You are on page 1of 31

Accepted Manuscript

Determination of triazole pesticides in aqueous solution based on


magnetic graphene oxide functionalized MOF-199 as solid phase
extraction sorbents

Guangyang Liu, Lingyun Li, Xiaodong Huang, Shuning Zheng, Donghui


Xu, Xiaomin Xu, Yanguo Zhang, Huan Lin

PII: S1387-1811(18)30267-1
DOI: 10.1016/j.micromeso.2018.05.023
Reference: MICMAT 8925

To appear in: Microporous and Mesoporous Materials

Received Date: 20 June 2017


Revised Date: 10 May 2018
Accepted Date: 14 May 2018

Please cite this article as: G. Liu, L. Li, X. Huang, S. Zheng, D. Xu, X. Xu, Y. Zhang, H.
Lin, Determination of triazole pesticides in aqueous solution based on magnetic
graphene oxide functionalized MOF-199 as solid phase extraction sorbents, Microporous
and Mesoporous Materials (2018), doi: 10.1016/j.micromeso.2018.05.023.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a
service to our customers we are providing this early version of the manuscript. The
manuscript will undergo copyediting, typesetting, and review of the resulting proof before
it is published in its final form. 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.
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
1 Determination of triazole pesticides in aqueous solution based on

2 magnetic graphene oxide functionalized MOF-199 as solid phase

3 extraction sorbents
‡ ‡ *
4 Guangyang Liu , Lingyun Li , Xiaodong Huang, Shuning Zheng, Donghui Xu ,

5 Xiaomin Xu, Yanguo Zhang, Huan Lin

6 Institute of Vegetables and Flowers, Chinese Academy of Agricultural Sciences, Key

7 Laboratory of Vegetables Quality and Safety Control, Ministry of Agriculture of

8 China, Beijing 100081, P. R. China



9 Co-authors with equal contributions.
*
10 Corresponding author: Prof. Donghui Xu, E-mail: xudonghui@caas.cn; Tel.: +86
11 010 82106963; fax: +86 010 62146160.
12

13 Abstract: A novel magnetic copper-based metal organic framework

14 (M-MOF-199) has been successfully prepared using Fe3O4–graphene

15 oxide (GO) nanocomposite for solid phase extraction of five triazole

16 pesticides from environment water samples. High-performance liquid

17 chromatography coupled with tandem mass spectrometry was used for

18 sample quantification and determination. The results suggest that Fe4O3–

19 GO is a thin single-layer structure anchoring Fe3O4 and M-MOF-199 is

20 coated on its surface. Under the optimized conditions, the limits of

21 detection of the triazole pesticides are 0.05–0.1 ( S/N = 3) with correlation

22 coefficients larger than 0.992. The recoveries in spiked water samples are

23 72.3%–91.53% with relative standard deviations rang ing from 1.5% to

24 9.1%. Furthermore, the results show that M-MOF-199 is a promising


ACCEPTED MANUSCRIPT

25 hybrid adsorbent for magnetic dispersive solid phase extraction and

26 removal of triazole pesticides from environmental water samples.

27 Keywords: metal organic framework; magnetic graphene oxide; triazole

28 pesticide; solid phase extraction; determination

29

30 1. Introduction

31 Triazole pesticides are a type of broad-spectrum bactericidal

32 fungicide, and they are widely used to protect crops, vegetables, and fruits

33 [1, 2]. Because triazole pesticides are moderately lipophilic, highly stable,

34 and have long chemical and photochemical half-lives, they can be easily

35 transported and accumulate in environmental soil and water [3, 4].

36 Moreover, triazole pesticides will affect human health by disturbing

37 endocrine activity [5, 6]. Therefore, an effective and sensitive detection

38 method to analyze trace levels of triazole pesticides in the environment is

39 urgently required.

40 In recent years, several methods have been used to determine

41 triazole pesticides in environmental media and food samples, including

42 gas chromatography (GC) [7, 8], high-performance liquid

43 chromatography (HPLC) [9, 10], and liquid chromatography (LC)–

44 tandem mass spectrometry (MS) (LC–MS/MS) [11, 12]. Despite their

45 high sensitivity, selectivity, and efficiency, most of these methods are

46 vulnerable to matrix interference from the analytical sample [13, 14].


ACCEPTED MANUSCRIPT

47 Sample preparation techniques that enrich and pre-concentrate the

48 analytes in the sample matrix are important [15, 16].

49 Common sample preparation techniques include liquid–liquid

50 extraction, solid phase extraction, solid phase microextraction, and

51 magnetic solid phase extraction (MSPE) [17-19]. MSPE is a simple, rapid,

52 centrifugation-free (or filtration-free), and environmentally friendly

53 technique for pesticide pretreatment by mechanical separation [20]. A

54 magnetic sorbent is added into the solution for rapid adsorption of the

55 target, and it is then eluted from the magnetic sorbent for further analysis

56 [21].

57 Magnetic metal organic frameworks (M-MOFs) are one of the most

58 important magnetic sorbent materials. They have recently attracted much

59 interest and been extensively investigated in pesticide–MSPE research

60 [22, 23]. MOFs are a new type of three-dimensional crystalline sorbent

61 material formed by coordinating metal ions with organic ligands [24].

62 Owing to their tunable porosity, structure flexibility, high surface area,

63 and adsorption capacity, MOFs are used for drug delivery, catalysis, gas

64 storage, sensors, and separation [25, 26]. Many functionalized

65 nanomaterials are used for magnetic sorbent preparation, such as metal

66 nanoparticles, silica nanoparticles, quantum dots, polymers, multiwalled

67 carbon nanotubes, and graphene oxide (GO) [27, 28]. Among these

68 materials, GO has shown great potential to enhance the selectivity and


ACCEPTED MANUSCRIPT

69 adsorption capacity of MOFs by forming unique reactive adsorption

70 centers [29, 30]. Magnetic GO composites show high adsorption, rapid

71 separation, controllable magnetic properties, high structure stability, and

72 water-resistant ability [31]. However, little research has been performed

73 on the use of magnetic GO functionalized MOF composites for

74 adsorption and removal of triazole pesticides from environmental water

75 samples.

76 In this study, under the assumption that the benzene rings and five-

77 membered heterocycles in triazole pesticides may form hydrogen bonds,

78 hydrophobic interactions, electrostatic interactions, or π-π

79 stacking/interactions with GO–MOFs [32], we aimed t o combine the

80 benefits of M–GO and MOFs for rapid separation and highly selective

81 adsorption of triazole pesticides. A novel magnetic copper-based MOF

82 based on M–GO (M-MOF-199) was prepared and characte rized in detail.

83 The static adsorption capacity and mechanism was preliminarily

84 investigated and the MSPE procedure was optimized. The obtained

85 material was then investigated for extraction of triazole pesticides from

86 spiked water samples by MSPE followed by HPLC coupled with tandem

87 MS (HPLC–MS/MS) determination.

88

89 2. Experimental section

90 2.1 Chemicals and materials


ACCEPTED MANUSCRIPT

91 Flusilazole (molecular size 329.76, pKa 3.87), fenbuconazole

92 (molecular size 336.82, pKa 3.79), myclobutanil (molecular size 288.78,

93 pKa 2.89), penconazole (molecular size 284.18, pKa 3.72), and

94 epoxiconazole (molecular size 329.76, pKa 3.3) were obtained from

95 Sigma-Aldrich (St Louis, MO, USA). GO and H3BTC were purchased

96 from Aladdin Industrial Corporation (Shanghai, China). FeCl3·6H2O,

97 FeCl2·4H2O, and Cu(CH3COO)·H2O were purchased from Sinopharm

98 Chemical Reagent Co. Ltd. All other materials were of analytical reagent

99 grade and purchased from the Beijing Chemical Reagent Factory (Beijing,

100 China).

101 2.2 Apparatus

102 The size and morphology of M-MOF-199 were observed by

103 transmission electron microscopy (TEM) using a JEM-200CX

104 transmission electron microscope (JEOL, Japan) and scanning electron

105 microscopy (SEM) using a JSM-6300 scanning electron microscope

106 (JEOL, Japan). Fourier transform infrared (FT-IR) spectroscopy was

107 performed with a FT-IR-8400 spectrometer (Shimadzu, Japan). X-ray

108 powder diffraction (XRD) was performed with a D8 Advance X-ray

109 powder diffractometer (Bruker, Germany). The magnetic properties

110 were investigated using a Lake Shore 7410 vibrating sample

111 magnetometer (VSM, USA). A Shimadzu HPLC–MS/MS inst rument

112 (LC-30A, MS8050, Japan) equipped with an Eclipse XDB-C18 column


ACCEPTED MANUSCRIPT

113 (2.1 mm × 150 mm, 3.5 µm) was used for determination of the triazole

114 pesticides.

115 2.3 Preparation of Fe3O4–GO

116 Fe3O4–GO nanoparticles were obtained according to a prev iously

117 reported method [33] with the following procedure. GO (0.25 g) was

118 dispersed in ultrapure water (240 mL) in a three-necked flask by

119 ultrasonic dispersion for 1 h. FeCl3·6H2O (1.8 g) and FeCl2·4H2O (0.8 g)

120 were then added into the GO solution under N2 protection. The mixture

121 was heated at 80 °C with vigorous mechanical stirri ng for 1 h. After 1 h,

122 10 mL of 28% ammonium hydroxide was added and the mixture was

123 vigorously stirred for another 30 min. Finally, the Fe3O4–GO

124 nanoparticles were separated by an external magnetic field, washed

125 several times with ethanol and water, and then dried in a vacuum oven at

126 60 °C for 24 h.

127 2.4 Preparation of M-MOF-199

128 M-MOF-199 was prepared according to a reported method [34]. The

129 procedure was as follows. Fe3O4–GO nanoparticles (0.2 g) were dispersed

130 in ethanol solution (30 mL) containing 20 mM mercaptoacetic acid with

131 vigorous shaking for 1 h. The black precipitate was then magnetically

132 separated and washed with ultrapure water and ethanol. The Fe3O4–GO

133 nanoparticles were dispersed in dimethylformamide (DMF)/ethanol

134 solution (150 mL, v/v, 1/1) containing H3BTC (785 mg) with constant
ACCEPTED MANUSCRIPT

135 stirring for 30 min. Subsequently, ultrapure water (75 mL) containing

136 Cu(CH3COO)2·H2O (1425 mg) was added and the solution was stirred at

137 room temperature for 3 h. Triethylamine (1.0 mL) was then added

138 dropwise under stirring for another 3 h. The green precipitate was

139 collected by magnetic separation and washed several times with DMF

140 and ethanol to remove unreacted chemicals. Finally, M-MOF-199 was

141 dried in a vacuum oven at 60 °C for 24 h.

142 2.5 Adsorption experiments

143 A stock methanol solution containing 50 mg/L of five triazole

144 pesticides was prepared. The working solutions with specific

145 concentrations were prepared by diluting the stock solution with ultrapure

146 water.

147 To investigate the adsorption capacity of M-MOF-199, 20 mg of

148 M-MOF-199 was added to 4 mL of aqueous solutions containing the five

149 triazole pesticides (flusilazole, fenbuconazole, myclobutanil, penconazole,

150 and epoxiconazole) at various concentrations (200, 2000, or 5000 µg/L).

151 The supernatant was obtained and diluted with ultrapure water for further

152 quantitative detection by HPLC–MS/MS.

153 The static binding capacity of M-MOF-199 for triazole pesticide

154 adsorption was calculated using the follows equation:

155 Qe = (C0 − Ce) × V/M


ACCEPTED MANUSCRIPT

156 where Qe is the static binding capacity (µg/mg), C0 is the initial triazole

157 pesticide concentration (µg/L), Ce is the equilibrium triazole pesticide

158 concentration (µg/L), V is the total volume of the initial triazole pesticide

159 solution (L), and M is the mass of M-MOF-199 (mg).

160 2.6 Optimization for MSPE

161 The effect of the amount of M-MOF-199 on extraction was

162 investigated in the range 5–30 mg. Different amount of M-MOF-199 was

163 added to 4 mL of aqueous solutions containing the five triazole pesticides

164 at 2000 µg/L. After incubating for 15 min, the supernatant was obtained

165 and diluted with ultrapure water for further quantitative detection by

166 HPLC–MS/MS.

167 Extraction time profiles of the triazole pesticides for extraction times

168 ranging from 5 to 30 min were studied. 4 mL of 2000 µg/L triazole

169 pesticides solution were incubated with 20 mg M-MOF-199 for different

170 time in a 15 mL centrifuge tube. Then, HPLC–MS/MS w as used to

171 determine the amount of triazole pesticides in the supernatant.

172 Several solvents (methanol, acetonitrile, acetone, and acetonitrile–

173 acetic acid (v/v, 9:1)) were investigated for desorption of triazole

174 pesticides and to obtain the highest recovery for the MSPE approach. 20

175 mg of M-MOF-199 was mixed with 4 mL of aqueous solutions

176 containing the five triazole pesticides at 2000 µg/L for 15 min.
177 M-MOF-199 was collect and then ultrasonically treated with different
ACCEPTED MANUSCRIPT

178 solvents for 10 min to eluate the triazole pesticides.

179 2.7 MSPE of the water samples

180 First, the tap and well water samples were centrifuged and filtered

181 through a 0.22 µm Millipore syringe filter to remov e insoluble

182 substances. The water samples were then spiked with triazole pesticides to

183 achieve final triazole concentrations of 100, 200, or 500 µg/L. The spiked

184 tap water samples (4 mL) were incubated with 20 mg M-MOF-199 for 15

185 min in a 15 mL centrifuge tube. M-MOF-199 was then collect by a

186 magnet and the supernatant was removed. Subsequently, acetone (3 mL)

187 was added into the centrifuge tube and the mixture was ultrasonically

188 treated for 10 min to eluate the triazole pesticides from M-MOF-199. A

189 sample of the eluent solution (1 mL) was collected and transferred to a 5

190 mL centrifuge tube. The solution was concentrated using a mild nitrogen

191 stream and the residues were diluted to 200 µL. After filtering through a

192 0.22 µm Millipore syringe filter, 5 µL of the solution was analyzed by

193 HPLC–MS/MS.

194

195 3. Results and discussion

196 3.1 Characterization of M-MOF-199

197 The morphologies of F Fe3O4, Fe3O43–GO, and M-MOF-199 were

198 investigated by TEM and SEM. TEM images of Fe3O4 and Fe3O4–GO,

199 and a SEM image of M-MOF-199 are shown in Fig. 1. The average
ACCEPTED MANUSCRIPT

200 diameter of Fe3O4 microspheres is about 20 nm with a uniform size

201 distribution. Fig. 1(b) shows that F Fe3O4–GO has a smooth surface and

202 sheet structure with a mean diameter of 150 nm. Furthermore, Fe3O4

203 nanoparticles are coated on the GO surface. The TEM image of

204 M-MOF-199 (Fig. 1(c)) shows a relatively uniform porous structure.

205 Moreover, the shape and crystal structure of the Cu–BTC MOFs are

206 preserved. However, the MOFs do not have sharp edges and they are

207 even rounded at the corners. This special structure may improve the

208 triazole pesticide adsorption capacity of M-MOF-199.

209 The magnetic properties of Fe3O4, Fe3O4–GO, and M-MOF-199

210 were investigated at room temperature using a VSM, and the results are

211 shown in Fig. 2. Because both the remanence and coercivity values of all

212 of the samples are zero, the samples are superparamagnetic, which means

213 that M-MOF-199 will disperse well in aqueous solution and can be

214 rapidly separated under an external magnetic field. The saturated

215 magnetization values of Fe3O4, Fe3O4–GO, and M-MOF-199 are 68.25,

216 51.32, and 11.16 emu/g, respectively.

217 The porosity of M-MOF-199 was confirmed by determining the N2

218 adsorption–desorption isotherm at 77 K, as shown in Fig. 3a. The

219 isotherm shows some N2 adsorption and slight hysteresis of the

220 desorption curve at medium relative pressures (0.4 < P/P0 < 0.8), which

221 indicates the presence of mesopores. Moreover, the isotherm also shows
ACCEPTED MANUSCRIPT

222 an almost vertical increase at high relative pressures (0.8 < P/P0 < 1.0),

223 which confirms the presence of macropores. The Brunauer–Emmett–

224 Teller surface area and pore volume of M-MOF-199 are 182.6 m 2 g−1 and

225 0.841 mL g−1 , respectively. In addition, the distribution of pore sizes is

226 also presented in Fig. 3b. The results show that M-MOF-199 has high

227 surface area and large pore volume, which would enable the absorbent to

228 have a high triazole pesticide adsorption capacity.

229 The XRD spectra of Fe3O4, Fe3O4–GO, and M-MOF-199 are shown

230 in Fig. 4. The characteristic peaks of Fe3O4 are observed for all of the

231 samples, and they agree well with the diffraction data of Fe3O4 (JCPDS

232 card: 19-629). The diffraction peak at 2θ = 26.2° is the characteristic peak

233 of GO, which is also seen in the patterns of Fe3O4–GO and M-MOF-199.

234 Moreover, the diffraction peaks of the Cu–BTC MOF a t 10.5°, 13.7°,

235 15.2°, and 21.6° are in accordance with the literat ure. The results suggest

236 that all of the samples are highly crystalline materials and the crystalline

237 structure remains the same during the modification process.

238 Fig. 5 shows the FT-IR spectra of Fe3O4, Fe3O4–GO, and

239 M-MOF-199. A peak at 579 cm−1 attributed to the Fe–O stretching

240 vibration is observed in all of the spectra, which suggests that Fe3O4 was

241 successfully deposited on GO and M-MOF-199. In the spectrum of

242 Fe3O4–GO, the characteristic peaks of GO at 1223 cm −1 and 1035 cm−1

243 for the C–O stretching vibration and at 1622 cm −1 for C=C confirm
ACCEPTED MANUSCRIPT

244 successful preparation of Fe3O4–GO. From the spectrum of M-MOF-199,

245 the characteristic peaks of Fe3O4–GO are observed, and the typical peaks

246 at 1380 and 1502 cm−1 are attributed to the C=O group and benzene ring

247 substitution group of H3BTC, respectively. These results indicate that

248 M-MOF-199 was successfully prepared on the surface of Fe3O4–GO.

249 3.2 Adsorption capacity study

250 The static binding capacities of M-MOF-199 for triazole pesticide

251 adsorption were investigated and the result was shown in Fig. 6. For an

252 initial concentration of 5000 µg/L, the adsorption capacities of flusilazole,

253 fenbuconazole, myclobutanil, penconazole, and epoxiconazole are 3.56,

254 3.42, 3.24, 3.74, and 3.48 µg/mg, respectively. The reason for the removal

255 of these pesticides might be because there are some nitrogen-containing

256 groups, hydrophobic groups, and delocalized large π bonds from both

257 benzene rings and five-membered heterocycles in these five triazole

258 pesticides. Hydrogen bonding, hydrophobic interactions, electrostatic

259 interactions, and π– π stacking between M-MOF-199 and the above

260 functional groups enable M-MOF-199 to remove triazole pesticides.

261 3.3 Optimization of the conditions for MSPE of the triazole pesticides

262 The MSPE performance for the five triazole pesticides is affected by

263 several parameters, such as the amount of M-MOF-199, extraction time,

264 and desorption solvent. In this section, we optimize these factors.

265 3.3.1 Effect of the amount of M-MOF-199


ACCEPTED MANUSCRIPT

266 The effect of the amount of M-MOF-199 on extraction was

267 investigated to achieve satisfactory recovery. As shown in Fig. 7, the

268 extraction efficiency of the triazole pesticides rapidly increases with

269 increasing amount of M-MOF-199 when the amount of M-MOF- 199 is in

270 the range 5–15 mg. The extraction efficiency then s lightly increases with

271 increasing amount of M-MOF-199 as the amount of sorbent changes from

272 20 to 30 mg, which indicates that the triazole pesticides adequately

273 adsorb on M-MOF-199 when the amount is 20 mg. Therefore, 20 mg of

274 M-MOF-199 was used for extraction and MSPE.

275 3.3.2 Effect of the extraction time

276 MSPE is an equilibrium-based solid phase extraction process and

277 therefore the extraction time will influence the extraction recovery of the

278 sorbent. Extraction time profiles of the triazole pesticides for extraction

279 times ranging from 5 to 30 min are shown in Fig. 8. The extraction

280 recovery rapidly increases when the time increases from 5 to 15 min.

281 When the contact time is 15 min, the recoveries of all of the triazole

282 pesticides reach their equilibrium, which suggests that 15 min is

283 sufficient for triazole pesticide extraction. Therefore, 15 min was chosen

284 for the MSPE procedure.

285 3.3.3 Effect of the desorption solvent

286 The type of desorption solvent can affect the desorption efficiency of

287 triazole pesticides adsorbed on M-MOF-199. Several solvents (methanol,


ACCEPTED MANUSCRIPT

288 acetonitrile, acetone, and acetonitrile–acetic acid (v/v, 9:1)) were

289 investigated for desorption of triazole pesticides and to obtain the highest

290 recovery for the MSPE approach. As shown in Fig. 9, under the same

291 extraction and elution conditions, acetone shows the best desorption

292 efficiency for the triazole pesticides. Therefore, acetone was selected as

293 the desorption solvent for the further experiments.

294 3.4 Method validation

295 To investigate the performance of the developed method for triazole

296 pesticide detection, a series of aqueous solution containing the five

297 triazole pesticides at various concentrations were prepared for calibration.

298 The quantitative analytical parameters of the method, such as the linear

299 range, regression equation, correlation coefficient (r2), and limit of

300 detection (LOD) are summarized in Table 1. Good linearity is achieved

301 for triazole pesticide concentrations of 0.25–1000 µg/L with correlation

302 coefficients larger than 0.992 and LODs of 0.05–0.1 µg/L (S/N = 3). The

303 total ion chromatograms of the standard solution with triazole pesticide

304 concentrations of 200 µg/L are shown in Fig. 10.

305 3.5 Analysis of triazole pesticides in water samples

306 The developed method was used to determine triazole pesticides in

307 spiked tap water and well water samples by HPLC–MS/ MS to validate

308 the applicability of the method. As shown in Table 2, the triazole

309 pesticide recoveries from spiked 100, 200, and 500 µg/L tap water
ACCEPTED MANUSCRIPT

310 samples are 72.3%–87.93% with relative standard dev iations (RSDs) in

311 the range 1.5%–9.1%. The recoveries from spiked 100 , 200, and 500

312 µg/L well water samples are 73.74%–91.53% with RSDs in t he range

313 1.9%– 7.9%.

314

315 4. Conclusions

316 A novel magnetic copper-based MOF (M-MOF-199) has been

317 successfully synthesized using Fe3O4–GO nanocomposite as the magnetic

318 core and the support. M-MOF-199 was characterized and used as a

319 sorbent for magnetic dispersive solid phase extraction of five triazole

320 pesticides from tap and well water samples. Combining HPLC–MS/MS

321 with M-MOF-199, we have established a simple, rapid, and sensitive

322 detection method for triazole pesticides. The high surface area and pore

323 volume endow M-MOF-199 with high adsorption capacity for triazole

324 pesticides. Moreover, the current MSPE method based on M-MOF-199 is

325 easy to operate without filtration procedures or centrifugation. The

326 established method exhibits satisfactorily LODs, recoveries, and

327 acceptable RSDs, which suggests that M-MOF-199 is promising for

328 extraction of triazole pesticides and other analytes from environmental

329 aqueous solutions.

330

331 Acknowledgments
ACCEPTED MANUSCRIPT

332 This work was supported by the National Natural Science Foundation

333 of China (No. 31701695), the project of risk assessment on vegetable

334 products (GJFP2017002) and the project of risk assessment on plant

335 growth regulator of fruit and vegetables (GJFP201701401).

336

337 References

338 [1] É. A. S. Silva, V. Lopez-Avila and J. Pawliszy n, J. Chromatogr. A.

339 1313(2013) 139-146.

340 [2] Y. Li, J. Zhang, B. Peng, S. Li, H. Gao and W. Zhou, Anal. Methods.

341 5(2013) 2241-2248.

342 [3] H. Wang, X. Yang, L. Hu, H. Gao, R. Lu, S. Zhang and W. Zhou,

343 New. J. Chem. 40(2016) 4696 -4704.

344 [4] Q. Wei, Z. Song, J. Nie, H. Xia, F. Chen, Z. Li and M. Lee, J. Sep.

345 Sci. 39(2016) 4603-4609.

346 [5] M. M. Abolghasemi, S. Hassani and M. Bamorowat, Microchim.

347 Acta. 183(2016) 889-895.

348 [6] Z. Zhang, W. Jiang, Q. Jian, W. Song, Z. Zheng, D. Wang and X.

349 Liu, Food Chem. 168(2015)

350 396-403.

351 [7] M. A. Farajzadeh, S. Sheykhizadeh and P. Khorram, Food

352 Anal.Methods. 7(2014) 1229-1237.

353 [8] J. Nie, F. Chen, Z. Song, C. Sun, Z. Li, W. Liu and M. Lee, Anal.
ACCEPTED MANUSCRIPT

354 Bioanal. Chem. 408(2016) 7461-7471.

355 [9] M. Yang, X. Xi, X. Wu, R. Lu, W. Zhou, S. Zhang and H. Gao, J.

356 Chromatogr. A. 1381(2015) 37-47.

357 [10] Q. Zhang, M. Tian, M. Wang, H. Shi and M. Wang, J. Agr. Food

358 Chem. 62(2014) 2809-2815.

359 [11] T. Chai, Q. Jia, S. Yang and J. Qiu, J. Sep. Sci. 37(2014) 595-601.

360 [12] Y. Fu, T. Yang, J. Zhao, L. Zhang, R. Chen and Y. Wu, Food Chem.

361 218(2017) 192-198.

362 [13] H. Su, Y. Lin, Z. Wang, Y.-L. E. Wong, X. Chen and T.-W. D. Chan,

363 J. Chromatogr. A. 1466(2016) 21-28.

364 [14] A. Periat, I. Kohler, A. Thomas, R. Nicoli, J. Boccard, J.-L.

365 Veuthey, J. Schappler and D.

366 Guillarme, J. Chromatogr. A. 1439(2016) 42-53.

367 [15] Q. Miao, J. Wang, J. Nie, H. Wu, Y. Liu, Z. Li and M. Qian, Anal.

368 Methods. 8(2016) 5296-5303.

369 [16] L. Xie, R. Jiang, F. Zhu, H. Liu and G. Ouyang, Anal. Bioanal.

370 Chem. 406(2014) 377-399.

371 [17] Y. Wen, L. Chen, J. Li, D. Liu and L. Chen, TrAC. Trend. Anal.

372 Chem. 59(2014) 26-41.

373 [18] R. Heydari, M. Hosseini and R. Rezaeepour, J. Iran. Chem.Soc.

374 (2017) 1-8.

375 [19] R. Rodríguez, J. Avivar, L. O. Leal, V. Cerdà and L. Ferrer, TrAC.


ACCEPTED MANUSCRIPT

376 Trend. Anal. Chem. 76(2016) 145-152.

377 [20] J. Nasiri, M. R. Naghavi, E. Motamedi, H. Alizadeh, M. R. F.

378 Moghadam, M. Nabizadeh and A. Mashouf, J. Chromatogr. B.

379 1043(2017) 96-106.

380 [21] J. Ma, Z. Yao, L. Hou, W. Lu, Q. Yang, J. Li and L. Chen, Talanta,

381 161(2016) 686-692.

382 [22] F. Maya, C. P. Cabello, R. M. Frizzarin, J. M. Estela, G. Turnes

383 and V. Cerdà, TrAC. Trend. Anal. Chem. (2017) .

384 [23] Y. Shao, L. Zhou, C. Bao, J. Ma, M. Liu and F. Wang, Chem. Eng.

385 J. 283(2016) 1127-1136.

386 [24] Y. Cui, B. Li, H. He, W. Zhou, B. Chen and G. Qian, Accounts

387 Chem. Res. 49(2016) 483-493.

388 [25] Y. Liu, Z. Gao, R. Wu, Z. Wang, X. Chen and T.-W. D. Chan, J.

389 Chromatogr. A, 1479(2017) 55-61.

390 [26] P. Falcaro, R. Ricco, A. Yazdi, I. Imaz, S. Furukawa, D. Maspoch,

391 R. Ameloot, J. D. Evans and C. J. Doonan, Coordin. Chem. Rev.

392 307(2016) 237-254.

393 [27] P. Rocío-Bautista, P. González-Hernández, VPino,. J. Pasán and A.

394 M. Afonso, TrAC. Trend. Anal. Chem. (2017).

395 [28] Y.-W. Zhang, Z. Li, Q. Zhao, Y.-L. Zhou, H.-W. Liu and X.-X.

396 Zhang, Chemical Communications, 50(2014) 11504-11506.

397 [29] Q. Yang, J. Wang, W. Zhang, F. Liu, X. Yue, Y. Liu, M. Yang, Z. Li


ACCEPTED MANUSCRIPT

398 and J. Wang, Chem. Eng. J, 2017, 313, 19-26.

399 [30] Q. Chen, X. Li, X. Min, D. Cheng, J. Zhou, Y. Li, Z. Xie, P. Liu, W.

400 Cai and C. Zhang, J. Electroanal.Chem, 789(2017) 114-122.

401 [31] J. Yu, C. Mu, B. Yan, X. Qin, C. Shen, H. Xue and H. Pang, Mater.

402 Horiz. (2017).

403 [33] Z. Hasan and S. H. Jhung, J.Hazard. Mater. 283(2015) 329-339.

404 [33] Y. Huang, Y. Wang, Q. Pan, Y. Wang, X. Ding, K. Xu, N. Li and Q.

405 Wen, Anal. Chim. Acta. 877(2015)90-99.

406 [34] J. Zhou, X. Li, L. Yang, S. Yan, M. Wang, D. Cheng, Q. Chen, Y.

407 Dong, P. Liu and W. Cai, Anal. Chim. Acta. 899(2015) 57-65.
ACCEPTED MANUSCRIPT

Tables

Table 1 Retention time, linearity, limit of detection for triazole pesticides.

Table 2 Detection of triazole pesticides and method recoveries in water samples.

Triazole Retention time Linear range Regression LOD


pesticides (min) (µg/L) R2 (µg/L)
equation
Flusilazole 6.983 0.25-1000 Y=15110.8X + 0.05
0.9994
1298.35
Fenbuconazole 6.878 0.25-1000 Y=4465.34X + 0.1
0.9994
54.8386
Myclobutanil 6.523 0. 25-1000 Y=14510.3X + 0.05
0.9993
180.454
Penconazole 7.153 0. 25-1000 Y=7369.22X + 0.05
0.9993
713.309
Epoxiconazole 6.826 0.25-1000 Y=20081.1X + 0.05
0.9992
1190.62

Table 1
ACCEPTED MANUSCRIPT

Triazole Spiked Tap water (µg/L) Well water (µ g/L)


pesticides (µg/L) Found Recovery (%± Found Recovery (%±
(µg/L) RSD) (µg/L) RSD)
0 0 0
Flusilazole 100 80.19 80.19±2.7 78.48 78.48±1.9
200 149.22 74.61±3.1 183.06 91.53±2.4
500 417.33 83.47±1.6 432.76 86.55±3.4
0 0 0
Fenbuconazole 100 85.68 85.68±5.3 73. 74 73.74±4.2
200 144.61 72.3±3.0 178.44 89.22±4.8
500 439.66 87.93±5.7 453.92 90.78±2.1
0 0 0
Myclobutanil 100 72.73 72.73±1.9 85.62 85.62±2.6
200 157.34 78.67±1.5 177.54 88.77±3.8
500 415.05 83.01±4.7 422.17 84.43±7.6
0 0 0
Penconazole 100 75.85 75.85±3.1 79.29 79.29±2.5
200 170.88 85.44±1.7 169.79 84.9±5.3
500 414.51 82.9±2.2 395.18 79.04±7.9
0 0 0
Epoxiconazole 100 78.45 78.45±4.7 74.52 74.52±7.4
200 175.07 87.54±8.0 148.04 74.02±5.8
500 373.13 74.63±9.1 423.42 84.68±3.3

Table 2
ACCEPTED MANUSCRIPT

Figure captions

Fig.1 TEM images of a) Fe3O4 and b) Fe3O4–GO, and SEM image of c) M-MOF-199.

Fig. 2 Magnetic hysteresis loops of a) Fe3O4, b) Fe3O4–GO– β-CD, and c)

M-MOF-199.

Fig. 3 a) N2 sorption –desorption isotherm of M-MOF-199, b) the distribution of pore

sizes of M-MOF-199 and c) the molecular structures of five triazole pesticides.

Fig. 4 XRD pattern of a) Fe3O4, b) Fe3O4–GO, and c) M-MOF-199.

Fig. 5 FT-IR spectra of a) Fe3O4, b) Fe3O4–GO, and c) M-MOF-199.

Fig. 6 Adsorption capacity of triazole pesticides on M-MOF-199.

Fig. 7 Effect of amount of M-MOF-199 on the extraction efficiency for triazole

pesticides.

Fig. 8 Effect of extraction time on the extraction efficiency for triazole pesticides of

M-MOF-199.

Fig. 9 Effect of desorption solvent on the desorption efficiency for triazole pesticides

of M-MOF-199.

Fig. 10 TICs of 200 µg/L standard solution and peak identification: 1) myclobutanil, 2)

epoxiconazole, 3) fenbuconazole, 4) flusilazole, and 5) penconazole.


ACCEPTED MANUSCRIPT

Fig. 1

Fig. 2
ACCEPTED MANUSCRIPT

Fig.3
ACCEPTED MANUSCRIPT

Fig.4

Fig.5
ACCEPTED MANUSCRIPT

Fig.6

Fig.7
ACCEPTED MANUSCRIPT

Fig.8

Fig.9
ACCEPTED MANUSCRIPT

Fig. 10
ACCEPTED MANUSCRIPT

1 A novel magnetic copper-based MOF was successfully prepared using

2 Fe3O4–GO.

4 M-MOF-199 was used for the adsorption of triazole pesticides.

6 Fe3O4–GO has a thin layer structure with M-MOF-199 coate d on its

7 surface.

9 HPLC-MS-MS was used for sample quantification and determination.

10

11 The LOD for triazoles was 0.05-0.1 µg/L and recover ies were

12 72.3-91.53%.

13

14
15
16
17
18

You might also like