High-Capacity and Photoregenerable Composite ... - ACS Publications

Sep 14, 2016 - AC), for highly efficient adsorption and photodegradation of a representative ... TNTs@AC offered a maximum Langmuir adsorption capacit...
0 downloads 0 Views 4MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

A high-capacity and photo-regenerable composite material for efficient adsorption and degradation of phenanthrene in water Wen Liu, Zhengqing Cai, Xiao Zhao, Ting Wang, Fan Li, and Dongye Zhao Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 14 Sep 2016 Downloaded from http://pubs.acs.org on September 14, 2016

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 free 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 accessible to all readers and 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 37

Environmental Science & Technology

1

A high-capacity and photo-regenerable composite material for

2

efficient adsorption and degradation of phenanthrene in water

3

Wen Liu,† Zhengqing Cai,† Xiao Zhao,† Ting Wang,‡ Fan Li,† Dongye Zhao†,*

4



5

University, Auburn, AL 36849, USA

6



7

Department of Environmental Engineering, Peking University, Beijing 100871, China

Environmental Engineering Program, Department of Civil Engineering, Auburn

The Key Laboratory of Water and Sediment Sciences, Ministry of Education,

8 9 10 11 12 13 14 15 16 17 18 19 20 21 22

* Corresponding author

23

D. Zhao, Tel: +01-334-844-6277, Fax: +01-334 844 6290,

24

E-mail: [email protected] 1

ACS Paragon Plus Environment

Environmental Science & Technology

25

Abstract

26

We report a novel composite material, referred to as activated charcoal supported

27

titanate nanotubes (TNTs@AC), for highly efficient adsorption and photodegradation

28

of a representative polycyclic aromatic hydrocarbon (PAH), phenanthrene.

29

TNTs@AC was prepared through a one-step hydrothermal method, and is composed

30

of an activated charcoal core and a shell of carbon-coated titanate nanotubes.

31

TNTs@AC offered a maximum Langmuir adsorption capacity of 12.1 mg/g for

32

phenanthrene (a model PAH), which is ~11.5 times higher than the parent activated

33

charcoal. Phenanthrene was rapidly concentrated onto TNTs@AC, and subsequently

34

completely photodegraded under UV light within two hours. The photo-regenerated

35

TNTs@AC can then be reused for another adsorption-photodegradation cycle without

36

significant capacity/activity loss. TNTs@AC performed well over a wide range of pH,

37

ionic strength, and dissolved organic matter. Mechanistically, the enhanced adsorption

38

capacity is attributed to the formation of carbon-coated ink-bottle pores of the titanate

39

nanotubes, which are conducive to capillary condensation; in addition, the modified

40

micro-carbon facilitates transfer of excited electrons, thereby inhibiting recombination

41

of the electron-hole pairs, resulting in high photocatalytic activity. The combined high

42

adsorption capacity, photocatalytic activity and regenerability/reusability merit

43

TNTs@AC a very attractive material for concentrating and degrading a host of

44

micro-pollutants in the environment.

45

2

ACS Paragon Plus Environment

Page 2 of 37

Page 3 of 37

Environmental Science & Technology

46 47

1. Introduction Polycyclic aromatic hydrocarbons (PAHs) are well-known water pollutants for

48

their detrimental biological effects, toxicity, mutagenecity and carcinogenicity.1,

49

Phenanthrene is one of the most widely detected PAHs in the environment, especially

50

at oil or coal tar contaminated sites and printing or dyeing wastewaters.3,

51

Phenanthrene has been used as a model PAH in many environmental studies due to its

52

universality, persistency and toxicity.1, 5-7

53

2

4

The adsorption/desorption behaviors for both conventional (e.g., activated carbon)

54

and emerging (e.g., carbon nanotubes (CNTs)) materials have been widely studied

55

and documented.8-18 In recent years, many researchers have reported that carbon

56

nanomaterials (CNMs) can effectively adsorb PAHs.8-17 For instance, single-walled

57

and multi-walled CNTs were found to offer 2−4 orders of magnitudes higher

58

adsorption capacity for phenanthrene than fullerene.8 In terms of adsorption

59

mechanisms, Chen et al. proposed that the strong adsorption of nitroaromatics by

60

CNTs was due to π−π interactions between nitroaromatic molecules (electron

61

acceptors) and the highly polarizable graphene sheets (electron donors).16 While these

62

emerging materials may offer improved performances than conventional ACs, their

63

much higher material cost prohibits their practical applications. Moreover, the

64

potential nano-toxicity, and occupational and environmental health risks of CNMs

65

remain under active investigations.19-21

66

Both conventional and emerging carbonaceous adsorbents have been designed to

67

adsorb/concentrate hydrophobic contaminants without chemical transformation. Often

68

times, such adsorption-based technologies are puzzled by some key technical

69

obstacles, such as regeneration/reusability of spent adsorbents and treatment of spent

70

regenerant.22, 23 To facilitate “green” and cost-effective regeneration and reuse of the 3

ACS Paragon Plus Environment

Environmental Science & Technology

71

spent adsorbents, it has been desirable to develop composite materials that combine

72

high adsorption capacity and reactivity. For instance, catalysts-modified CNMs have

73

been shown to be able to adsorb and catalytically degrade organic compounds,23-25

74

and the chemical transformation also regenerates the material.

75

Photocatalytic degradation of persistent organic pollutants including PAHs has

76

picked strong momentum.26, 27 While innovative photocatalysts have been consistently

77

sought, TiO2 remains the most widely used photocatalyst for its low cost and sound

78

photo-activity.2, 7, 28, 29 Derived from TiO2, titanate nanotubes (TNTs) were prepared

79

through hydrothermal treatment,30 and have drawn immediate interests for their large

80

specific surface area, nano-scale structures, ion-exchange property and good

81

photoelectric response.31, 32 TNTs are not only excellent adsorbents for heavy metals,

82

but also promising photocatalysts.33-39 However, due to the hydrophilic nature, TNTs

83

can hardly adsorb hydrophobic hydrocarbons such as PAHs, which also severely

84

inhibits photocatalytic degradation of PAHs.28, 40

85

While carbon nanofiber or graphene have been used to enhance the

86

photocatalytic activity of TNTs,41,

42

87

AC-supported TNTs to our knowledge. Taking advantage of the high adsorption

88

capacity of AC and photocatalytic activity of TNTs, we conceived a novel composite

89

material, referred to as TNTs@AC, by depositing TNTs onto a common activated

90

charcoal. the newly synthesized TNTs@AC is expected to show the following

91

synergistic effects: 1) the high adsorption capacity of AC will concentrate the target

92

organic pollutants onto the surface of TNTs@AC, facilitating the subsequent

93

photocatalytic degradation; 2) the high photocatalytic activity of TNTs will facilitate

94

effective degradation of the adsorbed pollutants, which also regenerates the spent

95

TNTs@AC; 3) the hydrothermal treatment during the material synthesis may

there has been no information available on

4

ACS Paragon Plus Environment

Page 4 of 37

Page 5 of 37

Environmental Science & Technology

96

facilitate micro-AC coating on TNTs, and the AC-amended TNTs will enhance both

97

the adsorption capacity and kinetics; and 4) AC on TNTs may serve as electron

98

shuttles and prevent recombination of the excited holes and electrons, and thus

99

enhance the photodegradation efficacy.

100

As such, the overall goal of this study was to develop a novel bi-functional

101

material

that

102

photodegradation/regeneration for rapid and complete removal of PAHs (and possibly

103

other trace organic contaminants) in water. Specifically, using phenanthrene as a

104

model PAH, this works aimed to: (1) synthesize and characterize the desired

105

TNTs@AC; (2) test the adsorption kinetics and capacity of TNTs@AC; (3) examine

106

the effects of various water chemistry conditions on adsorption, including pH, ionic

107

strength

108

photo-degradation/regeneration efficiency and material reusability; (5) elucidate the

109

underlying mechanisms.

110

2. Materials and Methods

111

2.1. Chemicals

and

offers

natural

both

high

organic

adsorption

matters

capacity

(NOMs);

(4)

and

efficient

evaluate

the

112

All chemicals were of analytical grade or higher. Nano-TiO2 (P25, 80% anatase

113

and 20% rutile) was purchased from Degussa (Evonik) of Germany. Sodium

114

hydroxide (GR) and absolute ethanol were obtained from Acros Organics (Fair Lawn,

115

NJ, USA). A DARCO granular activated charcoal (20−40 mesh) was acquired from

116

Sigma-Aldrich (St. Louis, MO, USA) and used as received. The key considerations

117

for the choice of AC included: 1) moderate adsorption affinity toward the target

118

contaminants so that adsorbed contaminants are available for subsequent

119

photodegradation, 2) some carbon can be released to modify the TNTs during the

5

ACS Paragon Plus Environment

Environmental Science & Technology

120

hydrothermal treatment to facilitate photodegradation, and 3) relatively larger pore

121

size to avoid pore clogging. Phenanthrene (Table S1 in Supporting Information (SI))

122

was purchased from Alfa Aesar (Ward Hill, MA, USA), and a stock solution of 2 g/L

123

was prepared by dissolving phenanthrene in methanol. A standard leonardite humic

124

acid (LHA, IHSS 1S104H) containing 64% of total organic carbon (TOC) was chosen

125

as the model NOM.41

126

2.2. Synthesis and Characterization of TNTs@AC

127

TNTs@AC was synthesized through a one-step hydrothermal method based on

128

our previous studies on TNTs preparation.37, 38 Typically, 2.4 g AC and 1.2 g TiO2

129

were mixed with 66.7 mL of a 10 mol/L NaOH solution. After stirred for 12 h, the

130

mixture was transferred into a Teflon reactor with a stainless steel cover, and then

131

heated at 130 °C for 72 h. The black precipitate (TNTs@AC) was then separated and

132

washed with deionized (DI) water till pH ~7.5−8.5, and then dried at 80 °C for 4 h.

133

For comparison, plain TNTs were also prepared separately via the same procedure but

134

without AC,37, 38 and a sample of amended AC was also prepared by subjecting the

135

parent AC to the same hydrothermal treatment without TiO2.

136 137

Methods on material characterization are provided in Section S1 in SI. 2.3. Adsorption Kinetic and Isotherm Experiments

138

Phenanthrene adsorption kinetic and equilibrium experiments were carried out to

139

gauge the adsorption rate and capacity of TNTs@AC. Details on the experimental

140

methods are provided in Section S2 in SI.

141

2.4. Photo-Regeneration and Reuse of TNTs@AC

142

Regeneration of spent TNTs@AC was performed through photodegradation of

143

phenanthrene adsorbed on TNTs@AC. Upon adsorption equilibrium, the mixture was

6

ACS Paragon Plus Environment

Page 6 of 37

Page 7 of 37

Environmental Science & Technology

144

left still for 1 h to allow the spent TNTs@AC to settle by gravity (>99% of

145

TNTs@AC settled). Then, ~90% of the supernatant was removed, and the residual

146

solid-liquid mixture was transferred into a glass photo-reactor with a quartz cover.

147

The reactor was then placed under UV light (365 nm, 1.42 mW/cm2). After 60 minʼs

148

UV irradiation, the liquid was decanted and the solid was extracted for remaining

149

phenanthrene using 20 mL methanol at 80 °C for 4 h.43 The regenerated TNTs@AC

150

was then reused in another cycle to adsorb and degrade phenanthrene, and the

151

adsorption-regeneration cycles were repeated 5 times to probe reusability of the

152

material.

153

2.5. Chemical Analysis

154

Methods for chemical analysis are provided in Section S3 in SI.

155

3. Results and Discussion

156

3.1. Characterizations of TNTs@AC

157

Figure 1 presents scanning electron microscope (SEM) images of the parent AC

158

and TNTs@AC. While the surface of AC appeared bulky, flat and smooth (Figure 1a),

159

the surface of TNTs@AC appeared rather rough and full of clusters of aggregates

160

(Figure 1b). A close-up of the surface revealed that the tubular TNTs formed an

161

interweaved network spreading throughout the surface (Figure 1c). The length of the

162

nanotubes stretched up to hundreds of nanometers. The energy-dispersive X-ray

163

spectra (EDS) (Figures 1d and 1e) reveal four major elements C, O, Na and Ti on the

164

surface of TNTs@AC, indicating that TNTs were not just simply coated on AC,

165

rather the nanotubes are intermingled with AC, i.e., some AC is also coated on TNTs

166

(further confirmed below).

167

[Figure 1]

7

ACS Paragon Plus Environment

Environmental Science & Technology

168

Figure S1 displays the X-ray diffractometer (XRD) patterns of neat TNTs, AC

169

and TNTs@AC. For neat TNTs, the peak at 9.4°, 24.4°, 28.1°, 48.2° and 61.5° are all

170

assigned to sodium tri-titanate,34, 37, 44 with a basic structure of NaxH1-xTi3O7 (x =

171

0−0.75, depending on the remaining sodium).33, 34, 37 The tri-titanate is composed of

172

corrugated ribbons of triple edge-sharing [TiO6] as a skeletal structure and

173

H+/Na+ located in interlayers.33,

174

interlayer distance (crystal plan (020)) of TNTs.33, 34, 37 For AC, the two peaks at 26°

175

and 43° are attributed to the diffractions of crystal planes of graphite (002) and (100),

176

respectively.45-47 For TNTs@AC, all the peaks observed for TNTs remained, and in

177

addition, the graphite (002) peak was observed, confirming the SEM finding that AC

178

is covered by TNTs with some AC coated on the surface TNTs. The Si impurities

179

(quartz/cristoballite-SiO2) in the raw AC were removed in TNTs@AC upon the

180

hydrothermal-alkaline treatment and the subsequent washing process (Section 2.2).

34, 37

In addition, the peak at 9.4° represents the

181

Figure S2 displays X-ray photoelectron spectroscopy (XPS) spectra of AC and

182

TNTs@AC, and Table S2 lists the corresponding atomic compositions. 7.1% of Ti

183

and 1.7% of Na are detected for TNTs@AC. Based on the Na/Ti ratio and the general

184

molecular formula of NaxH1-xTi3O7 for TNTs,34, 37 the compositions of the synthetic

185

TNTs can be identified as Na0.7H1.3Ti3O7. Based on carbon content in AC (82.1%) and

186

Ti mass added, the overall mass ratio of AC to TNTs in the composite material is

187

~1.7:1. The high resolution spectra of C 1s appeared similar before and after the

188

hydrothermal treatment (Figure S2b), while the C atomic percent associated with the

189

π−π bond increased from 9.9% for AC to 13.0% for TNTs@AC (Table S3), indicating

190

that TNTs@AC may offer stronger adsorption of aromatic organic compounds

191

through π−π interactions.48 The high resolution spectra of O 1s (Figure S2c) reveal

192

that the lattice O increased from 22.6% for AC (due to inorganic oxide impurities 8

ACS Paragon Plus Environment

Page 8 of 37

Page 9 of 37

Environmental Science & Technology

193

such as SiO2) to 72.8% for TNTs@AC (due to [Ti−O6]) (Table S3),46, 48, 49 confirming

194

accumulation of TNTs on AC. The O peak at 532.3 eV in TNTs@AC is assigned to

195

Ti−O/C−O, which suggests formation of a linkage of C−O−Ti between TNTs and AC.

196

The Fourier transform infrared spectroscopy (FTIR) spectra (Figure S3 and Section

197

S4 in SI) also confirm the new peak at 1081 cm-1 that is assigned to C−O−Ti bond.47

198

In addition, the C−OH peak of AC at 1091 cm-1 was not only much lowered, but also

199

shifted to 1097 cm-1 in TNTs@AC,50, 51 indicating decreased carbon-oxygen groups.

200

The hydrothermal treatment and loading of TNTs lowered the measured BET

201

surface area from 566.1 m2/g for AC to 471.6 m2/g for TNTs@AC, and the pore

202

volume from 0.61 to 0.52 cm3/g (Table S4). Considering the compositions of

203

TNTs@AC (AC:TNTs mass ratio of 1.7: 1) and the specific surface area of TNTs

204

(272.3 m2/g),38 the measured BET surface area of TNTs@AC is very close to the

205

calculated value 470.6 m2/g, suggesting that the hydrothermal treatment did not

206

significantly alter the AC surface area. However, the measured pore volume is much

207

lower than the theoretical value of 0.85 cm3/g of TNTs@AC (calculated as the

208

weighted average of the mean pore volumes for neat AC and TNTs), which supports

209

the postulate that some micro-AC may have intruded into the pores of TNTs during

210

the hydrothermal treatment. TNTs@AC exhibits a bimodal pore size distribution

211

profile with a primary peaking at ~4 nm and a secondary peaking at 2−2.5 nm (Figure

212

S4b), which are attributed to the pores of AC and conversion of larger pores (>10 nm)

213

of TNTs into more micropores in TNTs@AC, respectively.52 The pore size

214

distribution (Figure S4b) also indicates that most of the larger pores (>10 nm) in AC

215

disappeared in TNTs@AC, suggesting that these macro-pores may also be blocked

216

due to the leached AC particles and/or growth of TNTs at the mouth of the

217

macro-pores. It is less likely for TNTs to intrude into the internal finer pores due to

9

ACS Paragon Plus Environment

Environmental Science & Technology

218

their interwoven tubular structure and size exclusion. The AC coating on the

219

nanotubes reduced the mean pore diameter of TNTs from 18.6 to 3.7 nm of

220

TNTs@AC. Figure S5 shows that zeta potential of TNTs@AC is much less negative

221

than neat TNTs, indicating that AC coating on TNTs shielded part of the functional

222

groups (−OH/−ONa) on TNTs.35, 37 The pHPZC values were measured to be 3.1 for

223

TNTs@AC, 2.6 for TNTs and 6.8 for AC (Table S4).38

224

3.2. Adsorption and Desorption of Phenanthrene

225

Figure 2a shows adsorption kinetics of phenanthrene by TNTs@AC. TNTs@AC

226

displayed rapid uptake rate. The adsorption equilibrium was reached in 180 min, with

227

a high removal efficiency of 96.8% at equilibrium, and most (>92%) of the adsorption

228

capacity was filled in the first 60 min. In contrast, the parent AC showed much slower

229

kinetics and lower phenanthrene capacity (74.9% removal at 600 min). Furthermore,

230

the hydrothermally treated AC showed only slightly enhanced kinetics and

231

equilibrium uptake compared to the original AC, and much lower capacity than

232

TNTs@AC. These observations indicate that the TNTs play an important role in

233

phenanthrene adsorption by providing more accessible sites and added adsorption

234

capacity. Table S5 shows that the pseudo-second-order model best-fits the

235

experimental kinetic data (R2 =1) for TNTs@AC, whereas the intraparticle diffusion

236

model performs worst (see Section S6 for the models), which differs from standard

237

AC where film or intraparticle diffusion often controls the adsorption rate,53

238

suggesting that the rate-controlling step for TNTs@AC is due to chemical

239

interactions.54

240

[Figure 2]

241

Figure 2b compares the adsorption isotherms of phenanthrene by TNTs@AC,

242

parent AC, treated AC, and neat TNTs in the low concentration range of