Effective Recovery of Vanadium from Oil Refinery ... - ACS Publications

Feb 5, 2018 - Department of Chemical and Biomolecular Engineering, National University of Singapore, 4 Engineering Drive 4, 117585, Singapore. ‡. NU...
0 downloads 0 Views 4MB Size
Subscriber access provided by UNIVERSITY OF TOLEDO LIBRARIES

Article

Effective Recovery of Vanadium from Oil Refinery Waste into Vanadium-Based Metal–Organic Frameworks Guowu Zhan, Wei Cheng Ng, Wenlin Yvonne Lin, Shin Nuo Koh, and Chi-Hwa Wang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b04989 • Publication Date (Web): 05 Feb 2018 Downloaded from http://pubs.acs.org on February 9, 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

Effective Recovery of Vanadium from Oil Refinery

2

Waste into Vanadium-Based Metal–Organic

3

Frameworks

4

Guowu Zhan,† Wei Cheng Ng,‡ Wenlin Yvonne Lin,† Shin Nuo Koh,§ and Chi-Hwa Wang*†

5 †

6

Department of Chemical and Biomolecular Engineering, National University of Singapore, 4

7 8

Engineering Drive 4, 117585, Singapore ‡

NUS Environmental Research Institute, National University of Singapore, 1 Create Way, Create

9 10

Tower #15-02, 138602, Singapore §

Sembcorp Industries Ltd., 30 Hill Street #05-04, 179360, Singapore

11

Submitted to

12

Environmental Science & Technology

13

*Corresponding author, Tel.: +65-65165079, Fax: +65-67791936, E-mail address:

14

[email protected]

15

ACS Paragon Plus Environment

1

Environmental Science & Technology

16

Page 2 of 22

Graphical TOC Entry

17

18

ACS Paragon Plus Environment

2

Page 3 of 22

Environmental Science & Technology

19

ABSTRACT. Carbon black waste, an oil refinery waste, contains a high concentration of

20

vanadium (V) leftover from the processing of crude oil. For the sake of environmental

21

sustainability, it is therefore of interest to recover the vanadium as useful products instead of

22

disposing of it. In this work, V was recovered in the form of vanadium-based metal–organic

23

frameworks (V-MOFs) via a novel pathway by using the leaching solution of carbon black waste

24

instead of commercially available vanadium chemicals. Two different types of V-MOFs with

25

high levels of crystallinity and phase purity were fabricated in very high yields (>98%) based on

26

a coordination modulation method. The V-MOFs exhibited well-defined and controlled shapes

27

such as nanofibers (length: >10 µm) and nanorods (length: ~270 nm). Furthermore, the V-MOFs

28

showed high catalytic activities for the oxidation of benzyl alcohol to benzaldehyde, indicating

29

the strong potential of the waste-derived V-MOFs in catalysis applications. Overall, our work

30

offers a green synthesis pathway for the preparation of V-MOFs by using heavy metals of

31

industrial waste as the metal source.

32

KEYWORDS. Metal–organic frameworks, vanadium recovery, waste, catalysts, oil refinery.

33 34

■ Introduction

35

With the rapid growth of world population, the amount of resources consumed is increasing,

36

and inevitably so is the amount of industrial waste generated in order to meet the world’s

37

demand. These industrial wastes may contain components that still pose good market value but

38

not recovered due to technological or economic limitations.1 An economically feasible solution

39

for the recovery of metals from industrial wastes should therefore be developed as it is

40

considered an important aspect to move towards a sustainable society.2-4 For instance, vanadium

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 22

41

is present in crude oil, coal, oil shale and tar sands, and therefore is an unavoidable waste in the

42

oil refinery industry and oil-fired electrical power stations.5, 6According to the 2016 BP

43

Statistical Review of World Energy,7 Singapore is capable of processing 1.514 million barrels of

44

crude oil per day (1.6% share of the world’s total). In Singapore, the oil refinery industry

45

generates about 30 tons of carbon black waste (a type of petroleum coke) daily as an unavoidable

46

by-product from the cracking of crude oil. The carbon black waste was found to contain a high

47

concentration of V (>14000 ppm), which could be extracted by chemical leaching.8, 9 In 2013,

48

the global vanadium production from natural resources containing V (e.g., vanadium ore) was

49

roughly 151,000 tons, up 11% year on year. The expanding demand for vanadium has

50

encouraged researchers to recover the vanadium from the industrial wastes. Accordingly, the

51

carbon black waste can be a potential source of V metal, and we are interested in recovering and

52

converting the waste metal ions into vanadium based metal–organic frameworks (abbreviated as

53

V-MOFs).

54

Metal–organic frameworks (MOFs) or supramolecular coordination polymers with one-, two-,

55

or three-dimensional crystalline networks are a new class of functional self-assembly materials in

56

which the metal ions/clusters are bridged through diverse organic ligands.10 Generally, the V-

57

MOFs contain trivalent or tetravalent V as metal nodes.11, 12 MOFs have been reported to be

58

useful in a wide variety of applications such as CO2 adsorption,13-15 heterogeneous catalysts,16-18

59

cathode materials for lithium ion batteries,19 magnetic sensors,20 and membrane separation.21

60

The conventional approach of fabricating MOFs generally requires the use of expensive metal

61

ion precursors.22 In addition, the use of costly and non-reusable organic solvents further impedes

62

an economically feasible production of MOFs for industrial applications.23 For this

63

consideration, the proposed pathway of vanadium recovery into V-MOFs would be conducted in

ACS Paragon Plus Environment

4

Page 5 of 22

Environmental Science & Technology

64

aqueous solutions by directly using the leachate of carbon black waste as the solvent. This

65

strategy, besides being environmentally friendly, would also provide a feasible way to reduce the

66

cost of production by eliminating the need for organic solvents.

67

The coordination of multicarboxylic acids to metal ions is the most representative assembly

68

manner in building MOFs, in which the M−O−M serves as primary structural motif to form

69

extended networks.24-26 To implement this idea, in this work, the aqueous leaching solution of

70

carbon black waste was used to prepare V-MOFs, with multicarboxylic acids as the building

71

blocks (Figure 1a). Depending on the organic ligands employed, two different V-MOFs with

72

different crystal morphologies and orientations were synthesized, i.e., V-BDC product (one-

73

dimensional nanofiber) when using 1,4-benzenedicarboxylate (BDC2−) linker, and V-NDC

74

product (nanorod) when using 1,4-naphthalenedicarboxylate (NDC2−) linker. Additionally, we

75

demonstrated that the as-obtained V-MOFs are efficient heterogeneous catalysts for selective

76

benzyl alcohol oxidation in liquid-phase (see Figure 1b), which is an important oxidation

77

reaction from the viewpoint of establishing a green process for benzaldehyde production.

78

■ MATERIALS AND METHODS

79

Materials. Carbon black waste was collected from an oil refinery on Jurong Island in Singapore.

80

The

81

Benzenedicarboxylic acid (H2BDC, Aldrich, 98%), 1,4-Naphthalenedicarboxylic acid (H2NDC,

82

Alfa Aesar, 98%), vanadium chloride (Aldrich, 97%), cetyltrimethylammonium bromide

83

(CTAB, Fluka, 96%), sodium dithionite (Aldrich, 85%), hydrochloric acid (VWR Chemicals,

84

32%), sodium hydroxide (Merck, 99%), benzyl alcohol (Alfa Aesar, 99%), tertbutyl

following

chemicals

were

used

as

received

without

further

purification:1,4-

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 22

85

hydroperoxide (TBHP, Aldrich, 70% in H2O), n-dodecane (Alfa Aesar, 99%), ethyl acetate

86

(Merck, 99.5%), and toluene (J.T.Baker, 99.5%). Deionized water was used in all experiments.

87

Carbon black waste leaching solution. The moisture content in the as-received carbon black

88

waste was about 45.7 wt% (see the TGA curve in Figure S1). So, the sample was first dried in an

89

electric oven at 150oC for 12 h. Then, 2.5 g of the dry solid was dispersed in 50 mL of NaOH (1

90

M) by sonication (150 W) for 1 h. Afterward, the leaching solution was collected by a membrane

91

filter (Nylon, pore size: 0.45 µm) under vacuum filtration condition.

92

Synthesis of V-MOFs. For the synthesis of V-BDC, 4 mL of the leaching solution was mixed

93

with 4 mL of HCl solution (1 M). Then, 1 mL of sodium dithionite aqueous solution (0.5 M) was

94

added. The color of the solution changed immediately upon the addition of the reducing agent.

95

After stirring for 2 min, 80 mg of H2BDC linkers and 0.1 g of CTAB were added. The mixture

96

was thoroughly stirred at room temperature for 10 min before hydrothermal treatment in a

97

Teflon-lined steel autoclave (capacity: 25 mL) at 200oC for 15 h. The solid was collected by

98

centrifugation and washing for two times with water/ethanol to remove sulfurous impurity. As

99

for the synthesis of V-NDC, 80 mg of H2NDC was used following the same approach. Note that

100

V-MOFs prepared without adding CTAB were labeled as V-BDC-i and V-NDC-i, respectively (i

101

refers to the irregular shape of the products).

102

Catalytic performance evaluations. Firstly, V-MOFs were activated in static air at 280oC for

103

10 h before the tests. Afterwards, the benzyl alcohol oxidation reaction was carried out in a

104

magnetically stirred flask (capacity: 50 mL). Typically, the initial reaction mixture was

105

composed of 1 mmol of benzyl alcohol, 1 mmol of n-dodecane (as an internal standard), 2.5

106

mmol of TBHP, 15 mL of toluene, and 30 mg of V-MOF catalyst. At a specific time, 0.2 mL of

ACS Paragon Plus Environment

6

Page 7 of 22

Environmental Science & Technology

107

the sample was extracted with 1 mL of ethyl acetate. Catalyst solids were separated from the

108

reaction mixture by syringe filters (PTFE, pore size: 0.45µm). The extracts were subsequently

109

analyzed by a gas chromatography (GC).

110

Characterization Techniques. Morphologies of V-MOF samples were characterized by

111

scanning electron microscopy (SEM, JSM-6700F) and transmission electron microscopy (TEM,

112

JEM-2010). The crystallographic information was identified by selected-area electron diffraction

113

(SAED, JEM-2100F) and X-ray diffraction (XRD, Bruker D8 Advance) equipped with a Cu Kα

114

radiation source. The elemental mapping was carried out by energy-dispersive X-ray

115

spectroscopy (EDX, Oxford Instruments, Model 7426). Brunauer–Emmett–Teller (BET) specific

116

surface areas of the samples, were calculated using N2 physisorption isotherms at 77 K

117

(Quantachrome NOVA-3000 system). The chemical compositions of the samples were analyzed

118

by X-ray photoelectron spectroscopy (XPS, AXIS-HSi, Kratos Analytical). Thermogravimetric

119

analysis (TGA) studies were carried out on a thermobalance (TGA-2050, TA Instruments) with

120

flowing air atmosphere (flow rate: 50 mL/min) at a temperature ramping rate of 10°C/min. Metal

121

concentrations in leaching solutions were measured by inductively coupled plasma-atomic

122

emission spectrometry (ICP-OES, Optima 7300DV, Perkin Elmer).

123

resonance (NMR) solution spectra were recorded on a Bruker AV500 (500MHz) spectrometer.

124

■ RESULTS AND DISCUSSION

125

Synthesis of V-MOFs. The carbon black waste contains a significant amount of heavy metals,

126

such as vanadium (~14500 ppm), nickel (~3200 ppm), and iron (~1500 ppm). Chemical leaching

127

with sodium hydroxide is an effective method to recover the vanadium ions from the carbon

128

black waste. The concentration of V in the obtained leaching solution was determined as 705

51

V nuclear magnetic

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 22

129

ppm, and only trace amounts of Ni and Fe metal ions were detected (see Table S1). In this work,

130

5 wt% of carbon black waste solid was used in the leaching process. Therefore, the leaching

131

efficiency of vanadium was about 97.2%. It is worth noting that the leachate obtained may be

132

further recycled through a cycling leaching process to increase the concentration of V in the final

133

leaching solution (see Figure 1). Since the vanadium ions in the carbon black waste leaching

134

solution were mainly in +5 oxidation state as monomeric anion VO43− (refer to 51V NMR data in

135

Figure S2), VV was first reduced to VIII species by using sodium dithionite as a reducing agent.

136

This approach was similar to a previous study on the reductive transformation of V2O5 powders

137

to vanadium-based MOFs.27 Sodium dithionite is a widely used reductant in industry. It was

138

chosen due to its high driving force for reducing VV to VIII ions.28 After only 1 second of

139

reaction, the vanadate ions could be totally reduced as visualized from the color change of the

140

leaching solution from yellow to gray. It should be noted that some sulfite and sulfate impurities

141

may be present in the original MOF product,29 but these impurities can be easily removed by

142

water washing treatment, as checked by thermogravimetric analysis (TGA) shown in Figure S3.

143

Based on Figure 2a, one can see that the amount of sodium dithionite added is crucial to the

144

transformative yield of vanadium. As a control, without the addition of sodium dithionite, no

145

MOFs was formed because VV cannot be used to build MOFs. It was found that 0.6 mL of

146

sodium dithionite (0.5 M) was the optimal value to achieve ~99% yield of V-BDC. Similarly, the

147

yield for V-NDC was around 99.4% under the same circumstances. As expected, there was no

148

solid product formed in the absence of organic linker even though the reducing agent was added.

149

This observation further confirms that the coordination reactions between leaching solution and

150

organic linkers produce MOFs in good yield. In addition, the production of V-MOFs was easily

151

scalable using the same set of synthesis parameters (see the result in Figure S4). However, it was

ACS Paragon Plus Environment

8

Page 9 of 22

Environmental Science & Technology

152

found that ascorbic acid was not an effective reducing agent for the transformation, as a

153

significant amount (99%) of vanadium ions remained in the solution.

154

Morphology control of V-MOFs. The morphology and size of the prepared V-BDC and V-

155

NDC products were observed by SEM and TEM, as displayed in Figure 3. Interestingly, the

156

derived V-BDC and V-NDC showed distinctly different shapes of nanofiber and nanorod,

157

respectively. For example, the length of V-BDC nanofiber was typically larger than tens of

158

micrometers with an average diameter of 40 nm. On the other hand, V-NDC exhibited short

159

nanorod shape with an average length of 270 nm and diameter of 64 nm. EDX elemental maps

160

further confirmed the chemical compositions of the prepared V-MOFs and also showed that V,

161

C, and O were evenly distributed across the entire structures (Figure 3d,h). Albeit no difference

162

in vanadium recovery rate, V-MOFs samples synthesized without CTAB (viz., V-BDC-i, and V-

163

NDC-i) had an irregular shape with a broad size distribution, as presented in Figure S5. It is well

164

known that CTAB, a capping agent, has frequently been used to control the morphology of MOF

165

crystals by adjusting growth rates of different facets via the electrostatic interactions between

166

CTA+ and the deprotonated carboxylic groups.30 In addition, it was also found that CTAB would

167

facilitate the dissolution of organic linkers in aqueous solution, probably due to their van der

168

Waals interactions. In control experiments, we also used VCl3 as precursors for preparing V-

169

MOFs. Indeed, the as-synthesized V-BDC and V-NDC displayed similar shapes of nanofibers

170

and nanorods, respectively (Figure S6). However, in the case of fabrications from VCl3, the

171

vanadium yields for V-BDC-i, V-BDC, V-NDC-i, and V-NDC were only 49.7%, 29.9%, 44.8%,

172

and 35.1%, respectively (Figure 2b). The enhanced yields by using leaching solutions in the

173

recovery process were probably due to two factors: (i) the lower pH value of the synthesis

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 22

174

solution offered a more suitable environment for recovery, and (ii) the reduction process

175

promoted the coordination between VIII ions and the organic ligands.

176

Crystal structures of V-MOFs. According to XRD pattern in Figure 2c, the designed V-BDC

177

is [VIII(OH)(BDC)](H2BDC)x, which is known as MIL-47as and the asterisk marks the Bragg

178

peaks belonging to the included guest H2BDC ligands. Upon activation in static air at 280oC for

179

10 h, MIL-47as can be converted to VIV(O)(BDC) (or known as MIL-47, one of the most studied

180

V-MOF materials) with high levels of crystallinity by removing the free organic linkers, along

181

with the oxidation of VIII ions to VIV ions and the conversion of VIII–OH bonds to vanadyl

182

(VIV=O) bonds. Interestingly, N2 physisorption measurements at 77 K revealed a drastic increase

183

in the BET surface area and pore volume as the filled linkers were removed upon activation

184

(SBET: 6.91 vs 408.3 m2 g-1; pore volume: 0.02 vs 0.21 cm3 g-1 referring to Figure S7). Besides

185

this, the orthorhombic crystal system of V-BDC was also confirmed from the SAED pattern

186

(Figure 4a), in which the series of spots can be attributed to (022), (101), and (123)

187

crystallographic planes with the respective d-spacing of 0.55 nm, 0.61 nm, and 0.35 nm.12 The

188

angle between the (022) and (101) spots is 71°, consistent with the theoretical value in the crystal

189

structure. The SAED pattern also indicates that the nanofibers observed are oriented almost

190

along the [11-1] zone axis with a single crystalline nature, and the reflections exhibit the

191

most intense diffraction spots. As also verified from XRD pattern (Figure 2d), the prepared V-

192

NDC is isostructural to [Al(OH)(NDC)].31 The presence of several peaks in the range (5 to 40o)

193

is due to the substitution of (AlIIIOH)2+ sites by (VIIIOH)2+ sites.27 Different from V-BDC, there

194

was no significant change in XRD patterns of V-NDC before and after activation treatments

195

(Figure S8), indicating that crystalline structure of V-NDC was guest-free and the larger H2NDC

196

molecules were unlikely possible to be captured within the pores. The BET surface area of V-

ACS Paragon Plus Environment

10

Page 11 of 22

Environmental Science & Technology

197

NDC after activation treatment was around 300 m2 g-1, comparable to the value of Al-NDC.32

198

Likewise, the tetragonal crystal system of V-NDC was confirmed from the SAED pattern,31 in

199

which the reflections of (200), (101), and (301) are clearly visible (Figure 4b). The surface angle

200

between the (200) and (101) facets is 73°, matching well with the crystal structure. Similar to

201

most other MOFs,33 the V-BDC and V-NDC frameworks would partially degrade under high

202

energy electron-beam irradiation, as the mentioned SAED diffraction spots were no longer

203

evident after 30 seconds of illumination (Figure S9). In both V-MOFs, small changes in MOF

204

crystal structures were found when CTAB surfactant was added to modulate the product shapes.

205

TGA and XPS Characterizations. The thermal stability of the as-prepared V-MOFs was

206

studied by TGA. Results shown in Figure 2e demonstrated that V-MOF structures were quite

207

stable up to 300°C in air atmosphere. There was no noticeable difference between the different

208

MOF crystals, and it was observed that the frameworks decomposed in the temperature range of

209

300–400 °C, which led to the formation of V2O5 as residue. The molar ratios of V metal ions to

210

organic linkers (either BDC2− or NDC2−) were close to 1: 1 in all the four MOFs, which matched

211

well with the formulas. Moreover, the reduction of vanadate in leaching solution by sodium

212

dithionite was further confirmed from XPS characterization of the V-MOFs. As revealed in

213

Figure 2f, all the V2p3/2 levels exhibited a binding energy of 515.8±0.2 eV, indicating the

214

presence of VIII rather than VV ions in the frameworks. The difference in binding energies (∆)

215

between the O1s and V2p3/2 levels was around 15 eV, which matched well with the reported

216

value (14.71 eV).20, 34 The XPS V2p3/2 levels of V-MOFs after activation treatments (280oC for

217

10 h) are shown in Figure S10, indicating that vanadium ions in the frameworks were oxidized to

218

VIV having a binding energy of 516.6 eV.35 The XPS peak deconvolution analysis indicates that

219

the ratios of VIV/(VIV+VIII) were 88.9% and 90.0% in V-BDC and V-NDC samples, respectively.

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 22

220

Catalytic performances of V-MOFs. Benzaldehyde is an important intermediate to many

221

organic compounds, and it can be produced from the catalytic oxidation of benzyl alcohol in

222

liquid-phase.36, 37 Herein, the catalytic performance of our prepared V-MOFs (V-BDC, V-BDC-i,

223

V-NDC, and V-NDC-i) was evaluated for this reaction. Firstly, a control study of the oxidative

224

reaction in the absence of V-MOFs showed only minor background reaction (