Impact of Pore Size on Fenton Oxidation of Methyl Orange Adsorbed

Mar 29, 2017 - The authors gratefully acknowledge funding from the Canada Research ... (K) values for OTC were more than 1 order of magnitude higher t...
0 downloads 0 Views 2MB Size
Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)

Article

Impact of pore size on Fenton oxidation of methyl orange adsorbed on magnetic carbon materials: Trade-off between capacity and regenerability Ye Xiao, and Josephine M. Hill Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b00089 • Publication Date (Web): 29 Mar 2017 Downloaded from http://pubs.acs.org on March 30, 2017

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 30

Environmental Science & Technology

1

Impact of pore size on Fenton oxidation of methyl orange adsorbed on magnetic

2

carbon materials: Trade-off between capacity and regenerability

3

Ye Xiao, Josephine M. Hill*

4 5 6

Department of Chemical & Petroleum Engineering, Schulich School of Engineering, University

7

of Calgary, 2500 University Dr NW, Calgary, AB, Canada, T2N 1N4

8 9 10

*Corresponding Author Tel: +1 403 210 9488; fax: 1 403 284 4852; e-mail: [email protected]

11

1 ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 30

12

Abstract: The economic clean-up of wastewater continues to be an active area of research. In

13

this study, the influence of pore size on regeneration by Fenton oxidation for carbon materials

14

with adsorbed methyl orange (MO) was investigated. More specifically three carbon supports,

15

with pore sizes ranging from mainly microporous to half microporous-half mesoporous to mainly

16

mesoporous, were impregnated with γ-Fe2O3 to make them magnetic and easy to separate from

17

solution. The carbon samples were characterized before adsorption and after regeneration with

18

hydrogen peroxide at 20 °C. In addition, adsorption kinetics and isotherms were collected, and

19

the Weber-Morris intraparticle diffusion model and Freundlich isotherm model fit to the data.

20

The adsorption capacity increased with increasing microporosity while the regeneration

21

efficiency increased with increasing mesoporosity. Further experiments with varying

22

regeneration and adsorption conditions suggested that the regeneration process may be diffusion

23

limited. The MO adsorbed in the micropores was strongly adsorbed and difficult to remove

24

unlike the MO adsorbed in the mesopores, which could be reacted under relatively mild

25

conditions. Thus, there was a trade-off between adsorption capacity and regeneration.

26 27 28

Keywords: regeneration; Fenton oxidation; carbon; methyl orange; pore size

2 ACS Paragon Plus Environment

Page 3 of 30

Environmental Science & Technology

29

1. Introduction

30

Adsorption technologies, especially those using carbon materials such as activated carbon (AC)

31

1

32

been widely investigated for the treatment of water contaminated by organic pollutants. The

33

adsorption process concentrates the organic pollutants on the carbon matrix, which can then be

34

disposed or regenerated. Because of the potential for the leaching of pollutants during the

35

disposal process and high price of activated carbon (>$1500/ton 5), it may be more

36

environmentally and economically beneficial to regenerate the used carbon materials. Thus, this

37

study focuses on improving the regeneration process by studying the relationship between pore

38

size and regeneration efficiency of three carbon-based adsorbents with different physical

39

structures.

40

The current commercial thermal regeneration processes require high temperatures (800-900 °C)

41

and suffer from loss of carbon, high energy intensity, and high cost 6-7. Desorption-based

42

technologies may alleviate these problems 7-9; the adsorbates are desorbed intact by manipulating

43

pH, solvent polarity, pressure, and temperature (lower than 300 °C). Many adsorbates, however,

44

will not desorb at the lower temperatures and so degradation-based advanced oxidation

45

technologies, of which Fenton oxidation is the most common, are used to mineralize the

46

adsorbates. Although some novel electrochemical advanced oxidation methods reported over

47

90% regeneration efficiencies 10-16, the regeneration efficiencies of Fenton oxidation (without

48

electrochemistry) are generally less than ~60% 17-21, with exceptions for the treatment of volatile

49

organic compounds 19, 22.

50

Some improvements in regeneration efficiency for the Fenton oxidation process have been

51

achieved by introducing UV light 18, increasing the Fe content 21, using an acid pretreatment 19,

, carbon nanotubes 2, three-dimensional graphene-based macrostructures 3, and biochar 4, have

3 ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 30

52

increasing temperature 19, and reducing particle size 20. Despite the improvements, the

53

regeneration efficiencies were still less than 70% after one cycle. The loss of adsorption capacity

54

was attributed to (i) damage of the carbon matrix by the strongly oxidizing hydroxyl radicals

55

(•OH), (ii) blockage of pores by the oxidation products, and (iii) incomplete removal of the

56

adsorbates 8. The first two reasons are inconsistent with the often observed stable adsorption

57

capacity after the first adsorption-regeneration cycle 23. That is, carbon deterioration and pore

58

blockage should continuously decrease the adsorption capacity with each subsequent

59

regeneration cycle. The third reason has been related to the slow diffusion of H2O2 within the

60

activated carbon 20. Most adsorbents have high (>1000 m2/g) surface areas to maximize

61

adsorption capacity, but these high surface areas can only be achieved with microporous

62

materials, which will have diffusion limitations. Indeed, the highest regeneration efficiencies (>

63

90%) were reported for mesoporous carbon materials 24-25 but a direct study of the role of pore

64

structure has not been done.

65

In this paper, we propose that the pore structure of the carbon materials play a critical role during

66

the Fenton oxidation regeneration process. To test this hypothesis, three samples with different

67

sized pores were cycled through adsorption and regeneration steps. The samples were powders

68

with particle sizes below 0.1 mm that were impregnated with iron oxide to facilitate separation

69

from solution. Methyl orange (MO) was chosen as the adsorbate because it is a typical organic

70

pollutant likely to be irreversibly adsorbed at the testing temperatures (20 °C), removing the

71

complication of desorption 26-28, and easily detected by ultraviolet/visible (UV/Vis)

72

spectrophotometry. Adsorption isotherms and kinetics were collected, the regeneration and

73

adsorption conditions varied, and the samples characterized at various stages throughout the

74

experiments.

4 ACS Paragon Plus Environment

Page 5 of 30

Environmental Science & Technology

75

2. Materials and methods

76

2.1 Materials

77

Mesoporous carbon (>99.95%, denoted as SMC) and activated charcoal (denoted as NORIT)

78

were purchased from Sigma-Aldrich (St Louis, MO, USA), while activated carbon Colorsorb G5

79

(denoted as G5) was obtained from Jacobi Carbon company (Columbus, OH, USA). These three

80

carbon samples were modified using ferrous sulfate (FeSO4•7H2O, ≥ 99.0%, Anachemia,

81

Montreal, QC, Canada), iron (III) chloride (FeCl3, 97%, Sigma-Aldrich), and sodium carbonate

82

(Na2CO3, ≥ 99.5%, EM Science, Gibbstown, NJ, USA). Methyl orange (MO, C14H14N3NaO3S,

83

ACS reagent grade, Ricca Chemical Company, Arlington, TX, USA) was the model organic

84

pollutant. Regeneration was done with a hydrogen peroxide solution (H2O2, 30 wt% in water,

85

Sigma-Aldrich) diluted to different concentrations.

86

2.2 Preparation and characterization of magnetic samples

87

The magnetic samples were prepared according to a method reported previously 29-30 with slight

88

modifications. First, ~0.19 g of FeCl3 and ~0.66 g of FeSO4•7H2O were dissolved in 20 mL of

89

water at 70 °C. The solution was maintained at this temperature with constant stirring while ~1.0

90

g of carbon (SMC, G5, or NORIT) was added to form a suspension, which was subsequently

91

stirred for 30 min, before the drop-wise addition of 60 mL of Na2CO3 solution (13.5 g/L). After 1

92

h, the suspension was cooled to room temperature and aged for 24 h. The suspension was filtered

93

and washed with deionized water. The solid on the filter paper was collected and dried at 50 °C

94

overnight before storing. Iron oxide without carbon addition was also prepared by the same

95

procedure. The solids were named as follows: Fe2O3, Fe2O3/SMC, Fe2O3/G5 and Fe2O3/NORIT

96

and each contained approximately 20 wt% Fe, which was a sufficient loading to be able to easily

97

magnetically separate them from solution. 5 ACS Paragon Plus Environment

Environmental Science & Technology

98

The crystal structures of the prepared samples were determined by an X-ray diffractometer

99

(Multiflex, Rigaku, USA) with a scan rate of 2 °/min at 40 kV and 30 mA from 10° to 90° 2-

Page 6 of 30

100

theta. The particle size of the deposited iron oxide was estimated with the Scherrer equation. The

101

physical properties of the samples were determined by N2 adsorption at -196 °C. The BET

102

surface area and total pore volume were determined from the N2 adsorption isotherms within

103

partial pressure ranges of 0.02-0.3 and 0.96-0.97, respectively. The pore size distributions and

104

micropore volumes were derived from the N2 adsorption isotherms using the two-dimensional

105

non-local density functional theory (2D-NLDFT) method. The metal loadings were determined

106

with thermogravimetric analysis (SDT Q600, TA Instruments-Waters LLC, USA) by heating the

107

samples (~ 6 mg) at 10 °C/min to 800 °C and holding at this temperature for 10 min in flowing

108

air (50 mL/min) while the mass was monitored.

109

2.3 Adsorption isotherms

110

Methyl orange (MO) adsorption isotherms were collected using batch adsorption with an

111

incubator shaker (VWR symphony 5000I, Henry Troemner LLC, USA). Several magnetic

112

samples (8 mg) were weighed and placed in separate glass vials. Then 20 mL of MO solutions

113

with different concentrations were added to the vials, which were then placed in the shaker and

114

shaken at 250 rpm and 25 °C for 24 h. After adsorption, the magnetic samples were separated

115

from the solution with a magnet, and the MO concentration in the solution determined by a UV-

116

Vis spectrophotometer (EVOLUTION 220, Thermo scientific, USA). The experimental errors

117

were less than 3.0% at a 95% confidence level based on duplicate experiments.

118

2.4 Adsorption and Fenton oxidation regeneration

6 ACS Paragon Plus Environment

Page 7 of 30

Environmental Science & Technology

119

In a typical experiment, the first step consisted of mixing 0.1 g of fresh magnetic sample and 100

120

mL of MO solution in a beaker, and stirring the resulting suspension at 20 °C on a stir plate. To

121

obtain the adsorption kinetics, 0.5 mL of solution was withdrawn at various times, and its MO

122

concentration determined with the UV/Vis spectrophotometer. After 24 h of stirring, the

123

suspension was separated with a magnet. The upper solution layer was decanted from the beaker

124

leaving the saturated sample behind. Then 40 mL of deionized water was added to the beaker,

125

and the pH of the suspension was adjusted to 3.0 by adding a dilute HNO3 solution. Next, 0.4 mL

126

of 30 wt% H2O2 solution, which was sufficient as indicated by the small changes in H2O2

127

concentration after regeneration (Figure S1), was added and the new suspension stirred at 20 °C

128

for 8 h. The suspension was then magnetically separated, and the upper layer decanted. The

129

remaining solid was the regenerated magnetic sample. The second cycle involved returning to

130

the first step but using the regenerated sample rather than fresh sample. After regeneration, the

131

prefix “R-” was added to the sample name (e.g. R-Fe2O3/SMC).

132

3. Results and discussion

133

3.1 Characterization of samples

134

The XRD patterns (Figure S2 in Supporting Information) of the prepared samples had six peaks

135

at 30.4°, 35.7°, 43.4°, 53.8°, 57.3°, and 62.9° 2-theta, corresponding to the (220), (311), (400),

136

(422), (511), and (440), respectively, planes of maghemite (γ-Fe2O3). Maghemite is magnetic

137

and the average particle sizes were 23 nm, 20 nm, and 18 nm on Fe2O3/SMC, Fe2O3/G5, and

138

Fe2O3/NORIT, respectively, as calculated by the Scherrer equation using the MDI Jade 5.0

139

software. The particle size was similar to that obtained by Oliveira et al. 29 for Fe2O3 supported

140

on Aldrich Darco G60 AC. The peak at 26.2° 2-theta in the XRD pattern of Fe2O3/SMC

7 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 30

141

corresponds to the (002) plane of graphite (#08-0415). The patterns of the samples supported on

142

G5 and NORIT have a peak at 21° 2-theta that may be associated with silica (#74-0201).

143

Table 1 contains the physical properties of the samples and their metal oxide loadings. The BET

144

surface areas, total pore volumes, and micropore volumes consistently increased in the following

145

order: Fe2O3/SMC < Fe2O3/G5 < Fe2O3/NORIT. After the addition of Fe2O3, the surface areas

146

and pore volumes decreased (except for the pore volume of the Fe2O3/SMC sample). The

147

majority of this decrease was a result of the change in composition of the samples – the metal

148

loadings were approximately 20 wt% and, thus, the specific surface areas were expected to

149

decrease to ~80% of the values for the bare carbon supports if the added Fe2O3 only increased

150

the mass and did not contribute to the porosity. The fact that the total pore volumes of the SMC-

151

supported samples were identical (0.22 cm3/g) suggests that the added Fe2O3 was porous.

152

Comparing the pore size distributions of Fe2O3/SMC and SMC (Figure S3), there were

153

additional pores in the size ranges of 10-20 nm and 30-40 nm for the magnetic sample. As the

154

Fe2O3 crystal sizes (18-23 nm) and metal loadings were similar (~20-22%) on all magnetic

155

samples (Table 1 and Figure S4), the decreased pore volumes on the G5- and NORIT-supported

156

samples may have partially been a result of pores being blocked by the added Fe2O3. Analysis of

157

the XRD peaks only provides an average crystal size and not the distribution. Figure 1 illustrates

158

the pore size differences of the magnetic samples. Fe2O3/SMC has mesopores with pores ranging

159

in diameter from 3-20 nm and 28-50 nm, while Fe2O3/NORIT has mainly micropores 1-2 nm in

160

diameter. Fe2O3/G5 had a mixture of micropores and mesopores up to ~20 nm. Both NORIT and

161

G5 had some pores of diameter 28-50 nm but considerably fewer than SMC.

162

The surface areas and pore volumes were lower after regeneration (again with the exception of

163

the pore volume of the SMC-supported sample, Table 1 and Figure 1). The decreases for the 8 ACS Paragon Plus Environment

Page 9 of 30

Environmental Science & Technology

164

Fe2O3/SMC sample were relatively small compared to the decreases for the other two samples.

165

Blockage of the smallest pores is visible in the pore size distributions in Figure 1 (see insets in

166

each graph). In addition, the measured ash content decreased after regeneration for the G5- and

167

NORIT-supported samples. These results are consistent with MO being trapped in the

168

micropores and increasing the combustible carbon content of the regenerated samples R-

169

Fe2O3/G5 and R-Fe2O3/NORIT (Figure S4).

170

3.2 Adsorption kinetics

171

The rate of MO adsorption was determined on the fresh and regenerated magnetic samples as

172

shown in Figure 2. Within ~30 min, 80% of the total MO adsorption capacity was reached on

173

the fresh Fe2O3/NORIT and Fe2O3/G5 samples, and these adsorption capacities were 362 mg/g

174

and 178 mg/g, respectively. After regeneration, 460 min and 230 min were required to reach

175

80% of the total capacities and these capacities were reduced to 30.7 mg/g and 64.0 mg/g,

176

respectively. In contrast, the adsorption on the Fe2O3/SMC sample was much faster, with almost

177

complete adsorption in 30 min on both the fresh and regenerated samples but the total capacity

178

was lower at 26.5 mg/g. These results confirmed that a mixing time of 1 day (1440 min) was

179

sufficient to reach equilibrium and obtain the adsorption isotherms.

180

The Weber-Morris intraparticle diffusion 31-32 model was applied to the data in Figure 2 and the

181

resulting model parameters are given in Table 2. The model fit was good in terms of correlation

182

coefficients (R2 >0.92). According to the Weber-Morris adsorption kinetics model, if the

183

adsorption process is only controlled by an intraparticle diffusion process, a straight line through

184

the origin will be obtained by plotting  versus  . 33. For MO adsorption on all samples, these

185

plots were not linear (Figure S5). When the collected data was divided into three steps, governed

186

by film diffusion, mesopore diffusion, and micropore diffusion, consistent with previous dye 9 ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 30

187

adsorption studies 32, 34-35, the fits were linear in each step (Table 2 and Figure S5). The

188

diffusion rate constants increased after regeneration for Fe2O3/SMC, while they decreased by 31-

189

74% for Fe2O3/G5 and 64-97% for Fe2O3/NORIT, consistent with the previously discussed data

190

(Table 1, Figures 1 and 2).

191

3.3 Adsorption isotherms

192

The adsorption isotherms for MO on the magnetic samples before and after regeneration are

193

shown in Figure 3 (note the change in the y-axis range between Figures 3a and b). Consistent

194

with the characterization data described above, the adsorption capacities of Fe2O3/NORIT and

195

Fe2O3/G5 reduced significantly after regeneration (Figure 3b). In contrast, the isotherms for

196

Fe2O3/SMC as prepared and after regeneration were essentially the same.

197

As used in many research publications 10, 17, 23, 36, the regeneration efficiency was calculated with

198

the following equation:

199

         % =  × 100%



(1)



200

where  is the adsorption capacity of fresh carbon (mg/g),  is the adsorption capacity of

201

regenerated carbon (mg/g) and the conditions for adsorption are kept constant. As seen in Figure

202

3, the isotherm shape can change after regeneration, which complicates the use of Eqn. (1) and

203

may lead to an overestimation of the regeneration efficiency. To avoid this issue, high initial MO

204

concentrations of 100 mg/L, 400 mg/L, and 600 mg/L were used with Fe2O3/SMC, Fe2O3/G5,

205

and Fe2O3/NORIT, respectively, according to the isotherms (Figure 3) 36. Under these

206

adsorption conditions (Figure 2), the calculated regeneration efficiencies were 97%, 36%, and

207

8.5% for Fe2O3/SMC, Fe2O3/G5, and Fe2O3/NORIT, respectively. The regeneration efficiencies

10 ACS Paragon Plus Environment

Page 11 of 30

Environmental Science & Technology

208

decreased in the order of Fe2O3/SMC > Fe2O3/G5 > Fe2O3/NORIT - the reverse order of their

209

micropore volumes (Table 1).

210

To quantify the change in isotherm shape after regeneration, the Freundlich and Langmuir

211

models were applied to the data in Figure 3 and the resulting parameters are given in Table 2

212

and Table S1. The isotherm data was better fit by the Freundlich model according to the values

213

of R2, which were between 0.86 and 1.0 for the Freundlich model and between 0.71 and 0.94 for

214

the Langmuir model. The fits were significantly poorer for the R-Fe2O3/NORIT sample than for

215

the other samples. The values of the surface heterogeneity indicator (7.7 < n < 17, Table 2)

216

indicated the non-uniform nature of the adsorption sites on both fresh and regenerated magnetic

217

samples 37, which is common for adsorption on carbon substrates 38. After regeneration, the value

218

of n decreased for the Fe2O3/SMC and Fe2O3/G5 samples, and increased for the Fe2O3/NORIT

219

sample. Although the fit was poorer with the Langmuir model, the change in the separation

220

factor (RL) calculated for this model 37 was consistent with adsorption being less favourable after

221

regeneration for all samples but more so for the G5- and NORIT-supported samples than for the

222

SMC-supported samples. The similar adsorption capacities and isotherm parameters of SMC

223

samples before and after regeneration indicated that the Fenton oxidation process had little effect

224

on the carbon surface properties. There were no visible changes in sample morphology indicated

225

by SEM (Figure S6), which is consistent with the literature that states little or no reduction

226

adsorption capacity resulted from oxidation of the carbon surface 20, 39. Thus, the sites on which

227

adsorption was the strongest on the G5 and NORIT materials were not regenerated. Based on the

228

characterization data, these sites are located in the micropores in which the molecules would

229

have more interaction with the carbon substrate, and hence, a stronger adsorption.

230

3.4 Effect of adsorption and regeneration conditions 11 ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 30

231

The regeneration and adsorption conditions were varied to study their influence on the MO

232

capacity and regeneration efficiencies and the results are shown in Figure 4 (note the different y-

233

axis ranges in Figures 4a, b, and c). As the adsorption was fastest on the Fe2O3/SMC sample, the

234

regeneration time was reduced from 8 h to 4 h and to 1 h. As shown in Figure 4a, reducing the

235

time by a factor of two had a relatively small impact on the adsorption capacity while reducing

236

the time by a factor of eight had a much larger impact – capacities of 25.8 mg/g, 24.1 mg/g and

237

13.6 mg/g, respectively. To confirm that the impact of desorption was minor, a regeneration

238

experiment with only deionized water (no H2O2) was performed (Figure S7). After desorption,

239

less than 5% of adsorbed MO was desorbed into the water for all the carbon samples.

240

The regeneration time, initial H2O2 concentration, and volume of H2O2 solution (reported as a

241

ratio of H2O2 to MO) were increased separately for Fe2O3/G5 regeneration and all these factors

242

improved the regeneration efficiencies as shown in Figure 4b. A similar increase in adsorption

243

capacity (56 mg/g to ~76 mg/g) was achieved by increasing the regeneration time from 8 to 16 h

244

or the initial H2O2 concentration from 0.3 to 1.5 wt.%. Increasing the amount of available H2O2

245

(by increasing the volume of solution while maintaining the initial H2O2 concentration) resulted

246

in a smaller increase in adsorption capacity - 55.7 mg/g to 64.0 mg/g. In addition, the influence

247

of shaking speed on the regeneration efficiency was investigated (Figure S8). There were no

248

differences in the adsorption capacities of regenerated Fe2O3/G5 with (250 rpm) or without

249

shaking suggesting that the Fenton oxidation regeneration of MO saturated carbon was not

250

controlled by diffusion in the bulk solution.

251

Finally, the MO concentration was varied with the Fe2O3/NORIT sample, which had the highest

252

initial (i.e., before regeneration) adsorption capacity. Increasing the initial MO concentrations

253

from 500 mg/L to 520 mg/L to 600 mg/L increased the adsorption capacities from 342 mg/g to 12 ACS Paragon Plus Environment

Page 13 of 30

Environmental Science & Technology

254

351 mg/g to 362 mg/g, respectively, but also decreased the adsorption capacities after

255

regeneration from 63.7 mg/g to 50.7 mg/g to 30.7 mg/g, respectively (Figure 4c). Note, the

256

regeneration conditions were varied within this set of experiments. A longer regeneration time

257

(20 h) was used for the second experiment (MO initial concentration of 520 mg/L), and a larger

258

volume of H2O2 solution (540 mL) was used for the third experiment (MO initial concentration

259

of 600 mg/L). The application of more rigorous regeneration conditions, which were effective

260

for Fe2O3/SMC and Fe2O3/G5, were not able to compensate for the additional MO adsorption on

261

the Fe2O3/NORIT sample. Extrapolating the adsorption isotherm for Fe2O3/NORIT (Figure 3) to

262

higher equilibrium concentrations, this sample should have additional adsorption capacities of

263

~40 mg/g, 39 mg/g, and 22 mg/g (i.e., equilibrium capacities at 440 mg/L, 470 mg/L, and 570

264

mg/L). Thus, even without regeneration, the sample would have adsorbed more MO on a second

265

exposure to the MO solutions, complicating the calculation of a regeneration efficiency and

266

possibly explaining some of the discrepancies in the literature.

267

3.5 Multiple adsorption-regeneration cycles

268

The Fe2O3/SMC and Fe2O3/G5 samples were tested for stability over multiple

269

adsorption/regeneration cycles and the results are shown in Figure 5. The adsorption capacity of

270

Fe2O3/SMC was 27 mg/g initially and then 24 mg/g and 23 mg/g in cycles 2 and 3, respectively

271

(Figure 5a). The regeneration between the third and fourth cycles was improved by leaving the

272

sample overnight after the 4 h regeneration and solution decanting. The solution remaining in the

273

pores continued the Fenton oxidation overnight removing more of the MO and resulting in

274

higher adsorption capacities of 32 mg/g after the fourth and fifth cycles. After the fifth cycle, the

275

original regeneration method was employed and the adsorption capacity in the sixth cycle was 25

276

mg/g, which corresponds to a regeneration efficiency of 93%. The Fe2O3/G5 sample, which had 13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 30

277

more micropores than the Fe2O3/SMC sample, had a more significant decrease in adsorption

278

capacity in the first two cycles from 97 mg/g to 78 mg/g, with a more gradual decrease in

279

capacity between cycles 3 (73 mg/g) and 6 (65 mg/g) (Figure 5b).

280

3.6 Relationship between regeneration efficiency and pore structure

281

To better understand the effect of pore structure on the Fenton oxidation process, regeneration

282

efficiency was plotted versus microporosity (%) and the Weber-Morris intraparticle diffusion

283

rate constants from Tables 1 and S2 (Figure 6). The regeneration efficiency was inversely

284

proportional to microporosity (Figure 6a) - the higher the fresh sample microporosity, the less

285

effective the regeneration. The relationship between regeneration efficiency and the parameters

286

of the Weber-Morris interparticle diffusion model were varied (Figures 6b, c, and d). A linear

287

relationship between regeneration efficiency and 

288

of 0.982), and there appeared to be a positive trend (R2 of 0.75) with 

289

no trend with 

290

strongly in the micropores and not being removed during the regeneration – an explanation that

291

is in line with the studies that report regeneration efficiencies of less than 60% for microporous

292

AC regenerated with Fenton oxidation 17-21, 40. As shown in Figure 1, the micropores were

293

blocked after regeneration.

294

Several studies state that electrostatic interactions govern the adsorption process for MO

295

especially when metals or metal oxides were involved 23, 41-42, but these interactions could not

296

explain the high MO adsorption capacity at high pH conditions when the adsorbent surface was

297

negatively charged 23, 43. Other mechanisms including hydrogen bonding, pore filling, and van

298

der Waals interactions were also proposed for MO adsorption 43-44. Dispersion forces could

!,%

!,#

(mesopore diffusion) was observed (R2 !,$ ,

(film diffusion) but

(micropore diffusion). These results are consistent with the MO binding

14 ACS Paragon Plus Environment

Page 15 of 30

Environmental Science & Technology

299

dominate the interaction as MO and the surface groups on AC are dissimilar 45. Due to the

300

anisotropy of dispersion force as well as repulsive steric forces, the MO molecule would adsorb

301

with the orientation of closest distance to the carbon surface 45. In addition, for solution

302

concentrations below ~300 mg/L, MO anions tend to generate dimers 46. Adsorption of a MO

303

dimer could result in complete blockage of a micropore of ~1.6 nm in diameter according to the

304

MO molecular size (1.61 nm × 0.61 nm × 0.52 nm) 44, and leave the benzenesulfonate (-SO3-) or

305

dimethylanilinium (-N+(CH3)2) side exposed. This orientation is illustrated in Figure 7.

306

In the Fenton oxidation process, no ferric ions were detected in the solutions after 8 h

307

regeneration of the samples, which was consistent with literature reports that only a small

308

amount of Fe (