Magnetic supramolecular polymer: Ultrahigh and highly selective Pb(II) capture from aqueous solution and battery wastewater

Magnetic supramolecular polymer: Ultrahigh and highly selective Pb(II) capture from aqueous solution and battery wastewater

Journal Pre-proof Magnetic supramolecular polymer: Ultrahigh and highly selective Pb(II) capture from aqueous solution and battery wastewater Zongwu W...

6MB Sizes 0 Downloads 5 Views

Journal Pre-proof Magnetic supramolecular polymer: Ultrahigh and highly selective Pb(II) capture from aqueous solution and battery wastewater Zongwu Wang, Jing Zhang, Qing Wu, Xuexue Han, Mengna Zhang, Wei Liu, Xinding Yao, Jinglan Feng, Shuying Dong, Jianhui Sun PII:

S0045-6535(20)30235-6

DOI:

https://doi.org/10.1016/j.chemosphere.2020.126042

Reference:

CHEM 126042

To appear in:

ECSN

Received Date: 28 November 2019 Revised Date:

13 January 2020

Accepted Date: 26 January 2020

Please cite this article as: Wang, Z., Zhang, J., Wu, Q., Han, X., Zhang, M., Liu, W., Yao, X., Feng, J., Dong, S., Sun, J., Magnetic supramolecular polymer: Ultrahigh and highly selective Pb(II) capture from aqueous solution and battery wastewater, Chemosphere (2020), doi: https://doi.org/10.1016/ j.chemosphere.2020.126042. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2020 Published by Elsevier Ltd.

CRediT authorship contribution statement Zongwu Wang: Conceptualization, Investigation, Writing-Original Draft, Writing-Review & Editing. Jing Zhang: Data Curation, Formal analysis. Qing Wu: Validation, Supervision. Xuexue Hana: Investigation. Mengna Zhang: Investigation. Wei Liu: Resources, Visualization. Xinding Yao: Resources, Visualization. Jinglan Feng: Visualization, Writing - Review & Editing. Shuying Dong: Resources, Visualization, Writing-Review & Editing. Jianhui Suna: Project administration, Supervision, Writing-Review & Editing.

Magnetic supramolecular polymer: Ultrahigh and highly selective Pb(II)

1 2

capture from aqueous solution and battery wastewater

3

Zongwu Wanga,b, Jing Zhanga,c, Qing Wua, Xuexue Hana, Mengna Zhanga, Wei Liub, Xinding Yaob,

4

Jinglan Fenga, Shuying Donga, *, Jianhui Suna, *

5

a

6

Henan Key Laboratory for Environmental Pollution Control, School of Environment, Henan

7

Normal University, Xinxiang, Henan, 453007, P. R. China

8

b

9

Engineering Technology Research Center of Green Coating Materials, Kaifeng, Henan, 475004, P.

MOE Key Laboratory of Yellow River and Huai River Water Environmental and Pollution Control,

School of Environment Engineering, Yellow River Conservancy Technical Institute, Henan

10

R. China

11

c

Sanmenxia Polytechnic, Sanmenxia, Henan, 472000, P. R. China

12 13

*Corresponding authors. E-mail: [email protected] (S. Dong), [email protected] (J. Sun).

14

Tel.: +86-373-3325971

15 16 17 18 19 20 21 22

1

23

Abstract

24

For the practical capture of heavy metal ions from wastewater, fabricating environmental

25

friendly adsorbents with high stability and super adsorption capacity are pursuing issue. In this

26

work, we develop magnetic supramolecular polymer composites (M-SMP) by using a simple

27

two-step hydrothermal method. Systematical characterizations of morphological, chemical and

28

magnetic properties were conducted to confirm the formation of M-SMP composites. The resulting

29

M-SMP composites were applied to remove Pb(II) from aqueous solution and from real battery

30

wastewater, and easy separation was achieved using a permanent magnet. By investigating the

31

effects of various parameters, we optimized their operating condition for Pb(II) adsorption by the

32

M-SMP. The uptake of Pb(II) onto M-SMP fitted well the pseudo-second-order and Langmuir

33

isotherm models, and favourable thermodynamics showed a spontaneous endothermic process.

34

The SMP endowed M-SMP with ultrahigh adsorption capacity for Pb(II) (946.9 mg g–1 at pH = 4.0,

35

T = 298 K), remarkable selectivity, satisfactory stability and desirable recyclability. In

36

Pb-contaminated lead-acid battery industrial wastewater, the concentration of Pb(II) declined from

37

18.070 mg L–1 to 0.091 mg L–1, which meets the current emission standard for the battery industry.

38

These merits, combined with simple synthesis and convenient separation, make M-SMP an

39

outstanding scavenger for the elimination of industrial Pb(II) wastewater.

40

Keywords: Magnetic supramolecular polymer; Pb(II); Adsorption; Lead-acid battery wastewater

41

2

42

1. Introduction

43

On account of the growing industrial activities and urbanization, the discharging of heavy

44

metals to environment poses continuous and increasing threats to the quality of the environment,

45

the food chain and human health seriously due to its high toxicity, bioaccumulation and

46

non-degradability (Ji et al., 2019; Li et al., 2019). Therefore, it is strategically and ecologically

47

imperative to explore reliable technology to tackle this global dilemma. Among the diverse

48

technologies, the adsorption method stands out for its simple operation, cost-effective and free of

49

toxic sludge (Fu and Wang, 2011; Wu et al., 2015; Xu et al., 2018). However, the lack of

50

adsorbents with excellent comprehensive performance has become a hindrance to the application

51

of adsorption technology. To date, much efforts have been devoted to the fabrication of various

52

engineered materials, including carbon, functionalized-silica, clay mineral, polymer-based and

53

metal-organic frameworks etc., to removal heavy metal ions from wastewater (Da'na, 2017; Dai et

54

al., 2018; Songwut Lapwanit et al., 2016; Wen et al., 2018; Yadav et al., 2019; Yang et al., 2019).

55

Because of their low toxic and environment-friendly qualities, polymeric materials are one of

56

the most promising candidates. Then also have excellent regulatable chemical functionality,

57

dimensional stability and extraordinary adsorption property. Furthermore, considering the ease of

58

separating a magnetic sample after treatment, magnetic polymeric adsorbents are becoming a focus

59

of research (Huang et al., 2018; Venkateswarlu and Yoon, 2015). Therefore, it is important to

60

develop a creative synthesis method and screen suitable raw material for producing high-efficiency

61

magnetic polymeric materials.

62

Among many supramolecular precursors, trithiocyanuric acid (TTCA) is the main ingredient

63

of commercial highly potent chelating agent TMT15, which is widely adopted in the industrial

64

wastewater management owing to its strong complex ability of sulfur ligands with many soft

65

Lewis acidic metal ions (Pb2+, Cd2+, Hg2+, and others) (Fu et al., 2019; Huang et al., 2018).

66

However, the aggregation, pH limitations and difficulty of recovery in practice seriously restrict its 3

67

application value. Organodisulfide polymer has been investigated as an excellent heavy metal

68

scavenger, but suffers from a cumbersome process and limited final yield, and requires the usage

69

of toxic reagents, resulting in a typical impediment to its large-scale application for heavy metal

70

recovery (Ko et al., 2017; Ko et al., 2018; Lin et al., 2019). Recent research indicates that the

71

melamine (MA)-based dendrimer also presents high metal capture performance owing to the

72

strong complexation between -NH2 groups and soft metal ions (Al et al., 2017; Sharahi and

73

Shahbazi, 2017). However, it also has a complicated synthetic process, and its potential ecological

74

toxicity does not meet the requirements of commercial application.

75

For these reasons, it is still imperative to develop a smart strategy to overcome these

76

problems while retaining the merits of TTCA and MA. There have been a few attempts to

77

sequentially add TTCA and MA to aqueous solution for precious metal recovery (Nagai et al.,

78

2015; Nagai et al., 2016). Nevertheless, to the best of our knowledge, the related studies for the

79

synthesis of adsorbent material and investigation of adsorption behaviour have not been reported.

80

Therefore, the development of a straightforward strategy to simultaneously address these issues

81

and further improve adsorption ability is attractive but remains a great challenge.

82

Hydrogen bonds abandoning the conventional covalent bonds is a popular synthetic strategy

83

used widely in organic chemical synthesis (Shalom et al., 2013; Whitesides et al., 1991).

84

Interestingly, the supramolecular polymer (SMP) derived from MA and TTCA comprises a

85

hexagonal network stacked in three dimensions with channels. This could be constructed by a

86

simple procedure via the highly hydrogen-bonding action of N–H···S and N–H···N, forming

87

precipitates in water with satisfactory chemical stability. The SMP in non-covalent combination is

88

the reorganization of MA and TTCA at molecular level (Fan et al., 2017; Ranganathan et al., 1999;

89

Whitesides et al., 1995). Considering the high complexation capacity and recovery property for

90

industrial and practical applications, we constructed magnetic supramolecular polymer (M-SMP)

91

by incorporating SMP and magnetic Fe3O4 nanoparticles (NPs) to solve the aforementioned 4

92

problems.

93

The objectives of this current study were (1) to synthesize M-SMP composites by a simple

94

hydrothermal method for effective capture of Pb(II) ions in acidic wastewater samples, (2) to

95

characterize the morphology and composition systematically by various microscopic and

96

spectroscopic techniques, (3) to investigate the adsorption performance of Pb(II) on M-SMP

97

composites through routine experiments, (4) to reveal the plausible interaction mechanism by the

98

results of batch experiments coupled with Fourier transform infrared (FT-IR) and x-ray

99

photoelectron spectroscopy (XPS) characterization, and (5) to explore the application of M-SMP

100

composites for attenuating Pb(II) concentrations in lead-acid battery wastewater. Given all that,

101

the as-prepared M-SMP materials possess considerable practical applications for heavy metal

102

contaminated industrial wastewater clean-up.

103

2. Experimental procedures

104

2.1 Materials and instrumentation

105

All chemicals in this experiment were analytical grade and were received without further

106

purification. Detailed information on the adsorbents and apparatus are listed in Supplementary

107

Information (SI) (Text S1).

108

2.2 Synthesis of magnetic supramolecular polymer (M-SMP) composites

109

Magnetic supramolecular polymer (M-SMP) composites were prepared successfully by a

110

simple two-step method. In the first step, the Fe3O4 nanoparticles (NPs) were obtained according

111

to previous work with some modification (Text S2). In the second step, melamine (315 mg) and

112

Fe3O4 (200 mg) were dispersed in 30 mL water and heated at 70 oC by sonication for 1 h to form a

113

homogeneous solution. Afterwards, trithiocyanuric acid (443 mg) was mixed thoroughly with 40

114

mL water, accompanied sonication for 0.5 h at 70 oC. Then the above two solutions were mixed

115

together, transferred into Teflon-lined autoclave, and maintained at 100 oC for another 2 h. Finally,

116

the resulting grey solid was collected by magnetic separation, cleaned with deionized water, and 5

117

dried at 60 oC for 12 h, which is denoted as M-SMP. The light yellow SMP was also prepared

118

without Fe3O4 NPs using the same method. The process is shown in Fig. 1a. Digital photographs

119

of raw materials and as-prepared composites are shown in Fig. S1, and the proposed route of the

120

synthesis and the chemical equation of M-SMP is presented in Fig. S2.

121 6

122

Fig. 1. (a) Schematic illustration of the fabrication procedure for M-SMP. (b) SEM images of SMP. (c) SEM

123

images of M-SMP. (d) TEM images of SMP. (e) TEM images of M-SMP. The top-left inset is a high-resolution

124

TEM image and the bottom-right is the corresponding SAED pattern.

125

2.3 Batch adsorption experiments

126

For batch adsorption procedures, 25 mg M-SMP solid was dispersed into 25 mL desired Pb(II)

127

solution in a 50 mL flask iodine, and the suspension was shaken under 120 rpm at the desired

128

temperature in a thermostatic shaker bath. Finally, the solid was separated from the mixing

129

solution via a permanent magnet, and the remaining metal ion concentrations were detected by

130

inductively coupled plasma-mass spectrometry (ICP-MS). The removal efficiency R (%) and

131

adsorption capacity qe (mg g–1) were calculated according to the relevant equations (Text S3).

132

2.4 Industrial battery wastewater samples

133

The industrial wastewater samples were obtained from the lead-acid battery factory located in

134

Xinxiang city (35.20oN, 113.51oE), Henan province, China. The mass concentration of Pb(II)

135

determined by ICP-MS was 18.070 mg L-1, Other quality indicators were listed in Table S1.

136

3. Results and discussion

137

3.1 Material characterizations

138

The morphologies of SMP and M-SMP were observed by scanning electron microscope

139

(SEM), as shown in Fig 1b-c. Compared with the SMP with the smooth surface (Fan et al., 2017),

140

the M-SMP composites illustrate a rod-like structure with abundant Fe3O4 nanoparticles wrapped

141

inside supramolecular polymer. This was further confirmed by transmission electron microscopy

142

(TEM) of SMP (Fig. 1d) and M-SMP (Fig. 1e). Furthermore, the surface of the M-SMP was

143

slightly rough compared with that of the SMP. Notably, the high-resolution TEM (HR-TEM)

144

image in the top-left inset of Fig. 1e shows that M-SMP composites have interlayer spacing of

145

about 0.253 nm, matching well with the (311) plane of Fe3O4 (Wang et al., 2017). The

146

corresponding selected area electron diffraction (SAED) (bottom-right inset of Fig. 1e) further

147

verified the crystalline structure of Fe3O4. 7

148 149

Fig. 2. FT-IR spectra (a), XRD patterns (b), XPS spectra (c) and Raman shift spectra (d) of the intermediate and

150

final as-prepared samples.

151

The FT-IR spectra of raw materials and composites are shown in Fig. 2a. For TTCA, the

152

characteristic peaks located at 1100, 1520, 2655 and 2900-3150 cm–1 represented the C=S, C-N,

153

S-H, and N-H groups, respectively. For MA, the peaks of 3426 and 810 cm–1 are ascribed to the

154

N-H stretching vibration and the triazine ring vibration. These results were in good agreement with

155

the literature reported (Jun et al., 2013; Ko et al., 2017; Li et al., 2016; Shalom et al., 2013).

156

Although the characteristic peaks of supramolecular polymers derived from MA and TTCA in the

157

SMP and the M-SMP both showed almost the sum of those of virgin MA and TTCA, some new

158

peaks appeared owing to the hydrogen bonding between TTCA and MA in composites. Compared

159

with the signal of TTCA and MA, the hydrogen bonding of N–H···S and N–H···N resulted in the

8

160

N-H stretching vibration shifting from 3426 cm–1 to 3418 cm–1 accompanied with the enhanced

161

intensity, the blue shift of the C=S stretching vibration from 1100 cm–1 to 1135 cm–1, and the

162

triazine ring vibration shifting from 810 cm–1 to 780 cm–1 (Jun et al., 2013; Li et al., 2016; Yang et

163

al., 2016). Surprisingly, a new peak at 1633 cm–1 was the signal of thione (Fu and Huang, 2018).

164

The peak at 554 cm–1 reflected the stretching vibration of the Fe-O bond (Song et al., 2019),

165

indicating the successful incorporation of Fe3O4 component. The formation of supramolecular

166

polymer composites was also demonstrated by the high crystallinity nature of SMP in Fig. 2b

167

(Shalom et al., 2013). In the x-ray diffraction (XRD) pattern of SMP in contrast to those of MA

168

and TTCA (Fig. S3), three significant peaks at 12.25o, 13.06o, and 18.41o can be indexed to the

169

in-planar packing. Also, the well-resolved peak at 24.58o with a 0.362 nm d-spacing corresponded

170

to the graphite-like structure of 2-D SMP sheets (Fan et al., 2017; Jun et al., 2013; Shalom et al.,

171

2013), showing the formation of a new supramolecular polymer arrangement (Feng et al., 2014).

172

The characteristic peaks of M-SMP sample offsetting the partial signals of Fe3O4 was in line with

173

those of SMP, indicating the chemical stability of SMP in the incorporation of Fe3O4 (Fig. 2b).

174

Furthermore, there was no difference in the XRD patterns of M-SMP before and after immersing

175

the M-SMP in water for 12 h, confirming the stability of as-prepared M-SMP in aqueous solution.

176

The surface composition and the chemical state of the element were investigated by XPS. For

177

the SMP, the dominant peaks centered at 162.1, 227, 288.2 and 399.7 eV confirmed the

178

coexistence of S, C and N elements (Fig. 2c), corresponding to S2p, S2s, C1s and N1s signals,

179

respectively (Fu and Huang, 2018; Huang et al., 2019). In the high-resolution XPS spectra (Fig.

180

S4a), there were three forms of C1s, including C bonds of the aromatic trithiol (25.0% of the total

181

carbon) at 288.7 eV, the triazine ring (48.6%) at 287.8 eV and the thione (26.4%) at 284.8 eV in

182

M-SMP. The high-resolution N1s XPS spectra of the M-SMP (Fig. 3a) were also deconvoluted by

183

the aromatic trithiol (61.6.0% of the total nitrogen) at 399.3 eV, the -NH- (31.3%) at 400.2 eV and

184

the thione (7.3%) at 400.8 eV (Dinda and Kumar, 2015; Fu and Huang, 2018; Fu et al., 2019; 9

185

Kucharski and Szukiewicz, 2000). Fig. 3b showed the fitted S2p spectra of S2p3/2, S2p1/2 in aromatic

186

trithiol, and S2p3/2 in thione, which were situated at 161.9, 162.6 and 163.2 eV, respectively (Kim et

187

al., 2015; Ko et al., 2018; Wang et al., 2017). Compared with the spectrum of SMP, the existence

188

of the Fe2p signal and the enhancement of O1s intensity accompanying the high-resolution spectra

189

of Fe2p and O1s (Fig. S4b-c) demonstrated that Fe3O4 was successfully incorporated with SMP.

190

Furthermore, the M-O bond at 531.2 eV confirmed the as-prepared M-SMP was not a simple solid

191

mixing process of SMP and Fe3O4 (Tian et al., 2019).

192

Fig. 2d and Fig. S5 depicted the Raman shift spectra of these samples. The dominant band at

193

687 cm–1 was the characteristic Raman signal of MA (Wang et al., 2016; Zhang et al., 2019),

194

which was attributed to the deformation modes of the in-plane triazine ring. Raman modes

195

centered at 448, 1125 and 1275 cm–1 were assigned to the N-C-S, C=S and C-N groups,

196

respectively, which originated from the direct evidence of TTCA structure (Kim et al., 2015; Li et

197

al., 2016). After the reaction of TTCA with MA, the characteristic peak of MA shifted to 691 cm–1

198

(Fig. 2d and Fig. S5a), and the representative peaks of TTCA shifted to 440, 1162 and 1264 cm–1

199

(Fig. 2d and Fig. S5b) in the SMP and the M-SMP. Furthermore, the fact that there were no

200

marked differences between the SMP and the M-SMP except for signal intensity confirmed that

201

the introduction of Fe3O4 did not affect the formation of hydrogen bonds.

202

The magnetization measurement was investigated by vibrating sample magnetometer (VSM)

203

(Fig. 3c). The M-SMP showed superparamagnetic nature with saturation magnetization (Ms) of

204

19.86 emu g–1, which was lower than that of naked Fe3O4 owing to a certain amount of

205

non-magnetic SMP. In Fig. 3c, inset (L) and Inset (R) are digital photographs of the heavy metal

206

aqueous solution with dispersed M-SMP before and after magnetic separation. Notably, the

207

M-SMP showed rapid response to the magnetic field, meeting the requirement of magnetic

208

separation. The synergy of the SMP and Fe3O4 nurtured the composite into a promising adsorbent

209

with easy separation. The thermal stability of as-prepared M-SMP and SMP was determined using 10

210

thermal gravimetric analysis (TGA) under N2 atmosphere (Fig. 3d). It can be seen that the quality

211

of M-SMP and SMP remains constant without quality loss up to 300 oC, demonstrating the higher

212

thermal stability of M-SMP. Compared with the 100% weight loss of SMP, approximate 25% of

213

the stable quality for M-SMP was due to the existence of thermally stable Fe3O4 when the

214

temperature was above 750 oC, which was consistent with the formulation of the material in this

215

study. The specific surface area and pore size of M-SMP were determined by the N2

216

adsorption-desorption isotherms measurement shown in Text S4, Fig S6 and Table S2. The pore

217

diameter between 2 and 50 nm will help to provide faster accessibility between heavy metal ions

218

and functional groups containing N and S atoms in the M-SMP (Ko et al., 2017).

219 220

Fig. 3. High-resolution XPS spectra of N1s (a) and S2p (b). Hysteresis loop of M-SMP and Fe3O4 (c), TGA curves

221

of M-SMP and SMP (d). At the bottom right of (c), the L and R insets are digital photographs of the heavy metal

222

aqueous solution with dispersed M-SMP before and after magnetic separation.

223

3.2 Adsorption experiments 11

224

3.2.1 Effect of pH

225

To avoid the formation of Pb(OH)2 precipitation and the dissociation of hydrogen bonds in

226

M-SMP, adsorption experiments with M-SMP were performed over a wide pH range from 1.0 to

227

6.0 (Nagai et al., 2015; Nassar, 2010). The result of zeta potential measurement showed that the

228

surface charge of M-SMP was negative over the operating pH range, and the effect of different pH

229

on the adsorption performance was also investigated (Fig. 4a). When the pH solution was below

230

2.0, Pb2+ ions were difficult to be immobilized on the M-SMP because of the strong competition

231

effects between H+ and Pb(II). Notably, the adsorption capacity of Pb(II) on M-SMP increased

232

dramatically with the pH ranging from 2.0 to 4.0, which was mainly attributed to the rising

233

electrostatic attraction between the sharply increasing negatively charged M-SMP surface sites and

234

positively charged Pb(II) species. When pH was tuned from 4.0 to 6.0, the adsorption capacity of

235

M-SMP toward Pb(II) decreased slightly. Furthermore, because most industrial wastewater

236

containing heavy metals is considered acidic, the higher adsorption capacity of the as-prepared

237

adsorbent under acidic conditions indicated that the M-SMP has better practical application

238

prospect (Li et al., 2016). Given these conditions, we tested the performance of M-SMP toward

239

Pb(II) at pH 4.0 in the following experiments.

12

240 241

Fig. 4. (a) Effect of pH and Zeta potential (T = 298 K, C0 = 200 mg L–1). (b) Effect of M-SMP dosage on Pb(II)

242

adsorption (pH = 4.0, T = 298 K, C0 = 200 mg L–1). (c) Effect of contact time on Pb(II) uptake onto M-SMP. (d)

243

Adsorption capacities (qe) at different temperatures with initial Pb(II) concentrations from 55 to 4500 mg L–1

244

(pH = 4.0), and (e) adsorption isotherm and removal efficiencies of Pb(II) onto M-SMP (pH = 4.0, T = 298 K). (f)

245

Co-existing heavy metal ions removal performance from a mixed solution of heavy metal ions and the relevant

246

heavy metal ions removal performance individually (pH = 4.0, T = 298 K, C0 = 200 mg g–1). 13

247

3.2.2 Effect of dosage

248

The adsorption capacity of Pb(II) on different M-SMP dosage was displayed in Fig. 4b.

249

Interestingly, the highest adsorption capacity was achieved, and the removal process presented the

250

lowest efficiency when the dosage was 0.2 g L–1. This could be caused by insufficient adsorption

251

sites concerning the abundant target ions. When the dosage varied from 0.2 to 1.2 g L–1, there was

252

an apparent effect on the inhibition of adsorption capacity. Meanwhile, the promotion of removal

253

efficiency was discovered, which could be attributed to the constant amount of target ions and the

254

increasing number of binding sites (Tang et al., 2018). Compared with the fixed amount of metal

255

ions, the available adsorption sites on M-SMP became excessive with continuous increasing

256

dosage (above 1.0 g L–1). Consequently, the removal efficiency increased slightly, and the

257

adsorption capacity decreased sharply. Taking the two factors into account simultaneously, 1.0 g

258

L–1 was the appropriate dosage for the preconcentration of 200 ppm Pb(II).

259

3.2.3 Adsorption kinetics

260

Fig. 4c shows the effect of contact time on Pb(II) adsorption onto M-SMP at 298 K. The

261

adsorption capacity increased significantly from 0 to 189.4 mg g–1 within the first 100 min, and the

262

removal efficiency was up to 95% accordingly. Later, as the time progressed after 100 min, the

263

adsorption amount and the adsorption efficiency maintained high level. The removal of Pb(II) on

264

M-SMP took place in two stages: in the first stage within 100 min, the rapid removal might be

265

attributed to the ample adsorption active sites provided by M-SMP. In the second stage after 100

266

min, the adsorption sites became exhausted, and the adsorption rate was dominated by the

267

transport of Pb(II) onto the M-SMP from the exterior to the interior, resulting in a relatively slow

268

increase stage. Based on the kinetic result, a shaking time of 100 min was considered appropriate

269

in the following experiments.

270

For better insight into the adsorption process, three classical kinetic models, including the

271

pseudo-first-order, pseudo-second-order and intra-particle kinetic models, were used to simulate 14

272

the adsorption data (Text S5). The fitting of the model-estimated result and experimental data was

273

assessed by the linear correlation coefficients (R2). As shown in Fig. S7, the relevant parameters

274

were fitted by the plots of ln(qe-qt) vs t, t/qt vs t and qt vs t1/2, as listed in Table S3. From the

275

intra-particle kinetic model (Fig. S7c), the entire adsorption process of mass transfer had

276

experienced three stages: the instantaneous diffusion from solution to M-SMP surface, the interior

277

surface diffusion stage, and the final equilibrium stage. The R2 value of pseudo-second-order

278

equation for Pb(II) removal was 0.9998, which was higher than that of the other two equations,

279

indicating the pseudo-second-order equation can perfectly describe the kinetics for Pb(II) removal

280

from the M-SMP. Furthermore, the calculated data of qe,exp from the pseudo-second-order model at

281

298 K was 201.5 mg g–1, which was very close to the experimental one, qe,cal (Table S4). Given

282

these findings, the strong surface complexation between Pb(II) and the M-SMP was mainly

283

controlled by chemisorption (Huang et al., 2019; Tang et al., 2018).

284

3.2.4 Effect of initial Pb(II) concentrations and adsorption isotherms

285

The adsorption performance of Pb(II) onto M-SMP was further evaluated by varying initial

286

Pb(II)concentrations from 55 to 4500 mg L–1. As shown in Fig. 4d-e, the adsorption capacities

287

clearly increased with the increasing concentrations, originating from the fact that the driving force

288

provided by the high concentration of metal ions overcame the mass transfer resistance from the

289

solution to solid phases. The isotherm model is an effective tool to better investigate the relevant

290

adsorption equilibrium at various temperatures. Therefore, three representative isotherms

291

(including the Langmuir, Freundlich and Temkin isotherm models) were engaged to depict the

292

adsorption equilibrium Pb(II) onto the M-SMP (shown in Text S6). All the related parameters

293

tabulated in Table S5 can be calculated from the linear plots of ce/qe vs ce, lnqe vs lnce and qe vs

294

lnqe, as shown in Fig. S8-10. According to the R2 values, Langmuir isotherm model matched the

295

adsorption data better than those on other isotherm models. Therefore, this phenomenon implied

296

that the adsorption of Pb(II) on the M-SMP was monolayer adsorption behavior owing to the 15

297

identical adsorption activity on the M-SMP. The calculated maximum adsorption capacity (qm) is

298

946.9 mg g–1 for Pb(II), superior to that of recently reported magnetic materials (Table S6).

299

Furthermore, even at low concentration, there was also a higher equilibrium adsorption capacity,

300

indicating that the M-SMP has excellent adsorption performance in low concentration solution

301

(Fig. 4e) and demonstrating that the as-prepared M-SMP exhibited great potential application

302

prospect in low-concentration wastewater purification.

303

3.2.5 Effect of temperature and thermodynamic analysis

304

To better understand the effect of temperature on the adsorption performance, the

305

thermodynamic studies were carried out, and detailed information is presented in Fig. 4d and Text

306

S7. It is found that no distinct influence was observed below 1400 mg L–1, whereas the adsorption

307

capacities were promoted slightly at a higher temperature when the initial Pb(II) concentrations

308

remained in the range of 1400-4500 mg L–1, indicating that higher temperatures were more

309

favourable for the increasing Pb(II) uptake. In other words, this phenomenon demonstrated the

310

endothermic nature of the adsorption process. The relevant parameters are listed in Table S7.

311

Clearly, all negative standard free energy changes (∆G) values and the decreasing trend along with

312

rising temperature indicated the spontaneity of the adsorption system. The positive enthalpy

313

change (∆H) value demonstrated that the adsorption process of Pb(II) onto M-SMP is an

314

endothermic reaction. Furthermore, the positive value of entropy change (∆S) suggested that the

315

randomness on the interface between Pb(II) ions and M-SMP was increased during the adsorption

316

process.

317

3.2.6 Effect of coexisting ions

318

The effect of various coexisting cations on the adsorption performance was conducted at pH

319

4.0. Five typical kinds of metal ions [Ni(II), Zn(II), Cd(II), Mg(II), and Ca(II)] associated with

320

Pb(II) competition were used to test the specificity of the M-SMP. After adsorption equilibrium,

321

the residual efficiency of Pb(II) individually was 0.60%, and that in the presence of co-existence 16

322

ions is 5.25%, which were both higher than that of other divalent ions (Fig. 4f). The result

323

indicated that the M-SMP exhibited high removal efficiency for Pb(II) individually or with the

324

presence of the competing divalent cations, confirming the high selectivity of the M-SMP toward

325

Pb(II). The phenomenon is attributed to the strong coordination bonds of Pb···N and Pb···S for

326

sharing lone pair electronic (Wang et al., 2019; Wang et al., 2020). In other words, heavy metal

327

ions (the Lewis soft-acid) forming strong interaction with S or N (the Lewis soft-base) plays an

328

important role in the adsorption process.

329

3.3 Regeneration and comparison with commercial adsorbents

330

Easy reusability of the M-SMP is another crucial factor for the wide application of adsorbent

331

to minimize the wastewater treatment cost and reduce secondary pollution to the environment. For

332

the desorption experiment, 0.05 M EDTA-2Na and 0.5 M HCl solutions were used to elute the

333

absorbed Pb(II) ions from the M-SMP, as listed in Table S8. The desorption efficiencies of 0.05 M

334

EDTA-2Na were significantly better than those of 0.5 M HCl solution due to the high chelating

335

ability of EDTA-2Na toward Pb(II). Therefore, 0.05 M EDTA-2Na was selected as the desorbent

336

in the subsequent regeneration experiment. The grey solids were collected by the magnet

337

separation. There was only slightly decrease of Pb(II) adsorption capacities after five cycles of

338

regeneration experiments (Fig. 5a), maintaining a high removal efficiency of 92%. What' more, the

339

saturation magnetization value of M-SMP (Fig. S11) maintained a constant value without obvious

340

decreasing, indicating that the magnetic property of the M-SMP is quite stable. Therefore, the

341

resulting composites had excellent stability, good reproducibility and easy regeneration.

342

17

343

344 345

Fig. 5. (a) Recycles performance of Pb(II) onto M-SMP (pH = 4.0, T = 298 K, C0 = 200 mg L–1), (b) FT-IR

346

spectra of M-SMP before and after Pb(II) adsorption. The specific high-resolution XPS spectra of Pb4f (c),N1s (d)

347

and S2p (e) after Pb(II) adsorption.

348

18

349

Aiming at comparatively evaluating the adsorption performance, the adsorption experiments

350

were performed by using commercial adsorbent including such as silica gel, activated carbon,

351

aluminium oxide, D401 and D402 resins under the same conditions, instead of the synthesized

352

adsorbent in this study (Table S9). In both low and high initial Pb(II) concentrations, the uptake of

353

Pb(II) onto the M-SMP was higher than those of most commercial adsorbents. Similarity, the

354

removal percentage was clearly superior to above commercial adsorbents, especially in the case of

355

high Pb(II) concentration. These results demonstrated that M-SMP composites are a promising

356

candidate for capturing Pb(II) from wastewater.

357

3.4 Plausible adsorption mechanism

358

To reveal the plausible interaction mechanism of Pb(II) with the adsorbent, FT-IR and XPS

359

surveys of the M-SMP before and after adsorption were performed. In FT-IR spectra of Pb(II)

360

loaded M-SMP compared with pristine M-SMP, the represented peaks shifted to lower

361

wavenumbers (Fig. 5b), including C=S (from 1135 cm–1 to 1128 cm–1), C-N (from 1520 cm–1 to

362

1480 cm–1), N-H (from 3418 cm–1 to 3410 cm–1), and thione (from 1633 cm–1 to 1626 cm–1). The

363

complete disappearance of signal (780 cm–1) and the weakness of strong peaks (1128 cm–1, 1626

364

cm–1 and 3410 cm–1) in the M-SMP suggested the coordination bonds of Pb···N and Pb···S (Fu

365

and Huang, 2018; Fu et al., 2019). Furthermore, the good reusability confirmed that the C-S bond

366

was not broken. Fig. 2c showed the full XPS spectrum after Pb(II) adsorption. Compared with the

367

spectrum before adsorption, a new signal of Pb4f after adsorption demonstrated the presence of

368

Pb(II) cation in the adsorbed M-SMP sample. Compared with purified Pb(II) with the binding

369

energies Pb4f 5/2 at 139.6 eV and Pb4f 7/2 at 144.5 eV (Peng et al., 2014), the remarkable binding

370

energy shifts for Pb4f 5/2 (from 139.6 eV to 139.3 eV) and for Pb4f 7/2 (from 144.5 to 142.9 eV) after

371

adsorption demonstrated the strong affinity between functional groups and Pb(II) (Fig. 5c) (Wang

372

et al., 2016). In contrast with the intensity of N1s signal in Fig 3a, the weakness of that after

373

adsorption (Fig. S12a) and the becoming smaller of binding energy value for three peaks suggested 19

374

that the formation of Pb···N (Fig. 5d) (Fu and Huang, 2018). In high-resolution XPS scan of S2p in

375

the aromatic trithiol form after adsorption, the area of S2p 1/2 was significantly increased from

376

19.1% to 21.6%, whereas that of S2p 3/2 was reduced from 16.7% to 6.7% (Fig. 5e, Fig. S12b and

377

Table S10). These phenomena confirmed the strong Pb2+···S2- bonding interactions, which was

378

consistent with previous literatures (Wang et al., 2017; Zhuang et al., 2018). Due to the uniform

379

array of MA and TTCA in the M-SMP, S and N atoms derived from MA and TTCA were

380

homogeneously distributed over the entire M-SMP, providing abundant adsorption active sites for

381

the target heavy metals in all solid-liquid contact interfaces. Therefore, the as-prepared M-SMP

382

possess superior heavy metal capture ability (Fu and Huang, 2018; Xu et al., 2018). In addition to

383

the main role of complexation interaction, electrostatic attraction and ion exchange are also

384

possible causes, according to the effect of pH on adsorption in this study.

385

3.5 Application to real battery industry wastewater

386

To verify the applicability of the M-SMP to Pb(II) extraction from lead-acid battery industry

387

wastewater, we performed the same adsorption experiment as in our previous work (Wang et al.,

388

2019; Wang et al., 2020). Keeping in mind the complexity of real wastewater, it is necessary to

389

explore the effects of various solution pH and adsorbent dosage on the performance of Pb(II)

390

removal. The adsorption results were diverse in acidic conditions (Fig. 6a). Although there is

391

abundant free Pb(II) in acidic condition due to its weak binding ability to bind with the other

392

wastewater components (Table S1), the equilibrium concentration after adsorption was still high,

393

indicating the poor adsorption ability of M-SMP toward Pb(II) at pH = 1.2, which is consistent

394

with the results mentioned previously (Wang et al., 2019; Wang et al., 2020). Considering both the

395

chemical stability of adsorbent and the efficiency of adsorption, the effect of the M-SMP dosage

396

on Pb(II) capture was subsequently performed at pH= 4.0. Fig. 6b shows that the ideal dosage was

397

1.6 g L–1, higher than that on Pb(II) removal in simulated aqueous solution. This phenomenon may

398

be caused by the competition for adsorption active sites between Pb(II) and the other battery 20

399

industry wastewater components (Table S1). Furthermore, a substantially constant mass before and

400

after adsorption was maintained (Table S11).

401 402

Fig. 6. Effect of pH (a) and dosage (b) for Pb(II) removal in lead-acid battery industry wastewater.

403 404

Pleasant results were obtained under desired operating conditions, and the concentration of

405

Pb(II) in wastewater was reduced from 18.070 mg L–1 to 0.091 mg L–1, which was well below the

406

less than 0.70 mg L–1 concentration limit of China’s Emission Standard of Pollutants for Battery

407

Industry (GB 30484-2013). Furthermore, the removal efficiency reached 99.50%. These results

408

show that the as-prepared composites possessed excellent Pb(II) capture ability in real battery

409

industry wastewater and confirm that the M-SMP has potentially low-cost and ultrahigh adsorption

410

ability for Pb(II) cleanup from wastewater.

411

4. Conclusions

412

In conclusion, we successfully fabricated a magnetic supramolecular polymer composite via a

413

two-step method. Based on the systematic characterizations and batch adsorption experimental

414

results, superior adsorption uptake (946.9 mg L–1 to Pb(II) at pH = 4.0, T = 298 K), and favourable

415

reusability make the M-SMP an efficient adsorbent in the treatment of heavy metal pollution. In

416

addition, the accumulation of Pb(II) on the M-SMP was attributed mainly to the formation of

417

Pb···N and Pb···S bonding, which was jointly demonstrated by investigation of FT-IR and

418

analyzing of XPS. Furthermore, the exploration of capturing Pb(II) in lead-acid battery wastewater 21

419

was performed, and desired results were achieved. In consideration of its easy preparation and

420

separation, rapid adsorption rate and ultrahigh Pb(II) adsorption capacity, the as-prepared M-SMP

421

could be a competitive purifier adsorbent in the remediation of industrial wastewater containing

422

heavy metal.

423

Conflicts of interest

424

There are no conflicts to declare.

425

Acknowledgements

426

This work was supported by the NSFC (Grants No. 21677047, U1604137 and 51808200), Science

427

and Technology Key Program of Henan Province (No.172102310698), Natural Science foundation

428

of Henan Province (No.182300410154), the China Postdoctoral Science Foundation (Grant No.

429

2018M630825) and Plan for University Scientific Innovation Talent of Henan Province

430

(19HASTIT046).

431

References

432

Al Hamouz,O.C.S., Adelabu, I.O., Saleh, T.A., 2017. Novel cross-linked melamine based

433

polyamine/CNT composites for lead ions removal. J. Environ. Manage. 192, 163-170.

434

Da'na E., 2017. Adsorption of heavy metals on functionalized-mesoporous silica: A review.

435

Microporo. Mesopor. Mat. 247, 145-157.

436

Dai, Y.H., Liang, Y., Xu, X.Y., Zhao, L., Cao, X.D., 2018. An integrated approach for simultaneous

437

immobilization of lead in both contaminated soil and groundwater: Laboratory test and numerical

438

modeling. J. Hazard. Mater. 342, 107-113.

439

Dinda, D., Saha, S.K., 2015. Sulfuric acid doped poly diaminopyridine/graphene composite to

440

remove high concentration of toxic Cr(VI). J. Hazard. Mater. 291, 93-101.

441

Fan, Q.J., Liu, J.J., Yu, Y.C., Zuo, S.L., Li, B.S., 2017. A simple fabrication for sulfur doped 22

442

graphitic carbon nitride porous rods with excellent photocatalytic activity degrading RhB dye.

443

Appl. Surf. Sci. 391, 360-368.

444

Feng, L.L., Zou, Y.C., Li, C.G., Gao, S., Zhou, L.J., Sun, Q.S., Fan, M.H., Wang, H.J., Wang, D.J.,

445

Li, G.D., Zou, X.X., 2014. Nanoporous sulfur-doped graphitic carbon nitride microrods: A durable

446

catalyst for visible-light-driven H2 evolution. Int. J. Hydrogen. Energ. 39, 15373-15379.

447

Fu, F.L., Wang, Q., 2011. Removal of heavy metal ions from wastewaters: a review. J. Environ.

448

Manage. 92, 407-418.

449

Fu, W., Huang, Z.Q., 2018. One-pot synthesis of a two-dimensional porous Fe3O4/poly(C3N3S3)

450

network nanocomposite for the selective removal of Pb(II) and Hg(II) from synthetic wastewater.

451

ACS Sustainable Chem. Eng. 6, 147851-114794.

452

Fu, W., Wang, X.Y., Huang, Z.Q., 2019. Remarkable reusability of magnetic Fe3O4-encapsulated

453

C3N3S3 polymer/reduced graphene oxide composite: A highly effective adsorbent for Pb and Hg

454

ions. Sci. Total Environ. 659, 895-904.

455

Huang, X., Yang, J.Y., Wang, J.K., Bi, J.T., Xie, C., Hao, H.X., 2018. Design and synthesis of

456

core-shell Fe3O4@PTMT composite magnetic microspheres for adsorption of heavy metals from

457

high salinity wastewater. Chemosphere 206, 513-521.

458

Huang, Y., Xia, S.Y., Lyu, J.J., Tang, J.C., 2019. Highly efficient removal of aqueous Hg2+ and

459

CH3Hg+ by selective modification of biochar with 3-mercaptopropyltrimethoxysilane. Chem. Eng.

460

J. 360, 1646-1655.

461

Ji, J.J., Chen, G., Zhao, J., 2019. Preparation and characterization of amino/thiol bifunctionalized

462

magnetic nanoadsorbent and its application in rapid removal of Pb (II) from aqueous system. J.

463

Hazard. Mater. 368, 255-263. 23

464

Jun, Y.S., Lee, E.Z., Wang, X.C., Hong, W.H., Stucky, G.D., Thomas, A., 2013. From

465

melamine-cyanuric acid supramolecular aggregates to carbon nitride hollow spheres. Adv. Funct.

466

Mater. 23, 3661-3366.

467

Kim, H., Lee, J.P., Ahn, H.M., Kim, O., Park, M.J., 2015. Synthesis of three-dimensionally

468

interconnected sulfur-rich polymers for cathode materials of high-rate lithium-sulfur batteries. Nat.

469

Commun. 6, 1-9.

470

Ko D., Lee, J.S., Patel, H.A., Jakobsen, M.H., Hwang, Y.H., Yavuz, C.T., Hansen, H.C.B.,

471

Andersen, H.R., 2017. Selective removal of heavy metal ions by disulfide linked polymer

472

networks. J. Hazard. Mater. 332, 140-148.

473

Ko, D., Mines, P.D., Jakobsen, M.H., Yavuz, C.T., Hansen, H.C.B., Andersen, H.R., 2018.

474

Disulfide polymer grafted porous carbon composites for heavy metal removal from stormwater

475

runoff. Chem. Eng. J. 348, 685-692.

476

Kucharski, M., Szukiewicz, B.C., 2000. Reactions of trithiocyanuric acid with oxiranes. I.

477

synthesis of polyetherols. J. Appl. Polym. Sci. 76, 439-445.

478

Li, P.G., Wang, J.X., Li, X.T., Zhu, W.J., He, S.F., Han, C.Y., Luo, Y.M., Ma, W.H., Liu, N.S.,

479

Dionysiou, D.D., 2019. Facile synthesis of amino-functional large-size mesoporous silica sphere

480

and its application for Pb2+ removal. J. Hazard. Mater. 378, 120664.

481

Li, X.P., Bian, C.Q., Meng, X.J., Xiao, F.S., 2016. Design and synthesis of an efficient nanoporous

482

adsorbent for Hg2+ and Pb2+ ions in water. J. Mater. Chem. A 4, 5999-6005.

483

Lin, G., Wang, S.X., Zhang, L.B., Hu, T., Cheng, S., Fu, L.K., Xiong, C., 2019. Enhanced and

484

selective adsorption of Hg2+ to a trace level using trithiocyanuric acid-functionalized corn bract.

485

Environ. Pollut. 244, 938-946. 24

486

Nagai, D., Kuribayashi, T., Tanaka, H., Morinaga, H., Uehara, H., Yamanobe, T., 2015. A facile,

487

selective, high recovery system for precious metals based on complexation between melamine and

488

cyanuric acid. RSC Adv. 5, 30133-30139.

489

Nagai, D.S., Nagashima, A., Mori, M., 2016. A facile and high-recovery system for palladium(II)

490

Ion based on complexation between trithiocyanuric acid and melamine. Chem. Lett. 45, 1165

491

-1167.

492

Nassar, N.N., 2010. Rapid removal and recovery of Pb(II) from wastewater by magnetic

493

nanoadsorbents. J. Hazard. Mater. 184, 538-546.

494

Peng, Q.M., Guo, J.X., Zhang, Q.R., Xiang, J.Y., Liu, B.Z., Zhou, A.G., Liu, R.P., Tian, Y.J., 2014.

495

Unique lead adsorption behavior of activated hydroxyl group in two-dimensional titanium carbide.

496

J. Am. Chem. Soc. 136, 4113-4116.

497

Ranganathan, A., Pedireddi, V.R., Rao C.N.R., 1999. Hydrothermal synthesis of organic channel

498

structures: 1:1 hydrogen-bonded adducts of melamine with cyanuric and trithiocyanuric acids. J.

499

Am. Chem. Soc. 121, 1752-1753.

500

Shalom, M., Inal, S., Fettkenhauer, C., Neher, D., Antonietti, M., 2013. Improving carbon nitride

501

photocatalysis by supramolecular preorganization of monomers. J. Am. Chem. Soc. 135,

502

7118-7121.

503

Sharahi, F.J., Shahbazi, A., 2017. Melamine-based dendrimer amine-modified magnetic

504

nanoparticles as an efficient Pb(II) adsorbent for wastewater treatment: Adsorption optimization by

505

response surface methodology. Chemosphere 189, 291-300.

506

Song, S., Zhang, S., Huang S.Y., Zhang, R., Yin, L., Hu, R.Z., Wen, T., Zhuang, L., Hu, B.W.,

507

Wang, X.K., 2019. A novel multi-shelled Fe3O4@MnOx hollow microspheres for immobilizing 25

508

U(VI) and Eu(III). Chem. Eng. J. 355, 697-709.

509

Lapwanit, S., Trakulsujaritchok, T., Nongkhai, P.N., 2016. Chelating magnetic copolymer

510

composite modified by click reaction for removal of heavy metal ions from aqueous solution.

511

Chem. Eng. J. 289 286-295.

512

Tang, N., Niu, C.G., Li, X.T., Liang, C., Guo, H., Lin, L.S., Zheng, C.W., Zeng, G.M., 2018.

513

Efficient removal of Cd2+ and Pb2+ from aqueous solution with amino- and thiol-functionalized

514

activated carbon: Isotherm and kinetics modeling. Sci. Total Environ. 635, 1331-1313 1344.

515

Tian, H., He, J.H., Hu, M.H., 2019. A selectivity-controlled adsorbent of molybdenum disulfide

516

nanoshe armed with superparamagnetism for rapid capture of mercury ions. J. Colloid Interf. Sci.

517

551, 251-260.

518

Venkateswarlu, S., Yoon, M.Y., 2015. Core-shell ferromagnetic nanorod based on amine polymer

519

composite (Fe3O4@DAPF) for fast removal of Pb(II) from aqueous solutions. ACS Appl. Mater.

520

Inter. 7, 25362-25372.

521

Wang, J., Zhang, W.T., Yue, X.Y., Yang, Q.F., Liu, F.B., Wang, Y.R., Zhang, D.H., Li, Z.H., Wang,

522

J.L., 2016. One-pot synthesis of multifunctional magnetic ferrite–MoS2–carbon dot nanohybrid

523

adsorbent for efficient Pb(II) removal. J. Mater. Chem. A. 4, 3893-3900.

524

Wang, J., Wang, P.Y., Wang, H.H., Dong, J.F., Chen, W.Y., Wang, X.X., Wang, S.H., Hayat, T.,

525

Alsaedi, A., Wang, X.K., 2017. Preparation of molybdenum disulfide coated Mg/Al layered double

526

hydroxide composites for efficient removal of chromium(VI). ACS Sustain. Chem. Eng. 5,

527

7165-7174.

528

Wang, Q.W., Dong, S.Y., Zhang, D., Yu, C.F., Lu, J., Wang, D., Sun, J.H., 2017. Magnetically

529

recyclable visible-light-responsive MoS2@Fe3O4 photocatalysts targeting efficient wastewater 26

530

treatment. J. Mater. Sci. 53, 1135-1147.

531

Wang, R., Xu, Y., Wang, R.J., Wang, C.Y., Zhao, H.Z., Zheng, X.Q., Liao, X., Cheng, L., 2016. A

532

microfluidic chip based on an ITO support modified with Ag-Au nanocomposites for SERS based

533

determination of melamine. Microchim Acta 184, 279-287.

534

Wang, Z.W., Wu, Q., Zhang, J., Zhang, H., Feng, J.L., Dong, S.Y., Sun, J.H., 2019. In situ

535

polymerization of magnetic graphene oxide-diaminopyridine composite for the effective

536

adsorption of Pb(II) and application in battery industry wastewater treatment. Environ. Sci. Pollut.

537

R., https://doi.org/10.1007/s11356-019-06511-1.

538

Wang, Z.W., Zhang, J., Wen, T., Liu, X.L., Wang, Y.F., Yang, H.Y., Sun, J.Y., Feng, J.L., Dong,

539

S.Y., Sun, J.H., 2020. Highly effective remediation of Pb(II) and Hg(II) contaminated wastewater

540

and

541

https://doi.org/10.1016/j.scitotenv.2019.134341.

542

Wen J., Fang Y., Zeng G.M., 2018. Progress and prospect of adsorptive removal of heavy metal

543

ions from aqueous solution using metal-organic frameworks: A review of studies from the last

544

decade. Chemosphere 201, 627-643.

545

Whitesides, G.M., Mathias, J.P., Seto, C.T., 1991. Molecular self-assembly and nanochemistry: A

546

chemical strategy for the synthesis of nanostructures. Science 254, 1312-1319.

547

Whitesides, G.M., Simanek, E.E., Mathias, J.P., Seto, C.T., Chin, D.N., Mammen, M., Gordon,

548

D.M., 1995. Noncovalent synthesis: Using physical-organic chemistry to make aggregates. Acc.

549

Chem. Res. 28, 37-44.

550

Wu, Q., Cui, Y.R., Li, Q.L., Sun, J.Y., 2015. Effective removal of heavy metals from industrial

551

sludge with the aid of a biodegradable chelating ligand GLDA. J. Hazard. Mater. 283, 748-754.

soil

by

flower-like

magnetic

MoS2

27

nanohybrid.

Sci.

Total

Environ.,

552

Xu, W.B., Song, Y., Dai, K., Sun, S., Liu, G.Y., Yao, J.R., 2018. Novel ternary nanohybrids of

553

tetraethylenepentamine and graphene oxide decorated with MnFe2O4 magnetic nanoparticles for

554

the adsorption of Pb(II). J. Hazard. Mater. 358, 337-345.

555

Yadav, V.B., Gadi, R., Kalra, S., 2019. Clay based nanocomposites for removal of heavy metals

556

from water: A review. J. Environ. Manage. 232, 803-817.

557

Yang, X.D., Wan, Y.S., Zheng, Y.L., He, F., Yu, Z.B., Huang, J., Wang, H.L., OK, Y.S., Jiang, Y.S.,

558

Gao, B., 2019. Surface functional groups of carbon-based adsorbents and their roles in the removal

559

of heavy metals from aqueous solutions: A critical review. Chem. Eng. J. 366, 608-621.

560

Yang, Z.K., Lin, L., Liu, Y.N., Zhou, X., Yuan, C.Z., Xu, A.W., 2016. Supramolecular

561

polymers-derived nonmetal N,S-codoped carbon nanosheets for efficient oxygen reduction

562

reaction. RSC Adv. 6, 52937-52944.

563

Zhang, C.M., You, T.T., Yang, N., Gao, Y.K., Jiang, L., Yin, P.G., 2019. Hydrophobic paper-based

564

SERS platform for direct-droplet quantitative determination of melamine. Food Chem. 287,

565

363-368.

566

Zhuang, Y.T., Zhang, X., Wang, D.H., Yu, Y.L., Wang, J.H., 2018. Three-dimensional molybdenum

567

disulfide/graphene hydrogel with tunable heterointerfaces for high selective Hg(II) scavenging. J.

568

Colloid Interf. Sci. 514, 715-722.

28

Highlights: 

Magnetic supramolecular polymer was fabricated by facile hydrothermal strategy.



As-synthesized composites demonstrated ultrahigh adsorption capacity toward Pb(II).



High-performance was mainly account of coordination bonds of Pb···N and Pb···S.



Application in real Pb(II)-contaminated battery wastewater were performed.

Declaration of interests ☑ The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. ☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: