You are on page 1of 11

View Article Online / Journal Homepage / Table of Contents for this issue

Chem Soc Rev


Cite this: Chem. Soc. Rev., 2012, 41, 42074217 www.rsc.org/csr
Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

Dynamic Article Links

TUTORIAL REVIEW

Adsorption of organic molecules on rutile TiO2 and anatase TiO2 single crystal surfaces
Andrew G. Thomas*a and Karen L. Syresb
Received 29th February 2012 DOI: 10.1039/c2cs35057b The interaction of organic molecules with titanium dioxide surfaces has been the subject of many studies over the last few decades. Numerous surface science techniques have been utilised to understand the often complex nature of these systems. The reasons for studying these systems are hugely diverse given that titanium dioxide has many technological and medical applications. Although surface science experiments investigating the adsorption of organic molecules on titanium dioxide surfaces is not a new area of research, the eld continues to change and evolve as new potential applications are discovered and new techniques to study the systems are developed. This tutorial review aims to update previous reviews on the subject. It describes experimental and theoretical work on the adsorption of carboxylic acids, dye molecules, amino acids, alcohols, catechols and nitrogen containing compounds on single crystal TiO2 surfaces.

1. Introduction
Titanium dioxide is utilised in a range of technological applications including as a pigment, a biosensor support,1 photocatalyst2 and in the Gra tzel photovoltaic cell3 to name but a few. In addition, it is also found at the surface of Ti and Ti alloy
a

School of Physics and Astronomy and the Photon Science Institute, The University of Manchester, Oxford Road, Manchester M13 9PL. E-mail: andrew.g.thomas@manchester.ac.uk b School of Chemistry, The University of Nottingham, University Park, Nottingham, NG7 2RD. E-mail: karen.syres@nottingham.ac.uk

biomaterials where its presence is thought to give rise to the excellent osseointegration properties of these materials.4 The surface properties and interactions of TiO2, particularly with small molecules such as water, oxygen, formate and methanol for example,5,6 have been widely studied for over thirty years. In many of the applications described above TiO2 is either intentionally functionalised or expected to interact with organic molecules. In the case of dye-sensitised solar cells (DSSCs) the titania surface is coated with a dye molecule, the most ecient of which to date is the N3-dye. However, new dyes which utilise the process of singlet ssion (SF), whereby two charge carriers are produced following absorption of a single photon,

Andrew Thomas received his BSc in Chemistry from the University of Manchester and his MSc in instrumentation and Analytical Science from UMIST, before obtaining his PhD from the University of Liverpool. Following this he held three research appointments at UMIST before becoming an Experimental Ocer in the Department of Physics. He was made a Research Fellow in Physics at UMIST in 2001 and moved to the Andrew G. Thomas Photon Science Institute at The University of Manchester following the merger of UMIST and The Victoria University of Manchester in 2004.
This journal is
c

Karen Syres obtained her MPhys degree from the University of Manchester before obtaining her PhD under the supervision of Andrew Thomas and Wendy Flavell. Following this she held a one-year PhD plus scholarship before moving to The School of Chemistry at the University of Nottingham where she is currently a Post-doctoral Research Fellow. Karen L. Syres

The Royal Society of Chemistry 2012

Chem. Soc. Rev., 2012, 41, 42074217

4207

View Article Online

are now being sought.7 In order to determine suitable SF dyes the TiO2-organic bonding, energy level alignment and charge injection rates must be fully characterised. Whilst dye sensitised solar cells are already commercially available the electricity can not be stored. In order to overcome the energy storage problem TiO2 is also being investigated as a substrate for articial photosynthesis. Here light energy is used to drive photocatalytic redox reactions to produce hydrogen, or other fuels from sunlight.8 Targeted biomaterials based upon TiO2 nanoparticles, which are designed to cluster at the site of tumours, have been functionalised with polyethylene glycol to evade the bodys immune system. Here molecules such as dopamine or dihydroxy phenylalanine (DOPA) have been used as the anchor groups.9 Furthermore, studies have shown dopamine to be eective both as an anchor molecule to bind DNA to titania nanoparticles and to enhance charge separation.10 DOPA and dopamine are members of the catechol group of chemicals and along with the simplest catechol, pyrocatechol, have been shown to shift the optical absorption spectrum of TiO2 from the ultraviolet (UV) to the visible part of the electromagnetic spectrum. The interactions of TiO2 at the surface of biomaterials are numerous and extremely complex. As the molecules become larger their interactions also become more complex as the number of potential bonding groups on the molecule increases.11 Due to its ready availability, the rutile TiO2 (110) surface is the most widely studied of the three structural phases of TiO2 (anatase, rutile and brookite). Many technological applications consist of TiO2 in the anatase form since this is the phase adopted by nanoparticulate TiO2.12 This is thought to be due to the fact that the anatase TiO2 (101) surface has the lowest surface energy.13 Rutile TiO2 (110) in its (1 1) termination is considered a prototypical metal oxide surface and much is now understood about this surface. There have also been a number of studies of the (100), (001) and (011) rutile TiO2 surfaces. Anatase surfaces have been less widely studied as high quality single crystals have been dicult to obtain. However over the past ten years or so the quality of anatase single crystals and growth of thin lms has improved.1416 This has allowed comparisons, particularly between the crystallographically equivalent rutile TiO2 (110) and anatase TiO2 (101) surfaces to be made.17 Subtle dierences in both the surface electronic structure and adsorption strengths have been observed, although the adsorption geometries of the molecules that have been studied on both surfaces are similar.1820 This review seeks to give a summary of adsorption of organic molecules on TiO2 from both an experimental and theoretical perspective. It will concentrate mainly on the adsorption on highly-idealised vacuum-prepared single crystal surfaces. We will also try to avoid repeating work described in the comprehensive review by Diebold5 and the more recent critical review by Pang et al.6 Rather, we hope to bring those articles up to date as well as describing why this area is so heavily researched. In order to fully appreciate the adsorption and interaction of organic molecules with the various surfaces we shall begin with a very brief overview of the geometric and electronic structure of the most widely studied rutile and anatase TiO2 surfaces.
4208 Chem. Soc. Rev., 2012, 41, 42074217

2. Clean surfaces
The reviews by Diebold5 and Pang et al.6 describe the structures of clean TiO2 surfaces in some detail. Here we wish only to give a basic understanding of the structure. It is well established that TiO2 single crystal surfaces can be routinely prepared in vacuum by Ar+ ion sputter/anneal cycles. The age and history of TiO2 single crystals has a large eect on the surface. The rutile TiO2 (110) 1 1 surface consists of rows of bridging oxygen ions (Obr) with 5 fold Ti4+ ions (Ti5c) and in plane oxygen ions. This structure has been widely conrmed experimentally using various diraction and scanning probe microscopy (SPM) techniques.6 With reference to the surface chemistry and electronic structure some of the most interesting information obtained concerns surface oxygen vacancies.5 It has long been known that O-vacancies at the TiO2 (110) surface lead to the formation of a band gap state around 1 eV below the Fermi level.17 Although this peak is not removed by water adsorption it is removed by treatment with molecular oxygen. SPM has indicated that the peak is associated with O-vacancies in the bridging oxygen rows. However, the charge is distributed over several Ti atoms and only a small amount remains on the topmost surface atoms.21 Apart from the (110) surface the three most commonly investigated rutile TiO2 surfaces are the (100), (001) and (011) surfaces. The (100) surface forms a corrugated (1 1) structure with rows of oxygens at the outermost surface.5 Annealing at higher temperatures leads to a (1 3) reconstruction thought to consist of (110) microfacets.5 The (001) surface is inherently unstable due to the large number of broken bonds at the surface. Low energy electron diraction (LEED) data suggests the surface forms (011) facets at the surface to minimise the surface energy. The TiO2 (011) surface undergoes a (2 1) reconstruction which exhibits Ti5c and two-fold coordinated oxygen atoms at the surface.22 The twofold coordinated oxygen atoms partially block the Ti5c sites potentially blocking adsorption at these sites from gas phase molecules.23 In addition, this surface seems resilient to the formation of surface oxygen vacancies. The anatase TiO2 (101) surface is the most stable and most frequently observed surface of anatase TiO2. It has a sawtooth structure with fully co-ordinated six-fold and under co-ordinated ve-fold Ti atoms. As with the rutile phase, surface preparation in vacuum by Ar+ ion etching and annealing to 700 1C leads to a well-ordered surface. Again, similarly to the rutile (110) surface, photoemission spectra of the valence band region often show a feature in the band gap region at a binding energy of 1 eV. However, unlike rutile this peak is thought to arise from subsurface O-vacancies.24 The anatase TiO2 (001) surface is known to undergo a (1 4)(4 1) reconstruction when prepared in vacuum. A number of dierent models have been suggested for this structure including the added molecule model (ADM),25,26 added and missing row models, which result in formation of planes in the [103]27 or [014]20,28 directions, and a model based on [101] microfacets.20,29 We will turn now to the main focus of this tutorial review, namely the adsorption of organic molecules on TiO2 surfaces.
This journal is
c

Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

The Royal Society of Chemistry 2012

View Article Online

3. Carboxylic acids, dicarboxylic acids and anhydrides


Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

Simple carboxylic acid adsorption on TiO2 surfaces has been widely studied. The motivation for these studies lies in technological applications of TiO2, such as DSSCs, and as model molecules for understanding the (photo)catalytic activity of TiO2 surfaces. We shall discuss photosensitising dyes and their ligands and amino acids separately in sections 4 and 5. We will concentrate here on simple carboxylic acids, anhydrides and di-carboxylic acids. 3.1 Monocarboxylic acids

3.1.1 Adsorption on the rutile TiO2 (110) surface. The review by Pang et al.6 discussed the adsorption of formic acid on rutile TiO2 (110) thoroughly thus we shall summarise only the main points here. Suce to say, that formic acid adsorbs at room temperature on all TiO2 surfaces which have been studied. The adsorption occurs dissociatively as formate and is considered to be a model for all monocarboxylic acids adsorbed on TiO2. In addition, formate, acetate, propionate and trimethyl acetate (TMA) form (2 1) overlayers at saturation coverage.6 Chemical-state specic scanned-energy mode photoelectron diraction (PhD) carried out by Sayago et al. concluded that the formate is bound to two surface Ti5c atoms in a bridging geometry. This geometry (A) is shown in Fig. 1.6 Although other experiments and theoretical treatments support this geometry there are also results which suggest alternative geometries. One of these involves lling of an Obr site by one of the carboxyl oxygen atoms with the other bound to Ti5c (B in Fig. 1)6 and another involves bonding through only one formate oxygen atom (site C in Fig. 1).30 It has been suggested that experimental conditions such as the surface preparation or history of the rutile TiO2 (110) crystal, and the sample temperature during dosing may all play a part in determining the adsorption geometry.6 A near edge absorption ne structure spectroscopy (NEXAFS) study of formate, acetate and propionate supported the existence of minority adsorption sites involving bridging-oxygen vacancies. The molecules were found to exhibit dierent twist angles

relative to the [001] azimuth. This observation was ascribed to a majority of molecules being adsorbed to Ti5c along the [001] azimuth and a smaller proportion to Obr vacancy sites, roughly perpendicular to the [001] azimuth (B in Fig. 1). The molecules were also found to be roughly perpendicular to the TiO2 (110) surface.31 The proton lost from from formic acid upon adsorption is thought to bind to bridging oxygen atoms to form a bridging hydroxyl. However, Lyubinetsky et al. suggested that for TMA overlayers the proton is only weakly bound to the bridging oxygens and may oscillate between two oxygen atoms.6 Benzoic acid adsorption on the rutile TiO2 (110) surface has been studied by LEED, NEXAFS, electron stimulated desorption ion angular distribution (ESDIAD), X-ray photoelectron spectroscopy (XPS) and scanning tunneling microscopy (STM).3235 Like the aliphatic acids it is found to adsorb dissociatively in a bridging bidentate geometry. However, unlike the simple aliphatic acids benzoic acid is found to form dimers by rotation of the phenyl ring. This then allows the formation of bonds between the hydrogen atoms of one ring and the p-system of its neighbour along the direction.32,33 This geometry is also thought to allow interaction of another hydrogen atom in the molecule with the bridging oxygen rows, via hydrogen bonds. Pyridine carboxylic acids are found to adsorb to the rutile TiO2 (110) surface in a similar manner to benzoic acid.36 A detailed study of the adsorption of picolinic acid, nicotinic acid and isonicotinic acid showed that slow deposition of these molecules onto the (110) surface led to the formation of monolayer dimer structures. These are formed by interaction between the nitrogen lone pair and hydrogen atoms of nearest neighbours. It was found that this interaction was strongest for isonicotinic acid where tilting of the molecule led to increased interaction between the molecules. Interestingly this work showed that the rate of molecular deposition onto the surface was critical to the degree of ordering.34,35 With regard to thermal and photodissociation of small carboxylic acids on the rutile TiO2 (110) surface, there have been many studies. These are discussed in depth in the reviews of Diebold,5 Pang et al.6 and Henderson2 and the reader is referred to these works for more details. 3.1.2 Other rutile surfaces. Adsorption of small carboxylic acids, on the rutile TiO2 (100) (1 3) and TiO2 (100) (1 1) show the adsorption is similar to that seen for the rutile (110) surface, i.e. the acid adsorbs dissociatively. Similar results were obtained for adsorption on the (001) surface.5 More recently there have been a few experimental and theoretical studies of carboxylic acid adsorption on the TiO2 (011) (2 1) surface23,37,38 which suggest adsorption in a bridging bidentate mode.38 This bonding mode is favoured despite the larger TiTi distances found on the (011) surface relative to the (110) surface.37 Temperature programmed desorption (TPD) measurements suggest the molecule decomposes above 500 1C with ketene, CO and CH4 being the main decomposition products. At temperatures lower than 500 1C the acetate recombines with an adsorbed proton and desorbs intact.38 Quah et al. studied the photoexcited decomposition of acetic acid on the rutile TiO2 (011) surface using XPS and TPD.38 UV illumination in the absence of oxygen gas reproduced the results found for the TiO2 (110) surface, i.e. C 1s XPS spectra
Chem. Soc. Rev., 2012, 41, 42074217 4209

Fig. 1 The three adsorption geometries deduced for formate on the rutile TiO2 (110) surface. A is the geometry thought to be adopted by the majority of adsorbed formate, deduced from PhD and NEXAFS.6 B is a minority species inferred from NEXAFS and STM measurements and C from STM.6

This journal is

The Royal Society of Chemistry 2012

View Article Online

Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

showed no change. However, in the presence of O2 gas, C 1s peaks due to adsorbed acetate were found to decrease in intensity, indicating loss of acetate from the surface. The cross section for acetate removal was shown to decrease with decreasing oxygen partial pressure. The main decomposition products were found to be ethane and methane. Ethane production was reduced if the background oxygen partial pressure was not replenished. It was suggested that this occurred because surface oxygen was used up in the reaction and replaced by gas phase oxygen. The production of ethane could be restored by the introduction of more oxygen. Preparation of a reduced TiO2 (011) surface by electron bombardment showed no dierence in photocatalytic activity towards adsorbed acetate. 3.1.3 Single crystal anatase TiO2 surfaces. Tanner et al. have studied adsorption of acetic and formic acids on the anatase TiO2 (001) (1 4)(4 1) reconstructed surface epitaxially grown on a (100) Nb-doped SrTiO3 substrate.26,29 STM showed that the acids adsorb on the fully oxidised surface.26 Neither acetate nor formate species could be desorbed by heating up to 700 1C. Above 700 1C, CO and CO2 were observed in TPD experiments. Both acids were found to be adsorbed dissociatively to form a (4 2) structure at saturation coverage. This structure was thought to arise due to adsorption of the deprotonated carboxyl groups to undercoordinated Ti atoms at the surface. Flashing the surfaces to around 850 1C gave dierent results for formate and acetate. Formate appeared particularly resistant to annealing and formed strongly bound patches. As the temperature was increased above 850 1C it was found that a new phase was formed which the authors describe as a disordered (4 4) overlayer.29 Acetate, on the other hand, was simply lost from the surface at 850 1C, resulting in a small coverage of acetate and decomposition products. TPD measurements from anatase TiO2 (001) which had been Ar+ ion etched prior to exposure to formic or acetic acid showed desorption of a number of species between 300 and 750 1C. Formate decomposed to form CO, formaldehyde (H2CO), CO2 and water. Acetate gave rise to CO, ketene (CH2CO) and smaller amounts of water and acetic acid,29 similar to the rutile surfaces described above. 3.2 Dicarboxylic acids

Fig. 2 Possible binding modes of a dicarboxylic acid to a TiO2 surface.

The interest in dicarboxylic acids lies in the fact that bonding via two acid groups to the TiO2 surface may lead to more ecient charge transfer between an adsorbed molecule and the oxide surface. In addition, similar molecules are of interest in preventing crystal growth of particular surface planes, or limiting the size of crystals in the preparation of nanoparticles. This is achieved if the bonding to the surface is suciently strong to prevent further reaction of the crystal with the growth medium. Fig. 2 shows the various ways in which a dicarboxylic acid may be expected to adsorb on a TiO2 surface. As in the case of monocarboxylic acids, a chelating mode is unlikely at a Ti5c site since this would lead to 7-fold coordinated Ti ions.39 To the best of the authors knowledge the only experimental studies of dicarboxylic acid adsorption on single crystal surfaces are those of dyes and dye ligands
4210 Chem. Soc. Rev., 2012, 41, 42074217

used in DSSC systems which are discussed below. The adsorption of oxalic acid40 on TiO2 has been the subject of some theoretical work. The surfaces were modelled with simple TiO2 polymers to represent rutile and anatase surfaces. The calculations suggest the most stable adsorption geometry (i.e. the structure which gives the highest adsorption energy) involves deprotonation of both carboxyl groups. The adsorption then takes place through two oxygen atoms in a bridging mode. Adsorption was found to be stronger on the anatase surface than the rutile surface.40 Malonic acid adsorbed on P25 degussa particulate TiO2 has been studied using attenuated total reection infrared spectroscopy. This work suggested adsorption of malonic acid on the TiO2 surface occurred via one bridging bidentate and one monodentate carboxylate group as shown in Fig. 2j. Illumination of the surface with UV showed decomposition of the malonic acid rst to oxalic acid and eventually to CO and H2O.41 3.3 Anhydrides The adsorption of acetic anhydride on the rutile TiO2 (110) surface at room temperature has been studied by XPS, LEED and high resolution electron energy loss spectroscopy (HREELS).42 Unlike acetic acid, the anhydride does not have a terminal proton but instead has two carbonyl groups in an almost co-planar structure. The distance between the two carbonyl groups
This journal is
c

The Royal Society of Chemistry 2012

View Article Online

Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

Scheme 1

is 0.265 nm which is slightly smaller than the separation between Ti5c atoms in the (110) surface (B0.296 nm). A direct XPS comparison of adsorption of acetic anhydride and acetic acid showed that 12% more carbon was present on the anhydride dosed surface relative to the acetic acid dosed surface. Theoretical work has suggested that the strength of carboxylic acid adsorption was partly due to the formation of bridging hydroxyl groups (OHbr) stabilising the carboxylate moiety.43 However, the results for anhydride adsorption suggest the presence of the proton has little eect on the stability. LEED and HREELS spectra indicated that acetic anhydride adsorbs dissociatively on the surface to form acetate ions. For an intact acetic anhydride molecule one would expect a p(3 1) LEED pattern,42 however a p(2 1) LEED pattern was observed. This is also the LEED pattern one obtains at saturation coverage of acetic acid.32 The dissociative adsorption process involves a surface bridging oxygen species as shown in Scheme 1a. The authors rationalise that the part of the anhydride which attaches to the surface bridging oxygen acts in a similar manner to the proton in the acetic acid, i.e. to reduce the negative charge of the bridging oxygen rows.42 In doing so it allows the negatively charged acetate moiety to approach the Ti5c atom. However, attempts to conrm the requirement to neutralise the bridging oxygen atoms using methyl acetate indicated that methyl acetate did not adsorb on the surface. This is despite the fact that the methyl fragment should be able to neutralise Obr as shown in Scheme 1b. The authors concluded that it was the requirement for the bridging geometry which governed the adsorption mechanism, resulting in two acetate adsorption. The authors also suggest some of the acetate species which had formed on Obr could convert to occupy two Ti5c sites resulting in a bridging oxygen vacancy. This vacancy site is usually associated with the presence of Ti3+ at the surface. However, here no redox reaction occurs since the surface oxygen binds to the oxygen decient acetate moiety (Scheme 1b) thus the Ti atom retains its +4 oxidation state. This, they argue leads to an excess positive charge on the bridging oxygen row which stabilises the bridging adsorption geometry in a similar way to the presence of OHbr. Wilson et al. have studied adsorption of maleic anhydride (MA) on the rutile TiO2 (001) surface.44 The motivation lies in its wide use in chemical synthesis and also more recently as a potential dye anchor system for DSSCs. It was suggested from the desorption products observed in TPD that adsorption on this surface occurred via ring cleavage at the central oxygen atom.
This journal is
c

Fig. 3 Proposed adsorption geometry for maleic anhydride on the anatase TiO2 (101) surface determined from photoemission and NEXAFS spectroscopy. The arrow indicates the surface bridging oxygen to which the molecule binds, resulting in a doubly-bidentate bridging geometry. Adapted from ref. 45 with permission of the author.

This results in a bidentate chelating adsorption mode involving a surface oxygen atom. TPD showed decomposition products which varied as a function of the surface stoichiometry with CO, CO2 and ketene dominating the TPD spectra for both surface treatments. Acetylene was also found to be desorbed from the surface. MA adsorption on the (101), (001) and (100) anatase TiO2 surfaces has been studied by Johansson et al.45 Adsorption on the (101) and (100) surfaces also occurs via ring opening and attachment through the three oxygen atoms of the molecule and a surface oxygen atom.45 However, in this case the adsorption mode was deduced to be a bridging geometry. Adsorption on the TiO2 (101) surface involves three surface titanium atoms as shown in Fig. 3. On the anatase TiO2 (001) surface, O 1s XPS spectra showed two peaks arising from MA which is indicative of oxygen in two dierent chemical environments. Although the chemical shifts of these two peaks were dierent from that of a multilayer of MA, the authors suggest that the ring opening process does not occur on the (001) surface.45

4. Photosensitising dye molecules


The adsorption of dye-sensitising molecules and charge transfer between these molecules and titania surfaces is of great fundamental and technological interest from the point of view of DSSCs. The most widely used dye for these TiO2 nanoparticle cells is the so-called N3 dye, ruthenium di-2,2 0 -bipyridyl4,4 0 -dicarboxylic acid diisocyanate. This large molecule is labile so that the normal method of adsorption onto clean surfaces in ultra-high vacuum (evaporation by heating in vacuum) is not possible. Instead the molecule has been deposited by electrospray deposition46 in UHV or by removing vacuum prepared crystals capped with small molecules and
Chem. Soc. Rev., 2012, 41, 42074217 4211

The Royal Society of Chemistry 2012

View Article Online

5. Amino acids
The adsorption of amino acids on titania surfaces is of interest from the point of view of biomaterials and biosensors. Amino acids have also been investigated to control the size and shape of TiO2 nanoparticles.51 The success of Ti based biomaterials is attributed to the layer of passivating oxide at the surface and it is this surface which will be exposed to the biological environment and govern the success or failure of the implant. Functional biomaterials based upon TiO2 nanoparticles are also being investigated. Amino acid adsorption is also fundamentally interesting since TiO2 is an amphoteric oxide. Amino acids, of course, have an acidic carboxyl group and a basic amine group. One may therefore expect there to be some competition as to which group will bind to the surface. There are some diculties in using photoelectron spectroscopy to investigate amino acid adsorption due to the instability of the molecules under high intensity radiation. One of the earliest studies of glycine adsorption on rutile TiO2 (110) showed that the molecule dissociated on the surface under synchrotron radiation.52,53 It has been known for some time that pyridine does not form a strong bond to the TiO2 surface (see below). Experimentally, this behaviour is echoed in amino acid adsorption where binding on TiO2 single crystal surfaces in vacuum occurs in a similar manner to that seen for carboxylic acids.39,5457 Tonner has carried out DFT calculations of proline and glycine adsorbed on the rutile TiO2 (110) surface. The results showed adsorption through the carboxyl group, with proton transfer to the surface.58 They also showed hydrogen bonding via the amine group which led to further stabilisation of the adsorption as shown in Fig. 5. Szieberth et al. performed DFT calculations of glycine adsorption on the anatase (101) surface.59 The highest adsorption energy was obtained for a model where the carbonyl oxygen of glycine bonds to a Ti5c site. In this model the hydroxyl group forms a hydrogen bond to a twofold coordinated oxygen ion and the amine group is bonded to Ti5c via the nitrogen lone pair. However, the authors point out that the energy dierence between this model and one involving adsorption via the deprotonated carboxyl group, or indeed solely via the amine group, is very small thus it is possible all three modes of adsorption may be present.

Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

Fig. 4 Proposed adsorption geometry of the N3 dye on the rutile TiO2 (110) surface deposited by the electrospray method. Figure adapted from ref. 48 with permission of the author.

dropping the dye from ethanolic solution.47,48 Electrospray experiments suggests the dye bonds via one of the bi-isonicotinic acid (BINA) ligands in a bidentate geometry with further bonding through the sulfur atom of one of the thiocyanate groups as shown in Fig. 4. A density functional theory (DFT) study of a simpler dye molecule (cis(CO)-trans(I)-Ru-(4,4 0 -dicarboxylate2,2 0 -bipyridine)(CO)2I2 on the anatase TiO2 (101) surface also suggests bidentate bonding via the BINA ligand is the most stable conguration.49 A number of studies have looked at the BINA ligand adsorbed on the rutile TiO2 (110) surface19,34,35 and the anatase TiO2 (101) and (001) surfaces.20 Regarding the adsorption geometry of the ligand the mode of adsorption is the same on both the rutile and anatase surfaces studies, i.e. both acid groups become deprotonated and the molecule adsorbs in a doubly bidentate geometry. There were some slight dierences found in the adsorption angle with respect to the surface normal. On the rutile TiO2 (110) surface the tilt angle was found to be 251 from the surface normal with an azimuthal twist of 441 relative to the [001] crystallographic direction. On anatase TiO2 (101) the tilt angle was around 201 from the surface normal with a twist of around 401 from the [010] azimuth.20 On the anatase (001) surface the tilt angle was much larger at 531 but the authors pointed out that this may be due to the surface reconstruction. Charge transfer from the BINA ligand to the TiO2 (110) surface was found to occur in o 3 fs.50 For the complete dye molecule deposited by the electrospray method, electrons were found to be injected from the third lowest unoccupied molecular orbital (LUMO+3) to the surface in less than 16 fs.46
4212 Chem. Soc. Rev., 2012, 41, 42074217

Fig. 5 Proposed adsorption geometry for glycine on the rutile TiO2 (110) surface determined from DFT calculations.58 Hydrogen bonding occurs between the nitrogen atom and the proton lost (marked with an arrow) from the carboxylic acid group.

This journal is

The Royal Society of Chemistry 2012

View Article Online

Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

Fleming and co-workers have carried out studies of glycine and proline adsorbed on rutile TiO2 (110) single crystal surfaces. Proline is found to adsorb in a bidentate geometry via the carboxyl group. They observed the presence of NH and NH2+ in proline suggesting two adsorption structures a deprotonated anionic form and a zwitterionic form. Upon heating the zwitterion was lost from the surface due to loss of the additional proton in the amine. Glycine adsorption on the rutile TiO2 (110) has been studied by STM.56 Again it was found that the molecule adsorbed preferentially via deprotonation of the carboxyl group and formed a (2 1) overlayer similar to acetic and formic acids. Glycine adsorption on the TiO2 (011) surface studied using XPS found similar results. As for the case of proline, there was evidence of coexistence of anionic and zwitterionic glycine. The latter was again found to be lost upon heating due to conversion to the anion. Adsorption of phenylalanine on the rutile TiO2 (110) surface suggests the molecule adsorbs in a bidentate geometry following deprotonation of the acid group.60 There was little evidence for formation of the zwitterionic state in agreement with DFT calculations for glycine and proline adsorbed on this surface.58 The authors also observed the possible formation of hydrogen bonds between the amine group of the rst layer and the carboxyl group of a second layer. Finally, in this section we point out that for peptides it is likely to be the side groups of the amino acids which control the overall adsorption mechanism. This is because the terminal carboxyl and amine groups will form a small proportion of the overall molecule.11

6. Alcohols
The decomposition of adsorbed alcohol molecules has been the subject of many studies aiming to understand the nature of photochemical reactions on the TiO2 surface. In addition, methanol and ethanol production from photocatalytic reactions are of particular interest for use as fuel. Since the adsorption of small alcohols (methanol, ethanol and propanol) adsorbed on rutile TiO2 surfaces is covered comprehensively elsewhere,5,6 we shall mention only the main points and more recent ndings. 6.1 Adsorption on the rutile TiO2 (110) surface

single crystal5 found, like Henderson et al., that molecular methanol was the majority surface species. It was found that under UV light methoxy is more reactive than molecularly adsorbed methanol for hole-mediated photo-oxidation. UV light was found to lead to decomposition of methanol to formaldehyde and a surface OH group. Onda et al. studied the adsorption of methanol onto a rutile TiO2 (110) surface using two-photon photoemission (2PPE).61 Following adsorption of methanol onto the surface they observed a resonance which they assign to a charge transfer from the reduced Ti5c+4-d ions to the free H atoms on surface hydroxyl groups. They believe that the charge-transfer state is stabilised by the presence of the methanol molecules at the Ti5c+4-d sites. Gamble et al. used TPD and XPS to study the decomposition of ethoxy groups on a TiO2 (110) surface in the presence water or hydroxyl groups.5 Deuterated ethanol was found to adsorb dissociatively on the surface to create ethoxy groups (ethoxy groups were also formed by adsorption of tetraethoxysilane). They found that ethoxy groups bound to surface Ti atoms can be readily removed from the surface by combination with surface hydroxyl groups. They were desorbed from the surface as ethanol gas (B250400 K). However, ethoxy groups bound to bridging oxygen vacancies at the surface cannot react with water or hydroxyl groups on the surface (below B450 K). Jayaweera et al. used XPS to investigate the photoreaction under UV light of ethanol adsorbed on a rutile TiO2 (110) single crystal.5 They found that ethanol adsorbs dissociatively through its oxygen atom to one titanium atom on the surface. They suggest that due to steric eects and repulsion between neighbouring molecules that ethanol saturates at 0.5 molecules of ethanol to every one Ti atom, which also agreed with their estimation based on the attenuation of the Ti 2p signal. Under UV irradiation and an O2 atmosphere they observed a decrease in the peaks associated with ethanol adsorption and the rise of a peak due to CH3COO and HCOO. It is understood these species are formed by chemical reactions triggered by the photo-excited electron. 6.2 Alcohols on other TiO2 single crystal surfaces Kim and Barteau reported that on the TiO2 (001) surface methanol adsorbs molecularly and dissociatively at 200 K. Below room temperature molecular methanol desorbs from n et al. found that methanol coverage the surface.5 Roma on the (100) and (110) surfaces increased with the number of defects created on the surface by electron or Ar+ bombardment.5 Kim and Barteau have also studied the adsorption and decomposition of ethanol, n-propanol and isopropanol on a TiO2 (001) surface using TPD and XPS. They found the alcohols adsorbed molecularly and dissociatively at 200 K but only dissociatively at room temperature.5 Methanol adsorption on anatase TiO2 (101) has been studied by Diebolds group using TPD and XPS.15 It was found that methanol adsorbed molecularly on this surface with no signs of methoxy formation. The dierence to adsorption on the rutile surface was attributed to the diculty in forming surface oxygen vacancies on the anatase TiO2 (101) surface.
Chem. Soc. Rev., 2012, 41, 42074217 4213

Henderson et al.5 carried out an extensive study of methanol on a TiO2 (110) surface using TPD, HREELS, static secondary ion mass spectroscopy (SSIMS) and LEED. The majority of methanol molecules were found to adsorb molecularly to the surface, with evidence of some dissociative adsorption (creating methoxy), particularly at oxygen vacancy sites. It was found that exposing the TiO2 surface to O2 resulted in more methoxy groups on the surface due to cleavage of the CH3OH bond. It was also found that co-adsorbed water had little eect on methanol at the surface. An earlier study by Henderson et al., reported a high cross-section for electronstimulated decomposition of methanol-related adsorbed species on TiO2 (110).5 Therefore, methods such as LEED, where lowenergy electrons are red at the surface, prove dicult in characterising the methanolTiO2 interface. More recently, a TPD study of photochemical hole scavenging reactions of methanol adsorbed on a rutile TiO2 (110)
This journal is
c

The Royal Society of Chemistry 2012

View Article Online

7. Catechols
Catechols (benzenediols) such as pyrocatechol and dopamine adsorbed on TiO2 surfaces are being studied for many applications. Adsorbing pyrocatechol onto TiO2 nanoparticles is thought to result in interesting charge transfer processes62 and dopamine is widely used as a bridging molecule which facilitates electron/hole transfer between TiO2 nanoparticles and a biological system.63 In addition to their charge transfer properties, catechols are used in many systems because they form strong bonds with TiO2. Dopamine and L-dihydroxyphenylalanine (L-DOPA) are being used to attach molecules, which would not normally adsorb strongly on TiO2. This is done by grafting the molecule via the amine side group resulting in a strong bridge between the TiO2 and the molecule. 7.1 Pyrocatechol

The simplest catechol is pyrocatechol (1,2-benzene-diol) shown in Fig. 6. The pyrocatechol-TiO2 system is of interest since it has potential applications in solar cells.64 Pyrocatechol does not absorb light below 4.2 eV (300 nm) which is much larger than the 3.2 eV (370 nm) band gap of TiO2. However, in pyrocatechol-sensitised TiO2 nanoparticles there is an absorption shift to 3 eV (420 nm). It has been proposed by Persson et al.62 that instead of the electron being excited in the adsorbate and then being transferred into the TiO2 conduction band, there is a direct pyrocatechol-to-TiO2 charge transfer. This means the electron is directly photoinjected from the pyrocatechol into the conduction band of the TiO2 without the participation of excited states in the pyrocatechol.62 This direct charge transfer is believed to be an excitation from the p orbital in the pyrocatechol to the Ti 3d levels at the bottom of the conduction band of the TiO2. The three possible ways a catechol can adsorb on a TiO2 surface are shown in Fig. 6. In the bidentate chelating structure, both the oxygen atoms are bonded to the same titanium atom. In the bridging bidentate structure, each oxygen atom is bonded to a dierent titanium atom on the surface. In a monodentate structure, only one of the oxygen atoms is bonded to a titanium atom.

Redfern et al.65 have shown in theoretical calculations that the pyrocatechol molecule should adsorb on the anatase (101) surface in a bidentate bridging structure. They also showed that on a defect site (common in nanoparticles) that it would adsorb in a chelating bidentate structure. In a combined theoretical and experimental study, Li et al. investigated the correlation between bonding geometry and band gap states for pyrocatechol adsorbed on rutile TiO2 (110).66 They used STM to elucidate the bonding geometry and UV photoemission to investigate the presence of band gap states in the TiO2. In conjunction with DFT calculations they were able to interpret their experimental results. They propose that pyrocatechol adsorbs on the rutile TiO2 (110) surface in a mixed monolayer coverage of both monodentate and bridging bidentate structures and that the two can easily convert from one structure to another via proton exchange between the pyrocatechol and the surface. They also proposed that only pyrocatechol adsorbed in a bridging bidentate geometry introduces states into the band gap of the TiO2. Fig. 7 shows the calculated adsorption geometry of pyrocatechol on rutile TiO2 (110) proposed by Li et al. This arrangement shows adjacent catechol molecules tilted in opposite directions in a mixture of bridging bidentate and monodentate adsorption geometries. Using angle-resolved UPS they calculated that the pyrocatechol molecules adsorbed in a bridging bidentate geometry (which give rise to the band gap state) are tilted by 15301 from the surface normal. Liu et al.66 studied the organization of pyrocatechol on an anatase TiO2 (101) surface using STM and DFT calculations. They found that isolated pyrocatechol molecules prefer to adsorb at step defects on the surface. On terraces they found both monodentate and bidentate structures present, with monodentate favoured at low-coverage and bidenate favoured at higher coverages. They propose that monodentate pyrocatechol is mobile at room temperature and can move to preferential adsorption sites, but bidentate pyrocatechol is much less mobile. In addition, they observed the formation of onedimensional islands that change shape over time without breaking up. This is caused by individual molecules hopping to the next site along, followed shortly after by neighbouring molecules.67

Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

Fig. 6 Possible bonding modes of pyrocatechol to a TiO2 surface. A. Bidentate chelating, B. Bidentate bridging, C. Monodentate.

Fig. 7 Calculated adsorption geometry of pyrocatechol on rutile TiO2 (110) suggested by STM measurements. Adapted with permission of the author from ref. 74.

4214

Chem. Soc. Rev., 2012, 41, 42074217

This journal is

The Royal Society of Chemistry 2012

View Article Online

In a further investigation, the same group deduced that these pyrocatechol islands are responsible for a band-gap state found in the valence band spectra of the pyrocatechol-dosed TiO2 surface.68
Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

to the surface (i.e. standing up on the surface). Experimental and computational results indicated the appearance of new unoccupied states upon adsorption of dopamine, which could be due to hybridisation between the dopamine and the TiO2 surface.

7.2

Dopamine

Dopamine adsorbed on TiO2 has been widely studied for biological and environmental applications such as photodegradation of bacteria, targeted biomaterials, anti-fouling materials and bioelectronics.9,10,63,69 In these applications, dopamine is often used to anchor other molecules, such as polymer chains, to the surface of TiO2 nanoparticles. In many of these applications, dopamine facilitates electron/hole transfer across the interface between the TiO2 and the biological system. Dopamine has been employed as an anchoring molecule between DNA and TiO2 nanoparticles with an aim towards DNA-sequence recognition.10,63 The system creates a lightinduced charge separation capable of carrying information about the electronic properties of the biomolecules. The dopamine molecules in this system allow charge separation across the interface. Dopamine-modied TiO2 was found to be more ecient at charge separation than carboxyl-group-modied TiO2. Dopamine-modied TiO2 was found to be more photoactive because of the higher reducing power of the delocalized electrons and/or an increased absorption of visible light due to a shift of the absorption edge in the dopamine modied particles.70 In addition, using dopamine to anchor carboxyethylb-cyclodextrin resulted in very ecient charge separation. This is interesting for applications of dye-sensitised solar cells where dye molecules are often attached to TiO2 via carboxyl groups. As with pyrocatechol, there are few studies of dopamine on model single crystal anatase surfaces. Vega-Arroyo et al.71 carried out a theoretical study of the TiO2/dopamine-DNA system. They found that the dopamine molecules provide a strong covalent bond to the TiO2 surface and they facilitate charge separation. The dopamine/TiO2 interface also provides an electronic transition in a suitable range for photoexcitation. Calculations of dopamine on undercoordinated sites (defect sites) on the anatase TiO2 (101) surface carried out by this group,71 concluded that the dopamine molecule would adsorb in a chelating bidentate structure following deprotonation. This is in agreement with results for the pyrocatechol molecule on a defect site.65 Vega-Arroyo et al. also calculated that dissociative chelating bidentate adsorption of dopamine on corner/defects sites was more favourable than dissociative monodentate adsorption or molecular adsorption. On an anatase TiO2 (101) surface with few defects calculations suggested dopamine adsorbs in a bridging bidentate geometry.71 The adsorption of dopamine on an anatase TiO2 (101) single crystal was studied by Syres et al.72 using photoemission and NEXAFS. It was found that dopamine adsorbs in a bridging or chelating bidentate geometry. Valence band photoemission spectra indicated that the adsorption of dopamine removes the band gap state at the surface present on the clean TiO2 surface. Carbon K edge NEXAFS spectra indicated that the dopamine molecules orientate themselves with their phenyl rings normal
This journal is
c

8. Nitrogen containing compounds


Nitrogen containing compounds are of interest due to their many industrial uses in the manufacture of dyes, explosives and polymers. In addition, knowledge of how these materials interact with particular materials is also of interest in the development of sensors for detection of trace amounts of explosives. We have discussed amino acids and pyridine carboxylic acids above and here will concentrate on other nitrogen containing compounds. 8.1 Aliphatic amines Farfan-Arribas and Madix used amine adsorption to determine the Lewis acidity of surface Ti4+ ions in rutile TiO2 (110) surfaces.73 Ethylamine and diethylamine were both found to adsorb on the stoichiometric surface through formation of an NTi bond. On defected surfaces bonding also occurred at oxygen-vacancy (VO) sites. In both cases the amines remained intact but on the defected surface less amine was adsorbed. The authors suggested that this may be due to adsorption at the defect sites blocking adsorption at more than one neighbouring Ti5c site. They also found that the desorption activation energy decreased in the order diethylamine > ethylamine 4 ammonia which is also the order of decreasing Lewis basicity. This, the authors suggested, was evidence of adsorption driven by a Lewis acidbase interaction with the Ti4+ ions behaving as Lewis acids. 8.2 Pyridine Suzuki et al.74,75 have studied the adsorption of pyridine and 2,6-dimethylpyridine (2,6-DMP) on the rutile TiO2 (110) surface. Unlike the aliphatic amines described above, neither pyridine nor 2,6-DMP form a strong bond to Ti5c on terraces of the TiO2 (110) surface but are rather weakly bound with the plane of the ring parallel to the surface.75 However, an STM study of a rutile TiO2 (110) surface with single-atom height steps found that pyridine would attach to four-fold coordinated Ti atoms at the step edges. Pyridine molecules at the step edges were much less mobile than those on the terraces. The adsorbed pyridine molecules were able to exchange between these step adsorption sites and the terraces.74 8.3 Other nitroaromatics Li et al. used STM, XPS and LEED to study the conversion of azobenzene to aniline at anatase (101) and rutile (110) TiO2 surfaces.76 The interaction of these molecules with TiO2 surfaces is driven by work which showed that TiO2 supported Au nanoparticles act as a high yield catalyst for synthesis of azobenzene via oxidation of anilines and reduction of nitroaromatic compounds to aniline.77 Li et al. compared azobenzene and aniline adsorption on TiO2 and all three techniques gave almost identical results. They propose that the NQN bond in azobenzene is cleaved by the TiO2 surface resulting in a phenyl
Chem. Soc. Rev., 2012, 41, 42074217 4215

The Royal Society of Chemistry 2012

View Article Online

Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

Fig. 8 Schematic reaction model of azobenzene and aniline on TiO2 determined from STM and photoelectron spectroscopy. Adapted from ref. 84 with permission of the author. Black spheres represent carbon atoms, purple spheres are nitrogen atoms and white spheres represent hydrogen atoms.

As outlined in the introduction TiO2 is still being heavily researched for novel applications involving photochemical reactions, which will require functionalisation by new dyes or biologically active species. In biomaterials, a fundamental understanding of the success of Ti based implants can only be obtained by studying how molecules interact with TiO2 surfaces in a similar way to that achieved in catalysis. The advent of new techniques for deposition of molecules, such as electrospray methods, will allow the study of more complex molecules adsorbed on TiO2 surfaces. In addition, novel methods for preparation of TiO2 single crystal surfaces under ambient conditions are being investigated. Coupled with the availability of environmental SPM, sum frequency spectroscopy and high pressure XPS adsorption of molecules under more technologically realistic conditions can be studied. For many processes this will allow the study of adsorption in the presence of solvents, and in particular water which is likely to play a major role in adsorption in technological applications since it will interact both with the surface and the molecules of interest. Finally, the use of ultrafast lasers and X-ray and UV lasers are allowing us to probe surface chemical reactions and electron dynamics on the fs time scale. These techniques oer the opportunity to monitor adsorption and surface reactions in real time as well as monitoring charge transfer between molecular overlayers and the substrate.

References
imide species adsorbed to the surface. This can then be converted to aniline on reaction with hydrogen. Conversely, when aniline is evaporated onto the surface it loses one or more hydrogen atoms, and also adsorbs as a phenyl imide species (see Fig. 8). Hence, azobenzene and aniline give almost identical results, forming ordered superstructures of phenyl imide on the TiO2 surfaces studied. As further proof of this mechanism the authors carried out a further study using ultraviolet photoemission spectroscopy (UPS) to compare the electronic structure of azobenzene and aniline adsorbed on anatase (101) and rutile (110) TiO2 surfaces.18 They found that at saturation coverage the UPS spectra of the two adsorbed molecules are identical, proving that the NQN bond in azobenzene is cleaved by the TiO2 surface, resulting in an adsorbed phenyl imide species. In addition, they discovered that at low coverages of azobenzene adsorbed on anatase, photon irradiation converted azobenzene from a at-lying molecule to two upright phenyl imide species. They propose that the NQN bond cleavage is facilitated by a photon-induced transcis isomerization. Ramalho et al. studied the adsorption of cis- and transisomers of azobenzene on the rutile TiO2 (110) surface using DFT based methods and a periodic model.78 They found that the cis- conformation is the most stable when adsorbed. In addition, they found that the TiO2 surface is important in reducing the endothermic character of the reaction.
1 Q. Xie, Y. Y. Zhao, X. Chen, H. M. Liu, D. G. Evans and W. S. Yang, Biomaterials, 2011, 32, 65886594. 2 M. A. Henderson, Surf. Sci. Rep., 2011, 66, 185297. 3 B. O Regan and M. Gra tzel, Nature, 1991, 353, 737740. 4 B. Kasemo and J. Lausmaa, Crc Critical Reviews in Biocompatibility, 1986, 2, 335380. 5 U. Diebold, Surf. Sci. Rep., 2003, 48, 53229. 6 C. L. Pang, R. Lindsay and G. Thornton, Chem. Soc. Rev., 2008, 37, 23282353. 7 M. B. Smith and J. Michl, Chem. Rev., 2010, 110, 68916936. 8 K. Kalyanasundaram and M. Gra tzel, Curr. Opin. Biotechnol., 2010, 21, 298310. 9 T. Kotsokechagia, F. Cellesi, A. Thomas, M. Niederberger and N. Tirelli, Langmuir, 2008, 24, 69886997. 10 T. Rajh, Z. Saponjic, J. Q. Liu, N. M. Dimitrijevic, N. F. Scherer, M. Vega-Arroyo, P. Zapol, L. A. Curtiss and M. C. Thurnauer, Nano Lett., 2004, 4, 10171023. 11 S. Ko ppen, O. Bronkalla and W. Langel, J. Phys. Chem. C, 2008, 112, 1360013606. 12 X. Q. Gong, A. Selloni, M. Batzill and U. Diebold, Nat. Mater., 2006, 5, 665670. 13 M. Lazzeri, A. Vittadini and A. Selloni, Physical Review B, 2002, 65, 1. 14 W. Hebenstreit, N. Ruzycki, G. S. Herman, Y. Gao and U. Diebold, Phys. Rev. B: Condens. Matter, 2000, 62, R16334R16336. 15 G. S. Herman, Z. Dohnalek, N. Ruzycki and U. Diebold, J. Phys. Chem. B, 2003, 107, 27882795. 16 R. Hengerer, B. Bolliger, M. Erbudak and M. Gra tzel, Surf. Sci., 2000, 460, 162169. 17 A. G. Thomas, W. R. Flavell, A. K. Mallick, A. R. Kumarasinghe, D. Tsoutsou, N. Khan, C. Chatwin, S. Rayner, G. C. Smith, R. L. Stockbauer, S. Warren, T. K. Johal, S. Patel, D. Holland, A. Taleb and F. Wiame, Physical Review B, 2007, 75. 18 S. C. Li, Y. Losovyj, V. K. Paliwal and U. Diebold, J. Phys. Chem. C, 2011, 115, 1017310179. 19 L. Patthey, H. Rensmo, P. Persson, K. Westermark, L. Vayssieres, A. Stashans, A. Petersson, P. A. Bruhwiler, H. Siegbahn, S. Lunell and N. Martensson, J. Chem. Phys., 1999, 110, 59135918. 20 A. G. Thomas, W. R. Flavell, C. Chatwin, S. Rayner, D. Tsoutsou, A. R. Kumarasinghe, D. Brete, T. K. Johal, S. Patel and J. Purton, Surf. Sci., 2005, 592, 159168.

9. Conclusion and outlook


We hope that the preceding review has given the reader a avour of the numerous technologically important molecules adsorbed on TiO2 single crystal surfaces which have been studied.
4216 Chem. Soc. Rev., 2012, 41, 42074217

This journal is

The Royal Society of Chemistry 2012

View Article Online


21 P. Kruger, S. Bourgeois, B. Domenichini, H. Magnan, D. Chandesris, P. Le Fevre, A. M. Flank, J. Jupille, L. Floreano, A. Cossaro, A. Verdini and A. Morgante, Phys. Rev. Lett., 2008, 100, 055501. 22 T. J. Beck, A. Klust, M. Batzill, U. Diebold, C. Di Valentin and A. Selloni, Phys. Rev. Lett., 2004, 93, 4. 23 J. G. Tao, T. Luttrell, J. Bylsma and M. Batzill, J. Phys. Chem. C, 2011, 115, 34343442. 24 Y. B. He, O. Dulub, H. Z. Cheng, A. Selloni and U. Diebold, Physical Review Letters, 2009, 102, 4. 25 M. Lazzeri and A. Selloni, Phys. Rev. Lett., 2001, 87. 26 R. E. Tanner, A. Sasahara, Y. Liang, E. I. Altman and H. Onishi, J. Phys. Chem. B, 2002, 106, 82118222. 27 G. S. Herman, M. R. Sievers and Y. Gao, Phys. Rev. Lett., 2000, 84, 33543357. 28 Y. Liang, S. P. Gan, S. A. Chambers and E. I. Altman, Phys. Rev. B: Condens. Matter, 2001, 63. 29 R. E. Tanner, Y. Liang and E. I. Altman, Surf. Sci., 2002, 506, 251271. 30 M. Aizawa, Y. Morikawa, Y. Namai, H. Morikawa and Y. Iwasawa, J. Phys. Chem. B, 2005, 109, 1883118838. 31 A. Gutierrez-Sosa, P. Martinez-Escolano, H. Raza, R. Lindsay, P. L. Wincott and G. Thornton, Surf. Sci., 2001, 471, 163169. 32 Q. Guo, I. Cocks and E. M. Williams, Surf. Sci., 1997, 393, 111. 33 Q. Guo and E. M. Williams, Surf. Sci., 1999, 433435, 322326. 34 J. Schnadt, J. N. OShea, L. Patthey, J. Schiessling, J. Krempasky, rtensson and P. A. Bruhwiler, Surf. Sci., 2003, 544, M. Shi, N. Ma 7486. 35 J. Schnadt, J. Schiessling, J. N. OShea, S. M. Gray, L. Patthey, M. K. J. Johansson, M. Shi, J. Krempasky, J. Ahlund, rtensson and P. A. Bruhwiler, P. G. Karlsson, P. Persson, N. Ma Surf. Sci., 2003, 540, 3954. 36 J. N. OShea, Y. Luo, J. Schnadt, L. Patthey, H. Hillesheimer, J. Krempasky, D. Nordlund, M. Nagasono, P. A. Bruhwiler and N. Martensson, Surf. Sci., 2001, 486, 157166. 37 P. R. McGill and H. Idriss, Surf. Sci., 2008, 602, 36883695. 38 E. L. Quah, J. N. Wilson and H. Idriss, Langmuir, 2010, 26, 64116417. 39 G. J. Fleming, K. Adib, J. A. Rodriguez, M. A. Barteau and H. Idriss, Surf. Sci., 2007, 601, 57265731. 40 A. Fahmi, C. Minot, P. Fourre and P. Nortier, Surf. Sci., 1995, 343, 261272. 41 I. Dolamic and T. Burgi, J. Phys. Chem. B, 2006, 110, 1489814904. 42 H. Ashima, W. J. Chun and K. Asakura, Surf. Sci., 2007, 601, 18221830. 43 S. P. Bates, G. Kresse and M. J. Gillan, Surf. Sci., 1998, 409, 336349. 44 J. N. Wilson, D. J. Titheridge, L. Kieu and H. Idriss, J. Vac. Sci. Technol., A, 2000, 18, 18871892. 45 E. M. J. Johansson, S. Plogmaker, L. E. Walle, R. Scholin, A. Borg, A. Sandell and H. Rensmo, J. Phys. Chem. C, 2010, 114, 1501515020. 46 L. C. Mayor, J. Ben Taylor, G. Magnano, A. Rienzo, C. J. Satterley, J. N. OShea and J. Schnadt, J. Chem. Phys., 2008, 129. 47 A. Sasahara, C. L. Pang and H. Onishi, J. Phys. Chem. B, 2006, 110, 47514755. 48 K. L. Syres, A. G. Thomas, D. J. H. Cant, S. J. O. Hardman and A. Preobrajenski, Surf. Sci., 2011, 606, 273277. 49 M. Haukka and P. Hirva, Surf. Sci., 2002, 511, 373378. 50 J. Schnadt, P. A. Bruhwiler, L. Patthey, J. N. OShea, S. Sodergren, M. Odelius, R. Ahuja, O. Karis, M. Bassler, P. Persson, H. Siegbahn, S. Lunell and N. Martensson, Nature, 2002, 418, 620623. 51 K. Kanie and T. Sugimoto, Chem. Commun., 2004, 15841585. 52 E. Soria, I. Colera, E. Roman, E. M. Williams and J. L. de Segovia, Surf. Sci., 2000, 451, 188196. 53 E. Soria, E. Roman, E. M. Williams and J. L. de Segovia, Surf. Sci., 1999, 433, 543548. 54 G. J. Fleming, K. Adib, J. A. Rodriguez, M. A. Barteau, J. M. White and H. Idriss, Surf. Sci., 2008, 602, 20292038. 55 G. J. Fleming and H. Idriss, Langmuir, 2004, 20, 75407546. 56 T. Z. Qiu and M. A. Barteau, J. Colloid Interface Sci., 2006, 303, 229235. 57 J. N. Wilson, R. M. Dowler and H. Idriss, Surf. Sci., 2011, 605, 206213. 58 R. Tonner, ChemPhysChem, 2010, 11, 10531061. 59 D. Szieberth, A. Maria Ferrari and X. Dong, Phys. Chem. Chem. Phys., 2010, 12, 1103311040. 60 A. G. Thomas, W. R. Flavell, C. P. Chatwin, A. R. Kumarasinghe, S. M. Rayner, P. F. Kirkham, D. Tsoutsou, T. K. Johal and S. Patel, Surf. Sci., 2007, 601, 38283832. 61 K. Onda, B. Li, J. Zhao and H. Petek, Surf. Sci., 2005, 593, 3237. 62 P. Persson, R. Bergstrom and S. Lunell, J. Phys. Chem. B, 2000, 104, 1034810351. 63 J. Q. Liu, L. de la Garza, L. G. Zhang, N. M. Dimitrijevic, X. B. Zuo, D. M. Tiede and T. Rajh, Chem. Phys., 2007, 339, 154163. 64 W. R. Duncan and O. V. Prezhdo, J. Phys. Chem. B, 2005, 109, 365373. 65 P. C. Redfern, P. Zapol, L. A. Curtiss, T. Rajh and M. C. Thurnauer, J. Phys. Chem. B, 2003, 107, 1141911427. 66 S. C. Li, J. G. Wang, P. Jacobson, X. Q. Gong, A. Selloni and U. Diebold, J. Am. Chem. Soc., 2009, 131, 980984. 67 S. C. Li, L. N. Chu, X. Q. Gong and U. Diebold, Science, 2010, 328, 882884. 68 S. C. Li, Y. Losovyj and U. Diebold, Langmuir, 2011, 27, 86008604. 69 N. M. Dimitrijevic, E. Rozhkova and T. Rajh, J. Am. Chem. Soc., 2009, 131, 28932899. 70 N. M. Dimitrijevic, Z. V. Saponjic, D. M. Bartels, M. C. Thurnauer, D. M. Tiede and T. Rajh, J. Phys. Chem. B, 2003, 107, 73687375. 71 M. Vega-Arroyo, P. R. LeBreton, T. Rajh, P. Zapol and L. A. Curtiss, Chem. Phys. Lett., 2005, 406, 306311. 72 K. Syres, A. Thomas, F. Bondino, M. Malvestuto and M. Gra tzel, Langmuir, 2010, 26, 1454814555. 73 E. Farfan-Arribas and R. J. Madix, J. Phys. Chem. B, 2003, 107, 32253233. 74 S. Suzuki, Y. Yamaguchi, H. Onishi, K. Fukui, T. Sasaki and Y. Iwasawa, Catal. Lett., 1998, 50, 117123. 75 S. Suzuki, Y. Yamaguchi, H. Onishi, T. Sasaki, K. Fukui and Y. Iwasawa, J. Chem. Soc., Faraday Trans., 1998, 94, 161166. 76 S. C. Li and U. Diebold, J. Am. Chem. Soc., 2010, 132, 64+. 77 A. Grirrane, A. Corma and H. Garcia, Science, 2008, 322, 16611664. 78 J. P. Prates Ramalho and F. Illas, Chem. Phys. Lett., 501, 379384.

Published on 19 April 2012. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 29/07/2013 15:22:01.

This journal is

The Royal Society of Chemistry 2012

Chem. Soc. Rev., 2012, 41, 42074217

4217

You might also like