Sie sind auf Seite 1von 6

LWT - Food Science and Technology 61 (2015) 1e6

Contents lists available at ScienceDirect

LWT - Food Science and Technology


journal homepage: www.elsevier.com/locate/lwt

Selection of photocatalytic bactericidal titanium dioxide (TiO2)


nanoparticles for food safety applications
Veerachandra K. Yemmireddy, Yen-Con Hung*
Department of Food Science and Technology, University of Georgia, 1109 Experiment Street, Griffin, GA 30223-1797, USA

a r t i c l e i n f o a b s t r a c t

Article history: The main objective of this study was to develop a systematic testing protocol for selecting bactericidal
Received 15 September 2014 TiO2 nanoparticles (NPs). Photocatalytic bactericidal activity of TiO2 NPs at 1 mg/mL concentration was
Received in revised form tested against Escherichia coli O157:H7. The effect of source of NPs (three different commercial samples
19 November 2014
referred as T1, T2 and T3), bacterial cell wash conditions (1 vs 3 wash), volume of reaction mixture (10, 20
Accepted 21 November 2014
and 30 mL) and intensity of UVA light (1 vs 2 mW/cm2) on bactericidal activity has been determined.
Available online 28 November 2014
Sample T3 was found to be the most effective among the tested TiO2 samples. Increasing the number of
cell washes from 1 to 3 increased the log reduction (2.91 vs 4.57). Increasing the light intensity increased
Keywords:
TiO2
the overall log reduction (3.27 vs 4.22). Decreasing the volume of reaction mixture increased the log
Nanoparticles reduction. This study has identified the best testing protocol for evaluating TiO2 NP bactericidal efficacy
Photocatalyst as single wash of bacterial cells with a reaction mixture volume of 20 mL and UVA light intensity of
Bactericidal activity 2 mW/cm2. In addition, it was found that photocatalytic oxidation of organic dyes can be used a quick
Escherichia coli O157:H7 and easier way to screen bactericidal TiO2 NPs prior to actual microbiological tests.
© 2014 Elsevier Ltd. All rights reserved.

1. Introduction or is restricted by regulations such as pharmaceutical and food


industries (Skorb et al., 2008). In addition, TiO2 becomes super-
Advanced oxidation processes based on heterogenous photo- hydrophilic upon irradiation with UV light and this functionality is
catalysis using photocatalytic nanoparticles (NPs) is gaining reversible and depends on the light exposure (Chen & Mao, 2007).
popularity in food safety applications. Heterogenous photocatalysis These properties of TiO2 help to reduce the usage of cleaning agents
utilizes light along with a semiconductor NPs to produce reactive and to shorter cleaning cycles in the food industry.
oxygen species (ROS) which can inactivate bacteria and degrade a Since the photochemical sterilization of Escherichia coli using Pt-
wide range of chemical contaminants (Mills & Le Hunte, 1997). Of TiO2 was reported by Matsunaga, Tomoda, Nakajima, and Wake
the available semiconductor NPs which can be used as photo- (1985), TiO2 photocatalysts have extensively studied to disinfect a
catalysts, TiO2 is generally considered to be the best semiconductor broad spectrum of microorganisms including viruses, bacteria,
photocatalyst available at present (Mills & Lee, 2002) due to its fungi, algae, and cancer cells (Kim, Kim, Cho, & Cho, 2003). The
strong oxidizing power at ambient temperature and pressure, sta- bactericidal properties of TiO2 are attributed to the high redox
ble, non-toxic, cheap and readily available. TiO2 has been approved potential of the surface species also known as ROS, such as hydroxyl
by the Food and Drug Administration (FDA) for use in human food, radical (OH), superoxide radical (O$ 2 ), hydrogen peroxide (H2O2)
drugs, cosmetics, and food contact materials (Chorianopoulos, formed by photo-excitation. The type and the source of TiO2 plays
Tsoukleris, Panagou, Falaras, & Nychas, 2011). TiO2 photocatalysts an important role during bacterial inactivation because the rate of
generate strong oxidizing power when illuminated with UV light of formation of ROS is a function of the particle size, crystalline phase,
wavelength less than 385 nm. TiO2-mediated photooxidation the isoelectric point, specific surface area of the nanostructure
shows promise for the elimination of microorganisms in areas (Hitkova et al., 2012). On the other hand, the biological parameters
where the use of chemical cleaning agents or biocides is ineffective of the microorganisms such as microbial species, growth phase,
initial cell density etc., are also important and photocatalytic
disinfection process may vary depending on the light intensity, the
* Corresponding author. Tel.: þ1 770 412 4739; fax: þ1 770 412 4748. wavelengths and experimental conditions (Hitkova et al., 2012).
E-mail addresses: kranti@uga.edu (V.K. Yemmireddy), yhung@uga.edu
Several commercial photocatalytic TiO2 products available on the
(Y.-C. Hung).

http://dx.doi.org/10.1016/j.lwt.2014.11.043
0023-6438/© 2014 Elsevier Ltd. All rights reserved.
2 V.K. Yemmireddy, Y.-C. Hung / LWT - Food Science and Technology 61 (2015) 1e6

market and notably, Degussa P25 TiO2, which is considered a 230 rpm and 37  C for 16 h. Following the incubation, cells were
standard and often used as a comparison in scientific experimen- harvested by sedimentation either once (4000  g for 12 min) or
tation for determining photocatalytic activity (Mills & Le Hunte, resuspended and sedimented three times (3200  g for 10 min) in
1997). Hoffmann, Martin, Choi, and Bahnemannt (1995) have sug- sterile phosphate-buffered saline (PBS, pH 7.2) in order to deter-
gested that the anatase/rutile structure of P25 promotes charge- mine the effect of cell harvesting conditions on the photocatalytic
pair separation and inhibits recombination. The different disinfection efficacy. The harvested cells were re-suspended in
electron-hole pair recombination lifetimes and interfacial electron- 10 mL PBS and equal volume of each strain suspension were
transfer rate constants may be due to the different preparation combined to obtain 10 mL of a five strain cocktail containing
methods of the samples that result in different crystal defect approximately 108 CFU/mL. Cell concentration was adjusted by
structures and surface morphologies (Chen & Mao, 2007). measuring the absorbance of bacterial suspension at 600 nm using
Several studies claimed high bactericidal activity of synthesized UV/Vis spectrophotometer (Beckman DU520, Beckman Coulter Inc.,
and commercial TiO2 NPs that are equivalent to Degussa P-25 Brea, CA, USA) and confirmed by plating 100 mL portions of the
(Hitkova et al., 2012; Kim et al., 2003). However, none of these appropriate serial dilution on tryptic soy agar (TSA) (Difco Labo-
studies have considered important factors like bacterial cell har- ratories) plates incubated at 37  C for 24 h.
vesting conditions and reaction mixture volume that may nega-
tively influence the results of photocatalytic disinfection. Hence
developing a systematic testing protocol to evaluate bactericidal 2.3. Photocatalytic experiments
efficacy of NPs helps proper selection of TiO2 NPs for food safety
related applications. The overall objective of this study was to The photocatalytic disinfection experiments were carried out in
develop a systematic testing protocol to determine the bactericidal a sterile glass petri-dish (90  18 mm2; diameter  depth) mounted
efficacy of different TiO2 nanoparticles in suspension. Specific ob- on a magnetic stirrer (Model# H1190M, Hanna Instruments,
jectives include: i) to determine the effect of type and source of TiO2 Smithfield, RI, USA) together placed in a photocatalytic disinfection
NPs on bactericidal activity, ii) to determine the effect of cell har- chamber (Fig. 1). Aqueous suspension of TiO2 NPs of 9, 18 and 27 mL
vesting conditions, light intensity and volume of suspension on volume were added into the petri-dish along with 1, 2 and 3 mL of
bactericidal activity. bacterial cultures, respectively. In this way, the effect of volume of
suspension (10, 20 and 30 mL) on the photocatalytic bactericidal
activity was investigated for each commercial TiO2 sample indi-
2. Materials and methods
vidually. The initial concentration of the bacterial culture in the
suspension was fixed at approximately 107 CFU/mL. The petri-dish
2.1. Nanoparticles
with bacteria-NP suspension was illuminated with a UVA light
system fitted with four 40 W lamps (American DJ®, Model UV Panel
Total three different types of commercial TiO2 NPs (referred as
HP™, LL-UV P40, Los Angeles, CA, USA) from the top under
T1, T2 and T3) of known characteristics (Table 1) were used in this
continuous stirring at medium speed with a magnetic stirbar
study. Samples T1 and T2 were of anatase crystal phase with
(3.8 cm length  1.25 cm diameter) (Fig. 1). The intensity of the
10e25 nm size and >99% purity. While sample T3 is mixture of
light was measured using a UV radiometer (Peak sensitivity
anatase/rutile phase (~80:20 wt %) with ~21 nm size and >99.5%
365 nm, UVP®, Upland, CA, USA). The light intensity reaching the
purity. The details of the nanoparticle characteristics as provided by
surface of the sample was adjusted to either 1 or 2 mW/cm2 (±0.15)
the individual manufacturer were listed in Table 1. Aqueous sus-
by changing the distance between the light source and the sample.
pensions of TiO2 NPs at 1 mg/mL concentration were prepared by
A positive (photocatalyst in dark) and a negative (without photo-
sonication in water-bath (Model # FS30, Fisher Scientific, Waltham,
catalyst under UVA light) control samples were also included. All
MA, USA) for about 1 h at 23  C using sterile deionized water. The
the experiments were conducted at room temperature using in-
suspensions were prepared fresh every time prior to each photo-
door air as oxidant. A sample of 1 mL was withdrawn from the
catalytic disinfection experiment.

2.2. Bacterial strains and inoculum preparation

Five strains of E. coli O157:H7 isolated from different sources:


E009 (beef), EO932 (cattle), O157-1 (beef), O157-4 (human), and
O157-5 (human) were used in this study. All bacterial strains were
stored at 70  C in tryptic soy broth (TSB) (Difco, Becton Dickinson,
Sparks, MD, USA) containing 20% glycerol. Prior to the experiment,
cultures were activated at least twice by growing them overnight in
10 mL of TSB at 37  C. Later each bacterial stain was cultured
separately in 10 mL of TSB and kept on a shaking incubator at

Table 1
Characteristics of commercial TiO2 NPs.

TiO2 samplea Crystal phase Purity (%) Size (nm) Specific surface True
area (m2/g) density
(g/cm3)

T1 Anatase 99.5 10e25 50e150 NA


T2 Anatase >99 10e25 200e240 3.9
T3 Anatase-Rutile 99.5 ~21 35e65 NA
Fig. 1. Schematic of photocatalytic disinfection set-up; 1. Wooden chamber; 2. UVA
a
Commercial TiO2 samples T1: Sky Spring Nanomaterial's; T2: US Research light bulbs (40 W each); 3. Magnetic stirrers; 4. Glass petridish with stirbar; 5. Height
Nanomaterials; T3: Degussa P-25 from Aldrich. to adjust light intensity.
V.K. Yemmireddy, Y.-C. Hung / LWT - Food Science and Technology 61 (2015) 1e6 3

treatment solution at every 30 min for 3 h and added into 9 mL particular, the difference in the log reduction among the tested
sterile PBS. Appropriate serial dilutions of the samples were pre- samples might be attributed to the differences in the surface
pared and the surviving bacteria from the control and treatments characteristics of the individual TiO2 NPs (Table 1). Sample T1 and
were enumerated by spiral plating 50 mL of each dilution on TSA. T2 used in this study have a wide particle size distribution of
The plates were incubated at 37  C for 24 h, and colonies were 10e25 nm and sample T3 has an average particle size of 21 nm. Size
counted and recorded as log CFU per mL. It was noticed that under of the particle is an important factor which influences the quantum
tested conditions, positive and negative controls had negligible yield of ROS responsible for the photocatalytic disinfection.
effect on log reduction. Changes in particle size influence photoactivity through changes in
Further, the photocatalytic activity of TiO2 samples was evalu- surface area, light scattering and light absorptivity. Gerischer
ated by degradation of methylene blue (MB) solution. The photo- (1995) demonstrated that the quantum yield increases with
decay rate was measured by using 20 mL of TiO2 aqueous solution decreasing illumination intensity and size of the particle during
(1 mg/mL) saturated with MB dye (20 mg/L) in a petridish illumi- photocatalysis. Bui et al. (2008) pointed out that the differences in
nated with UVA light at 2 mW/cm2 under continuous stirring as particle size, surface area and charge among different varieties of
described earlier. A 3 mL of the sample was collected at every TiO2 in powder affect the photocatalytic efficiency. In current study,
60 min for 3 h and the TiO2 NPs were separated by centrifuging the though sample T1 and T2 have similar size distribution
suspension at 4000 rpm for 15 min at 4  C. The photodecay rate of (10e25 nm), T2 has showed more bactericidal activity compared to
MB was determined by measuring the absorbance of supernatant at sample T1 under similar test conditions. This difference could be
664 nm using UVeVis spectrophotometer. Residual concentration attributed to the high particle specific surface area (SSA) of T2
of MB (mg/L) due to TiO2 photocatalytic activity was calculated by (200e240 m2/g) compared to T1 (50e150 m2/g). Increasing the
using standard MB curve. particle surface area provides high relative OH ion coverages for
hydroxyl radical (OH$) formation which is an important ROS
2.4. Data analysis responsible for photocatalytic disinfection. In contrast, sample T3
with mixed phase of anatase (~80%) and rutile (~20%), an average
All the experiments were replicated three times. Data were particle size of ~21 nm and relatively low SSA (35e65 m2/g) have
subjected to an analysis of variance with a completely randomized showed highest bactericidal activity than T1 and T2.
factorial design. Statistical analysis was performed using SAS As reported in several studies, high bactericidal activity of mixed
(2008) General Linear Model procedure performed with SAS Soft- phase of anatase-rutile TiO2 powder, particularly Degussa P-25
ware Release 9.3 (SAS Institute). T-tests were used for pairwise (sample T3) attributed to the generation of high amount of ROS
comparisons. Least significant difference of means tests was done such as OH$, O-2 and H2O2 due to effective charge separation by
for multiple comparisons, and all tests were performed with a level avoiding electronehole pair recombination during photocatalytic
of significance 0.05. disinfection process. Gumy et al. (2006) demonstrated that neither
a high surface area nor a high aggregate size can be the sole
3. Results and discussion properties of a TiO2 photocatalyst leading to optimal E. coli inacti-
vation. Another study reported that Degussa P-25 obtained by
3.1. Effect of type and source of TiO2 flame pyrolysis (Degussa, 1997) is an effective photocatalyst for
bacterial inactivation. Bickley, Gonzalez-Carreno, Lees, Palmisano,
The reduction trend of TiO2 (T1, T2 and T3) samples using single and Tilley (1991) reported that the dynamic process of Degussa P-
wash of bacterial cells, 20 mL bacteria-NP suspension and 2 mW/ 25 preparation would create a complex variety of multiphased
cm2 UVA light intensity over a 3 h photocatalytic disinfection particles characterized by juxtaposition of anatase with rutile
treatment was shown in Fig. 2. Under tested conditions, Sample T3 phases. This condition results in increased charge separation and
has exhibited highest log reduction (5.78) followed by T2 (2.59) and slows down electron (e) e hole (hþ) pair recombination rate
T1 (1.81), respectively. This indicates that the type and source of which in-turn results in high photocatalytic activity. Nguyen, Amal,
TiO2 has played an important role on the log reduction. In and Beydoun (2005) reported for the same mass of NPs, dispersed
particle sizes of a commercial TiO2 PC 500 are approximately three
times larger than TiO2 P-25. As a result P-25 provides a larger
surface area per unit weight for contact with bacteria than PC 500
and exhibits improved efficiency. This indicates that the prepara-
tion method of the TiO2, NP surface structure, size, and crystallo-
graphic structure all played an important role during the interfacial
charge transfer between TiO2 and E. coli leading to disinfection
(Gumy et al., 2006). The results of current study further support the
importance of type and source of TiO2 on bactericidal activity.
Degussa P-25 was found to be the most effective among tested
commercial TiO2 NPs.
However, several studies that reported high bactericidal activity
of other synthesized or commercial TiO2 NPs did not accounted for
the effect of other influencing factors such as bacterial culture
harvesting method and volume of suspension used during photo-
catalytic disinfection. These factors are important and need to be
considered when selecting TiO2 NPs for food safety applications.

3.2. Effect of cell harvesting conditions


Fig. 2. Effect of TiO2 source and bacterial cell harvesting conditions on the log
reduction UVA: Without NP UVA alone; T1: Sky Spring Nanomaterial's; T2: US The reduction trend of samples T1, T2 and T3 with respect to
Research Nanomaterials; T3: Degussa P-25. number of cell washes at 2 mW/cm2 UVA intensity and 20 mL
4 V.K. Yemmireddy, Y.-C. Hung / LWT - Food Science and Technology 61 (2015) 1e6

reaction mixture volume was shown in Fig. 2. Using single wash of effect on efficacy of individual TiO2 samples. Increasing the UVA
bacterial cells, sample T1 and T2 showed reductions of only 0.14, intensity from 1 to 2 mW/cm2 significantly increased the log
0.62 log CFU/mL after 120 min and 1.81, 2.51 log after 180 min reduction of T1 (2.04e3.04), T2 (2.99e3.94) and T3 (4.77e5.67).
treatment, respectively. Both samples required at least 90 min for Benabbou, Derriche, Delix, Lejeune, and Guillard (2007) reported a
initiating photocatalytic disinfection. Whereas, sample T3 being decrease in light intensity from 3.85 to 0.48 mW/cm2 increased the
more reactive right after 30 min treatment showed a reductions of time necessary to totally inactivate E. coli (3 log) from 90 to 180 min.
3.12 and 5.78 log CFU/mL after 120 and 180 min, respectively. It Increasing the light intensity increases the amount of photon
should be noted that depending on the reactivity of NP there will be generation which results in more electronehole pairs formation,
a lag in bacterial cell killing. This is in part attributed to the time eventually leading to the formation of more OH radicals (Marugan,
required for the ROS to react with bacterial cell membrane and van Grieken, Cassano, & Alfano, 2010). On the other hand, Cho,
facilitate damage and destruction of intracellular components and Chung, Choi, and Yoon (2004) reported the existence of linear
eventual death of cell. It is clearly visible that sample T3 is the most correlation between inactivation of E. coli and OH concentration.
reactive among the tested TiO2 samples. This linear dependence of reaction rate with the photon flux is only
However, increasing the number of cell washes from 1 to 3 found at low intensities of irradiation, because at high intensities
showed significant change in the reduction efficacy of all three TiO2 the concentration of charge carriers is so high that recombination is
samples. Sample T1 and T2 started killing bacterial cells from more favored, limiting the efficiency of the process (Herrmann,
around 30 min (a decrease of initial reactivity time at 90 min for 1 1999). However, it is believed that the tested intensity range
wash) and showed a reduction of 3.28 and 4.48 log in 3 h respec- (1e2 mW/cm2) of current study operating in the linear region
tively which is about 81 and 78% respective increase from the corresponding to the optimal light utilization. This implies that the
reduction at 1 wash treatment. Sample T3 is least benefited by saturation of TiO2 acting as the photosensitizer has not been
increasing number of cell washes as it showed no difference in the reached when increasing intensity from 1 to 2 mW/cm2. Benabbou
reduction potential at the end of 3 h treatment. However, sample T3 et al. (2007) noticed an induction period in the first 10 min for the
showed better reduction between the treatment times 90 (98% lower intensities, which suggests that the self-defense and auto-
increase) and 120 (44% increase) min when compared with 1 wash repair mechanism for protecting the bacteria were more efficient
treatment. These results indicate possible damage to the bacterial at a low intensity, but still insufficient as regard the accumulation of
cell membrane during additional centrifugation/sedimentation irradiation. Similar results were also observed in the current study
prior to the photocatalytic disinfection treatment might have with an initial lag period of ~30 min at low intensity (i.e. 1 mW/
enhanced reduction efficacy of TiO2 NPs. cm2) for the photocatalytic disinfection (data not shown). To
Peterson, Sharma, van der Mei, and Busscher (2012) stated that further evaluate the photocatalytic disinfection efficacy of TiO2 NPs,
centrifugation in essence involves compacting bacteria into a pellet, the effect of volume of suspension has also been investigated for
causing collisions against each other that result in shear forces on optimal light utilization.
the bacterial cell surface, which may lead to cell surface damage
with a potential effect on the outcome of surface-sensitive exper-
iments. Gilbert, Evans, Evans, Duguid, and Brown (1991) reported a 3.4. Effect of volume of suspension
decrease of 25 and 40% in the viability of exponential-phase Pseu-
domonas aeruginosa following centrifugation at 5000  g and Effect of volume of TiO2 NP-bacterial suspension (10, 20 and
10,000  g respectively. Similar experiments with stationary and 30 mL) on photocatalytic bactericidal activity for 3 h was shown in
exponential phase E. coli cells greatly altered biocide sensitivity Table 2. Statistical analysis of the data revealed that the overall log
(Gilbert, Pemberton, & Wilkinson, 1990). Pembrey, Marshall, and reduction of TiO2 is higher (4.32) at lower volume (10 mL) when
Schneider (1999) reported loss of viability and modification of compared to higher volumes (20 and 30 mL). No significant dif-
physicochemical cell surface properties of E. coli or Staphylococcus ference in the reduction has been observed between 20 and 30 mL
epidermis by high-speed centrifugation (15,000  g) when volume of suspensions at the end of 3 h treatment time. As ex-
compared to harvesting at (5000  g). Similarly, subjecting the pected, all the tested TiO2 samples (T1, T2 and T3) exhibited high
bacterial cells to multiple washing steps in current study might reduction at lower volume of suspension (10 mL). However, volume
have damaged the outer cellular membrane and makes it more does not showed any effect on overall reduction potential of least
susceptible to ROS attack during photocatalytic disinfection treat- effective TiO2 sample T1 at the end of 3 h (avg ~ 2.5 log). Whereas,
ment. Especially, increasing the number of cell washes resulted sample T2 has showed significantly higher reduction (4.30) at
higher bactericidal efficacies by less effective commercial samples 10 mL and almost similar reductions at 20 and 30 mL (~3 log). The
T1 and T2 (Fig. 2). This situation may leads to false prediction of most effective sample T3 has only showed significant difference in
bactericidal efficacy of TiO2 NPs while selecting NPs for food safety the reduction between 10 mL (5.64 log) and 30 mL (4.70 log).
applications such as coating on food contact and non-food contact
surfaces. The results of this study suggest that bacterial harvesting
conditions are at-most important to accurately determine bacteri- Table 2
cidal activity of photocatalytic NPs. It is recommended to use less Effect of light intensity and volume on bactericidal activity of TiO2 NPs.
severe harvesting conditions depending on the type of test
Variable Log reductiona (CFU/mL) after 3 h treatment
organism.
Overall T1 T2 T3

3.3. Effect of light intensity Intensity (mW/cm ) 2

1 3.27B 2.04B 2.99B 4.77B


2 4.22A 3.04A 3.94A 5.67A
Effect of UVA light (365 nm peak wavelength) intensity on the
Volume (mL)
log reduction of TiO2 samples (T1, T2 and T3) was shown in Table 2. 10 4.32A 3.02A 4.30A 5.64A
Increasing the light intensity from 1 mW/cm2 to 2 mW/cm2 20 3.55B 2.32A 3.09B 5.32AB
increased overall log reduction of TiO2 NPs from 3.27 to 4.22. 30 3.36B 2.28A 3.01B 4.70B
Although the trend of bactericidal efficacy is unchanged (i.e. a
Mean values with the same superscript in the same column within the same
T3 > T2 > T1), the intensity of UVA light has showed significant variable combination are not significantly different (p > 0.05).
V.K. Yemmireddy, Y.-C. Hung / LWT - Food Science and Technology 61 (2015) 1e6 5

These results indicate that volume has significant effect on the operational variables such as catalyst concentration or light in-
log reduction of individual TiO2 NPs. We expected improved effi- tensity. However, different microbiological aspects (osmotic stress,
cacy of least effective TiO2 sample (T1) by optimal light utilization repairing mechanism, regrowth, bacterial adhesion to TiO2 surface,
at lower volume of 10 mL. However, as per the current study, no etc) makes disinfection kinetics significantly much complex than
significant difference in the log reduction has been observed by the oxidation of chemical compounds (Marugan et al., 2010). Hence
decreasing the suspension volume from 30 mL to 10 mL for least certain analogies and differences exist between photocatalytic
effective sample T1. Even though it is beyond the scope of our study oxidation and photocatalytic disinfection. However, photocatalytic
to understand the agglomerate size of NPs in suspension, its effect degradation of organic dyes such as MB can be used a quick and
on ROS generation potential of individual TiO2 samples should not easier test to screen bactericidal TiO2 NPs for food safety applica-
be ruled out. If agglomerates were formed even providing more tions prior to actual microbiological tests.
light flux by decreasing the volume of suspension will not improve
photocatalytic disinfection efficacy. To support this hypothesis, 4. Conclusions
Gumy et al. (2006) reported that out of several surface properties,
aggregate size of several commercial NPs in suspension played an Photocatalytic disinfection efficacy of three different commer-
important role during the interfacial charge transfer between TiO2 cial TiO2 NPs to inactivate E. coli O157:H7 has been systematically
and E. coli leading to bacterial abatement. Agglomerated condition investigated. Type and source of TiO2 has showed significant effect
reduces effective surface area of NP available for bacteria to come on log reduction. Among the tested commercial TiO2 samples, T3
in-contact while stirring the suspension during photocatalytic (Degussa P-25) was found to be the most efficient photocatalyst
disinfection. Further we observed, a minimum of 20 mL volume of followed by T2 and T1. The same trend has been observed for
suspension is necessary for effective mixing of NP and bacteria photocatalytic degradation of MB solution. Increasing the number
while conducting bactericidal efficacy tests in suspension. of bacterial cell washes from 1 wash to 3 wash prior to photo-
catalytic disinfection treatment increased the log reduction of even
3.5. Comparison of photocatalytic oxidation and photocatalytic least effective TiO2 samples. It is preferred to use less severe cell
disinfection rates of TiO2 harvest conditions for accurate determination of bactericidal effi-
cacy of TiO2 NPs. As expected, increasing the light intensity
The photocatalytic degradation rate of MB by three commercial increased the overall log reduction of all TiO2 samples. Volume of
TiO2 samples was shown Fig. 3. Under tested conditions, sample T3 suspension has showed variable effect on efficacy of tested TiO2
was most efficient (26% decay) followed by sample T2 (10%) and NPs. As per current study, using 20 mL of suspension with single
T1(1%) in the degradation of MB solution by photocatalytic oxida- wash of cells at less severe harvest conditions and 2 mW/cm2 UVA
tion. Similar photocatalytic disinfection trend of TiO2 samples (T1, intensity was found to be best testing protocol for evaluating
T2 and T3) was observed for E. coli inactivation (Fig. 3). This shows bactericidal efficacy of TiO2 NPs.
that the photocatalytic degradation potential of organic com-
pounds such as MB can be used as a prior screening test to indi-
Acknowledgments
rectly predict bactericidal efficacy of TiO2 NPs since the mechanism
of both the processes depend on ROS generation rate. Chen, Yang,
Funding for this study was provided by Agriculture and Food
Xu, Wu, and Zhang (2009) reported apparent correlation between
Research Initiative grant no 2011-68003-30012 from the USDA
the photocatalytic processes of decomposing formaldehyde and
National Institute of Food and Agriculture, Food Safety: Food Pro-
inactivating E. coli. They noticed similar trend with respect to key
cessing Technologies to Destroy Foodborne Pathogens Program-
parameters such as light intensity, initial concentration, and type of
(A4131).
nanomaterial on the effect of photodegradation and disinfection.
This study concluded analogy is potential method to evaluate the
antimicrobial effect based on organic compound degradation effect. References
Similarly, Marugan et al. (2010) reported that analogies between
Benabbou, A. K., Derriche, Z., Delix, C., Lejeune, P., & Guillard, C. (2007). Photo-
photocatalytic degradation of chemicals and microorganism inac- catalytic inactivation of Escherichia coli: effect of concentration of TiO2 and
tivation were agreeable only when analyzing the effect of microorganism, nature, and intensity of UV irradiation. Applied Catalysis B:
Environmental, 76(3e4), 257e263.
Bickley, R. I., Gonzalez-Carreno, T., Lees, J. S., Palmisano, L., & Tilley, R. J. D. (1991).
A structural investigation of titanium dioxide photocatalysts. Journal of Solid
State Chemistry, 92(1), 178e190.
Bui, T. H., Felix, C., Pigeot-Remy, S., Herrmann, J. M., Lejeune, P., & Guillard, C. (2008).
Photocatalytic inactivation of wild and hyper-adherent E. coli strains in pres-
ence of suspended or supported TiO2. Influence of the Isoelectric point, of the
particle size and of the adsorptive properties of titania. Journal of Advanced
Oxidation Technologies, 11(3), 510e518.
Chen, X., & Mao, S. S. (2007). Titanium dioxide nanomaterials: synthesis, properties,
modifications, and applications. Chemical Reviews, 107, 2891e2959.
Chen, F., Yang, X., Xu, F., Wu, Q., & Zhang, Y. (2009). Correlation of photocatalytic
bactericidal effect and organic matter degradation of TiO2. Part I: observation of
phenomena. Environmental Science & Technology, 43(4), 1180e1184.
Cho, M., Chung, H., Choi, W., & Yoon, J. (2004). Linear correlation between inacti-
vation of E. coli and OH radical concentration in TiO2 photocatalytic disinfection.
Water Research, 38, 1069e1077.
Chorianopoulos, N. G., Tsoukleris, D. S., Panagou, E. Z., Falaras, P., & Nychas, G.-J. E.
(2011). Use of titaniumdioxide (TiO2) photocatalysts as alternative means for
Listeria monocytogenes biofilm disinfection in food processing. Food Microbi-
ology, 28, 164e170.
Degussa, A. G. P. (1997). Highly dispersed metal-oxides prepared by the Aerosil pro-
cedure No. 56, D-1600. Frankfurt 11, Germany.
Fig. 3. Comparison of photocatalytic degradation rate of methylene blue and photo- Gerischer, H. (1995). Photocatalysis in aqueous-solution with small TiO2 particles
catalytic disinfection rate of E. coli O157:H7 among different TiO2 NPs. T1: Sky Spring and the dependence of the quantum yield on particle-size and light-intensity.
Nanomaterial's; T2: US Research Nanomaterials; T3: Degussa P-25. Electrochimica Acta, 40(10), 1277e1281.
6 V.K. Yemmireddy, Y.-C. Hung / LWT - Food Science and Technology 61 (2015) 1e6

Gilbert, P., Evans, D. J., Evans, E., Duguid, I. G., & Brown, M. R. W. (1991). Surface Matsunaga, T., Tomoda, R., Nakajima, T., & Wake, H. (1985). Photoelectrochemical
characteristics and adhesion of Escherichia coli and Staphylococcus epidermis. sterilization of microbial-cells by semiconductor powders. Fems Microbiology
Journal of Applied Bacteriology, 71, 72e77. Letters, 29(1e2), 211e214.
Gilbert, P., Pemberton, D., & Wilkinson, D. E. (1990). Synergism within polyhexa- Mills, A., & Le Hunte, S. (1997). An overview of semiconductor photocatalysis.
methylene biguanide biocide formulations. Journal of Applied Bacteriology, 69, Journal of Photochemistry and Photobiology A: Chemistry, 108, 1e35.
593e598. Mills, A., & Lee, S.-K. (2002). A web-based overview of semiconductor
Gumy, D., Morais, C., Bowen, P., Pulgarin, C., Giraldo, S., Hadju, R., et al. (2006). photochemistry-based current commercial applications. Journal of Photochem-
Catalytic activity of commercial of TiO2 powders for the abatement of the istry and Photobiology A: Chemistry, 152, 233e247.
bacteria (E. coli) under solar simulated light: influence of the isoelectric point. Nguyen, V. N. H., Amal, R., & Beydoun, D. (2005). Photocatalytic reduction of sele-
Applied Catalysis B: Environmental, 63, 76e84. nium ions using different TiO2 Photocatalysts. Chemical Engineering Science, 60,
Herrmann, J. M. (1999). Heterogeneous photocatalysis: fundamentals and applica- 5759e5769.
tions to the removal of various types of aqueous pollutants. Catalysis Today, 53, Pembrey, R. S., Marshall, K. C., & Schneider, R. P. (1999). Cell surface analysis
115e129. techniques: what do cell preparation protocols do to cell surface properties?
Hitkova, A., Stoyanova, N., Ivanova, M., Sredkova, V., Popova, R., Iordanova, A., et al. Applied and Environmental Microbiology, 65, 2877e2894.
(2012). Study of antibacterial activity of nonhydrolytic synthesized TiO2 against Peterson, B. W., Sharma, P. K., van der Mei, H. C., & Busscher, H. J. (2012). Bacterial
E. coli, P. Aeruginosa and S. Aureus. Journal of Optoelectronics and Biomedical cell surface damage due to centrifugal compaction. Applied and Environmental
Materials, 4(1), 9e17. Microbiology, 78(1), 120e125.
Hoffmann, M. R., Martin, S. T., Choi, W., & Bahnemannt, D. W. (1995). Environmental Skorb, E. V., Antonouskaya, L. I., Belyasova, N. A., Shchukin, D. G., Mo € hwald, H., &
applications of semiconductor photocatalysis. Chemical Reviews, 95(1), 69e96. Sviridov, D. V. (2008). Antibacterial activity of thin-film photocatalysts based on
Kim, B., Kim, D., Cho, D., & Cho, S. (2003). Bactericidal effect of TiO2 photocatalyst on metal-modified TiO2 and TiO2:In2O3 nanocomposite. Applied Catalysis B,
selected food-borne pathogenic bacteria. Chemosphere, 52(1), 277e281. 84(1e2), 94e99.
Marugan, J., van Grieken, R., Cassano, A. E., & Alfano, O. M. (2010). Kinetic modelling
of the photocatalytic inactivation of bacteria. Water Science and Technology,
61(6), 1547e1553.

Das könnte Ihnen auch gefallen