Nanocellulose from Spinifex as an Effective Adsorbent to Remove

Jan 23, 2018 - This approach represents a sustainable and cost-effective solution to tackle heavy metal contamination problems (such as cadmium) in dr...
1 downloads 8 Views 15MB Size
Subscriber access provided by READING UNIV

Article

Nanocellulose from Spinifex as an Effective Adsorbent to Remove Cadmium(II) from Water Priyanka R. Sharma, Aurnov Chattopadhyay, Sunil K Sharma, LiHong Geng, Nasim Amiralian, Darren J. Martin, and Benjamin S. Hsiao ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.7b03473 • Publication Date (Web): 23 Jan 2018 Downloaded from http://pubs.acs.org on January 24, 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 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.

ACS Sustainable Chemistry & Engineering 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 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

1

Nanocellulose from Spinifex as an Effective Adsorbent to Remove

2

Cadmium(II) from Water

3 4

Priyanka R. Sharma1, Aurnov Chattopadhyay2, Sunil K. Sharma1, Lihong Geng1,

5

Nasim Amiralian3, Darren Martin3, Benjamin S. Hsiao1*

6 1

7

Department of Chemistry, Stony Brook University, 100 Nicolls Road, Stony Brook, New York 11794-3400, United States

8 2

9 10 11

3

University High School, 4771 Campus Dr., Irvine, CA 92612 United States

Australian Institute for Bioengineering and Nanotechnology, Corner College and Cooper Rds, The University of Queensland, QLD 4072, Australia

12 13 14 15 16 17 18 19 20 21 22

* Corresponding author E-mail: [email protected]; Tel: +1(631)632-7793

23 1 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

24

Abstract

25

Nanocelluloses, in the form of carboxycellulose nanofibers, with low crystallinity (CI ~

26

50 %), high surface charge (-68 mV) and hydrophilicity (static contact angle 38°), were prepared

27

from an untreated (raw) Australian spinifex grass using the nitro-oxidation method employing

28

nitric acid and sodium nitrite. The resulting nanofibers (NOCNF) were found to be an effective

29

medium to remove Cd2+ ions (cadmium(II)) from water. For example, a low concentration of

30

NOCNF suspension (0.20 wt%) could remove Cd2+ ions over a large concentration range (50-

31

5,000 ppm) in a relatively short time period (≤ 5 minutes). The results showed that at low Cd2+

32

concentrations (below 500 ppm), the remediation mechanism was dominated by interactions

33

between carboxylate groups on the NOCNF surface and Cd2+ ions, which also acted as a

34

crosslinking agent to gel the NOCNF suspension. At high Cd2+ concentrations (above 1,000

35

ppm), the remediation mechanism was dominated by the mineralization process of forming

36

Cd(OH)2 nanocrystals, which was verified by TEM and WAXD. Based on the Langmuir

37

isotherm model, the maximum Cd2+ removal capacity of NOCNF was around 2,550 mg/g,

38

significantly higher than those of any adsorbents reported in the literature. NOCNF exhibited the

39

highest removal efficiency of 84%, when the Cd2+ concentration was 250 ppm. This study

40

demonstrated a simple pathway to convert underutilized biomass into valuable absorbent

41

nanomaterials that can effectively remove cadmium(II) ions from water.

42 43 44

Keywords Spinifex, carboxycellulose, nanofibers, cadmium(II) removal

45

2 ACS Paragon Plus Environment

Page 2 of 50

Page 3 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

46

ACS Sustainable Chemistry & Engineering

Introduction

47 48

Cadmium is one of the top ten chemicals posing a major public health problem to the

49

society mainly through water contamination.1 In specific, cadmium is a confirmed carcinogen,

50

mutagen and teratogen.2,3 As cadmium is a common component in many electronic devices and

51

other products, such as batteries, solar cells, paints and pigments,4 it can enter the water sources

52

through industrial waste and run-offs. When human is exposed to cadmium by consumption of

53

contaminated foods or water,5 cadmium can be adsorbed by the lungs or gastrointestinal tract and

54

transported in the bloodstream to other parts of body and accumulated in the liver and kidneys.

55

The biochemical pathways indicate that cadmium can bind to proteins, non-protein sulfhydryl

56

groups and various other macromolecules, such as metallothionein, in kidneys.6 According to a

57

report from the Agency for Toxic Substances and Disease Registry (ATSDR),5 the acute oral

58

toxicity dose of cadmium is in between 1,500 and 8,900 mg (i.e., 20 and 30 mg/kg) that would

59

lead to human fatalities. Although fatalities due to the cadmium exposure are rare, even a short-

60

term exposure is known to cause severe gastrointestinal irritation, resulting in vomiting,

61

abdominal pain and diarrhea.5 There have been several recent studies reporting the increasing

62

level of cadmium contaminant in some regions of Africa, Asia and South America.7-9 The

63

problem is not unique only in these regions, whereas similar challenges also exist in other parts

64

of the world. It is thus imperative to identify sustainable, cost-effective and environmental-

65

friendly solutions to deal with this challenge.

66 67

The goal of this study is to demonstrate that carboxycellulose nanofibers (CNF)

68

extracted from a raw biomass by the newly developed nitro-oxidation method can be used as an

3 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

69

effective adsorbent to remove cadmium(II) ions from water. This approach represents a

70

sustainable and cost-effective solution to tackle the heavy metal contamination problems (such as

71

cadmium) in drinking water. In previous studies, nanocelluloses prepared by acid hydrolysis and

72

oxidation methods (e.g. carboxymethylation, TEMPO) have been utilized as means to take out

73

heavy metal ions from water. For example, nanocrystalline cellulose (CNC) produced by acid

74

hydrolysis, incorporated with succinic acid, was used to remove lead(II) and cadmium(II) ions

75

from water;10 CNF containing carboxylate groups prepared by oxidation methods was also used

76

for chromium(III), nickle(II) and uranyl(II) ions removal.11,12 In addition, thiol modified CNF

77

was found to be effective for chromium(VI) and lead(II) removal,13 and amine-modified CNC

78

was capable of remove negatively charged chromium(VI) metal ions from water.14

79 80

Recently, Carpenter et al. reported the properties and life cycle assessment of different

81

nanocelluloses and compared them with carbon nanotubes (CNT)15. Thakur et al. also described

82

the applications of cellulosic derivatives for water purification and compared their effectiveness

83

with that of CNT.16 Both studies showed that CNF appeared to be a superior adsorption medium

84

over CNT for water purification, due to CNF’s higher functionality, lower cost, greater

85

sustainability and better safety. In addition, CNF has been shown to be a new and improved

86

barrier material for membrane applications, due to its good chemical resistance, hydrophilicity

87

(leading to low fouling tendency) and high porosity (leading to high flux). For example, in our

88

laboratory, thin-film nanofibrous composite (TFNC) membranes, containing multi-layered

89

fibrous scaffolds including a barrier layer made of TEMPO mediated CNF (diameter ~ 5 nm), an

90

electrospun polyacrylonitrile (PAN) nanofibrous mid-layer (diameter ~ 150 nm), and a non-

91

woven polyethylene terephthalate (PET) substrate (diameter~ 20 µm), were demonstrated. These

4 ACS Paragon Plus Environment

Page 4 of 50

Page 5 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

92

membranes exhibited a significantly higher flux (i.e., 2-10 X) for separation of oil/water

93

emulsions than commercial ultrafiltration (UF) membranes with similar rejection capability.17-20

94 95

There are many other naturally occurring sorbent materials that have also been

96

demonstrated to be able to remove small contaminants, such as heavy meal ions and dye

97

molecules, from water.21-27 For example, materials such as ZnO, AgO, TiO2 and alumina,

98

especially in the nanoscale form, have all been shown to be effective adsorption media for water

99

remediation.28-43 From the sustainability perspective, we are particularly interested in the use of

100

underutilized biomasses, such as agriculture waste, grass and weed, as resources to extract

101

valuable nanomaterial (CNF) for water purification because of their abundance and up-cycling

102

potential.

103 104

In this study, the chosen biomass is a raw (untreated) grass from Australia, named

105

spinifex. There are 69 species of spinifex grasses (genus Triodia) in the arid/semi-arid regions of

106

Australia, which covers over one quarter of the continent.44 These grasses are abundant but

107

underutilized. There are plenty of other similar arid grasses exist in Africa, Asia and South

108

America, where many of these regions are also experiencing the cadmium contamination

109

problems 7-9 due to the rapid development of electronic industry. Hence, the goal of this study is

110

to demonstrate that spinifex grass, as a model system for other arid grasses, is a good biomass

111

source to extract functional carboxycellulose nanofibers, whereby the extracted nanofibers could

112

be used as effective adsorbent media to remove cadmium ions from water.

113

5 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

114

It has been reported that typical spinifex grass has a high content of hemicellulose (~44

115

%), where this characteristics might lead to an easy nanofiber fibrillation tendency when treated

116

with mild alkali conditions followed by a mechanical process (e.g. high pressure

117

homogenization).45,46 The resulting cellulose nanofibers have been successfully demonstrated as

118

an ingredient for nanopaper with high toughness, and as an efficient reinforcement for latex

119

natural rubber and thermoplastic polyurethane nanocomposites.45,47,48 However, spinifex

120

nanofibers produced by using the above process (combined alkali and homogenization

121

treatments) were found to possess no or small surface charge, where their utility as an adsorbent

122

material for water purification is limited.

123 124

To extract CNF from untreated spinifex grasses, we employed the nitro-oxidation

125

method, involving the mixtures of nitric acid and sodium nitrite. This method has been

126

successfully used to prepare charged CNF (abbreviated as NOCNF hereafter) directly from

127

untreated (or raw) jute fibers,49. Nitro-oxidation is a simple approach that can oxidize biomass

128

and produce carboxylated nanocellulose with significant reduction in chemicals, water, and

129

energy consumption. In addition, the effluent from the nitro-oxidation treatment can be recycled

130

to nitrogen-rich fertilizer, thus offering great potential to advance the nexus of food, energy and

131

water. Recently, carboxylated cellulose nanocrystals with a maximum zeta potential of -45 mV,

132

have also been extracted using the hydrothermal treatment of MCC (microcrystalline cellulose)

133

in the presence of nitric acid and hydrochloric acid.50 We envision that the use of nitro-oxidation

134

method to extract functional nanocelluloses will become more popular due to its simplicity.

135

6 ACS Paragon Plus Environment

Page 6 of 50

Page 7 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

136

In the present work, the effectiveness of NOCNF extracting from spinifex to remove

137

heavy metal ions from water was demonstrated in the study of Cd2+ removal. The extracted

138

NOCNF was found to have medium crystallinity (CI = 53 %), high surface charge (-68 mV),

139

carboxylate content (0.86 mmol/g) and hydrophilicity (static contact angle = 38°). The

140

mechanism of Cd2+ removal by NOCNF was explored by using the combined Fourier transform

141

infra-red (FTIR) spectroscopy, scanning electron microscopy/ energy dispersive X-ray

142

spectroscopy (SEM/EDS), confocal microscopy and transmission electron microscopy (TEM)

143

techniques to characterize the floc coagulant formed upon the mixing of metal ion solutions at

144

different concentrations and a NOCNF suspension (0.2 wt%). The results indicated that at low

145

Cd2+ concentrations (below 500 ppm), the remediation process was dominated by the

146

interactions between Cd2+ ions and COO- groups on the NOCNF surface, where the adsorption

147

pathway prevails. In a way, Cd2+ ions also behaved as an effective crosslinking agent to gel

148

NOCNF in suspension. At high Cd2+ concentrations (above 1,000 ppm), the remediation

149

mechanism was dominated by the mineralization of cadmium hydroxide crystals within the

150

gelled NOCNF scaffold. The combination of these two processes greatly enhanced the

151

adsorption capacity of NOCNF.

152 153

The quantitative determination of the adsorption capacity of NOCNF against the Cd2+

154

ions was accomplished by the static adsorption study using the inductively coupled plasma mass

155

spectroscopy (ICP-MS) technique with different cadmium concentrations and under varying pH

156

levels. It was found that NOCNF possessed a maximum removal capacity (Qm) of 2550 mg/g,

157

which, to the best of our knowledge, is about 40% higher than the most effective adsorbent

158

reported in the literature.51 Finally, we demonstrated that the NOCNF-cadmium floc containing

7 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

159

mineralized cadmium hydroxide could be easily filtered using the low-cost gravity-driven

160

microfiltration or decantation methods. We truly believe that the simple nitro-oxidation approach

161

to extract functional carboxylated nanocelluloses from underutilized raw biomass, such as

162

spinifex grass, can provide a practical pathway to tackle the toxic metal ions (such as

163

cadmium(II)) contamination problems for drinking water purification.

164 165

Experimental

166 167

Materials

168 169

Spinifex grass was obtained from The University of Queensland, Australia. Nitric acid

170

(ACS reagent, 70 %), sodium nitrite (ACS reagent ≥ 97 %), sodium bicarbonate, cadmium

171

nitrate heptahydrate, sodium hydroxide, hydrochloric acid (36% assay) were purchased from

172

Fisher Scientific. The above chemicals were used without further purification. Sodium nitrite,

173

nitric acid, sodium bicarbonate and spinifex grass were used for the preparation of NOCNF,

174

where cadmium nitrate heptahydrate, sodium hydroxide and hydrochloric acid were used for the

175

cadmium remediation studies.

176 177

Preparation and Characterization of NOCNF

178 179

NOCNF were prepared from spinifex grass using the nitro-oxidation method, which has

180

been reported earlier by us.49 In brief, 2 g of water washed spinifex grass was placed in a three-

181

neck round bottom flask, where 10 mL of nitric acid (65-70 wt%, 0.239 mol) was slowly added

8 ACS Paragon Plus Environment

Page 8 of 50

Page 9 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

182

until the grass was completely soaked. Subsequently, 0.028 mol of sodium nitrite was added to

183

the mixture under a continuous stirring (using a magnet stirrer). Upon the addition of sodium

184

nitrite, red fumes were generated. To prevent the escape of these red fumes, the mouths of the

185

round bottom flask were immediately closed. This reaction was allowed to continue at 50°C for

186

12 h, and then the reaction was quenched by adding 250 mL of water. After that, the reaction

187

mixture was transferred into a beaker, and further quenched by adding 250 mL of distilled water.

188

The suspended product was allowed to settle down at the bottom of the beaker, where the upper

189

portion was decanted off to remove the excess acid. The above decantation procedure was

190

repeated 2-3 times until fibers started to suspend in water. In the next stage, the fiber suspension

191

was centrifuged at 3,000 xg for 10 min, until the pH of the supernatant reached above 2.5.

192

Subsequently, the bottom suspension was transferred to a dialysis bag (6-8 kDa) in deionized

193

water for 4-5 days until the conductivity of water reached below 5 µS. After dialysis, the

194

suspension was again centrifuged at 3,000 xg for 5 min, to separate microsized and nanosized

195

fibers. The pH level of the resulting NOCNF suspension was found to be 5.30, which might be

196

due to the presence of remnant (unconverted) carboxyl acid groups in the NOCNF. The yield of

197

carboxycellulose nanofibers obtained by the above procedure was 10 wt% and that of

198

carboxycellulose microfibers was 20 wt%. The carboxyl groups (COOH) on NOCNF were

199

converted to carboxylate groups (COONa) by treating the fiber suspensions with 8% sodium

200

bicarbonate solution at room temperature for 30 min.

201 202

The functional groups on the NOCNF surface were determined using Fourier transform

203

infrared (FTIR) spectroscopy (PerkinElmer Spectrum One instrument-ATR mode), 13C CPMAS

204

NMR (Bruker Utrashield 500WB plus), conductometric titration and Zeta probe analyzer

9 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

205

(Colloidal Dynamics) instruments. The morphology, polydispersity and crystal structure of

206

NOCNF were characterized by transmission electron microscopy (TEM, FEI Tecnai G2 Spirit

207

BioTWIN instrument), scanning electron microscopy (SEM, Zeiss LEO 1550 SFEG-SEM) with

208

energy dispersive X-ray spectroscopy (EDS) capability, and wide-angle X-ray diffraction

209

(WAXD, Benchtop Rigaku MiniFlex 600) techniques. The specific surface area of NOCNF was

210

measured by Novatouch LX2 (Quantachrome Instruments). The contact angle of NOCNF film

211

was measured using the Future Digital Scientific Corp. (FD) contact angle instrument (Model no.

212

OCA 15 EC). Descriptions of the above instruments and corresponding sample preparation

213

schemes for characterization are listed in Supporting Information.

214 215

Determination of Cadmium(II) Remediation Mechanism and Maximum Removal Capacity

216 217

The cadmium(II) remediation mechanism by NOCNF was determined by the following

218

experiments. Cd2+ solutions with concentrations ranging from 50 to 5,000 ppm were prepared,

219

which were subsequently mixed with a NOCNF suspension of fixed concentration. In specific, a

220

2 mL of 0.2 wt% NOCNF suspension was mixed with a 2 mL of Cd2+ solution at different

221

concentration. Upon mixing, a floc appeared and precipitated to the bottom of the vessel by

222

manual shaking. The non-flocculated portion was diluted by a factor of 1,000 to reach the

223

dilution level below 100 ppb to characterize the removal efficiency, where the flocculated

224

portion was separated by using a 0.1µ size microfiltration filter. The extracted floc sample,

225

containing aggregates of cadmium hydroxide and NOCNF, was characterized by TEM,

226

SEM/EDS, WAXD, and confocal microscopy (Leica TCS SP8X, also described in Supporting

227

Information).

10 ACS Paragon Plus Environment

Page 10 of 50

Page 11 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

228 229

The adsorption efficiency and adsorption capacity of NOCNF against cadmium(II) were

230

determined as follows. The non-floc portion, after dilution with a 2 wt% nitric acid solution to

231

the level below 100 ppb, while maintaining the same pH, was characterized by the inductively

232

coupled plasma mass spectroscopy (ICP-MS) technique (Supporting Information). The

233

adsorption efficiency for NOCNF was calculated based on the difference in Cd2+ concentration

234

before and after the mixing with the NOCNF suspension divided by the original Cd2+

235

concentration (Ce). This efficiency was a function of the Cd2+ concentration. To determine the

236

adsorption capacity of NOCNF, an ideal adsorption capacity was first calculated based on the

237

available mass (grams) of NOCNF in suspension and the available Cd2+ ions. The experimental

238

adsorption capacity (Qe) was the product of the adsorption efficiency and the ideal adsorption

239

capacity. As a result, the experimental adsorption capacity of NOCNF was also a function of the

240

Cd2+ concentration. The effect of pH value on the cadmium(II) adsorption efficiency by NOCNF

241

was further evaluated. In this study, a constant 10,000 ppm solution of cadmium(II) at different

242

pH values (3, 5, 7, 9, and 11) were prepared for the mixing with the 0.2 wt% NOCNF

243

suspension.

244 245

The maximum adsorption or removal capacity (Qm) of NOCNF against cadmium(II) was

246

determined using the Langmuir Isotherm model, which is based on a monolayer adsorption on

247

the active site of adsorbent and has the following expression:

248

஼௘

஼௘



= ொ௠ + ொ௠௕ ொ௘

Eq. 1

11 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

249

where Ce is the equilibrium (or original) concentration of cadmium(II) and Qe is the adsorption

250

capacity of cadmium(II) by NOCNF at equilibrium. The value of Qm (and b – the Langmuir

251

constant) was calculated from the slope of the linear plot based on Ce/Qe versus Ce.52

252 253

Results and Discussion

254 255

Characterization of NOCNF Extracted from Spinifex Using the Nitro-Oxidation Method

256 257

Figure 1(i) illustrates the FTIR spectra of untreated but water-washed spinifex grass and

258

extracted NOCNF using the nitro-oxidation method. In these spectra, the characteristic peaks of

259

cellulose at 3340 cm-1 due to the O-H stretching and 2900 cm-1 due to the CH and CH2 stretching

260

were seen. The prominent peak (highlighted in yellow) in NOCNF at 1581 cm-1 was indicative of

261

the carboxylate groups, which confirmed the conversion of -CH2OH at the C6 position in the

262

anhydroglucose unit to -COO-. Other peaks at 1372, 1150, 1100, and 1030 cm-1 were due to the

263

stretching and bending vibrations in the glycosidic bonds of cellulose. In the FTIR spectrum of

264

untreated spinifex grass, the peaks at 1739, 1515, 1460, 1240 and 810 cm-1 could be attributed to

265

the symmetrical streching of the C=C bond in aromatic groups of lignin, xylan and glucomannan

266

of hemicellulose.

267 268

It was interesting to note that the intensity of the peak assosiated to the C-H stretching

269

groups at 2900 cm-1 in NOCNF was found to be decreased, while the intensity of COONa peak

270

at 1581 cm-1 was notably increased, when compared to those in untreated spinifex grass. This

271

further confirmed the oxidation of anhydroglucose units at the C6 position by the nitro-oxidation

12 ACS Paragon Plus Environment

Page 12 of 50

Page 13 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

272

method. The significant decrease (or disappearance) in intensity at 1739, 1515, 1460, 1240, and

273

810 cm-1, also verified that the nitro-oxidation method effectively reduced or even removed the

274

hemicellulsoe and lignin content from the cell walls of spinifex grass. Table 1 summarizes some

275

physical and chemical properties of NOCNF extracted from spinifex using the nitro-oxidation

276

method. It was seen that the carboxylate content of spnifex-based NOCNF measured by the

277

conductometric tititration method was 0.86 mmol/g; the surface charge of NOCNF in suspension

278

measured by zetaprobe analyzer was -68 mV; and the specific surface area obtained for the

279

freeze dried NOCNF sample was 5.7 m2/g (this was comparatively lower than that of NOCNF

280

obtained from wood, algae and bacetrial cellulose using other extraction methods,52 which was

281

probably due to a lower degree of polymerization by using the nitro-oxidation method). In

282

addition, the static contact angle of the extracted NOCNF film was 38° (Figure S1 in Supporting

283

Information), indicating the hydrophillic nature of NOCNF.

284 285

Solid state

13

C CPMAS NMR spectra of untreated spinifex grass and the extracted

286

NOCNF are shown in Figure 1(ii). The spectrum of spinifex grass (Figure 1(ii)A) clearly

287

indicated the presence of cellulose, hemicellulose, and lignin components. For example, the

288

small peaks at 21.8 and 173 ppm could be attributed to the methyl and carboxyl carbons in the

289

acetyl groups of glucoroxylans in hemicellulose, whereas the small peak at 56.6 ppm was due to

290

the methoxyl (-OCH3) carbons in lignin.54 In this spectra (spinifex grass), the presence of

291

cellulose was also evident: the peak at 104.60 ppm was due to the C1 carbons, the peaks at 89.4

292

and 84.1 ppm (doublet) were due to crystalline and amorphous regions of the C4 carbons, and

293

the sharp doublet peaks with highest intensity in between 72.4-75.2 ppm, could be contributed to

294

the C2, C3 and C5 carbons. The doublet peaks at 65.2 and 63 ppm were due to the crystalline

13 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

295

and amorphous regions of the C6 carbons in the cellulose component of spinifex. However, in

296

extracted NOCNF, the

297

difference. For example, the disappearance of peaks at 153, 56.6 and 21.4 ppm, corresponding

298

to the C3, C4 carbons of syringyl units of lignin and methyl groups of glucoroxylans in

299

hemicellulose, confirmed the significant reduction/removal of lignin and hemicellulose content

300

in spinifex grass by the nitro-oxidation method. The intensity increases and broadening of the

301

carbonyl peak at 173 ppm, along with the reduction in the amorphous region of the C6 peak at

302

about 63 ppm, provided the evidence of oxidation of the cellulose moiety at the C6 position.

303

Interestingly, the crystalline peaks of the C6 carbons (65.2 ppm) and C4 carbons (89 ppm) all

304

became sharper in NOCNF, indicating that the oxidation mainly took place in the amorphous

305

region of cellulose. These analyses were consistent with the analysis of FTIR spectral results.

13

C CPMAS NMR spectrum (Figure 1(ii)B) exhibited some notable

306 307

WAXD measurements of spinifex grass and extracted NOCNF were also carried out and

308

the results are illustrated in Figure 2(i). These patterns indicate that both samples exhibited the

309

characteristic signature of a cellulose I structure, with three distinct diffraction peaks at 2θ angles

310

of 16.64° (spinifex)/16.86°(NOCNF), 22.45°(spinifex)/22.77°(NOCNF), and 34.80°, which

311

could be indexed as the (110), (200), and (004) refelctions.55,56 The crystallinity index (CI)

312

calculated using the “Segal equation” (details given in Supplementary Information) was about

313

53% for NOCNF, and about 50% for untreated spinifex grass. The CI of spinifex-based NOCNF

314

was found to be higher than that of jute-based NOCNF49 extracted by the nitro-oxidation

315

method, which was probably due to the different nature of the biomass.

316

14 ACS Paragon Plus Environment

Page 14 of 50

Page 15 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

317

A typical TEM image of NOCNF is shown in Figure 2(ii), which exhibited long filament-

318

like fibers entangled in a random fashion. Using a modified ImageJ software, it was possible to

319

estimate the average fiber length and fiber width. Based on the measurements from 20 individual

320

filaments, the average NOCNF fiber length was 190 ± 90 nm, the average fiber width was 4.0 ±

321

1.5 nm and the polydispersity index (PDI) was 0.35 ± 0.01 (the results also are listed in Table 1)

322

for the spinifex-based nanocellulose sample extracted by the nitro-oxidation method. The

323

average fiber width was in excellent agreement with that of NOCNF extracted from untreated

324

jute fibers, but the average length was shorter than that from jute.49

325 326

Explore the Cadmium(II) Removal Mechanism by NOCNF

327 328

The remediation mechanism of cadmium(II) by NOCNF was determined by

329

characterization of floc formed the mixing of cadmium(II) solutions at different concentration

330

with a 0.2 wt% NOCNF suspension. The visual examination that elucidates the interactions

331

between NOCNF and Cd2+ ions is shown Figure 3(i), where the left photo was a cadmium(II)

332

nitrate solution (Cd2+~5,000 ppm) that was clear and transparent. However, upon the addition of

333

5 mL of NOCNF suspension, a white precipitate or floc appeared at the bottom of the bottle in a

334

very short period of time (< 2 min) (the right photo in Figure 3(i)). The floc sample contained

335

aggregates of NOCNF and cadmium contaminant, which was analyzed carefully to determine the

336

remediation mechanism at different cadmium(II) concentration.

337 338

FTIR spectra of pure NOCNF and the floc obtained by mixing 500 ppm of cadmium(II)

339

nitrate solution and NOCNF suspension (0.2 wt%) are illustrated in Figure 3(ii). In Figure 3(ii)A,

15 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

340

the NOCNF spectrum depicted the main characteristic peaks at 3400 cm-1 from the hydroxyl (-

341

OH) stretching, 2883 cm-1 from the C-H symmetrical stretching, 1232 cm-1 from the C-OH

342

bending at the C6 position, 1204 cm-1 from the C-O-C symmetric stretching, and 1581 cm-1 from

343

the COO- stretching. However, the floc spectrum (Figure 3(ii)B) exhibited some notable

344

difference. For example, the COO- stretching peak was shifted to 1630 cm-1 versus the initial

345

1581 cm-1 in NOCNF, which is in agreement with the occurrence of interactions between COO-

346

and Cd2+ ions. As one Cd2+ ion can effectively interact with two COO- groups, cadmium(II) can

347

be considered as a good crosslinking agent for the charged NOCNF particles in water. The peak

348

at 3400 cm-1 in the floc spectrum also displayed a significant broadening and intensification

349

when compared with that of pure NOCNF. As this peak was due to the hydroxyl (-OH)

350

stretching, its broadening and intensification could no longer be attributed only to the presence of

351

water in NOCNF, but due to the formation of cadmium hydroxide (Cd(OH)2), which is discussed

352

later.

353 354

The steady state shear viscosity measurement of the cadmium(II) nitrate aqueous solution

355

(Cd2+=500 ppm), the NOCNF suspension at concentration 0.1 wt%, and the floc portion of the

356

mixture containing Cd2+ ions (500 ppm) and NOCNF (0.1 wt%) in suspension was carried out to

357

understand the interactions between Cd2+ ions and NOCNF. The viscosity results are illustrated

358

in Figure S2 (Supporting Information). It was seen that the cadmium nitrate solution possessed

359

very low viscosity, in which instability was observed at low shear rates. In contrast, the NOCNF

360

suspension (0.1 wt%) behaved as a Newtonian fluid with slight shear-thinning tendency at high

361

shear rates. However, the floc of Cd2+ and NOCNF aggregates displayed a drastic increase in

362

viscosity and pronounced shear-thinning behavior. In fact, the floc portion of the mixture

16 ACS Paragon Plus Environment

Page 16 of 50

Page 17 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

363

behaved exactly like a gel. This observation supported the notion that Cd2+ ions functioned as an

364

effective crosslinking agent to bind negatively charged NOCNF particles together, similar to the

365

hydrogel formation induced by ionic interactions between the oppositely charged molecules.57

366

Furthermore, the enhanced shear-thinning behavior in the floc implied that the total attraction

367

forces between Cd2+ and NOCNF, even through strong ionic interactions, were not sufficiently

368

high to resist the deformation forces at high shear rates. The above viscosity results are

369

complementary to the FTIR results (Figure 3(ii)), also indicating that Cd2+ is an effective

370

crosslinking agent to gel the NOCNF suspension.

371 372

It was interesting to note that the actual Cd2+ adsorption capacity of NOCNF in

373

suspension was much higher than the expected value based on the available carboxylate groups

374

on NOCNF. The very high adsorption capacity was due to the combined effects of cadmium(II)

375

adsorption by the charged NOCNF scaffold (as a polyelectrolyte) and the mineralization of

376

cadmium hydroxide crystals through nucleation and growth pathways. The mineralization of

377

metal ions in polyelectrolyte has been well documented. For example, Hirasawa et al. reported

378

that the mixing of lead sulfate solution with polyethyleimine (PEI) polyelectrolyte could lead to

379

nanosized lead crystals.58 In this system, the PEI component clearly acted as a nucleating agent

380

in the aqueous environment. Initially, no visible changes were observed upon the addition of lead

381

sulfate to the PEI polyelectrolyte. However, with the increasing lead sulfate content, the lead

382

concentration suddenly displayed a sharp decrease, which was termed as the ‘bend point’ by the

383

authors. Subsequently, the tranparent solution became to a milky suspension, containing a large

384

amount of nanocrystals. In the present study, a similar behavior was also observed after the

385

addition of Cd+2 ions into the NOCNF suspension.

17 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

386 387

The cadmium(II) remediation mechanism by NOCNF was revealed by the SEM/EDS

388

measurements. Figure 4 shows SEM images of two floc samples formed upon the addition of

389

Cd2+ solutions having two different concentrations: (i) 500 ppm and (ii) 1,000 ppm, to the 0.2

390

wt% NOCNF suspension. In Figure 4(i), the image of the floc with lower Cd2+ concentration

391

(500 ppm) exhibited the clear appearance of agglomerated nanofibers, partially covered with

392

white clouds of cadmium compound (as seen in the EDS inset), indicating the adsorption of Cd2+

393

ions onto the NOCNF surface. In fact, no evidence (by X-ray diffraction) of cadmium

394

mineralization was observed at this concentration. The quantitative evidence of the presence of

395

cadmium (Cd) compound was supported by the EDS spectrum, where the carbon (C), oxygen

396

(O) and sodium (Na) peaks were due to NOCNF, and the very large silicon (Si) peak was due to

397

the use of silicon wafer as the support substrate. The similar intensity of the peaks associated

398

with NOCNF and Cd indicated that mass ratio between NOCNF and the adsorbed cadmium was

399

similar. This suggests that at cadmium concentrations ≤500 ppm, the remediation mechanism

400

was mainly due to the adsorption of Cd+2 ions onto the NOCNF surface.

401 402

However, SEM image of the floc sample obtained upon the mixing of a cadmium(II)

403

solution with a higher concentration (1,000 ppm) and the NOCNF suspension showed very

404

different morphology (Figure 4(ii)). In this image, the fiber appearance was less visible, while

405

thick cloudy structures associated with cadmium hydroxide crystals became dominant in some

406

regions. The corresponding EDS spectrum (inset) showed the Cd peak with much higher

407

intensity, while other peaks (C, O, Na and Si) were maintained at the same level. This indicated

408

the mineralization of cadmium hydroxide crystals occurred. It also indicates that the adsorbed

18 ACS Paragon Plus Environment

Page 18 of 50

Page 19 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

409

cadmium on the NOCNF surface acted as a nucleus, which initiated the growth of cadmium

410

hydroxide crystals in the presence of high Cd+2 concentration. In this stage, the remediation

411

mechanism is mainly dominated by the crystal growth of cadmium hydroxide.

412 413

The formation of cadmium hydroxide nanocrystals could be confirmed by Figure 5,

414

which illustrates (A) TEM image and (B) corresponding WAXD profile of the floc sample

415

formed by mixing of 1,250 ppm Cd2+ solution and 0.2 wt% NOCNF suspension at pH = 7. In

416

TEM measurement, the floc sample was not stained with uranyl acetate to avoid the overlap of

417

contrast. In Figure 5A, it was seen that the NOCNF scaffold acted as a nucleating network in the

418

formation of cadmium hydroxide nanocrystals (black dots), where the average size of these

419

nanocrystals was in the range of 3-5 nm. The results are in full agreement with the data reported

420

in the previous literature, where the presence of cellulose nanofibers could induce the formation

421

of metal nanocrystals (ferrite, Ag, Au, ZnO, TiO2).59-61 Figure 5B shows the corresponding

422

WAXD profile of the floc sample in Figure 5A. Distinct diffraction peaks were seen, which

423

could be indexed by the unit cells of Cellulose I structure and cadmium hydroxide. In specific,

424

the peaks at the 2θ angle of 16.86° and 22.77° could be indexed by (110), (200) diffractions from

425

the Cellulose I structure, whereas the peaks at the 2θ angle of 29.2°, 31.9°, 35.4°, 38.9°, 42.8°,

426

48.09°, 55.4°, and 56.4° could be indexed by (001), (200), (220), (131ത), (111), (310), (201) and

427

(421ത) diffractions of cadmium hydroxide, respectively.62

428 429

Confocal images of NOCNF cast from the 0.2 wt% suspension, cadmium nitrate deposit

430

from a 1,000 ppm solution, and a floc sample prepared by mixing the NOCNF suspension and

431

1,000 ppm solution, are shown in Figure 6. Figure 6(i) represents the NOCNF aggregates with

19 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

432

NOCNF being stained by the acid blue dye, while Figure 6(ii) represents the cadmium nitrate

433

deposit with Cd2+ ions being stained by the LeadmiumTM AM dye (this dye has been commonly

434

used to detect lead or cadmium ions in living cells because of its good binding tendency with

435

these metal ions63). The two dyes have different excitation wavelengths: the acid blue dye has an

436

excitation at λ = 580 nm, and the LeadmiumTM AM dye has an excitation at λ = 405-490 nm. As

437

a result, their imaging by different lasers could be used to differentiate the regions of NOCNF

438

and Cd2+ ions at a macroscopic level. In specific, the laser with a wavelength λ = 580 nm was

439

used to take the image in Figure 6(i), which exhibited the agglomeration of NOCNF rather than

440

individual nanofibers; the laser with wavelength λ = 405-490 nm was used to take the image in

441

Figure 6(ii), exhibiting light shadow of cadmium nitrate deposit. Figure 6(iii) represents the floc

442

sample, prepared by mixing stained NOCNF with the acid blue and stained Cd2+ ions with the

443

LeadmiumTM AM dye, taking at λ = 580 nm. In this image, the same type of red textures (due to

444

the agglomeration of NOCNF) was seen as that in Figure 6(i). However, when imaging the same

445

floc sample at λ = 470 nm (Figure 6(iv)), a very different texture was seen when compared that

446

in Figure 6(ii). In specific, the floc images in Figure 6(iv) exhibited a granular texture (marked

447

by red circles) with the granules having much stronger intensity than the diffuse texture in the

448

image of cadmium nitrate deposit (Figure 6(ii)). This granular texture was also not seen in Figure

449

6(iii), in which the imaging was taken to highlight the stained NOCNF. The granular regions in

450

Figure 6(iv) were due to the clusters of collapsed NOCNF scaffold bonded by cadmium

451

hydroxide nanocrystals. These results further confirmed the occurrence of cadmium hydroxide

452

mineralization within the NOCNF scaffold during the remediation process.

453 454

Determine the Maximum Cadmium(II) Removal Capacity of NOCNF

20 ACS Paragon Plus Environment

Page 20 of 50

Page 21 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

455 456

The maximum removal capacity of NOCNF against cadmium(II) was determined as

457

follows. Based on the procedures outlined in the Experimental section, the ideal adsorption

458

capacity, adsorption efficiency and experimental adsorption capacity (Qe) of NOCNF for the

459

removal of cadmium(II) in the concentration range of 50 – 5,000 ppm were determined, and the

460

results are shown in Table 2. It was found that the adsorption (or removal) efficiency of NOCNF

461

against cadmium(II) in the 50-1,250 ppm region (except 100 ppm) was between 74 and 84 %.

462

However, the efficiency of NOCNF in the 2,500 and 5,000 ppm region was between 14 and

463

17%. Based on Table 2, the values of the original Cd2+ concentration (Ce), experimental

464

adsorption capacity (Qe) and Ce/Qe are summarized in Table 3. It was found that the plot of Qe

465

against Ce almost exhibited a linear relationship (Figure 7(i)). These results were rearranged

466

according to Eq. 1, where the plot of Ce/Qe versus Ce could be fitted with the Langmuir

467

Isotherm model, as shown in Figure 7(ii). The slope of the fit is proportional to the reciprocal of

468

the maximum adsorption capacity (Qm), which was about 2,550 mg/g (R2= 0.986).

469 470

The maximum adsorption capacity (Qm) of the spinifex-based NOCNF was compared

471

with other commercially available and experimental adsorbents, which are listed in Table 4. It

472

was found that the Qm of NOCNF in this study was about 40% higher than the best commercial

473

adsorbent based on magnesium oxide;51 it was also several times higher than those of porous

474

chitosan beads64 and carboxymethyl functionalized cellulose,65 all effective cadmium(II)

475

adsorbents, probably due to the much larger available surface area in NOCNF. The combined

476

advantage of the very large surface to volume ratio (nanofibers) and abundant charged groups

477

(COO-) in NOCNF was even more pronounced when compared to raw wood,66 coffee grounds,67

21 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 50

478

agriculture waste (e.g. corn stalk, maize cob and bagasse) and biowaste (e.g. oil cake and shell

479

dust),66-71

480

nanocomposites,82 where the Qm value of spinifex-based NOCNF was significantly higher. In

481

fact, extracted NOCNF also exhibited higher efficiency than the other carboxylated biopolymers,

482

such as carboxylated alginic acid,83 esterified saw dust bearing carboxyl group,84 carboxylated

483

chitosan,85 carboxylated cellulose nanocrystals10 and cellulose nanocrystal modified using

484

sodium succinate.10

synthetic

polymer,72

biochar,73

activated

charcoal,74-80

styofoam,81

and

485 486

The adsorption efficiency of NOCNF for removal of Cd2+ ions from water was found to

487

be pH dependent.

To demonstrated this effect, the adsorption efficiency of NOCNF for

488

cadmium(II) removal was measured at different pH level in a very high Cd+2 concentration

489

(10,000 ppm), where the result is shown in Figure 8. It was found that the highest adsorption

490

efficiency was achieved at pH = 7, which would enable practical applications. This behavior

491

could be explained as follows. At the very low pH values, the effective charge density became

492

lower as some carboxylate groups got turned into neutral carboxylic acid groups, while at very

493

high pH values, some NOCNF could be denatured. The observation of the high adsorption

494

capacity of NOCNF occurred at neutral pH is also consistent with the results obtained from

495

biosorption of olive stones.86

496 497

To further evaluate the usage of NOCNF as an adsorbent material to remove Cd2+ ions

498

from water, a freeze-dried NOCNF sample (in powder form) was used in the following

499

experiment. In specific, 0.056 g of NOCNF was re-dispersed in 150 mL of water, where the pH

500

value of the milky NOCNF suspension (Figure 9(i)) was found to be 5.30 and the corresponding

22 ACS Paragon Plus Environment

Page 23 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

501

conductivity was 30.40 µs. The milky appearance indicated that the re-dispersion of freeze-dried

502

NOCNF powder would take time and was in the microscopic level. However, when 10 mL of

503

cadmium nitrate solution (1,500 ppm) was added to this NOCNF suspension, a precipitation

504

formed rather quickly. For example, after 5 minutes of decanting, the coagulant of NOCNF and

505

cadmium hydroxide crystals completely precipitated to the bottom of the flask (Figure 9(ii)). The

506

removal of cadmium(II) by these freeze-dried NOCNF sample was complete, but the pH value of

507

the suspension in Figure 9(ii) was found to decrease to 4.50 and the conductivity increase to

508

220.4 µs. This might be due to the following reason. Upon the addition of cadmium nitrate

509

solution in the NOCNF suspension, the formation of nitric acid (HNO3) could occur, resulting

510

from the release of nitrate (NO3-) ions from cadmium nitrate and H+ ions from the COOH group

511

of NOCNF (in lower pH value). This reaction would lead to the decrease in pH and increase in

512

conductivity. This observation was quite different from a study of cadmium(II) removal from

513

cadmium acetate solution using olive stones.86 In that study, the precipitation of cadmium

514

hydroxide (Cd(OH)2) crystals took place at pH between 10 and 11, in which the maximum

515

removal efficiency of cadmium(II) was seen. The precipitate in Figure 9(ii) (due to the coagulant

516

of NOCNF and cadmium hydroxide crystals) could be easily removed by gravity driven

517

microfiltration process, where Figure 9(iii) illustrates the transparent solution after filtration

518

using a course filter paper (average pore size around 40 µm).

519 520

Conclusions

521 522

In this study, carboxycellulose nanofibers (NOCNF) extracted from an arid grass spinifex

523

using the nitro-oxidation method were demonstrated to be an effective flocculating or

23 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

524

coagulating agent for removal of cadmium(II) from water. The maximum adsorption (removal)

525

capacity of the studied NOCNF was found to be 2,550 mg/g, determined by the Langmuir

526

isotherm model using the data from a static adsorption study. This capacity was significantly

527

higher than those reported in the literature for cadmium(II) removal. The removal pathway was

528

found to contain two mechanisms: at low cadmium(II) concentrations (< 500 ppm), the removal

529

was dominated by the charge interactions between the COO- groups and Cd2+ ions, whereas at

530

high cadmium(II) concentrations (> 1,000 ppm), the removal was dominated by the formation of

531

Cd(OH)2 nanocrystals. As Cd2+ ions behaved as an effective crosslinking agent to gel the

532

NOCNF suspension, the precipitate of NOCNF/Cd(OH)2 aggregates could be easily removed by

533

gravity-driven (or very low energy) microfiltration using filters of large pore size (e.g. 40 µm),

534

or by decantation. The nitro-oxidation method was found to be a simple and effective means to

535

extract CNF from spinifex, which could be viewed as a model for many underutilized arid

536

grasses (e.g. agave) or even agriculture waste.

537

adsorbent to purify the cadmium(II) contaminated water because of its very large surface to

538

volume ratio and abundant carboxylate groups. The spinifex-based NOCNF system can be

539

viewed as a mode system to tackle the emerging cadmium pollution problems in many part of the

540

world, where NOCNF can be extracted from a wide range of underutilized biomasses.

The resulting NOCNF is an outstanding

541 542

Supporting Information

543 544

Detailed descriptions on instruments used, such as Fourier transform infra-red 13

545

spectroscopy (FTIR),

C nuclear magnetic resonance (NMR), conductometric titration method,

546

zeta potential measurement, transmission electron microscopy (TEM), scanning electron

24 ACS Paragon Plus Environment

Page 24 of 50

Page 25 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

547

microscopy (SEM), wide-angle X-ray diffraction (WAXD), surface area measurement by the

548

Brunauer-Emmett-Teller (BET) method, confocal microscopy, inductively coupled plasma mass

549

spectroscopy (ICP-MS), contact angle measurement (including a figure illustrating the contact

550

angle measurement of NOCNF extracted from spinifex fibers) and rheological measurements

551

with figure showing the steady state shear viscosity data of cadmium(II) nitrate solution,

552

NOCNF suspension, and the floc portion of the cadmium(II)/NOCNF mixture.

553 554

Acknowledgments

555 556

The authors would like to thank the SusChEM Program of the National Science

557

Foundation (DMR-1409507) for the financial support. The authors would also like to thank

558

Susan von Horn (iLab-Stony Brook University) and Dr. Chung-Chueh Chang and Ya-Chen

559

Chuang (ThINC facility at AERTC, Stony Brook University, USA) for conducting the TEM

560

measurements, Dr. Jim Quinn (Materials Science and Engineering- Stony Brook University) for

561

the SEM measurements and Dr. David Hirschberg (School of Marine and Atmospheric Science,

562

Stony Brook University) for conducting the ICP-MS analysis.

563 564

References

565

1. Soisungwan, S. Long-term exposure to cadmium in Food and Cigarette Smoke, Liver effects

566 567 568

and Hepatocellular carcinoma. Current Drug Metabolism. 2012, 13(3), 257 – 271. 2. Degraeve, N. Carcinogenic, Tetratogenic and Mutagenic effects of cadmium. Mutat. Res. 1981, 1, 115-135.

25 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

569

3. Yong, J.; Alan, C.; Robert, S.; Hanan, A. R.; Jack, T.; Thomas, K.; Michael, R.; Dmitry, G.

570

Cadmium is a mutagen that acts by inhibiting mismatch repair. Nature Genet. 2003, 34, 326-

571

329.

572

4. https://www.osha.gov/SLTC/cadmium/ Retrieved on 19th February 2017.

573

5. Agency for Toxic Substances and Disease Registry. 1989. Toxicological profile for

574 575 576 577 578

cadmium. ATSDR/U.S. Public Health Service, ATSDR/TP-88/08. 6. Goyer. R. Toxic effects of metals. In: Amdur, M.O., J.D. Doull and C.D. Klaassen, Eds. Casarett and Doull's Toxicology. 4th ed. Pergamon Press, New York. 1991, 623-680. 7. Anetor, J. Rising environmental cadmium levels in developing countries: threat to genome stability and health. J. Physiol. Sci. 2012, 27(2), 103-15.

579

8. Debajit, D.; Hari, S. Copper (Cu), Zinc (Zn) and Cadmium (Cd) Contamination of

580

Groundwater in Dikrong River Basin, Paumpare District, Arunachal Pradesh, India. IOSR

581

JESTFT 2015, 9(10), 20-23.

582

9. Annabella, G.; Lucas, S.; Lucrecia, F. Cadmium toxicity assessment in juveniles of the

583

Austral South America amphipod Hyalella curvispina. Ecotoxicol. Environmen. Saf. 2012,

584

79, 163-169.

585

10. Yu, X.; Tong, S.; Ge, M.; Wu, L.; Zuo, J.; Cao, C.; Song, W. Adsorption of heavy metal ions

586

from aqueous solution by carboxylated cellulose nanocrystals. J. Environ. Sci. 2013, 25 (5)

587

933–943.

588 589 590 591

11. Srivastava, S.; Kardam, A.; Raj, K. R. Nanotech reinforcement onto cellulose fibers: Green remediation of toxic metals. Int. J. Green Nanotechnol. 2012, 4, 46-53. 12. Ma, H.; Hsiao, B. S.; Chu, B. Ultrafine cellulose nanofibers as efficient removal of UO22+ in water. ACS Macro. Lett. 2012, 1, 213-216.

26 ACS Paragon Plus Environment

Page 26 of 50

Page 27 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

592

13. Yang, R.; Aubrecht, K. B.; Ma, H. Y.; Wang, R.; Grubbs, R. B.; Hsiao, B. S.; Chu, B. Thiol-

593

modified cellulose nanofibrous composite membranes for chromium(VI) and lead(II)

594

adsorption. Polymer 2014, 55, 1167-1176.

595

14. Sing, K.; Arora, J. K.; Sinha, T. J. M.; Srivastava, S. Functionalization of nanocrystalline

596

cellulose for decontamination of Cr(III) and Cr(VI) from aqueous system: Comutational

597

modeling approach. Clean Techn. Environ. Policy 2014, 16.

598 599 600 601

15. Carpenter, A. W.; de Lannoy, C. F.; Wiesner, M. R. Cellulose nanomaterials in water treatment technologies. Environ. Sci. Technol. 2015, 49, 5277-5287. 16. Thakur, V. K.; Voicu, S. I., Recent advances in cellulose and chitosan based membranes for water purification: A concise review. Carbohydr. Polym. 2016, 146, 148-65.

602

17. Wang, R.; Guan, S.; Sato, A.; Wang, X.; Wang, Z.; Yang, R.; Hsiao, B. S.; Chu, B.,

603

Nanofibrous microfiltration membranes capable of removing bacteria, viruses and heavy

604

metal ions. Journal of Membrane Science 2013, 446, 376-382.

605 606 607 608 609 610

18. Ma, H.; Burger, C.; Hsiao, B. S.; Chu, B., Nanofibrous microfiltration membrane based on cellulose nanowhiskers. Biomacromolecules 2012, 13 (1), 180-186. 19. Ma, H.; Burger, C.; Hsiao, B. S.; Chu, B., Ultrafine polysaccharide nanofibrous membranes for water purification. Biomacromolecules 2011, 12 (4), 970-976. 20. Ma, H.; Burger, C.; Hsiao, B. S.; Chu, B., Ultra-fine cellulose nanofibers: new nano-scale materials for water purification. Journal of Materials Chemistry 2011, 21 (21), 7507.

611

21. Gupta, V. K.; Kumar, R., Nayak, A.; Saleh, T. A.; Barakat, M. A. Adsorptive removal of

612

dyes from aqueous solution onto carbon nanotubes: A review. Adv. Colloid Interface Sci.

613

2013, 193-194, 24–34.

27 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

614

22. Gupta, V. K.; Mittal, A.; Jhare, D.; Mittal, J. Batch and bulk removal of hazardous colouring

615

agent Rose Bengal by adsorption techniques using bottom ash as adsorbent. RSC Adv. 2012,

616

2, 8381–8389.

617 618

23. Gupta, V. K.; Nayak, A.; Agarwal, S. Bioadsorbents for remediation of heavy metals: Current status and their future prospects. Environ. Eng. Res. 2015, 20 (1), 001-018.

619

24. Mittal, A.; Mittal, J.; Malviya, A.; Kaur, D.; Gupta, V. K. Decoloration treatment of a

620

hazardous triarylmethane dye, Light Green SF(Yellowish) by waste material adsorbents. J.

621

Colloid Interface Sci. 2010, 342, 518-527.

(13)

622

25. Gupta, V. K.; Mittal, A.; Jhare, D.; Mittal, J. Batch and bulk removal of hazardous colouring

623

agent Rose Bengal by adsorption techniques using bottom ash as adsorbent. RSC Advances

624

2012, 2, 8381–8389.

625 626

26. Gupta, V. K.; Sharma, S. Removal of Zinc from Aqueous Solutions Using Bagasse Fly Ash -a Low Cost Adsorbent. Ind. Eng. Chem. Res. 2003, 42, 6619-6624

627

27. Mohammadi, N.; Khani, H.; Gupta, v. K.; Amereh, E.; Agarwal, S. Adsorption process of

628

methyl orange dye onto mesoporous carbon material-kinetic and thermodynamic studies. J.

629

Colloid Interface Sci. 2011, 362, 457-462.

630

28. Saravanan, R.; Karthikeyan, S.; Gupta, V. K.; Sekaran, G.; Narayanan, V.; Stephen, A.

631

Enhanced photocatalytic activity of ZnO/CuO nanocomposite for the degradation of

632

textile dye on visible light illumination. Mater. Sci. Eng. C 2013, 33, 91-98.

633

29. Saravanan, R.; Thirumal, E.; Gupta, V. K.; Narayanan, V.; Stephen, A. The photocatalytic

634

activity of ZnO prepared by simple thermal decomposition method at various temperatures.

635

J. Mol. Liq. 2013, 177, 394.

28 ACS Paragon Plus Environment

Page 28 of 50

Page 29 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

636

30. Gupta, V. K.; Jain, R.; Nayak, A.; Agarwal, S.; Shrivastava, M. Removal of the hazardous

637

dye-Tartrazine by photodegradation on titanium-dioxide surface. Mater. Sci. Eng. C 2011,

638

31, 1062-1067.

639

31. Jain, A. K.; Gupta, V. K.; Bhatnagar, A.; Suhas. A Comparative Study of Adsorbents

640

Prepared from Industrial Wastes for Removal of Dyes. Sep. Sci. Technol. 2003, 38 (2), 463-

641

481.

642

32. Saleh, T. A.; Gupta, V. K.

Functionalization of tungsten oxide into MWCNT and its

643

application for sunlight-induced degradation of rhodamine B. J. Colloid Interface Sci. 2011,

644

362, 337-344.

645

33. Saravanan, R.; Sacari, E.; Gracia, F.; Khan, M. M., Mosquera, E.; Gupta, V. K. Conducting

646

PANI stimulated ZnO system for visible light photocatalytic degradation of coloured dyes. J.

647

Mol. Liq. 2016, 221, 1029-1033.

648

34. Rajendran, S.; Khan, M. M.; Gracia, F.; Qin, J. Gupta, V. K.; Arumainathan, S. Ce3+-ion-

649

induced visible-lightphotocatalytic degradation and electrochemical activity of ZnO/CeO2

650

nanocomposite. Sci. Rep. 2016, 6 (31641), 1-11.

651

35. Saravanan, R.; Karthikeyan, S. Gupta, V. K.; Sekaran, G.; Narayanan, V.; Stephen, A.

652

Enhanced photocatalytic activity of ZnO/CuO nanocomposite for the degradation of textile

653

dye on visible light illumination. Mater. Sci. Eng. C 2013, 33, 91-98.

654

36. Saravanan, R.; Thirumal, E.; Gupta, V.K.; Narayanan, V. Stephen, A. The photocatalytic

655

activity of ZnO prepared by simple thermal decomposition method at various temperatures.

656

J. Mol. Liq. 2013, 177, 394-401.

29 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

657

37. Saleh, T. A.; Gupta, V. K. Synthesis and characterization of alumina nano-particles

658

polyamide membrane with enhanced flux rejection performance. Sep. Purif. Technol. 2012,

659

89 245–251.

660

38. Karthikeyan, S.; Gupta, V. K.; Boopathy, R.; Titus, A.; Sekaran, G. A new approach for the

661

degradation of high concentration of aromatic amine by heterocatalytic Fenton oxidation:

662

Kinetic and spectroscopic studies. J. of Mol. Liq. 2012, 173, 153-163.

663

39. Asfaram, A.; Ghaedi, M.; Agarwal, S.; Tyagib, I.; Gupta, V. K. Removal of basic dye

664

Auramine-O by ZnS: Cu nanoparticles loaded on activated carbon: optimization of

665

parameters using response surface methodology with central composite design. RSC Adv.

666

2015, 5, 18438- 18450.

667

40. Devaraj, M; Saravanan, R.; Deivasigamani, R. K.; Gupta, V. K.; Gracia, F.; Jayadevan, S.

668

Fabrication of novel shape Cu and Cu/Cu2O nanoparticles modified electrode for the

669

determination of dopamine and paracetamol. J. Mol. Liq. 2016, 221, 930–941.

670

41. Hubbard, A. T. (Editor). Encyclopedia of Surface and Colloid Science. 2004.

671

42. Saleh, T. A.; Gupta, V. K. Photo-catalyzed degradation of hazardous dye methyl orange by

672

use of a composite catalyst consisting of multi-walled carbon nanotubes and titanium

673

dioxide. J. Colloid Interface Sci. 2012, 371, 101-106.

674

43. Saravanan, R.; Gupta, V. K.; Mosquera, E.; Gracia, F.; Narayanan, V., Stephen, A., Visible

675

light induced degradation of methyl orange using Ag0.333V2O5 nanorod catalysts by facile

676

thermal decomposition method. J. Saudi Chem. Soc. 2015, 19 (5), 521-527.

677 678

44. Alan F. M. "Grasslands-Spinifex". Te Ara - the Encyclopedia of New Zealand 2009, Retrieved on 31 January 2010.

30 ACS Paragon Plus Environment

Page 30 of 50

Page 31 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

679

45. Amiralian, N.; Annamalai, P. K.; Memmott, P.; Martin, D. J. Isolation of cellulose

680

nanofibrils from Triodia pungens via different mechanical methods. Cellulose 2015, 22(4),

681

2483-2498.

682

46. Amiralian, N.; Annamalai, P. K.; Memmott, P.; Taran, E.; Schmidt, S.; Martin, D. J.

683

Reinforcement of natural rubber latex using lignocellulosic nanofibers isolated from spinifex

684

grass. J. RSC Adv. 2015, 5, 32124–32132.

685

47. Hosseinmardi, A.; Annamalai, P. K.; Wang, L.; Martin, D. J.; Amiralian, N. Reinforcement

686

of natural rubber latex using lignocellulosic nanofibers isolated from spinifex grass.

687

Nanoscale 2017, 9 (27), 9510-9519.

688

48. Mohd Amin, K. N.; Amiralian, N.; Annamalai, P. K.; Edwards, G.; Chaleat, C.; Martin, D. J.

689

Scalable processing of thermoplastic polyurethane nanocomposites toughened with

690

nanocellulose. Chem. Eng. J. 2016, 302, 406-416.

691

49. Sharma, P. R.; Joshi, R.; Sharma, S.K.; Hsiao, B. S. A Simple Approach to Prepare

692

Carboxycellulose Nanofibers from Untreated Biomass. Biomacromolecules 2017, 18(8),

693

2333-2342.

694

50. Cheng, M.; Qin, Z.; Chen, Y.; Ran, Z. Facile one-step extraction and oxidative carboxylation

695

of cellulose nanocrystals through hydrothermal reaction by using mixed inorganic acids.

696

Cellulose, 2017, 24, 3243-3254.

697

51. Cao, C.; Qu, J.; Wei, F.; Liu, H; Song, W. Superb adsorption capacity and mechanism of

698

flowerlike magnesium oxide nanostructures for lead and cadmium ions. ACS Appl. Mater.

699

Interfaces, 2012, 4, 4283-4287.

700 701

52. Wang, J.; Kuo, Y. Preparation of fructose-mediated (polyethylene glycol/chitosan) membrane and adsorption of heavy metal ions. J. Polym. Sci. 2007, 105, 1480-1489.

31 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

702

53. Stefelova, J.; Slova k, V.; Siqueira, G.; Olsson, R. T.; Tingaut, P.; Zimmermann, T.; Sehaqui,

703

H. Drying and pyrolysis of cellulose nanofibers from wood, bacteria, and algae for char

704

application in oil absorption and dye adsorption. ACS Sustainable Chem. Eng. 2017, 5,

705

2679−2692.

706

54. Reddy, K. O.; Ashok, B.; Reddy, K. R. N.; Feng, Y. E.; Zhang, J.; Rajulu, A. V. Extraction

707

and characterization of novel lignocellulosic fibers from Thespesia lampas plant. Int. J.

708

Polym. Anal. Charact. 2014, 19, 48–61.

709 710 711 712 713 714

55. Isogai, A.; Usuda, M.; Kato, T.; Uryu, T.; Atalla, R. H. Solid-state CP/MAS carbon-13 NMR study of cellulose polymorphs. Macromolecules 1989, 22(7), 3168-1393. 56. Chidambareswaran, P. K.; Sreenivasan, S.; Patil, N. B.; Sundaram, V.; Srinathan, B. X-ray diffraction studies on cotton/jute blends. J. Appl. Polym. Sci. 1976, 20, 3443-3448. 57. Hennick, W. E.; van Nostrum, C. F. Novel crosslinking methods to design hydrogels. Adv. Drug Deliv. Rev., 2012, 64, 223-236.

715

58. Hirasawa, I.; Mikani, T.; Katayama, A.; Sakuma, T. Strategy to obtain nm size crystals

716

through precipattion in the presence of polyelectrolyte. Chem. Eng. Technol. 2006, 29 (2),

717

212-214.

718

59. Galland, S.; Andersson, R. L.; Salajkov, M.; Strom,V; Olsson, R. T.; Berglund, L. A.

719

Cellulose nanofibers decorated with magnetic nanoparticles – synthesis, structure and use in

720

magnetized high toughness membranes for a prototype loudspeaker. J. Mater. Chem. C 2013, 1

721

(47), 7963-7972.

722 723

60. Adel, A. M. Incorporation of nano-metal particles with paper matrices. Interdiscip. J. Chem. 2016, 1 (2), 36-46.

32 ACS Paragon Plus Environment

Page 32 of 50

Page 33 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

724

61. Boufi, S.; Ferraria, A. M.; Botelho, doRego. A. M.; Battaglini, N.; Herbst, F.; Vilar, M. R.

725

Surface functionalisation of cellulose with noble metals nanoparticles through a selective

726

nucleation. Carbohydr. Polym. 2011, 86 (4), 1586-1594.

727 728

62. Qu, P.; Shancheng, Y.; Meng, H. Controllabe growth of cadmium hydroxide nanostructures by hydrothermal method. Solid State Sciences 2010, 12, 83–87.

729

63. https://tools.thermofisher.com/content/sfs/manuals/mp10024.pdf retrieved 17 August 2017.

730

64. Rorrer, G. L.; Hsien, T. Y. Synthesis of porous-magnetic chitosan beads for removal of

731

cadmium ions from waste water. Ind. Eng. Chem. Res. 1993, 32, 2170-2178.

732

65. Wei, W.; Sok, K.; Myung-He, S.; Bediako, K.; Yeoung-Sang, Y. J. Carboxymethyl cellulose

733

fiber as a fast binding and biodegradable adsorbent of heavy metals. Taiwan. Inst. Chem.

734

Eng. 2015, 57, 104-110.

735 736

66. Mohd, R.; Othman, S.; Rokiah, H.; Anees, A. Removal of cadmium(II) from aqueous solutions by adsorption using meranti wood. Wood Sci. Technol. 2012, 46 (1), 221-241.

737

67. Mohd, R.; Othman, S.; Rokiah, H.; Anees, A. Adsorption of cadmium from aqueous solution

738

onto untreated coffee grounds: equilibrium, kinetics and thermodynamics. J. Hazard. Mater.

739

2010, 186, 124-134.

740

68. Liuchun, Z.; Chaofei, Z.; Zhi, D.; Hui, Z.; Xiaoyun, Y.; Congqiang, L. Preparation of

741

cellulose derived from corn stalk and its application for cadmium ion adsorption from

742

aqueous solution. Carbohydr. Polym. 2012, 90, 1008-1015.

743 744

69. Azouaou, N.; Sadaoui, Z.; Mokaddem, H. Removal of cadmium from aqueous solution by adsorption on vegetable wastes. J. Appl. Sci. 2008, 8, 4638-4643.

33 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

745

70. Garg, U.; Kaur, M. P.; Jawa, G. K.; Sud, D.; Garg, V. K. Removal of cadmium(II) from

746

aqueous solutions by adsorption on agricultural waste biomass. J. Hazard. Mater. 2008, 154,

747

1149–1157.

748

71. Hossain, A.; Bhattacharyya, S. R.; Aditya, G. Biosorption of cadmium by waste shell dust of

749

fresh water mussel Lamellidens marginalis: Implications for metal bioremediation. ACS

750

Sustainable Chem. Eng. 2015, 3, 1-8.

751

72. Deng, S.; Zhang, G.; Liang, S.; Wang, P. Microwave assisted preparation of thiol-

752

functionalized polyacrylonitrile fiber for the selective and enhanced adsorption of mercury

753

and cadmium from water. ACS Sustainable Chem. Eng. 2017, 5, 6054-6063.

754

73. Liang, J.; Li, X.; Yu, Z.; Zeng, G.; Luo, Y.; Jiang, L.; Yang, Z.; Qian, Y.; Wu, H.

755

Amorphous MnO2 modified biochar derived from aerobically composted swine manure for

756

adsorption of Pb(II) and Cd(II). ACS Sustainable Chem. Eng. 2017, 5, 5049-5058.

757

74. Guo, Z.; Zhang, X.; Kang, Y.; Zhang, J. Biomass derived carbon sorbents for Cd(II) removal:

758

Activation and adsorption mechanism. ACS Sustainable Chem. Eng. 2017, 5, 4103-4109.

759

75. Kula, I.; Uğurlu, M.; Karaoğlu, H.; Celik, A. Adsorption of Cd(II) ions from aqueous

760

solutions using activated carbon prepared from olivestone by ZnCl2 activation. Bioresour.

761

Technol. 2008, 99 (3), 492−501.

762

76. Mohan, D.; Singh, K. P. Single-and multi-component adsorption of cadmium and zinc using

763

activated carbon derived from bagasse agricultural waste. Water Res. 2002, 36(9),

764

2304−2318.

765

77. Kobya, M.; Demirbas, E.; Senturk, E.; Ince, M. Adsorption of heavy metal ions from

766

aqueous solutions by activated carbon prepared from apricot stone. Bioresour. Technol.

767

2005, 96 (13), 1518−1521.

34 ACS Paragon Plus Environment

Page 34 of 50

Page 35 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

768 769 770 771

ACS Sustainable Chemistry & Engineering

78. Wilson, K.; Yang, H.; Seo, C. W.; Marshall, W. E. Select metal adsorption by activated carbon made from peanut shells. Bioresour. Technol. 2006, 97 (18), 2266−2270. 79. Jia, Y.; Thomas, K. Adsorption of cadmium ions on oxygen surface sites in activated carbon. Langmuir 2000, 16 (3), 1114−1122.

772

80. Boudrahem, F.; Soualah, A.; Aissani-Benissad, F. Pb(II) and Cd(II) removal from aqueous

773

solutions using activated carbon developed from coffee residue activated with phosphoric

774

acid and zinc chloride. J. Chem. Eng. Data 2011, 56 (5), 1946−1955.

775

81. Mahmoud, M. E.; Abdou, A. E.; Ahmed, S. B. Conversion of waste styrofoam into

776

engineered adsorbents for efficient removal of cadmium, lead and mercury from water. ACS

777

Sustainable Chem. Eng. 2016, 4 (3), 819−827.

778

82. Tang, J.; Mu, B.; Zheng, M.; Wang, A. One-step calcination of the spent bleaching earth for

779

the efficient removal of heavy metal ions. ACS Sustainable Chem. Eng. 2015, 3 (6),

780

1125−1135.

781 782

83. Jeon, C.; Park, J. Y.; Yoo, Y. J. Characteristics of metal removal using carboxylated alginic acid. Water Res. 2002, 36, 1814-1824.

783

84. Marchetti, V.; Cleament, A.; Gearardin, P.; Loubinoux, B. Synthesis and use of esterified

784

sawdusts bearing carboxyl group for removal of cadmium(II) from water. Wood Sci. Technol.

785

2000, 34, 167-173.

786

85. Lv, K. L.; Du, Y. L.; Wang, C. M. Synthesis of carboxylated chitosan and its adsorption

787

properties for cadmium(II), lead(II) and copper(II) from aqueous solutions. Water Sci.

788

Technol. 2009, 60.2, 467-474.

789 790

86. Blazquez, G.; Hernainz, F.; Calero, M.; Ruiz-Nunez, L. F. Removal of cadmium ions with olive stones: the effect of some parameters. Process Biochem. 2005, 40, 2649-2654.

791 35 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 50

792

Table 1. General properties of NOCNF extracted from spinifex using the nitro-oxidation

793

method. Charge

- 68 mV

COO-

Size

content

(L/D) nm

0.86

190±90/

mmol/g

4.0±1.5

PDI

0.35±0.01

794 795

PDI: polydispersity index; CI: crystallinity index; SA: surface area

796 797

36 ACS Paragon Plus Environment

CI

SA

(%)

(m2/g)

53

5.7

Page 37 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

798

Table 2. Calculated ideal adsorption capacity and experimental adsorption capacity of NOCNF

799

for the cadmium(II) removal in the Cd+2 concentration range of 50-5,000 ppm. Final Cd+2

Original Cd+2

Adsorption

Original

Original NOCNF

Ideal

Experimental

conc. by

conc. by

efficiency

quantity

(0.2 wt%)/2mL

adsorption

adsorption

ICPMS

ICPMS

Cd2+(mg)

(g)

capacity

capacity

(ppb)

(ppb)

(mg/g)

Qe (mg/g)

4,300

5,000

0.14

20

0.004

5,000

736

2,075

2,500

0.17

10

0.004

2,500

428

325

1,250

0.74

5

0.004

1,250

233

110

500

0.78

2

0.004

500

99

40

250

0.84

1

0.004

250

53

100

125

0.20

0.5

0.004

125

26

9.5

50

0.81

0.2

0.004

50

10

800 801

Total amount of NOCNF used in 2 mL of 0.2 wt% suspension = 0.004 g

802

Adsorption efficiency = (original Cd2+ conc. - final Cd2+ conc.) / original Cd2+ conc.

803

Ideal adsorption capacity = milligrams of Cd2+in solution / grams of NOCNF in suspension

804

Experimental adsorption capacity = adsorption efficiency x ideal adsorption capacity

805

Qe: experimental adsorption capacity; Ce: original concentration of Cd2+ in ppm

806

37 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

807

Table 3. The relationship between the values of Ce/Qe versus Ce; the results could be fitted by the

808

Langmuir adsorption model to determine the maximum adsorption capacity (Qm) of NOCNF. 809

Ce (ppm)

Qe (mg/g)

Ce/Qe

810

5,000

736

6.80

811

2,500

428

5.84

812

1,250

233

5.36

813

500

99

5.08

814

250

53

4.76

815

125

26

4.9

816

50

10

5

817

38 ACS Paragon Plus Environment

Page 38 of 50

Page 39 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

818

Table 4. Comparison of the Qm value obtained for NOCNF with those of different adsorbents

819

reported in the literature. Type of adsorbent

Maximum adsorption

Reference

capacity (mg /g) NOCNF

2,550

This study

flower – like magnesium oxide

1,500

51

chitosan beads

518-188

64

epichlorohydrin cross-linked carboxymethyl

150

65

meranti wood

150 – 175

66

untreated coffee grounds

15 – 17

67

corn stalk

3.39

68

maize cob

105

69

jatropha oil cake

86

70

sugarcane bagasse

69

70

waste shell dust of water mussel

18.18

71

thiol functionalized polyacrylonitrile fibers

350.6

72

MnO2-biochar

45.8

73

activated charcoal

8-40

74-80

styrofoam

50-70

81

attapulgite/ carbon nanocomposites

46

82

carboxylated alginic acid

50-100

83

esterified saw dust bearing carboxyl group

198

84

cellulose

39 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 40 of 50

carboxylated chitosan

720-800

85

carboxylated cellulose nanocrystals

1.9

10

cellulose nanocrystals modified using

345

10

sodium succinate 820 821

40 ACS Paragon Plus Environment

Page 41 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 1. (i) FTIR spectra and (ii) 13C CPMAS NMR spectra of spinifex fibers and NOCNF.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. (i) WAXD profiles of (A) spinifex and (B) NOCNF; (ii) TEM image of NOCNF extracted from spinifex using the nitro-oxidation method.

ACS Paragon Plus Environment

Page 42 of 50

Page 43 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 3. (i) Comparative images of two samples: (left) 5 mL of solution based on 5,000 ppm of cadmium nitrate and distilled water, (right) 5 mL of a NOCNF suspension (0.20 wt%) mixed the cadmium nitrate solution (5,000 ppm) at pH 7, (ii) FTIR spectra of (A) NOCNF and (B) floc obtained from the mixture of cadmium nitrate solution (500 ppm of Cd2+) and NOCNF suspension (0.20 wt%).

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4. (i) SEM of the floc obtained from the mixture of NOCNF suspension and 500 ppm of cadmium nitrate solution (inset: EDS spectra); (ii) SEM of the floc obtained from the mixture of NOCNF suspension and 1,000 ppm of cadmium nitrate solution (inset: EDS spectra).

ACS Paragon Plus Environment

Page 44 of 50

Page 45 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 5. (i) TEM image of the floc containing NOCNF and Cd(OH)2 nanocrystals (black dots); (ii) corresponding WAXD profile (reflection peaks were indexed by the unit cells of cellulose I and Cd(OH)2) of the floc obtained by mixing a 1,250 ppm of Cd2+ solution and NOCNF suspension (0.20 wt%) at pH = 7.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60



Figure 6. Confocal images of (i) NOCNF, stained with the acid blue (using the laser with = 580 nm), (ii) cadmium nitrate deposited with a 1,000 ppm solution, where Cd2+ ions were stained by LeadmiumTM AM dye (using the laser with = 405-490 nm); (iii) the floc obtained by mixing stained NOCNF (with acid blue) and stained Cd2+ ions (with LeadmiumTM AM dye) taken at =580 nm; (iv) the same floc taken at =470 nm (the red circles indicated the aggregates of NOCNF and cadmium hydroxide nanocrystals).

ACS Paragon Plus Environment

Page 46 of 50

Page 47 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering



(i)

(ii) Figure 7. (i) Adsorption of Cd2+ ions by the NOCNF suspension at the Cd+2 concentration between 500 and 5,000 ppm, (ii) the fitting of the adsorption data using the Langmuir isotherm model (Qe is the adsorption capacity measured in milligrams of Cd2+ per gram of NOCNF, Ce is the original Cd2+ concentration measured in mg of cadmium(II) per liter of solution).

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 48 of 50



Figure 8. The effect of pH value on the Cd2+ adsorption efficiency of NOCNF at the cadmium concentration of 10,000 ppm.

ACS Paragon Plus Environment

Page 49 of 50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 9. Photographs of (i) freeze dried NOCNF (0.056 g) dispersed in 150 mL of water, (ii) 5 minutes after the addition of 10 mL cadmium nitrate solution (1,500 ppm) in the (i) NOCNF suspension, and (iii) filtration of (ii) through a filter paper (average pore size = 40 µm) using the gravity filtration method.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table of Content Graphic

Carboxylated nanofibers (NOCNF) extracted from spinifex using nitro-oxidation approach was efficient in removing cadmium(II) from water via adsorption and mineralization.

1

ACS Paragon Plus Environment

Page 50 of 50