Open access

Applications of Mesoporous Ordered Semiconductor Materials — Case Study of TiO2

Written By

Antonio E. H. Machado, Karen A. Borges, Tatiana A. Silva, Lidiaine M. Santos, Mariana F. Borges, Werick A. Machado, Bruno P. Caixeta, Marcela Dias França, Samuel M. Oliveira, Alam G. Trovó and Antonio O.T. Patrocínio

Submitted: 11 June 2014 Published: 06 May 2015

DOI: 10.5772/59602

From the Edited Volume

Solar Radiation Applications

Edited by Segun Raphael Bello

Chapter metrics overview

2,777 Chapter Downloads

View Full Metrics

1. Introduction

1.1. Titanium dioxide (TiO2)

TiO2 is a promising material for technological applications for its versatility [1-4], abundance, low toxicity, good chemical stability, photosensitivity and photostability [5-8]. In nature, it is found mainly in the mineral ilmenite [9], that can be processed industrially by two different routes [10]: The first, involves the reaction of the concentrate of ilmenite with hot sulphuric acid, resulting in the formation of sulphates of titanium, Fe(II) and Fe(III), being these last eliminated by centrifugation, after cooling. The final solution is then purified and hydrolyzed to produce pure TiO2 [11]. The other usual way of obtaining consists in combining the ore with coke and gaseous chlorine under heating, resulting in CO2 and a spongy material rich in TiCl4. The reaction product is subjected to successive fractional distillation, with the formation of TiCl2 and TiCl3, due to stability of titanium in other degrees of oxidation. The different precursors of titanium are hydrolyzed, forming titanium dioxide [12].

In 1972, [13], obtained, for the first time, success in decomposition of water under irradiation with light and without application of any electric potential. They reported that, in trials where they employed n-type TiO2 as anode and Platinum as cathode, when TiO2 electrode illuminated under short-circuit conditions, hydrogen evolved from the Platinum electrode, while oxygen evolved from the anode. Since then, studies aimed at discovering and exploring different possibilities of technological application for TiO2 have been given great importance [14-22].

As technological applications for TiO2 can be cited its use in ultraviolet radiation absorbing filters [3, 23], in chemical sensors for gases [24-26], as a bactericide [27], in biomaterials for bone implants [28], in environmental photocatalysis [8, 24, 29, 30], in the photocatalytic hydrogen evolution [17, 31-36], in dye-sensitized solar cells [21, 37-44], among other.

The photocatalytic efficiency of TiO2 depends on its structural and morphological characteristics, which are related to the method of synthesis used in the preparation of nanoparticles [18, 21, 29, 45]. To be photoactive, favoring the photocatalysis process, besides being mainly consisting of anatase crystalline phase, the TiO2 must possess high specific surface area, good porosity, with high sized pores [35, 45, 46]. In this context, the search for TiO2 particles that have differentiated features, with catalytic properties potentiated, constitutes a field of intense activity [19, 20, 47-50].

Advertisement

2. Porosity and porous materials

The porosity, related to the presence of cavities, channels or interstices is of great importance since it is related to the ability of materials to interact with atoms, ions, molecules, and nanoparticles not only at their surfaces but also throughout the bulk [51, 52]. Thus, the control of porosity is very important if the objective is the development of new materials [53, 54].

The pores are classified as closed or open, considering the ability of porous materials to interact with their neighborhoods. Figure 1 shows various types of closed (a) or opened (b, c, f) pores.

Figure 1.

Schematic representation of a porous solid, presenting the most common types of pores: (a) closed; (b, c, d) opened; (e) interconnected; (f) surface roughness. Font: [55].

Open pores have continuous channels that communicate to the outer surface of the material, generating a cross-linked structure. Closed pores are inactive for the flow of liquids, gases and other substrates, being totally isolated from their neighborhood. These pores are related to macroscopic properties such as mechanical resistance and thermal conductivity [55]. The pores can also be interconnected, (e). Otherwise used to classify the pores takes into account its format: bottleneck, (b), cylindrical, (c), and funnel type, (d). The surface roughness, represented by (f), is also a type of porosity [55].

Figure 2.

Representation of the six main types of isotherm: I, microporous solids; II, non-porous solids; III, macroporous solids; IV and V, mesoporous solids; VI, non-porous solids with uniform surface.

IUPAC recommends a quantitative division of the pores in three classes according to their mean diameter: micropores, lower than 2 nm; mesopores, between 2 and 50 nm; macropores, higher than 50 nm [56]. This pore size classification is based on measurements of adsorption-desorption of N2 in boiling temperature and in the statistical width of layers of N2 adsorbed. The analysis of these data is usually done by using the Brunauer/Emmett/Teller method (BET), proposed by Stephen Brunauer, Paul Hugh Emmett e Edward Teller [57-59]. Applying BET, it is possible to describe the form of the adsorption and desorption isotherms for a specific solid. Knowing the format of the isotherm it is possible to define its porosity [60], Figure 2.

Microporous solids usually present a type I isotherm, whereas the isotherms II and III are related to non-porous solids finely divided or macroporous solids. Already the isotherms type IV and V present a hysteresis loop, a characteristic of mesoporous materials. The type IV hysteresis represents materials with uniform porosity, while type V hysteresis is referred to pores with non-defined forms and sizes. Finally, the type VI hysteresis is related to non-porous solids with almost uniform surface [60, 61].

Advertisement

3. Mesoporous nanostructured TiO2

Organized mesoporous nanostructured titanium dioxide (TiO2) is one of the most promising semiconductor materials, considering its current technological applications [19-21, 44, 60, 62-70]. The presence of pores in nanostructured materials tends to enhance their physical and chemical properties. Compared with bulk TiO2, the uniform channels in mesoporous TiO2 do not only increase the density of active sites with high accessibility but also facilitate the diffusion of reactants and products; the high surface area and large pore volume provides enhanced capability for dye loading and pollutant adsorption [60]. It is also possible to produce mesoporous TiO2 using template-free methods. However, these materials usually present disordered mesostructures as a result of the irregular packing of building blocks [60].

Advertisement

4. Surfactants as drive agents in obtaining mesoporous ordered materials

Certain surfactants are capable to act as drive agents, as structural templates, in the synthesis of a semiconductor oxide, promoting its organization. This process occurs due to the spontaneous organization of the critical nuclei of the oxide formed around the structures defined by the surfactant, which control the growth of the microstructures initially formed, resulting in ordered mesoporous materials. The structural control and manipulation of the physical characteristics of a mesoporous material depends on the nature and concentration of the surfactant employed in the synthesis process [21, 71].

For a well-succeeded action of the directing agent in the synthesis, the model and the inorganic species need to interact [71]:

  1. S+I-: cationic surfactants are used as structural drivers for inorganic species negatively charged;

  2. S-I+: anionic surfactants are used as structural drivers for positively charged inorganic species;

  3. S+X-I+: the surfactant and inorganic species are positively charged in the presence of species of opposite charge acting as counter ion;

  4. S-X+I-: the surfactant and inorganic species are negatively charged in the presence of a species of opposite charge, which acts as a counter ion;

  5. Sº Iº: the surfactant has no load and the medium favors the neutrality of the inorganic species. In this case, the interaction between the species occurs through hydrogen bonds or dipole.

The concentration of surfactant in the reaction medium must be higher than the critical micelle concentration (CMC), which favors the auto mounting of micelles, originating bi and three-dimensional arrays responsible for empty spaces that will produce the nanoporous material [71].

Advertisement

5. Synthesis of mesoporous TiO2

Among the methodologies of synthesis, the methods of sol-gel and homogeneous precipitation best favor the control of morphology [58, 72, 73]. The term sol-gel involves several synthetic methodologies based on hydrolytic processes [19, 20, 59, 74]. Hydrolysis and condensation reactions involving the precursor allows the formation of colloidal particles (sol) and subsequent formation of three-dimensional networks (gel). This process ensures a good homogeneity for the reaction product, leading, in general, to metastable phases, amorphous or not [50]. In the case of amorphous materials, a step of crystallization for oxides formed is required.

The use of solvothermal conditions is an alternative for conversion of the amorphous material to crystalline without the use of high temperatures [60, 73]. The solvothermal method consists in the dissolution of a metallic precursor in anhydrous benzyl alcohol and heating at temperatures above 423K. Under these conditions, the alcoholic hydroxyl group connects partially to the metal ion starting a polycondensation reaction. The steric hindrance, due to the greater volume of the organic group, tends to control the formation of nanoparticles which tend to stabilize in smaller sizes and narrower size dispersions than the obtained in the sol-gel process [75].

The hydrothermal method is a particularity of the solvothermal method in which the solvent is water. Under hydrothermal conditions the solubility of the amorphous particles increases significantly and the crystallization can occur concurrently with processes of re-dissolution and re-precipitation in the crystalline core [50, 76, 77]. This method has been widely used, since small variations in the main synthesis parameters (pH, concentration of precursor, among others) can cause significant changes in morphology, size, chemical constitution, between other properties of the synthesized nanostructures.

The homogeneous precipitation method has also been studied enough due to its great versatility. In this case, chlorinated precursors of titanium or even titanium tetraisopropoxide react with a base (sodium or ammonium hydroxide) forming titanium hydroxide. This compound is converted through the loss of water in amorphous TiO2, which, under different thermal treatments, gives rise to the three known crystalline forms [78].

The combination between these strategies of synthesis and the use of molecular templates warrants a better control in obtaining ordered mesoporous materials. In this sense, two approaches can be highlighted: the soft template and hard template methods [60].

Figure 3.

Representative scheme of the Soft-templating (A) and Hard-templating (B) methods, used in the synthesis of mesoporosos materials.

The soft-templating method consists in the use of a template in the liquid phase, being widely used in the sol-gel processes based on titanium precursors. Due to its high chemical reactivity these precursors tend to suffer hydrolysis, giving rise to dense precipitates very quickly. The self-structuration of surfactants can produce mesostructures at the sol particle. This occurs by the spontaneous organization of molecules around automounted micelle structures, usually in aqueous medium, giving rise to colloidal nanocrystals.

Studies have suggested that two mechanisms are related to the formation of mesoporous materials via soft-templating method [71, 79-81]. The first, known as Cooperative Self-Assembly (CSA), and the second Liquid Crystal Template (LCT). In CSA, there is a simultaneous aggregation between the inorganic precursor and the surfactant, with the formation of a liquid crystal-like phase with cubic, hexagonal or lamellar arrangement, containing both micelles and the inorganic precursor. On the other hand, in the LCT mechanism the concentration of surfactant is so high that a liquid-crystalline phase is formed without the presence of inorganic precursors. In both mechanisms, after formation of the material is necessary to remove the template for the obtaining of porosity. For this, the procedure most commonly used is the calcination [71].

Figure 4.

Formation of mesoporous structures: (A) Cooperative self-assembly process (CSA), and (B) Liquid–crystal templating process (LCT).

Chiola, Ritsko and coworkers, reported in 1971 the obtaining of mesoporous silica [82]. However, the special characteristics of the produced materials did not were at first explained because the absence of adequate methods of analysis. In 1990, scientists from Mobil Oil Corporation [83], reported obtaining of mesoporous silicate molecular sieves, denoted by M41s, synthesized using a supramolecule-templating process, giving rise to a new and fascinating research field. Subsequently, efforts have been concentrated in studies related to synthesis and applications of mesoporous materials [9, 72, 84-91].

In 1995, the preparation of mesoporous TiO2 using sol-gel method associated with a surfactant as structural template were at first reported [92]. Since then, efforts have been spent on the proposition of methods of synthesis of ordered mesoporous TiO2, aiming to contribute to the improvement of the photocatalytic activity of this material, aiming its different applications [93-96].

Nanometric structures of mesoporous TiO2 were produced using the sol-gel methodology, under acidic (pH 3 to 4) and alkaline (pH 7 to 9) media, using cetyl trimethyl ammonium bromide (CTAB), sodium dodecyl benzenesulphonate (DBS) and nonylphenol ethoxylate as surfactants (NPE) [72]. The effect of the surfactant and pH on morphology, particle size, surface area, and pore size distribution was studied being the synthesized oxides evaluated considering the degradation of naphthalene. Using CTAB under acidic pH the nanometric material is 100% anatase, while under alkaline pH the material was produced in two crystalline phases, anatase (84.7%) and brookite (15.3%). In the presence of DBS, under acidic pH, 83% of the crystalline phase was anatase and 17% rutile, while under alkaline pH, 100% was anatase. However, using NPE a mixture 94.7% anatase and 5.3% rutile was produced under acidic pH, while under alkaline pH the phases are anatase and brookite (8.28%).The authors suggested that the good result obtained for the degradation of naphthalene (97% up to 4 h under irradiation in the visible - λ > 400 nm) using the mesoporous material obtained in the presence of CTAB in acidic medium is due to the surface area (86.7 cm3 g-1) and to the mean pore volume (0.0569 cm3 g-1).

A variant of the Soft Templating method is the Evaporation Induced Self Assembly (EISA) [97, 98]. It was originally designed for the preparation of fine films of mesostructured silica [79], and has been successful in obtaining different mesoporous materials (TiO2, ZrO2, Al2O3, Nb2O5, Ta2O5, WO3, HfO2, SnO2, etc) [79, 80]. In this method, organic solvents such as ethanol, propanol, etc, are used as reaction medium. By virtue of the use of non-aqueous medium and block copolymer surfactants, the rate of hydrolysis well as redox reactions and phase transformation are minimized. EISA presents advantages when compared to aqueous sol-gel reaction, such as more control and homogeneity in the formation of particles, high crystallinity in reaction temperatures between 373 and 573K.

Figure 5.

Scheme presenting the most representative steps of Evaporation Induced Self-Assembly (EISA).

A methodology capable of producing monodisperse spheres of TiO2 at room temperature, using titanium (IV) butoxide (TBT) mixed to ethylene glycol, used as template, was proposed in [99]. The mixture was subsequently subjected to a conventional hydrothermal treatment. Thus, it was possible to obtain nanoparticles with surface area up to 233.2 m2 g-1 and mean pore volume of 0.27 cm3 g-1. The photocatalytic activity of the synthesized material was evaluated by degradation of methylene blue, Red MX-5B and phenolphthalein under UV irradiation, when were obtained kinetic constants of degradation of 6.7, 13.2, and 14.7 w% min-1, respectively, attributed to the greater surface area of the catalyst. The synthesis of TiO2 spheres based on EISA, using titanium (IV) tetra isopropoxide and Triton X-100 has been also reported, aiming their application in DSSCs [100]. The particle size could be controlled changing the Ti/H2O molar ratio and adjusted between 500 and 1100 nm, with a surface area of 56-151 cm2 g-1.

Using the EISA strategy excellent catalysts based on CeO2, capable to oxidize harmful gases, as CO, NOx and hydrocarbons, in non-toxic gases, were synthesized [81]. The tribloc copolymer P123 and polyvinyl pyrrolidone (PVP) were used as templates. In this way, were obtained catalysts with different pore sizes (about 40 μm, 2 μm and < 0.3 μm) and specific surface area of 32.5 m2 g-1.

Damages to the mesoporous structure can occur due to the presence of residual templates, as for example the blocking of pores or even poisoning of catalytic sites on the surface of the mesoporous material. Calcination has been pointed as a way to promote the elimination of some surfactants [101]. Non-ionic surfactants, as Triton X-100, can be easily removed by washing of the oxide formed [102].

In the hard-templating method the solid is formed around a template, usually zeolites and meso and macroporous silica, via interactions between the template and the precursor [60]. This methodology is more complex when compared to the soft-templating method. However, it is highly effective in obtaining mesoporous structures [103]. By applying this methodology nanostructures are formed on surface of the templates [104].

The obtaining of uniform nanospheres of TiO2 based on hard-templating methodology, using silica as template and (NH4)2TiF6 as precursor, has been proposed [105]. This material presented surface area of 54.2 m2 g-1 and pore volume of 0.15 cm3 g-1. The photocatalytic activity of this material was evaluated regarding the degradation of the dyes methylene blue, rhodamine B and methyl-orange. The levels of degradation were respectively 95%, 95% and 90%. This result was attributed to the high surface area and charge separation in these monodisperse materials.

In reference [106] the obtaining of mesoporous TiO2 is proposed using the combination of hard-and soft-tamplating methods. Using mesoporous silica in the first step (hard-templating) and Pluronic P123 as driver agent in the soft-templating stage, to control of morphology of the particles, was possible the production of three-dimensionally ordered mesopores with surface area of 145 m2 g-1 and pore volume of 0.246 cm3 g-1. Due to the porous of the material obtained and the three-dimensional periodicity of the spheres, the authors highlight the possible applications that these materials can have, as in environmental photocatalysis, solar cells and in lithium rechargeable batteries. Finally, the authors highlight the ease with which this methodology can be extended for the synthesis of other mesoporous metal oxides.

Figure 6 presents the main differences between the hard and soft template methods, while Table 1 presents the structure of the main surfactants suggested as driver agents for the synthesis of mesoporous materials.

Figure 6.

Comparison between hard and soft templating [107].

Surfactant Structure References
Polyethylene glycol [108, 109]
Triton X-100 [109-112]
Ethyl cellulose [77, 113, 114]
1-hexadecyl [115]
Acetylacetone [115]
Isopropanol [115]
Sodium dodecyl sulfate [116]
Cetyltrimethyl
ammonium bromide
[5, 110, 117, 118]
Cetyltrimethyl
ammonium chloride
[5]
Hexamethylene
tetramine
[118]
Benzalkonium chloride [5]
Polyvinyl
pyrrolidone
[118]
Pluronic P123 [62]

Table 1.

Main surfactants suggested as driver agents for the synthesis of mesoporous materials and their chemical structures.

Advertisement

6. Technological applications of mesoporous TiO2

6.1. Environmental photocatalysis

One of the potential applications of heterogeneous photocatalysis is the environmental remediation [8, 20, 74, 119-122]. Different semiconductor oxides have shown to be capable of triggering photocatalytic, highlighting the TiO2 [19, 20, 123]. In this context, nanostructured materials tend to favor the photocatalytic applications of these materials [60].

Kim and Kwak reported studies where the photocatalytic activity of a mesoporous TiO2 prepared using di-block copolymers via sol-gel synthesis was evaluated, using methylene blue as oxidizable substrate [50]. The materials with small crystallite size (about 5.1 nm), high surface area (about 210 m2 g-1) and small pore size distribution were the ones who exhibited the best photocatalytic activities, using ultraviolet radiation, with apparent constants of degradation around 0.093 min-1. The particles synthesized presented spherical morphology with surface areas dependent on the type of di-block copolymer used.

In [58] is related a sol-gel synthesis of a TiO2 with intermediate porosity, under hydrothermal conditions using a water-soluble cationic surfactant, obtaining a mesoporous TiO2 with maximum pore size of 6.9 nm and maximum surface area of 284 m2 g-1. The synthesized material had its photocatalytic activity evaluated on degradation of methylene blue. The synthesized product showed to be superior (95%) to TiO2 P25 (24%), using irradiation with UV radiation under the same reaction conditions. The greater photocatalytic efficiency was attributed to the higher surface area, when compared to this parameter measured to TiO2 P25 (50 m2 g-1), and the ordination of the synthesized material, attributed to the use of the surfactant.

The synthesis of a TiO2 supported on SBA-15 mesoporous silica, reported in [124], resulted in materials efficient to degrade photocatalytic phenols. SBA-15 silica presents hexagonal arrangement and bi-dimensional pores with diameter up to 30 nm. The materials obtained have surface area of 361.4 m2 g-1 and pore volume of 1.217 cm3 g-1.

6.2. Hydrogen production

Hydrogen production via heterogeneous photocatalysis based in the electronic excitation of a semiconductor material using UV-Visible radiation is promising for the production of clean and sustainable energy [19, 54, 125]. Generally speaking, this process is based on the electronic excitation of the photocatalyst. From there, the photo-generated hole at the valence band reacts readily with a sacrificial reagent, while in conduction band electrons are trapped by a co-catalyst, being used in the reduction of H+ ions, with the production of hydrogen [126].

Mesoporous TiO2 of intermediate porosity have generally high surface area and high density of active sites, which tends to facilitate the diffusion of reagents, favoring the conversion of solar energy in photocatalytic reactions [41, 93, 127].

A porous composite based on TiO2 incorporated to silica MCM-48 (Mobil Crystalline Materials n° 48) was proposed as photocatalyst for hydrogen production [128]. The authors reported that the tetrahedrically coordinated TiO2 act as active sites on photocatalysis of water reducing the potential, facilitating the formation of gaseous hydrogen, even in the absence of a co-catalyst.

Microspheres of TiO2 doped with carbonates has also been proposed for hydrogen production under visible (λ > 400 nm) irradiation [129]. A rate of H2 production of 0.2 mmol h-1 g-1 was achieved using the photocatalyst charged with 1 wt% Pt. The photocatalytic performance was attributed to the characteristics of the mesoporous structure such as the diameter of the microspheres, between 0.5 and 4 μm, pore size in the range between 3 and 11 nm and high surface area (500 m2 g-1).

A mesoporous TiO2-SiO2 mixed oxide with molar ratio of 97:3, calcined at 773 K was also proposed for hydrogen production [66]. It produced hydrogen at a rate of 0.27 cm3 h-1 g-1, an expressive value when compared to the rate achieved using TiO2 P25 (0.17 cm3 h-1 g-1). This mixed oxide has characteristics that favored its photocatalytic activity, such as a surface area of 162 m2 g-1, pore diameter of 4.3 nm and pore volume of 0.24 cm3 g-1.

6.3. Electrodes in Lithium ion batteries

Mesoporosos materials have been used in the production of electrodes for Lithium ion batteries. Physical properties such as particle size, porosity and pore size determine the performance of this class of batteries [127]. The mesoporous structure favors the transport of electrolyte, facilitating the contact between the surface of the electrode and the electrolyte and shortening the path of diffusion of the Li+ ions.

These batteries have a long life cycle, rapid loading and unloading capacity, being capable to store high energy densities in a compact and lightweight container. Because of this, it has been used as power sources for portable electronic devices [127, 130, 131].

Mesoporous TiO2 is a promising material in this kind of battery by offering numerous advantages, such as high potential of ions insertion, low cost, low toxicity, easy synthesis and stability to pH variation [60]. Anatase is generally regarded as the most suitable polymorph of TiO2 for Li+ insertion, superior to rutile and with more stable structure than brookite. However, anatase with large particle sizes present poor performance due to its low capabilities to promote ionic diffusion and electronic conductivity [132, 133]. Thus, the control of particle size and porosity during its synthesis is imperative for obtaining materials with desirable characteristics.

The particle size influence on the dynamic process of storage of Li+ions, ensuring a significant increase on battery performance when the particle size is reduced. For example, the discharge capacity of TiO2 rutile with particles of 300 nm is 110 mAh g−1, while this same parameter for particle sizes of 30 and 15 nm is, respectively, 338 and 378 mAh g−1 [134].

Studies involving the application of mesoporous microspheres of anatase with regular porosity showed that pore structure significantly influences on specific capacity, speed capability and performance cycle of the batteries [60]. The materials with higher surface area had the best performance. It was also observed that when the pore size was very small, the transport of Li+ in the electrolyte was heavily restricted.

6.4. Solar cells

Solar cells have a great advantage when compared to conventional power generation systems, since in these solar energy can be directly converted into electric. The traditional solar cells are basically constituted by two layers of semiconductor materials, an n-p type pair. When photons with appropriate energy reach the semiconductor pair, electrons are excited producing electricity [135].

The first solar cells were based on crystalline silicon. Actually, studies are being conducted with the goal of developing efficient photovoltaic cells and low cost, since the crystalline silicon-based cells are of very high cost.

The solar cells can be classified as:

  1. First generation solar cells, where are the silicon solar cells used commercially, which today correspond to about 86% of the solar cells market [136];

  2. Second generation solar cells. More profitable than the first generation solar cells, but still less efficient;

  3. Third generation solar cells, where are the dye-sensitized organic cells;

  4. Hybrid solar cells [137].

In the organic solar cells are used semiconductor organic polymers or small and medium organic molecules, such as phthalocyanines, fullerenes, poly-(p-phenylenevinylene) (PPV). These cells can be produced on flexible substrates, and are promising in terms of production costs. On the other hand, they still have low efficiency, as well as some limitations, such as the possibility of degradation of the organic component.

In 2009, Park and co-workers developed an organic solar cell based on a polymer-fullerene composite, with a reasonable energetic efficiency (6.5%) [138].

Dye-sensitized solar cells, DSCs, Figure 7, are devices constituted by a semiconductor material, a sensitizer (dye), a conductive glass, an electrolyte and a platinum counter-electrode. Unlike conventional systems, where semiconductor assumes both the task of absorbing light as the charge carriage, in DSCs such functions are separated: the light is absorbed by a dye anchored to the surface of a semiconductor, and the charge separation occurs at the semiconductor-electrolyte interface [139].

Figure 7.

Simplified diagram showing the functioning of DSCs.

A great leap in this area was given by [140] by using mesoporous nanocristalline TiO2 films in place of single crystals of oxide semiconductor as substrates for the adsorption of the sensitizer dye. The use of mesoporous films resulted in a significant increase in the conversion efficiency of incident light in current, IPCE, which went from 0.13 to 88% [141]. The principal reason for this was the increase in the amount of dye adsorbed due to the greater surface area in general available in the mesoporous films. Using mesoporous films, the overall efficiency of conversion of solar cells sensitized by dye reached 7% in the decade of 90 [140]. Actually, the best cells exceed the 13% [142, 143], confirming the feasibility of commercial applications of these devices.

The I-/I3-redox pair dissolved in nitriles is thermodynamically capable to reduce the photo oxidized dye on the surface of TiO2. The difference between the energy of the semiconductor conduction band and the redox potential of the electrolyte determines the maximum open-circuit potential than a DSC can achieve. Additionally, the kinetics of reduction is one or two orders of magnitude faster than the process of recombination between the electron in the conduction band of the oxide and the photo oxidized dye [144]. This difference makes the charge separation process in the TiO2-sensitizer-electrolyte interface to be very efficient. The disadvantages of using liquid electrolytes based on I-/I3-redox pair include the toxicity of species of iodine and the inner filter effect caused by the strong absorption of light in the visible region by I3-, which makes DSCs efficiency dependent on the exposed face.

The system TiO2/Ru(II) complex/(I-/I3-) is relatively simple and efficient, allowing a considerable advance in the understanding of the charge separation process from absorption of sunlight in sensitized semiconductors. However, this system presents challenges to be overcome, to its large-scale implementation. There stands out the high cost of the Ru(II) complexes, the high toxicity of the electrolyte and the and the need for an efficient sealing of the device in order to prevent the leakage of the electrolyte and the consequent loss of activity. Even so, modules up to 6000 cm2 have been produced [145].

Alternatively, several other materials have been proposed to build DSCs. Organic compounds, such as indole and anthocyanins derivatives also have been proposed as sensitizers [146-150]. Recently, [151] described DSCs possessing efficiencies higher than 12% using a Zn(II) porphyrine as sensitizer and Co(II)/(III) complexes as redox pair. Another approach that has attracted much attention involves the use of solid systems that, at first, dispense step of sealing and increase solar cell stability in real conditions of operation [152-155].

A major breakthrough in the development of solid DSCs has been achieved with the use of perovskites, as the CH3NH3PbI3, as sensitizers [156-158]. These compounds are chemically very stable and can be produced in situ from precursors solutions and have a wide range of absorption, from visible to near infrared. When adsorbed on surface of TiO2 films, CH3NH3PbI3 is capable of injecting electrons efficiently and be regenerated by hole transporter materials (HTM) as spiro-MeOTAD (2,2',7,7'-tetraquis [N,N-di(4-metoxyphenyl)amino]-9,9'-spirobifluorene). DSCs with efficiencies higher than 14% have been described using this approach [156]. It is interesting that the use of this same material in DSCs with Ru(II) complexes results in conversion efficiencies between 2 and 3% [159, 160]. This makes clear that it is not enough to simply replace one or another component in the DSCs to achieve greater efficiencies. It is necessary a good understanding of the characteristics of each part of the interfaces of the device and of the different processes of electron transfer that occur between the components.

Figure 8.

Decrease of electronic recombination on the FTO/electrolyte interface due to the application of the blocking layer.

In all approaches described above, it is common the need to deposit multiple layers of semiconductor oxides with different functions. Besides the nanoporous layer, responsible for dye adsorption and transport of electrons, there is the blocking layer deposited on the surface of the substrate which is responsible for reduction of the electronic recombination on the substrate-electrolyte interface, Figure 8.

Additionally, scattering layers, composed of particles with size between 400 and 700 nm can be deposited, which contributes to a greater use of the incident light. Several proposals can be found in the literature in relation to different layers of metal oxides in DSCs. In these are included the use of different techniques of deposition and also various materials. The research groups working in this area agree that the deposition of different semiconductor oxide layers with different morphologies is indispensable for production of high efficiency DSCs [143, 161]. However, there is still no consensus on the best deposition technique to be employed and what better composition to be used for the different electrolytes.

Recently, we reported the production of contact-blocking layers using the layer-by-layer (LbL) technique, which showed excellent performance in DSCs based on liquid electrolytes [162-164]. This technique stands out for its low cost, possibility of control of composition, thickness and morphology of films and the possibility to be employed on a large scale. It was noted, for example, that the use of a mixture between TiO2 and more insulating oxides as Nb2O5 results in an increase of all the photoelectrochemical parameters of the DSCs [165].

Advertisement

7. Conclusion

In this chapter, aspects related to obtaining and application of mesoporous nanostructured materials in photocatalytic processes had been addressed, emphasizing its application in advanced oxidative processes, increasing the overall efficiency of conversion in dye sensitized solar cells, manufacturing of electrodes for lithium-ion batteries and hydrogen production, with focus on TiO2. Although many advances have occurred, some challenges still needed to be overcome so that these materials become more efficient and economically viable. On the other hand, the potential that mesoporous materials demonstrated and the improvements already achieved promote the study and development of what is a promising source of technological applications.

Advertisement

Acknowledgments

To Fundação de Amparo à Pesquisa do Estado de Minas Gerais (FAPEMIG), Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES), and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Brazilian agencies for research funding and grants, by the support and funding.

References

  1. 1. Bahng SH, Kwon NH, Kim HC, Siddique A, Kang HJ, Lee JY, et al. Simple Synthesis of Water-Dispersible and Photoactive Titanium Dioxide Nanoparticles Using Functionalized Poly(ethylene oxide)s. Macromolecular Research 2014;22(4) 445-456. DOI: 10.1007/s13233-014-2062-5.
  2. 2. Hidalgo D, Messina R, Sacco A, Manfredi D, Vankova S, Garrone E, et al. Thick mesoporous TiO2 films through a sol–gel method involving a non-ionic surfactant: Characterization and enhanced performance for water photo-electrolysis. International Journal of Hydrogen Energy 2014. DOI: 10.1016/j.ijhydene.2014.02.163.
  3. 3. Sanchez-Quiles D, Tovar-Sanchez A. Sunscreens as a source of hydrogen peroxide production in coastal waters. Environmental Science & Technology 2014;48(16) 9037-9042. DOI: 10.1021/Es5020696.
  4. 4. Xi BJ, Chu XN, Hu JY, Bhatia CS, Danner AJ, Yang H. Preparation of Ag/TiO2/SiO2 films via photo-assisted deposition and adsorptive self-assembly for catalytic bactericidal application. Applied Surface Science 2014;311 582-592. DOI: 10.1016/j.apsusc.2014.05.112.
  5. 5. Yeh SW, Ko HH, Chiang HM, Chen YL, Lee JH, Wen CM, et al. Characteristics and properties of a novel in situ method of synthesizing mesoporous TiO2 nanopowders by a simple coprecipitation process without adding surfactant. Journal of Alloys and Compounds 2014;613 107-116. DOI: 10.1016/j.jalicom.2014.05.227.
  6. 6. Wang ZY, Yao N, Hu X. Single material TiO2 double layers antireflection coating with photocatalytic property prepared by magnetron sputtering technique. Vacuum 2014;108 20-26. DOI: 10.1016/j.vacuum.2014.05.009.
  7. 7. Chen KC, Wang YH. The effects of Fe-Mn oxide and TiO2/alpha-Al2O3 on the formation of disinfection by-products in catalytic ozonation. Chemical Engineering Journal 2014;253 84-92. DOI: 10.1016/j.cej.2014.04.111.
  8. 8. A. Oliveira, E. M. Saggioro, T. Pavesi, J. C. Moreira, L. F. V. Ferreira. Solar photochemistry for environmental remediation-advanced oxidation processes for industrial wastewater treatment. In: Saha S (ed.) Molecular Photochemistry-Various Aspects. Rijeka: InTech; 2012. p. 195-223.
  9. 9. Wu MM, Shen Y, Gu F, Xie YA, Zhang JC, Wang LJ. Preparation and photoelectric properties of mesoporous ZnO films. Journal of Sol-Gel Science and Technology 2010;53(2) 470-474. DOI: 10.1007/s10971-009-2099-7.
  10. 10. Xiong XH, Wang ZX, Wu FX, Li XH, Guo HJ. Preparation of TiO2 from ilmenite using sulfuric acid decomposition of the titania residue combined with separation of Fe3+with EDTA during hydrolysis. Advanced Powder Technology 2013;24(1) 60-67. DOI: 10.1016/j.apt.2012.02.002.
  11. 11. Sasikumar C, Rao DS, Srikanth S, Ravikumar B, Mukhopadhyay NK, Mehrotra SP. Effect of mechanical activation on the kinetics of sulfuric acid leaching of beach sand ilmenite from Orissa, India. Hydrometallurgy 2004;75(1-4) 189-204. DOI: 10.1016/j.hydromet.2004.08.001.
  12. 12. El-Hazek N, Lasheen TA, El-Sheikh R, Zaki SA. Hydrometallurgical criteria for TiO2 leaching from Rosetta ilmenite by hydrochloric acid. Hydrometallurgy 2007;87(1-2) 45-50. DOI: 10.1016/j.hydromet.2007.01.003.
  13. 13. Fujishima A, Honda K. Eletrochemical photolysis of water at semiconductor electrode. Nature 1972(238) 37-38.
  14. 14. He LF, Wang CD, Yao XL, Ma RG, Wang HK, Chen PR, et al. Synthesis of carbon nanotube/mesoporous TiO2 coaxial nanocables with enhanced lithium ion battery performance. Carbon 2014;75 345-352. DOI: 10.1016/j.carbon.2014.04.013.
  15. 15. Hu LY, Wang JY, Zhang JX, Zhang QY, Liu ZH. An N-doped anatase/rutile TiO2 hybrid from low-temperature direct nitridization: enhanced photoactivity under UV-/visible-light. Rsc Advances 2014;4(1) 420-427. DOI: 10.1039/C3ra44421j.
  16. 16. Lan X, Wang LZ, Zhang BY, Tian BZ, Zhang JL. Preparation of lanthanum and boron co-doped TiO2 by modified sol-gel method and study their photocatalytic activity. Catalysis Today 2014;224 163-170. DOI: 10.1016/j.cattod.2013.10.062.
  17. 17. Liu GQ, Liu LC, Song JR, Liang JD, Luo QZ, Wang DS. Visible light photocatalytic activity of TiO2 nanoparticles hybridized by conjugated derivative of polybutadiene. Superlattices and Microstructures 2014;69 164-174. DOI: 10.1016/j.spmi.2014.02.013.
  18. 18. Neubert S, Ramakrishnan A, Strunk J, Shi HY, Mei B, Wang LD, et al. Surface-Modified TiO2 Photocatalysts Prepared by a Photosynthetic Route: Mechanism, Enhancement, and Limits. Chempluschem 2014;79(1) 163-170. DOI: 10.1002/cplu.201300277.
  19. 19. A. E. H. Machado, A. O. T. Patrocinio, M. D. França, L. M. Santos, K. A. Borges, L. F. Paula. Metal oxides for photoinduced hydrogen production and dye-sensitized solar cell applications. In: Méndez-Vilas A. (ed.) Materials and processes for energy: communicating current research and technological developments. Badajoz: Formatex; 2013. p. 867-879.
  20. 20. A. E. H. Machado, L. M. Santos, K. A. Borges, P. S. Batista, V. A. B. Paiva, P. S. Müller Jr., et al. Potential Applications for Solar Photocatalysis: From Environmental Remediation to Energy Conversion. In: Babatunde E. B. (ed.) Solar Radiation. Rijeka: InTech; 2012. p. 339-378.
  21. 21. F. Sordello, V. Maurino, C. Minero. Improved photochemistry of TiO2 inverse opals and some examples. In: Saha S (ed.) Molecular Photochemistry-Various Aspects. Rijeka: InTech; 2012. p. 63-87.
  22. 22. M. Valenzuela. Photocatalytic deposition of metal oxides on semiconductor particles: a review. In: Saha S (ed.) Molecular Photochemistry-Various Aspects. Rijeka: InTech; 2012. p. 25-41.
  23. 23. Jaroenworaluck A, Pijarn N, Kosachan N, Stevens R. Nanocomposite TiO2-SiO2 gel for UV absorption. Chemical Engineering Journal 2012;181 45-55. DOI: 10.1016/j.cej.2011.08.028.
  24. 24. Affam AC, Chaudhuri M. Degradation of pesticides chlorpyrifos, cypermethrin and chlorothalonil in aqueous solution by TiO2 photocatalysis Journal of Environmental Management 2013;130 160-165. DOI: 10.1016/j.jenvman.2013.08.058.
  25. 25. Chen JW, Shi JW, Wang X, Ai HY, Cui HJ, Fu ML. Hybrid metal oxides quantum dots/TiO2 block composites: Facile synthesis and photocatalysis application. Powder Technology 2013;246 108-116. DOI: 10.1016/j.powtec.2013.05.014.
  26. 26. Hirakawa T, Sato K, Komano A, Kishi S, Nishimoto CK, Mera N, et al. Specific properties on TiO2 photocatalysis to decompose isopropyl methylphosphonofluoridate and dimethyl methylphosphonate in Gas Phase. Journal of Photochemistry and Photobiology a-Chemistry 2013;264 12-17. DOI: 10.1016/j.jphotochem.2013.04.012.
  27. 27. A. Eremenko, N. Smirnova, I. Gnatiuk, O. Linnik, N. Vityuk, Y. Mukha, et al. Silver and gold nanoparticles on sol-gel TiO2, ZrO2, SiO2 surfaces: optical spectra, photocatalytic activity, bactericide properties. In: Cuppoletti J (ed.) Nanocomposites and Polymers with Analytical Methods. Rijeka: InTech; 2011. p. 51-83.
  28. 28. V. S. Viteri, E. Fuentes. Titanium and Titanium Alloys as Biomaterials. In: Gegner J (ed.) Tribology-Fundamentals and Advancements. Rijeka: InTech; 2013. p. 155-181.
  29. 29. Grcic I, Vujevic D, Zizek K, Koprivanac N. Treatment of organic pollutants in water using TiO2 powders: photocatalysis versus sonocatalysis. Reaction Kinetics Mechanisms and Catalysis 2013;109(2) 335-354. DOI: 10.1007/s11144-013-0562-5.
  30. 30. Machado AEH, França MD, Velani V, Magnino GA, Velani HMM, Freitas FS, et al. Characterization and evaluation of the efficiency of TiO2/zinc phthalocyanine nanocomposites as photocatalysts for wastewater treatment using solar irradiation. International Journal of Photoenergy 2008;2008. DOI: 10.1155/2008/482373.
  31. 31. Fujishima A, Honda K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972;238(5358) 37-+. DOI: 10.1038/238037a0.
  32. 32. Gao P, Sun DD. Hierarchical sulfonated graphene oxide-TiO2 composites for highly efficient hydrogen production with a wide pH range. Applied Catalysis B-Environmental 2014;147 888-896. DOI: 10.1016/j.apcatb.2013.10.025.
  33. 33. Melian EP, Suarez MN, Jardiel T, Rodriguez JMD, Caballero AC, Arana J, et al. Influence of nickel in the hydrogen production activity of TiO2. Applied Catalysis B-Environmental 2014;152 192-201. DOI: 10.1016/j.apcatb.2014.01.039.
  34. 34. Mendez JAO, Lopez CR, Melian EP, Diaz OG, Rodriguez JMD, Hevia DF, et al. Production of hydrogen by water photo-splitting over commercial and synthesised Au/TiO2 catalysts. Applied Catalysis B-Environmental 2014;147 439-452. DOI: 10.1016/j.apcatb.2013.09.029.
  35. 35. Taboada E, Angurell I, Llorca J. Dynamic photocatalytic hydrogen production from ethanol-water mixtures in an optical fiber honeycomb reactor loaded with Au/TiO2. Journal of Catalysis 2014;309 460-467. DOI: 10.1016/j.jcat.2013.10.025.
  36. 36. Wu ZF, Zhang WH, Xiong F, Yuan Q, Jin YK, Yang JL, et al. Active hydrogen species on TiO2 for photocatalytic H2 production. Physical Chemistry Chemical Physics 2014;16(15) 7051-7057. DOI: 10.1039/C4cp00697f.
  37. 37. Hagfeldt A, Gratzel M. Light-Induced Redox Reactions in Nanocrystalline Systems. Chemical Reviews 1995;9549-68.
  38. 38. Hagfeldt A, Gratzel M. Molecular photovoltaics. Accounts Chemistry Reserarch 2000(33) 269-277.
  39. 39. Jayaweera PVV, Perera AGU, Tennakone K. Why Gratzel`s cells works so well. Inorganica Chimica Acta 2008(361) 707-711.
  40. 40. Peter LM. Dye-sensitized nanocristalline solar cells. Physical Chemistry Chemical Physics 2007;9 2630-2642.
  41. 41. Chang HM, Yang YJ, Li HC, Hsu CC, Cheng IC, Chen JZ. Preparation of nanoporous TiO2 films for DSSC application by a rapid atmospheric pressure plasma jet sintering process. Journal of Power Sources 2013;234 16-22. DOI: 10.1016/j.jpowsour.2013.01.113.
  42. 42. Sannino D, Vaiano V, Ciambelli P, Zama I, Gorni G. Evaluation of N719 amount in TiO2 films for DSSC by thermogravimetric analysis. Journal of Thermal Analysis and Calorimetry 2013;111(1) 453-458. DOI: 10.1007/s10973-012-2436-x.
  43. 43. Zeng Z, Zhang B, Liu XJ, Peng X, Meng SX, Feng YQ. The impact of porphyrin meso-position steric effect on porphyrin density on TiO2 surface of DSSC. Advanced Materials and Engineering Materials Ii 2013;683 522-525. DOI: 10.4028/www.scientific.net/AMR.683.522.
  44. 44. Paula LF, Amaral RC, Iha NYM, Paniago RM, Machado AEH, Patrocinio AOT. New layer-by-layer Nb2O5-TiO2 film as an effective underlayer in dye-sensitised solar cells. Rsc Advances 2014;4(20) 10310-10316. DOI: 10.1039/C4ra00058g.
  45. 45. Zhu XD, Wang YJ, Zhou DM. TiO2 photocatalytic degradation of tetracycline as affected by a series of environmental factors. Journal of Soils and Sediments 2014;14(8) 1350-1358. DOI: 10.1007/s11368-014-0883-7.
  46. 46. Kumar N, Maitra U, Hegde VI, Waghmare UV, Sundaresan A, Rao CNR. Synthesis, Characterization, Photocatalysis, and Varied Properties of TiO2 Cosubstituted with Nitrogen and Fluorine. Inorganic Chemistry 2013;52(18) 10512-10519. DOI: 10.1021/Ic401426q.
  47. 47. Gupta SK, Singh J, Anbalagan K, Kothari P, Bhatia RR, Mishra PK, et al. Synthesis, phase to phase deposition and characterization of rutile nanocrystalline titanium dioxide (TiO2) thin films. Applied Surface Science 2013;264 737-742. DOI: 10.1016/j.apsusc.2012.10.113.
  48. 48. Idigoras J, Berger T, Anta JA. Modification of Mesoporous TiO2 Films by Electrochemical Doping: Impact on Photoelectrocatalytic and Photovoltaic Performance. Journal of Physical Chemistry C 2013;117(4) 1561-1570. DOI: 10.1021/Jp306954y.
  49. 49. Ji YF, Zhou L, Ferronato C, Salvador A, Yang X, Chovelon JM. Degradation of sunscreen agent 2-phenylbenzimidazole-5-sulfonic acid by TiO2 photocatalysis: Kinetics, photoproducts and comparison to structurally related compounds. Applied Catalysis B-Environmental 2013;140 457-467. DOI: 10.1016/j.apcatb.2013.04.046.
  50. 50. Kim DS, Kwak SY. The hydrothermal synthesis of mesoporous TiO2 with high crystallinity, thermal stability, large surface area, and enhanced photocatalytic activity. Applied Catalysis a-General 2007;323 110-118. DOI: 10.1016/j.apcata.2007.02.010.
  51. 51. Davis ME. Ordered porous materials for emerging applications. Nature 2002;417(6891) 813-821. DOI 10.1038/Nature00785.
  52. 52. Lee ES, Lee KM, Yoon SI, Ko YG, Shin DH. Influence of CNT incorporation on the photovoltaic behavior of TiO2 films formed by high-voltage electrophoretic deposition. Current Applied Physics 2013;13 S26-S29. DOI: 10.1016/j.cap.2013.01.013.
  53. 53. Kondo JN, Domen K. Crystallization of mesoporous metal oxides. Chemistry of Materials 2008;20(3) 835-847. DOI: 10.1021/Cm702176m.
  54. 54. Zheng XL, Kuang Q, Yan KY, Qiu YC, Qiu JH, Yang SH. Mesoporous TiO2 Single Crystals: Facile Shape-, Size-, and Phase-Controlled Growth and Efficient Photocatalytic Performance. Acs Applied Materials & Interfaces 2013;5(21) 11249-11257. DOI: 10.1021/Am403482g.
  55. 55. Ferreira OP, Alves OL, Macedo JD, Gimenez ID, Barreto LS. Ecomaterials: Development and application of functional porous materials for environmental protection. Química Nova 2007;30(2) 464-467. DOI: 10.1590/S0100-40422007000200039.
  56. 56. McCusker LB, Liebau F, Engelhardt G. Nomenclature of structural and compositional characteristics of ordered microporous and mesoporous materials with inorganic hosts-(IUPAC recommendations 2001). Pure and Applied Chemistry 2001;73(2) 381-394. DOI: 10.1351/pac200173020381.
  57. 57. Guo SY, Han S, Chi B, Pu J, Li J. Synthesis of shape-controlled mesoporous titanium phosphate nanocrystals: The hexagonal titanium phosphate with enhanced hydrogen generation from water splitting. International Journal of Hydrogen Energy 2014;39(6) 2446-2453. DOI: 10.1016/j.ijhydene.2013.12.007.
  58. 58. Kao LH, Hsu TC, Cheng KK. Novel synthesis of high-surface-area ordered mesoporous TiO2 with anatase framework for photocatalytic applications. Journal of Colloid and Interface Science 2010;341(2) 359-365. DOI: 10.1016/j.jcis.2009.09.058.
  59. 59. Yang XH, Fu HT, Yu AB, Jiang XC. Large-surface mesoporous TiO2 nanoparticles: Synthesis, growth and photocatalytic performance. Journal of Colloid and Interface Science 2012;38 774-783. DOI 10.1016/j.jcis.2012.06.080.
  60. 60. Li W, Wu ZX, Wang JX, Elzatahry AA, Zhao DY. A Perspective on Mesoporous TiO2 Materials. Chemistry of Materials 2014;26(1) 287-298. DOI: 10.1021/Cm4014859.
  61. 61. Kim DS, Han SJ, Kwak SY. Synthesis and photocatalytic activity of mesoporous TiO2 with the surface area, crystallite size, and pore size. Journal of Colloid and Interface Science 2007;316(1) 85-91. DOI: 10.1016/j.jcis.2007.07.037.
  62. 62. Blin JL, Stebe MJ, Roques-Carmes T. Use of ordered mesoporous titania with semi-crystalline framework as photocatalyst. Colloids and Surfaces a-Physicochemical and Engineering Aspects 2012;407 177-185. DOI: 10.1016/j.colsurfa.2012.05.029.
  63. 63. Crossland EJW, Noel N, Sivaram V, Leijtens T, Alexander-Webber JA, Snaith HJ. Mesoporous TiO2 single crystals delivering enhanced mobility and optoelectronic device performance. Nature 2013;495(7440) 215-219. DOI: 10.1038/Nature11936.
  64. 64. Herrero C, Quaranta A, Leibl W, Rutherford AW, Aukauloo A. Artificial photosynthetic systems. Using light and water to provide electrons and protons for the synthesis of a fuel. Energy & Environmental Science 2011;4(7) 2353-2365. DOI: 10.1039/C0ee00645a.
  65. 65. Kandiel TA, Ivanova I, Bahnemann DW. Long-term investigation of the photocatalytic hydrogen production on platinized TiO2: an isotopic study. Energy & Environmental Science 2014;7(4) 1420-1425. DOI: 10.1039/C3ee41511b.
  66. 66. Rungjaroentawon N, Onsuratoom S, Chavadej S. Hydrogen production from water splitting under visible light irradiation using sensitized mesoporous-assembled TiO2-SiO2 mixed oxide photocatalysts. International Journal of Hydrogen Energy 2012;37(15) 11061-11071.
  67. 67. Taffa DH, Kathiresan M, Arnold T, Walder L, Erbacher M, Bauer D, et al. Dye sensitized membranes within mesoporous TiO2 Photocurrents in aqueous solution. Journal of Photochemistry and Photobiology a-Chemistry 2010;216(1) 35-43. DOI: 10.1016/j.jphotochem.2010.09.003.
  68. 68. Tan HQ, Zhao Z, Niu M, Mao CY, Cao DP, Cheng DJ, et al. A facile and versatile method for preparation of colored TiO2 with enhanced solar-driven photocatalytic activity. Nanoscale 2014;6(17) 10216-10223. DOI: 10.1039/C4nr02677b.
  69. 69. Zhang RY, Elzatahry AA, Al-Deyab SS, Zhao DY. Mesoporous titania: From synthesis to application. Nano Today 2012;7(4) 344-366. DOI: 10.1016/j.nantod.2012.06.012.
  70. 70. Zhou XF, Lu J, Jiang JJ, Li XB, Lu MN, Yuan GT, et al. Simple fabrication of N-doped mesoporous TiO2 nanorods with the enhanced visible light photocatalytic activity. Nanoscale Research Letters 2014;9 DOI: 10.1186/1556-276x-9-34.
  71. 71. Pal N, Bhaumik A. Soft templating strategies for the synthesis of mesoporous materials: Inorganic, organic-inorganic hybrid and purely organic solids. Advances in Colloid and Interface Science 2013;189 21-41. DOI: 10.1016/j.cis.2012.12.002.
  72. 72. Abdel-Azim SM, Aboul-Gheit AK, Ahmed SM, El-Desouki DS, Abdel-Mottaleb MSA. Preparation and Application of Mesoporous Nanotitania Photocatalysts Using Different Templates and pH Media. International Journal of Photoenergy 2014;2014. DOI: 10.1155/2014/687597.
  73. 73. Song HJ, Chen T, Sun YL, Zhang XQ, Jia XH. Controlled synthesis of porous flower-like TiO2 nanostructure with enhanced photocatalytic activity. Ceramics International 2014;40(7) 11015-11022. DOI: 10.1016/j.ceramint.2014.03.108.
  74. 74. Patrocinio AOT, El-Bacha AS, Paniago EB, Paniago RM, Iha NYM. Influence of the Sol-Gel pH Process and Compact Film on the Efficiency of TiO2-Based Dye-Sensitized Solar Cells. International Journal of Photoenergy 2012. DOI: 10.1155/2012/638571.
  75. 75. Niederberger M, Garnweitner G, Buha J, Polleux J, Ba JH, Pinna N. Nonaqueous synthesis of metal oxide nanoparticles: Review and indium oxide as case study for the dependence of particle morphology on precursors and solvents. Journal of Sol-Gel Science and Technology 2006;40(2-3) 259-266. DOI: 10.1007/s10971-006-6668-8.
  76. 76. Ribeiro C, Malagutti, A. R., Mendonça, V. R. E Mourão, A. J. L. Nanoestruturas em fotocatálise: uma revisão sobre estratégias de síntese de fotocatalisadores em escala nanométrica. Química Nova 2009;32(8).
  77. 77. Muniz EC, Goes MS, Silva JJ, Varela JA, Joanni E, Parra R, et al. Synthesis and characterization of mesoporous TiO2 nanostructured films prepared by a modified sol-gel method for application in dye solar cells. Ceramics International 2011;37(3) 1017-1024. DOI: 10.1016/j.ceramint.2010.11.014.
  78. 78. Kojima T, Sugimoto T. Formation Mechanism of Amorphous TiO2 Spheres in Organic Solvents 3. Effects of Water, Temperature, and Solvent Composition. Journal of Physical Chemistry C 2008;112(47) 18445-18454. DOI: 10.1021/Jp802957e.
  79. 79. Mahoney L, Koodali RT. Versatility of evaporation-induced self-assembly (EISA) method for preparation of mesoporous TiO2 for energy and environmental applications. Materials 2014;7(4) 2697-2746. DOI: 10.3390/Ma7042697.
  80. 80. Arconada N, Castro Y, Duran A, Hequet V. Photocatalytic oxidation of methyl ethyl ketones over sol-gel mesoporous and meso-structured TiO2 films obtained by EISA method. Applied Catalysis B-Environmental 2011;107(1-2) 52-58. DOI: 10.1016/j.apcatb.2011.06.036.
  81. 81. Ouyang J, Li XY, Jin J, Yang HM, Tang AD. Surface status and reduction behavior of porous ceria (CeO2) via amended EISA method. Journal of Alloys and Compounds 2014;606 236-241. DOI: 10.1016/j.jallcom.2014.04.048.
  82. 82. Chiola V, Ritsko JE, Vanderpool CD, Process for producing low-bulk density silica. Patent: US 3556725D A. US; 19/01/1971.
  83. 83. Yanagisawa T, Shimizu T, Kuroda K, Kato C. Trimethylsilyl derivatives of alkyltrimethylammonium-kanemite complexes and their conversion to microporous SiO2 materials. The Bulletin of the Chemical Society of Japan 1990;63(5) 1535–1537 DOI: 10.1246/bcsj.63.1535.
  84. 84. Badawy MI, Ghaly MY, Ali MEM. Photocatalytic hydrogen production over nanostructured mesoporous titania from olive mill wastewater. Desalination 2011;267(2-3) 250-255. DOI: 10.1016/j.desal.2010.09.035.
  85. 85. Bian YR, Wang XH, Hu ZY. Synthesis of mesoporous titania thin films by a simple route at low-temperature via plasma treatment. Journal of Materials Science 2013;48(11) 4088-4096. DOI: 10.1007/s10853-013-7221-0.
  86. 86. Zhou W, Li W, Wang JQ, Qu Y, Yang Y, Xie Y, et al. Ordered Mesoporous Black TiO2 as Highly Efficient Hydrogen Evolution Photocatalyst. Journal of the American Chemical Society 2014;136(26) 9280-9283. DOI: 10.1021/Ja504802q.
  87. 87. Xiang P, Li X, Wang H, Liu GH, Shu T, Zhou ZM, et al. Mesoporous nitrogen-doped TiO2 sphere applied for quasi-solid-state dye-sensitized solar cell. Nanoscale Research Letters 2011;6. DOI: 10.1186/1556-276X-6-606.
  88. 88. Wu ZY, Tao YF, Lin Z, Liu L, Fan XX, Wang Y. Hydrothermal Synthesis and Morphological Evolution of Mesoporous Titania-Silica. Journal of Physical Chemistry C 2009;113(47) 20335-20348. DOI: 10.1021/jp9037842.
  89. 89. Wen ZH, Wu W, Liu Z, Zhang H, Li JH, Chen JH. Ultrahigh-efficiency photocatalysts based on mesoporous Pt-WO3 nanohybrids. Physical Chemistry Chemical Physics 2013;15(18) 6773-6778. DOI: 10.1039/c3cp50647a.
  90. 90. Wang YZ, Zhu SP, Chen XR, Tang YG, Jiang YF, Peng ZG, et al. One-step template-free fabrication of mesoporous ZnO/TiO2 hollow microspheres with enhanced photocatalytic activity. Applied Surface Science 2014;307 263-271. DOI: 10.1016/j.apsusc.2014.04.023.
  91. 91. Shamaila S, Khan A, Sajjad L, Chen F, Zhang JL. Mesoporous titania with high crystallinity during synthesis by dual template system as an efficient photocatalyst. Catalysis Today 2011;175(1) 568-575. DOI: 10.1016/j.cattod.2011.03.041.
  92. 92. Antonelli DM, Ying JY. Synthesis of Hexagonally Packed Mesoporous TiO2 by a Modified Sol-Gel Method. Angewandte Chemie-International Edition in English 1995;34(18) 2014-2017. DOI: 10.1002/anie.199520141.
  93. 93. Joo JB, Lee I, Dahl M, Moon GD, Zaera F, Yin YD. Controllable Synthesis of Mesoporous TiO2 Hollow Shells: Toward an Efficient Photocatalyst. Advanced Functional Materials 2013;23(34) 4246-4254. DOI: 10.1002/adfm.201300255.
  94. 94. Lopez A, Acosta D, Martinez AI, Santiago J. Nanostructured low crystallized titanium dioxide thin films with good photocatalytic activity. Powder Technology 2010;202(1-3) 111-117. DOI: 10.1016/j.powtec.2010.04.025.
  95. 95. Ramasamy E, Jo C, Anthonysamy A, Jeong I, Kim JK, Lee J. Soft-template simple synthesis of ordered mesoporous titanium nitride-carbon nanocomposite for high performance dye-sensitized solar cell counter electrodes. Chemistry of Materials 2012;24(9) 1575-1582. DOI: 10.1021/Cm203672g.
  96. 96. Tran TH, Nosaka AY, Nosaka Y. Adsorption and decomposition of a dipeptide (Ala-Trp) in TiO2 photocatalytic systems. Journal of Photochemistry and Photobiology a-Chemistry 2007;192(2-3) 105-113. DOI: 10.1016/j.jphotochem.2007.05.011.
  97. 97. Lu YF, Fan HY, Stump A, Ward TL, Rieker T, Brinker CJ. Aerosol-assisted self-assembly of mesostructured spherical nanoparticles. Nature 1999;398(6724) 223-226. DOI:10.1038/18410.
  98. 98. Yang H, Coombs N, Sokolov I, Ozin GA. Free-standing and oriented mesoporous silica films grown at the air-water interface. Nature 1996;381(6583) 589-592. DOI: 10.1038/381589a0.
  99. 99. Yang XH, Fu HT, Yu AB, Jiang XC. Large-surface mesoporous TiO2 nanoparticles: Synthesis, growth and photocatalytic performance. Journal of Colloid and Interface Science 2012;387 74-83. DOI: 10.1016/j.jcis.2012.06.080.
  100. 100. Wu MT, Chow TJ. TiO2 particles prepared by size control self-assembly and their usage on dye-sensitized solar cell. Microporous and Mesoporous Materials 2014;196 354-358. DOI: 10.1016/j.micromeso.2014.05.035.
  101. 101. Gao J, Li HR, Rong H, Dai YH. Large pore nanocrystalline TiO2 films for quasi-solid state dye-sensitized solar cells. Nanotechnology and Precision Engineering, Pts 1 and 2 2013;662 177-181. DOI: 10.4028/www.scientific.net/AMR.662.177.
  102. 102. Wang Y, Jiang ZH, Yang FJ. Preparation and photocatalytic activity of mesoporous TiO2 derived from hydrolysis condensation with TX-100 as template. Materials Science and Engineering B-Solid State Materials for Advanced Technology 2006;128(1-3) 229-233. DOI: 10.1016/j.mseb.2005.12.004.
  103. 103. Rodriguez-Abreu C, Vilanova N, Solans C, Ujihara M, Imae T, Lopez-Quintela A, et al. A combination of hard and soft templating for the fabrication of silica hollow microcoils with nanostructured walls. Nanoscale Research Letters 2011;6. DOI: 10.1186/1556-276x-6-330.
  104. 104. Hu CY, Xu YJ, Duo SW, Li WK, Xiang JH, Li MS, et al. Preparation of Inorganic Hollow Spheres Based on Different Methods. Journal of the Chinese Chemical Society 2010;57(5A) 1091-1098. DOI: 10.1002/jccs.201000154.
  105. 105. Tang H, Zhang D, Tang GG, Ji XR, Li CS, Yan XH, et al. Low temperature synthesis and photocatalytic properties of mesoporous TiO2 nanospheres. Journal of Alloys and Compounds 2014;591 52-57. DOI: 10.1016/j.jallcom.2013.12.176.
  106. 106. Chen JF, Hua ZJ, Yan YS, Zakhidov AA, Baughman RH, Xu LB. Template synthesis of ordered arrays of mesoporous titania spheres. Chemical Communications 2010;46(11) 1872-1874. DOI: 10.1039/B915706a.
  107. 107. Ma TY, Liu L, Yuan ZY. Direct synthesis of ordered mesoporous carbons. Chemical Society Reviews 2013;42(9) 3977-4003. DOI: 10.1039/C2cs35301f.
  108. 108. Wan J, Lei Y, Zhang YR, Leng Y, Liu J. Study on TiO2 photoelectrode to improve the overall performance of dye-sensitized solar cells. Electrochimica Acta 2012;59 75-80. DOI: 10.1016/j.electacta.2011.10.040.
  109. 109. Wyrwas B, Chrzanowski L, Lawniczak L, Szulc A, Cyplik P, Bialas W, et al. Utilization of Triton X-100 and polyethylene glycols during surfactant-mediated biodegradation of diesel fuel. Journal of Hazardous Materials 2011;197 97-103. DOI: 10.1016/j.jhazmat.2011.09.060.
  110. 110. Batigoc C, Akbas H, Boz M. Thermodynamics of non-ionic surfactant Triton X-100-cationic surfactants mixtures at the cloud point. Journal of Chemical Thermodynamics 2011;43(12) 1800-1803. DOI: 10.1016/j.jct.2011.06.005.
  111. 111. Patel V, Ray D, Aswal VK, Bahadur P. Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth. Colloids and Surfaces a-Physicochemical and Engineering Aspects 2014;450 106-114. DOI: 10.1016/j.colsurfa.2014.03.015.
  112. 112. Xu S, Zhou CH, Yang Y, Hu H, Sebo B, Chen BL, et al. Effects of Ethanol on Optimizing Porous Films of Dye-Sensitized Solar Cells. Energy & Fuels 2011;25(3) 1168-1172. DOI: 10.1021/Ef101546a.
  113. 113. Dhungel SK, Park JG. Optimization of paste formulation for TiO2 nanoparticles with wide range of size distribution for its application in dye sensitized solar cells. Renewable Energy 2010;35(12) 2776-2780. DOI: 10.1016/j.renene.2010.04.031.
  114. 114. Li Y, Wang WN, Zhan ZL, Woo MH, Wu CY, Biswas P. Photocatalytic reduction of CO2 with H2O on mesoporous silica supported Cu/TiO2 catalysts. Applied Catalysis B-Environmental 2010;100(1-2) 386-392. DOI: 10.1016/j.apcatb.2010.08.015.
  115. 115. Nguyen-Phan TD, Pham HD, Kim S, Oh ES, Kim EJ, Shin EW. Surfactant removal from mesoporous TiO2 nanocrystals by supercritical CO2 fluid extraction. Journal of Industrial and Engineering Chemistry 2010;16(5) 823-828. DOI: 10.1016/j.jiec.2010.05.005.
  116. 116. Spataru T, Preda L, Osiceanu P, Munteanu C, Anastasescu M, Marcu M, et al. Role of surfactant-mediated electrodeposited titanium oxide substrate in improving electrocatalytic features of supported platinum particles. Applied Surface Science 2014;288 660-665. DOI: 10.1016/j.apsusc.2013.10.092.
  117. 117. Mali SS, Kim H, Shim CS, Patil PS, Kim JH, Hong CK. Surfactant free most probable TiO2 nanostructures via hydrothermal and its dye sensitized solar cell properties. Scientific Reports 2013;3. DOI: 10.1038/Srep03004.
  118. 118. Rahman MYA, Umar AA, Roza L, Salleh MM. Effect of optical property of surfactant-treated TiO2 nanostructure on the performance of TiO2 photo-electrochemical cell. Journal of Solid State Electrochemistry 2012;16(5) 2005-2010. DOI: 10.1007/s10008-011-1605-3.
  119. 119. Hashimoto K, Irie H, Fujishima A. TiO2 photocatalysis: A historical overview and future prospects. Japanese Journal of Applied Physics Part 1-Regular Papers Brief Communications & Review Papers 2005;44(12) 8269-8285. DOI: 10.1143/Jjap.44.8269.
  120. 120. Hoffmann MR, Martin ST, Choi WY, Bahnemann DW. Environmental Applications of Semiconductor Photocatalysis. Chemical Reviews 1995;95(1) 69-96. DOI: 10.1021/Cr00033a004.
  121. 121. Ibhadon AO, Fitzpatrick P. Heterogeneous Photocatalysis: Recent Advances and Applications. Catalysts 2013;3(1) 189-218. DOI: 10.3390/Catal3010189.
  122. 122. Oregan B, Gratzel M. A Low-Cost, High-Efficiency Solar-Cell Based on Dye-Sensitized Colloidal TiO2 Films. Nature 1991;353(6346) 737-740. DOI: 10.1038/353737a0.
  123. 123. Khataee AR, Zarei M, Ordikhani-Seyedlar R. Heterogeneous photocatalysis of a dye solution using supported TiO2 nanoparticles combined with homogeneous photoelectrochemical process: Molecular degradation products. Journal of Molecular Catalysis a-Chemical 2011;338(1-2) 84-91. DOI: 10.1016/j.molcata.2011.01.028.
  124. 124. Zhao S, Su D, Che J, Jiang BY, Orlov A. Photocatalytic properties of TiO2 supported on SBA-15 mesoporous materials with large pores and short channels. Materials Letters 2011;65(23-24) 3354-3357. DOI 10.1016/j.matlet.2011.07.053.
  125. 125. Feng CG, Li YZ, Liu X. Photocatalytic Degradation of Imidacloprid by Phosphotungstic Acid Supported on a Mesoporous Sieve MCM-41. Chinese Journal of Chemistry 2012;30(1) 127-132. DOI: 10.1002/cjoc.201180453.
  126. 126. Schneider J, Bahnemann DW. Undesired Role of Sacrificial Reagents in Photocatalysis. Journal of Physical Chemistry Letters 2013;4(20) 3479-3483.
  127. 127. Shin JY, Samuelis D, Maier J. Sustained Lithium-Storage Performance of Hierarchical, Nanoporous Anatase TiO2 at High Rates: Emphasis on Interfacial Storage Phenomena. Advanced Functional Materials 2011;21(18) 3464-3472. DOI: 10.1002/adfm.201002527.
  128. 128. Zhao D, Budhi S, Rodriguez A, Koodali RT. Rapid and facile synthesis of Ti-MCM-48 mesoporous material and the photocatalytic performance for hydrogen evolution. International Journal of Hydrogen Energy 2010;35(11) 5276-5283. DOI: 10.1016/j.ijhydene.2010.03.087.
  129. 129. Liu B, Liu LM, Lang XF, Wang HY, Lou XW, Aydil ES. Doping high-surface-area mesoporous TiO2 microspheres with carbonate for visible light hydrogen production. Energy & Environmental Science 2014;7(8) 2592-2597. DOI: 10.1039/C4ee00472h.
  130. 130. Armand M, Tarascon JM. Building better batteries. Nature 2008;451(7179) 652-657. DOI: 10.1038/451652a.
  131. 131. Dunn B, Kamath H, Tarascon JM. Electrical Energy Storage for the Grid: A Battery of Choices. Science 2011;334(6058) 928-935. DOI: 10.1126/science.1212741.
  132. 132. Guo YG, Hu YS, Maier J. Synthesis of hierarchically mesoporous anatase spheres and their application in lithium batteries. Chemical Communications 2006;2006(26) 2783-2785. DOI: 10.1039/B605090e.
  133. 133. Wang Y, Liu SQ, Huang KL, Fang D, Zhuang SX. Electrochemical properties of freestanding TiO2 nanotube membranes annealed in Ar for lithium anode material. Journal of Solid State Electrochemistry 2012;16(2) 723-729. DOI: 10.1007/s10008-011-1417-5.
  134. 134. Deng D, Kim MG, Lee JY, Cho J. Green energy storage materials: Nanostructured TiO2 and Sn-based anodes for lithium-ion batteries. Energy & Environmental Science 2009;2(8) 818-837. DOI: 10.1039/B823474d.
  135. 135. Dimroth F, Kurtz S. High-efficiency multijunction solar cells. Mrs Bulletin 2007;32(3) 230-235. DOI: 10.1557/Mrs2007.27.
  136. 136. Peter LM. Dye-sensitized nanocrystalline solar cells. Physical Chemistry Chemical Physics 2007;9(21) 2630-2642. DOI: 10.1039/B617073k.
  137. 137. Ito S, Nazeeruddin MK, Liska P, Comte P, Charvet R, Pechy P, et al. Photovoltaic characterization of dye-sensitized solar cells: Effect of device masking on conversion efficiency. Progress in Photovoltaics 2006;14(7) 589-601. DOI: 10.1002/Pip.683.
  138. 138. Park SH, Roy A, Beaupré S, Cho S, Coates N, Moon JS, et al. Bulk heterojunction solar cells with internal quantum efficiency approaching 100%. Nature Photonics 2009;3 297-302. DOI:10.1038/nphoton.2009.69.
  139. 139. Gratzel M. Solar energy conversion by dye-sensitized photovoltaic cells. Inorganic Chemistry 2005;44(20) 6841-6851. DOI: 10.1021/Ic0508371.
  140. 140. O'Regan B, Grätzel M. A Low-Cost, High-Efficiency Solar-Cell Based on Dye-Sensitized Colloidal TiO2 Films. Nature 1991;353(6346) 737-740. 10.1038/353737a0.
  141. 141. Gratzel M. Photoelectrochemical cells. Nature 2001;414(6861) 338-344. DOI: 10.1038/35104607.
  142. 142. Chiba Y, Islam A, Watanabe Y, Komiya R, Koide N, Han LY. Dye-sensitized solar cells with conversion efficiency of 11.1%. Japanese Journal of Applied Physics, Part 2: Letters & Express Letters 2006;45(24-28) L638-L640.
  143. 143. Kroon JM, Bakker NJ, Smit HJP, Liska P, Thampi KR, Wang P, et al. Nanocrystalline dye-sensitized solar cells having maximum performance. Progress in Photovoltaics 2007;15(1) 1-18. DOI: 10.1002/Pip.707.
  144. 144. Kalyanasundaram K, Gratzel M. Applications of functionalized transition metal complexes in photonic and optoelectronic devices. Coordination Chemistry Reviews 1998;177 347-414. DOI: 10.1016/S0010-8545(98)00189-1.
  145. 145. Hinsch A, Veurman W, Brandt H, Loayza Aguirre R, Bialecka K, Flarup Jensen K. Worldwide first fully up-scaled fabrication of 60 × 100 cm2 dye solar module prototypes. Progress in Photovoltaics 2012;20(6) 698-710. DOI: 10.1002/pip.1213.
  146. 146. Higashijima S, Miura H, Fujita T, Kubota Y, Funabiki K, Yoshida T, et al. Highly efficient new indoline dye having strong electron-withdrawing group for zinc oxide dye-sensitized solar cell (vol 67, pg 6289, 2011). Tetrahedron 2011;67(43) 8421-8421. DOI: 10.1016/j.tet.2011.08.092.
  147. 147. Kuang D, Uchida S, Humphry-Baker R, Zakeeruddin SM, Grätzel M. Organic dye-sensitized ionic liquid based solar cells: Remarkable enhancement in performance through molecular design of indoline sensitizers. Angewandte Chemie-International Edition 2008;47(10) 1923-1927.
  148. 148. Patrocinio AOT, Murakami Iha NY. Em busca da sustentabilidade: Células solares sensibilizadas por extratos naturais. Quimica Nova 2010;33(3) 574-578.
  149. 149. Patrocinio AOT, Mizoguchi SK, Paterno LG, Murakami Iha NY. Efficient and low cost devices for solar energy conversion: efficiency and stability of some natural dye sensitized solar cells. Synthetic Metals 2009;159 2342-2344. DOI: 10.1016/j.synthmet.2009.08.027.
  150. 150. Shahid M, Shahid ul I, Mohammad F. Recent advancements in natural dye applications: a review. Journal of Cleaner Production 2013;53(0) 310-331. DOI: http://dx.doi.org/10.1016/j.jclepro.2013.03.031.
  151. 151. Yella A, Lee H-W, Tsao HN, Yi C, Chandiran AK, Nazeeruddin MK, et al. Porphyrin-Sensitized Solar Cells with Cobalt (II/III)-Based Redox Electrolyte Exceed 12 Percent Efficiency. Science 2011;334(6056) 629-634. 10.1126/science.1209688.
  152. 152. Adithi U, Thomas S, Uma V, Pradeep N. Electrical Characterization of Dye sensitized Nano Solar cell using Natural Pomegranate juice as Photosensitizer. In: Chauhan AK, Murli C, Gadkari SC (ed.) Solid State Physics 2013. p. 208-209.
  153. 153. Zhang Q, Liu X. Dye-Sensitized Solar Cell Goes Solid. Small 2012;8(24) 3711-3713. 10.1002/smll.201201759.
  154. 154. Bach U, Daeneke T. A Solid Advancement for Dye-Sensitized Solar Cells. Angewandte Chemie-International Edition 2012;51(42) 10451-10452. DOI: 10.1002/anie.201205437.
  155. 155. Yanagida S, Yu YH, Manseki K. Iodine/Iodide-Free Dye-Sensitized Solar Cells. Accounts of Chemical Research 2009;42(11) 1827-1838. DOI: 10.1021/Ar900069p.
  156. 156. Park N-G. Organometal Perovskite Light Absorbers Toward a 20% Efficiency Low-Cost Solid-State Mesoscopic Solar Cell. Journal of Physical Chemistry Letters 2013;4(15) 2423-2429. DOI: 10.1021/jz400892a.
  157. 157. Burschka J, Pellet N, Moon SJ, Humphry-Baker R, Gao P, Nazeeruddin MK, et al. Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature 2013;499(7458) 316-+. DOI: 10.1038/Nature12340.
  158. 158. Bisquert J. The Swift Surge of Perovskite Photovoltaics. Journal of Physical Chemistry Letters 2013;4(15) 2597-2598. DOI: 10.1021/jz401435d.
  159. 159. Krüger J, Plass R, Cevey L, Piccirelli M, Grätzel M, Bach U. High efficiency solid-state photovoltaic device due to inhibition of interface charge recombination. Applied Physics Letters 2001;79(13) 2085-2087. DOI: http://dx.doi.org/10.1063/1.1406148.
  160. 160. Krüger J, Plass R, Grätzel M, Matthieu H-J. Improvement of the photovoltaic performance of solid-state dye-sensitized device by silver complexation of the sensitizer cis-bis(4,4′-dicarboxy-2,2′bipyridine)-bis(isothiocyanato) ruthenium(II). Applied Physics Letters 2002;81(2) 367-369. DOI: http://dx.doi.org/10.1063/1.1490394.
  161. 161. Ito S, Murakami TN, Comte P, Liska P, Grätzel C, Nazeeruddin MK, et al. Fabrication of thin film dye sensitized solar cells with solar to electric power conversion efficiency over 10%. Thin Solid Films 2008;516(14) 4613-4619. DOI: 10.1016/j.tsf.2007.05.090.
  162. 162. Patrocinio AOT, El-Bacha AS, Paniago EB, Paniago RM, Murakami Iha NY. Influence of the Sol-Gel pH Process and Compact Film on the Efficiency of TiO2-Based Dye-Sensitized Solar Cells. International Journal of Photoenergy 2012;638 5711-5717. DOI: 10.1155/2012/638571.
  163. 163. Patrocinio AOT, Paterno LG, Murakami Iha NY. Role of Polyelectrolyte for Layer-by-Layer Compact TiO2 Films in Efficiency Enhanced Dye-Sensitized Solar Cells. Journal of Physical Chemistry C 2010;114(41) 17954-17959. DOI: 10.1021/jp104751g.
  164. 164. Patrocinio AOT, Paterno LG, Murakami Iha NY. Layer-by-layer TiO2 films as efficient blocking layers in dye-sensitized solar cells. Journal of Photochemistry and Photobiology a-Chemistry 2009;20 523-527. DOI: 10.1016/j.jphotochem.2009.04.008.
  165. 165. Paula LF, Amaral RC, Murakami Iha NY, Paniago RM, Machado AEH, Patrocinio AOT. New layer-by-layer Nb2O5-TiO2 film as an effective underlayer in dye-sensitised solar cells. Rsc Advances 2014;4(20) 10310-10316. DOI: 10.1039/C4RA00058G.

Written By

Antonio E. H. Machado, Karen A. Borges, Tatiana A. Silva, Lidiaine M. Santos, Mariana F. Borges, Werick A. Machado, Bruno P. Caixeta, Marcela Dias França, Samuel M. Oliveira, Alam G. Trovó and Antonio O.T. Patrocínio

Submitted: 11 June 2014 Published: 06 May 2015