Electron traps and their effect on the surface chemistry of TiO … · Electron traps and their...

6
Electron traps and their effect on the surface chemistry of TiO 2 ð110Þ Anthoula C. Papageorgiou a , Nikolaos S. Beglitis b , Chi L. Pang a , Gilberto Teobaldi c , Gregory Cabailh a , Qiao Chen a,d , Andrew J. Fisher b , Werner A. Hofer c,1 , and Geoff Thornton a,1 a London Centre for Nanotechnology and Department of Chemistry, University College London, 17-19 Gordon Street, London WC1H 0AH, UK; b London Centre for Nanotechnology and Department of Physics, University College London, 17-19 Gordon Street, London WC1H 0AH, UK; c Surface Science Research Centre, Department of Chemistry, University of Liverpool, Liverpool L69 3BX , UK; and d Department of Chemistry, School of Life Sciences, University of Sussex, Falmer, Brighton BN1 9QJ, UK Edited by Nicholas J. Turro, Columbia University, New York, NY, and approved December 4, 2009 (received for review October 2, 2009) Oxygen vacancies on metal oxide surfaces have long been thought to play a key role in the surface chemistry. Such processes have been directly visualized in the case of the model photocatalyst sur- face TiO 2 ð110Þ in reactions with water and molecular oxygen. These vacancies have been assumed to be neutral in calculations of the surface properties. However, by comparing experimental and simulated scanning tunneling microscopy images and spectra, we show that oxygen vacancies act as trapping centers and are ne- gatively charged. We demonstrate that charging the defect signif- icantly affects the reactivity by following the reaction of molecular oxygen with surface hydroxyl formed by water dissociation at the vacancies. Calculations with electronically charged hydroxyl favor a condensation reaction forming water and surface oxygen adatoms, in line with experimental observations. This contrasts with simula- tions using neutral hydroxyl where hydrogen peroxide is found to be the most stable product. photocatalysis STM/STS titania water molecular oxygen The rutile TiO 2 ð110Þ surface, which we use as a model photoca- talytic system here, is displayed as a ball model in Fig. 1A. It is characterized by alternate rows of fivefold coordinated Ti (Ti 5c ) and bridging-O atoms (O b ) that run in the [001] direction. Fig. 1B shows a typical scanning tunneling microscopy (STM) image of the surface. The Ti 5c rows appear bright, and the O b rows appear dark. Point defects are common on this surface, the vast majority being O b vacancies (O b -vac) and surface hydroxyls (OH b ). As can be seen in Fig. 1B, these defects appear as bright spots between the bright Ti 5c rows. O b -vac are known to be particularly reactive, with STM images providing compelling evidence that O 2 dissoci- ates at O b -vac (1, 2) leaving a healed vacancy and one oxygen adatom (O ad ) on Ti 5c according to ref. 3: 2Ti 3þ þ O b -vac þ O 0 2 ðgÞ 2Ti 4þ þ O 2b þ O 0 ad [1] where the reduction of one oxygen atom of O 2 ðgÞ to one bridging oxide species (O 2b ) is accomplished by oxidation of the two Ti 3þ sites associated with O b -vac to Ti 4þ (3), on the basis of a purely ionic model. (Formal charges are written in reactions 1 and 2 to highlight the redox processes involved.) The interaction of O 2 with OH b , on the other hand, is still a matter of controversy. Following the reaction of these species at temperatures 240 K, water is seen to desorb at 310 K in tem- perature programmed desorption (TPD) spectra (3, 4). Hender- son et al. (3) concluded that this water evolution is a consequence of the formation of oxygen adatoms (O ad ) at the surface as follows: 2Ti 3þ þ 2OH 1b þ O 0 2 ðgÞ 2Ti 4þ þ H 2 O 0 ðgÞþ 2O 2b þ O 0 ad [2] where the two Ti 3þ species provide the two electrons necessary to reduce one oxygen atom of O 2 ðgÞ to H 2 OðgÞ (3). In stark contrast to the TPD results, previous calculations find H 2 O 2 to be by far the most stable product (5). Moreover, on the basis of these calculations, water desorption is not expected up to the highest temperature computed, 350 K (5). This discrepancy provided the initial motivation for the present work. Results and Discussion We use STM to provide an additional experimental test of the picture that has emerged thus far. Fig. 1B shows a surface con- taining both O b -vac and OH b , alongside the same surface in Fig. 1C after it was exposed to 90 Langmuirs (L) O 2 at 300 K (1 L ¼ 1.33 × 10 6 mbar · s, 1 mbar ¼ 100 Pa). A number of small, bright spots can be seen on the Ti 5c sites (bright rows) in the latter image. The histogram of the height distribution of these bright spots, shown in Fig. 1D, indicates that these bright spots are almost entirely due to one final product. It should be noted that at lower O 2 exposures we see a number of different types of species on Ti 5c rows that are likely to arise from terminal hydroxyls (OH t ) and other metastable species such as O 2 H. These latter results are consistent with previous work (4, 6). Analysis of the images in Fig. 1 shows that before the reaction with O 2 there were about 72 OH b species and 38 O b -vacs. Fol- lowing the reaction with O 2 , about 118 new bright spots are seen on the bright Ti 5c rows in Fig. 1C. As it is well known that O ad form after exposure of O b -vac to O 2 (1, 2, 79), we conclude that O ad also forms from exposure of OH b to O 2 , in line with the conclusions of Henderson et al. (3). Moreover, the availability of only 72 H atoms rules out the possibility that the products in the present case are OH t (4, 6). There is, however, a subtle but crucial discrepancy between our observations and the conclusions of Henderson et al. (3). With a starting point of 38 O b -vacs and 72 OH b , we would expect each O b -vac to yield one O ad (reaction 1) and each OH b to yield half an O ad (reaction 2), i.e., only 74 O ad in total compared to the 118 found in the experiment. Because the availability of electrons from reduced TiO 2 is a prerequisite for O 2 ðgÞ dissociative adsorp- tion on TiO 2 ð110Þ (3, 5), the extra O ad found in Fig. 1C require the TiO 2 sample to supply more electrons than expected on the basis of the number of O b -vac (2Ti 3þ ) and OH b (1Ti 3þ ) initially present. Thus, the assignment of the adsorbates in Fig. 1C to O ad and the interpretation of TPD results (3) challenge the current Author contributions: A.J.F., W.A.H., and G. Thornton designed research; A.C.P., N.S.B., C.L.P., G. Teobaldi, G.C., and Q.C. performed research; A.C.P., N.S.B., C.L.P., G. Teobaldi, G.C., and Q.C. analyzed data; and A.C.P., N.S.B., C.L.P., G. Teobaldi, G.C., Q.C., A.J.F., W.A.H., and G. Thornton wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. 1 To whom correspondence should be addressed. E-mail:[email protected] or [email protected]. This article contains supporting information online at www.pnas.org/cgi/content/full/ 0911349107/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.0911349107 PNAS February 9, 2010 vol. 107 no. 6 23912396 CHEMISTRY Downloaded by guest on October 13, 2020

Transcript of Electron traps and their effect on the surface chemistry of TiO … · Electron traps and their...

Page 1: Electron traps and their effect on the surface chemistry of TiO … · Electron traps and their effect on the surface chemistry of TiO2ð110Þ Anthoula C. Papageorgioua, Nikolaos

Electron traps and their effect on thesurface chemistry of TiO2ð110ÞAnthoula C. Papageorgioua, Nikolaos S. Beglitisb, Chi L. Panga, Gilberto Teobaldic, Gregory Cabailha, Qiao Chena,d,Andrew J. Fisherb, Werner A. Hoferc,1, and Geoff Thorntona,1

aLondon Centre for Nanotechnology and Department of Chemistry, University College London, 17-19 Gordon Street, London WC1H 0AH, UK; bLondonCentre for Nanotechnology and Department of Physics, University College London, 17-19 Gordon Street, London WC1H 0AH, UK; cSurface ScienceResearch Centre, Department of Chemistry, University of Liverpool, Liverpool L69 3BX , UK; and dDepartment of Chemistry, School of Life Sciences,University of Sussex, Falmer, Brighton BN1 9QJ, UK

Edited by Nicholas J. Turro, Columbia University, New York, NY, and approved December 4, 2009 (received for review October 2, 2009)

Oxygen vacancies on metal oxide surfaces have long been thoughtto play a key role in the surface chemistry. Such processes havebeen directly visualized in the case of the model photocatalyst sur-face TiO2ð110Þ in reactions with water and molecular oxygen.These vacancies have been assumed to be neutral in calculationsof the surface properties. However, by comparing experimentaland simulated scanning tunneling microscopy images and spectra,we show that oxygen vacancies act as trapping centers and are ne-gatively charged. We demonstrate that charging the defect signif-icantly affects the reactivity by following the reaction of molecularoxygen with surface hydroxyl formed by water dissociation at thevacancies. Calculationswith electronically charged hydroxyl favor acondensation reaction formingwater and surface oxygen adatoms,in line with experimental observations. This contrasts with simula-tions using neutral hydroxyl where hydrogen peroxide is found tobe the most stable product.

photocatalysis ∣ STM/STS ∣ titania ∣ water ∣ molecular oxygen

The rutile TiO2ð110Þ surface, which we use as a model photoca-talytic system here, is displayed as a ball model in Fig. 1A. It ischaracterized by alternate rows of fivefold coordinated Ti (Ti5c)and bridging-O atoms (Ob) that run in the [001] direction. Fig. 1Bshows a typical scanning tunneling microscopy (STM) image ofthe surface. The Ti5c rows appear bright, and the Ob rows appeardark. Point defects are common on this surface, the vast majoritybeing Ob vacancies (Ob-vac) and surface hydroxyls (OHb). As canbe seen in Fig. 1B, these defects appear as bright spots betweenthe bright Ti5c rows. Ob-vac are known to be particularly reactive,with STM images providing compelling evidence that O2 dissoci-ates at Ob-vac (1, 2) leaving a healed vacancy and one oxygenadatom (Oad) on Ti5c according to ref. 3:

2Ti3þ þOb-vacþO02ðgÞ → 2Ti4þ þO2−

b þO0ad [1]

where the reduction of one oxygen atom of O2ðgÞ to one bridgingoxide species (O2−

b ) is accomplished by oxidation of the two Ti3þsites associated with Ob-vac to Ti4þ (3), on the basis of a purelyionic model. (Formal charges are written in reactions 1 and 2 tohighlight the redox processes involved.)

The interaction of O2 with OHb, on the other hand, is still amatter of controversy. Following the reaction of these species attemperatures ≤240 K, water is seen to desorb at ∼310 K in tem-perature programmed desorption (TPD) spectra (3, 4). Hender-son et al. (3) concluded that this water evolution is a consequenceof the formation of oxygen adatoms (Oad) at the surface asfollows:

2Ti3þ þ 2OH1−b þO0

2ðgÞ → 2Ti4þ þH2O0ðgÞ þ 2O2−b þO0

ad

[2]

where the two Ti3þ species provide the two electrons necessary toreduce one oxygen atom of O2ðgÞ to H2OðgÞ (3). In stark contrast

to the TPD results, previous calculations find H2O2 to be by farthe most stable product (5). Moreover, on the basis of thesecalculations, water desorption is not expected up to the highesttemperature computed, 350 K (5). This discrepancy provided theinitial motivation for the present work.

Results and DiscussionWe use STM to provide an additional experimental test of thepicture that has emerged thus far. Fig. 1B shows a surface con-taining both Ob-vac and OHb, alongside the same surface inFig. 1C after it was exposed to 90 Langmuirs (L) O2 at 300 K(1 L ¼ 1.33 × 10−6 mbar · s, 1 mbar ¼ 100 Pa). A number ofsmall, bright spots can be seen on the Ti5c sites (bright rows)in the latter image. The histogram of the height distribution ofthese bright spots, shown in Fig. 1D, indicates that these brightspots are almost entirely due to one final product.

It should be noted that at lower O2 exposures we see a numberof different types of species on Ti5c rows that are likely to arisefrom terminal hydroxyls (OHt) and other metastable species suchas O2H. These latter results are consistent with previous work(4, 6).

Analysis of the images in Fig. 1 shows that before the reactionwith O2 there were about 72 OHb species and 38 Ob-vacs. Fol-lowing the reaction with O2, about 118 new bright spots are seenon the bright Ti5c rows in Fig. 1C. As it is well known that Oadform after exposure of Ob-vac to O2 (1, 2, 7–9), we conclude thatOad also forms from exposure of OHb to O2, in line with theconclusions of Henderson et al. (3). Moreover, the availabilityof only 72 H atoms rules out the possibility that the productsin the present case are OHt (4, 6).

There is, however, a subtle but crucial discrepancy between ourobservations and the conclusions of Henderson et al. (3). With astarting point of 38 Ob-vacs and 72 OHb, we would expect eachOb-vac to yield one Oad (reaction 1) and each OHb to yield half anOad (reaction 2), i.e., only 74 Oad in total compared to the 118found in the experiment. Because the availability of electronsfrom reduced TiO2 is a prerequisite for O2ðgÞ dissociative adsorp-tion on TiO2ð110Þ (3, 5), the extra Oad found in Fig. 1C requirethe TiO2 sample to supply more electrons than expected on thebasis of the number of Ob-vac (2Ti3þ) and OHb (1Ti3þ) initiallypresent.

Thus, the assignment of the adsorbates in Fig. 1C to Oad andthe interpretation of TPD results (3) challenge the current

Author contributions: A.J.F., W.A.H., and G. Thornton designed research; A.C.P., N.S.B.,C.L.P., G. Teobaldi, G.C., and Q.C. performed research; A.C.P., N.S.B., C.L.P., G. Teobaldi,G.C., and Q.C. analyzed data; and A.C.P., N.S.B., C.L.P., G. Teobaldi, G.C., Q.C., A.J.F.,W.A.H., and G. Thornton wrote the paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.1To whom correspondence should be addressed. E-mail:[email protected] [email protected].

This article contains supporting information online at www.pnas.org/cgi/content/full/0911349107/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.0911349107 PNAS ∣ February 9, 2010 ∣ vol. 107 ∣ no. 6 ∣ 2391–2396

CHEM

ISTR

Y

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0

Page 2: Electron traps and their effect on the surface chemistry of TiO … · Electron traps and their effect on the surface chemistry of TiO2ð110Þ Anthoula C. Papageorgioua, Nikolaos

theoretical understanding of the surface chemistry of TiO2ð110Þ.Even using a hybrid HSE06 approach (10, 11), which is expectedto describe more accurately than other functionals the electronicstructure of metal oxides, we also predict that H2O2 is over-whelmingly the most stable product of the reaction betweenO2 and OHb. This result, which is inconsistent with the TPD data,mimics that found in previous simulations (5).

It turns out that we can only model Oad as the most stableproduct when we allow excess electronic charge to accumulateat the Ob-vac and OHb (see Fig. 2), the latter being formed byreaction of water molecules with Ob-vac. The presence of extraelectrons at Ob-vac∕OHb originates from occupation of addi-tional 3d electron states arising from polaronic distortion at bothOb-vac and OHb sites. As we discuss later, the excess electroniccharge is not pinned directly at the Ob-vac but is rather foundlocalized at a number of reduced Ti sites around the vacancy(see also Fig. S1). High affinities for both surface and bulk elec-tron polaron trapping have been reported for substrates such asHfO2 (13) and SiO2 (14). This has been used to explain a numberof surface properties, such as SiO2 discharge phenomena,although not the surface chemistry.

Various spectroscopies suggest that Ob-vac introduces elec-tronic states into the band gap about 1 eV below the conductionband (CB) onset (3, 15, 16). These band gap states (BGS) areknown to have d character (15). The BGS persist even whenthe Ob-vac are replaced with OHb formed by water dissociationat the Ob-vac (3, 15). In our calculations, regardless of the numb-er of electrons occupying the BGS [2 electrons (e) for Ob-vac(0),3e for Ob-vacð1−Þ, 4e for Ob-vacð2−Þ, 1e for OHbð0Þ, 2e forOHbð1−Þ, 3e for OHbð2−Þ] we find that the energy remains closeto 1 eV below the CB onset (Fig. 2A). The extra electrons, how-ever, drastically affect the spatial distribution of the associated 3delectronic density. While for Ob-vac(0), the simulations suggest asubsurface localization immediately beneath the defect site (Tisin Fig. 2B), the presence of one and two extra electrons inducesan electronic reorganization that moves electronic charge awayfrom the Ob-vac site and the corresponding Ob row (see Fig. 2Dand E). In line with other hybrid DFTresults for Ob-vac in high-kdielectrics (13, 17), an analysis of the Ob-vac induced changes inthe global electronic distribution reveals symmetry breaking atthe defect site (details are shown in Fig. S2). This is discussedin more detail below.

We now consider the possible origin of the additional electron-ic charge. Given the ease of electron transfer in TiO2 (18, 19) withreported barriers as low as 0.09 eV and mobilities as high as5.24 × 10−2 cm2 V−2 s−1, it is reasonable to expect interstitial Ti(Tiint)-donated electrons to diffuse comparatively freely withinthe system (20). To test the hypothesis that they will, in this case,diffuse toward the Ob-vac∕OHb sites at the surface, we also mod-eled a much larger unit cell (8 trilayers) with inclusion of oneinitially neutral Tiint atom. The increased size of the simulatedsystems here (288 atoms for the clean slab) prevent us fromapplying HSE06 so we chose instead to use a local spin density

Fig. 1. Reaction of O2 with TiO2ð110Þ. (A) Ball model of TiO2ð110Þ. Red andblue spheres denote O and Ti, respectively. The pink spheres are bridging Oatoms, which lie in the [001] azimuth of the substrate. Parallel Ti rows that liebetween the bridging-O rows are fivefold coordinated Ti atoms. Greenspheres indicate H atoms (from OHb). (B) 130 × 170 Å2 STM image(V ¼ 1.5 V, I ¼ 0.25 nA) of an as-prepared TiO2ð110Þ surface that containsOb-vac and OHb. OHb forms from dissociation of water from the residualvacuum at Ob-vac. An Ob-vac, an OHb, and an OHb pair are indicated. (C)The surface in B after exposure to ∼90 L O2 at 300 K. One of the bright spotsassigned to Oad is circled. B and C have been smoothed using Image SXM (12)v.1.75. (D) A histogram showing the height distribution of 276 bright spotsfound on the Ti5c rows fitted to two Gaussian curves. The data are taken froman unsmoothed, larger version of the image in C. The histogram indicatesthat the reaction products are almost entirely from one species.

Fig. 2. Calculated electronic structure of Ob-vac and OHb. (A) The total den-sity of states for the optimized layers in the presence of different amounts ofextra electronic charge. (B) The TiO2ð110Þ surface shown as a stick modelwhere the blue intersections indicate Ti sites and the red intersections indi-cate O sites. The arrow points at the Ob-vac and each Ti atom is labeled. Theglobal charge density of the BGS is shown in yellow (10−6 eÅ−3) for Ob-vac(0),Ob-vacð1−Þ, and Ob-vacð2−Þ in C, D, and E, respectively.

2392 ∣ www.pnas.org/cgi/doi/10.1073/pnas.0911349107 Papageorgiou et al.

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0

Page 3: Electron traps and their effect on the surface chemistry of TiO … · Electron traps and their effect on the surface chemistry of TiO2ð110Þ Anthoula C. Papageorgioua, Nikolaos

approximation ðLSDAÞ þU scheme already successfully appliedto Ob-vac on TiO2ð110Þ (21).

Fig. 3 shows the BGS spatial distribution for Ob-vac bothisolated and in the presence of one Tiint. Recently, Tiint specieshave been suggested as the main origin of titania BGS (9) withnegligible importance of Ob-vac (or OHb) sites. In contrast, ourresults shown in Fig. 3C andD demonstrate that a combination ofTiint and Ob-vac induce new BGS. For isolated Ob-vac, the BGS(and the associated Ti3þ sites) are symmetrically localized at orimmediately beneath the defect site (Fig. 3A). For one isolatedTiint (Fig. 3B), the calculations suggest an asymmetric localisationof the BGS at the Tiint site (one state) and on other separate,regular lattice Ti sites across the slab. Following refs. 22 and23, we assign the appearance of localized BGS, and the ensuinglocal rehybridization to the chemical reduction of specific Ti sitesfrom their stoichiometric oxidation state (i.e., reduction fromTi4þ to Ti3þ in an oversimplified ionic model).

Upon consideration of an Ob-vac together with one Tiint, cou-pling between the vacancy site and the Tiint-donated electrons isevidenced by the large changes in the BGS eigenvalues (Fig. 3D).By comparing the formation energies of an isolated Ob-vac(þ5.11 eV) and a Tiint-coupled vacancy (þ2.72 eV), the simula-tions suggest a net stabilization of 2.39 eV for Ob-vac within theadopted simulation cell. The concentration of Tiint modeled isabout an order of magnitude higher than experimental estimatesfrom the literature (20). This allows us to capture the qualitativebehavior but prevents us from addressing quantitatively thenumber of electrons globally transferred to the surface. Given

that the Tiint concentration simulated is about an order of mag-nitude greater than experimental estimates (20), this couplingshould be even more favorable in reality because the electrostaticrepulsion between BGS is expected to be much lower.

This BGS coupling drastically affects both the energy and, aswas the case for the HSE06 modeling for Ob-vacð2−Þ (Fig. 2), thespatial distribution of the BGS of Ob-vac. The presence of Tiint-donated BGS induces an electronic reorganization that moveselectronic charge away from the vacancy site and the correspond-ing Ob row (Fig. 3C). Note that this is a different role for Tiintthan that claimed in an earlier study, where electronic states as-sociated with Tiint were responsible for the BGS (9).

Analysis of the individual BGS for the coupled systems pro-vides insight into how the Tiint-donated electrons couple toOb-vac sites. As shown in Fig. 3C–F, when Tiint-donated electronsare modeled together with Ob-vac, hybridization between specificBGS (circled in Fig. 3C, E, and F) make Tiint-donated electronsavailable at the Ob-vac site thus locally charging Ob-vac with theextra electrons necessary to account for the additional Oadimaged in Fig. 1C. The same mechanism is also found for theinteraction of Tiint-donated electrons with OHb. In this case,comparison between the formation energies of OHb(þ0.32 eV) and OHb∕Tiint (−1.05 eV) suggests a net stabilizationof 1.37 eV. As with the Ob-vac∕Tiint system, the Tiint electronicsurplus is also characterized by symmetry breaking and hybridi-zation between specific BGS (circled in Fig. S3).

Although we are unable to address directly the number of elec-trons globally transferred to the surface, we can make estimatesbased on the upper limit of reported Tiint concentration:2 × 1019 cm−3, i.e., one per 800 TiO2 unit cells (20). In our esti-mation, we assume typical Ob-vac∕OHb defect concentrations of5 × 1013 cm−2 on a 1.0 × 0.5 × 0.1 cm3 sample, and we considerthat each Tiint generates four Ti3þ sites (Fig. 3B and ref. 24). Thismeans that to singly (doubly) charge Ob-vac∕OHb would onlyrequire the participation of 0.0017% (0.0034%) of the globalnumber of Tiint-donated electrons.

To further validate our model of the electronic structure, weprobed the electronic charge distribution associated with Ob-vac experimentally using scanning tunneling spectroscopy(STS) at ∼5 K. This low temperature was chosen to deliberatelyfreeze out polaronic hopping effects that are thought to averageout asymmetries that are predicted to arise at 0 K (17, 18, 25).During the acquisition of the STM image shown in Fig. 4A,the tip was immobilized at each point of the scan while tunnelingcurrent versus bias voltage (I-V) spectra were recorded. Thismethod of recording I-V spectra alongside STM images is knownas current imaging tunneling spectroscopy (CITS) (16), and itallows the I-V spectra to be displayed as current maps at eachvoltage.

The STM image in Fig. 4A was acquired at a sample bias ofþ2 V with a tunneling current of 0.03 nA. It clearly shows bothOHb and Ob-vac as bright spots between bright Ti5c rows. Thecurrent map atþ2 V has a similar appearance to the STM image,as one would expect. However, the current map at −2 V (Fig. 4B)has a very different appearance. Ob-vac and OHb both appear asdark spots centered between two bright rows. Bright features canbe seen on the bright Ti5c rows and a correlation analysis (Fig. 4C)shows that these lie diagonally adjacent to the Ob-vacs in thepositions of the second nearest Ti5c neighbors. Each Ob-vac tendsto be surrounded by one or two brighter features rather than fourequally bright features. Fig. 4D shows normalized conductancespectra ðdI∕dVÞ × ðV∕IÞ that approximate the local density ofstates. Each curve represents an average of 360 spectra corre-sponding to the bright features, Ob-vac, OHb, Ti5c, and Ob. Simi-lar spectra were recorded using the same tip with the tunnelingcurrent set between 0.015 nA and 0.7 nA. We also recordedsome spectra using an iridium tip. In neither case did we observea shift of features, which suggests that the effect produced by

Fig. 3. Electronic charge density distribution (10−5 eÅ−3) for the BGS of theconsidered 8 trilayer systems. (A) Ob-vac, (B) one Tiint between the fourth andfifth trilayers, (C) Ob-vac modeled together with one Tiint between the fourthand fifth TiO2ð110Þ trilayers (Ob-vacþ Tiint). Blue intersections indicate Tiatoms and red intersections O atoms. Tiint is shown as a blue sphere. The posi-tions of Ob-vac and Tiint are marked by the black arrows. Different colors havebeen used to distinguish between Tiint-donated (black, gray, green, orange),and surface Ob-vac induced (purple, blue) BGS. (D) Single-state energy leveldiagram with respect to the CB onset (E-ECB� ¼ 0) for the BGS in A–C aftervacuum level electrostatic alignment. Up and down arrows refer to themodeled spin of the specific state. The same BGS color labelling has beenused for all the displayed A–F panels. (E and F) For clarity, the hybridizedBGS of Ob-vacþ Tiint (orange and purple) circled in C, are redisplayed withthe omission of the orange in E and purple in F BGS.

Papageorgiou et al. PNAS ∣ February 9, 2010 ∣ vol. 107 ∣ no. 6 ∣ 2393

CHEM

ISTR

Y

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0

Page 4: Electron traps and their effect on the surface chemistry of TiO … · Electron traps and their effect on the surface chemistry of TiO2ð110Þ Anthoula C. Papageorgioua, Nikolaos

tip-induced band-bending is negligible. These STS data are inbroad agreement with recently reported 78 K results (27).

Fig. 5 shows our STM and current map simulations of Ob-vacat þ2 V and −2 V, respectively. These are derived from theHSE06 density of states. The simulations at þ2 V are similarto each other, the highest current being found at the Ob-vac(Fig. 5A, C, and E). In contrast, the simulations at −2 V arestrongly dependent on the Ob-vac charge. For both Ob-vac(0)and Ob-vacð1−Þ, the highest current is found in a spot (or spots)directly in line with the bridging O row, inconsistent with the ex-perimental contrast. The experimental distribution of the defectstates is only reproduced in simulations of Ob-vacð2−Þ where thehighest current is found on a second nearest Ti5c neighbor withlittle current along the Ob row. The current map for Ob-vacð2−Þat −2 V was deconstructed by separating the contributions into“slices” with energy windows of 0.2 eV (Fig. S4). In this way, wefind that the asymmetry originates between energies of 1.1 and1.5 eV, with the majority signal between 1.1 and 1.3 eV. This iden-tifies the asymmetry in the current map with the BGS introducedby Ob-vacð2−Þ. This asymmetry is not observed in data recorded

at 78 K (27). Presumably this arises because of significant polaro-nic hopping of the type recently identified at room tempera-ture (25).

STM and current map simulations based on the LSDA þUtreatment of Tiint∕Ob-vac introduced above (Fig. 3C) also fitthe experimental results well. The asymmetry can again be tracedto the BGS of Ob-vac, in this case coupled to the Tiint donatedelectrons (Fig. S5, S6).

We are now in a position to reexamine the theoretical descrip-tion of the reactivity of O2 with Ob-vac and OHb in the presenceof Tiint-donated electron surplus. The use of the HSE06 (10, 11)approach allows us to accurately describe the relative energies ofmolecular H2OðgÞ, O2ðgÞ, H2ðgÞ. Consequently, the level ofaccuracy of the calculated chemical potential and then thegrand-canonical formation energies are expected to be sufficientlyhigh to quantitatively address the thermodynamic stability of theconsidered systems (SI Text, Table SI).

To match with our STM experiments, we use the results of ourcalculations from a simulated temperature of 300 K, but we notethat the same trends are found at a large range of temperaturesaround 300 K. The formation energies of the pertinent adsor-bates (H2O2, H2O, O2H, OHt, and Oad) are shown in Fig. 6alongside those for Ob-vac and OHb. We stress that owing tothe necessity of enforcing a compensating background in thesimulation cells, only relative energy differences for one specificcharge state can be meaningfully compared. To illustrate thispoint, the energy of H2O2ð2−Þ can be compared with Oadð2−Þor O2Hð2−Þ but not with H2O2ð0Þ or Oadð0Þ.

The calculations indicate that when O2 reacts with Ob-vacð2−Þ(ΔGf ¼ 0.41 eV), O2 dissociates, filling the vacancy and formingone Oadð2−Þ (ΔGf ¼ −16.06 eV) with ΔG ¼ −16.47 eV[Ob-vacð2−Þ þObðgÞ → Oadð2−Þ þOb], consistent with experi-ments (1, 2). As for the reaction of O2 with OHb, we find that

Fig. 5. Modeled STM and CITS appearance for Ob-vac. Simulated STMimages (þ2 V, 10−7 eÅ−3) and current maps (−2 V, same height above thesurface as from the corresponding left-side topography) are shown forOb-vac(0) in A and B, for Ob-vacð1−Þ in C and D, and for Ob-vacð2−Þ in Eand F. Ti5c rows are indicated by black lines, and an X marks the Ob-vac.STM simulations were performed with a tip-surface distance of ∼5 Å [thedetailed procedure can be found in ref. (28)].

Fig. 4. Experimental STM and STS data. (A) ð44 ÅÞ2 STM image recordedsimultaneously with the STS. (B) ð44 ÅÞ2 CITS current map at −2 V. Thesquares in A and B show the positions of Ob-vac (green), OHb (purple),and some bright features associated with Ob-vac (black). One impurity is alsopresent and marked with a white square. Using a larger-scale image, thepositions of Ob-vac, OHb, and other impurities outside the area imaged inA and B are also indicated. A and B have been smoothed using ImageSXM (12) v.1.75. (C) A correlation map between Ob-vac and bright featuresin B. The center of the map represents the position of an Ob-vac shown as agreen square. The black rectangles represent unit cells centered on Ti5c atomsthat surround Ob-vac. The results are averaged between the four quadrantswith the numbers shown only in one quadrant. The results are expressed aspercentages that add to 100% when the numbers in all four quadrants aresummed. The darker the shading, the greater the probability of finding abright feature at the separation indicated by the map. (D) STS spectra repre-sented as LDOS plots by plotting ðdI∕dVÞ × ðV∕IÞ vs V (26). The LDOS plots aretaken from the bright features associated with Ob-vac (black), Ob-vac (green),OHb (purple), Ti5c (yellow), and Ob (red). Each curve is averaged from 180individual spectra taken from the CITS set shown in this work and another180 individual spectra taken from an equivalent CITS set recorded in an iden-tical area of the surface. The black squares in A and B indicate which brightfeatures contribute to the curves for bright features associated with Ob-vac;none were counted when they were also diagonally adjacent to OHb or inclose vicinity to impurities.

2394 ∣ www.pnas.org/cgi/doi/10.1073/pnas.0911349107 Papageorgiou et al.

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0

Page 5: Electron traps and their effect on the surface chemistry of TiO … · Electron traps and their effect on the surface chemistry of TiO2ð110Þ Anthoula C. Papageorgioua, Nikolaos

whereas formation of O2Hð2−Þ from OHbð2−Þ is spontaneous(OHbð2−Þ þO2ðgÞ → O2Hð2−Þ þOb, ΔG ¼ −4.75 eV), the for-mation of H2O2ð2−Þ (ΔGf ¼ −1.02 eV) is strongly unfavouredwith respect to Oadð2−Þ (ΔGf ¼ −16.06 eV). Thus the reactionbetween O2 and OHb must take place via formation ofO2Hð2−Þ that, as with other OxHyð2−Þ intermediates, eventuallyreacts to form Oadð2−Þ. O2 acts to scavenge all the initially avail-able electron excess. Consistent with available spectroscopic data(3, 9, 15) where exposure to O2 quenches the BGS, sampleoxidation in the calculation is accompanied by removal of Ti3þBGS and localisation of Oadð2−Þ states in the valence band,VB (Fig. 6). Similar results are obtained by simulating oneTiint in the presence of two Oad, as shown in Fig. S7. No BGSare found associated with the Oad; instead the electronic chargedensity contributes to the lattice VB. A Bader charge analysisreveals a difference of 10% between the modeled atomic chargesof these Oad products (0.9e) and the bridging Ob atoms (1e), thushighlighting the same oxide hybridization (22, 23) for both speciesin the simulation cell.

Hence, by accounting for TiO2ð110Þ surface-trapped electron-polarons, and the ensuing increase in the actual number of Ti3þ

species available at the surface, the experimentally observed for-mation of Oad from the reaction of O2 with OHb can be correctlydescribed by theory. Furthermore, the same electron traps couldbe involved in trapping photogenerated electrons with potentialimplications for the photocatalytic activity of the surface. Weexpect that electron trapping will be important in dictating thereaction pathways on other dielectric metal oxide surfaces. Thiswill have important consequences in areas such as corrosion,catalysis, and microelectronics.

MethodsExperimental. O2 reactions were performed at room temperature with a vari-able temperature Omicron ultra high vacuum (UHV) STM. STS measurementsemployed a liquid helium bath cryostat Omicron UHV STM in order to max-imize stability. CITS were imaged with 100 × 100 pixels, the component I-Vspectra comprising 40 equally spaced points between þ2 and −2 V. Positivebias indicates tunneling into empty sample states, with negative bias corre-sponding to tunneling out of filled sample states. Tungsten tips were alwaysused unless otherwise stated. Standard sample preparation procedures wereemployed that are described elsewhere (1).

Computational. Grand-canonical (μVT) formation energies (29) were calcu-lated within a DFT-PAW approach implemented in the VASP program (30).LSDAþ U simulations were carried out on the basis of 3 × 2 8 trilayers (8L)slabs with a local Ueff ¼ 5.5 eV correction for Ti atoms, as suggested fromembedded cluster configuration interaction results (21). As for the smaller 3 ×2 4 trilayer (4L) slabs, exchange and correlation terms were calculated atPW91-GGA (31) level for geometrical relaxation and on a hybrid HSE06(10, 11) basis for final single point calculations and STM simulations. Ti 3pand 3s semicore states were treated as valence states. A 3 × 4 × 1 (Γ only) spe-cial k-points grid, 15 Å vacuum range, and 400 eV plane-wave energy cutoffwere used for all the 4L (8L) simulations. The Γ-only finite sampling error onBGS dispersion (8L) was found to be converged within 0.1 eV against the finer3 × 4 × 1 grid. The two (six) topmost trilayers and the adsorbates were relaxedto maximum atomic forces of 0.01 eÅ−1. Spin polarization was enforcedaccounting for the number of extra electrons and optimising for the highestspin multiplicity.

As in previous investigations (28), STM and STS simulations were per-formed with a Tersoff-Hamann approach (32) implemented in the bSKANprogram (33). The calculated HSE06 Fermi energy was used as the zerovoltage point.

Bader charge analysis (34) was carried out on the basis of the total chargedensity i.e. accounting for both the electronic and ionic core charge.

ACKNOWLEDGMENTS. The authors thank Georg Kresse, Daniel Sanchez-Portal, Alex Shluger and David Muñoz Ramo for useful discussions and DavidHumphrey for assistance with some of the experiments. Support fromEngineering and Physical Sciences Research Council (EP/C541898/1), EU(STRP Project NANOCHEMSENS), and Royal Society (W.A.H.) is gratefullyacknowledged.

1. Bikondoa O, et al. (2006) Direct visualization of defect-mediated dissociation of wateron TiO2ð110Þ. Nat Mater 5:189–192.

2. Wendt S, et al. (2005) Oxygen vacancies on TiO2ð110Þ and their interaction with H2Oand O2 : A combined high-resolution STM and DFT study. Surf Sci 598:226–245.

3. Henderson MA, Epling WS, Peden CHF, Perkins CL (2003) Insights into photoexcitedelectron scavenging processes on TiO2 obtained from studies of the reaction of O2

with OH groups adsorbed at electronic defects on TiO2ð110Þ. J Phys Chem B107:534–545.

4. Zhang Z, et al. (2009) Water as a catalyst: Imaging reactions of O2 with partially andfully hydroxylated TiO2ð110Þ surfaces. J Phys Chem C 113:1908–1916.

5. Tilocca A, Di Valentin C, Selloni A (2005) O2 interaction and reactivity on a modelhydroxylated rutile(110) surface. J Phys Chem B 109:20963–20967.

6. Du Y, et al. (2009) Imaging consecutive steps of O2 reaction with hydroxylatedTiO2ð110Þ: Identification of HO2 and terminal OH intermediates. J Phys Chem C113:666–671.

7. EplingWS, Peden CHF, Henderson MA, Diebold U (1998) Evidence for oxygen adatomson TiO2ð110Þ resulting from O2 dissociation at vacancy sites. Surf Sci 412-413:333–343.

8. Rasmussen MD, Molina LM, Hammer B (2004) Adsorption, diffusion, and dissociationof molecular oxygen at defected TiO2ð110Þ: A density functional theory study. J ChemPhys 120:988–997.

9. Wendt S, et al. (2008) The role of interstitial sites in the Ti3d defect state in the bandgap of titania. Science 320:1755–1759.

10. Heyd J, Scuseria GE, Ernzerhof M (2003) Hybrid functionals based on a screenedCoulomb potential. J Chem Phys 118:8207–8215.

11. Heyd J, Scuseria GE, Ernzerhof M (2006) Erratum: Hybrid functionals based on ascreened Coulomb potential [J Chem Phys. 118, 8207 (2003)]. J Chem Phys 124:219906.

12. http://www.liv.ac.uk/~sdb/ImageSXM/.13. Muñoz Ramo D, Gavartin JL, Shluger AL, Bersuker G (2007) Spectroscopic properties of

oxygen vacancies in monoclinic HfO2 calculated with periodic and embedded clusterdensity functional theory. Phys Rev B 75:205336.

14. Giordano L, Sushko PV, Pacchioni G, Shluger AL (2007) Electron trapping at pointdefects on hydroxylated silica surfaces. Phys Rev Lett 99:136801.

15. Kurtz RL, Stockbauer R, Madey TE, Román E, De Segovia JL (1989) Synchrotronradiation studies of H2O adsorption on TiO2ð110Þ. Surf Sci 218:178–200.

16. Batzill M, Katsiev K, Gaspar DJ, Diebold U (2002) Variations of the local electronicsurface properties of TiO2ð110Þ induced by intrinsic and extrinsic defects. Phys RevB 66:235401.

17. Di Valentin C, Pacchioni G, Selloni A (2006) Electronic structure of defect states inhydroxylated and reduced rutile TiO2ð110Þ surfaces. Phys Rev Lett 97:166803.

18. Deskins NA, Dupuis M (2007) Electron transport via polaron hopping in bulk TiO2 : Adensity functional theory characterization. Phys Rev B 75:195212.

19. Deskins NA, Rousseau R, Dupuis M (2009) Localized electronic states from surfacehydroxyls and polarons in TiO2ð110Þ. J Phys Chem C 113:14583–14586.

20. Henderson MA (1999) A surface perspective on self-diffusion in rutile TiO2. Surf Sci419:174–187.

21. Calzado CJ, Hernández NC, Sanz JF (2008) Effect of on-site Coulomb repulsion term Uon the band-gap states of the reduced rutile (110) TiO2 surface. Phys Rev B 77:045118.

22. Resta R (2008) Physical chemistry: Charge states in transition. Nature 453:735.

Fig. 6. Plan view of surface species together with their grand-canonicalformation free energies (T ¼ 300 K, PO2

¼ 1.3 × 10−8 mbar, PH2O ¼1 × 10−11 mbar). Ti is shown blue, lattice O red, O from adsorbates orange,and H green. Bridging oxygen (Ob) atoms are shown larger to highlight them.The energies are in eV and the brackets after the energies indicate theelectronic charge of the surface species. Geometries are optimized (3 × 2

supercell) for the neutral state. The total density of states for the optimizedlayers (filled red) of Oadð2−Þ∕TiO2ð110Þ is displayed together with theOadð2−Þ-projected density of states (PDOS, filled orange) in the bottom-rightpanel.

Papageorgiou et al. PNAS ∣ February 9, 2010 ∣ vol. 107 ∣ no. 6 ∣ 2395

CHEM

ISTR

Y

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0

Page 6: Electron traps and their effect on the surface chemistry of TiO … · Electron traps and their effect on the surface chemistry of TiO2ð110Þ Anthoula C. Papageorgioua, Nikolaos

23. Raebiger H, Lany S, Zunger A (2008) Charge self-regulation upon changing theoxidation state of transition metals in insulators. Nature 453:763–766.

24. Iddir H, Öğüt S, Zapol P, Browning ND (2007) Diffusion mechanisms of native pointdefects in rutile TiO2 : Ab initio total-energy calculations. Phys Rev B 75:073203.

25. Krüger P, et al. (2008) Defect states at the TiO2ð110Þ surface probed by resonantphotoelectron diffraction. Phys Rev Lett 100:055501.

26. Gurlu O, Zandvliet HJW, Poelsema B (2004) Electronic properties of (2 × 1) and cð4 × 2Þdomains on Ge(001) studied by scanning tunneling spectroscopy. Phys Rev Lett93:066101.

27. Minato T, et al. (2009) The electronic structure of oxygen atom vacancy and hydroxylimpurity defects on titanium dioxide (110) surface. J Chem Phys 130:124502.

28. Teobaldi G, et al. (2007) Modelling STM images of TiO2ð110Þ from first-principles:Defects, water adsorption and dissociation products. Chem Phys Lett 437:73–78.

29. Van de Walle CG, Neugebauer J (2004) First-principles calculations for defects andimpurities: Applications to III-nitrides. J Appl Phys 95:3851–3879.

30. Kresse G, Furthmüller J (1996) Efficient iterative schemes for ab initio total-energycalculations using a plane-wave basis set. Phys Rev B 54:11169–11186.

31. Perdew JP, et al. (1992) Atoms, molecules, solids, and surfaces: Applications of thegeneralized gradient approximation for exchange and correlation. Phys Rev B46:6671–6687.

32. Tersoff J, Hamann DR (1985) Theory of the scanning tunneling microscope. Phys Rev B31:805–813.

33. Palotás K, Hofer WA (2005) Multiple scattering in a vacuum barrier obtained fromreal-space wavefunctions. J Phys: Condens Matter 17:2705–2713.

34. Henkelman G, Arnaldsson A, Jónsson H (2006) A fast and robust algorithm for Baderdecomposition of charge density. Comp Mater Sci 36:354–360.

2396 ∣ www.pnas.org/cgi/doi/10.1073/pnas.0911349107 Papageorgiou et al.

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0