Sie sind auf Seite 1von 13

View Online

Faraday Discuss. Chem. SOC.,1984, 78,151-163

Detection of the Intermediates of Colloidal Ti0,-catalysed Photoreactions


Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

BY DETLEF BAHNEMANN," ARNIM HENGLEIN AND LUBOMIR SPANHEL


Hahn-Meitner-Institut fur Kernforschung Berlin, Bereich Strahlenchemie, D-1000 Berlin 39, Federal Republic of Germany

Received 3rd May, 1984


Electrons and positive holes have been produced in colloidal TiOz particles upon illumination with near-u.v. light. In the absence of adsorbed reactants, the charge carriers recombine. In the presence of an adsorbed reactant for holes [such as SCN- or poly(viny1 alcohol)], an excess of long-lived electrons remains on the particles and finally reacts with a reactant in the bulk solution. In the presence of an adsorbed reactant for electrons (such as a platinum deposit or methyl viologen), an excess of long-lived holes remains and finally reacts with an oxidizable compound. The processes observed include the reduction of water to form HZ, the reduction of tetranitromethane and halothane and the oxidation of Br- and various organic compounds. Methyl viologen, MV2+, was found to be reduced to the radical M V + and also to be oxidized, to a lesser extent, to form a fluorescing compound. The absorption spectra of the long-lived excess of electrons and positive holes trapped at the interface were measured using flash photolysis.

The catalysis of redox reactions by colloidal metals and metal oxides has been known since the early days of colloid chemistry, although these effects were not studied systematically but in an incidental manner. In the early days of colloid chemistry, some sixty years ago, investigation of the chemistry of free radicals in solution had just started, i.e. radicals were recognized as being the intermediates in many redox processes. Furthermore, the theory of the solid state (and inorganic colloids are solids) began to advance. Today, these two fields are well understood and are the appropriate background for an investigation of the production and reactions of free radicals at the colloid/solution interface. These reactions can be studied in two ways. Radicals may be formed in the bulk solution using radiation or photochemistry and their reactions with a colloid can then be observed. For example, it has been found that many organic radicals rapidly transfer an electron to colloidal metal particles.' The electrons can be stored in the colloid and finally be transferred pairwise to the solvent or a solute. In these reactions the colloid acts as a catalyst for new radical reactions. Similarly, radicals may transfer an electron to a semiconductor colloid. Colloidal TiOz has been found to produce a blue colour upon the uptake of electrons2 and colloidal cadmium sulphide produces highly reactive cadmium metal.3 On the other hand, electrons are promoted to the conduction band, thus leaving positive holes in the valence band, when light is absorbed by the semiconductor colloid. This report deals solely with the latter kind of studies. It is known from solid-state physics that the diffusion length of these charge carriers is much longer than the diameter, ca. 25 nm, of our colloidal particles. This means that the charge carriers will rapidly reach the surface. They may be trapped there and finally recombine, or they may, and this is the topic of our investigation, react with the solvent or dissolved substances. Free radicals are often the intermediates in these interfacial reactions.
151

View Online

152

INTERMEDIATES OF

TiO2-CATALYSED PHOTOREACTIONS

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

Even larger suspended particles of a semiconductor can initiate chemical reactions upon light adsorption. For example, TiO, powder catalyses the photodecomposition of organic acids into CO, and alkyl radicals (Kolbe decarb~xylation),~ the oxidation of halide anions,' the reduction of precious-metal ions6 and the formation of amino acids in solutions containing methane and ammonia.' The literature on the catalytic effectsof TiO, particles is growing rapidly, with many of these reactions being studied from the point of view of using solar energy for the initiation of chemical reactions. However, attention should also be focused on fundamental studies in which the mechanism of these catalytic reactions is explored. It is advantageous to use TiO, as a colloid of small particle size ( s 2 5 nm), as the solutions of such particles are transparent and only attenuate light to a negligible extent by scattering, thus allowing the application of powerful optical methods which have proved to be so helpful in kinetic studies of homogeneous systems. The electrontransfer reactions at the interface are typical electrochemical processes, the small particles often being referred to as microelectrodes, although they are not part of an electrical circuit in the usual sense.
EX PER1 MENTAL
Colloidal Ti02 was prepared by saponification of titanium tetraisopropoxide in acidic aqueous solution. The solvent was evaporated under vacuum until a white powder remained. The final solution was obtained by dissolving 500mg of this powder in 1 dm3 of water. Colloidal platinum was prepared by reducing a 3 x mol dmP3 H2PtC16 solution with 1.7 X IK3mol dm-3 sodium citrate (for 1 h at 100 "C) and removing the excess of ions with Amberlite MBI ion-exchange resin. In order to produce a TiOz colloid with the particles carrying a small P deposit, the colloidal Pt solution was added to the Ti02 solution at the t required ratio. The solvent was evaporated under vacuum and a brown powder was obtained. This powder was then redissolved to obtain the final solution of Pt-covered Ti02. Details of these preparations have recently been described.8 The solutions were degassed before illumination using a water pump and then an oil diffusion pump. Hydrogen was determined using a Toepler pump, a McLeod manometer and a gas chromatograph. The changes in optical absorption and conductivity of the solutions were measured in flash-photolysis experiments with a frequency-doubled JK ruby laser ( A = 347.1 nm, 15 ns pulse width). The laser was operated at ca. 5% of its maximum power and 10-7-10-6 moles of photons were absorbed per Bash and dm3 of solution. The data from 16-64 flashes were digitized and transferred to PDP 1 1/40 computer. The digitized signals were analysed pn-line using a Tektronix 4010 interactive graphic display. The details have already been published, as have the details of the dosimetry method used for both the optical-absorption and conductivity measurements.* The continuous illuminations were carried out with a 450 W xenon lamp (distance 30 cm, 10 cm water filter, various colour filters). In the quantum-yield measurements a monochromator was used to select a f15 nm wavelength range around 360 nm. Actinometry was carried out with ferric oxalate.

RESULTS AND DISCUSSION


PROPERTIES O F COLLOIDAL

Ti0,

Fig. 1 shows the absorption spectrum of the TiOz solution. The absorption increases steeply at wavelengths <380 nm, which corresponds to the band-gap energy of 3.2 eV in macrocrystalline TiOz. On the standard electrochemical energy scale the upper edge of the valence band in macrocrystalline TiO, is located at 3.1 V and the lower edge of the conduction band at -0.1 V. Positive holes in TiOz are very

View Online

D . B A H N E M A N N , A. HENGLEIN A N D L. SPANHEL

153

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

300

400
A/nm

500

x )
mol dm-3 colloidal TiOz at

Fig. 1. Absorption spectrum of a solution containing 6.3 x pH 3.

strong oxidizing agents while the electrons have only moderate reducing power. Note also that colloidal T i 0 2 has an isoelectric point of ca. 3.5, i.e. at low pH the particles are positively charged (by chemisorbed H+) and at higher pH they are negatively charged (by chemisorbed OH-).9 As a consequence, the electronic levels in Ti02- are shifted towards more negative potentials with increasing pH (0.059 mV per pH unit). l o
FORMATION OF

H2

P CONCENTRATION A N D pH t This reaction has already been studied by Gratzel and coworkers" and we will here describe a few additional observations which seem to be important for an understanding of the mechanism. H2 is formed in substantial yields only in the presence of platinum and a scavenger for positive holes. The platinum is believed to pick u p the excess of electrons which remains when some of the positive holes are removed from the illuminated T i 0 2 particles. A useful scavenger is alcohol:
ROLE OF

TiO, h + + R,CHOH e-+Pt

__+

hv

Ti02(e- + h')
H + + R,COH

(1) (2)

H+
__*

Pt-H

__+

Pt+iH2.

(3)

The electrons rapidly form adsorbed hydrogen atoms on the platinum and finally combine to yield H2. In fig. 2 the quantum yield of H2 is plotted as a function of

View Online

154

INTERMEDIATES OF

TiOz-CATALYSED PHOTOREACTIONS

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

[ P ) rnol dm-3 t/

10-5

lo4

Fig. 2. Quantum yield of H2 formation in the illumination of a 6.3 x mol dm-3 colloidal TiOz solution containing various amounts of platinum deposit at pH 3. 0 , 1 moldm-3 I 5 x lov3mol dm-3 PVA. methanol; X, 1 mol d W 3 ethanol; 0,moI dm-3 propan-2-01;

+,

the concentration of platinum for various hole scavengers. The greatest yield achieved was 4.5%, which corresponds to 9% efficiency for the scavenging of charge carriers. At too high a platinum concentration on the Ti02 particles smaller yields were observed. The maximum hydrogen yield was obtained with Pt concentrations of ca. 3 x mol dm-3. Taking into account the size of the Pt and TiO, particles (3.2 and 25 nm, respectively) one calculates that each TiOz particle carries just one P particle. In the presence of too much platinum the recombination of charge t carriers apparently occurs more frequently. Fig. 3 shows how H2production depends on the p H of the solution. It seems surprising at first sight that the yield goes up with decreasing H+ concentration even though hydrogen ions are involved in the evolution of H2 [reaction (3)]. We have to conclude that the recombination of charge carriers is promoted by higher H' concentrations. With larger pieces of semiconductor there is a depletion layer at the surface, across which there exists a potential gradient which allows separation of the charge carriers. As the dimensions of the colloidal particles are much smaller than the thickness of such a depletion layer a potential gradient of this kind does not exist. However, the charges at the surface which are produced by the chemisorption processes Ti02H:+ (plus counteranions) for pH <3.5, Ti02+ nH" for pH >3.5, TiO,+nOH- G Ti02(OH):- (plus countercations) (4) (5)

may strongly influence the trapping of the charge carriers. Electrons may reside Ionger or be trapped more efficiently at a surface containing a lot of chemisorbed

View Online

D. B A H N E M A N N , A. HENGLEIN A N D L. SPANHEL

155

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

PH

Fig. 3. Rate of H2 formation as a function of pH. The solution contained 6.3 x lo-' mol dmW3 colloidal TiO,, 3 x lo-' mol dm-3 P and 1 mol dm-' ethanol. t

protons before they reach the platinum deposit, thus leading to enhanced recombination with the holes, and platinum may scavenge not only electrons but also holes.', With increasing concentration of chemisorbed protons the holes may not reach the surface to be trapped there, and as a consequence a larger fraction may react with the platinum deposit. The yield of H2 in the experiments shown in fig. 2 is also determined by back reactions of the products and the concentration of the scavenger. In the case of alcohols, which are not strongly adsorbed at the Ti02 particles, the yield of H2 is proportional to the alcohol concentration, and of the aliphatic alcohols studied methanol is by far the most hydrophilic compound and should consequently be rather well adsorbed on the hydrophilic Ti02 surface. Thus, it is not surprising that the highest H2 yields are observed with the methanol system. In the case of the thiocyanate ion, however, which is strongly adsorbed and a good scavenger of holes, as described below, no H2 formation could be detected. This is explained by efficient reoxidation of intermediate H atoms by semioxidized thiocyanate, (SCN) 1, which is also adsorbed on the positively charged particle.
REDUCTION OF TETRANITROMETHANE LONG- LIVED ELECTRONS

An efficient reduction of dissolved compounds not adsorbed at the colloidal TiO, particles occurs only in the presence of an adsorbed scavenger for positive holes. Tetranitromethane was used as the reactant for electrons:
(4) e-+C(N02), --* C(NO,),+NO,. This reaction can be followed by recording the light intensity at 350nm, where the stable product C(N02), absorbs strongly. In fig. 4 the concentration of this anion is shown as a function of the illumination time for solutions containing various sodium salts and polymers [PVA is poly(viny1 alcohol) and PAA is poly(acry1amide)]. Note that thiocyanate and citrate are the most efficient promoters of the reduction of tetranitromethane, and of the polymers, PVA is more active than PAA.

View Online

156

INTERMEDIATES OF

TiO2-CATALYSED

PHOTOREACTIONS

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

Fig. 4. Concentration of the trinitromethane anion as a function of time for the illumination mol dm-3 TiOz solution containing 5 x lop4mol dm-3 tetranitromethane and of a 6.3 x various additives (concentration of the added salts 0.1 rnol dm-3, concentration of the added polymers 5 x mol dm-3) at pH 2.

r: 0.15 ;
h

6 z
Q

01,
$1

I'

03 0

0.2

0.4

[C( NO,),]/ 10- rnol dm-3

'

0.6

0.8

1.0

Fig. 5. Quantum yield of the trinitromethane anion as a function of the concentration of tetranitromethane with 0.1 mol dmV3SCN'- present as hole scavenger at pH 2.

The efficiency of the hole scavenger is determined by both its ability to be adsorbed and its ability to be oxidized. For short illumination times, the concentration of C(N0,); formed is proportional to the time, whereas for longer times the curves bend towards the horizontal axis because of photodecomposition of C( NO2); at the colloidal TiO, particles., Reaction (6) requires the diffusion of tetranitromethane to the colloidal particles, as tetranitromethane is not adsorbed. Under continuous illumination conditions, a stationary concentration of an excess of electrons will be built up on the colloidal

View Online

D. BAHNEMANN. A. HENGLEIN AND L. SPANHEL

157

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

I 400

500
A/nm

600

700

Fig. 6. Absorption spectrum of a laser-flashed solution of 6.3 x 10-3 mol dm-3 Ti02 containing 5x rnol dm-3 PVA and time profile of the absorption signal (inset) at pH 10.

365nm C(NO,),

4
4

500 nm (SCN),

TiO,

2 TiO,(e-+h+)
SCN'
4

h++SCN;& e-+C(NO,),

SCN

(SCN),

C(NO,);+NO,

Fig. 7. Oscillographic traces and elementary reactions occurring in the laser-flash photolysis of a solution containing 6.3 x rnol dmP3Ti02,0.02 mol dm-3 SCN- and 5 x lov4rnol dmP3 C(NO& at pH 3.

particles. An excess electron may either react with tetranitromethane or recombine with the positive hole created by the next photon absorbed by the colloidal particle. At sufficiently high concentrations of tetranitromethane, all the electrons will react with it (or, more precisely, as many electrons will react as there are positive holes scavenged by the adsorbed hole scavenger). The yield will then become independent of the concentration of tetranitromethane. Fig. 5 shows the quantum yield of the formation of C( NO2)3 as a function of the concentration of tetranitromethane. Above ca.4 x mol dm-3 the yield is constant at 0.17 C( NO,), anions per photon absorbed. When the SCN- concentration is varied the C(NO,), yield initially increases linearly and then levels off at concentrations >0.3 mol dm-3 as the Langmuir adsorption limit is gradually reached. At this concentration of SCN- the quantum yield was 0.28. Laser-flash experiments with T i 0 2 sols containing a hole scavenger were also carried out. It has already been reported' that the absorption of excess electrons can be seen immediately after a laser flash. The spectrum of the excess of electrons is shown in fig. 6. It contains a broad band in the visible region, the maximum being at 630nm. The absorption signal is stable for seconds (see inset of fig. 6). An excess of electrons can also be produced on colloidal TiO, particles via electron transfer from organic radicals generated by y-radiation in the bulk solution;2 in these experiments larger TiO, particles (ca.100 nm) were used and the absorption

View Online

158

INTERMEDIATES OF

TiO2-CATALYSED PHOTOREACTIONS

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

maximum for the excess of electrons appeared at 750 nm. In the presence of electron scavengers like tetranitromethane or oxygen the absorption signal was not stable. Fig. 7 shows typical oscillographic curves for the absorptions observed in flashphotolysis experiments on a sol containing tetranitromethane and thiocyanate. Immediately after the flash, the 500nm absorption of oxidized thiocyanate, i.e. of the complexed radical (SCN);, is present. After the flash, the absorption of the (SCN)1 radical decays according to second-order kinetics. The calculated rate constant is independent of the presence of tetranitromethane and agrees with the value of 3.2 x lo9 dm3 mol-' s-l for the reaction 2(SCN)i-+

(SCN), +2(SCN-)

as measured by pulse r a d i o l y ~ i s . ' ~The absorption of C(N0,); is built up after -'~ the flash according to first-order kinetics, with the first-order rate constant increasing linearly with tetranitromethane concentration." These results demonstrate that the two reactions responsible for the optical changes take place independently of each other. The reaction of the holes h' with adsorbed SCN- occurs during the flash; the reaction of e- with C(NOJ4 occurs after the flash.
OXIDATION OF
LONG-LIVED POSITIVE HOLES

Br-

A N D O R G A N I C COMPOUNDS

When a TiO, sol containing an adsorbed scavenger for electrons is i luminated by a laser flash, the absorption of the excess of positive holes can be traced immediately after the flash. For the experiments shown in fig. 8, a TiO,. Pt sol was used in which the metal deposit acted as a scavenger for electrons. The absorption signal decayed after the flash (as shown by the inset of fig. 8). This decay was of multi-exponential order, which indicates the participation of various first-order processes with different lifetimes. The presence of oxygen did not alter the decay of the signal. However, the decay was faster in the presence of oxidizable compounds, as will be described below. The spectrum contains a broad band with its maximum at 430nm which is attributed to trapped holes. Chemically, a trapped hole is an 0'- radical anion. In aqueous solution this species absorbs in the ultraviolet, but in the T i 0 2 particles transitions into the conduction band may occur, the energies of which lie in the visible region. Note that the absorption shown in fig. 8 decreases rather steeply below 380nm, i.e. at the photon energy which corresponds to the band-gap energy of TiO,. Note also that the height of the absorption signal of the trapped holes depends on the age of the colloidal solution when it is kept in the light. For example, 30 min after the preparation of the solution

* A closer look at fig. 7 shows that part of the C(N0,); absorption is already present immediately after the 15 ns laser flash followed by the time-resolved build-up described above. A more detailed study with aqueous solutions containing only C( NO2), and SCN- showed that immediate formation of C( NO,); is also observed in the absence of TiO,. It was found that tetranitromethane and thiocyanate form a complex via
C(NO,),+ SCN[C(NO,),---SCN-] which has an absorption maximum at ca. 340 nm. An intramolecular electron transfer takes place once this complex absorbs a photon, leading to very fast formation of C(N0,); ( t l i 2 < 15 ns) [C(NO,),---SCN-]

+SCN-

hv

C(NO2); +NO, +(SCN),.

Details of this study will be published elsewhere."

View Online

D. BAHNEMANN, A. HENGLEIN AND L. SPANHEL

r---

159

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

400

500
h/nm

600

mol dm-3 TiOz and Fig. 8. Absorption spectrum of a laser-flashed Ti02-Pt colloid (3.8 X 1.6 x lo- mol dm-3 Pt) and decay of the absorption (inset) at pH 2.5.

nm

nm

Fig. 9. Schematic description of the reactions occurring in a solution containing the Ti02-Pt colloid and Br- anions.

the signal had decayed to 70% of its original value and after 4 h to 50%. No further change in the optical absorption of the holes was found when the solution was stored for additional 24 h. This is explained by the rough surface of the colloidal particles when they are freshly dissolved. Upon ageing in sunlight, the particles become polished, i.e. the surface becomes smoother and fewer centres for the trapping of holes are available. The Ti02-Pt powder which was used for the preparation of the solution could be stored in sunlight without loss in activity with respect to the photoproduction of holes after dissolution. Fig. 9 describes in a schematic manner what is expected when a sol containing the Ti02-Pt colloid and Br- ions is flashed. It is assumed that two Br-- anions are adsorbed at a TiOz particle and that three e---h+ pairs are formed by the laser flash. The electrons are scavenged by the Pt deposit to yield adsorbed H atoms and finally H2. Two of the holes react with adsorbed Br- during the laser flash, i.e. before being trapped. The third hole reacts after the flash, waiting for the arrival of a Brion from the bulk solution. Experimentally, the absorption signal of the holes immediately after the flash was found to be smaller in the presence than in the absence of Br- and the decay of the remaining absorption was faster. Fig. 10 shows the absorption of h measured at 475 nm (where Bri- has only a very small absorption) as a function of the KBr concentration. The signal becomes smaller with increasing salt concentration until it reaches a limiting value as the saturation region

View Online

160

INTERMEDIATES O F

Ti02-CATALYSED PHOTOREACTIONS

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

0'
0

2.5
[KBr]/ 1 0-2 rnol dm-3

'0 5.0

Fig. 10. Absorbances at 475 nm (holes) and 365 nm (BrJ immediately after the laser flash as function of the KBr concentration for a Ti02-Pt colloid (3.8 x lop3 mol dmp3 Ti02 and 1.6 X mol dmP3 Pt) at pH 2.5.

of the Langmuir adsorption is reached. The 365 nm absorption of oxidized bromide, which is present in the form Bri-, is also shown. It increases with the KBr concentration in a symmetric manner with respect to the decrease in the 475 nm absorption of the holes. Knowing the absorption coefficient of Bri- ( ~ ~ ~ ~ ~ , , , = 9 . 5 x lo3 dm3 mol-' cm-'),Ih one can readily calculate that of the holes: E~~~~~ 9.2 X lo3dm3 mol-' cm-*. (In this calculation the absorption of h' at 365 nm and of Br;- at 475 nm had to be taken into account.) Electron-deficient surface states are well known from studies of the electrochemistry and photoelectrochemistry of semiconductor electrodes where their existence is deduced from electric-current measurements. Studies on colloids complement these findings by characterizing the trapped holes through their optical absorption. The trapped holes can now be dealt with in flash-photolysis experiments, as are the intermediates of photoreactions in homogeneous solution. Their chemical reaction with added substances can be followed by fast optical absorption measurements, as described below. The lifetime of the holes in acid solution is relatively long; even after milliseconds some of the holes can still be observed. They probably disappear through reaction with OH- ions (see below) which are present at low concentrations in acid solutions. The multi-exponential order of the decay of the holes is taken a9 an indication that they are trapped in states of different energies (and lifetimes with respect to the oxidation of water). Fig. 11 shows the results obtained in experiments where the Ti02-Pt sol contained various organic additives. Fig. 1 l ( a ) shows the 475 nm absorption of the holes immediately after the laser flash as a function of the concentration of the additive. Note that citrate is most efficient in reacting with the holes during the flash, ethanol being the least efficient reactant. Fig. 11( 6 ) shows the time required for a drop of 25% of the residual absorption at 475 nm. Note that ethanol is now the most efficient reactant while citrate has little effect. Reaction with the holes during the laser flash requires the adsorption of the reactant at the colloidal particles. Citrate is more strongly adsorbed than ethanol, but on the other hand it is far less readily oxidized in homogeneous solution than ethanol. The holes which react during the flash are at such a high positive potential (ca. 3.1 V) that they can react with practically a11 the adsorbed organic molecules regardless of their structure. However, where the reaction after the flash is concerned, the holes are already trapped ( i e . they are at a less positive potential) and the ability of the organic compound to be oxidized determines the rate of reaction.
I

View Online

D. BAHNEMANN, A. HENGLEIN A N D L. SPANHEL

161

0.50
h

50
acetic acid
II)

.c

Y 0

4
Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

3.

5
0

0.35

\ r

5 2s

&
0.20
0

r .

0.5

1.0

1.5

n "0

0.5
concentration/mol dme3

1.0

concentration/mo! dm-3

Fig. 11. ( a ) Absorbance of the holes at 475 nm immediately after the laser flash as a function of the concentration of added organic compounds. ( b ) Reaction of the remaining holes: time for 25% decrease as a function of the concentration of the organic compound. p H 3 for ethanol and pH 6 for all other reagents.

Fig. 12. ( a ) Time profile of the 392 nm absorption and ( b ) the conductivity of a 6.3 X lop3mol dm-3 TiOz sol containing 5 x lod5mol dm-3 methyl viologen at pH 10.

OXIDATION OF

OH-

A N D METHYL VIOLOGEN

For the experiments shown in fig. 12, a TiO2 sol at pH 10 was used where the colloidal Darticles were neeativelv charged. The solution also contained methvl viologen, MV2+( 1,l '-dimet~yl-4-4'-dipy~dylium . . chloride), a compound often.usid . . .. . . . . . as an electron scavenger. MV" is strongly adsorbed at the colloidal particles in alkaline solution. The 392nm absorption of the radical cation MV'+ (which is a stable radical in the absence of air) was present immediately after the laser flash [curve ( a ) in fig. 121. The conductivity of the solution was also recorded [curve ( b ) ] : it decreased after the flash following first-order kinetics, the half-life being 8.5 ps. With increasing OH- concentration the pseudo-first-order rate constant of this decrease became greater, as can be seen from fig. 13 (upper curve). The decrease in conductivity is attributed to the oxidation of OH- ions by positive holes after the flash, i.e. to the first step in the evolution of oxygen or hydrogen peroxide at the TiO, particles. As expected, the decrease in conductivity was not observed in the absence of methyl viologen. Methyl viologen interfered with e--h+ recombination, leaving an excess of positive holes on the colloidal particles which finally oxidized OH-. Experiments were also carried out with platinum as the electron scavenger. The conductivity again decreased after the flash, the rate of this decrease, however, being approximately three times smaller than for the Ti02-MV2+colloid. This may result from the fact that the TiO, particles on which MV2+is adsorbed carry less overall
--.7+ *

View Online

162

INTERMEDIATES OF

TiO2-CATALYSED PHOTOREACTIONS

- 4
LCI

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

1
0

v1

2
1 0

[OH-]/ 1 0-3 rnol dm-3

Fig. 13. Pseudo-first-order rate constant of the decrease in conductivity after flashing, (, 0 ) a Ti02-Pt sol (3.8 x lop3rnol dm-3 Ti02 and 1.6 x lop5 mol dm-3 pt) and (---, 0)
a TiOz sol (6.3 X lop3mol dmP3 Ti02) containing 5 x mol dm-3 methyl viologen.

0.3
t^ 0.2 >
v

0.1

0
[OH-]/

0.5
rnol dm-3

1
mol dm-3 methyl viologen.

Fig. 1 . Quantum yield of MV" as a function of the OH- concentration for a 6.3 X 4
mol dmF3Ti02 sol containing 5 x

negative charge. OH- ions from the bulk can therefore more easily diffuse to the trapped holes. The lower curve in fig. 13 shows the pseudo-first-order rate constant at various OH- concentrations. Finally, fig. 14 shows the quantum yield for the reduction of MV2+ as a function of the OH- concentration. Large yields, 4 = 0.35, are observed at the higher OH- concentrations. The lower yields below mol dm-3 are interpreted as being caused by less adsorption of MV2+ at the less negatively charged colloidal particles. The consumption of OH- as measured by the decrease in conductivity was the same as the yield of MV" at all OHconcentrations investigated. An interesting side reaction was observed for the T i 0 2- MV2+system in alkaline solution. The solution became strongly fluorescent after some illumination, A, ,, of the fluorescence band lying at 5 17 nm. This effect became less pronounced when a solution was used which also contained poly(viny1 alcohol). As this polymer is a hole scavenger, it is concluded that methyl viologen not only scavenges electrons but also holes, although to a much lesser degree, the product having fluorescent properties. As will be described elsewhere," fluorescing products are also found in the oxidation of methyl viologen in homogeneous solution by 'OH radicals. The

View Online

D. BAHNEMANN, A. HENGLEIN AND L. SPANHEL

163

oxidation product in the heterogeneous photo-oxidation at TiOz was found to be I ',2'-dihydro- I , 1'-dimethyl-2'-oxo-4,4'-bipyridylium chloride.
TWO-ELECTRON REDUCTION OF HALOTHANE

Downloaded by University of Nottingham on 28 July 2011 Published on 01 January 1984 on http://pubs.rsc.org | doi:10.1039/DC9847800151

In studies on microheterogeneous systems, reductions in which two electrons are transferred to a molecule are of special interest. The reduction of water, for example, requires two electrons. The kinetics of two-electron-transfer processes in both colloidal metal and colloidal semiconductor solutions have been studied in this l a b ~ r a t o r y . ' ~ ~ ' ~ Monig et al. have shown that halothane, CF,-CHClBr, undergoes an interesting reduction in homogeneous solution in the presence of free radicals, XH.*' If an electron donor D such as vitamin C is also present, the product of reduction is the F- anion, which can readily be traced by ion chromatography and specific fluoride electrodes: CF,-CHClBr+XH -+ CF,-CHCl+X+H'+Br(7) CF,-CHCI+D
-+

F-+CF,=CHCl+D'.

(8)

In the presence of oxygen, F- is not formed as the intermediate radical CF,-CHC1 is intercepted. Monig and coworkers also found that halothane in an illuminated Ti02-Pt sol forms F- even in the presence of oxygen. This can be understood in terms of the rapid pick-up of a second electron by the intermediate radical CF,-CHC1 at the Pt deposit of the colloid. Halothane thus appears to be an interesting probe of two-electron-transfer reactions at an interface. The details of these investigations will be published elsewhere.*' Mrs Gunkel is thanked for her help in the preparation of the colloids and in carrying out steady-state illuminations and Mr Kormann is thanked for his help in the hydrogen measurements. We also thank Miss Chapman for a critical reading of the manuscript.

' ' '


lo I'

l3
14

"
16

l7
18

l9

2o
21

A. Henglein, J. Phys. Chem., 1979, 83, 2209; A. Henglein and J. Lilie, J. Am. Chem. SOC., 1981, 103, 105. A. Henglein, Ber. Bunsenges. Phys. Chem., 1982, 86, 241. M. GutiCrrez and A. Henglein, Ber. Bunsenges. Phys. Chem., 1983, 87, 474. B. Kraeutler and A. J. Bard, J. Am. Chem. SOC.,1978, 100, 5985. B. Reichmann and C. E. Byvik, J. Phys. Chem., 1981, 85, 2255. H. Hada, Y. Yonezawa and M. Saikawa, Bull. Chem. SOC.Jpn., 1982, 55, 2010. H. Reiche and A. J. Bard, J. Am. Chem. Soc., 1979, 101, 3127. D. Bahnemann, A. Henglein, J. Lilie and L. Spanhel, J. Phys. Chem., 1984, 88, 709. P. Salvador and C. GutiCrrez, J. Electrochem. SOC.,1984, 131, 326. E. C. Dutoit, F. Cardon and W. P. Gomes, Ber. Bunsenges. Phys. Chem., 1976, 80, 475. D. Duonghong, E. Borgarello and M. Gratzel, J. Am. Chem. Soc., 1981, 103, 4685. A. Henglein, J. Phys. Chem., 1982, 86, 2291. J. H. Baxendale, P. L. T. Bevan and D. A. Scott, Trans. Faraday SOC., 1968, 64,2389. G. E. Adams, J. W. Boag, J. Currant and D. B. Michael, in Pulse Radiolysis, ed. M. Ebert, J. P. Keene, A. J. Swallow and J. H. Baxendale (Academic Press, London, 1965), pp. 117-129. D. Bahnemann, to be published. B. Cereek, M. Ebert, C. W. Gilbert and A. J. Swallow, in Pulse Radiolysis, ed. M. Ebert, J. P. Keene, A. J. Swallow and J. H. Baxendale (Academic Press, London, 1965), pp. 83-98. D. Bahnemann and Ch-H. Fischer, to be published. A. Henglein, Ber, Bunsenges. Phys. Chem., 1980, 84, 253. A. Henglein, M. GutiCrrez and Ch-H. Fischer, Ber. Bunsenges. Phys. Chem., 1984, 88, 170. J. Monig, K. Krischer and K-D. Asmus, Chem. Biol. Interactions, 1983, 45, 43. K-D. Asmus, D. Bahneniann, R. Chapman and J. Monig, to be published.

Das könnte Ihnen auch gefallen