Sie sind auf Seite 1von 9

Water Research 94 (2016) 341e349

Contents lists available at ScienceDirect

Water Research
journal homepage: www.elsevier.com/locate/watres

Review

Application of ultraviolet light-emitting diodes (UV-LEDs) for water


disinfection: A review
Kai Song, Madjid Mohseni, Fariborz Taghipour*
Department of Chemical and Biological Engineering, The University of British Columbia, 2360 East Mall, Vancouver, BC V6T 1Z3, Canada

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

Article history: Ultraviolet (UV) disinfection is an effective technology for the inactivation of pathogens in water and is of
Received 28 October 2015 growing interest for industrial application. A new UV source d ultraviolet light-emitting diode (UV-LED)
Received in revised form d has emerged in the past decade with a number of advantages compared to traditional UV mercury
29 February 2016
lamps. This promising alternative raises great interest in the research on application of UV-LEDs for
Accepted 1 March 2016
Available online 2 March 2016
water treatment. Studies on UV-LED water disinfection have increased during the past few years. This
article presents a comprehensive review of recent studies on UV-LEDs with various wavelengths for the
inactivation of different microorganisms. Many inconsistent and incomparable data were found from
Keywords:
UV disinfection
published studies, which underscores the importance of establishing a standard protocol for studying
Ultraviolet light-emitting diode (UV-LED) UV-LED inactivation of microorganisms. Different UV sensitivities to UV-LEDs and traditional UV lamps
Water treatment were observed in the literature for some microorganisms, which requires further investigation for a
Inactivation effectiveness better understanding of microorganism response to UV-LEDs. The unique aspects of UV-LEDs improve
inactivation effectiveness by applying LED special features, such as multiple wavelengths and pulsed
illumination; however, more studies are needed to investigate the influencing factors and mechanisms.
The special features of UV-LEDs offer the flexibility of novel reactor designs for a broad application of UV-
LED reactors.
© 2016 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
2. Inactivation effectiveness of UV-LEDs at various wavelengths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
3. Microorganism response to UV-LEDs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
4. Time-response data on UV-LED disinfection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
5. Mechanism of inactivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
6. Future research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
7. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348

1. Introduction people around the world lack access to a safe drinking water source
and are threatened by waterborne diseases annually (Hatami, 2013;
Drinking water safety is an important issue worldwide, espe- WHO, 2014). The development of efficient water treatment tech-
cially in most developing countries and rural areas. Millions of nologies, especially inactivation of pathogenic microorganisms in
water, is of great significance for human health and well-being.
Ultraviolet (UV) radiation can effectively inactivate various mi-
croorganisms in water (Hijnen et al., 2006) and has been increas-
* Corresponding author.
ingly used for water disinfection. UV radiation has numerous
E-mail address: fariborz.taghipour@ubc.ca (F. Taghipour).

http://dx.doi.org/10.1016/j.watres.2016.03.003
0043-1354/© 2016 Elsevier Ltd. All rights reserved.
342 K. Song et al. / Water Research 94 (2016) 341e349

advantages over conventional chemical disinfection (e.g., chlori- inactivation kinetics, may not apply directly to UV-LEDs. Moreover,
nation or ozonation), such as no chemical addition, no harmful few studies currently exist that investigate the application of UV-
disinfection by-products (DBPs) formation, and no introduction of LEDs to water disinfection. Therefore, a comprehensive research
disinfectant-resistance to bacteria (Mori et al., 2007). UV disinfec- on UV-LEDs for water disinfection is essential to better understand
tion has been recommended as a substitute for chemical additives the feasibility and future applications of this technology.
for surface water treatment (USEPA, 2006). Currently, there are This review presents and discusses published data from the
more than 7000 municipal UV disinfection installations in the literature, aiming to give an overall understanding of UV-LED water
world (Muller, 2011), and small household UV disinfection systems disinfection. Using the results of this review, some of the main
are also available (Brownell et al., 2008). It is estimated that the future research directions are identified. To the best of the authors'
global market for UV disinfection equipment has a potential to knowledge, this is the first review paper on the application of newly
reach $2.8 billion by 2020 (Allied Analytics LLP, 2014). The main UV emerging UV-LEDs for water disinfection.
sources for current UV disinfection systems are low- or medium-
pressure mercury lamps (Chevremont et al., 2013a). Although 2. Inactivation effectiveness of UV-LEDs at various
these lamps are widely used in water treatment systems, there are wavelengths
still many issues with them. The major concern is that the UV lamps
are fragile and contain toxic mercury, which is hazardous to the There has been a lack of uniformity in research materials and
environment and requires proper disposal (Chevremont et al., methods reported for UV-LED disinfection studies over the last
2013b; Close et al., 2006). Moreover, these lamps require signifi- decade, making comparisons difficult. Table 1 summarizes the main
cant amounts of energy to operate due to a low wall plug efficiency results from the published work on UV-LED water disinfection. The
of around 15e35% and have a relatively short lifetime of about data on UV dose and log inactivation were obtained from each
10,000 h (Autin et al., 2013; Chatterley and Linden, 2010). study, and the UV dose response, as the value of UV dose/log inac-
In the past few years, with the rapid development and tivation (unit: mJ/cm2 per log inactivation), was calculated to
improvement of the semiconductor industry, UV light-emitting evaluate and compare the effectiveness of disinfection among
diodes (UV-LEDs) have emerged as a new source for UV radiation different studies.
generation. An LED is a semiconductor device that utilizes semi- The germicidal efficiency of UV radiation was reported to be
conducting materials to create a pen junction (hole and electron). highly dependent on the wavelength, and the spectral sensitivity of
The electrons and holes recombine at the junction to emit radia- microorganisms was found to not necessarily follow the DNA
tion, and the wavelength of the radiation depends on the semi- absorbance spectrum (Chen et al., 2009; Mamane-Gravetz et al.,
conducting materials. Commercial visible LEDs have been available 2005). Therefore, UV wavelength is an essential factor for micro-
for nearly 50 years and have diverse applications, especially in the organism inactivation, and the effectiveness may vary from
lighting industry, due to the increasingly higher efficiency and microorganism to microorganism (Linden et al., 2001; Vilhunen
lower cost (Ibrahim et al., 2014). Recently, the newly emerging UV- et al., 2009). Different UV wavelengths in the range of
LEDs have followed a similar track and are expected to be 315e400 nm (UVA), 280e315 nm (UVB), and <280 nm (UVC) have
economically viable in the coming years (Harris et al., 2013). been applied for microorganism inactivation by UV-LEDs. From the
UV-LEDs at various wavelengths can be manufactured using published data, sorted by wavelength (Table 1), the influence of UV-
different semiconducting materials. The most frequently used LED wavelengths on inactivation effectiveness can be evaluated.
materials are III-nitride, including gallium nitride (GaN), Hamamoto et al. (2007) and Mori et al. (2007) applied UVA
aluminium gallium nitride (AlGaN), and aluminum nitride (AlN) radiation with 365-nm UVA-LEDs on Escherichia coli and obtained
(Khan et al., 2005). The wavelength of GaN-based UV-LEDs can be UV dose responses of 55,263 and 13,846 mJ/cm2 per log inactiva-
as short as 365 nm, which is in the near UV region (Taniyasu et al., tion, respectively (Table 1). These values are high considering that
2006b). However, the AIN UV-LEDs are reported to emit UV radi- typically the UV dose required by 254-nm mercury lamps for 4 log
ation at 210 nm (deep UV), which is the shortest wavelength among E. coli inactivation is about 8 mJ/cm2, which makes the UV dose
semiconductors (Taniyasu et al., 2006a). A wavelength from 210 to response only 2 mJ/cm2 per log inactivation (Bolton and Cotton,
365 nm (covering from deep UV to near UV regions) is available 2008). With the help of titanium dioxide (TiO2), Xiong and Hu
from the emission of AlGaN, which consists of AIN and GaN in (2013) established a photocatalytic disinfection system using 365-
appropriate proportions (Taniyasu and Kasu, 2010). Because the nm UV-LEDs, and the results still showed a high UV dose
wavelength is found to be an essential factor for water disinfection response of 229 mJ/cm2 per log inactivation for E. coli inactivation.
efficiency (Vilhunen et al., 2009), the ability of UV-LEDs to offer a These results demonstrated that 365-nm UVA-LEDs alone are not
variety of wavelengths is well aligned with the needs of efficient effective for microorganism inactivation.
disinfection, making it a potential option. Oguma et al. (2013) used 310-nm UVB-LEDs for E. coli inacti-
In addition to diversity in wavelengths, UV-LEDs possess several vation and reported a UV dose response of 94.8 mJ/cm2 per log
unique advantages such as environmental friendliness (no mer- inactivation, which is much lower than that required for 365-nm
cury), compactness and robustness (more durable), faster start-up UVA-LEDs. Nonetheless, this result was far greater than the
time (no warm-up time), potentially less energy consumption, values reported for UVC-LEDs for various microorganisms, which
longer lifetime, and the ability to turn on and off with high fre- ranged from 1.0 to 30.5 mJ/cm2 per log inactivation (Table 1).
quency (Wurtele et al., 2011). It is predicted that by 2020, UV-LEDs Therefore, UVB-LEDs alone are also not highly effective for micro-
will operate at 75% wall plug efficiency with a lifetime longer than organism inactivation.
100,000 h, comparable to the operating parameters of current Aoyagi et al. (2011) selected 255-nm and 280-nm UVC-LEDs to
visible LEDs (Autin et al., 2013; Ibrahim et al., 2014). All these fac- study the inactivation effects on bacteriophages 4X174, Qb, and
tors make UV-LEDs a promising alternative to conventional UV MS2. The results showed that the UV dose responses at 280 nm for
mercury lamps for water disinfection. However, due to the sub- these three microorganisms were 2.8, 28.7, and 30.5 mJ/cm2 per log
stantial differences between the traditional mercury lamps and inactivation, respectively (Table 1), which were all higher than
newly emerging UV-LEDs, the numerous research results and those at 255 nm (1.7, 12.5, and 12.8 mJ/cm2 per log inactivation,
established methodologies on water disinfection by mercury respectively). The results indicated that UVC-LEDs at 255 nm could
lamps, such as experimental protocols, reactor designs, and be more effective than at 280 nm for the inactivation of
K. Song et al. / Water Research 94 (2016) 341e349 343

Table 1
Summary of dose response data for water disinfection by UV-LEDs.

Wavelength Microorganism Disinfection UV dose (mJ/ Log Dose response (mJ/cm2 per log Reference
(nm) medium cm2) inactivation inactivation)

250 B. subtilis Water 59.2 3 19.7 Morris, 2012


254 Mesophilic bacteria Water 0.73 0.8 1.0 Chevremont et al., 2012a
255 4X174 Water 6.4 3.7 1.7 Aoyagi et al., 2011
255 Qb Water 30 2.4 12.5 Aoyagi et al., 2011
255 MS2 Water 41 3.2 12.8 Aoyagi et al., 2011
255 MS2 Water 60 2.3 26.1 Bowker et al., 2011
255 T7 Water 20 3.9 5.1 Bowker et al., 2011
255 E. coli Water 9 2.7 3.3 Bowker et al., 2011
265 E. coli Water 20 3.4 5.9 Chatterley and Linden,
2010
265 E. coli Water 10.8 4 2.7 Oguma et al., 2013
265 Pseudomonas Biofilm in tube 7.8 4 2.0 Bak et al., 2010
aeruginosa
269 B. subtilis Water 40 5.9 6.8 Wurtele et al., 2011
275 MS2 Water 60 2.1 28.6 Bowker et al., 2011
275 T7 Water 20 4.7 4.3 Bowker et al., 2011
275 E. coli Water 9 3.8 2.4 Bowker et al., 2011
280 E. coli Water 13.8 4 3.5 Oguma et al., 2013
280 Mesophilic bacteria Water 1.37 1.4 1.0 Chevremont et al., 2012a
280 4X174 Water 8.9 3.2 2.8 Aoyagi et al., 2011
280 Qb Water 43 1.5 28.7 Aoyagi et al., 2011
280 MS2 Water 58 1.9 30.5 Aoyagi et al., 2011
282 B. subtilis Water 60 7.2 8.3 Wurtele et al., 2011
310 E. coli Water 56.9 0.6 94.8 Oguma et al., 2013
365 E. coli Water 315,000 5.7 55,263 Hamamoto et al., 2007
365 E. coli Water 54,000 3.9 13,846 Mori et al., 2007
365 E. coli Water/TiO2 688 3 229 Xiong and Hu, 2013
365 Mesophilic bacteria Water 4.22 0.3 12.5 Chevremont et al., 2012a
405 Mesophilic bacteria Water 25.58 0.3 88.0 Chevremont et al., 2012a
254/365 Mesophilic bacteria Water 4.95 2.4 2.1 Chevremont et al., 2012a
280/365 Mesophilic bacteria Water 5.59 3.5 1.6 Chevremont et al., 2012a
254/405 Mesophilic bacteria Water 26.31 2.2 11.9 Chevremont et al., 2012a
280/405 Mesophilic bacteria Water 26.95 3.5 7.7 Chevremont et al., 2012a

bacteriophages 4X174, Qb, and MS2. same as long as the product of fluence rate and exposure time is the
Another study on UVC-LEDs at 255 nm and 275 nm was re- same. At the same time, these authors found a deviation from the
ported by Bowker et al. (2011) for the inactivation of three micro- reciprocity law for E. coli inactivation that showed a higher fluence
organisms: MS2, T7, and E. coli. Similar UV dose responses were rate and shorter exposure time resulted in a higher log inactivation
obtained from this study for the 255-nm and 275-nm wavelengths than a lower fluence rate and a longer exposure time despite the
(26.1, 5.1, and 3.3 mJ/cm2 per log inactivation versus 28.6, 4.3, and total UV fluence (UV dose) being the same. This phenomenon could
2.4 mJ/cm2 per log inactivation for MS2, T7, and E. coli, respectively; be attributed to UV disinfection depending not only on photo-
Table 1). For MS2, the UV dose response at 255 nm was slightly chemical reactions but also on biological processes (Harm, 1980).
lower than that at 275 nm, indicating that UVC-LED at 255 nm was Therefore, the deviation from the reciprocity law may occur on
more effective than at 275 nm. However, for T7 and E. coli, 255 nm certain microorganisms, like E. coli, because the biological pro-
resulted in slightly higher UV dose responses than at 275 nm, cesses induced by UV radiation may vary with different microor-
suggesting that 275 nm was more effective. The results for MS2 and ganisms, and some of the organisms may be more sensitive to the
T7 were consistent with their action spectra, considering that the fluence rate.
action spectrum of MS2 has a peak around 260 nm and the peak for Although some of the UV-LED dose responses reported from
T7 is around 270 nm (Mamane-Gravetz et al., 2005; Ronto et al., different studies are in agreement, there are many cases where the
1992; Beck et al., 2015). The results did not agree with the E. coli results are inconsistent. One such example is the inconsistency in
action spectrum given that 255 nm is closer to the peak around results between studies by Aoyagi et al. (2011) and Bowker et al.
260e265 nm and therefore 255 nm is expected to be more effective (2011). Both studies used 255-nm UVC-LEDs for MS2 inactivation.
than 275 nm (Bolton and Cotton, 2008). The higher effectiveness of However, in the former case, a 41-mJ/cm2 UV dose provided a 3.2
275 nm may be attributed to the higher output power of the 275- log inactivation, whereas in the latter, a 60-mJ/cm2 UV dose ach-
nm UV-LEDs, resulting in a higher fluence rate and shorter expo- ieved only 2.3 log inactivation. Similar inconsistent results were
sure time to reach the same UV dose, which is proposed to be more also observed between the studies by Chatterley and Linden (2010)
effective for E. coli inactivation than the combination of a lower and Oguma et al. (2013), where both applied 265 nm to E. coli and
fluence rate and longer exposure time (Bowker et al., 2011; Harm, reported UV dose responses of 5.9 and 2.7 mJ/cm2 per log inacti-
1980; Sommer et al., 1998). Basically, UV inactivation of microor- vation, respectively. These disagreements might result from the
ganisms is believed to follow the BunseneRoscoe reciprocity law, differences among materials and experimental conditions in the
which states that the photochemical effect depends only on the different studies. Different UV-LEDs in these studies have various
total energy dose, i.e., the product of fluence rate and exposure time radiation patterns, such as emission spectra, viewing angle, and
(Murata and Osakabe, 2013). This was verified by Sommer et al. radiation distribution. Various apparatuses were applied for UV-
(1998) who reported that for inactivation of bacteriophages MS2, LED water disinfection, and different methods were used to
4X174, and B40-8 by mercury UV lamps, log inactivation is the determine the UV dose from the UV-LEDs. Currently, there is no
344 K. Song et al. / Water Research 94 (2016) 341e349

(A) initiative has been announced, undertaken by a working group of


(B)
the IUVA Manufacturers Council, to present a consistent method-
ology for the determination and benchmarking of UVC output from
LEDs (IUVA, 2015).
UV lamp Chevremont et al. (2012a) studied the inactivation effect of
coupled UVA- and UVC-LEDs on mesophilic bacteria. The UV dose
Collimated UV-LED
responses for UVA alone at 365 nm and 405 nm were 12.5 and
beam
88 mJ/cm2 per log inactivation, which are much higher than those
for the UVC alone at 254 nm and 280 nm (both 1.0 mJ/cm2 per log
Water sample
inactivation). However, when combining UVA- and UVC-LEDs, the
container and
UV dose responses at 254/365 nm, 280/365 nm, 254/405 nm, and
Water sample stirrer
280/405 nm all sharply decreased (2.1, 1.6, 11.9, and 7.7 mJ/cm2 per
container and log inactivation, respectively), indicating that the combination of
stirrer UVA- and UVC-LEDs is more efficient than UVA-LEDs alone.
Fig. 1. Schematic diagram for typical UV lamp collimated beam apparatus (A) and UV-
Moreover, in terms of log inactivation, it was found that the com-
LED bench scale apparatus (B). bination of 280 nm and 365 nm provided higher log inactivation
than the sum of each alone (3.5 > 1.4 þ 0.3 ¼ 1.7). The other three
combinations showed similar phenomena: 254/365 nm
consistent methodology for obtaining the UV-LED dose response of (2.4 > 0.8 þ 0.3 ¼ 1.1), 254/405 nm (2.2 > 0.8 þ 0.3 ¼ 1.1), and 280/
microorganisms, and there is no standard protocol for determining 365 nm (3.5 > 1.4 þ 0.3 ¼ 1.7). This synergistic effect from the
the UV dose delivered by UV-LEDs to a sample solution. Therefore, wavelength combinations was also reported by Nakahashi et al.
the discrepancy among different studies is inevitable. (2014), who combined 254 nm with 365 nm for Vibrio para-
The standardized protocol for microorganism inactivation by haemolyticus inactivation. However, Oguma et al. (2013) combined
conventional UV mercury lamps has been well established, and the 265-nm, 280-nm, and 310-nm UV-LEDs for E. coli inactivation and
results under the standardized procedure are comparable (Bolton did not observe a synergistic effect. On the contrary, these re-
and Linden, 2003; Kuo et al., 2003). However, the UV lamp proto- searchers found that combined wavelengths were less efficient
col is not expected to be directly applicable to UV-LEDs because of than each wavelength applied separately, which could have resul-
the substantial differences between mercury lamps and UV-LEDs. ted from different indicator microorganisms and wavelength
The bench-scale UV disinfection experiments for mercury lamps combinations, as well as an inefficient thermal management of the
usually utilize a collimated beam apparatus (Fig. 1A). The key part experimental setup (Oguma et al., 2013). These studies suggest that
of this apparatusda collimated beamdis used to provide a uniform combination of selected wavelengths might be a promising way to
irradiation field on a surface area by collimating a parallel beam to improve the disinfection efficiency of UV-LEDs, but more studies
vertically project on the water surface, so that the irradiance on the are needed because the experimental setup and conditions may be
surface of the sample can be accurately measured by a radiometer, influential.
enabling the UV dose to be determined with necessary corrections Although a medium-pressure mercury lamp can provide poly-
(Bolton and Linden, 2003). To ensure the uniformity of the beam chromatic radiation, its spectrum, which includes UVA, UVB, UVC,
and accuracy of the measurement, the length of the collimated and visible light, is fixed and cannot be adjusted. Different effects
beam has to be at least 20 cm (Blatchley, 1997). Such a long colli- and mechanisms for medium-pressure mercury lamp inactivation
mated beam is not practical for UV-LEDs due to their low radiant of microorganisms have been attributed to its polychromatic radi-
power. Currently, the output power of a UVC-LED is only several ation. However, within the very broad range of wavelengths in its
milliwatts, which is much lower than that of low-pressure mercury spectrum, it is difficult to identify and distinguish which wave-
UV lamps (typically 40 W) or high-output medium-pressure mer- length or which combination of specific wavelengths the additional
cury lamps (up to 30 kW). As a result, the UV-LEDs have to be close effects and mechanisms result from. UV-LEDs provide great flexi-
to a water sample in which inactivation is performed for deliverable bility for wavelength combinations and fluence rate due to their
UV energy. Various apparatus can be applied for UV-LED inactiva- unique feature of wavelength diversity. This flexibility offers a new
tion of microorganisms. Typically, the UV-LEDs are located directly approach to tailor wavelength combinations and the radiant power
above the water sample (Fig. 1B), and the distance between UV- of each peak wavelength for optimal inactivation and to identify the
LEDs and the water sample is usually no more than 2 cm additional mechanism of a particular combination and fluence rate.
(Chevremont et al., 2012a, 2012b; Hamamoto et al., 2007; Hwang The ability to turn on and off with a high frequency is another
et al., 2013; Li et al., 2010; Mori et al., 2007; Oguma et al., 2013; unique feature of UV-LEDs, enabling adjustable UV pulsed illumi-
Vilhunen et al., 2009). Considering that a UV-LED is a point nation. Such a feature makes UV-LEDs desirable for potentially
source and emits hemispheric radiation, the UV emission from a enhancing the inactivation effectiveness by pulsed irradiation. Li
UV-LED is neither parallel nor vertical to the water surface within et al. (2010) reported enhanced germicidal effects of pulsed UV-
such a short distance between the UV-LED and the water sample LED irradiation by applying UVA-LED at 365 nm for inactivation
(Fig. 1B). As a result, uniform irradiance is not expected on the of Candida albicans and E. coli biofilms. They found that pulsed
water sample surface, which leads to difficulties for the accurate irradiation had significantly greater germicidal ability than
determination of UV dose. Further, the radiant power of UV-LEDs continuous irradiation with a maximum at 100 Hz and 75% duty
can be significantly affected by operating parameters, such as the rate (the percentage of the exposure time in total operating time)
applied current and voltage, and by thermal management during under the same UV dose. Wengraitis et al. (2013) applied pulsed
the operation. Therefore, there is a need for the development of a UVC radiation by UV-LEDs at 272 nm for E. coli disinfection on agar
standardized protocol for a UV-LED-based inactivation study of plates and found it to be more effective than continuous illumi-
microorganisms, especially for the accurate determination of UV nation; the log inactivation for pulsed UVC radiation at 272 nm with
dose and the proper operation of the system. This necessity has 1 Hz and 10% duty rate was 3.8 times higher than that of continuous
been recognized by researchers and industry leaders in the field of illumination based on the same UV dose, indicating the high effi-
UV-LEDs. Recently, an International Ultraviolet Association (IUVA) ciency of pulsed UV radiation.
K. Song et al. / Water Research 94 (2016) 341e349 345

The conventional UV source for generating pulsed UV is a xenon inactivation.


lamp. The enhanced germicidal effect of xenon lamp-pulsed UV has The k values for bacteriophage 4X174 by 255-nm and 280-nm
been demonstrated by studies for food decontamination and water UV-LEDs were reported as 0.578 and 0.360 cm2/mJ, respectively
disinfection (Bohrerova et al., 2008; Elmnasser et al., 2007; Gomez- (Aoyagi et al., 2011), which were higher than all other microor-
Lopez et al., 2007; Oms-Oliu et al., 2010). However, the pulses ganisms in Table 2, suggesting that 4X174 is very vulnerable to UVC
generated by a xenon lamp are different from those of UV-LEDs in radiation. The k value for 4X174 by 254 nm was also much higher
terms of spectrum, intensity, frequency, and duty rate. The wave- than that for most of the other microorganisms in the studies
length distribution of xenon lamp pulses ranges from 100 nm to concerning UV lamps. However, the k value of UV lamps at 254 nm
1000 nm, including UV, visible light, and infrared with a peak po- was closer to that of UV-LEDs at 280 nm instead of 255 nm. This
wer up to 35 MW. Usually the duration of each pulse is from disagreement may result from the differences between these two
nanoseconds to milliseconds, and typically 1 to 20 pulses are UV sources, including the fact that the UV fluence rate of conven-
emitted per second (Oms-Oliu et al., 2010). As discussed previously, tional UV lamps is much higher than that for UV-LEDs, which may
UV-LEDs have selectable wavelengths with relatively low radiant affect the microorganism responses (Harm, 1980). On the other
power and are capable of highly adjustable pulsed illumination hand, 4X174 inactivation by UV lamps showed no sensitivity to
with broad ranges for frequency and duty rate. The numerous different fluence rates based on the study by Sommer et al. (1998).
studies on xenon lamp-pulsed UV reveal that these pulse patterns Nonetheless, further research is needed to investigate the UV-LED
play important roles for enhanced germicidal effect (Elmnasser wavelength effect given the dearth of literature on this aspect.
et al., 2007). Therefore, due to the differences in pulse patterns, E. coli inactivation was investigated with UVC-LEDs at different
the direct applicability of the findings on xenon lamp-pulsed UV to wavelengths (255, 265, 272, 275, and 280 nm) in several studies
UV-LED pulsed illumination is not expected, and more studies on (Bowker et al., 2011; Chatterley and Linden, 2010; Oguma et al.,
UV-LED pulsed illumination are needed to confirm the enhanced 2013; Wengraitis et al., 2013). The k values ranged from 0.170 to
germicidal effect. 0.422 cm2/mJ, which were slightly lower than the values for 4X174,
indicating that E. coli is quite sensitive to UVC radiation. The wide
3. Microorganism response to UV-LEDs range of k values suggests that UV sensitivity of E. coli largely de-
pends on the wavelength. Moreover, although these studies used
Although UV radiation is effective against most pathogenic E. coli as the indicator microorganism, the strains of E. coli were
microorganisms in water, different microorganisms have different different, which could be a factor that results in the different
responses to UV radiation due to various UV resistances and pro- inactivation rate constants because UV sensitivity may vary with
cess conditions. The sensitivity of microorganisms to UV radiation different strains of a species, especially for E. coli (Malley et al.,
can be evaluated by the inactivation rate constant k (cm2/mJ) from 2004; Sommer et al., 1998, 2000). The average k value from UV
the linear portion of the relationship between log inactivation and lamp studies for E. coli inactivation was 0.506 cm2/mJ, which was
the applied UV dose (Hijnen et al., 2006): higher than the values found using UV-LEDs, with 275 nm resulting
in the closest value, rather than the 255-nm UV-LED. A possible
Log inactivation ¼ k  UV dose explanation may be the difference of fluence rate from UV lamps
and UV-LEDs since a few studies have found that the fluence rate
The k values for various microorganisms were calculated based can affect E. coli inactivation (Bowker et al., 2011; Harm, 1980;
on the data from each study (summarized in Table 1), and are listed Sommer et al., 1998). As previously discussed, although UV-LEDs
in Table 2. The data from conventional UV mercury lamps at 254 nm are usually placed very close to the water sample for UV exposure
are also shown in Table 2 for comparison. A high k value means the (Fig. 1), the fluence rate is still much lower than that of UV lamps
microorganism is UV sensitive and requires a low UV dose for

Table 2
UV sensitivity of microorganisms for UVC-LEDs and low-pressure UV lamps.

Microorganism Wavelength (nm) UV source k (cm2/mJ) Reference

4X174 255 UV-LEDs 0.578 Aoyagi et al., 2011


4X174 280 UV-LEDs 0.360 Aoyagi et al., 2011
4X174 254 UV lamps 0.396 Hijnen et al., 2006
E. coli 11229 255 UV-LEDs 0.300 Bowker et al., 2011
E. coli K12 29425 265 UV-LEDs 0.170 Chatterley and Linden, 2010
E. coli K12 IFO 3301 265 UV-LEDs 0.370 Oguma et al., 2013
E. coli 11229 275 UV-LEDs 0.422 Bowker et al., 2011
E. coli K12 IFO 3301 280 UV-LEDs 0.290 Oguma et al., 2013
E. coli 254 UV lamps 0.506 Hijnen et al., 2006
T7 255 UV-LEDs 0.195 Bowker et al., 2011
T7 275 UV-LEDs 0.235 Bowker et al., 2011
T7 254 UV lamps 0.232 Hijnen et al., 2006
Bacillus subtilis 250 UV-LEDs 0.051 Morris, 2012
Bacillus subtilis 269 UV-LEDs 0.148 Wurtele et al., 2011
Bacillus subtilis 282 UV-LEDs 0.120 Wurtele et al., 2011
Bacillus subtilis 254 UV lamps 0.059 Hijnen et al., 2006
Qb 255 UV-LEDs 0.080 Aoyagi et al., 2011
Qb 280 UV-LEDs 0.035 Aoyagi et al., 2011
Qb 254 UV lamps 0.084 Hijnen et al., 2006
MS2 255 UV-LEDs 0.078 Aoyagi et al., 2011
MS2 255 UV-LEDs 0.038 Bowker et al., 2011
MS2 275 UV-LEDs 0.035 Bowker et al., 2011
MS2 280 UV-LEDs 0.033 Aoyagi et al., 2011
MS2 254 UV lamps 0.055 Hijnen et al., 2006
346 K. Song et al. / Water Research 94 (2016) 341e349

due to the low output power of current UV-LEDs, which could be an much more effective for a specific microorganism. Unlike UV lamps,
influencing factor. Bowker et al. (2011) compared a 254-nm UV UV-LEDs can be designed to emit a particular wavelength that
lamp with 275-nm and 255-nm UV-LEDs for E. coli inactivation. The targets a specific pathogen of concern. In addition, the selected
fluence rates were 0.34, 0.094e0.11, and 0.049e0.060 mW/cm2, wavelengths could be combined by applying various UV-LEDs to
respectively. The results showed that log inactivation for the same produce a synergistic effect (Taghipour, 2013). Therefore, a
UV dose increased concurrently with the increase in fluence rate: comprehensive investigation of microorganism responses to UV-
from the 255-nm UV-LED to the 275-nm UV-LED to the 254-nm UV LEDs with different wavelengths and wavelength combinations is
lamp. These results support the hypothesis that the combination of essential to take full advantage of UV-LEDs for water disinfection.
a higher fluence rate and shorter exposure time may result in Additionally, from the comparison with conventional UV mer-
higher E. coli inactivation as previously discussed, which could cury lamps at 254 nm, some microorganisms showed different UV
explain the influence of different UV sources. sensitivities to different UV sources at the same wavelength, indi-
Bacteriophage T7 was found to have k values of 0.195 and cating that the characteristics and performance of UV sources, such
0.235 cm2/mJ by 255-nm and 275-nm UV-LEDs, respectively as the fluence rate, may play important roles in microorganism
(Bowker et al., 2011), which were slightly lower than those of E. coli, inactivation. Microorganisms were previously shown to exhibit
indicating that T7 was also sensitive to UVC radiation. The k value higher UV sensitivities to medium-pressure mercury lamps than to
for 275 nm was slightly higher than that of 255 nm, suggesting that low-pressure mercury lamps (Hijnen et al., 2006). However, few
T7 was more sensitive to 275 nm than to 255 nm, in agreement studies compare UV-LEDs and UV lamps. Thus, more studies are
with the action spectrum of T7 bacteriophage with a peak around required to better understand how different UV sources affect
270 nm (Ronto et al., 1992; Beck et al., 2015). The k value of a UV microorganism inactivation using UV-LEDs compared with UV
lamp for T7 was 0.232 cm2/mJ, which was close to that found with lamps.
the 275-nm UV-LED but far from the 255-nm UV-LED, implying
that T7 may also be sensitive to fluence rate, as previously 4. Time-response data on UV-LED disinfection
discussed.
Bacillus subtilis spores were inactivated in two studies (Morris, Some studies reported time-response data for microorganism
2012; Wurtele et al., 2011) by UV-LEDs with three different wave- inactivation by UV-LEDs instead of doseeresponse data (Table 3).
lengths (250, 269, and 282 nm). The k values for 269 nm and The UV exposure times ranged from 30 s to 2 h, and log inactivation
282 nm were 0.148 and 0.120 cm2/mJ, respectively, but the k value ranged from 2 to 7; both results showed a wide range and seemed
decreased dramatically to 0.051 cm2/mJ for 250 nm. These results incomparable. Because no quantitative data on UV dose were
seemed to be related to the action spectrum of B. subtilis spores, provided in these studies, it was difficult to analyze and compare
which have a peak around 265 nm and a trough around 240 nm the effectiveness of UV-LED inactivation. The difficulty was most
(Mamane-Gravetz et al., 2005). The k value for B. subtilis spores by a likely caused by the lack of an established standard method for the
UV lamp was 0.059 cm2/mJ, which was similar to that of a 250-nm measurement of UV dose associated with the newly emerging UV-
UV-LED, showing a consistency in the UV sensitivity of B. subtilis LEDs, which are substantially different from UV mercury lamps as
spores between the two different UV sources. discussed previously. This issue demonstrates the significance of
Bacteriophage Qb inactivation by 255-nm and 288-nm UV-LEDs UV-LED characterization and the need for standard protocols for
resulted in k values of 0.080 and 0.035 cm2/mJ, respectively (Aoyagi UV-LED microorganism inactivation and water disinfection.
et al., 2011), which were lower than most of the microorganisms
listed in Table 2, indicating that bacteriophage Qb was highly 5. Mechanism of inactivation
resistant to UVC radiation. The k value at 255 nm was significantly
higher than that at 280 nm, which agreed well with the action Many studies have discussed the mechanism of microorganism
spectrum of Qb having a peak between 260 and 265 nm (Beck et al., inactivation by UV-LEDs, and various mechanisms have been pro-
2015). The k value for the 255-nm UV-LED agreed well with the posed. Much like the conventional mercury-based UV lamps, UVC
values from low-pressure UV lamp inactivation studies for Qb radiation has proven to be efficient for microorganism inactivation,
(0.084 cm2/mJ), implying that Qb had the same UV sensitivity at and most studies on UV-LED disinfection used UVC-LEDs. UVC ra-
different UV sources. diation is believed to have direct germicidal effects by acting
Bacteriophage MS2 inactivation was reported in two studies directly on the DNA of microorganisms, leading to the formation of
(Aoyagi et al., 2011; Bowker et al., 2011) using UV-LEDs at three pyrimidine dimers and preventing them from reproducing without
different wavelengths (255, 275, and 280 nm). The k values for intermediate steps (Chatterley and Linden, 2010; Chevremont et al.,
280 nm from the former study and 255, 275 nm from the latter 2012a, 2012b; Hamamoto et al., 2007). Because DNA mainly ab-
study were close (0.033, 0.038, and 0.035 cm2/mJ, respectively) and sorbs UV radiation from 200 to 300 nm with an absorbance peak
among the lowest of all the microorganisms inactivated by UVC around 260 nm (Wurtele et al., 2011), UVC-LEDs, especially those
radiation, indicating that MS2 was highly resistant to UVC radia- with the wavelengths around 260 nm, are the most efficient for
tion. The k value for 255 nm from the former study was 0.078 cm2/ microorganism inactivation (Table 1). However, direct damage to
mJ, which was higher than the other values. This result matched the DNA might be reparable by DNA-repair mechanisms, such as
action spectrum of MS2, which has a peak around 260 nm photo-reactivation and dark repair (Oguma et al., 2001, 2013; Sanz
(Mamane-Gravetz et al., 2005; Beck et al., 2015). Although these et al., 2007). Since DNA repair is undesirable for microorganism
two studies showed different k values for the 255-nm UV-LED on inactivation, it is necessary to weaken or prevent the repair. DNA
MS2 inactivation, the average was 0.058 cm2/mJ, which was close to repair might be prevented by damaging the repair enzymes, which
that for MS2 inactivation by conventional UV lamps (0.055 cm2/mJ). are proposed to be more vulnerable to high UV intensities (Sommer
This finding suggested that MS2 might also be insensitive to et al., 1998). Moreover, the absorption spectrum of protein has a
different fluence rates from UV-LEDs and conventional UV lamps. peak around 280 nm, which might help damage repair enzymes
These studies have demonstrated the diversity of UV responses and prevent DNA repair (Kalisvaart, 2004).
for different microorganisms to various wavelengths. Because UV Although UVA radiation is poorly absorbed by DNA and is less
response largely depends on the indicator microorganisms and efficient in inducing damage on DNA (Sinha and Hader, 2002), it
applied wavelength by UV-LEDs, a particular wavelength could be still has the ability to inactivate microorganisms (Kalisvaart, 2004).
K. Song et al. / Water Research 94 (2016) 341e349 347

Table 3
Summary of time-response data for water disinfection by UV-LEDs.

Wavelength (nm) Microorganism Disinfection medium Exposure time Log inactivation Reference

254 E. coli Water 30 s 3.5 Chevremont et al., 2012b


260 E. coli Water 50 min 2.5 Nelson et al., 2013
269 E. coli Water 5 min 3e4 Vilhunen et al., 2009
276 E. coli Water 5 min 3e4 Vilhunen et al., 2009
280 E. coli Water 30 s 7 Chevremont et al., 2012b
365 E. coli Water 30 s 2.7 Chevremont et al., 2012b
365(Pulse) E. coli Biofilm 1h 3 Li et al., 2010
365(Pulse) Candida albicans Biofilm 1h 3 Li et al., 2010
365 Vibrio parahaemolyticus Water 6 min 1 Nakahashi et al., 2014
375 E. coli Biofilm 2h 2 Hwang et al., 2013
375 Streptococcus mutans Water 15 min 4 Kim et al., 2007
405 E. coli Water 30 s 3.3 Chevremont et al., 2012b
254/365 E. coli Water 30 s 7 Chevremont et al., 2012b
280/365 E. coli Water 30 s 7 Chevremont et al., 2012b
280/405 E. coli Water 30 s 7 Chevremont et al., 2012b

The current commercially available UVA-LEDs have much higher using a combination of UVA-LEDs and TiO2 (Xiong and Hu, 2013).
output power and energy efficiency than UVC-LEDs (Aoyagi et al., The mechanism for residual disinfecting effect was proposed to be
2011; Harris et al., 2013; Muramoto et al., 2014), resulting in the cumulative damage of cellular components by reactive oxygen
many studies on the application of UVA-LEDs for microorganism species or stable oxidants, such as H2O2, which could prevent the
inactivation. The inactivation mechanism of UVA radiation has not reproduction of damaged microorganisms (Pablos et al., 2013;
been studied as widely as UVC radiation because the frequently Rincon and Pulgarin, 2004, 2007; Shang et al., 2009).
used UV mercury lamps can only emit UVC radiation at 254 nm. The Many studies have been conducted on microorganism inacti-
main mechanism of UVA inactivation involves an indirect effect by vation mechanisms using different wavelengths, including UVC and
reactive intermediates and oxidative damage to DNA and other UVA, with a focus on the effect on DNA damage. However, there is
cellular components (Chatterley and Linden, 2010; Chevremont little information in the literature on DNA damage and inactivation
et al., 2012a, 2012b; Hamamoto et al., 2007; Hwang et al., 2013). mechanisms using a combination of different UV wavelengths
The reactive intermediates are proposed to be reactive oxygen (Nakahashi et al., 2014). Because a few studies have reported the
species (ROS), which are created by UVA radiation via photooxi- synergistic effect of combining particular wavelengths
dation of oxygen. Studies showed that addition of mannitol and (Chevremont et al., 2012a; Nakahashi et al., 2014), identifying the
catalase significantly protected microorganisms from 365-nm mechanisms is essential. Chevremont et al. (2012a) argued that
UVA-LEDs radiation by scavenging hydroxyl radicals (OH) and coupled wavelengths combined two UV properties: UVC induces
hydrogen peroxide (H2O2), respectively. Therefore, hydroxyl radi- direct damage on DNA, but such DNA damage can be repaired by
cals and hydrogen peroxide are believed to be the major reactive enzyme photolyase, whereas the oxidative damage to bacterial
oxygen species involved in UVA disinfection (Hamamoto et al., membranes by UVA cannot be repaired. The research on an
2007; Li et al., 2010). These reactive intermediates induce oxida- oxidative DNA product, 8-hydroxy-20 -deoxyguanosine (8-OHdG),
tive damage to DNA, proteins, and cell membranes and cause which was induced by UVA alone, and a thymine dimer, cyclo-
growth delay (Eisenstark, 1987; Oppezzo and Pizarro, 2001; butane pyrimidine dimers (CPDs), which was induced by UVC
Pizarro, 1995; Pizarro and Orce, 1988; Ramabhadran and Jagger, alone, suggested that the combination of UVA and UVC suppressed
1976; Sinha and Hader, 2002). This process takes more time than one or more recovery systems for DNA damage, such as CPDs, and
the direct damage produced by UVC radiation (Chatterley and oxidative stress from UVA may play a key role in this synergistic
Linden, 2010). Although the indirect damage by UVA radiation is effect (Nakahashi et al., 2014). It is proposed that the coupled
less efficient than the direct damage by UVC radiation for micro- wavelengths of UVA and UVC may also reduce reactivation after
organism inactivation, the damage by UVA is believed to be irrep- exposure due to the combined effects of two types of UV wave-
arable, whereas the damage by UVC is reparable through DNA- lengths (Chevremont et al., 2012a).
repair mechanisms (Oguma et al., 2013; Xiong and Hu, 2013). UV The mechanism of pulsed UV illumination by xenon lamps with
damage by low-pressure mercury lamps, which emit UVC radiation high energy and a broad spectrum has been widely studied for
at 254 nm, can be repaired relatively easily, but UV damage induced inactivation and food decontamination. Yet, its effect is not well
by medium-pressure mercury lamps, which produce UVC and UVA understood, and there is little research on pulsed illumination by
radiation together, is difficult to repair (Oguma et al., 2002; Zimmer UV-LEDs. Pulsed UV illumination has more instantaneous energy
and Slawson, 2002). Therefore, the prevention of microorganism than continuous UV illumination (Li et al., 2010), and additional
self-repair would be an advantage of UVA radiation for microor- inactivation mechanisms have been proposed, including photo-
ganism inactivation. Furthermore, UVA radiation has higher chemical, photothermal, and photophysical effects (Elmnasser
penetrability and can penetrate farther into the solution for a better et al., 2007). Other than DNA damage by UV, it is believed that
disinfection of turbid water and wastewater (Chevremont et al., pulsed UV treatment can prevent DNA repair due to inactivation of
2012a; Mori et al., 2007). the DNA-repair system and other enzymatic functions (Elmnasser
UVA radiation alone is less efficient for disinfection, but UVA et al., 2007). The UV pulse with more instantaneous energy may
radiation coupled with TiO2 could efficiently produce reactive ox- lead to localized overheating and membrane destruction
ygen species for microorganism inactivation (Marugan et al., 2010). (Krishnamurthy et al., 2007). Furthermore, the repeated and con-
Because UVA-LEDs have high output power, they are desirable for stant disturbance from high intensity pulses could induce cell
photocatalytic disinfection with TiO2. An interesting phenomenon structure damage such as cell wall rupture, membrane damage, and
called “residual disinfecting effect” was reported, in which further cellular content leakage (Krishnamurthy et al., 2010). As a result of
inactivation of E. coli was observed after a photocatalytic process these additional effects, pulsed UV illumination is reported to be 4
348 K. Song et al. / Water Research 94 (2016) 341e349

to 6 times faster than continuous UV illumination for equivalent using UV-LEDs.


inactivation levels (Fine and Gervais, 2004). These proposed The unique characteristics of UV-LEDs, particularly multiple
mechanisms are mainly based on studies of xenon lamps pulsed UV wavelengths and pulsed illumination, could improve inactivation
illumination with high energy and a broad spectrum. Due to the effectiveness under optimal conditions. However, the influencing
different pulses generated by xenon lamps and UV-LEDs, the factors and mechanisms involved need to be further investigated
applicability of these mechanisms for UV-LED pulsed illumination for a more efficient application of UV-LEDs. The special features of
still needs further examination. UV-LEDs as small-point UV sources with adjustable radiation pat-
terns provide great flexibility for novel reactor designs and new
6. Future research applications. The design of new UV-LED reactors, however, has to
be performed taking into account three phenomena: the reactor
The research on water disinfection by newly emerging UV-LEDs hydrodynamics, radiation distribution, and kinetics. Each of these
is limited. Because UV-LEDs are believed to be a promising alter- phenomena may be implemented with a higher degree of freedom
native to traditional UV mercury lamps (Muramoto et al., 2014), using UV-LEDs as the radiation source compared to UV lamps.
more research is required to better understand the application of
UV-LEDs for water disinfection. Further, some unique features of Acknowledgements
UV-LEDs such as wavelength diversity and pulsed illumination and
their effects on microorganism inactivation must be explored and The authors are grateful to the Natural Science and Engineering
understood. Given that UV-LEDs are compact and the radiation Research Council (NSERC) of Canada for financial support. K. Song is
patterns, such as emission wavelength, viewing angle, and radia- also grateful for a scholarship from the China Scholarship Council
tion distribution, are adjustable, they can enable creative reactor (CSC).
designs through the optimization of flow and radiation distribu-
tion, as well as reactor geometry and kinetics (Taghipour, 2013). The References
following are a few suggested areas for further investigation:
Allied Analytics LLP, 2014. Global UV disinfection equipment market e size, share,
global trends, company profiles, analysis, segmentation and forecast,
1) UV-LED characterization and a standard protocol for microor- 2013e2020. Res. Mark. http://www.researchandmarkets.com/reports/3066124/
ganism inactivation are necessary to obtain reliable quantitative .
information on the effectiveness of UV-LEDs. Specifically, a Aoyagi, Y., Takeuchi, M., Yoshida, K., Kurouchi, M., Yasui, N., Kamiko, N., Araki, T.,
Nanishi, Y., 2011. Inactivation of bacterial viruses in water using deep ultraviolet
standard method for UV-dose determination of UV-LEDs is
semiconductor light-emitting diode. J. Environ. Eng. ASCE 137 (12), 1215e1218.
essential for accurate and reliable UV doseeresponse data, Autin, O., Romelot, C., Rust, L., Hart, J., Jarvis, P., MacAdam, J., Parsons, S.A.,
which can be compared among different studies. Jefferson, B., 2013. Evaluation of a UV-light emitting diodes unit for the removal
2) Multiple wavelengths and pulsed illumination by UV-LEDs can of micropollutants in water for low energy advanced oxidation processes.
Chemosphere 92 (6), 745e751.
have significant impacts on inactivation effectiveness, but more Bak, J., Ladefoged, S.D., Tvede, M., Begovic, T., Gregersen, A., 2010. Disinfection of
studies are required for the fundamental understanding of these Pseudomonas aeruginosa biofilm contaminated tube lumens with ultraviolet C
phenomena, as well as determining the optimal condition. light emitting diodes. Biofouling 26 (1), 31e38.
Beck, S.E., Wright, H.B., Hargy, T.M., Larason, T.C., Linden, K.G., 2015. Action spectra
3) The additional inactivation mechanisms by different wave- for validation of pathogen disinfection in medium-pressure ultraviolet (UV)
length combinations and pulsed illumination require further systems. Water Res. 70, 27e37.
investigation, which could lead to the design of more efficient Blatchley, E.R., 1997. Numerical modelling of UV intensity: application to
collimated-beam reactors and continuous-flow systems. Water Res. 31 (9),
disinfection systems using tailored combinations of selected 2205e2218.
wavelengths and pulsation modes by UV-LEDs. Bohrerova, Z., Shemer, H., Lantis, R., Impellitteri, C.A., Linden, K.G., 2008. Compar-
4) Research on reactor designs for UV-LED water disinfection sys- ative disinfection efficiency of pulsed and continuous-wave UV irradiation
technologies. Water Res. 42 (12), 2975e2982.
tem is highly encouraged for the practical application of this
Bolton, J.R., Cotton, C.A., 2008. The Ultraviolet Disinfection Handbook. American
technology. The unique characteristics of UV-LEDs compared to Water Works Association, Denver.
traditional UV mercury lamps, such as compactness, portability, Bolton, J.R., Linden, K.G., 2003. Standardization of methods for fluence (UV dose)
determination in bench-scale UV experiments. J. Environ. Eng. ASCE 129 (3),
robustness, wavelength diversity, and pulsed illumination,
209e215.
provide flexible and diverse options for novel reactor designs, Bowker, C., Sain, A., Shatalov, M., Ducoste, J., 2011. Microbial UV fluence-response
which could also open the door to new applications of UV-LED assessment using a novel UV-LED collimated beam system. Water Res. 45 (5),
reactors. 2011e2019.
Brownell, S.A., Chakrabarti, A.R., Kaser, F.M., Connelly, L.G., Peletz, R.L., Reygadas, F.,
Lang, M.J., Kammen, D.M., Nelson, K.L., 2008. Assessment of a low-cost, point-
of-use, ultraviolet water disinfection technology. J. Water Health 6 (1), 53e65.
7. Conclusions Chatterley, C., Linden, K., 2010. Demonstration and evaluation of germicidal UV-
LEDs for point-of-use water disinfection. J. Water Health 8 (3), 479e486.
Chen, R.Z., Craik, S.A., Bolton, J.R., 2009. Comparison of the action spectra and
Newly emerging UV-LEDs provide a promising alternative for relative DNA absorbance spectra of microorganisms: information important for
water disinfection due to many advantages over traditional mer- the determination of germicidal fluence (UV dose) in an ultraviolet disinfection
of water. Water Res. 43 (20), 5087e5096.
cury lamps. The comparison of microorganism response to UV- Chevremont, A.C., Boudenne, J.L., Coulomb, B., Farnet, A.M., 2013a. Fate of carba-
LEDs and conventional UV lamps reveals that some microorgan- mazepine and anthracene in soils watered with UV-LED treated wastewaters.
isms may be sensitive to different UV sources, likely due to the Water Res. 47 (17), 6574e6584.
Chevremont, A.C., Boudenne, J.L., Coulomb, B., Farnet, A.M., 2013b. Impact of wa-
difference in the UV source radiation patterns and the fluence rates. tering with UV-LED-treated wastewater on microbial and physico-chemical
Inactivation studies of several microorganisms using UV-LED of the parameters of soil. Water Res. 47 (6), 1971e1982.
same wavelengths and comparable fluence rates, however, still Chevremont, A.C., Farnet, A.M., Coulomb, B., Boudenne, J.L., 2012a. Effect of coupled
UV-A and UV-C LEDs on both microbiological and chemical pollution of urban
show considerable discrepancies among published results. The
wastewaters. Sci. Total Environ. 426, 304e310.
inconsistent and incomparable reported results on water disinfec- Chevremont, A.C., Farnet, A.M., Sergent, M., Coulomb, B., Boudenne, J.L., 2012b.
tion by UV-LEDs along with the substantial differences between Multivariate optimization of fecal bioindicator inactivation by coupling UV-A
UV-LEDs and UV lamps highlight the importance and necessity of and UV-C LEDs. Desalination 285, 219e225.
Close, J., Ip, J., Lam, K.H., 2006. Water recycling with PV-powered UV-LED disin-
having a standard protocol for UV-LED microbial inactivation fection. Renew. Energy 31 (11), 1657e1664.
studies, especially a standard method for UV-dose determination Eisenstark, A., 1987. Mutagenic and lethal effects of near-ultraviolet radiation (290-
K. Song et al. / Water Research 94 (2016) 341e349 349

400 nm) on bacteria and phage. Environ. Mol. Mutagen 10 (3), 317e337. 1397e1403.
Elmnasser, N., Guillou, S., Leroi, F., Orange, N., Bakhrouf, A., Federighi, M., 2007. Nelson, K.Y., McMartin, D.W., Yost, C.K., Runtz, K.J., Ono, T., 2013. Point-of-use water
Pulsed-light system as a novel food decontamination technology: a review. Can. disinfection using UV light-emitting diodes to reduce bacterial contamination.
J. Microbiol. 53 (7), 813e821. Environ. Sci. Pollut. Res. 20 (8), 5441e5448.
Fine, F., Gervais, P., 2004. Efficiency of pulsed UV light for microbial decontamina- Oguma, K., Katayama, H., Mitani, H., Morita, S., Hirata, T., Ohgaki, S., 2001. Deter-
tion of food powders. J. Food Prot. 67 (4), 787e792. mination of pyrimidine dimers in Escherichia coli and Cryptosporidium parvum
Gomez-Lopez, V.M., Ragaert, P., Debevere, J., Devlieghere, F., 2007. Pulsed light for during UV light inactivation, photoreactivation, and dark repair. Appl. Environ.
food decontamination: a review. Trends Food Sci. Technol. 18 (9), 464e473. Microbiol. 67 (10), 4630e4637.
Hamamoto, A., Mori, M., Takahashi, A., Nakano, M., Wakikawa, N., Akutagawa, M., Oguma, K., Katayama, H., Ohgaki, S., 2002. Photoreactivation of Escherichia coli after
Ikehara, T., Nakaya, Y., Kinouchi, Y., 2007. New water disinfection system using low- or medium-pressure UV disinfection determined by an endonuclease
UVA light-emitting diodes. J. Appl. Microbiol. 103 (6), 2291e2298. sensitive site assay. Appl. Environ. Microbiol. 68 (12), 6029e6035.
Harm, W., 1980. Biological Effects of Ultraviolet Radiation. Cambridge University Oguma, K., Kita, R., Sakai, H., Murakami, M., Takizawa, S., 2013. Application of UV
Press, New York, pp. 127e130. light emitting diodes to batch and flow-through water disinfection systems.
Harris, T.R., Pagan, J., Batoni, P., 2013. Optical and fluidic co-design of a UV-LED Desalination 328, 24e30.
water disinfection chamber. ECS Trans. 45 (17), 11e18. Oms-Oliu, G., Martin-Belloso, O., Soliva-Fortuny, R., 2010. Pulsed light treatments
Hatami, H., 2013. Importance of water and water-borne diseases: on the occasion of for food preservation: a review. Food Bioprocess Technol. 3 (1), 13e23.
the world water day (March 22, 2013). Int. J. Prev. Med. 4 (3), 243. Oppezzo, O.J., Pizarro, R.A., 2001. Sublethal effects of ultraviolet A radiation on
Hijnen, W.A.M., Beerendonk, E.F., Medema, G.J., 2006. Inactivation credit of UV ra- Enterobacter cloacae. J. Photochem. Photobiol. B Biol. 62 (3), 158e165.
diation for viruses, bacteria and protozoan (oo)cysts in water: a review. Water Pablos, C., Marugan, J., van Grieken, R., Serrano, E., 2013. Emerging micropollutant
Res. 40 (1), 3e22. oxidation during disinfection processes using UV-C, UV-C/H2O2, UV-A/TiO2 and
Hwang, K.S., Jeon, Y.S., Choi, T.I., Hwangbo, S., 2013. Combination of light emitting UV-A/TiO2/H2O2. Water Res. 47 (3), 1237e1245.
diode at 375 nm and photo-reactive TiO2 layer prepared by electrostatic Pizarro, R.A., 1995. UV-A oxidative damage modified by environmental conditions
spraying for sterilization. J. Electr. Eng. Technol. 8 (5), 1169e1174. in Escherichia coli. Int. J. Radiat. Biol. 68 (3), 293e299.
Ibrahim, M.A.S., MacAdam, J., Autin, O., Jefferson, B., 2014. Evaluating the impact of Pizarro, R.A., Orce, L.V., 1988. Membrane damage and recovery associated with
LED bulb development on the economic viability of ultraviolet technology for growth delay induced by near-UV radiation in Escherichia coli K-12. Photochem.
disinfection. Environ. Technol. 35 (4), 400e406. Photobiol. 47 (3), 391e397.
IUVA, 2015. Proposed testing protocol for measurement of UV-C LED lamp output. Ramabhadran, T.V., Jagger, J., 1976. Mechanism of growth delay induced in Escher-
IUVA NEWS 17 (2), 7. ichia coli by near ultraviolet radiation. Proc. Natl. Acad. Sci. U. S. A. 73 (1),
Kalisvaart, B.F., 2004. Re-use of wastewater: preventing the recovery of pathogens 59e63.
by using medium-pressure UV lamp technology. Water Sci. Technol. 50 (6), Rincon, A.G., Pulgarin, C., 2004. Bactericidal action of illuminated TiO2 on pure
337e344. Escherichia coli and natural bacterial consortia: post-irradiation events in the
Khan, M.A., Shatalov, M., Maruska, H.P., Wang, H.M., Kuokstis, E., 2005. III-nitride UV dark and assessment of the effective disinfection time. Appl. Catal. B Environ. 49
devices. Jpn. J. Appl. Phys. 44 (10), 7191e7206. (2), 99e112.
Kim, B.H., Kim, D., Cho, D.L., Lim, S.H., Yoo, S.Y., Kook, J.K., Cho, Y.I., Ohk, S.H., Rincon, A.G., Pulgarin, C., 2007. Absence of E. coli regrowth after Fe3þ and TiO2 solar
Ko, Y.M., 2007. Sterilization effects of a TiO2 photocatalytic film against a photoassisted disinfection of water in CPC solar photoreactor. Catal. Today 124
Streptococcus mutans culture. Biotechnol. Bioprocess Eng. 12 (2), 136e139. (3e4), 204e214.
Krishnamurthy, K., Demirci, A., Irudayaraj, J.M., 2007. Inactivation of Staphylococcus Ronto, G., Gaspar, S., Berces, A., 1992. Phage T7 in biological UV dose measurement.
aureus in milk using flow-through pulsed UV-Light treatment system. J. Food J. Photochem. Photobiol. B Biol. 12 (3), 285e294.
Sci. 72 (7), M233eM239. Sanz, E.N., Davila, I.S., Balao, J.A.A., Alonso, J.M.Q., 2007. Modelling of reactivation
Krishnamurthy, K., Tewari, J.C., Irudayaraj, J., Demirci, A., 2010. Microscopic and after UV disinfection: effect of UV-C dose on subsequent photoreactivation and
spectroscopic evaluation of inactivation of Staphylococcus aureus by pulsed UV dark repair. Water Res. 41 (14), 3141e3151.
light and infrared heating. Food Bioprocess Technol. 3 (1), 93e104. Shang, C., Cheung, L.M., Ho, C.M., Zeng, M.Z., 2009. Repression of photoreactivation
Kuo, J., Chen, C.L., Nellor, M., 2003. Standardized collimated beam testing protocol and dark repair of coliform bacteria by TiO2-modified UV-C disinfection. Appl.
for water/wastewater ultraviolet disinfection. J. Environ. Eng. ASCE 129 (8), Catal. B Environ. 89 (3e4), 536e542.
774e779. Sinha, R.P., Hader, D.P., 2002. UV-induced DNA damage and repair: a review. Pho-
Li, J., Hirota, K., Yumoto, H., Matsuo, T., Miyake, Y., Ichikawa, T., 2010. Enhanced tochem. Photobiol. Sci. 1 (4), 225e236.
germicidal effects of pulsed UV-LED irradiation on biofilms. J. Appl. Microbiol. Sommer, R., Haider, T., Cabaj, A., Pribil, W., Lhotsky, M., 1998. Time dose reciprocity
109 (6), 2183e2190. in UV disinfection of water. Water Sci. Technol. 38 (12), 145e150.
Linden, K.G., Shin, G., Sobsey, M.D., 2001. Comparative effectiveness of UV wave- Sommer, R., Lhotsky, M., Haider, T., Cabaj, A., 2000. UV inactivation, liquid-holding
lengths for the inactivation of Cryptosporidium parvum oocysts in water. Water recovery, and photoreactivation of Escherichia coli O157 and other pathogenic
Sci. Technol. 43 (12), 171e174. Escherichia coli strains in water. J. Food Prot. 63 (8), 1015e1020.
Malley, J.P., Ballester, N.A., Linden, K.G., Mofidi, A., Bolton, J.R., Margolin, A.B., Taghipour, F., 2013. Is UV-LED the future of ultraviolet water purification? IUVA
Crozes, G., Cushing, B., Mackey, E., Lane, J.M., Janex, M.L., 2004. Inactivation of NEWS 15 (4), 23e26.
Pathogens with Innovative UV Technologies. American Research Foundation Taniyasu, Y., Kasu, M., 2010. Improved emission efficiency of 210-nm deep-
and American Water Works Association. ultraviolet aluminum nitride light-emitting diode. NTT Tech. Rev. 8 (8), 1e5.
Mamane-Gravetz, H., Linden, K.G., Cabaj, A., Sommer, R., 2005. Spectral sensitivity Taniyasu, Y., Kasu, M., Makimoto, T., 2006a. An aluminium nitride light-emitting
of Bacillus subtilis spores and MS2 coliphage for validation testing of ultraviolet diode with a wavelength of 210 nanometres. Nature 441 (7091), 325e328.
reactors for water disinfection. Environ. Sci. Technol. 39 (20), 7845e7852. Taniyasu, Y., Kasu, M., Makimoto, T., 2006b. Aluminum nitride deep-ultraviolet
Marugan, J., van Grieken, R., Pablos, C., Sordo, C., 2010. Analogies and differences light-emitting diodes. NTT Tech. Rev. 4 (12), 54e58.
between photocatalytic oxidation of chemicals and photocatalytic inactivation USEPA, 2006. Ultraviolet Disinfection Guidance Manual for the Final Long Term 2
of microorganisms. Water Res. 44 (3), 789e796. Enhanced Surface Water Treatment Rule. Office of Water, Washington DC. EPA
Mori, M., Hamamoto, A., Takahashi, A., Nakano, M., Wakikawa, N., Tachibana, S., 815-R-06e007.
Ikehara, T., Nakaya, Y., Akutagawa, M., Kinouchi, Y., 2007. Development of a new Vilhunen, S., Sarkka, H., Sillanpaa, M., 2009. Ultraviolet light-emitting diodes in
water sterilization device with a 365 nm UV-LED. Med. Biol. Eng. Comput. 45 water disinfection. Environ. Sci. Pollut. Res. 16 (4), 439e442.
(12), 1237e1241. Wengraitis, S., McCubbin, P., Wade, M.M., Biggs, T.D., Hall, S., Williams, L.I.,
Morris, J.P., 2012. Disinfection of Bacillus subtilis Spores Using Ultraviolet Light Zulich, A.W., 2013. Pulsed UV-C disinfection of Escherichia coli with light-
Emitting Diodes (MS thesis). emitting diodes, emitted at various repetition rates and duty cycles. Photo-
Muller, J., 2011. Seeing the Light: the Benefits of UV Water Treatment. Water Online. chem. Photobiol. 89 (1), 127e131.
http://www.wateronline.com/doc/seeing-the-light-the-benefits-of-uv-0001. WHO, 2014. Progress on Drinking Water and Sanitation. http://apps.who.int/iris/
Muramoto, Y., Kimura, M., Nouda, S., 2014. Development and future of ultraviolet bitstream/10665/112727/1/9789241507240_eng.pdf.
light-emitting diodes: UV-LED will replace the UV lamp. Semicond. Sci. Technol. Wurtele, M.A., Kolbe, T., Lipsz, M., Kulberg, A., Weyers, M., Kneissl, M., Jekel, M.,
29 (8). 2011. Application of GaN-based ultraviolet-C light emitting diodes e UV LEDs e
Murata, Y., Osakabe, M., 2013. The Bunsen-Roscoe reciprocity law in ultraviolet-B- for water disinfection. Water Res. 45 (3), 1481e1489.
induced mortality of the two-spotted spider mite Tetranychus urticae. J. Insect Xiong, P., Hu, J.Y., 2013. Inactivation/reactivation of antibiotic-resistant bacteria by a
Physiol. 59 (3), 241e247. novel UVA/LED/TiO2 system. Water Res. 47 (13), 4547e4555.
Nakahashi, M., Mawatari, K., Hirata, A., Maetani, M., Shimohata, T., Uebanso, T., Zimmer, J.L., Slawson, R.M., 2002. Potential repair of Escherichia coli DNA following
Hamada, Y., Akutagawa, M., Kinouchi, Y., Takahashi, A., 2014. Simultaneous exposure to UV radiation from both medium- and low-pressure UV sources
irradiation with different wavelengths of ultraviolet light has synergistic used in drinking water treatment. Appl. Environ. Microbiol. 68 (7), 3293e3299.
bactericidal effect on Vibrio parahaemolyticus. Photochem. Photobiol. 90 (6),

Das könnte Ihnen auch gefallen