. Photooxidation of organic impurities in water using thin films of titanium dioxide. J. Phys. Chem. 1987; 91:3328–3333.

McLintock, S., Ritchie, M. Reactions on titanium dioxide; photoadsorption and oxidation of ethylene and propylene. Trans. Faraday Soc. 1965; 61:1007–1016.

Mellott, N.P., Durucan, C., Pantano, C.G., Guglielmi, M. Commercial and laboratory prepared titanium dioxide thin films for self-cleaning glasses: photocatalytic performance and chemical durability. Thin Solid Films. 2006; 502:112–120.

Mills, A., Lee, S.K. A web-based overview of semiconductor photochemistry-based current commercial applications. J. Photochem. Photobiol. A. 2002; 152:233–247.

Nazari, A., Riahi, S. The effects of TiO2 nanoparticles on properties of binary blended concrete. J. Compos. Mater. 2010; 45:1181–1188.

O’Regan, B., Gratzel, M. A low-cost, high-efficiency solar cell based on dyesensitized colloidal TiO2 films. Nature. 1991; 353:737–740.

Ohtani, B., Environmental applications of photocatalysis. Special Lectures on Environmental Science II. 2010 Available at: pcat.cat.hokudai.ac.jp/class/es2012/20120405b_BO_Sapporo.pdf [(accessed August 2012)].

Page, K., Palgrave, R.G., Parkin, I.P., Wilson, M., Savin, S.L.P., Chadwick, A.V. Titania and silver–titania composite films on glass – potent antimicrobial coatings. J. Mater. Chem. 2007; 17:95–104.

Palmisano, G., Augugliaro, V., Pagliaro, M., Palmisano, L., Photocatalysis: a promising route for 21st century organic chemistry. Chem. Commun. 2007:3425–3437.

Paz, Y. Application of TiO2 photocatalysis for air treatment: patents’ overview. Appl. Catal. B. 2010; 99:448–460.

Polo, A., Diamanti, M.V., Bjarnsholt, T., Høiby, N., Villa, F., Pedeferri, M.P., Cappitelli, F. Effects of photoactivated titanium dioxide nanopowders and coating on planktonic and biofilm growth of Pseudomonas aeruginosa. Photochem. Photobiol. 2011; 87:1387–1394.

Riahi, S., Nazari, A. Physical, mechanical and thermal properties of concrete in different curing media containing ZnO2 nanoparticles. Energy Build. 2011; 43:1977–1984.

Ruot, B., Plassais, A., Olive, F., Guillot, L., Bonafous, L. TiO2-containing cement pastes and mortars: Measurements of the photocatalytic efficiency using a rhoda-mine B-based colourimetric test. Sol. Energy. 2009; 83:1794–1801.

Sanchez, F., Sobolev, K. Nanotechnology in concrete – a review. Construct. Build. Mater. 2010; 24:2060–2071.

Schiavello, M., Basic concepts in photocatalysis. M. Schiavello. Photocatalysis and Environment Trends and Applications, Dordrecht, Kluwer. 1988:351–360.

Schrauzer, G.N., Guth, T.D. Photocatalytic reactions. 1. Photolysis of water and photoreduction of nitrogen on titanium dioxide. J. Am. Chem. Soc. 1977; 99:7189–7193.

Shapovalov, V.I. Nanopowders and films of titanium oxide for photocatalysis: a review. Glass Phys. Chem. 2010; 36:121–157.

Shibata, T., Irie, H., Hashimoto, K. Enhancement of photoinduced highly hydrophilic conversion on TiO2 thin films by introducing tensile stress. J. Phys. Chem. B. 2003; 107:10696–10698.

Sievers, M. Advanced oxidation processes. In: Wilderer, P., eds. Treatise on Water Science, vol. 4. Oxford: Elsevier; 2011:377–408.

Toma, F.L., Bertrand, G., Klein, D., Coddet, C. Photocatalytic removal of nitrogen oxides via titanium dioxide. Environ. Chem. Lett. 2004; 2:117–121.

Wang, R., Hashimoto, K., Fujishima, A., Chikuni, M., Kojima, E., Kitamura, A., Shimo-higoshi, M., Watanabe, T. Light-induced amphiphilic surfaces. Nature. 1997; 388:431–432.

Watanabe, T., Kojima, E., Norimoto, K., Saeki, Y. Fabrication of TiO2 photocatalytic tile and practical applications. Fourth Euro Ceramics. 1995; 11:175–180.

Wysocka-Krol, K., Wieliczko, A., Podbielska, H., Silver doped nanomaterials and their possible use for antibacterial photodynamic activityMohseni H., Agahi M.H., Razeghi M., eds. Biosensing and Nanomedicine IV – Proceedings of SPIE, Bellingham, SPIE-Int Soc Optical Engineering. 2011:809915.

Yu, C.M. Deactivation and Regeneration of Environmentally Exposed Titanium Dioxide (TiO2) Based Products. Department of Chemistry, Chinese University of Hong Kong; 2003.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset