+ All Categories
Home > Documents > Early-Stage Deactivation of Platinum-Loaded TiO2 Using In ...was suspended in 50 vol% aqueous...

Early-Stage Deactivation of Platinum-Loaded TiO2 Using In ...was suspended in 50 vol% aqueous...

Date post: 24-Feb-2021
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
8
Early-Stage Deactivation of Platinum-Loaded TiO 2 Using In Situ Photodeposition during Photocatalytic Hydrogen Evolution Greta M. Haselmann and Dominik Eder* Technische Universitä t Wien, Institut fü r Materialchemie, Getreidemarkt 9, 1060, Vienna, Austria * S Supporting Information ABSTRACT: This work unravels a sudden deactivation of Pt/TiO 2 (P25) during the initial stages of photocatalytic H 2 evolution from aqueous solution that, until now, has gone unnoticed, using a unique combination of in situ photodeposition of Pt with an on-line gas detector system. Utilizing a set of techniques, including high- resolution transmission electron microscopy (HRTEM), X-ray photoelectron spectroscopy (XPS), UV-visible diuse reectance spectroscopy (DRS-UV-vis), X-ray powder diraction (XRD), Raman spectroscopy, and physisorption, we were able to attribute this deactivation to a shift in mechanism, accompanied by an increase in CO concentration. Key to this phenomenon is the ratio of Pt atoms to oxygen vacancies, which were created through ultrasonic pretreatment and in situ UV irradiation in the bulk and surface, respectively. We also observed a potential additional contribution to the deactivation by encapsulation of the Pt nanoparticles, indicating that strong metalsupport interaction (SMSI) may indeed happen in aqueous and ambient conditions. Furthermore, we encourage implementing the concept of a dynamiccatalyst to photochemistry that opens up a new approach toward understanding the complex mechanisms and kinetics in heterogeneous photocatalysis. KEYWORDS: photocatalysis, TiO 2 , deactivation, mechanism, defects, SMSI, photodeposition, hydrogen INTRODUCTION Heterogeneous photocatalysis covers a range of cutting-edge applications that address important socioeconomic areas such as energy, 1,2 environment, 3,4 hygiene and disinfection, 5,6 and recently also gained considerable impact on the development of novel green processes. 7 The comparability of photocatalytic activities generally constitutes a major challenge, because they are dependent on a variety of parameters, including reaction temperature, light intensity, and amount and type of sacricial agent, as well as specic reactor setups. Very often, this problem is addressed by using a common benchmark system as reference and, less consistently, by comparing quantum eciency values, which are less system-sensitive than actual activity values. This renders it crucial to have a good benchmark system that ensures stable and reliable activity levels. TiO 2 in particular of type P25so far has been the most widely studied material in photocatalysis. P25 is a commercial product that is commonly synthesized by ame-spray pyrolysis and has demonstrated high photocatalytic activities, making it a highly popular reference material. P25 is a mixed-phase compound that consists of anatase, rutile, and amorphous phases. The phase composition is typically characterized by an anatase:rutile ratio of 80:20, while the presence of an amorphous phase is very often neglected in the literature. Ohtani et al. determined the ratio of anatase, rutile, and amorphous phase in a sample by selective dissolution to be 78:14:8, while simultaneously noting the inconsistency in composition between dierent samples. 8 Still, when composited with platinum nanoparticles as co- catalysts, TiO 2 remains one of the most active photocatalysts for both oxidation (e.g., dye degradation, water purication) and reduction (e.g., hydrogen evolution) reactions. Therefore, Pt/TiO 2 has evolved as the most common reference photo- catalyst. Here, we report on a detailed investigation of the initial stages of photocatalytic hydrogen formation using Pt-loaded P25 in which we observed an unexpected sudden deactivation, which diers distinctly from previously reported passivation of photocatalysts. We were able to observe this eect due to the unique combination of in situ photodeposition of Pt with an on-line gas detector system. Employing a CO detector, in addition to H 2 and CO 2 , allowed us to pinpoint this deactivation to a shift in mechanism that is accompanied by an increased formation of CO. Deactivation of Pt/P25 has drastic consequences and may even be the reason for the large deviations within reported literature results, i.e., activities may have been obtained from already deactivated samples and thus underestimate the actual potential of TiO 2 . Received: March 16, 2017 Revised: June 1, 2017 Published: June 6, 2017 Research Article pubs.acs.org/acscatalysis © 2017 American Chemical Society 4668 DOI: 10.1021/acscatal.7b00845 ACS Catal. 2017, 7, 46684675
Transcript
Page 1: Early-Stage Deactivation of Platinum-Loaded TiO2 Using In ...was suspended in 50 vol% aqueous methanol solution (200 mL). Pt was photodeposited in situ from an aqueous H 2 PtCl 6 solution.

Early-Stage Deactivation of Platinum-Loaded TiO2 Using In SituPhotodeposition during Photocatalytic Hydrogen EvolutionGreta M. Haselmann and Dominik Eder*

Technische Universitat Wien, Institut fur Materialchemie, Getreidemarkt 9, 1060, Vienna, Austria

*S Supporting Information

ABSTRACT: This work unravels a sudden deactivation of Pt/TiO2(P25) during the initial stages of photocatalytic H2 evolution fromaqueous solution that, until now, has gone unnoticed, using a uniquecombination of in situ photodeposition of Pt with an on-line gasdetector system. Utilizing a set of techniques, including high-resolution transmission electron microscopy (HRTEM), X-rayphotoelectron spectroscopy (XPS), UV-visible diffuse reflectancespectroscopy (DRS-UV-vis), X-ray powder diffraction (XRD), Ramanspectroscopy, and physisorption, we were able to attribute thisdeactivation to a shift in mechanism, accompanied by an increase inCO concentration. Key to this phenomenon is the ratio of Pt atomsto oxygen vacancies, which were created through ultrasonicpretreatment and in situ UV irradiation in the bulk and surface, respectively. We also observed a potential additionalcontribution to the deactivation by encapsulation of the Pt nanoparticles, indicating that strong metal−support interaction(SMSI) may indeed happen in aqueous and ambient conditions. Furthermore, we encourage implementing the concept of a“dynamic” catalyst to photochemistry that opens up a new approach toward understanding the complex mechanisms and kineticsin heterogeneous photocatalysis.

KEYWORDS: photocatalysis, TiO2, deactivation, mechanism, defects, SMSI, photodeposition, hydrogen

■ INTRODUCTION

Heterogeneous photocatalysis covers a range of cutting-edgeapplications that address important socioeconomic areas suchas energy,1,2 environment,3,4 hygiene and disinfection,5,6 andrecently also gained considerable impact on the development ofnovel green processes.7 The comparability of photocatalyticactivities generally constitutes a major challenge, because theyare dependent on a variety of parameters, including reactiontemperature, light intensity, and amount and type of sacrificialagent, as well as specific reactor setups. Very often, this problemis addressed by using a common benchmark system asreference and, less consistently, by comparing quantumefficiency values, which are less system-sensitive than actualactivity values. This renders it crucial to have a good benchmarksystem that ensures stable and reliable activity levels. TiO2inparticular of type P25so far has been the most widely studiedmaterial in photocatalysis. P25 is a commercial product that iscommonly synthesized by flame-spray pyrolysis and hasdemonstrated high photocatalytic activities, making it a highlypopular reference material. P25 is a mixed-phase compoundthat consists of anatase, rutile, and amorphous phases. Thephase composition is typically characterized by an anatase:rutileratio of 80:20, while the presence of an amorphous phase isvery often neglected in the literature. Ohtani et al. determinedthe ratio of anatase, rutile, and amorphous phase in a sample byselective dissolution to be 78:14:8, while simultaneously notingthe inconsistency in composition between different samples.8

Still, when composited with platinum nanoparticles as co-catalysts, TiO2 remains one of the most active photocatalystsfor both oxidation (e.g., dye degradation, water purification)and reduction (e.g., hydrogen evolution) reactions. Therefore,Pt/TiO2 has evolved as the most common reference photo-catalyst.Here, we report on a detailed investigation of the initial

stages of photocatalytic hydrogen formation using Pt-loadedP25 in which we observed an unexpected sudden deactivation,which differs distinctly from previously reported passivation ofphotocatalysts. We were able to observe this effect due to theunique combination of in situ photodeposition of Pt with anon-line gas detector system. Employing a CO detector, inaddition to H2 and CO2, allowed us to pinpoint thisdeactivation to a shift in mechanism that is accompanied byan increased formation of CO. Deactivation of Pt/P25 hasdrastic consequences and may even be the reason for the largedeviations within reported literature results, i.e., activities mayhave been obtained from already deactivated samples and thusunderestimate the actual potential of TiO2.

Received: March 16, 2017Revised: June 1, 2017Published: June 6, 2017

Research Article

pubs.acs.org/acscatalysis

© 2017 American Chemical Society 4668 DOI: 10.1021/acscatal.7b00845ACS Catal. 2017, 7, 4668−4675

Page 2: Early-Stage Deactivation of Platinum-Loaded TiO2 Using In ...was suspended in 50 vol% aqueous methanol solution (200 mL). Pt was photodeposited in situ from an aqueous H 2 PtCl 6 solution.

■ EXPERIMENTAL SECTION

The photocatalytic experiments were conducted in a quartzflow reactor vessel that was top-irradiated with a 200 Wsuperpressure Hg lamp (λ = 240−500 nm, intensity = 30 mW/cm2) and connected to an on-line gas analyzer (X-STREAMgeneral purpose gas analyzer, Emerson) with argon as carriergas. The photocatalyst (50 mg, Aeroxide P-25, Acros Organics)was suspended in 50 vol % aqueous methanol solution (200mL). Pt was photodeposited in situ from an aqueous H2PtCl6solution. The photocatalysts were recovered and dried in air at60 °C for characterization. Calculations for quantumefficiencies (QE) as well as absolute H2 evolution rates forPt/P25 are shown in the Supporting Information (Table S1).Ultrasonication (US) is commonly used in the literature todisperse the photocatalyst. To investigate the effect of suchharsh pretreatment, all experiments were conducted with andwithout ultrasonication, using a common ultrasound bath(VWR Ultrasonic cleaner, 45 kHz, 80 W) prior to thephotodeposition and the evolution reaction (no exposure ofphotocatalyst to air between US treatment and UV irradiation).Brunauer−Emmett−Teller (BET) specific surface area

measurements were carried out using Micromeritics ASAP2020 and ASAP 2010 systems. Powder XRD patterns wererecorded using a Bruker D8 Advance system with LynxEyeDetector and Cu Kα radiation (λ = 1.5406 Å). Samples werescanned at 2θ angles of 20°−80°. HRTEM images wereobtained on a Zeiss Libra FE 200 system that was operated at200 kV. DRS−UV-vis was measured on a UV-vis photo-spectrometer (Jasco, Model V-670) that was equipped with anUlbricht sphere inside a diffuse reflectance unit. Raman spectrawere measured on a Jobin Yvon Horiba LABRAM HR systemthat was equipped with a Ne:YAG laser (λ = 532 nm) as a

monochromatic light source, a charge-coupled device (CCD)for detection, and an optical microscope (Olympus, ModelBX41) to focus the laser beam. XPS measurements were carriedout on a K-Alpha X-ray photoelectron spectrometer that wasprovided by Thermo VG Scientific. Monochromatic Al Kα X-rays were used as an excitation source (∼75 W, 400 μm spotsize). The pass energy was 20−30 eV.

■ RESULTS AND DISCUSSION

Typically, reports on photocatalysis with Pt/P25 in comparablereaction conditions show a steady increase in the amounts ofevolved H2 and thus stable production rates. However, weobserved an unexpected deactivation of Pt/P25 during the earlystages of reaction (i.e., mostly within the first hour), which isdependent on several parameters, such as pretreatmentconditions (e.g., ultrasound, calcination), UV light intensity,and Pt loading.Figure 1A shows the time-dependent changes of H2

evolution rates for P25 with 0.4 wt % Pt with (red) andwithout (black) US in aqueous methanol solution. The rates inboth cases reach a maximum value already at ∼30 min, whichcoincides with the time required for filling the dead volume ofthe gas circuit with the evolving gas. However, there is a distinctdifference in rates between the samples with and without USafter this point in time. The curve without US can bedistinguished into two regimes: a gradual decrease by ∼44%over a time period of ∼10 h, followed by a sudden drop to arelatively stable value of 4.7 mmol h−1, which is 10 times lowerthan the maximum level (39 mmol h−1 g−1; QE280−500 nm =28%). In contrast, when P25 was ultrasonicated for only 1 minprior to the experiment, the H2 evolution rate decreasedimmediately after reaching maximum at 30 min, leveling off to

Figure 1. (A) Time dependence of the hydrogen evolution rate over P25-TiO2 loaded with 0.4 wt % Pt in 50 vol % methanol solution with (redcurve) and without (black curve) US pretreatment. (B) Effect of UV intensity on the hydrogen evolution rate: decreasing the intensity (orange)under otherwise same conditions prevented deactivation during the course of the experiment. (C) A 3-fold increase in UV intensity (orange)deactivates a formerly stable sample. CO2 values were multiplied by a factor of 10 for plotting to enable the reader to follow the curves better.

ACS Catalysis Research Article

DOI: 10.1021/acscatal.7b00845ACS Catal. 2017, 7, 4668−4675

4669

Page 3: Early-Stage Deactivation of Platinum-Loaded TiO2 Using In ...was suspended in 50 vol% aqueous methanol solution (200 mL). Pt was photodeposited in situ from an aqueous H 2 PtCl 6 solution.

an almost-identical post-deactivation value of 4.2 mmol h−1 g−1.Longer irradiation with US only resulted in slightly lower rates,while the point of time of deactivation remained the same (seeFigure S1 in the Supporting Information). Interestingly, post-deactivation rates compare much better to reported literaturevalues of ∼3−4 mmol h−1 g−1 for similar conditions9−11 thanthe 10-fold-higher pre-deactivation rates.We first assessed various well-known explanations for

deactivation that have been reported in the literature. Structuraland morphological changes of the photocatalyst can be ruledout, as physisorption experiments and XRD studies confirmedthat neither US nor UV irradiation had any significant effect onBET surface area (Table S2 in the Supporting Information),phase composition (i.e., anatase-to-rutile ratio) and averagecrystal size (see Table S2 and Figure S3 in the SupportingInformation). We were also able to exclude leaching of Ptduring the reaction as well as the presence of impuritiesdetectable by XPS measurements. HRTEM images of pristineas well as the pretreated samples (US and US/UV) revealed novisible differences in the morphology of the crystals or thedispersion of the Pt (see Figure S4 in the SupportingInformation). Methanol was used in large excess to avoidexhaustion of the sacrificial agent during an experimental run.Therefore, we assume that the treatment (i.e., ultra-

sonication) of the metal oxide suspension prior to addition ofthe Pt precursor had a crucial effect on the photocatalyst byinvoking metal−metal oxide interactions and inducingdeactivation. Indeed, ultrasound is known to create a veryharsh chemical environment by acoustic cavitation that is evenutilized to activate catalysts (i.e., “sonocatalysis”).12 Rapidimplosion of cavities leads to estimated temperatures andpressures of up to 5500 °C and several 100 atm inside thebubbles, while the surrounding liquid still reaches temperaturesof up to 2100 °C.13−16 In the case of water, the heat of cavityimplosion leads to H• and OH• radical formation that canoxidize or reduce inorganic compounds such as metal oxidesintroducing defects into the crystal lattice. In addition, there canalso be a variety of mechanical effects on solid particles arisingfrom ultrasound treatment, such as microjets and shock waves,which can both damage the solid surface.16,17

There have been many reports about the formation of defectsby either or both of these effects, such as oxygen vacancies inmetal oxides by ultrasonic treatment (e.g., SiO2,

18,19 Nb2O5,20

MoO3,21 ZnO, ZrO2, Fe2O3, SnO2

22). In particular, Osorio-Vargas et al. showed that oxygen vacancies are formed duringlow-frequency US treatment of TiO2.

23 Bellardita et al.thoroughly investigated US-reduced TiO2 and suggested thatoxygen vacancies were formed almost irreversibly in the bulk ofTiO2, which explains why XPS and Raman spectroscopyshowed no Ti3+ signal.24 They further demonstrated that UStreatment increased the catalyst’s activity for glucose con-version. The effect of oxygen vacancies on the photocatalyticactivity was also confirmed in other reports.25−27 Therefore, itis reasonable to consider oxygen vacancies as the origin of theobserved deactivation in our photocatalysts.In order to assess the effect of methanol during US

irradiation, we ultrasonicated P25 in pure water and addedmethanol afterward. Deactivation now occurred earlier and witha considerably lower maximum activity (Figure S2). This wouldfit well with methanol acting as a radical scavenger: thepresence of methanol may prevent reduction during USirradiation to some extent, which results in a better activity,compared to samples that were ultrasonicated in pure water.

In another set of experiments, we varied the intensity of theUV irradiation, which is also known to induce the formation ofdefects such as Ti3+ and oxygen vacancies.28−30 Indeed, whenthe intensity of the UV light was lowered by a factor of 10, wedid not observe any deactivation of the ultrasonicated sampleunder otherwise identical conditions anymore (Figure 1B). Incontrast, increasing the UV light intensity has led to animmediate deactivation, even of samples that were previouslyproducing H2 at relatively stable rates (Figure 1C).These results demonstrate that both US and UV irradiation

can facilitate deactivation of photocatalysts through theformation of oxygen vacancies, possibly located in the bulk orsubsurface region. This is supported by our observation that theH2 evolution rate remained steady at 39 mmol h−1 g−1

(QEλ<400 nm = 56%) for a prolonged duration, when exposedto oxidizing atmosphere (400 °C, 5 h, air) prior to the reaction,as it is known that oxidative treatments can heal oxygenvacancy defects (Figure 2A).31 It is important to note that

changes in the dispersion of Pt between the calcined andnoncalcined samples as well as the ultrasonicated andnonultrasonicated samples could be excluded with the help ofTEM (see Figure S5 in the Supporting Information).Raman spectroscopy, XPS, and DRS-UV-vis were used to

characterize the defect structure in our materials. Thus, we firstevaluated Pt-free P25 and investigated the effects of US and UVirradiation on defect formation. In particular, for the effect fromUS irradiation, we irradiated samples in pure water and in 50vol % aqueous methanol solution and dried them without anyexposure to the UV light source. For the effect from UV light,we followed the procedure of our photocatalytic experiments,yet without the photodeposition of Pt. Raman spectroscopyand XPS provided no evidence for surface oxygen vacancies or

Figure 2. (A) P25 oxidized in air at 400 °C for 5 h (green) preventsdeactivation during the experimental run. (B) Tauc plots of pristineP25 (black), P25 irradiated with US in water (blue) and methanolsolution (magenta), P25 irradiated with UV (orange) and the calcinedP25 (green) show a band gap shift (ca. 0.15 eV) that mainly occurs inthe anatase phase.

ACS Catalysis Research Article

DOI: 10.1021/acscatal.7b00845ACS Catal. 2017, 7, 4668−4675

4670

Page 4: Early-Stage Deactivation of Platinum-Loaded TiO2 Using In ...was suspended in 50 vol% aqueous methanol solution (200 mL). Pt was photodeposited in situ from an aqueous H 2 PtCl 6 solution.

hydroxylation in the irradiated samples, in accordance withBellardita et al.24 We attribute this to the ex situ nature andsurface sensitivity of these measurements and the possibility ofhealing of surface defects upon exposure to ambientatmosphere. However, DRS-UV-vis spectra show a significantincrease in absorbance at ∼320 nm for both treatments, as wellas a considerable redshift of the absorption edge by 15 and 25nm for UV and US, respectively (see Figure S6 in theSupporting Information). Such a behavior is typical of oxygenvacancy formation and supports the previous works on US-treated TiO2 that suggested formation of bulk oxygen vacanciesbeing stable after exposure to an aqueous phase.23,24

Interestingly, the Tauc plot shows that the redshift originatesfrom the anatase component and amounts to ∼0.15−0.20 eV,which reduces the effective band gap to match rutile (3.03 eV)(Figure 2B).32 This reduction in energy complies with thedifference between the energy levels of the conduction band inanatase and freely mobile electrons formed by oxygenvacancies.31 In addition, the absorption spectra are not affectedby the presence of methanol. The correlation between defectsand deactivation is further supported by the absence of anabsorption shift in the nondeactivating (heat-treated) sample.Next, we investigated the parameter of Pt size and

concentration. We conducted a series of experiments atdifferent Pt loadings ranging from 0.25 wt % to 1.5 wt %.Deactivation was strongly dependent on the amount of Pt. Forexample, at 0.25 wt %, the photocatalyst always deactivatedindependently on pretreatment with US, while no deactivationwas observed at loadings of ≥0.75 wt %, even after USpretreatment (Figure 3). At intermediate loadings, thedeactivation was dependent most strongly on the USpretreatment, as described above. It is important to note thatthe particle size distributions of all loadings were similar, exceptfor loadings of ≥1.0 wt %, which increased in size from 4 nm to4.5 nm (Figure S7 and Table S3 in the SupportingInformation). This is in accordance with the literature33 andsuggests a homogeneous nucleation of Pt in solution beforeadsorption to the TiO2 surface. Therefore, we anticipate thatdeactivation in our case is not noticeably affected by particlesize. The difference between the low (deactivating) and high(nondeactivating) Pt loading therefore lies in the total numberof Pt particles deposited on TiO2. Pt clusters preferably depositat surface oxygen vacancies;34 consequently, the total numberof free surface vacancies decreases with increasing Pt loading.Therefore, we argue that, instead of particle size, the relative

ratio of all oxygen vacancies to the number of Pt particles iscritical. While enough oxygen vacancies are available to inducedeactivation at low Pt concentrations, increasing the Pt loadingwill eventually lead to the point, in which additional UV

irradiation is required to create a critical amount of defects.This is consistent with our results (i.e., dependent onirradiation time/reaching of a critical point (Figure 1A) andlight intensity (Figures 1B and 1C)).In the case of Pt-loaded P25, DRS-UV-vis proved unsuitable

for defect characterization due to overlapping absorption of Pt(see Figures S8−S10 in the Supporting Information); hence,we used Raman spectroscopy to evidence the presence ofsurface defects in our samples. Figure 4 shows the typical

modes of P25: Eg (146 cm−1, 198 cm−1, 639 cm−1), B1g (398cm−1), and an A1g mode (515 cm−1 superimposed by 519cm−1).35,36 In contrast to the Pt-free samples, we observed adistinct peak broadening and shift of the Eg mode at 146 cm−1

with increasing Pt loading, which is characteristic of defects andnow became measurable due to the stabilization with Ptparticles.25,27,35,37

XPS can provide valuable information on the oxidation stateof a species. Figure 5 shows the Pt 4f spectra of the samplesloaded with 0.25, 0.4, and 1.5 wt % Pt as well as pristine P25 for

Figure 3. Gas evolution rates during photocatalytic hydrogen evolution for a series of P25 loaded with different amounts of Pt. Deactivation isobserved at low Pt loadings independent of the pretreatment with US (e.g., 0.25 wt %) but not at high Pt loadings (≥0.75 wt %).

Figure 4. Raman spectra of recovered Pt/P25 after sacrificial watersplitting, compared to pristine P25: (left) complete spectra with typicalP25 Raman modes; (right) closeup of Eg mode at 146 cm−1, showingshift and peak broadening, which indicates the presence of defects inthe Pt-loaded samples.

ACS Catalysis Research Article

DOI: 10.1021/acscatal.7b00845ACS Catal. 2017, 7, 4668−4675

4671

Page 5: Early-Stage Deactivation of Platinum-Loaded TiO2 Using In ...was suspended in 50 vol% aqueous methanol solution (200 mL). Pt was photodeposited in situ from an aqueous H 2 PtCl 6 solution.

reference: Pristine P25 shows a small signal at 76.0 eV that islikely due to a 3s shakeup of Ti, as we did not see any otherimpurities in our spectra. Compared to P25, samples loadedwith 0.25 and 0.4 wt % Pt show a slight increase in signalintensities that lie in the range of metallic Pt0 (Pt 4f7/2, 70.8 eV;Pt 4f5/2, 74.1 eV).38,39 Especially, the Pt 4f7/2 signal helps indistinguishing the Pt signals from the reference. Interestingly,1.5 wt % Pt-loaded samples contain a pronounced shoulder atthe higher binding energy side, which can only partially beattributed to the 3s shake-up signal. Therefore, we ascribedthese shoulders to a higher oxidation state of Pt.38,40,41 Forcomparison, we poisoned Pt/P25 with oxygen by bubblingoxygen gas through the reaction solution during photo-deposition of Pt. The observed Pt 4f signals correlate well toliterature values for Pt2+−O (Pt 4f7/2, 72.8; Pt 4f5/2, 76.1 eV)and are consistent with the shoulder of the 1.5 wt % loadedsample. As we can exclude O2 formation in our setup,adsorption of other oxygen-containing species onto Pt is thelikely explanation for the observed shift. These could eitherstem from the metal oxide (e.g., in form of suboxide species) orreactants and byproducts generated during the photocatalyticredox processes.Until this point, we found that

(1) Pt-loaded P25 deactivates with an increased UV intensityand US treatment;

(2) US and UV irradiation create oxygen vacancies in bulkand surface regions, both affecting the catalytic reaction;and

(3) Vacancies are saturated by Pt (e.g., at higher Pt loadingsand otherwise unchanged conditions), which is beneficialto preventing deactivation.

In view of these observations, we suggest that there is a criticalvacancy:Pt ratio and a dynamic interplay between bulk andsurface vacancies.We thus need to consider interactions between the metal

oxide and the metal co-catalyst, which, in contrast toheterogeneous catalysis, are typically overlooked in the fieldof heterogeneous photocatalysis. Strong metal−support inter-actions (SMSI), in particular, have been frequently discussed asa major source of deactivation in heterogeneous catalysis fornoble metals supported on metal oxides.34,42−45 It typicallyoccurs when the catalyst is exposed to vacuum or reducingatmospheres at high temperatures (>700 K), upon which thesurface of metal oxide is reduced and becomes highly mobileand thus capable of decorating the surface of metal nano-particles. The resulting decrease in accessible active surface areais responsible for a partial or complete loss of catalytic activity(“decoration model”). An additional contribution may stemfrom interfacial charge transfer between the metal and thereduced metal oxide (“electronic model”), which has beenobserved after reduction at temperatures as low as 473K.43,46−49 This phenomenon has not been reported in ambientconditions so far, and very few reports have discussed SMSI inthe broader context of photocatalysis (e.g., refs 50−53). Indeed,we have observed Pt nanoparticles that did contain a thin shellof presumably TiO2, thus suggesting metal encapsulation (seeFigures S11−14 in the Supporting Information). However,these results remain too inconclusive to ascribe the observeddeactivation to an encapsulation-based SMSI effect.A closer look at the product distributionin particular, CO2

and COduring our experiments revealed further insight intothe source of deactivation. CO2 curves were unobtrusive andfollowed the shape of H2 evolution in all cases, showing theirdirect relation in origin (Figures 1, 2A, 3). However, thedetected CO2 amounts were ∼6 times less than expected fromstoichiometry, which might be due to an incompletedegradation of methanol or dissolution of CO2 into thereaction solution.Furthermore, we employed an additional detector for an

often neglected byproduct of photocatalysis: CO. There havebeen differing reports on the presence or absence of CO asbyproduct during photocatalysis, when methanol is used assacrificial agent. While there are many reports that explicitlyreport that CO was not detected,11,54−58 other groups observedCO generation,10,59−62 albeit without connection to deactiva-tion phenomena. For example, CO was detected during steamreforming of methanol over Pt-loaded TiO2 by Naldoni et al.and Chiarello et al.59,60 Furthermore, Zou et al. detected COfor P25 loaded with 0.1 wt % Pt suspended in 12.5 vol %methanol solution.10 CO generation has also been investigatedby Schubert et al. for the vapor-phase decomposition of methylformate, which is a significant byproduct during methanoldegradation, and formic acid.61 Methyl formate decompositionwas found to result in a significant generation of CO, especiallycompared to formic acid degradation. This CO generation wassuggested to block active sites on the noble-metal co-catalystand thus result in significantly slower reaction kinetics incomparison with the degradation of formic acid. The authorsalso noted that addition of water led to an increased generationof H2 and CO2, according to the water−gas shift reaction. Acorrelation between CO and defects was suggested byKobayashi and Yamaguchi. In an ab initio study of CO onTiO2 surfaces, they showed an enhanced interaction betweenoxygen vacancies and CO on a TiO2 surface due to electron

Figure 5. XPS Pt 4f spectra of pristine P25 and Pt/P25 photocatalystsrecovered after photocatalytic experiments as well as Pt/P25 poisonedwith O2. The spectra show weak signals for the low loadings in therange of metallic Pt (Pt 4f7/2,70.8 eV; Pt 4f5/2, 74.1 eV). A pronouncedshoulder in the 1.5 wt % loaded samples at ∼72.8 eV indicates a partialoxidation of Pt.

ACS Catalysis Research Article

DOI: 10.1021/acscatal.7b00845ACS Catal. 2017, 7, 4668−4675

4672

Page 6: Early-Stage Deactivation of Platinum-Loaded TiO2 Using In ...was suspended in 50 vol% aqueous methanol solution (200 mL). Pt was photodeposited in situ from an aqueous H 2 PtCl 6 solution.

back-donation from surface Ti3+ into the π*-orbital ofmolecular CO.63

Here, we found that CO generation is in fact inverselycorrelated with H2 generation for deactivating samples. Theamounts of CO that were detected were small but showed aclear and reproducible trend. Nondeactivating samples did notshow any CO generation at all giving a clear correlationbetween the presence of CO and deactivation of Pt-loadedTiO2 (Figure 6A). We also found that deactivation of our

photocatalyst could be prevented, when the water concen-tration was increased, which is also in accordance with theobservation of Schubert et al. (Figure 6B). Furthermore, anadsorption of CO on Pt is a reasonable explanation for theabove-mentioned shift to a higher oxidation state of Ptobserved in the XPS Pt 4f spectra (see Figure 5).64

We suggest a shift in mechanism for hydrogen evolutionreaction with in situ photodeposited Pt onto TiO2 that istriggered by a critical ratio of oxygen vacancies to active Pt sitesand is accompanied by an increased generation of CO possiblyvia evolution of methyl formate as intermediate during thedecomposition of methanol.Our finding highlights the complexity of photocatalytic

systems and should raise awareness for the method of Ptdeposition, reaction conditions, and side reactions. Comparingthe activities of our “deactivated” samples with literature valuesfor comparable conditions suggests that P25 might have beenunderestimated in its full potential as reaction conditions werenot adequate.

■ CONCLUSIONSIn conclusion, we investigated a previously unnoticed, suddendeactivation of Pt/P25 during the initial stages of photocatalyticH2 evolution from aqueous solution. We attribute this to a shiftin mechanism that is triggered by a critical ratio of oxygenvacancies to Pt atoms and is accompanied by an increasedgeneration of CO possibly via evolution of methyl formate asan intermediate during the decomposition of methanol. Oxygenvacancies are created both, prior to the reaction (viaultrasonication) and in situ (upon UV irradiation). The preciseinterplay of oxygen vacancies in the bulk with the active Pt sites,possibly via bulk-to-(sub)surface diffusion, is yet to beunderstood.We emphasize that it is indeed crucial to consider this

phenomenon when designing and evaluating photocatalyticexperiments, since deactivation is dependent on a variety ofexperimental parameters. For example, in cases where Pt isdeposited onto P25 prior to photocatalytic experiments, e.g., viaincipient wetness, the sample is commonly exposed to reducingatmospheres or other heat treatments, which may have alreadydeactivated or stabilized the photocatalyst. Comparing theactivities of our “deactivated” samples with literature values forcomparable conditions suggests that P25 might have beenunderestimated in its full potential as reaction conditions werenot adequate. A more detailed investigation on the potentialcontribution of a SMSI effect by Pt encapsulation is currentlybeing conducted.Our work will advance the design of new photocatalysts for

real applications, such as degradation of organic molecules andartificial photosynthesis. Furthermore, we encourage imple-menting the concept of a “dynamic” catalyst to photochemistrythat opens up a new approach toward understanding thecomplex mechanisms and kinetics in heterogeneous photo-catalysis.

■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acscatal.7b00845.

Additional photocatalytic experiments, efficiency calcu-lations, supplemental characterization of recoveredphotocatalysts by XRD, BET, XPS, DRS-UV-vis, TEMfor particle size distributions and Raman data (PDF)

■ AUTHOR INFORMATIONCorresponding Author*Tel.: +43-1-58801, ext 165400. E-mail: [email protected].

ORCIDDominik Eder: 0000-0002-5395-564XAuthor ContributionsThe manuscript was written through contributions of allauthors. All authors have given approval to the final version ofthe manuscript.

FundingThe research leading to these results has received funding fromthe European Union Seventh Framework Programme underGrant Agreement No. 310184, CARINHYPH project.

NotesThe authors declare no competing financial interest.

Figure 6. (A) (Left) Additional CO detection shows that the rate ofH2 generation (black) decreases with an increasing CO generation rate(red); (right) for a stable sample, no CO could be detected. (B) H2generation rates for different methanol concentrations for Pt-loadedP25 (0.25 wt %).

ACS Catalysis Research Article

DOI: 10.1021/acscatal.7b00845ACS Catal. 2017, 7, 4668−4675

4673

Page 7: Early-Stage Deactivation of Platinum-Loaded TiO2 Using In ...was suspended in 50 vol% aqueous methanol solution (200 mL). Pt was photodeposited in situ from an aqueous H 2 PtCl 6 solution.

■ ACKNOWLEDGMENTSWe would like to thank Dr. T. Reuter and Dr. A. Schafer fromnanoAnalytics GmbH (Munster, Germany) for XPS measure-ments and Dr. M. Sauer from AIC at TU Vienna for XPSdiscussions and Dr. P. Gebhardt from IMC at TU Vienna forTEM measurements. We would also like to thank Prof. G.Rupprechter and Dr. K. Fottinger from IMC at TU Vienna forhelpful discussions.

■ REFERENCES(1) Chen, X.; Shen, S.; Guo, L.; Mao, S. S. Chem. Rev. 2010, 110,6503−6570.(2) Cherevan, A. S.; Gebhardt, P.; Shearer, C. J.; Matsukawa, M.;Domen, K.; Eder, D. Energy Environ. Sci. 2014, 7, 791−796.(3) Gaya, U. I.; Abdullah, A. H. J. Photochem. Photobiol., C 2008, 9,1−12.(4) Li, M.; Qiang, Z.; Pulgarin, C.; Kiwi, J. Appl. Catal., B 2016, 187,83−89.(5) Foster, H. A.; Ditta, I. B.; Varghese, S.; Steele, A. Appl. Microbiol.Biotechnol. 2011, 90, 1847−1868.(6) Kubacka, A.; Munoz-Batista, M. J.; Ferrer, M.; Fernandez-Garcia,M. Appl. Catal., B 2013, 140−141, 680−690.(7) Colmenares, J. C.; Xu, Y.-J. Heterogeneous Photocatalysis: FromFundamentals to Green Applications. 1st Edition; He, L. N., Rogers, R.D., Su, D., Tundo, P., Zhang, C., Eds.; Springer: Berlin/Heidelberg,2016.(8) Ohtani, B.; Prieto-Mahaney, O. O.; Li, D.; Abe, R. J. Photochem.Photobiol., A 2010, 216, 179−182.(9) Melian, E. P.; Lopez, C. R.; Mendez, A. O.; Díaz, O. G.; Suarez,M. N.; Dona Rodríguez, J. M.; Navío, J. A.; Fernandez Hevia, D. Int. J.Hydrogen Energy 2013, 38, 11737−11748.(10) Zou, J.-J.; He, H.; Cui, L.; Du, H.-Y. Int. J. Hydrogen Energy2007, 32, 1762−1770.(11) Chen, T.; Feng, Z.; Wu, G.; Shi, J.; Ma, G.; Ying, P.; Li, C. J.Phys. Chem. C 2007, 111, 8005−8014.(12) Suslick, K. S.; Skrabalak, S. E. Sonocatalysis. In Handbook ofHeterogeneous Catalysis, Vol. 4; Ertl, G., Knozinger, H., Schuth, F.,Weitkamp, J., Eds.; Wiley−VCH: Weinheim, Germany, 2008; pp2006−2017.(13) Flint, E. B.; Suslick, K. S. Science 1991, 253, 1397−1399.(14) Suslick, K. S.; Hammerton, D. A.; Cline, R. E. J. Am. Chem. Soc.1986, 108, 5641−5642.(15) Birkin, P. R.; Power, J. F.; Leighton, T. G. Chem. Commun.2001, 2230−2231.(16) Thompson, L. H.; Doraiswamy, L. K. Ind. Eng. Chem. Res. 1999,38, 1215−1249.(17) Suslick, K. S.; Casadonte, D. J.; Doktycz, S. J. Solid State Ionics1989, 32-33, 444−452.(18) Romanyuk, A.; Spassov, V.; Melnik, V. J. Appl. Phys. 2006, 99,034314.(19) Romanyuk, A.; Oelhafen, P.; Kurps, R.; Melnik, V. Appl. Phys.Lett. 2007, 90, 013118.(20) Li, D. G. Ultrason. Sonochem. 2015, 27, 296−306.(21) Jeevanandam, P.; Diamant, Y.; Motiei, M.; Gedanken, A. Phys.Chem. Chem. Phys. 2001, 3, 4107−4112.(22) Fan, C.; Yu, S.; Qian, G.; Wang, Z. RSC Adv. 2017, 7, 18785−18792.(23) Osorio-Vargas, P. A.; Pulgarin, C.; Sienkiewicz, A.; Pizzio, L. R.;Blanco, M. N.; Torres-Palma, R. A.; Petrier, C.; Rengifo-Herrera, J. A.Ultrason. Sonochem. 2012, 19, 383−386.(24) Bellardita, M.; Nazer, H. A. E.; Loddo, V.; Parrino, F.; Venezia,A. M.; Palmisano, L. Catal. Today 2017, 284, 92−99.(25) Chen, X.; Liu, L.; Yu, P. Y.; Mao, S. S. Science 2011, 331, 746−750.(26) Kong, M.; Li, Y.; Chen, X.; Tian, T.; Fang, P.; Zheng, F.; Zhao,X. J. Am. Chem. Soc. 2011, 133, 16414−16417.(27) Leshuk, T.; Parviz, R.; Everett, P.; Krishnakumar, H.; Varin, R.A.; Gu, F. ACS Appl. Mater. Interfaces 2013, 5, 1892−1895.

(28) Shultz, A. N.; Jang, W.; Hetherington, W. M., III; Baer, D. R.;Wang, L.-Q.; Engelhard, M. H. Surf. Sci. 1995, 339, 114−124.(29) Zheng, Z.; Huang, B.; Qin, X.; Zhang, X.; Dai, Y.; Whangbo, M.-H. J. Mater. Chem. 2011, 21, 9079−9087.(30) Takeuchi, M.; Martra, G.; Coluccia, S.; Anpo, M. J. Phys. Chem.C 2007, 111, 9811−9817.(31) Eder, D.; Kramer, R. Phys. Chem. Chem. Phys. 2003, 5, 1314−1319.(32) Scanlon, D. O.; Dunnill, C. W.; Buckeridge, J.; Shevlin, S. A.;Logsdail, A. J.; Woodley, S. M.; Catlow, C. R. A.; Powell, M. J.;Palgrave, R. G.; Parkin, I. P.; Watson, G. W.; Keal, T. W.; Sherwood,P.; Walsh, A.; Sokol, A. A. Nat. Mater. 2013, 12, 798−801.(33) Ohtani, B.; Iwai, K.; Nishimoto, S.-i.; Sato, S. J. Phys. Chem. B1997, 101, 3349−3359.(34) Horsley, J. A. J. Am. Chem. Soc. 1979, 101, 2870−2874.(35) Naldoni, A.; Allieta, M.; Santangelo, S.; Marelli, M.; Fabbri, F.;Cappelli, S.; Bianchi, C. L.; Psaro, R.; Dal Santo, V. J. Am. Chem. Soc.2012, 134, 7600−7603.(36) Jiang, X.; Zhang, Y.; Jiang, J.; Rong, Y.; Wang, Y.; Wu, Y.; Pan,C. J. Phys. Chem. C 2012, 116, 22619−22624.(37) Zheng, Z.; Huang, B.; Lu, J.; Wang, Z.; Qin, X.; Zhang, X.; Dai,Y.; Whangbo, M.-H. Chem. Commun. 2012, 48, 5733−5735.(38) Parkinson, C. R.; Walker, M.; McConville, C. F. Surf. Sci. 2003,545, 19−33.(39) Koudelka, M.; Monnier, A.; Sanchez, J.; Augustynski, J. J. Mol.Catal. 1984, 25, 295−305.(40) Ohtani, B. Chem. Lett. 2008, 37, 216−229.(41) Bashir, S.; Wahab, A. K.; Idriss, H. Catal. Today 2015, 240,242−247.(42) Tauster, S. J.; Fung, S. C.; Garten, R. L. J. Am. Chem. Soc. 1978,100, 170−175.(43) Tauster, S. J.; Fung, S. C.; Baker, R. T. K.; Horsley, J. A. Science1981, 211, 1121−1125.(44) Bond, G. C.; Burch, R.; Birkett, M. D.; Kuhn, A. T. In StrongMetal-Support Interactions; The Royal Society of Chemistry: London,1983; Vol. 6, pp 27−60.(45) Schauermann, S.; Nilius, N.; Shaikhutdinov, S.; Freund, H.-J.Acc. Chem. Res. 2013, 46, 1673−1681.(46) Pesty, F.; Steinruck, H.-P.; Madey, T. E. Surf. Sci. 1995, 339,83−95.(47) Hayek, K.; Kramer, R.; Paal, Z. Appl. Catal., A 1997, 162, 1−15.(48) Dulub, O.; Hebenstreit, W.; Diebold, U. Phys. Rev. Lett. 2000,84, 3646.(49) Jochum, W.; Eder, D.; Kaltenhauser, G.; Kramer, R. Top. Catal.2007, 46, 49−55.(50) Colmenares, J. C.; Magdziarz, A.; Aramendia, M. A.; Marinas,A.; Marinas, J. M.; Urbano, F. J.; Navio, J. A. Catal. Commun. 2011, 16,1−6.(51) Fujiwara, K.; Deligiannakis, Y.; Skoutelis, C. G.; Pratsinis, S. E.Appl. Catal., B 2014, 154−155, 9−15.(52) Wang, J.; Zhang, M.; Wang, K.; Zhang, J.; Wu, Z.; Jin, Z. Appl.Surf. Sci. 2008, 254, 5375−5379.(53) Naldoni, A.; Riboni, F.; Marelli, M.; Bossola, F.; Ulisse, G.; DiCarlo, A.; Pis, I.; Nappini, S.; Malvestuto, M.; Dozzi, M. V.; Psaro, R.;Selli, E.; Dal Santo, V. Catal. Sci. Technol. 2016, 6, 3220−3229.(54) Kandiel, T. A.; Ivanova, I.; Bahnemann, D. W. Energy Environ.Sci. 2014, 7, 1420−1425.(55) Navarro, R. M.; Arenales, J.; Vaquero, F.; Gonzalez, I. D.; Fierro,J. L. G. Catal. Today 2013, 210, 33−38.(56) Moon, S.-C.; Mametsuka, H.; Tabata, S.; Suzuki, E. Catal. Today2000, 58, 125−132.(57) Bamwenda, G. R.; Tsubota, S.; Nakamura, T.; Haruta, M. J.Photochem. Photobiol., A 1995, 89, 177−189.(58) Galin ska, A.; Walendziewski, J. Energy Fuels 2005, 19, 1143−1147.(59) Chiarello, G. L.; Aguirre, M. H.; Selli, E. J. Catal. 2010, 273,182−190.

ACS Catalysis Research Article

DOI: 10.1021/acscatal.7b00845ACS Catal. 2017, 7, 4668−4675

4674

Page 8: Early-Stage Deactivation of Platinum-Loaded TiO2 Using In ...was suspended in 50 vol% aqueous methanol solution (200 mL). Pt was photodeposited in situ from an aqueous H 2 PtCl 6 solution.

(60) Naldoni, A.; D’Arienzo, M.; Altomare, M.; Marelli, M.; Scotti,R.; Morazzoni, F.; Selli, E.; Dal Santo, V. Appl. Catal., B 2013, 130−131, 239−248.(61) Schubert, G. b.; Bansagi, T. s.; Solymosi, F. J. Phys. Chem. C2013, 117, 22797−22804.(62) Yang, X.; Salzmann, C.; Shi, H.; Wang, H.; Green, M. L. H.;Xiao, T. J. Phys. Chem. A 2008, 112, 10784−10789.(63) Kobayashi, H.; Yamaguchi, M. Surf. Sci. 1989, 214, 466−476.(64) Toyoshima, R.; Yoshida, M.; Monya, Y.; Suzuki, K.; Amemiya,K.; Mase, K.; Mun, B. S.; Kondoh, H. Phys. Chem. Chem. Phys. 2014,16, 23564−23567.

ACS Catalysis Research Article

DOI: 10.1021/acscatal.7b00845ACS Catal. 2017, 7, 4668−4675

4675


Recommended