Vous êtes sur la page 1sur 17

Also available online - www.brill.

nl/rci

Res. Chem. Intermed., Vol. 33, No. 35, pp. 359 375 (2007) VSP 2007.

The application of TiO2 photocatalysis for disinfection of water contaminated with pathogenic micro-organisms: a review
CATHY MCCULLAGH 1 , JEANETTE M. C. ROBERTSON 1 , DETLEF W. BAHNEMANN 2 and PETER K. J. ROBERTSON 1,
1 Centre

for Research in Energy and the Environment, School of Engineering, The Robert Gordon University, Schoolhill, Aberdeen AB10 1FR, UK 2 Institut fr Technische Chemie, Leibniz Universitt Hannover, Callinstrasse 3, D-30167 Hannover, Germany Received 1 February 2006; accepted 28 February 2006 AbstractThe use of semiconductor photocatalysis for treatment of water and air has been the topic of intense research activity over the past 20 years. This powerful process has also been extended to the disinfection of environments contaminated with pathogenic micro-organisms. This review summarizes recent developments concerned with the photocatalytic treatment of water contaminated with pathogenic micro-organisms presenting a potential hazard to animals and human beings. Keywords: Photocatalysis; TiO2 ; micro-organism; pathogen; disinfection.

INTRODUCTION

The use of TiO2 photocatalysis for remediation of contaminated water and air has been extensively reported over the past 23 years [1 9]. Since 1985, when Matsunaga [10] reported the rst application of TiO2 photocatalysis for the destruction of three different bacterial species, there has been a concerted range of research on the biological applications of the process and the destruction of bacterial species [11 15], viruses [10, 16], algae [17] and fungi [10, 18] have been reported. This has been a particularly exciting application of the process that could demonstrate particular effectiveness in water disinfection of potable supplies and help secure water sustainability in the 21st Century. The provision of safe drinking water is now globally one of the most serious challenges facing the planet. In particular areas such as the Middle East, Sub-Saharan Africa, large areas of Asia and Southern Europe have
To

whom correspondence should be addressed. E-mail: peter.robertson@rgu.ac.uk

360

C. McCullagh et al.

been experiencing particular problems. Disease is a common consequence of water shortage with diarrhoea caused by ingestion of contaminated water being a signicant problem and killer, particularly for children under ve years old. The problems with water availability may become even more severe in the next decades as a consequence of global climate change. While there will be increased availability in high latitudes of the Northern Hemisphere, reduced water availability has been predicted for southern Europe, the Middle East, central Asia and Africa [19, 20]. The Middle East, for example, currently has 1% of the Worlds available water and 5% of the population [21]. Developing nations will particularly be affected by water availability problems and there will be further pressure on water demand resulting from economic development and population growth. In many cases the problems, however, are not related to water availability but are due to water contamination. Of the 3 million deaths from water related disease each year it is estimated that 90% result from drinking water contaminated with pathogenic micro-organisms [22]. In addition new challenges in the form of bacterial resistance in drinking water are presenting serious problems to the water industry as existing technologies may not be effective in the removal of a range of bacterial species in the future. Titania photocatalysis has, however proven to be rather effective in removal of a wide range of chemical compounds and microbiological pathogens from potable water samples. In this paper we critically review a range of the literature on the application of photocatalysis for removal of biological species.

THE APPLICATION OF TIO2 PHOTOCATALYSIS FOR TREATMENT OF PATHOGENIC MICRO-ORGANISMS

From the outset of research on the disinfecting properties of TiO2 photocatalysts, a number of parameters have been found to be of crucial importance in optimising the process. These include TiO2 concentration, UV light intensity, microbial starting concentration, temperature, pH and aeration on experimental outcome [16, 23 27]. The greatest level of research activity has focussed on the destruction of Escherichia coli, which is a bacterium traditionally used as an indicator of faecal contamination of water. The destruction of a wide range of other species, including Lactobacillus acidophilus, Saccharomyces cerevisiae [10], Salmonella enterica serovar Enteritidis, Pseudomonas aeruginosa [23], Salmonella cholerasuis, Vibrio parahaemolyticus and Listeria monocytogenes [15], has also been reported. Disinfecting effect of TiO2 suspensions Many workers have investigated the use of aqueous suspensions of TiO2 photocatalysts. These suspensions required continuous stirring to ensure effective contact of the titania and target species in addition to preventing catalyst settlement. As is the case with photocatalytic destruction of chemical species the aeration of the reaction suspension is required to ensure effective scavenging of electrons and prevent

TiO2 photocatalysis for disinfection of contaminated water

361

electron/hole recombination. Furthermore, sonication of TiO2 suspensions has appeared to be very effective in increasing the rate of bacterial destruction [11, 16, 24]. Sonication also appeared to cause greater intracellular damage in damaged bacterial cells. The use of ultrasound was believed to increase catalyst surface area producing more active sites for reaction with target species. In addition, smaller TiO2 particles may enter bacterial cells which may already have damaged cell membranes and cause additional damage to intracellular components. The disadvantages of a TiO2 suspension are that at the end of the treatment period the catalyst needs to be separated from the liquid (water), and the slurry may interfere with UV light penetration through the solution. Matsunaga et al. [10] were the rst to report on the photocatalytic destruction of bacteria using TiO2 . They studied the decomposition of three species; L. acidophilus, Sm. cerevisiae and E. coli (103 cells/ml) using platinum-loaded titanium oxide and demonstrated that cell death occurred following incubation with TiO2 /Pt particles under irradiation for up to 120 min. A range of Gram-negative and Gram-positive bacterial species have subsequently been studied with E. coli nding favour with most authors. E. coli is extensively used as a treatment efciency indicator and, if not detected, the treated water is regarded as free from faecal contamination [25]. The complete destruction of 108 cells/ml of E. coli following irradiation of a Degussa P25 suspension was reported by Wei et al. [26]. In this study the effects of a number of parameters were investigated including TiO2 dosage, light intensity and irradiation under solar illumination. It was reported that cell viability decreased monotonically with an increase in the TiO2 dose. The rate of photocatalytic destruction was found to depend on light intensity, increasing proportionately with light intensity in the range between around 180 to 1660 E/s per m2 . Under direct solar illumination experiments similar results to those observed under simulated solar illumination in the laboratory were obtained. Huang et al. [27] reported similar ndings for the photocatalytic disinfection of E. coli when they investigated the effectiveness of colloidal TiO2 suspensions. The photocatalytic destruction of coliform bacteria and the polio virus in secondary wastewater efuent containing suspensions of TiO2 irradiated with either F40BL uorescent lights or sunlight was studied by Watts et al. [16]. The concentration of bacteria investigated was 5107 colony-forming units (CFU)/ml for total coliform bacteria and 1106 CFU/ml for faecal coliform bacteria, while the virus numbers investigated were 3000 plaqueforming units (PFU). A two log inactivation of the poliovirus was achieved in 30 min, while the inactivation of the coliform bacteria to a similar level took 150 min (Fig. 1). It was concluded that although photocatalysis appeared effective for inactivating bacteria and viruses, it may not be feasible compared with techniques such as chlorination and ozonation due to the potentially long contact times (over 150 min, compared to <60 min for plug ow chlorine contact chambers). Wist et al. investigated the destruction of E. coli in distilled water and in river water [28]. There was no increase in the cell count numbers 24 h after treatment in the case

362

C. McCullagh et al.

Figure 1. Inactivation of total and faecal E. coli forms at pH 8 under F40BL lights. Reproduced from Ref. [16] with permission.

of the distilled water. In the case of the river water, however, sample cell numbers increased when photocatalysis was stopped. Bacterial re-growth occurs as not all of the bacteria have been killed during the photocatalytic treatment period, thus bacteria continue to grow and reproduce after the treatment time. It may also be that following treatment the bacteria entered a viable but non-culturable stage and then recovered their culturability after a period of more favourable conditions, i.e., the dark. The effectiveness of TiO2 photocatalysis for the destruction of the food borne pathogenic bacteria, Salmonella cholerasuis, Vibrio parahaemolyticus and Listeria monocytogenes was investigated by Kim et al. [15]. Of these species S. cholerasuis was more susceptible to photocatalytic attack than the other two bacteria. This study also established that the cell killing activity depended on both illumination time and TiO2 suspension concentrations. The Gram-negative bacteria, Enterobacter cloacae, was studied by Ibez et al. as a model organism for treatment in TiO2 suspensions irradiated by with UVA light [11]. The optimum loading of TiO2 determined in previous studies for the destruction of E. coli was investigated for this system. Under the same conditions of TiO2 loading and irradiation destruction levels of greater than 99.9% were reported for E. coli, S. typhimurium, P. aeruginosa and E. cloacae after 40 min irradiation. The residual effect of TiO2 photocatalysis of E. coli K12 in deionised water was studied by Rincn et al. [29]. This work demonstrated that there was no bacterial re-growth up to 60 h after photocatalytic treatment, although this was found to be dependent on the light intensity used to irradiate the system. A residual bactericidal effect was also demonstrated by Huang et al. [27], although this group only studied samples which had been kept in the dark for 30 min after a 30-min irradiation period.

TiO2 photocatalysis for disinfection of contaminated water

363

Figure 2. Effect of UVA light and TiO2 on E. coli, S. enteridis and P. aeruginosa.

Robertson et al. [23] compared the effectiveness of TiO2 photocatalysis and UV-A photolysis for the destruction of E. coli, S. enterica and P. aeruginosa. They reported a substantial decrease in bacterial numbers when all three strains were exposed to UV-A light in a TiO2 suspension (Fig. 2). They concluded that TiO2 photocatalysis is a more effective disinfection technology when compared to UV-A irradiation. UV-A irradiation results in the production of small colony variants which were not observed with TiO2 photocatalysis (Fig. 3). It was proposed that the production of these small colony variants (SCV) of pathogenic bacteria following UV-A treatment of drinking water may represent a potential health hazard. In addition, Robertson suggested that as these SCV were not observed with the UVA/TiO2 system, thus, this potential hazard is not a risk when using this technology. This observation also indicated that the bactericidal mechanism was different with the UV-A/TiO2 process compared to when UV-A light is used alone. Figure 4 illustrates the results achieved by Huang et al. [30]. They reported that when E. coli cells underwent illumination for 15 min in the presence of 1 mg/ml TiO2 , almost all of the cells were still viable. Increasing the illumination time to beyond 30 min resulted in a loss of viability for 96% of the cells. They reported complete killing after 60 min illumination. The photocatalytic activity of 13 commercially available TiO2 powders has been reported by Gumy et al. [31]. As has been observed for many chemical species, Degussa P25 proved to be the most effective material. Additives. Rincn et al. [32] investigated the effect of additives on the photobactericidal effect of Degussa P25 TiO2 under solar illumination. They used E. coli as their model organism and found that the addition of H2 O2 positively affected

364

C. McCullagh et al.

Figure 3. E. coli colonies from UV-A-treated sample, showing both regular and small colony phenotypes. Reproduced from Ref. [23] with permission.

Figure 4. The effect of TiO2 photocatalytic reaction on cell viability. The survival curves were obtained from the viable count of TiO2 (1.0 mg/ml) and/or UV light (8 W/m2 ) treated E. coli cells. The initial cell concentration was 106 cfu/ml. (3) Cell + TiO2 in dark; ( ), cell + light; (), cell + TiO2 + light; (), cell + TiO2 , 30 min in light and 30 min in dark. Reproduced from Ref. [30] with permission.

TiO2 photocatalysis for disinfection of contaminated water

365

the inactivation of E. coli under both photolytic and photocatalytic conditions. The 2 2 addition of inorganic ions such as HCO 3 , HPO4 , Cl , NO3 and SO4 to the suspension affects the sensitivity of bacteria to sunlight in the presence and absence 2 of TiO2 . Addition of HCO 3 and HPO4 resulted in a decrease in photobactericidal effect; the other anions caused a moderate decrease. Rincn also reported that the presence of naturally occurring organic substances present in water have a negative effect on photocatalytic disinfection. Effect of pH. The pH of the solution during photocatalysis has an effect on the electrostatic charge of the TiO2 surface, which determines the density of TiOH+ 2 groups. Consequently, the adsorption on TiO2 of bacteria and therefore the activity towards destruction of bacteria by TiO2 photocatalysis is pH dependent. Microbial growth is also pH dependent, with a decrease in the growth rate observed when the pH is changed in either direction away from the optimum pH. The vast majority of human pathogens are neutrophiles, i.e., they prefer a neutral pH. Most of the reported photocatalytic reactions have been performed in the pH range 58, which does not seem to overtly affect experimental outcome [16, 24, 27]. Herrera Melin et al. [13] examined the photocatalytic destruction of coliform bacteria and Streptococcus faecalis in urban waste water using both direct solar and UV lamp light. They found no signicant differences in bacterial destruction between TiO2 photocatalysis and UV light only (either solar or UV lamp-light source) at natural pH (7.8). If, however, the pH was lowered to 5, the inactivation rate was increased in the presence of the catalyst. They also examined bacterial re-growth in natural waste water samples previously treated at pH 5 and pH 7.8 2448 h after photocatalytic treatment, a signicant rise in bacterial numbers was observed, although re-growth was slower in samples treated at pH 5. Watts et al. [16], on the other hand, found that the photocatalytic destruction of coliform bacteria and poliovirus in secondary wastewater efuent was unaffected by pH changes in the range 58. Rincn et al. [32] reported that during the disinfection of E. coli the bacteria were more sensitive towards photocatalysis at starting pH of 4.0 or 9.0 than at an initial pH of 7.0. Interestingly, in the study by Gumy et al. [31] the photocatalytic activity of P25 for destruction of E. coli was not found to be inuenced by pH; however, this was a strong inuencing factor in the other twelve materials examined. It was reported that the catalytic activity of the titania samples closely correlated the catalyst isoelectric point with the materials with lower isoelectric points demonstrating poorer photocatalytic activity for disinfection. Disinfecting effects of TiO2 thin lms The use of TiO2 lms has been investigated by several authors and found to be as effective as the suspended form [27, 32 34]. This method obviously overcomes the disadvantage of having to separate catalyst from liquid at the end of the experimental period. Sunada et al. [33] reported a mechanism for photokilling of E. coli on dip-coated TiO2 thin lms. They dropped E. coli on the TiO2 thin lm and

366

C. McCullagh et al.

Figure 5. The log plot of the survival of E. coli cells vs. illumination time. The cell suspension was incubated on TiO2 lm under UV illumination (1.0 mW/cm2 ). The initial cell concentrations were 2 105 CFU/ml ( ), 2 106 CFU/ml (), 2 107 CFU/ml ( ) and 2108 CFU/ml ( ), respectively. Survival was also determined for cells (2 105 CFU/ml) on the TiO2 lm in the dark () and on normal glass (soda-lime glass, SLG) without TiO2 lm under UV illumination ( ) (1.0 mW/cm2 ). The data shown in the gure are the average values of three experiments. Reproduced from Ref. [33] with permission.

illuminated from above through a Pyrex window in a air tight chamber. They observed a great decrease in viable cell numbers on the illuminated TiO2 lm, demonstrating its photokilling ability (Fig. 5). Kikuchi et al. [34] prepared thin lms of TiO2 on soda lime glass coated with a thin lm of silica. The TiO2 layer was prepared by dip coating into titanium isopropoxide. They applied these lms to the photocatalytic disinfection of E. coli and reported complete sterilisation of a solution from 3104 cells/ml starting concentration. However bacterial recovery after photobactericidal treatment was not reported. They also reported a membrane separated system where the E. coli suspension was separated from the TiO2 surface by a porous PTFE membrane. Total sterilisation of the solution was reported however the rate of photokilling was slower and experimental time increased by a factor of 4. They determined the role of active species eliciting the bactericidal effect to be hydrogen peroxide together with a cooperative effect due to other oxygen species. Huang et al. reported TiO2 -coated soda lime glass, the surface of which had been functionalised by SO3 H prior to being coated with TiO2 [27]. The TiO2 coating

TiO2 photocatalysis for disinfection of contaminated water

367

was prepared by immersing the glass in a TiO2 colloid for 4 h. They investigated the bactericidal effect of the thin lms towards the destruction of E. coli in air under natural sunlight. Khn et al. [18] coated plexiglass with Degussa P25 TiO2 and investigated its bactericidal effects towards the destruction of a series of hygiene relevant bacteria, E. coli, P. aeriginosa, S. aureus, E. faecium and C. albicans. They illuminated different concentrations of each species with UV-A light for an hour and reported a reduction in cell counts for each species. The susceptibility of the species investigated had the following order: E. coli > P. aeruginosa > S. aureus > E. faecium > C. albicans. This order is expected when the complexity of the cell walls of each is considered and if it is assumed that the primary site of photocatalytic attack is the cell wall. E. coli and P. aeruginosa are both Gram-negative bacteria and have thin cell walls, S. aureus and E. faecium are Gram-positive bacteria and have denser cell walls, while C. albicans has a thick eukaryotic cell wall. Saito et al. [14] suggested that OH radicals produced following activation of TiO2 attack the cell wall leading to punctures. Khn et al. [18] support this theory as they discovered that the spores of B. subtilis were not susceptible to photocatalytic attack after 60 min irradiation. They illustrate the partially destroyed C. albicans on a TiO2 -coated and irradiated plate using SEM (Fig. 6). The authors suggest that this may represent an intermediate stage towards complete mineralization. Rincn et al. [29] immobilised TiO2 on Naon membranes and reported the photobactericidal effect of thin lms towards the destruction of E. coli. The lms were prepared by dip-coating, Naon membranes were dipped in a solution of TiO2

Figure 6. SEM photograph of partially destroyed C. albicans on a TiO2 -coated and irradiated plate. Reproduced from Ref. [18] with permission.

368

C. McCullagh et al.

Degussa P25 for several hours. They also immobilised TiO2 P-25 powder, nonporous anatase, prepared from acetyl acetonate-isopropoxide, and rutile TiO2 on Pyrex glass using a solgel method. They reported that using the naon support resulted in a reduction of viable cell numbers only 1 log higher than when there was no catalyst present. The solgel-prepared lms were less efcient than using a suspension of TiO2 when similar concentrations were compared. Film thickness is an important parameter to consider, thin lms allow more depth of light penetration. Amzaga-Madrid et al. [12] prepared thin lms of TiO2 by spray pyrolysis. They prepared undoped lms and lms doped with Cu and Al, soda lime glass was used as the support. The lms showed a photoinduced bactericidal effect on P. aeruginosa (108 CFU/ml) with bacterial inhibition varying between 28 and 96%, depending on the composition of the lms. The undoped lms resulted in a greater bacterial inhibition, although the Cu-doped lms had the greatest quantum yield. Effect of an electrochemically applied potential on the photo-bactericidal effect of TiO2 thin lms This area within the photobactericidal genre is the least exploited with only a small number of authors reporting the disinfection of water using photoelectrochemistry. The application of an applied potential to enhance the photocatalytic effect of TiO2 has been reported since the early 1970s. Buttereld et al. [35] were one of the rst to report the application of it as a bactericidal technique. They used an electric eld enhanced photochemical reactor which was capable of disinfecting water containing faecal indicators. They investigated the destruction of E. coli and Cl. perfringens and reported the removal of over 3 log units (100%) of E. coli and 2 log units of Cl. perfringens, in 25 min. They suggested that Cl. perfringens death was caused by hydroxyl radicals produced by the photoelectrocatalytic system. Dunlop et al. [36] reported the photocatalytic disinfection of E. coli cells with an applied potential. They compared thin lms prepared from Degussa P25 (anatase/rutile mixture) and Aldrich TiO2 (anatase) electrophoretically. Under open circuit conditions they found that disinfection was not observed in the absence of TiO2 nor with UV-A alone. They also reported that the lm prepared from the Degussa P25 TiO2 produced superior results. They applied a potential of 1000 mV (SCE) and found a signicant increase in the disinfection rate of both lms, an increase in rate of 40% was recorded for the Degussa electrode and 80% for the Aldrich electrode (Fig. 7). They also investigated cell loading and found that increasing the cell loading increased the rate of disinfection with the effect of the applied potential more pronounced with higher cell loading. The key nding in this work was that the number of colony forming units was reduced by 99.996% after 120 min photocatalytic treatment. Previous studies have been hampered by bacterial recovery following photocatalytic disinfection, even though photocatalysis has been proven to cause cellular damage. Bacteria have several enzymatic activities that may protect them from oxidative damage, including superoxide dismutase and catalase. In addition, other enzymes, such as exonuclease III and recA

TiO2 photocatalysis for disinfection of contaminated water

369

Figure 7. Photocatalytic inactivation of E. coli using Degussa and Aldrich TiO2 electrodes. Percentage bacterial survival is plotted against disinfection time. Reproduced from Ref. [36] with permission.

protein, appear to be important in repairing DNA lesions resulting from oxidative damage [37]. Under certain conditions the applied intensity and dose of UV irradiation absorbed by E. coli is not enough to prevent enzymatic DNA repair which results in bacteria reactivation. In Dunlop et al.s work [36] they allowed 48 h for bacterial recovery from disinfection. Following this recovery period the number of CFU remained below detectable limits showing that photocatalytic disinfection had caused irreversible damage to the bacterial cells. They do err on the side of caution by highlighting the fact that bacteria may have entered a viable but not culturable state. Curtis et al. [38] investigated the use of solgel and thermally prepared TiO2 lms with an applied potential to destroy Cryptosporidium oocysts. They used the propidium iodide exclusion method to analyse for cell death. Since PI has a higher afnity for DNA than SYTO9 it is able to displace SYTO9 from the DNA. As a result, viable cells will appear green and dead appear red (damaged cells can sometimes appear orange). The thermal lms gave superior results compared to the sol gel prepared lms owing to the superior conversion of photochemical holes to hydroxyl radicals at the surface of the thermal lm. Thermal lms under black light activation caused a decrease of 50% of viable oocysts. Mechanism of cell death The mechanism of cell death following photocatalysis has been a topic of extensive research in recent years [10, 14, 30, 33, 39]. A common feature in all this mechanistic work appears to be initial cell membrane damage followed by destruction of

370

C. McCullagh et al.

intracellular components. Matsunaga et al. [10] demonstrated that Coenzyme A, a coenzyme involved in a variety of biochemical reactions, was photoelectrochemically oxidized, which inhibited cellular respiratory activity resulting in subsequent cell death. A rapid leakage of potassium ions from S. sobrinus following TiO2 photocatalysis, which occurred in parallel to cell death, was reported by Saito et al. [14]. The slow release of protein and RNA from cells during photocatalysis was also demonstrated by this group [14]. The photocatalytic destruction of E. coli endotoxin, which is an integral component of the cell membrane, on a TiO2 thin lm was studied by Sunada et al. [33]. The destruction of E. coli cells was found to be accompanied by degradation of endotoxin and this implied that TiO2 photocatalysis destroyed the outer membrane of the bacterial cell. It has also been proposed that E. coli cell death occurs by a lipid peroxidation mechanism [39]. Figure 8 illustrates a TEM of a dispersion of TiO2 Degussa P25 1 mg/l in contact with E. coli K-12 cells reported by Nadtochenko et al. [40]. The titania particles at the concentrations used are seen to form 2030 m aggregates of the basic crystal. The authors suggest that the surface interaction of TiO2 with E. coli occurring at pH 6 is favoured as E. coli is negatively charged between pH 3 and 9. Maness et al. [39] suggested that reactive oxygen species such as hydroxyl radicals, superoxide or hydrogen peroxide, gener-

Figure 8. TEM of a dispersion of TiO2 Degussa P-25 (1 mg/l) in contact with E. coli K-12 cells. Reproduced from Ref. [40] with permission.

TiO2 photocatalysis for disinfection of contaminated water

371

ated on the TiO2 surface mutually attack the polyunsaturated phospholipids in the bacterial cell membrane. This primary attack results in breakdown of the cell membrane and the photocatalyst particles may access the damaged cell via phagocytosis causing further photocatalytic damage inside the cell. A mechanism for photokilling of E. coli on dip-coated TiO2 thin lms has been proposed by Sunada et al. [33]. This group released E. coli on a thin TiO2 lm and illuminated from above through a Pyrex window in an air tight chamber. Two stages in the photokilling reaction for intact cells were suggested; an initial lower rate photokilling step followed by a higher rate process. Sunada et al. [33] found that spheroplasts which did not have a cell wall exhibited only a single step with a higher rate, suggesting that the cell wall of E. coli cell acted as a barrier to the photocatalytic process. Changes in the concentration of cell wall components following illumination further showed that the outer membrane operates as a barrier, while the peptidoglycan layer did not have a barrier function. It was concluded that the photocatalytic reaction was commenced with a partial decomposition of the outer membrane followed by disordering of the cytoplasmic membrane which resulted in cell death. The mechanism of photocatalytic cell death was studied by Huang et al. [30] using E. coli. On irradiating E. coli with near-UV light in a TiO2 suspension an immediate increase of permeability to small molecules such as o-nitrophenol -D-galactopyranosideside (OPNG), and the leakage of small molecules such as -D-galactosidase resulted after 20 min. A kinetic analysis demonstrated that cell wall damage occurred in less than 20 min, followed by progressive damage of the cytoplasmic membrane and intracellular components. Huang et al. [30] also reported that a continued bactericidal effect occurred after the UV illumination is switched off. Methods of analysis of cell destruction Selective vs. non-selective media. The choice of culture media for enumerating bacterial recovery following photocatalytic treatment is an important factor that may affect experimental results [41]. A clear difference in the responses of selective and non-selective media following photocatalysis of E. coli was reported by Rincn et al. [41]. The non-selective medium plate count agar (PCA) showed a 1000times higher response than that of the selective medium ECC (E. coli CHROMagar). The use of selective agars has been shown to reduce the growth of even the target organism and, therefore, give misleading results when compared to a non-selective agar. Consequently under certain circumstances, e.g., when treating natural water samples, when it is considered necessary to use a selective agar then it is also important to include a non-selective agar to substantiate results. Cell death reporting, standardisation required for comparison. The manner in which results of photocatalytic disinfection processes are reported is a very important consideration. Many authors report results as % survival rates or % inactivation rates; however, caution must be taken when interpreting such results.

372

C. McCullagh et al.

This is quite acceptable when considering the application of TiO2 photocatalysis for removal of chemical contaminants from water or air, however, when considering destruction of bacterial species it may be less appropriate. For example, a reduction of 1108 CFU/ml to 1103 CFU/ml equates to a 99.99% inactivation yield which on the face of it would appear an effective level of destruction. 1103 CFU/ml, however, still represents a signicant level of contamination and not only could signify a potential hazard but would allow re-growth of the micro-organisms. If results of photocatalytic destruction of micro-organisms are reported only as % inactivation this, therefore, may be seen as being misleading in terms of the experimental outcome. Authors rarely quote actual bacterial numbers when reporting results of photocatalytic disinfection processes, which if adopted as a practice would benet all researchers in this eld. Although many authors claim to have achieved complete destruction of pathogens in water others have not. It is crucial to bear in mind that, in terms of satisfactory rates of microbial destruction, there is no tolerable limit for pathogens in water intended for consumption, for preparing food or drink or for personal hygiene. Bacterial contamination falls under the category of pathogens. The EPA Maximum Contaminant Level (MCL) for coliform bacteria in drinking water is zero (or no) total coliform per 100 ml of water. Standardisation of the method of reporting photocatalytic destruction of pathogens is therefore a topic that will have to be considered in detail in the future. Standardisation of other photocatalytic processes has been an area of signicant activity. For example, since 2002, when the rst photocatalytically active selfcleaning window glass was introduced to the European market, many products have been developed and are now available possessing photocatalytic properties. At the moment, however, these self-cleaning products are only tested according to existing appropriate national or international standards that, in the case of the respective glazing quality, i.e., optical and energetic properties. In building applications, for example, the newly developed glasses must pass the EN-1096 standard. However, there are no standards to evaluate the self-cleaning performances of these products. During the last two years various other products equipped with self-cleaning surfaces have been introduced to the European and, in particular, to the German market: roof tiles, paints and ceramic tiles, to name but a few. However, quantitative measures to prove the functionality of these coatings or to even enable a comparison of the performance of products from different producers are still not available. Therefore, standardization in photocatalysis has become an important issue. Initiated by the Japanese Industrial Standard organization (JIS) a Technical Committee comprised of four sub-committees has been formed in Japan in 2002 to develop standards for photocatalytic air, water and surface cleaning as well as for the associated antibacterial effect. A rst standard directed towards the photocatalytic removal of nitrous oxides from ambient air has already been issued by JIS in January 2004 [42]. Meanwhile this standard testing method has also been proposed to the International Standard Organization ISO. In Europe, the German

TiO2 photocatalysis for disinfection of contaminated water

373

Standard organization DIN has formed a Working Committee to develop industrial standards for photocatalysis (NMP 293). Moreover, a European research project has been started recently aiming for the development of a European Norm (EN) for photocatalytically coated glazing (Self Cleaning Glass, NMP3-CT-2003-505952). In Germany, the DIN has formed the Working Committee NMP 293 in June 2004 to develop standards for photocatalytically coated surfaces. Currently, this Working Committee comprises of almost twenty members mainly from the respective industries. It was decided that the main directions to be followed by NMP 293 will be the development of testing methods for the photocatalytic degradation of chemical and biological model systems, for photoinduced surface wetting, as well as for the underlying photophysical effects. Even though all these different committees have meanwhile started their work, there are no standards to assess the antibacterial activity of photocatalytically coated products yet. Moreover, different methods are not only required to check the activity for the degradation of chemical and biological pollutants but also for gas phase applications, for the photocatalytic cleaning of water, and for the ultimate destruction of surface-adsorbed species, respectively.

CONCLUSIONS

As can be seen from the above discussion TiO2 photocatalysis is a highly versatile technique which has proven to be very effective for the treatment of a wide range of biological species in potable water. The mechanism of the process has been studied in some detail and an understanding of the photokilling process is emerging. As with many applications of photocatalysis for the removal of chemical contaminants from water there is still much work to be performed on standardising methods for assessment of the photocatalytic disinfection process and reporting levels of destruction. Photocatalytic oxidation has, however, started to make an important impact in environmental remediation of contaminated water and commercial applications of this versatile environmental treatment technology will multiply over the next decades. This can be exemplied by a recent report which detailed the scale-up of photocatalytic technology using a solar powered reactor utilising compound parabolic collectors [43]. This unit achieved effective disinfection of water contaminated with E. coli of between 30 and 60 min.

REFERENCES
1. P. V. Kamat, Chem. Rev. 93, 671 (1993). 2. A. Mills, R. H. Davies and D. Worsley, Chem. Soc. Rev., 417 (1993). 3. D. F. Ollis and H. Al-Ekabi, Photocatalytic Purication and Treatment of Water and Air, Volume 3. Elsevier, Amsterdam (1993). 4. P. Pichat, Catal Today 19, 313 (1994). 5. M. R. Hoffmann, S. T. Martin, W. Choi and D. W. Bahnemann, Chem. Rev. 95, 69 (1995).

374 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19.

C. McCullagh et al. A. Mills and S. Le Hunte, J. Photochem. Photobiol. A: Chem. 108, 1 (1997). M. L. Sauer and D. F. Ollis, J. Catal. 158, 570 (1996). R. M. Alberici and W. F. Jardim, Appl. Catal. B: Environ. 14, 55 (1997). P. Pichat, J. Disdier, C. Hoang-Van, D. Mas, G. Goutailler, C. Gaysse, Catal. Today 63, 636 (2000). T. Matsunaga, R. Tomoda, Nakajima and H. Wake, FEMS Microbiol. Lett. 29, 211 (1985). J. A. Ibez, M. I. Litter and R. A. Pizarro, J. Photochem. Photobiol. A: Chem. 157, 81 (2003). P. Amzaga-Madrid, G. V. Nevrez-Moorilln, E. Orrantia-Borunda and M. Miki-Yoshida, FEMS Microbiol. Lett. 211, 183 (2002). J. A. Herrera Melin, J. M. Doa Rodrguez, A. Viera Surez, E. Tello Rendn, C. Valds do Campo, J. Arana and J. Prez Pea, Chemosphere 41, 323 (2000). T. Saito, T. Iwase, J. Horie and T. Morioka, J. Photochem. Photobiol. B: Biol. 14, 369 (1992). B. Kim, D. Kim, D. Cho and S. Cho, Chemosphere 52, 277 (2003). R. J. Watts, S. Kong, M. P. Orr, G. C. Miller and B. E. Henry, Water Res. 29, 95 (1995). B. J. P. A. Cornish, L. A. Lawton and P. K. J. Robertson Appl. Catal. B: Environ. 25, 59 (2000). K. P. Khn, I. F. Chaberny, K. Massholder, M. Stickler, V. W. Benz, H.-G. Sonntag and L. Erdinger, Chemosphere 53, 71 (2003). J. T. Houghton, Y. Ding, D. J. Griggs, M. Noguer, P. J. van der Linden, X. Dai, K. Maskell and C. A. Johnson, Climate Change 2001: The Scientic Basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge (2001). Climate change: what we know and what we need to know, The Royal Society Policy Document 22/02, The Royal Society, London (2002). BBC News Report, available online at: http://news.bbc.co.uk/1/hi/sci/tech/2859937.stm A. Dufour, M. Snozzi, W. Koster, J. Bartram, E. Ronchi and L. Fewtrell, Assessing Microbial Safety of Drinking Water: Improving Approaches and Methods. IWA, London (2003). J. M. C. Robertson, L. A. Lawton and P. K. J. Robertson, J. Photochem. Photobiol. A: Chem. 175, 51 (2005). X. Z. Li, M. Zhang and H. Chua, Water Sci. Technol. 33, 111 (1996). K. H. Baker and D. S. Heron, Water Environ. Res. 71, 530 (1999). C. Wei, W. Lin, Z. Zainal, N. E. Williams, K. Zhu, A. P. Kruzic, R. L. Smith and K. Rajeshwar, Environ. Sci. Technol. 28, 934 (1994). N. Huang, Z. Xiao, D. Huang and C. Yuan, Supramol. Sci. 5, 559 (1998). J. Wist, Sanabria, C. Dierolf, W. Torres and C. Pulgarin, J. Photochem. Photobiol. A: Chem 147, 241 (2002). G. Rincn and C. Pulgarin, Appl. Catal. B: Environ. 44, 263 (2003). Z. Huang, P.-C. Maness, D. M. Blake, E. J. Wolfrum, S. L. Smolinski and W. A. Jacoby, J. Photochem. Photobiol. A: Chem. 130, 163 (2000). D. Gumy, C. Morais, P. Bowen, C. Pulgarin, S. Giraldo, R. Hajdu and J. Kiwi, Appl. Catal. B: Environ. 63, 76, 2005. A.-G. Rincn and C. Pulgarin, Appl. Catal. B: Environ. 51, 283 (2004). K. Sunada, T. Watanabe and K. Hashimoto, J. Photochem. Photobiol. A: Chem. 156, 227 (2003). Y. Kikuchi, K. Sunada, T. Iyoda, K. Hashimoto and A. Fujishima, J. Photochem. Photobiol. A: Chem. 106, 51 (1997). I. M. Buttereld, P. A. Christensen, T. P. Curtis and J. Gunlazuardi, Water Res. 31, 675 (1997). P. S. M. Dunlop, J. A. Byrne, N. Manga and B. R. Eggins, J. Photochem. Photobiol. A: Chem. 148, 355 (2002). E. Cabiscol, J. Tamarit and J. Ros, Int. Microbiol. 3, 3 (2000). T. P. Curtis, G. Walker, B. M. Dowling and P. A. Christensen, Water Res. 36, 2410 (2002). P. C. Maness, S. Smolinski, D. M. Blake, Z. Huang, E. J. Wolfrum and W. Jacoby, Appl. Environ. Microbiol. 65, 4094 (1999).

20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39.

TiO2 photocatalysis for disinfection of contaminated water

375

40. V. A. Nadtochenko, A.-G. Rincn, S. E. Stanca and J. Kiwi, J. Photochem. Photobiol. A: Chem. 169, 131 (2005). 41. A.-G. Rincn and C. Pulgarin, Appl. Catal. B: Environ. 49, 99 (2004). 42. JIS R 1701-1:2004, Fine ceramics (advanced ceramics, advanced technical ceramics) Test method for air purication performance of photocatalytic materials Part 1: Removal of nitric oxide. National Institute of Advanced Industrial Science and Technology and Japan Fine Ceramics Association, Tokyo (2004). 43. P. Fernndez, J. Blanco, C. Sichel and S. Malato, Catal. Today 101, 345 (2005).

Vous aimerez peut-être aussi