The Myth of Visible Light Photocatalysis Using ... - ACS Publications

The Myth of Visible Light Photocatalysis Using Lanthanide Upconversion. 1. Materials. 2. 3. Sushant P. Sahu†, Stephanie L. Cates‡, Hyoung-Il Kimâ€...
0 downloads 0 Views 1MB Size
Subscriber access provided by READING UNIV

Article

The Myth of Visible Light Photocatalysis Using Lanthanide Upconversion Materials Sushant P. Sahu, Stephanie L Cates, Hyoung-il Kim, Jae-Hong Kim, and Ezra L. Cates Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b05941 • Publication Date (Web): 06 Feb 2018 Downloaded from http://pubs.acs.org on February 7, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 22

Environmental Science & Technology

1

The Myth of Visible Light Photocatalysis Using Lanthanide Upconversion

2

Materials

3 Sushant P. Sahu†, Stephanie L. Cates‡, Hyoung-Il Kim‡, Jae-Hong Kim‡, Ezra L. Cates†*

4 5 6



Department of Environmental Engineering and Earth Sciences, Clemson University, Clemson, SC.

7



Department of Chemical and Environmental Engineering, Yale University, New Haven, CT.

8 9 10

ABSTRACT Upconversion luminescence is a nonlinear optical process achieved by certain engineered

11

materials, which allows conversion of low energy photons into higher energy photons. Of particular

12

relevance to environmental technology, lanthanide-based upconversion phosphors have appeared in

13

dozens of publications as a tool for achieving visible light activation of wide-band gap semiconductor

14

photocatalysts, such as TiO2, for degradation of water contaminants. Supposedly, the phosphor particles

15

act to convert sub-band gap energy photons (e.g., solar visible light) into higher energy ultraviolet

16

photons, thus driving catalytic aqueous contaminant degradation. Herein, however, we reexamined the

17

photophysical properties of the popular visible-to-UV converters Y2SiO5:Pr3+ and Y3Al5O12:Er3+, and

18

found that their efficiencies are not nearly high enough to induce catalytic degradations under the reported

19

excitation conditions. Furthermore, our experiments indicate that the false narrative of visible-to-UV

20

upconversion-sensitized photocatalysis likely arose due to coincidental enhancements of dye degradation

21

via direct electron injection that occur in the presence of dielectric-semiconductor (phosphor-catalyst)

22

interfaces. These effects were unrelated to upconversion and only occurred for dye solutions illuminated

23

within the chromophore absorption bands. We conclude that upconversion using Pr3+ or Er3+-activated

1 ACS Paragon Plus Environment

Environmental Science & Technology

24

systems is not a technologically appealing mechanism for visible light photocatalysis, and provide

25

experimental guidelines for avoiding future misinterpretation of these phenomena.

26

TOC GRAPHIC

27 28 29 30

INTRODUCTION Heterogeneous photocatalysis has maintained a strong foothold in water treatment technology

31

research since the explosion of studies involving TiO2 began in the 1990s. Among the many uses of

32

semiconductor photocatalysts proposed by academia in the environmental fields, advanced oxidation is

33

the most common; therein, production of hydroxyl radicals by catalyst suspensions in photoreactors is

34

seen as a “chemical free” alternative to H2O2 and O3-based unit processes for destruction of recalcitrant

35

water contaminants.1, 2 While the most effective photocatalytic materials have band gap energies (Eg) that

36

demand UV-range excitation wavelengths, many groups have pursued catalysts that are activated instead

37

by lower-energy visible light.3-6 These efforts are motivated by the prospect of replacing energy-intensive

38

UV lamp reactors with solar reactors, though the operational practicality of this concept has been

39

debated.7 The primary route to visible light activation of semiconductors is through the use of materials

40

with sufficiently low Eg or intra-band states to absorb such light and promote valence electrons into the

41

conduction band. Inherently, however, a lower Eg generally implies weaker redox potential of the

42

resulting conduction band electrons (e–cb), valence band holes (h+vb), or both.8

2 ACS Paragon Plus Environment

Page 2 of 22

Page 3 of 22

43

Environmental Science & Technology

To circumvent the weak activity of narrow-Eg materials, an alternative approach appearing with

44

increasing frequency in the literature is the use of upconversion luminescent (UC) phosphors that convert

45

the visible wavelengths into UV photons; these photons may then activate a more conventional wider-Eg

46

material, such as TiO2 (Figure 1).8-12 Though stronger redox potential is preserved, the overall quantum

47

yield of this approach is then intimately tied to the optical conversion efficiency of the UC system. The

48

current authors have previously developed visible-to-UVC converting materials and investigated their

49

application towards antimicrobial technologies.13-16 Despite the high sensitivity of microorganisms to

50

germicidal UVC, we found that the conversion efficiency of even one of the most effective lanthanide

51

(Ln3+) phosphor systems – X2-Y2SiO5:Pr3+,Li+ (YSO:Pr3+) – was prohibitively low under practical

52

excitation conditions, yielding 10-3 % conversion of blue light to UVC under fluorescent lamp

53

excitation.13, 14 (β-Y2Si2O7:Pr3+,Li+ was demonstrated to have 3.1× greater visible-to-UV conversion

54

efficiency than YSO:Pr3+, the highest yet reported, though it has not been studied for environmental

55

applications.17) We also experimented in the past with UC-sensitized photocatalysis using YSO:Pr3+

56

coupled with TiO2 and found that even under 1.4 W/cm2 laser excitation, results were inconsistent and no

57

obvious catalytic dye degradation resulting from upconverted light could be observed (unpublished). It

58

has thus been unexpected to subsequently witness a steady stream of publications over the past ten years

59

that report UC-photocatalyst composite systems that yield remarkable contaminant degradation rates

60

under low intensity (non-laser) visible light excitation conditions.

61 62

Figure 1. Mechanism of visible-to-UV upconversion sensitized photocatalysis disputed herein. Green

63

circle – UC phosphor particle; white circles – semiconductor photocatalyst particles.

64

3 ACS Paragon Plus Environment

Environmental Science & Technology

65

The first studies reporting UC-sensitized visible light photocatalysis (UC-PC) were authored by

66

Wang et al. in 2005 and 2006.18, 19 Though we count over 25 subsequent papers from these and other

67

authors focusing on this topic, we have opted to cite only a limited number as representatives in order to

68

minimize further validation of this body of work. Generally the studies share a common formula of

69

coupling a catalyst (usually TiO2 or ZnO) with a UC phosphor (usually YAlO3:Er3+ or Y3Al5O12:Er3+,

70

referred to herein as yttrium aluminum perovskite/YAP:Er3+ and yttrium aluminum garnet/YAG:Er3+,

71

respectively) via mixing and heating to produce a composite powder. The catalytic activities under visible

72

light exposure were then assessed by monitoring dye decolorization by catalyst suspensions via UV/vis

73

spectroscopy.9, 20, 21 In all cases, significant enhancement of decolorization rates by the UC-PC composite

74

were reported over the photocatalyst suspension alone, suggesting that such composites might be useful in

75

solar-driven water treatment processes. More recently, Wu et al. reported UC-PC degradation of

76

methylene blue (MB) by YSO:Pr3+ coupled with TiO2,22 with their findings contradicting the results of

77

our unpublished experiments. Though most early papers on UC-PC were confined to specialty journals,

78

the sheer number of them has resulted in substantial influence on the field. Now, visible-to-UV UC

79

phosphors have begun to appear in studies concerning more complex catalytic systems, and in journals of

80

broader scope, including ACS Applied Materials & Interfaces,23 and Environmental Science &

81

Technology.24, 25 These reports appear to treat UC-PC as a well-established phenomenon by citing

82

previous publications, and as a result the effect of UC in these experiments is largely assumed, rather than

83

proven experimentally therein.

84

In this work our objective was to disprove the role of UC in catalytic enhancements observed for

85

Ln3+ phosphor-semiconductor composites under visible light irradiation, with more emphasis on control

86

experiments than most previous studies. Though our past experience has involved YSO:Pr3+ and other

87

Pr3+-activated converters,13, 16, 17 we also included YAP:Er3+ and YAG:Er3+ in the present study due to

88

their prevalence in previous UC-PC literature.9, 10, 20, 21 Intended to result from this work is a rigorous

89

understanding of the performance of UC materials under environmentally relevant excitation conditions,

4 ACS Paragon Plus Environment

Page 4 of 22

Page 5 of 22

Environmental Science & Technology

90

thus allowing the technological value of UC-PC to be objectively interpreted by the environmental

91

engineering and applied materials chemistry communities.

92 93

EXPERIMENTAL METHODS

94

Stock Chemicals. Er(NO3)3•5H2O (> 99.99% purity) and LiNO3•xH2O (99.999%) were purchased from

95

Alfa-Aesar. Y2O3 (99.999% purity) and MB were purchased from Acros Organics. Methyl Orange (MO,

96

> 98% purity) and methyl red (MR) were supplied from TCI chemicals. Commercial TiO2 nanopowders

97

(Aeroxide P25) was purchased from Acros Organics, and potassium hydrogen phthalate was obtained

98

from EMD. Tetraethyl orthosilicate, TEOS (purity > 99.999%), Al (NO3)3•9H2O (99.997% purity),

99

Pr(NO3)3•6H2O (> 99.99%), and HNO3 (trace metal grade) were purchased from Sigma-Aldrich. High

100

purity phenol, ethanol, isopropanol, and anhydrous citric acid were obtained from Fisher Scientific and

101

were used as received. Deionized water of resistivity >18 MΩ.cm was used in all experiments.

102

Materials Synthesis. Powder X2-Y2SiO5 doped with 1.2 mol.% Pr3+ and 10 mol.% Li+ (i.e., prepared

103

assuming a composition of Y1.776Pr0.024Li0.2SiO4.8, referred to herein as YSO:Pr3+) was prepared using a

104

sol-gel decomposition synthesis, as described in our previous work.26 “Undoped” stoichiometric YSO: Li+

105

samples without luminescent Pr3+ activation were prepared following the same procedure, without

106

addition of Pr3+ ions. YAG:Er3+ phosphors of 1 mol.% Er3+ were prepared with molar Er:Y:Al ratio of

107

0.01:0.99:1.67 by a sol-gel decomposition technique similar to methods described previously.27, 28 Briefly,

108

stoichiometric amounts of Y(NO3)3 (4.373 mL, 1.77 M) and Er(NO3)3 (0.391 mL, 0.2 M) aqueous stock

109

solutions were mixed in a beaker. Separately, 4.80 g Al (NO3)3•9H2O was dissolved in 5 mL deionized

110

water, stirred for 1 h at room temperature, and then mixed dropwise with the Y3+/Er3+ solution. Citric acid

111

(11.80 g) was added in a molar ratio of citric acid: metal ion of 3:1 under stirring. The solution was

112

evaporated at ~85 ºC in a water bath until viscous gel was obtained and then dried at 130 ºC in an oven for

113

24 h and ground to obtain a precursor powder. The material was then heated at 1200 ºC for 2 h in air

114

using a tube furnace to obtain white YAG:Er3+(1%) sample. Undoped YAG containing no Er3+ ions was 5 ACS Paragon Plus Environment

Environmental Science & Technology

115

also prepared. A mixed-phase YAP/YAG phosphor material was also prepared according to Dong et al.,28

116

and we further tested physically mixed phosphor and TiO2 powder samples prepared without the sintering

117

step, in the same molar ratios as the composites.

118

YSO:Pr3+/TiO2 and YAG:Er3+/TiO2 composite materials were prepared by mixing P25 TiO2 with

119

UC phosphor powders at a molar ratio 1:0.16 (TiO2:phosphor). Powder mixtures were ground in a mortar

120

and pestle followed by annealing at 550 ºC for 3 h in a tube furnace. The 16 mol% phosphor

121

concentration was previously found to be optimum for dye removal by YAG:Er3+/TiO2 by Feng et al.10

122

Characterization. The UC photoluminescence spectra of phosphor powders were acquired on a

123

customized spectrometer by guiding a focused beam from an air-cooled argon ion laser with 488 nm line

124

filter, and using phase-sensitive detection with components described previously.13 The laser was focused

125

onto the center of each sample at a 45º angle from the sample face, and collection lenses aligned normal

126

to the sample guided the emissions through a 450 nm short pass filter to the detector. The powder X-ray

127

diffraction (XRD) patterns were collected on a Rigaku Miniflex using Cu Kα radiation (λ =1.54 Å). The

128

morphology of the UC materials and composites were studied by scanning electron microscopy (SEM),

129

using a SU6600 variable pressure SEM.

130

Chemical Analyses. Concentrations of MB were determined by absorbance at 664 nm, based on a

131

calibration curve, while 430 and 460 nm were used for MR and MO, respectively. A Varian Cary 50

132

and/or Cary 300 UV/Vis spectrophotometer were employed. Phenol concentrations were monitored via

133

HPLC using an Ultimate 3000 (Thermo Scientific) instrument coupled with variable wavelength UV

134

detector and Hypersil Gold C10 column (3 µm, 150 mm × 2.1 mm; using acetonitrile/water mixture

135

[32:68 v/v] as the mobile phase at a flow rate of 0.32 mL/min).

136

Photocatalytic Experiments. Photocatalytic degradation of dyes and phenol by various materials were

137

assessed under excitation by both a monochromatic laser and broad spectrum white light LED (W-LED)

138

at room temperature. An argon ion laser (Modulaser Stellar-Pro-L-ML 1000) was used in conjunction

139

with a 488 nm line filter, resulting in a 140 mW beam which was de-focused to a diameter matching the

140

reactor cuvette yielding an intensity of approximately 180 mW/cm2. The W-LED employed was a 6 ACS Paragon Plus Environment

Page 6 of 22

Page 7 of 22

Environmental Science & Technology

141

mounted 800 mW cold white LED (Thorlabs; spectrum shown in Figure S1), coupled with 450 nm long

142

pass filter to avoid any direct band gap excitation of TiO2 by violet or UV wavelengths, and collimated

143

onto the suspension cuvette using an achromatic lens. Some tests experiments were also conducted using

144

UVA radiation (λmax = 365 nm); suspensions were placed in quartz glass cuvette under stirring and

145

photoirradiated by six 4 W black-light-blue lamps housed in an air-cooled irradiation chamber. Initial pH

146

values of the suspensions were measured in all cases, but not adjusted, and ranged from 7.0-7.5.

147

Photocatalysis experiments used aqueous suspensions of 2.8 g/L of catalyst or control materials

148

and were magnetically stirred in the dark for 12 h (dyes) or 2 h (phenol) to establish an adsorption

149

equilibrium. During photoirradiation, aliquots were withdrawn periodically, centrifuged to remove the

150

solid catalysts and analyzed immediately by UV/vis or HPLC. Initial concentrations of 0.1 mM for dyes

151

and 0.13 mM for phenol were used. Control experiments were performed using undoped YAG/TiO2 and

152

YSO/TiO2, TiO2 alone, MB-only, and dark conditions. To study the effect of ·OH scavenging on MB

153

degradation efficiency, isopropanol (18 mM) was added to suspensions prior to stirring and

154

photoirradiation, and degradation kinetics were similarly monitored.

155 156

RESULTS AND DISCUSSION

157

Phosphor Characterization. According to XRD results (Figure S2), the YAG:Er3+ powder was single

158

phase garnet structure (Ia3̄d).29 Also based on the present data, “YAP:Er3+” powders synthesized

159

according to Dong et al.28 contained a mixture of perovskite and garnet phases. The perovskite phase is

160

known to be thermodynamically unstable and single-phase powders are synthetically challenging.30, 31

161

Diffraction patterns for YAP:Er3+ samples reported by Dong et al., as well as Wang et al., also showed

162

impurity peaks that included YAG.27, 28 As expected, this mixed-phase material resulted in much weaker

163

UC emission than single-phase YAG:Er3+, shown in Figure S3. Resultantly, YAG:Er3+ was deemed a

164

more legitimate Er3+-activated UC phosphor and used herein, despite YAP:Er3+ appearing more

7 ACS Paragon Plus Environment

Environmental Science & Technology

165

frequently in the UC-PC literature. The diffraction patterns of YSO:Pr3+ and undoped YSO matched our

166

previous reports (data not shown).26, 32 For UC emission spectroscopic analyses, 488 nm (“cyan”) was chosen for excitation, as this

167 168

wavelength coincides with intermediate state absorption maxima for both Pr3+- and Er3+-activated

169

systems.16, 33, 34 Figures 2A-B show the established mechanisms by which the two phosphors convert

170

visible light into UV via energy transfer UC,16, 35 and the UC emission spectra recorded herein are shown

171

in Figure 2C. Emission by YSO:Pr3+ was consistent with our earlier reports and results of other groups,

172

showing broad 255-360 nm emission assigned to the 4f5d3HJ/3FJ interconfigurational transitions.16, 26, 34

173

Emission by YAG:Er3+ showed sharper peaks characteristic of intra-4f transitions and qualitatively

174

consistent with previous analyses of Er3+ UC in single crystals.33, 36, 37 Most notably, the detected UV-

175

range UC peak of YAG:Er3+ at 319 nm (2P3/24I15/2)33 was much weaker than emission by the Pr3+-

176

activated phosphor. In fact, with respect to integrated peak area, the cyan-to-UV conversion efficiency of

177

YAG:Er3+ was 3 orders of magnitude lower than YSO:Pr3+.

178

Based on the energy levels of the respective activator ions, this finding is not surprising.

179

Upconversion by YSO:Pr3+ is known to be relatively efficient due to: (1) an energetically wide cluster of

180

3

181

the 3PJ and 1D2 states resulting in negligible nonradiative multiphonon relaxation from the intermediate

182

state;39 (3) action of both [3PJ,3PJ][4f5d,3H4] and [3PJ,1D2][4f5d,3H4] energy transfer UC mechanisms,

183

the latter of which involves the 1D2 singlet state which has an especially long lifetime of 27 µs (see ref. 16

184

for additional details); and lastly (4), the vast majority of emission from the doubly excited 4f5d manifold

185

is in the form of high-energy UV photons, rather than cascade-type visible light emission, owing to its

186

parity-allowed nature. In contrast, the 4f electronic structure of Er3+ results in the crowded and evenly-

187

spaced array of energy levels seen in Figure 2B, which are prone to multiphonon relaxation. Furthermore,

188

even the Er3+ ions excited to the 2P3/2 emitting level which relax radiatively do so primarily in the form of

189

>400 nm photons, with the UV emission peaks being much weaker by comparison.33, 36, 37 These data

PJ intermediate states resulting in relatively strong cyan/blue light absorption;34, 38 (2) a wide gap between

8 ACS Paragon Plus Environment

Page 8 of 22

Page 9 of 22

Environmental Science & Technology

190

show the first direct comparison of YSO:Pr3+ emission to Er3+-activated systems, and it is clear that the

191

former has far greater UV conversion efficiency.

192 193

Figure 2. (A) Mechanisms of visible-to-UVC conversion by YSO:Pr3+.16, 17 (B) Primary mechanism of

194

visible-to-UV conversion by Er3+-doped phosphors.35 (C) Upconversion emission spectra of phosphors

195

under 488 nm laser excitation (140 mW); inset shows YAG:Er3+ and undoped YAG spectra with adjusted

196

axes. The peaks seen at >430 nm are due to leakage of the excitation light from the monochromator.

197

To investigate UC-PC, composites were synthesized by sintering the phosphor powders together

198

with the well-known polymorphic commercial P25 TiO2 nanoparticles. This method was used by Wang et

199

al. with pure anatase TiO2 27 and we also believe it is theoretically optimal for UC-PC activity. As seen in

200

in the SEM micrographs in Figure 3, it yields a porous TiO2 particulate layer, bound to the phosphor

201

particles, that is likely conducive to contaminant and O2 diffusion to photoactivated catalyst sites. In

202

Figure S4, the UC emission spectrum of YSO:Pr3+ is compared to that of the YSO:Pr3+/TiO2 composite

203

material. No UV peak was observed for the composite, indicating that Pr3+ UC emission during focused

204

laser excitation is entirely absorbed by the surrounding semiconductor nanoparticles. To confirm that the

205

photocatalytic activity of the TiO2 was preserved upon sintering, MB degradation control experiments

206

were performed using YSO:Pr3+/TiO2 under UVA irradiation (365 nm) to directly excite the TiO2. Shown

207

in Figure S5, the photocatalyst component was confirmed to degrade MB effectively.

9 ACS Paragon Plus Environment

Environmental Science & Technology

208 209

Figure 3. SEM micrographs of powder phosphors materials and phosphor-TiO2 composites, including

210

(A) YSO:Pr3+, (B) YSO:Pr3+/TiO2, (C) YAG:Er3+, and (D) YAG:Er3+/TiO2.

211

Observation of UC-PC herein was first attempted using argon ion laser excitation of aqueous

212

YSO:Pr3+/TiO2 and YAG:Er3+/TiO2 suspensions at 488 nm. Such monochromatic excitation is ideal for

213

both Pr3+ and Er3+ visible-to-UV conversion, as portrayed in Figures 2A-B; however these conditions

214

resulted in no observable MB degradation as seen in Figure 4A. For reference, past UC-PC studies by

215

other authors typically used solar irradiation in their experiments,9, 20, 21 which presumably provided no

216

more than 100 mW/cm2 of total power density (AM 1.5 solar irradiance), only a small fraction of which

217

would resonate with the Ln3+ excitation transitions. The lack of any MB degradation under ~180 mW/cm2

218

488 nm laser excitation herein shows that neither phosphor composite material resulted in substantial

219

TiO2 activation via UC.

220

Given that most prior UC-PC studies used broad spectrum excitation, we similarly investigated

221

UC-PC activity of the suspensions irradiated by a W-LED equipped with a 450 nm long-pass filter. The

222

filter was used to reject 400-410 nm wavelengths which could potentially result in direct excitation of the

223

rutile portion of P25 TiO2. Degradation of MB is shown in Figure 4B, revealing that the composites do in

224

fact induce a mild catalytic effect with the W-LED illumination, while the TiO2-only control showed no 10 ACS Paragon Plus Environment

Page 10 of 22

Page 11 of 22

Environmental Science & Technology

225

photocatalytic enhancement over the MB photolysis results. However, we further tested composite

226

particles made using undoped YAG and YSO as controls (YAG/TiO2 and YSO/TiO2). Without the Ln3+

227

activator ions, these materials were not capable of UC and yet showed the same degree of MB

228

degradation enhancement as the doped phosphors (Figure 4B). It is clear then that catalytic action of the

229

composites under W-LED excitation is related solely to a photochemical phenomenon and not related to

230

UC emission. In addition to MB, we tested W-LED-induced degradation of two other commonly

231

employed dyes – MO and MR. Catalytic enhancements were observed for neither UC-capable Ln3+-doped

232

composites nor the undoped controls, as seen in Figure 4C. Catalytic degradation of phenol was also

233

assessed, in order to contrast the composites’ photoactivity against dyes to their activity towards an

234

organic species with no visible light absorption bands. The absorption spectra of phenol and MB are thus

235

provided in Figure S6. As seen in Figure 4D, phenol degradation was not observed for any of the

236

materials under W-LED irradiation, further refuting any generation of ·OH under these conditions.

237

11 ACS Paragon Plus Environment

Environmental Science & Technology

238

Figure 4. (A) Decoloration of MB by phosphor-TiO2 composites under 488 nm laser irradiation. (B) MB

239

decoloration by composites and controls under white LED irradiation. (C) Degradation of various dyes

240

by composite materials under white LED irradiation for 240 min; solid and striped bars depict Ln3+-

241

doped and undoped phosphor materials, respectively. (D) Degradation of phenol under white LED

242

irradiation. Error bars denote standard deviations of experiments performed in triplicate.

243

We hypothesized that the reason for catalytic MB degradation occurring under W-LED excitation

244

and not 488 nm excitation was a result of the former comprising wavelengths within the absorption range

245

of MB. Semiconductors such as TiO2 are well-known to catalyze dye degradation through an alternative

246

mechanism that does not involve band gap excitation to produce e–cb/h+vb pairs. Instead, upon absorption

247

of light by the dye (which may be of lower photon energy than the TiO2 Eg), the excited dye molecule

248

may transfer an electron into the TiO2 conduction band and become oxidized, in what is referred to as

249

photosensitized oxidation or direct electron injection.40-43 The process relies on the resultant e–cb on TiO2

250

becoming scavenged by O2 to produce O2·–, which may also contribute to dye decoloration.40-42 Since

251

formation of h+vb and ·OH production are not significantly involved in this mechanism, addition of an

252

·OH scavenger would not have any quenching effect on the degradation rates.

253

We performed photocatalysis experiments with YSO:Pr3+/TiO2 and isopropanol as an ·OH

254

scavenger 44 to probe the involvement of such a mechanism in the composite material suspensions. If UC

255

were in fact contributing to degradation by exciting TiO2 with emitted UV, then isopropanol addition

256

would scavenge ·OH and slow degradation kinetics. First we confirmed this quenching effect using direct

257

UVA excitation of the composites; the suppressed MB degradation kinetics resulting from 18 mM

258

isopropanol addition are seen in Figure 5. Under W-LED excitation, however, no change in kinetics was

259

observed with isopropanol addition. The result confirms that UV-related mechanisms did not occur when

260

visible light was used, excluding UC as a contributing factor.

261

12 ACS Paragon Plus Environment

Page 12 of 22

Page 13 of 22

Environmental Science & Technology

262 263

Figure 5. Effect of isopropanol ·OH scavenging on photocatalytic degradation of MB by YSO:Pr3+/TiO2

264

composites particles under W-LED (visible) and UVA irradiation. Error bars indicate standard

265

deviations of triplicate experiments.

266 267

Above we have shown that UV UC is not responsible for enhanced dye degradation rates by

268

phosphor/TiO2 composite materials; however, our experiments did indicate that the presence of the

269

YAG:Er3+, YSO:Pr3+, or undoped host components may result in statistically significant MB degradation

270

enhancements compared to TiO2 alone. This is enhancement is only observed when the system is excited

271

within the MB absorption range, and it does not require absorption by the Ln3+/dielectric component.

272

Control experiments using simple, unfused mixtures of phosphor and TiO2 particles resulted in no

273

enhancement (Figure S7), thus clearly indicating that the presence of epitaxial YSO/YAG–TiO2 interface

274

is required for the catalytic effect. Detailed explanation of this mechanism is beyond the scope of this

275

work; however, numerous existing reports explain that certain surface defects on TiO2 can enhance direct

276

electron injection. Improved degradation may arise due to increased adsorption of dye molecules onto the

277

catalyst or by introducing electron-trapping defects into the TiO2 surface structure.45-49 Incorporation of

278

Y3+ and the chemically similar La3+ into TiO2 as either dopants or as Y(La)2O3 nanoparticle

279

heterojunctions has specifically been shown to enhance charge transfer from dyes, likely by inducing

280

formation of oxygen vacancies.46-49 The catalytic enhancements by the composite materials studied here 13 ACS Paragon Plus Environment

Environmental Science & Technology

281

were most likely similarly related to the dielectric-semiconductor interface formed during the sintering

282

step, with defects and/or heterojunctions playing the key role rather than the bulk YSO or YAG

283

components. Degradation of MO or MR was not observed under the irradiation conditions tested,

284

possibly a result of poor adsorption due to neutral charge of these dyes, though the data do not rule out the

285

possibility of catalytic activity under more intense W-LED illumination.

286

In our opinion, the overwhelming majoring of UC-PC literature exhibits three critical

287

experimental flaws: first, the syntheses and quality of the UC materials are highly questionable – typically

288

using low purity stock chemicals and showing XRD patterns that indicate phase impurities (e.g., see ref.

289

21, Figure 1a and 1b therein). It is well known that optical properties –especially in UC materials– are

290

highly affected by certain transition metal impurities, necessitating the use of 99.99% purity or greater

291

stock chemicals.50 Secondly, in order to conclude that UC is responsible for the catalytic enhancements,

292

the use of the non-activated phosphor host material (e.g., undoped YAP or YAG) in control experiments

293

is required in order to confirm that optical – and not chemical – effects are responsible. Only the study by

294

Feng et al. included such a control, finding that their YAG:Er3+ sample enhanced MB degradation to a

295

greater extent than YAG under white fluorescent lamp excitation (~12% vs. ~8% degradation after 2 h),

296

though both performed better than TiO2-only (~5% degradation).10

297

Finally, UC-PC authors rarely include UC emission spectra of the materials in their publications,

298

which is certainly a minimum requirement for demonstrating UC capability prior to applying this

299

phenomena to environmental technology. Studies unrelated to photocatalysis have shown visible-to-UV

300

conversion by the same aforementioned Er3+-doped systems, though they used single crystals under pulse

301

laser excitation with sophisticated detection systems.35-37, 51 Furthermore, the UC spectrum of YSO:Pr3+

302

provided by Wu et al., as well as one in a more recent YSO:Pr3+ UC-PC paper, were measured with

303

fluorescence spectrometers equipped with a xenon lamps.22, 52 Based on our experience in characterizing

304

YSO:Pr3+, cyan or blue laser excitation in excess of 50 mW and phase-sensitive detection are required to

305

resolve the visible-to-UV anti-Stokes emission spectrum.13, 26 Their spectra instead appear to be normal 14 ACS Paragon Plus Environment

Page 14 of 22

Page 15 of 22

Environmental Science & Technology

306

Stokes emission that resulted from unintentional UV excitation of the sample by second order diffraction

307

from the source monochromator in the absence of appropriate long-pass filters. This same instrumental

308

blunder has been implicated in false reports of UC by carbon quantum dots,53 and illustrates the

309

importance of avoiding accidental UV excitation when attempting to measure upconverted visible light.

310

Overall, there is little convincing evidence that the powder UC materials used in past UC-PC experiments

311

can produce enough UV emission under incoherent visible light excitation to drive aqueous contaminant

312

degradation. As mentioned earlier, Cates et al. determined the blue-to-UV conversion of YSO:Pr3+ to be

313

on the order of 0.001%;13 thus, even if the full solar irradiance of 100 mW/cm2 could be converted by

314

such materials, the emitted UV intensity would amount to only 1 µW/cm2. Typical UV-driven advanced

315

oxidation processes, on the other hand, require intensities in the mW/cm2 range.54, 55 We emphasize that our conclusions apply only to alleged photocatalytic enhancement by Ln3+

316 317

visible-to-UV upconversion materials. Reports on other types of UC-PC do not consistently contain the

318

flaws described herein. Several studies, for example, have shown photocatalysis enhancement attributed

319

to infrared-to-UV conversion by Yb3+/Tm3+ systems, such as YF3:Yb3+,Tm3+ and β-NaYF4:Yb3+,Tm3+ .12,

320

56-58

321

is generally orders of magnitude more efficient than the UC processes central to this work,8 making the

322

approach much more feasible. Moreover, visible-to-visible UC by organic triplet-triplet annihilation

323

systems has also been shown to enhance the action of visible light-active photocatalysts through

324

upconversion of sub-band gap photons 11, 59-61 – again, this process is enabled by highly efficient

325

conversion which is easily detected spectroscopically.

326

ASSOCIATED CONTENT

327

Supporting Information. W-LED emission profile, XRD patterns, UC spectra of mixed-phase “YAP:Er3+”

328

and YSO:Pr3+/TiO2, UVA photocatalytic MB degradation data and degradation by physical mixtures,

329

absorption spectra of MB and Phenol.

Conversion of 980 nm infrared radiation to UVA wavelengths by Yb3+-sensitized fluoride phosphors

15 ACS Paragon Plus Environment

Environmental Science & Technology

330

AUTHOR INFORMATION

331

*Corresponding author. [email protected]. ph: 864-656-1540

332

The authors declare no competing financial interest.

333 REFERENCES

334 335

(1) Gaya, U. I.; Abdullah, A. H. Heterogeneous photocatalytic degradation of organic contaminants

336

over titanium dioxide: A review of fundamentals, progress and problems. J. Photochem. Photobiol. C.

337

2008, 9, (1), 1-12.

338 339 340 341 342

(2) Chong, M. N.; Jin, B.; Chow, C. W. K.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, (10), 2997-3027. (3) Burda, C.; Lou, Y.; Chen, X.; Samia, A. C. S.; Stout, J.; Gole, J. L. Enhanced nitrogen doping in TiO2 nanoparticles. Nano Lett. 2003, 3, (8), 1049-1051. (4) Chen, C.; Cai, W.; Long, M.; Zhou, B.; Wu, Y.; Wu, D.; Feng, Y. Synthesis of visible-light

343

responsive graphene oxide/TiO2 composites with p/n heterojunction. ACS Nano. 2010, 4, (11), 6425-

344

6432.

345 346 347 348 349 350 351

(5) Cho, Y.; Choi, W.; Lee, C.-H.; Hyeon, T.; Lee, H.-I. Visible light-induced degradation of carbon tetrachloride on dye-sensitized TiO2. Enviro. Sci. Technol. 2001, 35, (5), 966-970. (6) Kim, J.; Lee, C. W.; Choi, W. Platinized WO3 as an environmental photocatalyst that generates OH radicals under visible light. Enviro. Sci. Technol. 2010, 44, (17), 6849-6854. (7) Cates, E. L. Photocatalytic water treatment: So where are we going with this? Enviro. Sci. Technol. 2017, 51, (2), 757-758. (8) Cates, E. L.; Chinnapongse, S. L.; Kim, J.-H.; Kim, J.-H. Engineering light: Advances in

352

wavelength conversion materials for energy and environmental technologies. Enviro. Sci. Technol. 2012,

353

46, (22), 12316-12328.

16 ACS Paragon Plus Environment

Page 16 of 22

Page 17 of 22

Environmental Science & Technology

354

(9) Wang, J.; Li, J.; Liu, B.; Xie, Y.; Han, G.; Li, Y.; Zhang, L.; Zhang, X. Preparation of nano-sized

355

mixed crystal TiO2-coated Er3+:YAlO3 by sol–gel method for photocatalytic degradation of organic dyes

356

under visible light irradiation. Water Sc. Technol. 2009, 60, (4), 917-926.

357

(10) Feng, G.; Liu, S.; Xiu, Z.; Zhang, Y.; Yu, J.; Chen, Y.; Wang, P.; Yu, X. Visible light

358

photocatalytic activities of TiO2 nanocrystals doped with upconversion luminescence agent. J. Phys.

359

Chem. C. 2008, 112, (35), 13692-13699.

360 361 362

(11) Kim, J.-H.; Kim, J.-H. Encapsulated triplet–triplet annihilation-based upconversion in the aqueous phase for sub-band-gap semiconductor photocatalysis. J Am. Chem. Soc. 2012, 134, (42), 17478-17481. (12) Wang, W.; Huang, W.; Ni, Y.; Lu, C.; Xu, Z. Different upconversion properties of β-

363

NaYF4:Yb3+,Tm3+/Er3+ in affecting the near-infrared-driven photocatalytic activity of high-reactive TiO2.

364

ACS Appl. Mater. Inter. 2014, 6, (1), 340-348.

365 366 367 368

(13) Cates, E. L.; Cho, M.; Kim, J.-H. Converting visible light into UVC: Microbial inactivation by Pr3+-activated upconversion materials. Enviro. Sci. Technol. 2011, 45, (8), 3680-3686. (14) Cates, S. L.; Cates, E. L.; Cho, M.; Kim, J.-H. Synthesis and characterization of visible-to-UVC upconversion antimicrobial ceramics. Enviro. Sci. Technol. 2014, 48, (4), 2290-2297.

369

(15) Cates, E. L.; Kim, J.-H. Bench-scale evaluation of water disinfection by visible-to-UVC

370

upconversion under high-intensity irradiation. J. Photochem. Photobiol. B. 2015, 153, 405-411.

371

(16) Cates, E. L.; Wilkinson, A. P.; Kim, J.-H. Visible-to-UVC upconversion efficiency and

372 373 374

mechanisms of Lu7O6F9:Pr3+ and Y2SiO5:Pr3+ ceramics. J. Lumin. 2015, 160, (0), 202-209. (17) Cates, E. L.; Li, F. Balancing intermediate state decay rates for efficient Pr3+ visible-to-UVC upconversion: the case of β-Y2Si2O7:Pr3+. RSC Adv. 2016, 6, (27), 22791-22796.

375

(18) Wang, J.; Fu-yu, W. E. N.; Zhao-hong, Z.; Xiang-dong, Z.; Zhi-jun, P. A. N.; Lei, Z.; Lei, W.;

376

Liang, X. U.; Ping-li, K.; Peng, Z. Degradation of dyestuff wastewater using visible light in the presence

377

of a novel nano TiO2 catalyst doped with upconversion luminescence agent. J. Environ. Sci. 2005, 17, (5),

378

727-730.

17 ACS Paragon Plus Environment

Environmental Science & Technology

379

(190 Wang, J.; Zhang, G.; Zhang, Z.; Zhang, X.; Zhao, G.; Wen, F.; Pan, Z.; Li, Y.; Zhang, P.; Kang, P.

380

Investigation on photocatalytic degradation of ethyl violet dyestuff using visible light in the presence of

381

ordinary rutile TiO2 catalyst doped with upconversion luminescence agent. Water Res. 2006, 40, (11),

382

2143-2150.

383

(20) Wang, J.; Li, R.; Zhang, Z.; Sun, W.; Xu, R.; Xie, Y.; Xing, Z.; Zhang, X. Efficient photocatalytic

384

degradation of organic dyes over titanium dioxide coating upconversion luminescence agent under visible

385

and sunlight irradiation. Appl. Catal. A-Gen. 2008, 334, (1-2), 227-233.

386

(21) Wang, J.; Xie, Y.; Zhang, Z.; Li, J.; Chen, X.; Zhang, L.; Xu, R.; Zhang, X. Photocatalytic

387

degradation of organic dyes with Er3+:YAlO3/ZnO composite under solar light. Sol. Energ. Mat. Sol. C.

388

2009, 93, (3), 355-361.

389

(22) Wu, J.; Song, Y.; Han, B.; Wei, J.; Wei, Z.; Yang, Y. Synthesis and characterization of UV

390

upconversion material Y2SiO5:Pr3+, Li+/TiO2 with enhanced the photocatalytic properties under a xenon

391

lamp. RSC Adv. 2015, 5, (61), 49356-49362.

392

(23) Xu, J.; Brenner, T. J. K.; Chen, Z.; Neher, D.; Antonietti, M.; Shalom, M. Upconversion-agent

393

induced improvement of g-C3N4 photocatalyst under visible light. ACS Appl. Mater. Inter. 2014, 6, (19),

394

16481-16486.

395

(24) Zhou, D.; Xu, Z.; Dong, S.; Huo, M.; Dong, S.; Tian, X.; Cui, B.; Xiong, H.; Li, T.; Ma, D.

396

Intimate coupling of photocatalysis and biodegradation for degrading phenol using different light types:

397

Visible light vs UV light. Enviro. Sci. Technol. 2015. 49, (13), 7776-7783.

398

(25) Cates, E. L. Comment on “Intimate coupling of photocatalysis and biodegradation for degrading

399

phenol using different light types: Visible light vs UV light”. Enviro. Sci. Technol. 2015, 49, (21), 13075-

400

13076.

401 402

(26) Cates, E. L.; Wilkinson, A. P.; Kim, J.-H. Delineating mechanisms of upconversion enhancement by Li+ codoping in Y2SiO5:Pr3+. J. Phys. Chem. C. 2012, 116, (23), 12772-12778.

18 ACS Paragon Plus Environment

Page 18 of 22

Page 19 of 22

Environmental Science & Technology

403

(27) Wang, J.; Xie, Y.; Zhang, Z.; Li, J.; Li, C.; Zhang, L.; Xing, Z.; Xu, R.; Zhang, X. Photocatalytic

404

degradation of organic dyes by Er3+:YAlO3/TiO2 composite under solar light. Environ. Chem. Lett. 2010,

405

8, (1), 87-93.

406

(28) Dong, S.; Zhang, X.; He, F.; Dong, S.; Zhou, D.; Wang, B. Visible-light photocatalytic degradation

407

of methyl orange over spherical activated carbon-supported and Er3+:YAlO3-doped TiO2 in a fluidized

408

bed. J. Chem. Technol. Biot. 2015, 90, (5), 880-887.

409 410 411

(29) Dobrzycki, Ł.; Bulska, E.; Pawlak, D. A.; Frukacz, Z.; Woźniak, K. Structure of YAG crystals doped/substituted with erbium and ytterbium. Inorg. Chem. 2004, 43, (24), 7656-7664. (30) Premkumar, H. B.; Sunitha, D. V.; Nagabhushana, H.; Sharma, S. C.; Nagabhushana, B. M.; Rao,

412

J. L.; Gupta, K.; Chakradhar, R. P. S. YAlO3:Cr3+ nanophosphor: Synthesis, photoluminescence, EPR,

413

dosimetric studies. Spectrochim. Acta A. 2012, 96, 154-162.

414 415 416 417

(31) Li, X.; Li, J.-G.; Xiu, Z.; Huo, D.; Sun, X. Transparent Nd:YAG ceramics fabricated using nanosized γ-alumina and yttria powders. J. Am. Ceram. Soc. 2009, 92, (1), 241-244. (32) Cates, E. L.; Kim, J.-H. Upconversion under polychromatic excitation: Y2SiO5:Pr3+, Li+ converts violet, cyan, green, and yellow light into UVC. Opt. Mater. 2013, 35, (12), 2347-2351.

418

(33) Zu, N.; Yang, H.; Dai, Z. Different processes responsible for blue pumped, ultraviolet and violet

419

luminescence in high-concentrated Er3+:YAG and low-concentrated Er3+:YAP crystals. Physica B. 2008,

420

403, (1), 174-177.

421 422 423 424 425 426 427 428

(34) Hu, C.; Sun, C.; Li, J.; Li, Z.; Zhang, H.; Jiang, Z. Visible-to-ultraviolet upconversion in Pr3+:Y2SiO5 crystals. Chem. Phys. 2006, 325, (2-3), 563-566. (35) Xu, H.; Dai, Z.; Jiang, Z. Luminescence characteristics of ultraviolet upconversion from Er3+:YAG crystal by Ar+ laser (488 nm) excitation. Eur. Phys. J. D. 2001, 17, (1), 79-83. (36) Xu, H.; Jiang, Z. Dynamics of visible-to-ultraviolet upconversion in YAlO3:1% Er3+. Chem. Phys. 2003, 287, (1–2), 155-159. (37) Yang, H.-G.; Dai, Z.-W.; Sun, Z.-W. Ultraviolet and visible upconversion dynamics in Er3+:YAlO3 under 2H11/2 excitation. Chinese Phys. 2006, 15, (6), 1273. 19 ACS Paragon Plus Environment

Environmental Science & Technology

429 430 431

(38) Sun, C. L.; Li, J. F.; Hu, C. H.; Jiang, H. M.; Jiang, Z. K. Ultraviolet upconversion in Pr3+:Y2SiO5 crystal by Ar+ laser (488 nm) excitation. Eur. Phys. J. D. 2006, 39, (2), 303-306. (39) Malyukin, Y. V.; Masalov, A. A.; Zhmurin, P. N.; Znamenskii, N. V.; Petrenko, E. A.; Yukina, T.

432

G. Two mechanisms of 1D2 fluorescence quenching of Pr3+-doped Y2SiO5 crystal. Phys. Status Solidi B.

433

2003, 240, (3), 655-662.

434

(40) Nasr, C.; Vinodgopal, K.; Fisher, L.; Hotchandani, S.; Chattopadhyay, A. K.; Kamat, P. V.

435

Environmental photochemistry on semiconductor surfaces: Visible light induced degradation of a textile

436

diazo dye, naphthol blue black, on TiO2 nanoparticles. J. Phys. Chem. 1996, 100, (20), 8436-8442.

437 438 439 440 441 442

(41) Kuo, W. S.; Ho, P. H. Solar photocatalytic decolorization of methylene blue in water. Chemosphere. 2001, 45, (1), 77-83. (42) Rochkind, M.; Pasternak, S.; Paz, Y. Using dyes for evaluating photocatalytic properties: A critical review. Molecules. 2015, 20, (1), 88. (43) Konstantinou, I. K.; Albanis, T. A. TiO2-assisted photocatalytic degradation of azo dyes in aqueous solution: Kinetic and mechanistic investigations. Appl. Catal. B. 2004, 49, (1), 1-14.

443

(44) Minero, C.; Mariella, G.; Maurino, V.; Vione, D.; Pelizzetti, E. Photocatalytic transformation of

444

organic compounds in the presence of inorganic ions. 2. Competitive reactions of phenol and alcohols on

445

a titanium dioxide−fluoride system. Langmuir. 2000, 16, (23), 8964-8972.

446

(45) Wang, Q.; Chen, C.; Zhao, D.; Ma, W.; Zhao, J. Change of adsorption modes of dyes on

447

fluorinated TiO2 and its effect on photocatalytic degradation of dyes under visible irradiation. Langmuir.

448

2008, 24, (14), 7338-7345.

449

(46) Zhang, J.; Zhao, Z.; Wang, X.; Yu, T.; Guan, J.; Yu, Z.; Li, Z.; Zou, Z. Increasing the oxygen

450

vacancy density on the TiO2 surface by La-doping for dye-sensitized solar cells. J. Phys. Chem. C. 2010,

451

114, (43), 18396-18400.

452 453

(47) Choi, J.; Park, H.; Hoffmann, M. R. Effects of single metal-ion doping on the visible-light photoreactivity of TiO2. J. Phys. Chem. C. 2010, 114, (2), 783-792.

20 ACS Paragon Plus Environment

Page 20 of 22

Page 21 of 22

454

Environmental Science & Technology

(48) Himanshu, N.; Hailemichael, A.; Lebohang, M.; Madhavi, T.; Rao, T. K. G. Synthesis and

455

characterization of Y3+-doped TiO2 nanocomposites for photocatalytic applications. Nanotechnology.

456

2009, 20, (25), 255601.

457

(49) Liqiang, J.; Xiaojun, S.; Baifu, X.; Baiqi, W.; Weimin, C.; Honggang, F. The preparation and

458

characterization of La doped TiO2 nanoparticles and their photocatalytic activity. J. Solid State Chem.

459

2004, 177, (10), 3375-3382.

460 461 462 463

(50) Ropp, R. C. Luminescence and the Solid State. Second ed.; Elsevier: Amsterdam, The Netherlands, 2004. (51) Özen, G.; Forte, O.; Di Bartolo, B. Upconversion dynamics in Pr-doped YAlO3 and Y3Al5O12 laser crystals. Opt. Mater. 2005, 27, (11), 1664-1671.

464

(52) Gao, W.; Zhang, W.; Tian, B.; Zhen, W.; Wu, Y.; Zhang, X.; Lu, G. Visible light driven water

465

splitting over CaTiO3/Pr3+-Y2SiO5/RGO catalyst in reactor equipped artificial gill. Appl. Catal. B. 2018,

466

224, 553-562.

467 468 469 470

(53) Wen, X.; Yu, P.; Toh, Y.-R.; Ma, X.; Tang, J. On the upconversion fluorescence in carbon nanodots and graphene quantum dots. Chem. Comm. 2014, 50, (36), 4703-4706. (54) James, C. P.; Germain, E.; Judd, S. Micropollutant removal by advanced oxidation of microfiltered secondary effluent for water reuse. Sep. Purif. Technol. 2014, 127, (Supplement C), 77-83.

471

(55) Gerrity, D.; Ryu, H.; Crittenden, J.; Abbaszadegan, M. Photocatalytic inactivation of viruses using

472

titanium dioxide nanoparticles and low-pressure UV light. J. Environ. Sci. Heal. A. 2008, 43, (11), 1261-

473

1270.

474

(56) Li, T.; Liu, S.; Zhang, H.; Wang, E.; Song, L.; Wang, P. Ultraviolet upconversion luminescence in

475

Y2O3:Yb3+,Tm3+ nanocrystals and its application in photocatalysis. J. Mater. Sci. 2011, 46, (9), 2882-

476

2886.

477 478

(57) Qin, W.; Zhang, D.; Zhao, D.; Wang, L.; Zheng, K. Near-infrared photocatalysis based on YF3:Yb3+,Tm3+/TiO2 core/shell nanoparticles. Chem. Comm. 2010, 46, (13), 2304-2306.

21 ACS Paragon Plus Environment

Environmental Science & Technology

479

(58) Hou, Z.; Zhang, Y.; Deng, K.; Chen, Y.; Li, X.; Deng, X.; Cheng, Z.; Lian, H.; Li, C.; Lin, J. UV-

480

emitting upconversion-based TiO2 photosensitizing nanoplatform: Near-infrared light mediated in vivo

481

photodynamic therapy via mitochondria-involved apoptosis pathway. ACS Nano. 2015, 9, (3), 2584-2599.

482

(59) Kwon, O. S.; Kim, J.-H.; Cho, J. K.; Kim, J.-H. Triplet–triplet annihilation upconversion in CdS-

483

decorated SiO2 nanocapsules for sub-bandgap photocatalysis. ACS Appl. Mater. Inter. 2015, 7, (1), 318-

484

325.

485

(60) Kim, H.-i.; Kwon, O. S.; Kim, S.; Choi, W.; Kim, J.-H. Harnessing low energy photons (635 nm)

486

for the production of H2O2 using upconversion nanohybrid photocatalysts. Energ. Environ. Sci. 2016, 9,

487

(3), 1063-1073.

488

(61) Kim, H.-i.; Weon, S.; Kang, H.; Hagstrom, A. L.; Kwon, O. S.; Lee, Y.-S.; Choi, W.; Kim, J.-H.

489

Plasmon-enhanced sub-bandgap photocatalysis via triplet–triplet annihilation upconversion for volatile

490

organic compound degradation. Enviro. Sci. Technol. 2016, 50, (20), 11184-11192.

491

22 ACS Paragon Plus Environment

Page 22 of 22