Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Metal Oxides for Dye-Sensitized Solar Cells

Journal of The American Ceramic Society, 2009
The incessant demand for energy forces us to seek it from sustainable resources; and concerns on environment demands that resources should be clean as well. Metal oxide semiconductors, which are stable and environment friendly materials, are used in photovoltaics either as photoelectrode in dye solar cells (DSCs) or to build metal oxide p–n junctions. Progress made in utilization of metal oxides for photoelectrode in DSC is reviewed in this article. Basic operational principle and factors that control the photoconversion efficiency of DSC are briefly outlined. The d-block binary metal oxides viz. TiO2, ZnO, and Nb2O5 are the best candidates as photoelectrode due to the dissimilarity in orbitals constituting their conduction band and valence band. This dissimilarity decreases the probability of charge recombination and enhances the carrier lifetime in these materials. Ternary metal oxide such as Zn2SnO4 could also be a promising material for photovoltaic application. Various morphologies such as nanoparticles, nanowires, nanotubes, and nanofibers have been explored to enhance the energy conversion efficiency of DSCs. The TiO2 served as a model system to study the properties and factors that control the photoconversion efficiency of DSCs; therefore, such discussion is limited to TiO2 in this article. The electron transport occurs through nanocrystalline TiO2 through trapping and detrapping events; however, exact nature of these trap states are not thoroughly quantified. Research efforts are required not only to quantify the trap states in mesoporous metal oxides but new mesoporous architectures also to increase the conversion efficiency of metal oxide-based photovoltaics....Read more
Metal Oxides for Dye-Sensitized Solar Cells Rajan Jose, w Velmurugan Thavasi, and Seeram Ramakrishna NUS Nanoscience and Nanotechnology Initiative, Faculty of Engineering, National University of Singapore, 117576 Singapore The incessant demand for energy forces us to seek it from sus- tainable resources; and concerns on environment demands that resources should be clean as well. Metal oxide semiconductors, which are stable and environment friendly materials, are used in photovoltaics either as photoelectrode in dye solar cells (DSCs) or to build metal oxide pn junctions. Progress made in utiliza- tion of metal oxides for photoelectrode in DSC is reviewed in this article. Basic operational principle and factors that control the photoconversion efficiency of DSC are briefly outlined. The d-block binary metal oxides viz. TiO 2 , ZnO, and Nb 2 O 5 are the best candidates as photoelectrode due to the dissimilarity in orbitals constituting their conduction band and valence band. This dissimilarity decreases the probability of charge recombi- nation and enhances the carrier lifetime in these materials. Ter- nary metal oxide such as Zn 2 SnO 4 could also be a promising material for photovoltaic application. Various morphologies such as nanoparticles, nanowires, nanotubes, and nanofibers have been explored to enhance the energy conversion efficiency of DSCs. The TiO 2 served as a model system to study the properties and factors that control the photoconversion efficiency of DSCs; therefore, such discussion is limited to TiO 2 in this article. The electron transport occurs through nanocrystalline TiO 2 through trapping and detrapping events; however, exact nature of these trap states are not thoroughly quantified. Research efforts are required not only to quantify the trap states in mesoporous metal oxides but new mesoporous architectures also to increase the conversion efficiency of metal oxide-based photovoltaics. I. Introduction I N light of rising energy demand and depleting oil resources, alternate energy sources are actively sought. Overdependence on fossil fuels leaves us vulnerable to air pollutants, viz. CO, NO x , and SO x , and related health risks and global warming due to the increased green-house gas concentration. These issues raise severe concerns on the sustenance of life on the earth. For example a rise of B11C in global temperature would melt a significant portion of the polar ice thereby removing large area from lower lying lands by the seas and oceans; migration of the living organism poleward thereby causing population imbal- ance; appearance of new microorganisms and diseases; etc. 1 Therefore, a step forward in the pursuit of alternative energy sources is the harness of energy from carbon-free sources such as wind, geothermal, hydroelectricity, tidal, and solar energy. Our primary source of clean abundant energy is the sun. More solar energy strikes the Earth in 1 h (4.3 10 20 J) than all the energy consumed on the planet in a year (4.1 10 20 J). 2 Con- version of this tremendous energy into electrical power, which is the science and technology behind photovoltaics, is an issue be- fore scientists and engineers for long time. Existing types of solar cells could be divided into two distinct classes: semiconductor pn junction solar cells and excitonic so- lar cells (ESCs). Excellent monographs are published by Michel Gra¨ tzel 3 and Brian Gregg 4 that detail the fundamental differ- ences between the pn junction and ESCs. In the pn junction cells, light absorption results in the creation of free electrons, which are subsequently accelerated by the in-built electric field in the junction. The photovoltage in the pn junction is the differ- ence in the quasi-Fermi levels of n- and p-type regions. In the ESCs, light absorption results in the generation of a transiently localized excited state, known as exciton—usually a Frenkel type that cannot thermally dissociate into free carriers in the material it was formed due to high exciton binding energy (kT). Excitons are the characteristics of semiconductor ana- logues to Fermi fluids in metals and are often characterized as a mobile excited state with an exciton diffusion coefficient. An exciton can be viewed as a quasi particle with an electron in the conduction band (or lowest unoccupied molecular orbital (LUMO) in the case of molecules and nanoclusters) and a hole in the valence band (or highest occupied molecular orbital (HOMO) in the case of molecules and nanoclusters). When a semiconductor (molecule, crystals, or clusters) is anchored to another material whose conduction band (LUMO) lies at lower energy, then the exciton dissociates into mobile carriers at the interface of the materials system (Fig. 1). This process is the basis of ESCs. Examples of this type include organic solar cells, 5 dye solar cells—most popularly known as dye-sensitized solar cells (DSCs), 3 and quantum dot solar cells. 6 Conjugated poly- mers and/or organic materials such as C 60 , carbon nanotubes, etc. are the materials of choice in organic solar cells; in DSCs, a wide bandgap metal oxide semiconductor is anchored to a dye; and in quantum dot solar cells, semiconductor nanocrystals with size less than their exciton Bohr radius is used to harvest light. Progress in the photovoltaic conversion efficiency is reviewed annually; a most recent update is provided by Green et al. 7 Feature D. J. Green—contributing editor This work was partially funded from the Clean Energy Program Office, Economic Development Board of Singapore. w Author to whom correspondence should be addressed. nnijr@nus.edu.sg Manuscript No. 25165. Received August 28, 2008; approved October 30, 2008. J ournal J. Am. Ceram. Soc., 92 [2] 289–301 (2009) DOI: 10.1111/j.1551-2916.2008.02870.x r 2009 The American Ceramic Society
For convenience the conversion efficiency data have been summarized and shown in Table I. Section II of this article briefly explains the basic operational principle of DSCs, advantages of DSC over pn junctions, factors that control the photoconversion efficiency, techniques to characterize DSC, and basic equations; section III explain the metal oxides used in DSC, charge transport in DSC taking TiO 2 as a specific example, morphologies explored for DSC, and the situation when metal oxides are not sensitized with dyes. II. DSCs The sensitization of semiconductors using dyes dates back to 19th century by the work on the silver halide emulsion using erythrosine dyes. 8 Systematic studies to understand the mecha- nistic details of the dye sensitization started after a century using RuL 2 (NCS) 2 ;L 5 2.2 0 -bipyridyl-4,4 0 -dicarboxylic acid dye (N3) and wide bandgap semiconductors such as ZnO and SnO. 9 These early studies were fundamental in nature aimed to under- stand the charge transfer process between the LUMO of the dye and the conduction band of the semiconductor when this ma- terial system is immersed in an electrolyte. Commercial interest in DSC has been triggered by the discovery that mesoporous TiO 2 anchored to N3 dye gave significant energy conversion efficiency (ZB7%). 10 This breakthrough was due to large sur- face area of the mesoporous TiO 2 that allowed anchoring sig- nificantly high amount of dye molecules (B0.13 mmol/cm 2 ) onto it; thereby increasing the absorption cross-section. The DSCs are fabricated by sandwiching a dye-anchored mesoporous metal oxide, known as photoelectrode, between two conducting glass plates (such as fluorine-doped indium tin oxide (FTO)) in the presence of an electrolyte. One of the FTO plates is coated with a thin layer of metal oxide to prevent wetting of the FTO plate by liquid electrolytes and the other is coated with a thin layer of platinum that acts as catalyst (Fig. 2). The FTO is a cost-limiting factor of DSC; commercial DSCs use Ti thin films in the place of FTO. The DSC has potential for becoming a cost-effective means for producing electricity, capa- ble of competing with available solar electric technologies and, eventually, with today’s conventional power technologies. 11 Assuming DSCs to be single junction, these cells can have a theoretical conversion efficiency limit of B31%. In 2001, Gra¨ tzel’s group reported ZB10.4% using the black dye, 20 nm TiO 2 particles, and the iodide/triiodide electrolyte. 12 The same group had broken this record and reported ZB11.04% using N3 dye, 20 nm TiO 2 particles, and guanidine thiocyanate electrolyte. 13 In 2006, Han and colleagues from Sharp Co. (Japan) reported ZB11.1% using the black dye Fig. 1. A dye molecule (left) anchored onto a metal oxide semiconduc- tor nanocrystal (right). Upon absorption of a photon of sufficient energy, an electron leaves from the highest occupied molecular orbital (HOMO) (whose position is shown in black ellipse for the dye) to the lowest unoccupied molecular orbital (LUMO) (shown in blue circles) leaving behind a hole (h 1 ). This electron is injected into the conduction band of the metal oxide, which then diffuses and makes an electric current. Fig. 2. Schematic of dye solar cell. Mesoporous metal oxide (nanorods and spherical particles) act as photo electrode. Dye molecules are an- chored on the surface of metal oxide and upon photo-excitation of dye, the electron is injected into the conduction band of the electrode. The photoelectrode is percolated with an electrolyte whose redox potential supports the regeneration of dye after it gets reduced. Table I. State of the Art Performance Solar Cells Type of cell Description Area (cm 2 ) w Efficiency (%) pn (single junction) Silicon Crystalline 4.00 (da) 24.770.5 Polycrystalline 1.00 (ap) 20.370.5 Amorphous 1.07 (ap) 9.570.3 Nanocrystalline 1.20 (ap) 10.170.2 GaAs Crystalline 3.91 (t) 25.170.8 Thin film 1.00 (t) 24.570.5 Polycrystalline 4.01 (t) 18.270.5 InP Crystalline 4.02 (t) 21.970.7 pn (multijunction) GaInP/GaAs/Ge 3.99 (t) 32.071.5 GaInP/GaAs 4.0 (t) 30.3 Thin film CuInGaSe 2 1.00 (ap) 18.870.5 CuInGaSe 2 16.0 (ap) 16.670.4 CdTe 1.03 (ap) 16.570.5 Si 4.02 (ap) 16.670.4 Si (submodule) 96.3 (ap) 9.870.3 Excitonic Dye solar cells 1.00 (ap) 10.470.3 Dye solar cells 26.5 (ap) 6.270.2 Organic 1.00 (ap) 3.070.1 Quantum dot solar cells 6 0.2 1.7 w da, designated illumination area; ap, aperture area; t, total area. 290 Journal of the American Ceramic Society—Jose et al. Vol. 92, No. 2
J. Am. Ceram. Soc., 92 [2] 289–301 (2009) DOI: 10.1111/j.1551-2916.2008.02870.x r 2009 The American Ceramic Society Journal Metal Oxides for Dye-Sensitized Solar Cells Rajan Jose,w Velmurugan Thavasi, and Seeram Ramakrishna NUS Nanoscience and Nanotechnology Initiative, Faculty of Engineering, National University of Singapore, 117576 Singapore The incessant demand for energy forces us to seek it from sustainable resources; and concerns on environment demands that resources should be clean as well. Metal oxide semiconductors, which are stable and environment friendly materials, are used in photovoltaics either as photoelectrode in dye solar cells (DSCs) or to build metal oxide p–n junctions. Progress made in utilization of metal oxides for photoelectrode in DSC is reviewed in this article. Basic operational principle and factors that control the photoconversion efficiency of DSC are briefly outlined. The d-block binary metal oxides viz. TiO2, ZnO, and Nb2O5 are the best candidates as photoelectrode due to the dissimilarity in orbitals constituting their conduction band and valence band. This dissimilarity decreases the probability of charge recombination and enhances the carrier lifetime in these materials. Ternary metal oxide such as Zn2SnO4 could also be a promising material for photovoltaic application. Various morphologies such as nanoparticles, nanowires, nanotubes, and nanofibers have been explored to enhance the energy conversion efficiency of DSCs. The TiO2 served as a model system to study the properties and factors that control the photoconversion efficiency of DSCs; therefore, such discussion is limited to TiO2 in this article. The electron transport occurs through nanocrystalline TiO2 through trapping and detrapping events; however, exact nature of these trap states are not thoroughly quantified. Research efforts are required not only to quantify the trap states in mesoporous metal oxides but new mesoporous architectures also to increase the conversion efficiency of metal oxide-based photovoltaics. I. Introduction I N light of rising energy demand and depleting oil resources, alternate energy sources are actively sought. Overdependence on fossil fuels leaves us vulnerable to air pollutants, viz. CO, NOx, and SOx, and related health risks and global warming due to the increased green-house gas concentration. These issues raise severe concerns on the sustenance of life on the earth. For example a rise of B11C in global temperature would melt a significant portion of the polar ice thereby removing large area D. J. Green—contributing editor Manuscript No. 25165. Received August 28, 2008; approved October 30, 2008. This work was partially funded from the Clean Energy Program Office, Economic Development Board of Singapore. w Author to whom correspondence should be addressed. nnijr@nus.edu.sg from lower lying lands by the seas and oceans; migration of the living organism poleward thereby causing population imbalance; appearance of new microorganisms and diseases; etc.1 Therefore, a step forward in the pursuit of alternative energy sources is the harness of energy from carbon-free sources such as wind, geothermal, hydroelectricity, tidal, and solar energy. Our primary source of clean abundant energy is the sun. More solar energy strikes the Earth in 1 h (4.3  1020 J) than all the energy consumed on the planet in a year (4.1  1020 J).2 Conversion of this tremendous energy into electrical power, which is the science and technology behind photovoltaics, is an issue before scientists and engineers for long time. Existing types of solar cells could be divided into two distinct classes: semiconductor p–n junction solar cells and excitonic solar cells (ESCs). Excellent monographs are published by Michel Grätzel3 and Brian Gregg4 that detail the fundamental differences between the p–n junction and ESCs. In the p–n junction cells, light absorption results in the creation of free electrons, which are subsequently accelerated by the in-built electric field in the junction. The photovoltage in the p–n junction is the difference in the quasi-Fermi levels of n- and p-type regions. In the ESCs, light absorption results in the generation of a transiently localized excited state, known as exciton—usually a Frenkel type that cannot thermally dissociate into free carriers in the material it was formed due to high exciton binding energy ( kT). Excitons are the characteristics of semiconductor analogues to Fermi fluids in metals and are often characterized as a mobile excited state with an exciton diffusion coefficient. An exciton can be viewed as a quasi particle with an electron in the conduction band (or lowest unoccupied molecular orbital (LUMO) in the case of molecules and nanoclusters) and a hole in the valence band (or highest occupied molecular orbital (HOMO) in the case of molecules and nanoclusters). When a semiconductor (molecule, crystals, or clusters) is anchored to another material whose conduction band (LUMO) lies at lower energy, then the exciton dissociates into mobile carriers at the interface of the materials system (Fig. 1). This process is the basis of ESCs. Examples of this type include organic solar cells,5 dye solar cells—most popularly known as dye-sensitized solar cells (DSCs),3 and quantum dot solar cells.6 Conjugated polymers and/or organic materials such as C60, carbon nanotubes, etc. are the materials of choice in organic solar cells; in DSCs, a wide bandgap metal oxide semiconductor is anchored to a dye; and in quantum dot solar cells, semiconductor nanocrystals with size less than their exciton Bohr radius is used to harvest light. Progress in the photovoltaic conversion efficiency is reviewed annually; a most recent update is provided by Green et al.7 Feature 290 Vol. 92, No. 2 Journal of the American Ceramic Society—Jose et al. Fig. 1. A dye molecule (left) anchored onto a metal oxide semiconductor nanocrystal (right). Upon absorption of a photon of sufficient energy, an electron leaves from the highest occupied molecular orbital (HOMO) (whose position is shown in black ellipse for the dye) to the lowest unoccupied molecular orbital (LUMO) (shown in blue circles) leaving behind a hole (h1). This electron is injected into the conduction band of the metal oxide, which then diffuses and makes an electric current. For convenience the conversion efficiency data have been summarized and shown in Table I. Section II of this article briefly explains the basic operational principle of DSCs, advantages of DSC over p–n junctions, factors that control the photoconversion efficiency, techniques to characterize DSC, and basic equations; section III explain the metal oxides used in DSC, charge transport in DSC taking TiO2 as a specific example, morphologies explored for DSC, and the situation when metal oxides are not sensitized with dyes. II. DSCs The sensitization of semiconductors using dyes dates back to 19th century by the work on the silver halide emulsion using erythrosine dyes.8 Systematic studies to understand the mechanistic details of the dye sensitization started after a century using RuL2(NCS)2; L 5 2.20 -bipyridyl-4,40 -dicarboxylic acid dye (N3) and wide bandgap semiconductors such as ZnO and SnO.9 These early studies were fundamental in nature aimed to understand the charge transfer process between the LUMO of the dye and the conduction band of the semiconductor when this material system is immersed in an electrolyte. Commercial interest in DSC has been triggered by the discovery that mesoporous TiO2 anchored to N3 dye gave significant energy conversion efficiency (ZB7%).10 This breakthrough was due to large sur- Fig. 2. Schematic of dye solar cell. Mesoporous metal oxide (nanorods and spherical particles) act as photo electrode. Dye molecules are anchored on the surface of metal oxide and upon photo-excitation of dye, the electron is injected into the conduction band of the electrode. The photoelectrode is percolated with an electrolyte whose redox potential supports the regeneration of dye after it gets reduced. face area of the mesoporous TiO2 that allowed anchoring significantly high amount of dye molecules (B0.13 mmol/cm2) onto it; thereby increasing the absorption cross-section. The DSCs are fabricated by sandwiching a dye-anchored mesoporous metal oxide, known as photoelectrode, between two conducting glass plates (such as fluorine-doped indium tin oxide (FTO)) in the presence of an electrolyte. One of the FTO plates is coated with a thin layer of metal oxide to prevent wetting of the FTO plate by liquid electrolytes and the other is coated with a thin layer of platinum that acts as catalyst (Fig. 2). The FTO is a cost-limiting factor of DSC; commercial DSCs use Ti thin films in the place of FTO. The DSC has potential for becoming a cost-effective means for producing electricity, capable of competing with available solar electric technologies and, eventually, with today’s conventional power technologies.11 Assuming DSCs to be single junction, these cells can have a theoretical conversion efficiency limit of B31%. In 2001, Grätzel’s group reported ZB10.4% using the black dye, 20 nm TiO2 particles, and the iodide/triiodide electrolyte.12 The same group had broken this record and reported ZB11.04% using N3 dye, 20 nm TiO2 particles, and guanidine thiocyanate electrolyte.13 In 2006, Han and colleagues from Sharp Co. (Japan) reported ZB11.1% using the black dye Table I. State of the Art Performance Solar Cells Type of cell p–n (single junction) Description Silicon GaAs p–n (multijunction) Thin film Excitonic w InP GaInP/GaAs/Ge GaInP/GaAs CuInGaSe2 CuInGaSe2 CdTe Si Si (submodule) Dye solar cells Dye solar cells Organic Quantum dot solar cells6 da, designated illumination area; ap, aperture area; t, total area. Crystalline Polycrystalline Amorphous Nanocrystalline Crystalline Thin film Polycrystalline Crystalline Area (cm2)w Efficiency (%) 4.00 (da) 1.00 (ap) 1.07 (ap) 1.20 (ap) 3.91 (t) 1.00 (t) 4.01 (t) 4.02 (t) 3.99 (t) 4.0 (t) 1.00 (ap) 16.0 (ap) 1.03 (ap) 4.02 (ap) 96.3 (ap) 1.00 (ap) 26.5 (ap) 1.00 (ap) 0.2 24.770.5 20.370.5 9.570.3 10.170.2 25.170.8 24.570.5 18.270.5 21.970.7 32.071.5 30.3 18.870.5 16.670.4 16.570.5 16.670.4 9.870.3 10.470.3 6.270.2 3.070.1 1.7 February 2009 Metal Oxides for Solar Cells with increase in the haze of TiO2 electrode and the iodide/ triiodide electrolyte,14 which is currently the record holder. The solar cell efficiencies presented in Table I clearly demonstrates that the DSCs excel other ESCs and p–n junctions made using amorphous and nanocrystalline silicon. (1) Advantages/Disadvantages of DSC Over p–n Junctions The advantages of DSCs over p–n junction cells are as follows: 1. Less sensitive to impurities; fabrication in ordinary environments possible. 2. Easy fabrication process: inexpensive and scalable to nonvacuum- and low-temperature-based high volume manufacturing via continuous processes (e.g., screen printing, spraying, pressing, or roll-to-roll production) possible. 3. Operates optimally over a wide range of temperatures. 4. The efficiency is relatively insensitive to the angle of incident light. 5. Materials choice is not fixed; tunability of photovoltaic properties possible. 6. Cells can be made on lightweight and flexible or rigid substrates (e.g., plastic, fabric, metal, glass, and ceramic). 7. DSCs could be potentially less expensive (o$1.0/Wp) than p–n junctions.15 The major disadvantages of DSCs include the following: 1. The conversion efficiency of DSCs depends on a lot number of rate limiting factors, which are explained in the next section. 2. Less impressive cell lifetime due to the presence of volatile chemicals compared with p–n junctions. 3. One of the major disadvantages of DSCs for practical application is the transparent conducting glass plate onto which they are assembled. Research in this direction are in progress to build the DSCs on cost effective platforms such as metal layers.16 (2) Processes and Rate Limiters in DSC The functional diagram explaining the components of DSC and the processes in it is shown in Fig. 3. The photovoltaic effect in DSC occurs at the interface between a dye-anchored wide bandgap oxide semiconductor and an electrolyte. The HOMO of the dye molecules are usually formed by relatively weak lateral overlapping of atomic orbitals (p bonds) with loosely bond electrons (p electrons). Upon irradiation, the dye undergoes a p–p transition thereby exciting an electron to the LUMO of the dye (p electrons). The electron thus excited to the LUMO of the dye undergoes a nonradiative transition into the conduction Fig. 3. Functional diagram of dye solar cell (DSC). Each component in DSC is represented by isolated blocks, which slightly deviates from reality. For example, the electrolyte wet the metal oxide also. The dashed arrows show the possible losses in DSC. The vertical line on the right shows the energy scale in eV. The energies of each functional material are represented by horizontal lines. 291 band of the oxide material within few picoseconds thereby getting oxidized. Progress made in dyes for DSC has been recently reviewed by Neil.17 The oxidized dye is brought to the normal state by reducing it using a redox electrolyte together with a catalyst. A large number of electrolytes, both solids and liquids, are proposed for DSC applications, an account of which is available in a recent review.18 The electrons injected to the conduction band of the oxide material is transported through the mesoporous network to the counter electrode by doing a work equal to the energy difference between the conduction band of the electrode and redox potential of the electrolyte. There are at least eight fundamental processes: i. Photon absorption, which is determined by the absorption wavelength window, intensity of solar radiation at that window, and absorption cross-section of the dye (a). ii. Radiative recombination, i.e., the relaxation of the excited dye directly into its ground state, and its rate constant (k1). This process typically occurs in a time scale of several nanoseconds. iii. Exciton diffusion length (DEXT). iv. Interfacial electron transfer, i.e., injection of electron from the dye to the photoelectrode, and its rate constant (k2). The interfacial electron transfer occurs typically in a time scale of several picoseconds. v. Electron back transfer, i.e., the capturing of conduction band electrons by the oxidized species in the electrolyte, and its rate constant (k3). vi. Interfacial charge recombination, i.e., the capturing of conduction band electrons by the oxidized dye molecules, and its rate constant (k4). vii. Electron transport through the photoelectrode material controlled by (a) diffusion coefficient of electron (De), (b) phonon relaxation through which an electron loses its energy through electron–phonon interaction. viii. Redox potential of the electrolyte and rate constant of (k5) of electron transfer to the oxidized dye. Among these processes, interfacial charge transfer is three orders of magnitude faster than radiative recombination, i.e., k2/k1B1000; and electron back transfer is two orders of magnitude faster than interfacial charge recombination, i.e., k3/k4B100.19 Therefore, it is reasonable to assume that these two factors i.e., radiative recombination and interfacial electron transfer, make negligible contribution to the final conversion efficiency. The exciton is highly localized in dyes and nanoclusters such as quantum dots belong to strong confined regime, i.e., particle radius  exciton Bohr radius. In such cases the exciton diffusion could be neglected. However, if extended structures such quantum dots in the weak confinement regime, i.e., particle radiusBexciton Bohr radius, are used in place of dyes, the exciton dynamics should also be taken into account. Therefore, the problem of eight rate limiting process reduces into five for DSCs, viz., (i), (iv), (v), (vii), and (viii). Among the reduced five problems (i), (vii), and (viii) are controlled by the light harvester (dyes), photoelectrode, and electrolyte, respectively; and (iv) and (v) are processes at the light harvester– photoelectrode and photoelectrode–electrolyte interfaces, respectively. It follows from the above description that the metal oxide photoelectrode participate in more number of processes than electrolytes and the dyes do; and therefore, they play a significant role on the final energy conversion efficiency of DSC. (3) Characterization of DSC Three techniques are mainly used to characterize DSCs, viz. photoelectric current (I–V ) measurements, and electrochemical impedance spectroscopy (EIS), incident photon to conversion efficiency (IPCE). I–V: The electrical output power of solar cells is calculated from photoelectric current measurements. To compare solar cell characterized in different laboratories all over the world, the Z is measured under a set of standard conditions.20 Essentially these conditions specify that the temperature of the cell should be 292 Journal of the American Ceramic Society—Jose et al. 251C and that the solar radiation incident on the cell should have a total power density of 1000 W/m2, with a spectral power distribution characterized as AM1.5. The photovoltaic parameters for practical applications are open circuit voltage (VOC), short-circuit current density (JSC), fill factor (FF), Z. The VOC, as defined before, is proportional to the difference between the Fermi level of the TiO2 electrode and the electrochemical potential of the redox couple. This definition requires that the VOC for a given photoelectrode–electrolyte system to be a constant. However, the VOC was found to depend on the recombination rate in addition to the sensitizer and its adsorption mode. The sensitizers adsorption geometry induces substantial downshift of the conduction band energy of the metal oxide thereby reducing the VOC.21 The JSC is the current per unit active area of the device when the applied potential across the device is zero, i.e. ISC/A. Under standard conditions, the JSC is determined by (i) the amount of dye anchored onto the metal oxide and (ii) the electrochemical properties of the mesoporous network in the presence of an electrolyte. Recently, we have shown that the JSC depends on the molecular structure of the dye also.22 Gregg4 and Bisquert et al.23,24 contributed significantly to the understanding of charge transport through the mesoporous network and current density. The kinetic equation for the one-dimensional current density Jn(x) of electrons through DSC as4 Jn ðxÞ ¼ nðxÞmn fHUðxÞ þ HmðxÞg (2) where n(x) is the concentration of electrons, which depends on the amount of the electron injected by the dye; mn is the electron mobility; U and m are the electrical and chemical potential, respectively, which are constant for a given materials system. The carrier motion is through either diffusion (driven by concentration gradient) or drift (driven by an electric field) whose magnitudes are related by the Einstein’s equation D kT ¼ mn q (3) where D is the diffusion coefficient and mn is the mobility. The electron transport in DSC occurs in the presence of an electrolyte, which screens the macroscopic electric fields; and therefore, charge transport in DSC is primarily due to diffusion. Therefore, primary interest for each DSC constituted by a given materials system is to determine the electron diffusion coefficient. The FF measures the ideality of the device and is defined as the ratio of maximum power output to the product of VOC and JSC, i.e., FF ¼ IMAX VMAX JSC VOC (4) The FF is a measure of decrease in photocurrent with increase in photovoltage and could be qualitatively determined from the squareness of the I–V curve. Ideally, the power generated within the cell should dissipate at the external circuit. Parallel current paths such as electron back transfer (k3) and charge recombination (k4) within the device are possible cause of poor fill factor. An ideal cell should offer high internal resistance (41000 O) to these parallel currents, often characterized by a shunt resistance. Magnitude of the shunt resistance can be determined from the slope of the I–V curve near the short-circuit current point. Another possible cause of poor fill factor is the magnitude of series resistance at the cell–external circuit junction. High sheet resistance of the FTO plate and/or high resistance at this junction developed during processing of DSC are possible cause of high series resistance. The series resistance, the magnitude of which should be as low as possible, can be determined from the slop of the I–V curve near the open circuit voltage point. Vol. 92, No. 2 Fig. 4. Typical impedance spectrum (Nyquist plot or Cole–Cole plot) of DSC (bottom panel) in which the imaginary part of the complex impedance function is plotted against the real part. The electrical equivalent of a typical DSC is shown in the top panel; each circuit element is explained in the text. The Z is defined as Z¼ PMAX JSC VOC FF ¼ PIN PIN (5) i.e., higher JSC, VOC, and FF for lower solar irradiance are the key for increased Z. As stated before, the mesoporous metal oxide network directly influences all these parameters. Thus reducing the losses in the mesoporous network, a better DSC could be fabricated. EIS: The EIS is a useful tool to understand the kinetics of electrical transport in the DSC. Figure 4 shows a typical impedance spectrum of a DSC together with its electrical equivalent model.25 The DSC can be viewed electrically as a combination of resistance and capacitors. In the transmission model shown in Fig. 4, the rct is the charge-transfer resistance of the charge recombination process between electrons in the mesoporous metal oxide film and ions in the electrolyte; Cm is the chemical capacitance of the mesoporous metal oxide film; rt is the transport resistance of the electrons in the mesoporous metal oxide film; Zd is the Warburg element showing the Nernst diffusion of ions in the electrolyte; RPt and CPt are the charge-transfer resistance and double-layer capacitance at the counter electrode (platinized transparent conducting oxide (TCO) plate), respectively; RCTO and CCTO are the charge-transfer resistance and the corresponding double-layer capacitance at the exposed CTO– electrolyte interface, respectively; RCO and CCO are the resistance and the capacitance at the CTO–TiO2 contact, respectively; Rs is the series resistance, including the sheet resistance of the TCO glass and the contact resistance of the cell. Steady-state transport resistance through the mesoporous network, transient diffusion coefficient, chemical capacitance at the metal oxide– electrolyte interface, and recombination resistance could be evaluated using this single technique.26 IPCE: The IPCE, or action spectra, is defined as the number of electrons flowing through the external circuit under short circuit conditions per incident photon at a given wavelength. IPCEðlÞ ¼ nelectron ðlÞ IðlÞhc ¼ nphotons ðlÞ Pin ðlÞel (1) where I(l) is the measured current, Pin(l) is the input power, and l is the wavelength of irradiation in nanometer. Obviously a February 2009 Metal Oxides for Solar Cells 293 Fig. 6. (A) Ideal atomic arrangement in bulk anatase TiO2 where the atoms are symmetrically packed at regular lattices. When the size of the particles is reduced to nanometer level, the bond lengths and bond angle deviates from their equilibrium position (B). This deviation produces additional energy states in the electronic band of nanomaterials, which are the source of traps in nanocrystalline materials. Such defect has high density in nanocrystalline materials. Fig. 5. Plot showing the incident photon to current conversion efficiency as a function of the excitation wavelength for (a) (101) plane single crystal TiO2 and (b) mesoporous anatase when sensitized by the surface-anchored ruthenium complex cis-RuL2(SCN)2. The electrolyte consisted of a solution of 0.3M LiI and 0.03M I2 in acetonitrile. Data reproduced from Grätzel, Nature 414, 339 (2003). high fraction of the incident photons should be converted into mobile electrons for enhanced performance. A mesoporous network of metal oxide with high specific surface area is a prerequisite for high IPCE. Figure 5 compares the difference in IPCE when single crystal and mesoporous TiO2 was sensitized by the N3 dye. III. Photoelectrodes for DSC The exciton generated at the dye is separated into free carriers at the dye–photoelectrode interface. There are two driving forces for this charge separation: (i) the charge separation is energetically favored because the conduction band of the metal oxide is at lower energies than the LUMO of the dye; (ii) the charge separation is entropically favored because there is larger density of electronic energy states in the conduction band of a crystal than molecular orbital of a dye.27 Binary metal oxides such as TiO2, ZnO, Fe2O3, ZrO2, Nb2O5, Al2O3, and CeO2 and ternary compounds such as SrTiO3 and Zn2SnO4 have been tested for their use as photoelectrodes in DSC. Among them TiO2, ZnO, and Fe2O3 are 3d transition metal oxides, ZrO2 and Nb2O5 are 4d transition metal oxides, Al2O3 is a p-block metal oxide, and CeO2 is a f-block metal oxide. In TiO2, the Ti ions are in a distorted octahedral environment and formally have a Ti41(3d0) electronic confi- guration (Fig. 6(A)). The valence band of TiO2 is composed primarily of oxygen 2p orbitals hybridized with Ti 3d states, while the conduction band is made up of pure 3d orbital of titanium, i.e., electrons in valence band and conduction bands are at different parity; and therefore, the transition probability of electrons to the valence band is decreased which ultimately decrease the (e–h1) recombination probability. For Fe2O3, 3d-states are present both in the valence band and conduction band leading to states of similar parity and increase the (e–h1) recombination probability. In the case of ZnO, which has completely filled 3d orbitals (3d10), the valence band consists of only d orbitals and conduction band consists of hybridized s–p orbitals. The electronic configuration of ZnO again leads to the situation of dissimilar parity and reduced (e–h1) recombination probability. Thus in view of the electronic configuration and recombination probability TiO2 and ZnO are the best choice as photoelectrodes among 3d transition metal oxides. Stoichiometric Al2O3, ZrO2, and Nb2O5 are insulators at room temperature with bandgaps B6.6, B5, and B3.5 eV, respectively. Their room temperature conductivities are typically few mS/cm. However, the Nb2O5 become an n-type semiconductor at lower oxygen content and Nb2O4.978 has conductivity B3  103 S/cm.28 Similar to TiO2, conduction band of Nb2O5 consists of mainly 4d orbitals of niobium and valence band consists of 2p orbitals of oxygen and would lead to a situation of decreased recombination probability due to dissimilar parities. This discussion follows that TiO2, ZnO, and Nb2O5 are the best candidates in view of their electronic band structure. However, it should be noted that the band structure of nanomaterials is greatly deviated from their bulk counterpart. In agreement with this discussion, TiO2, ZnO, and Nb2O5 have been the most popular choices as photoelectrodes in DSCs. Among them, TiO2 was the material of choice since 1991 to address the fundamental issues of DSCs; therefore, it is considered as a model system. (1) TiO2 Electrodes TiO2 is a relatively cheap, abundant, nontoxic, biocompatible, oxide semiconducting material that is widely used in healthcare products as well as in paint. TiO2 exists naturally in three crystalline polymorphs, namely rutile (Eg 5 3.05 eV), anatase (Eg 5 3.23 eV), and brookite (Eg 5 3.26 eV). In addition, TiO2 exist in many other high pressure and metastable forms, viz., the intermediate orthorhombic,29 orthorhombic columbite30 and cotunnite types,29 monoclinic baddeleyite type,31 and cubic fluorite type.32 Rutile is the most common and stable TiO2 294 Vol. 92, No. 2 Journal of the American Ceramic Society—Jose et al. Fig. 7. Core/shell model and corresponding energy diagram. The shell induces potential barrier for electrons after injected into the core. This potential barrier prevent from electron to recombine with the electrolyte. polymorph. Both rutile and anatase have tetragonal structure with a 5 0.46 nm and c 5 0.29 nm (rutile); a 5 0.3782 nm and c 5 0.9502 nm (anatase). Brookite has orthorhombic structure with a 5 0.5456 nm, b 5 0.9182 nm, and c 5 0.5143 nm. Brookite is extremely difficult to synthesize in the laboratory but both anatase and rutile can be readily prepared. TiO2 is known as an n-type semiconductor, which contains donor-type defects such as oxygen vacancies and titanium interstitials. Presence of titanium vacancies in TiO2 cause p-type properties.33 The Fermi level of anatase is 0.1 eV higher than that of the rutile; therefore, anatase is more preferred for its applications in DSC. However, recent studies show that rutile scatters light more effectively and more chemically stable.34,35 Similar VOC has been reported for anatase and rutile despite the difference in their conduction band. The observed JSC using rutile was about 30% lesser than that of the anatase which was ascribed to the lower surface area; and therefore, decreased the amount of dye loading.34 The DSC designs using TiO2 were continuously renovated since 1991 to enhance the efficiency by increasing the lightscattering properties of the metal oxide film, suppressing charge recombination, improving the interfacial energetic, and altering the particle morphology.36 Most successful TiO2 architectures so far developed features25: (a) A TiO2 layer (thickness B100–200 nm) coating by immersing the FTO plate in precursors such as TiCl4 and subsequent sintering to prevent the electrolyte to reach the FTO plate. (b) A thick layer (thickness B12–14 mm) of mesoporous TiO2 of particle size 20–25 nm over the first monolayer. (c) A thin layer (thickness B2–5 mm) of TiO2 of particle size B400 nm to improve light scattering on the top of the mesoporous particle film. (A) Electron Transport in TiO2: Progress made in understanding the electron transport through nanocrystalline TiO2 both in its pure form and in the presence of an electrolyte is discussed elaborately by Bisquert24 as well as Watson and Meyer.37 The electron transport through any materials is controlled by its band structure; and therefore, inevitably linked to the bonding characteristics. The TiO2 is an ionic compound and exhibits strong electron–phonon coupling, known as polarons. For a sufficiently strong electron–phonon interaction, small polarons correspond to localized and self-trapped electrons are formed. In such cases, the electron transport typically occurs through thermally activated hopping from one electronic state to the next. Large polarons, with spatially extended wave functions, are formed for weaker coupling strength and exhibit band-type behavior. This is typically the mechanism of electron transport in bulk semiconductors. On the other hand, the electron transport in sintered nanocrystalline films depends on the properties of the individual nanoparticles, the extent of particle connectivity or electronic coupling between the particles, and the geometrical configuration of the assembly.24 The exciton Bohr diameter of TiO2 is 1.5 nm, below which TiO2 exhibit quantum size effects, i.e., splitting of its electronic band into discrete electronic levels. The TiO2 particles considered for DSC is typically in the 10–25 nm range i.e., much higher than its exciton Bohr radius of 1.5 nm. Therefore, the nanocrystalline film composed of 10 nm particles can be considered as a network of bulk crystals with a band of extended electronic states arising from the high surface area of nanocrystals. The surface states in nanomaterials, usually lie within the bandgap, arise from unsaturated bonds, deviations in the bond lengths and bond angle from that of the bulk materials,38 and presence of impurities (Fig. 7). However, in the case of TiO2 nature and origin of the surface states for the particles in the 10–25 nm range is not clear yet and intensive research is now required to determine them. To show the drastic changes in the electrical transport properties of TiO2 when particle size is decreased, four typical experiments39–42 have been summarized and shown in Table II. Electron mobility and effective diffusion coefficients for four TiO2 morphologies, i.e., single crystal, polycrystalline thin films, 25-nm-sized particles containing a mixture of anatase (70%) and rutile (30%), and colloidal 9-nm-sized particles, are considered here. The constraint here is that the experiments are not identical for a strict comparison and conclusion. As the particle size is much higher than its exciton Bohr diameter of 1.5 nm, they can be considered as bulk particles with high density of defects as shown in Fig. 7. The diffusion coefficient of single crystal anatase was B0.5 cm2/s,42 which was reduced to 0.1 cm2/s for crystalline thin films grown on glass substrate. The diffusion coefficient of nanoparticles of size r25 nm is about four orders of magnitude lower than that of the single crystals. Though all of the particles considered here could be characterized as bulk crystals, the reported diffusion coefficients in single crystals and nanocrystals differ in four orders of magnitude from which it could be inferred that the mechanism of electron transport is different. The poor diffusion coefficient of nanocrystalline TiO2 has been explained under the framework of multiple trapping mechanism. Under this mechanism, the electron transport through energy states arises from the surface atoms are slowed down by trapping–detrapping events. The direct hopping, which is faster than multiple trapping, between localized states is neglected in this case. An electron normally pass through B107 trapping events in nanocrystalline materials which ultimately reduces the diffusion coefficient.43 The density of trap states influences the kinetics of electron transport.44 The photocarrier transport measurements through TiO2 nanoparticles showed a dispersive decay that usually occurs in materials of high trap density.45 The carrier drift lengths (B100 nm) and charge screening lengths in the transport measurements were much larger than the particle size thereby indicating that the electrons move through the mesoporous network by trap filling, i.e., trapping and detrapping events.45 Table II. Comparison of Electron Mobilities of Anatase TiO2 as a Function of Particle Size Morphology 3 Single crystal (0.6 mm ) 1 mm films on glass substrate (crystalline films) Degussa 25 (25 nm particles with 70% anatase and 30% rutile) Colloidal TiO2 nanoparticles of size 9 nm D (cm2/s) Mobility (cm2/Vs) Technique 0.5 0.1 3  104 9  104 20 4 0.01 0.034 Four probe dc resistivity42 Hall measurements41 Terahertz spectroscopy39 Microwave conductivity40 February 2009 295 Metal Oxides for Solar Cells Table III. Photovoltaic Performances of the DSCs Those Used Composite, Doped, and Core/Shell Electrodes Geometry JSC (mA/cm2) Composition TiO2-metal oxide composite Doped TiO2 TiO2 core/metal oxide shell structures 95TiO2–05ZrO250 90TiO2–10GeO257 80TiO2–20Nb2O549 99TiO2–01ZrO256 TiO2:0.1%Nb2O5w TiO2/ZrO251 TiO2/Nb2O552 TiO2/ZnO54 TiO2/SiO251 TiO2/Al2O351 TiO2/MgO53 TiO2/SrTiO355 11.2 11.4 6.83 16.5 3.1 9.1 11.4 11.7 10.6 12.1 11.7 10.2 (10.1) (9.3) (7.89) (15.6) (2.2) (9.1) (10.2) (13.2) (9.1) (9.1) (10.2) (10.5) VOC (mV) 670 721 645 715 841 675 732 620 710 760 720 708 (710) (700) (600) (700) (777) (735) (659) (490) (735) (735) (640) (650) FF 0.72 0.69 0.42 0.69 0.69 0.60 0.56 0.52 0.58 0.61 0.54 0.58 (0.68) (0.84) (0.29) (0.64) (0.54) (0.55) (0.51) (0.40) (0.55) (0.55) (0.47) (0.54) Z (%) 5.8 5.7 1.9 8.1 1.8 3.6 4.7 4.5 4.4 5.6 4.5 4.4 (4.6) (5.5) (1.4) (7.0) (1.0) (3.7) (3.5) (3.3) (3.7) (3.7) (3.1) (3.8) w This work. Values in parentheses show performance of the control experiment that used TiO2. DSCs, dye solar cells. The electron transport in DSC through the mesoporous electrode is coupled with the electrolyte as there is a junction capacitor in the metal oxide–electrolyte interface (Fig. 4). The capacitive charging at this junction further retard the electron mobility, i.e., the electron diffusion is coupled with the ion movement in the electrolyte. This is known as ambipolar diffusion, which typically occurs when an electrical carrier is injected into a sea of oppositely charged mobile carriers. Under such conditions the injected carrier is referred to as a minority carrier. The ambipolar diffusion coefficient (Damp) is defined as Damp ¼ ðn þ pÞ ðn=Dp Þ þ ðp=Dn Þ (6) where n and Dn are the electron density and diffusion coefficient, respectively; and p and Dp are the hole density and diffusion coefficient. The DampBDn for p  n. A packet of electrons generated by light in DSC will diffuse naturally, carrying along it a cloud of holes such that the entire packet remains electrically neutral. Though this fact was considered by a number of authors before the year 2000,46,47 the ambipolar diffusion model was first applied to DSC by Kopidakis et al.48 By using TiO2 particles of 15–20 nm, Kopidakis et al.48 measured an ambipolar diffusion coefficient of 2  104 cm2/s which was 15% less than the corresponding electron diffusion coefficient. (B) Electrically Modified TiO2 Photoelectrodes: Several studies have initiated to modify the electrical transport properties of TiO2 using composite electrodes,49,50 core/shell structures,51–55and doping.56 Photovoltaic properties of the resulting DSC are presented in Table III. Eguchi et al.49 prepared various compositions in the TiO2–Nb2O5 system starting from pure TiO2 to pure Nb2O5 and observed that the FF increase steadily Fig. 8. The J–V curves of DSCs fabricated using Nb51-doped TiO2 as photoelectrodes. with Nb2O5 content. Increase in the FF indicates that the loss due to carrier recombination decreased, i.e., increase in the shunt resistance as noted in Section II(3), with increase in the Nb2O5 content. The VOC of the composite electrode cell was increased by increasing the Nb2O5 content; and the JSC increased by increasing the TiO2 content. Increase in the VOC is due to the higher conduction band energy of Nb2O5 compared with TiO2. The BET surface area of the composite materials was smaller than TiO2, which adversely affected the amount of dye anchored onto the composite electrodes and lowered the JSC. More recently Kitiyanan and Yoshikawa50 prepared TiO2–ZrO2 and TiO2–GeO257 composite electrodes. The TiO2–ZrO2 composite electrodes showed increase in all the photovoltaic parameters, whereas the FF decreased in the TiO2–GeO2 composite compared with cells fabricated using pure TiO2. Another approach to improve the electrical transport properties was to form a shell of wider bandgap material to TiO2, which was an idea brought from the bandgap engineering of compound semiconductors. If the bandgap positions are in a typical way as shown in Fig. 8, the resulting configuration confine electrons in the core material. The electron confinement would decrease the electron transfer probability to the electrolyte and thereby increasing the FF and JSC. Palomares et al.51 coated layers of ZrO2, Al2O3, and SiO2 onto 15 nm TiO2 particles and observed significant increase in the FF and JSC for core/shell electrode compared with the bare TiO2 electrode. The best result was reported with optimum shell thickness for Al2O3 shell whose conduction band energy (B3.23 eV) lies between that of the ZrO2 (B3.13 eV) and SiO2 (B2.97 eV). Although the FF was increased with increase in the shell thickness, the JSC was decreased. Zaban et al.52 reported enhancement in FF and JSC for TiO2/Nb2O5 core/shell nanostructures. Ahn et al.58 recently reported diffusion coefficient and carrier lifetime of mesoporous TiO2 and TiO2–Nb2O5 core/shell structure using transient photocurrent measurements. The diffusion length calculated for B10 mm TiO2 film was B17.3 mm, which increased to 23.8 mm in the TiO2/Nb2O5 core/shell structure. Jung et al.53 coated layers of mesoporous MgO onto 25 nm-TiO2 particles and observed increase in FF and JSC. The surface area of the TiO2/MgO core/shell structures was higher than that of the bare TiO2, therefore, the enhanced JSC could result from the increased dye loading also. TiO2 particles were also coated using materials of similar bandgap such as ZnO and SrTiO3 and observed higher FF for the core/shell structure.54,55 The VOC increased in this case also (Table III). In the case of higher bandgap shell materials, the increased VOC results from the higher conduction band energy of the shells. For materials of similar band energy such as ZnO, it should be anticipated that the dye anchoring up-shifted the conduction band energy.21 Durr et al.56 substituted Zr41 ions for Ti41 ions in TiO2 and observed that no secondary phase was segregated up to 2% of Zr in TiO2. The DSCs fabricated using Zr41:TiO2 showed 15% enhancement in efficiency compared with the pure TiO2. 296 Journal of the American Ceramic Society—Jose et al. Table IV. Photovoltaic Properties of Nb51-Doped TiO2 Electrode TiO2 TiO21500 ppm Nb51 TiO211000 ppm Nb51 JSC (mA/cm2) VOC (V) FF Z (%) 2.15 4.89 3.11 0.777 0.837 0.841 54.3 68.8 70.2 0.95 2.74 1.78 (a) Nb51-Doped TiO2: Following Durr et al.,56 we have doped Nb51 in TiO2 and used the resultant material as photoelectrode. Pure TiO2 and Nb51:TiO2 were synthesized from hydrolysis of titanium isopropoxide and a mixture of titanium isopropoxide and niobium ethoxide using acetic acid, respectively. The amount of Nb atoms was fixed at 500 and 1000 ppm. All the solutions yielded transparent gel upon evaporation. The gels were dried at room temperature and heated at 5001C for 30 min. A complete characterization of the powders thus obtained was pending at the time of writing this report. The TiO2 particles thus obtained were sprayed on FTO plates (area B0.28 cm2) and sintered at 5501C for 1 h. The TiO2coated FTO electrode was soaked in a solution containing ruthenium dye (RuL2(NCS)2  2H2O; L 5 2.20 -bipyridyl-4, 40 -dicarboxylic acid, 0.5 mM, N3 Solaronix, Aubonne, VD, Switzerland) in 1:1 volume mixture of acetonitrile and tertbutanol for 12 h at room temperature for dye anchoring. The soaked electrodes were washed with ethanol to remove nonanchored dye molecules and subsequently dried in air. The Pt-sputtered FTO glasses were used as counter electrodes. The Pt counter electrode and dye-anchored TiO2 electrodes were assembled into a sealed sandwich type cell using a sealing material (SX1170-25, film thickness 5 25 mm, Solaronix). The electrolyte solution, acetonitrile containing 0.1M lithium iodide, 0.03M iodine, 0.5M 4-tert-butylpyridine, and 0.6M 1-propyl-2,3-dimeth- Vol. 92, No. 2 yl imidazolium iodide, was injected into the cell through a hole drilled on the counter electrode via vacuum backfilling. The cell was first placed in a small vacuum chamber to remove inside air. By exposing the cell again to ambient pressure causes the electrolyte to be driven into the cell. The hole was then sealed with a piece of glass using epoxy adhesive. Performance of the solar cell was evaluated using the photocurrent measurements with a 100 mW/cm2 xenon lamp (Thermo Oriel Xenon Lamp 150 W: Model 66902, Stratford, CT) under global AM1.5 condition. Current density–voltage curves of the cells were obtained by using a potentiostat (PGSTAT30, Autolab PGSTAT30, Eco Chemie B.V. Utrecht, the Netherlands). Figure 8 shows typical I–V characteristic of the DSCs fabricated in the present work. Table IV shows the photovoltaic properties of the DSCs determined from the I–V curves. The efficiency of the cells was improved for 500 and 1000 ppm doping of Nb51 ions in TiO2. As anticipated, the VOC increased for both the sample containing Nb51 for 500 and 1000 ppm, indicating incorporation of Nb51 in TiO2 lattice. Both the fill factor and current density increased for 500 ppm Nb51 doping and decreased for 1000 ppm Nb51 ion doping. Impressive increments in photovoltaic parameters were observed for 500 ppm Nb51 doping compared with pure TiO2 cells under similar experimental conditions. Detailed investigation along this line is currently underway. (C) One-Dimensional TiO2 Nanostructures: The enhanced grain-boundary density and random network of nanoparticles increase the loss in DSC due to carrier recombination. For the usual nanoparticle network used in DSC, with porosity B60%, the average diffusion length has been estimated to be 15–20 mm.43 This inferior electron transport limits the electrode thickness for DSC and adversely affects the final conversion efficiency. One possible way to improve the diffusion coefficient, and therefore, transport properties is to grow the electrode materials as one-dimensional nanostructures such as nanotubes Fig. 9. Different particle morphologies (A) particles showing a bimodal distribution of size, (B) nanotubes adapted from Palomares et al.,51 (C) random nanofiber network, (D) nanofibers deposited onto FTO plate. February 2009 Metal Oxides for Solar Cells 297 nanofibers on the FTO plate.70 This technique allowed us fabrication of large area ( 20 cm2) TiO2 films on FTO plates. In addition, TiO2 nanofibers were directly developed on conducting glass plates by layer deposition method.71 The DSCs fabricated using the electrospun nanofibers gave efficiencies B6% using conventional N3 dyes. We have also tested metal-free indoline dyes as sensitizers in electrospun TiO2 nanofibers and found that Z of some of the indoline DSCs are comparable to that of the N3 dye.72 The nanorod-based DSC was further used to understand deeply the controlling factors of the JSC; and therefore, the Z.22 Through a combined theoretical and experimental investigation using TiO2 nanofibers we have shown that the electron injection yields from the dye to the photoelectrode is a factor that limits JSC and Z.22 The efficiency of charge injection in DSCs; and therefore JSC and Z, are determined by the extent to which the LUMO of the dye lie close to the photoelectrode. Fig. 10. Schematic of electron diffusion between (A) sintered spherical nanoparticles and (B) one-dimensional (1-D) nanostructures. Electron diffusion through spherical nanoparticles is according to random walk model. Though diffusion is in accordance with the random walk in 1-D structures also, the electrons are constrained to move directionally. The panel ‘‘C’’ shown in the box explains how reduction of diameter of the 1-D structure make the electron flow more channeled. and nanorods,59 nanowires,60 and nanofibers61 (Fig. 9). Figure 10 explains this concept of channeled electron transport through one-dimensional nanostructures. Recently Frank and colleagues62,63 compared the transport and recombination properties of nanotubes and spherical nanoparticles films used in DSCs by frequency-resolved photovoltage spectroscopy. Both morphologies displayed comparable transport times; however, recombination was much slower in the nanotube films. This observation indicates that the nanotube-based DSCs have significantly higher charge collection efficiencies than their nanoparticle-based counterparts.63 It was also showed that capillary stress in the dense TiO2 nanotube arrays could induce bundling and microcrack formation, which would affect the dye loading and electron transport.62 Transport became significantly faster when the bundling of nanotubes was reduced. The photoconversion efficiency above B9.3% has been reported for DSCs using single crystal like anatase nanowires.64 The DSCs fabricated using TiO2 nanotubes made by anodic oxidation60,65 and using block copolymer templates66 gave Z up to 5.4% and 10%, respectively. Most frequently used method for fabrication of nanotubes is the anodic oxidization of titanium.65 TiO2 nanowires are fabricated from different methods such as depositing TiO2 into anodic alumina membranes, electrophoresis, hydrothermal method, and oxidation of metallic titanium.60 Lengths of the nanotubes and nanowires thus obtained are limited to few micrometers. This drawback is eliminated in nanofibers which are normally produced in long lengths using a relative simple technique, known as electrospinning.67 The diameter controlled anatase TiO2 nanofibers are fabricated by controlled electrospinning of a polymeric solution and subsequent sintering of the as-spun fibers.68 The electrospun TiO2 nanofibers are polycrystalline and composed of densely packed TiO2 grains of several nanometers diameters. We have shown that the nanofibers are characterized by increased surface strain from capillary force, the magnitude of which increases with decrease in the fiber diameter.69 The surface strain in anatase TiO2 caused the rutile phase to nucleate at the grain boundaries at lower temperatures than that has been observed in nanoparticles of similar size.69 Unlike other one-dimensional nanostructures such as nanowires and nanotubes, which are grown directly onto the FTO plate, the random web structures of electrospun nanofiber impose adhesion difficulties with the FTO plate. We developed a nanorod spraying technique to solve the adhesion issues of the (2) ZnO Electrodes The ZnO is II–VI semiconductor (EgB3.37 eV) that crystallizes either in cubic zinc-blende, cubic rocksalt (NaCl), or hexagonal wurtzite structure. The zinc-blende phase can be stabilized only by growth on cubic substrates and the rocksalt structure forms at relatively high pressures.73 The hexagonal wurtzite is the lowest energy structure of ZnO. The conduction band energy of ZnO is 4.45 eV with respect to vacuum, similar to that of TiO2. Hexagonal ZnO has been actively sought as a replacement for the TiO2 electrodes.74–77 Hagfeldt and colleagues78 studied the electron transport in nanostructured and discussed their finding under the framework of the multiple trapping diffusion model similar to that applied for TiO2. The charge transport in ZnO also depended on the electrolyte used. For nanoporous ZnO in 0.5M LiClO4 in ethanol, the diffusion coefficient varied from 104 to 106 cm2/s when the potential was varied between 100 and 300 mV vs Ag/AgCl in ethanol. On the other hand diffusion coefficient for TiO2 in 0.7M LiClO4 in ethanol at 300 mV vs Ag/AgCl in ethanol was B2  105 cm2/s.46 The density of traps is reported to be less for ZnO compared with that of TiO2. This lower trap density makes the electrons to move faster through the ZnO nanoporous network. The carrier life time in ZnO is longer than that in TiO2.78 Despite the superior transport properties of ZnO, the DSCs constituted using ZnO gave inferior performance. The best efficiency so far reported for ZnO electrode is 4.1% by using with N719 sensitizer and iodide/triiodide electrolyte.79 The poor photovoltaic performance utilizing ZnO electrode comes from the stability of ZnO at acidic conditions. The protons derived from Ru complexes such as N3 and N719 make the dye loading solution relatively acidic and dissolve the surface of ZnO, generating Zn21/dye aggregates.80 Such aggregates lower the electron injection efficiency. This could be overcome by immersing the ZnO film in ethanolic solution under reflux. Besides nanoparticles, various one-dimensional nanostructure of ZnO including nanowires,81–83 nanotubes,84,85 nanorods,86 as well as electrospun ZnO nanofibers87 has been explored. From resistance measurements and by assuming electron concentration of 1–5  1018 cm3 and mobility 1–5  1018 cm2/V/s, Law et al.81 calculated diffusion coefficient of single crystalline ZnO nanowires as 0.05–0.5 cm2/s. This diffusion coefficient of onedimensional ZnO nanowires is several hundred times larger than that reported for TiO2 and ZnO nanoparticles. Pillarshaped and branched-shaped ZnO nanowire grown directly on FTO substrate by thermal evaporation at temperatures of 8001–10001C were also suggested for channeled electron transport similar to one-dimensional nanostructures.88 Compared with the pillar-shaped nanowire efficiency of 0.34%, a branched ZnO nanowire structure has better conversion efficiency of 0.46%. This is suggested by the increase in the dye absorption of the high-density branched structure. Owing to the fact that many branches can be formed from a single pillar nanowire, branched structures provide a much larger surface area for dye 298 Journal of the American Ceramic Society—Jose et al. loading. ZnO nanosheets were also grown from precursor templates in aqueous solutions.89 Nanosheets fabricated were accumulated on conducting glass substrates with submicrometer spacing, giving the film a macroporous nature. A high shortcircuit photocurrent JSC of 13.8 mA/cm2 is thus obtained. This was attributed to the effective dye loading of 1.1  107 mol/cm2, which is comparable to 1.3  107 mol/cm2 of a 10-mm-thick nanoporous TiO2 photoelectrode sensitized with the N3 dyes.88 (3) Nb2O5 Electrodes Several groups used Nb2O5 as photoelectrodes in DSSC as nanoparticles,49,90–93 nanobelts,94 and as TiO2–Nb2O5 bilayers,52,58 and as blocking layers that prevent the electron back transfer.95,96 The Nb2O5 is a wide bandgap semiconductor with bandgap energy 3.49 eV,97which is about 0.29 eV larger than that of anatase. Lenzmann et al.91 determined the conduction band energy of Nb2O5 from electrochemical measurements and assuming one-electron configuration to be at 0.2–0.3 eV above to that of anatase. Because of its larger bandgap and higher conduction band edge compared with anatase, DSCs fabricated using Nb2O5 as photoelectrode and N3 dye showed higher VOC than that achieved using anatase. However, Nb2O5 sensitized by mercurochrome dye showed VOC lower than that achieved using anatase.98 Hoshikawa et al.90 compared the impedance characteristics of DSCs using Nb2O5 photoelectrode with that of TiO2. They report that when the Nb2O5 particles with large BET surface area were used as the electrode, the internal resistance of the solar cell was lower. Sayama et al.99 used porous semiconductor films of TiO2, Nb2O5, ZnO, SnO2, In2O3, WO3, Ta2O5, and ZrO2 in DSC sensitized by the N3 dye. They observed that the Nb2O5 semiconductor cell had the next highest IPCEB18% compared with the TiO2 cell, for which IPCEB45% was reported under similar cell fabrication conditions, and showed the highest VOC among them. This observation was explained on the difference in the electronic structure of these semiconductors. The electrons are assumed to be transferred mainly through the conjugated orbitals of the ester linkage and semiconductor conduction band. The Nb2O5 has a monoclinic structure (space group P2) with a 5 2.038 nm, b 5 0.3824 nm, c 5 1.937 nm, and b 5 115.691.100 Orthorhombic Pbam symmetry with a 5 0.6168 nm, b 5 2.931 nm, and c 5 0.3936 nm has also been assigned to Nb2O5.101 Its large unit cell dimension could reduce the surface area thereby lowering JSC as has been observed experimentally. However, by developing proper chemical technique nanocrystalline Nb2O5 with BET surface area 25 g/cm2 could be prepared,28 even if it is much lower than that of mesoporous TiO2. More attention is required to develop mesoporous Nb2O5 as well as their onedimensional nanostructures. Similar to other materials, actual mechanism of charge transport through mesoporous Nb2O5 is to be explored. In view of its higher conduction band energy and high chemical stability, Nb2O5 could be a suitable alternative to TiO2 on the way of development of high-performance DSC. (4) Ternary Photoelectrode Materials Besides the simple binary metal oxide systems ternary metal oxide systems such as SrTiO391,102 and Zn2SnO4103,104 have been considered as photoelectrode materials in DSC. The SrTiO3 is a semiconductor with bandgap similar to TiO2 (3.2 eV). However, its conduction band is relatively higher than that of TiO2 which could result in a higher VOC compared with TiO2.91 Its high dielectric constant makes SrTiO3 to behave as electrically mesoporous even with a large particle size of B80 nm.105 The Z reported for DSCs using SrTiO3 is relatively poor, which is explained as due to the specific surface structure of SrTiO3. The energetically favored surface of SrTiO3 is (001) face, which has two possible terminations, SrO or TiO2. The SrO surface is more basic; and therefore, poorly adsorb the negatively charged carboxylate group of the sensitizer. The poor photovoltaic performance might result from the decreased dye loading onto the semiconductor. Vol. 92, No. 2 Two groups reported Zn2SnO4 as a photoelectrode in DSC almost same time.103,104 The Zn2SnO4 is a TCO with bandgap B3.6 eV and has inverse spinel structure with octahedrally coordinated Sn41 atoms; half of the Zn21 atoms are in tetrahedral coordination and the other half in octahedral coordination. Research on Zn2SnO4 initiated following an observation that Z of a DSC that used mixture of ZnO and SnO2 as working electrode enhanced dramatically.106 The Zn2SnO4 cells reported efficiency up to 3.8% which is close to the highest efficiency reported for ZnO fabricated under similar conditions (4.1%) and much higher than that reported for SnO2 (1.2%). In addition Zn2SnO4 cells also overcame the stability problem associated with ZnO against acidic dyes. In this sense, the ternary oxide (Zn2SnO4) is more attractive than its simple binary components (ZnO and SnO2) as the electrode material for DSSC. (5) Metal Oxide Solar Cells Development of photocurrents upon irradiation of a semiconductor electrode has long been observed in TiO2, SnO2, SrTiO3, Fe2O3, etc.107,108 If such electrodes are coupled with a metal counter electrode (Schottky barrier) or with another semiconductor electrode, the system produce photoelectric current. However, such metal oxide solar cells showed poor efficiency typically with JSC few hundred microamperes at 1 sun irradiation because these materials are characterized by poor absorption coefficient and low carrier mobility (B0.01 cm2/Vs). A major source that limits the JSC is the charge carrier recombination in metal oxides.109 Recently there is a revived interest in these materials for fabrication of metal oxide solar cells.110,111 Corma et al.111 observed photovoltaic effect in CeO2, Ce1xZrxO2, and Ce1xLaxO2 (x 5 0–0.3) nanoparticles of size B5 nm when they are sandwiched between two conducting glass plates and in the presence of iodide/triiodide electrolyte. The cells were not sensitized with a dye. The CeO2 has a bandgap of 3.2 eV arising from O2p-Ce4f transition; and therefore, bulk CeO2 is not regard as a photoactive semiconductor. However, nanocrystals of pure CeO2 and that doped with Zr41 and La31 exhibit a photovoltaic response.111 In the case of nanocrystalline CeO2, the presence of significant fraction of Ce atoms lie at the surface thereby leading to oxygen vacancies and defects. The observed photoconductivity could be explained in terms of hopping of electron and holes between these surface states. The cells prepared with La31-doped CeO2 gave VOC 5 0.9 V, Isc 5 1.5 mA, fill factor 0.41, and energy conversion efficiency of 1.4% under AM1.5 condition. IV. Conclusions The nanostructured metal oxide semiconductors, with particle size greater than their exciton Bohr radius, influences the photoconversion efficiency of DSCs. Among the beneficial include (i) its mesoporosity and large surface area allow anchoring a large amount of dye molecules which enhances the absorption cross-section, (ii) larger density of states in metal oxides than the molecular orbital of dye allows faster injection of electrons from the dye to the metal oxide. The detrimental factors are (i) inefficient electrical transport through nanocrystalline network and (ii) charge recombination within the materials and also with the electrolyte. Inefficient electrical transport in nanostructured metal oxide semiconductors arise from the trapping and detrapping of electrons at the surface atomic states in the electronic band. The surface atoms are a large fraction in nanostructured materials, so as the trap density. Although considerable progress has been made in understanding the charge transport through nanoparticles network, further efforts are required to exactly determine the trap states in TiO2 nanoparticles of size 10–25 nm. Trapping and detrapping reduces the kinetic energy of the flowing electrons, which ultimately results in inferior cell performance. Quantification of the traps and their subsequent removal could enhance the photoelectric conversion efficiency of DSCs beyond the present record of B11%. Geometry of DSCs is well February 2009 Metal Oxides for Solar Cells optimized such that over 85% electrons converted from the incident photons are collected at the external electrode, which makes the collection of high-energy electrons a more stringent requirement. Binary metal oxides such as TiO2, ZnO, Nb2O5, Fe2O3, ZrO2, Al2O3, and CeO2 and ternary compounds such as SrTiO3 and Zn2SnO4 have been tested for their use as photoelectrodes in DSC. Among them the d-block binary metal oxides viz., TiO2, ZnO, and Nb2O5 are the best candidates as photoelectrode due to the dissimilarity in orbitals constituting their conduction band and valence band. Various morphologies and approaches, such as doping, composites, core/shell structures, have been explored to enhance the electrical transport properties. Onedimensional morphologies show great promise in enhancing the electrical transport properties. Well-ordered one-dimensional nanowires synthesized through template-assisted methods increased the recombination time and allowed enhanced charge collection efficiency in DSCs. Doping of Nb2O5 and ZrO2 in TiO2 is found to be beneficial in increasing the open circuit voltage; however, the doped materials showed lesser surface area than their parent materials. The reduced surface area adversely affected the dye loading; and therefore, the short-circuit current density. Owing to this drawback, the doped materials or composites made from the constituent metal oxides did not yield high-efficiency DSCs. Efforts are to be undertaken to synthesize mesoporous Nb2O5 and Nb-doped TiO2 with high crystallinity to realize high-performance DSCs. In addition, cost-effective TCOs that are stable over a wide range of temperature is required for the commercial exploitation of DSCs. References 1 IPCC Panel Reports on Climate Change and Biodiversity, April 2002. N. S. Lewis, G. W. Crabtree, and A. J. Nozik., et al (eds), Basic Research Needs for Solar Energy Utilization, Basic Energy Sciences Workshop on Solar Energy Utilization, April 18–21. 2005. 3 M. Grätzel, ‘‘Photoelectrochemical Cells,’’ Nature, 414, 338–44 (2003). 4 B. A. Gregg, ‘‘Excitonic Solar Cells,’’ J. Phys. Chem. B, 107, 4688–98 (2003). 5 G. Yu, J. Gao, J. C. Hummelen, F. Wudl, and A. J. Heeger, ‘‘Polymer Photovoltaic Cells—Enhanced Efficiencies via a Network of Internal Donor–Acceptor Heterojunctions,’’ Science, 270 [5243] 1789–91 (1995). 6 P. Yu, K. Zhu, A. G. Norman, S. Ferrere, A. J. Frank, and A. J. Nozik, ‘‘Nanocrystalline TiO2 Solar Cells Sensitized with InAs Quantum Dots,’’ J. Phys. Chem. B, 110 [50] 25451–4 (2006). 7 M. A. Green, K. Emery, D. L. King, Y. Hoshikawa, and W. Warta, ‘‘Solar Cell Efficiency Tables (Version 29),’’ Prog. Photovolt: Res. Appl., 15, 35–40 (2007). 8 J. Moser, ‘‘Notiz über Verstärkung photoelektrischer Ströme Durch Optische Sensibilisirung,’’ Monatshefte für Chemie (Chemical Monthly), 8 [1] 373–3 (1887). 9 R. Memming and F. Schröppel, ‘‘Electron Transfer Reactions of Excited Ruthenium(II) Complexes in Monolayer Assemblies at the SnO2–Water Interface,’’ Chem. Phys. Lett., 62 [2] 207–10 (1979). 10 B. O. Regan and M. Grätzel, ‘‘A Low-Cost, High Efficiency Solar Cell Based on Dye-Sensitized Colloidal TiO2 Films,’’ Nature, 353, 737–9 (1991). 11 R. D. McConnell, ‘‘Assessment of the Dye-Sensitized Solar Cell,’’ Renewable Sustainable Energy Rev., 6 [3] 271–93 (2002). 12 Md K. Nazeeruddin, P. Pechy, T. Renouard, S. M. Zakeeruddin, R. Humphry-Baker, P. Comte, P. Liska, L. Cevey, E. Costa, V. Shklover, L. Spiccia, G. B. Deacon, C. A. Bignozzi, and M. Grätzel, ‘‘Engineering of Efficient Panchromatic Sensitizers for Nanocrystalline TiO2-Based Solar Cells,’’ J. Am. Chem. Soc., 123 [8] 1613–24 (2001). 13 M. Grätzel, ‘‘Solar Energy Conversion by Dye-Sensitized Photovoltaic Cells,’’ Inorg. Chem., 44 [20] 6841–51 (2005). 14 Y. Chiba, A. Islam, Y. Watanabe, R. Komiya, N. Koide, and L. Y. Han, ‘‘Dye-Sensitized Solar Cells with Conversion Efficiency of 11.1%,’’ Jpn. J. Appl. Phys. Part 2, 45, L638 (2006). 15 G. Smestad, C. Bignozzi, and R. Argazzi, ‘‘Testing of Dye Sensitized TiO2 Solar Cells I: Experimental Photocurrent Output and Conversion Efficiencies,’’ Sol. Energy Mater. Sol. Cells, 32 [3] 259–72 (1994). 16 K. Yohei, Y. Yorikazu, and H. Shuzi, ‘‘All-Metal-Electrode-Type Dye Sensitized Solar Cells (Transparent Conductive Oxide-Less Dye Sensitized Solar Cell) Consisting of Thick and Porous Ti Electrode with Straight Pores,’’ Appl. Phy. Lett., 92 [3] 033308 (2008). 17 R. Neil, ‘‘Optimizing Dyes for Dye-Sensitized Solar Cells,’’ Angew. Chem., 45 [15] 2338–45 (2006). 18 H. J. Snaith and L. Schmidt-Mende, ‘‘Advances in Liquid-Electrolyte and Solid-State Dye-Sensitized Solar Cells,’’ Adv. Mater., 19, 3187–200 (2007). 19 J. Ferber, R. Stangl, and J. Luther, ‘‘An Electrical Model of the Dye-Sensitized Solar Cell,’’ Sol. Energy Mater. Sol. Cells, 53 [1–2] 29–54 (1998). 20 , ‘‘Reporting Solar Cell Efficiencies in Solar Energy Materials and Solar Cells (Editorial),’’ Sol. Energy Mater. Sol. Cells, 92, 371–3 (2008). 2 21 299 F. De Angelis, S. Fantacci, A. Selloni, M. Grätzel, and M. K. Nazeeruddin, ‘‘Influence of the Sensitizer Adsorption Mode on the Open-Circuit Potential of Dye-Sensitized Solar Cells,’’ Nano Lett., 7, 3189–95 (2007). 22 R. Jose, A. Kumar, V. Thavasi, K. Fujihara, S. Uchida, and S. Ramakrishna, ‘‘Relationship Between the Molecular Structure of the Dyes and Photocurrent Density in the Dye-Sensitized Solar Cells,’’ Appl. Phy. Lett., 93, 023125 (2008). 23 J. Bisquert, F. Fabregat-Santiago, I. Mora-Sero, G. Garcia-Belmonte, E. M. Barea, and E. Palomares, ‘‘A Review of Recent Results on Electrochemical Determination of the Density of Electronic States of Nanostructured MetalOxide Semiconductors and Organic Hole Conductors,’’ Inorg. Chim. Acta, 361 [3] 684–98 (2008). 24 J. Bisquert, ‘‘Physical Electrochemistry of Nanostructured Devices,’’ Phys. Chem. Chem. Phys., 10 [1] 49–72 (2008). 25 Q. Wang, S. Ito, M. Grätzel, F. Fabregat-Santiago, I. Mora-Sero, J. Bisquert, T. Bessho, and H. Imai, ‘‘Characteristics of High Efficiency Dye-Sensitized Solar Cells,’’ J. Phys. Chem. B, 110 [50] 25210–21 (2006). 26 F. Fabregat-Santiago, J. Bisquert, G. Garcia-Belmonte, G. Boschloo, and A. Hagfeldt, ‘‘Influence of Electrolyte in Transport and Recombination in Dye-Sensitized Solar Cells Studied by Impedance Spectroscopy,’’ Sol. Energy Mater. Sol. Cells, 87 [1–4] 117–31 (2005). 27 D. Cahen, G. Hodes, M. Grätzel, J. F. Guillemoles, and I. Riess, ‘‘Nature of Photovoltaic Action in Dye-Sensitized Solar Cells,’’ J. Phys. Chem. B, 104 [9] 2053–9 (2000). 28 M. A. Aegerter, ‘‘Sol–Gel Niobium Pentoxide: A Promising Material for Electrochromic Coatings, Batteries, Nanocrystalline Solar Cells and Catalysis,’’ Sol. Energy Mater. Sol. Cells, 68, 401–22 (2001). 29 N. A. Dubrovinskaia, L. S. Dubrovinsky, R. Ahuja, V. B. Prokopenko, V. Dmitriev, H. P. Weber, J. M. Osorio-Guillen, and B. Johansson, ‘‘Experimental and Theoretical Identification of a New High-Pressure TiO2 Polymorph,’’ Phys. Rev. Lett., 87 [27] 275501 (2001). 30 J. Haines and J. M. Léger, ‘‘X-Ray Diffraction Study of TiO2 up to 49 GPa,’’ Physica B: Condensed Matter, 192 [3] 233–7 (1993). 31 H. Sato, S. Endo, M. Sugiyama, T. Kikegawa, O. Shimomura, and K. Kusaba, ‘‘Baddeleyite-Type High-Pressure Phase of TiO2,’’ Science, 251 [4995] 786–8 (1991). 32 M. Mattesini, J. S. de Almeida, L. Dubrovinsky, N. Dubrovinskaia, B. Johansson, and R. Ahuja, ‘‘High-Pressure and High-Temperature Synthesis of the Cubic TiO2 Polymorph,’’ Phys. Rev. B, 70 [21] 212101 (2004). 33 M. K. Nowotny, L. R. Sheppard, T. Bak, and J. Nowotny, ‘‘Defect Chemistry of Titanium Dioxide. Application of Defect Engineering in Processing of TiO2-Based Photocatalysts,’’ J. Phys. Chem. C, 112 [14] 5275–300 (2008). 34 N. G. Park, J. van de Lagemaat, and A. J. Frank, ‘‘Comparison of DyeSensitized Rutile- and Anatase-Based TiO2 Solar Cells,’’ J. Phys. Chem. B, 104 [38] 8989–94 (2000). 35 N. G. Park, G. Schlichthorl, J. van de Lagemaat, H. M. Cheong, A. Mascarenhas, and A. J. Frank, ‘‘Dye-Sensitized TiO2 Solar Cells: Structural and Photoelectrochemical Characterization of Nanocrystalline Electrodes Formed from the Hydrolysis of TiCl4,’’ J. Phys. Chem. B, 103 [17] 3308–14 (1999). 36 C. J. Barbe, F. Arendse, P. Comte, M. Jirousek, F. Lenzmann, V. Shklover, and M. Gratzel, ‘‘Nanocrystalline Titanium Oxide Electrodes for Photovoltaic Applications,’’ J. Am. Ceram. Soc., 80 [12] 3157–71 (1997). 37 D. F. Watson and G. J. Meyer, ‘‘Cation Effects in Nanocrystalline Solar Cells,’’ Coord. Chem. Rev., 248, 1391–406 (2004). 38 R. Jose, N. U. Zhanpeisov, H. Fukumura, Y. Baba, and M. Ishikawa, ‘‘Structure–Property Correlation of CdSe Clusters Using Experimental Results and First-Principles DFT Calculations,’’ J. Am. Ceram. Soc., 128 [2] 629–36 (2006). 39 E. Hendry, M. Koeberg, B. O’Regan, and M. Bonn, ‘‘Local Field Effects on Electron Transport in Nanostructured TiO2 Revealed by Terahertz Spectroscopy,’’ Nano Lett., 6 [4] 755–9 (2006). 40 J. E. Kroeze, T. J. Savenije, and J. M. Warman, ‘‘Electrodeless Determination of the Trap Density, Decay Kinetics, and Charge Separation Efficiency of DyeSensitized Nanocrystalline TiO2,’’ J. Am. Ceram. Soc., 126 [24] 7608–18 (2004). 41 H. Tang, K. Prasad, R. Sanjines, P. E. Schmid, and F. Levy, ‘‘Electrical and Optical Properties of TiO2 Anatase Thin Films,’’ J. Appl. Phys., 75 [4] 2042–7 (1994). 42 L. Forro, O. Chauvet, D. Emin, L. Zuppiroli, H. Berger, and F. Levy, ‘‘High Mobility n-Type Charge Carriers in Large Single Crystals of Anatase (TiO2),’’ J. Appl. Phys., 75 [1] 633–5 (1994). 43 K. D. Benkstein, N. Kopidakis, J. van de Lagemaat, and A. J. Frank, ‘‘Influence of the Percolation Network Geometry on Electron Transport in Dye-Sensitized Titanium Dioxide Solar Cells,’’ J. Phys. Chem. B, 107 [31] 7759–67 (2003). 44 K. Schwarzburg and F. Wiliig, ‘‘Influence of Trap Filling on Photocurrent Transients in Polycrystalline TiO2,’’ Appl. Phys. Lett., 58, 2520–2 (1991). 45 R. Könenkamp, R. Henninger, and P. Hoyer, ‘‘Photocarrier Transport in Colloidal TiO2 Films,’’ J. Phys. Chem., 97 [28] 7328–30 (1993). 46 A. Solbrand, H. Lindstrom, H. Rensmo, A. Hagfeldt, S. E. Lindquist, and S. Sodergren, ‘‘Electron Transport in the Nanostructured TiO2–Electrolyte System Studied with Time-Resolved Photocurrents,’’ J. Phys. Chem. B, 101 [14] 2514–8 (1997). 47 L. Dloczik, O. Ileperuma, I. Lauermann, L. M. Peter, E. A. Ponomarev, G. Redmond, N. J. Shaw, and I. Uhlendorf, ‘‘Dynamic Response of Dye-Sensitized Nanocrystalline Solar Cells: Characterization by Intensity-Modulated Photocurrent Spectroscopy,’’ J. Phys. Chem. B, 101 [49] 10281–9 (1997). 48 N. Kopidakis, E. A. Schiff, N. G. Park, J. van de Lagemaat, and A. J. Frank, ‘‘Ambipolar Diffusion of Photocarriers in Electrolyte-Filled, Nanoporous TiO2,’’ J. Phys. Chem. B, 104 [16] 3930–6 (2000). 49 K. Eguchi, H. Koga, K. Sekizawa, and K. Sasaki, ‘‘Nb2O5-Based Composite Electrodes for Dye-Sensitized Solar Cells,’’ J. Ceram. Soc. Jpn., 108, 1067–71 (2000). 300 50 Journal of the American Ceramic Society—Jose et al. A. Kitiyanan and S. Yoshikawa, ‘‘The Use of ZrO2 Mixed TiO2 Nanostructures as Efficient Dye-Sensitized Solar Cells’ Electrodes,’’ Mater. Lett., 59, 4038–40 (2005). 51 E. Palomares, J. N. Clifford, S. A. Haque, T. Lutz, and J. R. Durrant, ‘‘Control of Charge Recombination Dynamics in Dye Sensitized Solar Cells by the Use of Conformally Deposited Metal Oxide Blocking Layers,’’ J. Am. Ceram. Soc., 125 [2] 475–82 (2003). 52 A. Zaban, S. G. Chen, S. Chappel, and B. A. Gregg, ‘‘Bilayer Nanoporous Electrodes for Dye Sensitized Solar Cells,’’ Chem. Commun. (Cambridge), [22] 2231–2 (2000). 53 H. S. Jung, J. K. Lee, M. Nastasi, S. W. Lee, J. Y. Kim, J. S. Park, K. S. Hong, and H. Shin, ‘‘Preparation of Nanoporous MgO-Coated TiO2 Nanoparticles and Their Application to the Electrode of Dye-Sensitized Solar Cells,’’ Langmuir, 21 [23] 10332–5 (2005). 54 S. Roh, R. S. Mane, S. Min, W. Lee, C. D. Lokhande, and S. Han, ‘‘Achievement of 4.51% Conversion Efficiency Using ZnO Recombination Barrier Layer in TiO2 Based Dye-Sensitized Solar Cells,’’ Appl. Phy. Lett., 89, 253512 (2006). 55 Y. Diamant, S. G. Chen, O. Melamed, and A. Zaban, ‘‘Core–Shell Nanoporous Electrode for Dye Sensitized Solar Cells: The Effect of the SrTiO3 Shell on the Electronic Properties of the TiO2 Core,’’ J. Phys. Chem. B, 107 [9] 1977–81 (2003). 56 M. Durr, S. Rosselli, A. Yasuda, and G. Nelles, ‘‘Band-Gap Engineering of Metal Oxides for Dye-Sensitized Solar Cells,’’ J. Phys. Chem. B, 110B, 21899 2006. Erratum in J Phys Chem B Condens Matter Mater Surf Interfaces Biophys 110 [51] 26507. 57 A. Kitiyanan, T. Kato, Y. Suzuki, and S. Yoshikawa, ‘‘The Use of Binary TiO2–GeO2 Oxide Electrodes to Enhanced Efficiency of Dye-Sensitized Solar Cells,’’ J. Photochem. Photobiol. A, 179 [1–2] 130–4 (2006). 58 K. Ahn, M. Kang, J. Lee, B. Shin, and J. Lee, ‘‘Enhanced Electron Diffusion Length of Mesoporous TiO2 Film by Using Nb2O5 Energy Barrier for Dye-Sensitized Solar Cells,’’ Appl. Phys. Lett., 89 [1] 013103 (2006). 59 G. K. Mor, O. K. Varghese, M. Paulose, K. Shankar, and C. A. Grimes, ‘‘A Review on Highly Ordered, Vertically Oriented TiO2 Nanotube Arrays: Fabrication, Material Properties, and Solar Energy Applications,’’ Sol. Energy Mater. Sol. Cells, 90 [14] 2011–75 (2006). 60 M. Adachi, J. Jiu, and S. Isoda, ‘‘Synthesis of Morphology-Controlled Titania Nanocrystals and Application for Dye-Sensitized Solar Cells,’’ Curr. Nanosci., 3 [4] 285–95 (2007). 61 V. Thavasi, G. Singh, and S. Ramakrishna, ‘‘Electrospun Nanofibers in Energy and Environmental Applications,’’ Energy Environ. Sci., 1 [2] 205–21 (2008). 62 K. Zhu, T. B. Vinzant, N. R. Neale, and A. J. Frank, ‘‘Removing Structural Disorder from Oriented TiO2 Nanotube Arrays: Reducing the Dimensionality of Transport and Recombination in Dye-Sensitized Solar Cells,’’ Nano Lett., 7 [12] 3739–46 (2007). 63 K. Zhu, N. R. Neale, A. Miedaner, and A. J. Frank, ‘‘Enhanced ChargeCollection Efficiencies and Light Scattering in Dye-Sensitized Solar Cells Using Oriented TiO2 Nanotubes Arrays,’’ Nano Lett., 7 [1] 69–74 (2007). 64 M. Adachi, Y. Murata, J. Takao, J. Jiu, M. Sakamoto, and F. Wang, ‘‘Highly Efficient Dye-Sensitized Solar Cells with a Titania Thin-Film Electrode Composed of a Network Structure of Single-Crystal-Like TiO2 Nanowires Made by the ‘‘Oriented Attachment’’ Mechanism,’’ J. Am. Ceram. Soc., 126, 14943–9 (2004). 65 M. Paulose, K. Shankar, O. K. Varghese, G. K. Mor, and C. A. Grimes, ‘‘Application of Highly-Ordered TiO2 Nanotube-Arrays in Heterojunction Dye-Sensitized Solar Cells,’’ J. Phys. D: Appl. Phys., 39 [12] 2498–503 (2006). 66 M. Wei, Y. Konishi, H. Zhou, M. Yanagida, H. Sugihara, and H. Arakawa, ‘‘Highly Efficient Dye-Sensitized Solar Cells Composed of Mesoporous Titanium Dioxide,’’ J. Mater. Chem., 16, 1287–93 (2006). 67 R. Ramasheshan, S. Sundarrajan, R. Jose, and S. Ramakrishna, ‘‘Nanostructured Ceramics by Electrospinning,’’ J. Appl. Phys., 102 [12] 111101 (2007). 68 D. Li and Y. Xia, ‘‘Fabrication of Titania Nanofibers by Electrospinning,’’ Nano Lett., 3, 555–60 (2003). 69 A. Kumar, R. Jose, K. Fujihara, J. Wang, and S. Ramakrishna, ‘‘Structural and Optical Properties of Electrospun TiO2 Nanofibers,’’ Chem. Mater., 19, 6536–42 (2007). 70 K. Fujihara, A. Kumar, R. Jose, S. Ramakrishna, and S. Uchida, ‘‘Spray Deposition of Electrospun TiO2 Nanorods for Dye-Sensitized Solar Cell,’’ Nanotechnology, 18 [36] 365709 (2007). 71 Z. Rui, C. Y. Jiang, X. Liu, B. Liu, A. Kumar, and S. Ramakrishna, ‘‘Improved Adhesion of Interconnected TiO2 Nanofiber Network on Conductive Substrate and Its Application in Polymer Photovoltaic Devices,’’ Appl. Phys. Lett., 93 [1] 013102 (2008). 72 R. Jose, A. Kumar, V. Thavasi, and S. Ramakrishna, ‘‘Conversion Efficiency Versus Sensitizer for Electrospun Nanofibers for the Dye-Sensitized Solar Cells,’’ Nanotechnology, 19, 424004 (2008). 73 U. Ozgur, Y. I. Alivov, C. Liu, A. Teke, M. A. Reshchikov, S. Dogan, V. Avrutin, S. J. Cho, and H. Morkoc, ‘‘A Comprehensive Review of ZnO Materials and Devices,’’ J. Appl. Phys., 98 [4] 041301 (2005). 74 G. Redmond, D. Fitzmaurice, and M. Grätzel, ‘‘Visible Light Sensitization by Cis-Bis (Thiocyanato)Bis(2,20 -bipyridyl-4,40 -Dicarboxylato)Ruthenium (II) of Transparent Nanocrystalline ZnO Films Prepared by Sol–Gel Techniques,’’ Chem. Mater., 6 [5] 686–91 (1994). 75 H. Rensmo, K. Keis, H. Lindstrom, S. Sodergren, A. Solbrand, A. Hagfeldt, S. E. Lindquist, L. N. Wang, and M. Muhammed, ‘‘High Light-to-Energy Conversion Efficiencies for Solar Cells Based on Nanostructured ZnO Electrodes,’’ J. Phys. Chem. B, 101 [14] 2598–601 (1997). 76 Vol. 92, No. 2 T. N. Rao and L. Bahadur, ‘‘Photoelectrochemical Studies on Dye-Sensitized Particulate ZnO Thin-Film Photoelectrodes in Nonaqueous Media,’’ J. Electrochem. Soc., 144 [1] 179–85 (1997). 77 K. Keis, L. Vayssieres, S. E. Lindquist, and A. Hagfeldt, ‘‘Nanostructured ZnO Electrodes for Photovoltaic Applications,’’ Nanostruct. Mater., 12 [1–4] 487–90 (1999). 78 A. Solbrand, K. Keis, S. Södergren, H. Lindström, S. E. Lindquist, and A. Hagfeldt, ‘‘Charge Transport Properties in the Nanostructured ZnO Thin Film Electrode–Electrolyte System Studied with Time Resolved Photocurrents,’’ Sol. Energy Mater. Sol. Cells, 60 [2] 181–93 (2000). 79 K. Kakiuchi, E. Hosono, and S. Fujihara, ‘‘Enhanced Photoelectrochemical Performance of ZnO Electrodes Sensitized with N-719,’’ J. Photochem. Photobiol. A, 179 [1–2] 81–6 (2006). 80 K. Keis, J. Lindgren, S. E. Lindquist, and A. Hagfeldt, ‘‘Studies of the Adsorption Process of Ru Complexes in Nanoporous ZnO Electrodes,’’ Langmuir, 16 [10] 4688–94 (2000). 81 M. Law, L. E. Greene, J. C. Johnson, R. Saykally, and P. Yang, ‘‘Nanowire Dye-Sensitized Solar Cells,’’ Nat. Mater., 4 [6] 455–9 (2005). 82 J. B. Baxter and E. S. Aydil, ‘‘Nanowire-Based Dye-Sensitized Solar Cells,’’ Appl. Phys. Lett., 86 [5] 053114 (2005). 83 B. Pradhan, S. K. Batabyal, and A. J. Pal, ‘‘Vertically Aligned ZnO Nanowire Arrays in Rose Bengal-Based Dye-Sensitized Solar Cells,’’ Sol. Energy Mater. Sol. Cells, 91 [9] 769–73 (2007). 84 A. B. F. Martinson, W. E. Jeffrey, T. H. Joseph, and J. P. Michael, ‘‘ZnO Nanotube Based Dye-Sensitized Solar Cells,’’ Nano Lett., 7 [8] 2183–7 (2007). 85 S. H. Kang, J.-Y. Kim, Y. Kim, H. S. Kim, and Y.-E. Sung, ‘‘Surface Modification of Stretched TiO2 Nanotubes for Solid-State Dye-Sensitized Solar Cells,’’ J. Phys. Chem. C, 111 [26] 9614–23 (2007). 86 Y. F. Hsu, Y. Y. Xi, A. B. Djurisic, and W. K. Chan, ‘‘ZnO Nanorods for Solar Cells: Hydrothermal Growth Versus Vapor Deposition,’’ Appl. Phys. Lett., 92 [13] 133507 (2008). 87 I. D. Kim, J. M. Hong, B. H. Lee, D. Y. Kim, E. K. Jeon, D. K. Choi, and D. J. Yang, ‘‘Dye-Sensitized Solar Cells Using Network Structure of Electrospun ZnO Nanofiber Mats,’’ Appl. Phys. Lett., 91 [16] 163109 (2007). 88 D. I. Suh, S. Y. Lee, T. H. Kim, J. M. Chun, E. K. Suh, O. B. Yang, and S. K. Lee, ‘‘The Fabrication and Characterization of Dye-Sensitized Solar Cells with a Branched Structure of ZnO Nanowires,’’ Chem. Phys. Lett., 442 [4–6] 348–53 (2007). 89 K. Kakiuchi, E. Hosono, T. Kimura, H. Imai, and S. Fujihara, ‘‘Fabrication of Mesoporous ZnO Nanosheets from Precursor Templates Grown in Aqueous Solutions,’’ J. Sol–Gel. Sci. Technol., 39 [1] 63–72 (2006). 90 T. Hoshikawa, R. Kikuchi, K. Sasaki, and K. Eguchi, ‘‘Impedance Analysis of Electronic Transport in Dye-Sensitized Solar Cells,’’ Electrochemistry (Tokyo, Japan), 70 [9] 675–80 (2002). 91 F. Lenzmann, J. Krueger, S. Burnside, K. Brooks, M. Graetzel, D. Gal, S. Ruehle, and D. Cahen, ‘‘Surface Photovoltage Spectroscopy of Dye-Sensitized Solar Cells with TiO2, Nb2O5, and SrTiO3 Nanocrystalline Photoanodes: Indication for Electron Injection from Higher Excited Dye States,’’ J. Phys. Chem. B, 105 [27] 6347–52 (2001). 92 S. G. Chen, S. Chappel, Y. Diamant, and A. Zaban, ‘‘Preparation of Nb2O5 Coated TiO2 Nanoporous Electrodes and Their Application in Dye-Sensitized Solar Cells,’’ Chem. Mater., 13 [12] 4629–34 (2001). 93 P. Guo and M. A. Aegerter, ‘‘Ru(II) Sensitized Nb2O5 Solar Cell Made by the Sol–Gel Process,’’ Thin Solid Films, 351 [1,2] 290–4 (1999). 94 M. Wei, Z. Qi, M. Ichihara, and H. Zhou, ‘‘Synthesis of Single-Crystal Niobium Pentoxide Nanobelts,’’ Acta Mater., 56 [11] 2488–94 (2008). 95 J. Xia, N. Masaki, K. Jiang, and S. Yanagida, ‘‘Sputtered Nb2O5 as an Effective Blocking Layer at Conducting Glass and TiO2 Interfaces in Ionic Liquid-Based Dye-Sensitized Solar Cells,’’ Chem. Commun. (Cambridge, UK), [2] 138–40 (2007). 96 J. Xia, N. Masaki, K. Jiang, and S. Yanagida, ‘‘Fabrication and Characterization of Thin Nb2O5 Blocking Layers for Ionic Liquid-Based Dye-Sensitized Solar Cells,’’ J. Photochem. Photobiol. A, 188 [1] 120–7 (2007). 97 J. Sancho-Parramon, V. Janicki, and H. Zorc, ‘‘Compositional Dependence of Absorption Coefficient and Band-Gap for Nb2O5–SiO2 Mixture Thin Films,’’ Thin Solid Films, 516 [16] 5478–82 (2008). 98 K. Hara, T. Horiguchi, T. Kinoshita, K. Sayama, H. Sugihara, and H. Arakawa, ‘‘Highly Efficient Photon-to-Electron Conversion with Mercurochrome-Sensitized Nanoporous Oxide Semiconductor Solar Cells,’’ Sol. Energy Mater. Sol. Cells, 64 [2] 115–34 (2000). 99 K. Sayama, H. Sugihara, and H. Arakawa, ‘‘Photoelectrochemical Properties of a Porous Nb2O5 Electrode Sensitized by a Ruthenium Dye,’’ Chem. Mater., 10 [12] 3825–32 (1998). 100 Powder Diffraction File Card No. 37-1468. 101 Powder Diffraction File Card No. 27-1003. 102 S. Burnside, J. E. Moser, K. Brooks, M. Grätzel, and D. Cahen, ‘‘Nanocrystalline Mesoporous Strontium Titanate as Photoelectrode Material for Photosensitized Solar Devices: Increasing Photovoltage through Flatband Potential Engineering,’’ J. Phys. Chem. B, 103, 9328–32 (1999). 103 B. Tan, E. Toman, Y. Li, and Y. Wu, ‘‘Zinc Stannate (Zn2SnO4) DyeSensitized Solar Cells,’’ J. Am. Ceram. Soc., 129, 4162–3 (2007). 104 T. Lana-Villarreal, G. Boschloo, and A. Hagfeldt, ‘‘Nanostructured Zinc Stannate as Semiconductor Working Electrodes for Dye-Sensitized Solar Cells,’’ J. Phys. Chem. B, 111, 5549–56 (2007). 105 P. Balaya, J. Jamnik, J. Fleig, and J. Maier, ‘‘Mesoscopic Electrical Conduction in Nanocrystalline SrTiO3,’’ Appl. Phys. Lett., 88, 062109 (2006). 106 K. Tennakone, G. R. R. A. Kumara, I. R. M. Kottegoda, and V. P. S. Perera, ‘‘An Efficient Dye-Sensitized Photoelectrochemical Solar Cell Made from Oxides of Tin and Zinc,’’ Chem. Commun. (Cambridge, UK), 15–16 (1999). February 2009 Metal Oxides for Solar Cells 107 K. L. Hardee and A. J. Bard, ‘‘Semiconductor Electrodes,’’ J. Electrochem. Soc., 123, 1024–6 (1976). 108 A. Fujishima and K. Honda, ‘‘Electrochemical Photolysis of Water at a Semiconductor Electrode,’’ Nature, 238, 37–8 (1972). 109 U. Bjoerksten, J. Moser, and M. Grätzel, ‘‘Photoelectrochemical Studies on Nanocrystalline Hematite Films,’’ Chem. Mater., 6 [6] 858–63 (1994). 110 W. B. Ingler, J. P. Baltrus, and S. U. M. Khan, ‘‘Photoresponse of p-Type Zinc-Doped Iron(III) Oxide Thin Films,’’ J. Am. Ceram. Soc., 126 [33] 10238–9 (2004). 111 A. Corma, P. Atienzar, H. Garcia, and J. Y. Chane-Ching, ‘‘Hierarchically Mesostructured Doped CeO2 with Potential for Solar-Cell Use,’’ Nat. Mater., 3, 394–7 (2004). Prof. Seeram Ramakrishna is currently the Research Strategy VicePresident of the National University of Singapore (NUS). After receiving Ph.D. degree in Materials Science and Engineering from the University of Cambridge in 1992, Seeram joined at the Kyoto Institute of Technology as a visiting lecturer. He joined the NUS in 1996 as Lee Kuan Yew Fellow where he was promoted to Professor at the departments of Mechanical engineering as well as Bioengineering in 2003. He served/ serving in various capacities as Directors, Dean, chairman in advisory boards, and in the editorial board of over 15 scholarly journals. His research interests include biomaterials, energy, and environment. The ISI web of knowledge currently ranks him at 99 out of 3415 most cited materials scientists in the world. 301 Dr. Rajan Jose pursued doctoral research at the Council of Scientific and Industrial Research (CSIR), Trivandrum, India and received PhD degree in Physics from the Mahatma Gandhi University, India in 2002. He was a postdoctoral fellow at AIST, Japan (2003–2005) and Toyota Technological Institute, Japan (2005–2007). He joined the NUS in 2007 where he is working on nanostructured materials for excitonic solar cells and one-dimensional metal oxide nanostructures. His research interests include functional inorganic solids for electronic, photonic, and energy applications. Dr Velmurugan Thavasi is a Research Fellow at the NUS Nanoscience and Nanotechnology Initiative. He obtained MEng in Chemical Engineering and PhD in Chemistry from National University of Singapore. His research interests include synthesis of nanomaterials, interfaces engineering, and the study of excitonic solar cells and bio-nano devices. Email: velnanotech@gmail.com