Positive-Charge Functionalized Mesoporous Silica Nanoparticles as

Jun 22, 2017 - Key Laboratory of Integrated Pest Management in Crops, Ministry of Agriculture, Institute of Plant Protection, Chinese Academy of Agric...
0 downloads 17 Views 3MB Size
Subscriber access provided by University of Newcastle, Australia

Article

Positive-Charge Functionalized Mesoporous Silica Nanoparticles as Nanocarriers for Controlled 2, 4Dichlorophenoxy Acetic Acid Sodium Salt Release Lidong Cao, Zhaolu Zhou, Shujun Niu, Chong Cao, Xiuhuan Li, Yongpan Shan, and Qiliang Huang J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.7b01957 • Publication Date (Web): 22 Jun 2017 Downloaded from http://pubs.acs.org on June 25, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Agricultural and Food Chemistry 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 39

Journal of Agricultural and Food Chemistry

4

Positive-Charge Functionalized Mesoporous Silica Nanoparticles as Nanocarriers for Controlled 2, 4-Dichlorophenoxy Acetic Acid Sodium Salt Release

5

Lidong Cao,†,ǁ Zhaolu Zhou,†,ǁ Shujun Niu,‡ Chong Cao,† Xiuhuan Li,† Yongpan

6

Shan,† and Qiliang Huang*,†

7



8

Institute of Plant Protection, Chinese Academy of Agricultural Sciences, No. 2

9

Yuanmingyuan West Road, Haidian District, Beijing 100193, P. R. China.

1 2 3

Key Laboratory of Integrated Pest Management in Crops, Ministry of Agriculture,

10



11

Nongkeyuan New Village, An'ning District, Lanzhou 730070, P. R. China.

Institute of Plant Protection, Gansu Academy of Agricultural Sciences, No. 1

12 13

*Correspondence author: [email protected]; tel./fax: +86-10-6281-6909

14 15 16 17 18 19 20

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Due

to its relatively high

water solubility and

Page 2 of 39

mobility,

21

ABSTRACT:

22

2,4-dichlorophenoxy acetic acid (2,4-D) has a high leaching potential threatening the

23

surface water and groundwater. Controlled release formulations of 2,4-D could

24

alleviate the adverse effects on the environment. In the present study, positive-charge

25

functionalized mesoporous silica nanoparticles (MSNs) were facilely synthesized by

26

incorporating trimethylammonium (TA) groups onto MSNs via a post-grafting

27

method. 2,4-D sodium salt, the anionic form of 2,4-D, was effectively loaded into

28

these positively charged MSN-TA nanoparticles. The loading content can be greatly

29

improved to 21.7% compared to using bare MSNs as a single encapsulant (1.5%).

30

Pesticide loading and release patterns were pH, ionic strength and temperature

31

responsive, which were mainly dominated by the electrostatic interactions. Soil

32

column experiments clearly demonstrated that MSN–TA can decrease the soil

33

leaching of 2, 4-D sodium salt. Moreover, this novel nano-formulation showed good

34

bioactivity on target plant without adverse effects on the growth of non-target plant.

35

This strategy based on electrostatic interactions could be widely applied to charge

36

carrying agrochemicals using carriers bearing opposite charges to alleviate the

37

potential adverse effects on the environment.

38

KEYWORDS: Mesoporous silica nanoparticles, 2, 4-D sodium salt, electrostatic

39

interaction, controlled release, soil leaching, bioactivity

40 41 42

ACS Paragon Plus Environment

Page 3 of 39

Journal of Agricultural and Food Chemistry

43

INTRODUCTION

44

The extensive use of pesticides in agriculture has contributed significantly to

45

farmers’ income and food productivity. However, depending on the mode of

46

application and environmental conditions, more than 90% of the applied pesticides are

47

either lost in the environment or unable to reach the target area required for pest

48

control, which not only increases the cost but also brings about adverse impacts on the

49

environment.1 The 2,4-dichlorophenoxy acetic acid (2,4-D) is one of the most

50

commonly used herbicides worldwide for post-emergence control of broad-leaf weeds

51

due to its low cost and good selectivity.2 Due to its relatively high water solubility,

52

2,4-D exists predominantly in anionic form and is weakly retained by soil particles.3

53

Therefore, it has a high leaching potential threatening surface water and groundwater

54

particularly if heavy rains occur shortly after herbicide application.4 To overcome this

55

concern, the development of controlled release formulations (CRFs) of 2,4-D,

56

especially with nanomaterials as carriers, could be advantageous because CRFs allow

57

the use of minimal amounts of herbicide for the same period of activity, which will

58

reduce the leaching potential and environmental pollution.5

59

The performance of CRFs on controlling the release of herbicide is closely

60

related to the carrier materials. Many natural inorganic and organic polymers have

61

been used to prepare 2,4-D CRFs. Clay minerals including layered double

62

hydroxides,3 cationic surfactant-modified Arizona montmorillonites6 and bentonites,7,

63

8

64

and controlled release of 2,4-D from granule matrix formulations based on lignins,10

and organo-palygorskite9 are the most extensively investigated materials. Adsorption

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

65

activated carbon,11 and biochars12 have also been studied. Ethylcellulose as

66

microsphere matrix and Gelatin–Arabic gum complex as an envelope for CRFs of

67

2,4-D was elaborated by double encapsulation using a solvent evaporation technique

68

followed by the complex coacervation method.13 In addition to physical encapsulation

69

or absorption in polymeric matrix, chemical combination of 2,4-D with polymers

70

through covalent bonding provides another useful strategy for controlled release of

71

active ingredient. The 2,4-D was chemically caged by coupling with photoremovable

72

protecting groups of coumarin or perylene-3-ylmethanol derivatives, and controlled

73

release of 2,4-D was achieved by irradiating the caged compounds using UV-vis

74

light.14-16

75

Since Mobil’s discovery of MCM-41,17 research on mesoporous silica

76

nanoparticles (MSNs) has gained worldwide interest due to MSNs’ remarkable

77

properties, such as low cost, facile preparation, biocompatibility, large specific surface

78

area, tunable pore size for high loading capacity, and ability for targeted and

79

controlled release of cargo molecules with surface functionalization and polymer

80

coatings.18-20 Taking advantage of these unique properties, MSNs have attracted

81

widespread interest and are an ideal scaffolding for delivery systems. A slow release

82

formula of biological pesticide pyoluteorin was prepared using mesoporous silica as

83

carriers, giving an example of putting an unstable compound inside the pores to avoid

84

its fast degradation.21 The insecticide imidacloprid was effectively loaded into

85

unmodified MSNs for termite control, and the effect of pore size, specific surface area

86

and mesoporous structure on uptake and release of biocide was systematically

ACS Paragon Plus Environment

Page 4 of 39

Page 5 of 39

Journal of Agricultural and Food Chemistry

87

studied.22 Prado reported nanosized silica modified with carboxylic acid as a support

88

for the controlled release of the herbicides 2,4-D and picloram.23 For controlled

89

release in response to external stimuli, a novel redox-responsive decanethiol

90

gatekeeper was grafted onto MSNs to mediate the delivery of salicylic acid with

91

glutathione reducing agent.24 Other pesticides, such as metalaxyl,25 tebuconazole,26

92

vancomycin,27 and essential oil components28 were physically or chemically

93

combined with MSNs for controlled release. Recently, we have developed quaternized

94

chitosan-capped mesoporous silica nanoparticles as nanocarriers for controlled

95

pyraclostrobin release.29

96

Non-covalent (e.g. hydrophobic, hydrogen bonding, and ionic) interaction

97

between cargo molecules and carrier material mainly affect the loading capacity and

98

release profile.30, 31 Although research on the controlled release of pesticides using

99

MSNs as carriers has had some progress, the driving forces for pesticide loading and

100

release in most studies are hydrogen bonding and hydrophobic interactions. Ionic

101

interactions are the long-range interactions that involve the electrostatic attraction and

102

repulsion between oppositely-charged ions. These are studied less and therefore are

103

poorly exploited as tool for achieving satisfactory loading and controlled release of

104

pesticides. Taking full advantages of ionic interactions, positive charge functionalized

105

MSNs have been used for drug and gene delivery.32, 33 For charge carrying pesticide

106

molecules, the loading content should be expected to increase by strengthening the

107

electrostatic attraction through a modification of the surface of MSNs to bear more

108

opposite charges.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

109

During the pursuit of ideal carrier for controllable loading and release of 2,4-D to

110

alleviate the adverse effects on the environment, positive charge functionalized MSNs

111

have to be ideal carriers for 2,4-D sodium salt, the anionic form of 2,4-D. In the

112

present study, we report the synthesis of positively charged MSN (MSN-TA) by

113

incorporating trimethylammonium (TA) functional groups onto the pristine MSN

114

(P-MSN), and the adsorption of 2,4-D sodium salt into MSN-TA samples. The

115

parameters that impact loading content such as the ratio of carrier to pesticide,

116

temperature, pH value, various anions, as well as the release profiles were studied.

117

Furthermore, the ability to decrease soil leaching was studied in comparison to the

118

free pesticide 2,4-D sodium salt. Finally, we investigated the bioactivity of this

119

nano-sized CRFs of 2,4-D sodium salt against one dicot target plant cucumber

120

(Cucumis sativus L.) and one monocot non-target plant wheat (Triticum aestivum L.).

121 122

MATERIALS AND METHODS

123

Materials. N-Trimethoxysilylpropyl-N,N,N-trimethylammonium chloride (50%

124

in methanol), cetyltrimethylammonium bromide (CTAB, 99%), and 2,4-D-sodium salt

125

monohydrate (98%) were purchased from J&K Scientific Ltd., Beijing, China.

126

Tetraethyl orthosilicate (TEOS, 99%) was purchased from Fluorochem Ltd., Hadfield,

127

UK. All other chemicals and reagents were commercially available and used without

128

further processing. Deionized water was obtained from a Milli-Q water system

129

(Millipore Corporation, Bedford, MA, USA) and was utilized for all reactions and

130

treatment processes.

ACS Paragon Plus Environment

Page 6 of 39

Page 7 of 39

Journal of Agricultural and Food Chemistry

131

Synthesis of Positive-Charge Functionalized MSN (MSN-TA). The MSN-TA

132

sample with a hexagonal well-ordered pore structure was synthesized from pristine

133

MSN (P-MSN) by post-grafting during a two-step preparation. The first step of

134

P-MSN synthesis used a sol-gel method reported by Radu with minor modifications.34

135

Briefly, 3.0 g of CTAB was dissolved in 2000 mL of water, and then 10.5 mL of 2.0

136

M sodium hydroxide was slowly introduced into the CTAB solution at room

137

temperature under constant stirring with the stirring rate of 800 r/min. The mixture

138

was heated to 80°C in an oil bath, and then 15.0 mL of TEOS was added dropwise.

139

The solution was stirred vigorously for 6 h at 80 °C. The white solid that formed

140

during the process was collected, washed several times with ethanol and water, and

141

dried at 80°C overnight in an oven. To remove the surfactant, the as-synthesized white

142

powder was calcined at 550 °C for 5 h.

143

Positive-charge functionalization used a post-grafting synthesis according to

144

amino-functionalized mesoporous silica with a little modification.35 Specifically, 0.5 g

145

of P-MSN was re-suspended in 20 mL of anhydrous toluene solution. After vigorous

146

stirring

147

N,N,N-trimethylammonium chloride solution (50% in methanol) was added. The

148

resulting mixture was refluxed for 4 h under vigorous stirring. Samples were collected

149

by centrifuging at 10,000 rpm for 10 min, washed, and re-dispersed with deionized

150

water and ethanol several times. The nanoparticles were dried at 80°C overnight in an

151

oven.

152

for

30

min,

the

2.0

mL

of

N-trimethoxysilylpropyl-

Characterization. Fourier transform infrared (FTIR) spectroscopy was

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

153

conducted on a spectrometer (NICOLET 6700, Thermo Fisher Scientific, Waltham,

154

MA, USA) with a potassium bromide pellet and recorded over the spectral region of

155

400 to 4000 cm−1 at a spectral resolution of 4 cm-1. Thermogravimetric analysis (TGA)

156

was performed with a Perkin Elmer Pyris Diamond (Woodland, CA, USA) from 30 to

157

550 °C at 20 °C/min under a N2 atmosphere.

158

The specific surface area and pore characteristics of the samples were studied by

159

determining the nitrogen adsorption using a specific surface area and pore size

160

analyzer (TriStarII 3020, Micromeritics Instruments Corp, Norcross, GA, USA) at

161

−196 °C. The sample was outgassed at 10-3 Torr and 120 °C for about 6 h prior to the

162

adsorption experiment. From the adsorption data, the Brunauer-Emmett-Teller (BET)

163

equation36 was used to calculate the specific surface area at a relative pressure (P/P0)

164

of 0.06−0.22; the Barrett-Joyner-Halenda (BJH) model37 was used to estimate the

165

pore size distribution from the desorption branches of the isotherms.

166

The morphology and particle size of the prepared samples were characterized

167

using scanning electron microscopy (SEM, SU8000, Hitachi Ltd., Tokyo, Japan,

168

operated at 10 kV) and transmission electron microscopy (TEM, Tecnai G2, F20

169

S-TWIN, FEI, Oregon, USA, with an accelerating voltage of 200 kV). For SEM

170

observations, the samples were gold-plated and dried under vacuum prior to imaging.

171

The average particle size was determined by statistical analysis of the SEM images of

172

more than 200 nanoparticles. For TEM analysis, specimens were prepared by

173

dispersing the as-obtained powder in water and then placing a drop of the suspension

174

onto carbon-coated copper grids with air drying.

ACS Paragon Plus Environment

Page 8 of 39

Page 9 of 39

Journal of Agricultural and Food Chemistry

175

The zeta potential measurements were performed using distilled water as a

176

solvent on a ZetaSizer Nano ZS Analyzer (Zetasizer Nano ZS, Malvern Instruments

177

Ltd., Malvern, UK). The samples were prepared at 1 mg mL-1 to make intensities in

178

the range suitable for scattering. Different pH values were adjusted by the addition of

179

0.01 M HCl or NaOH. Before measurement, each sample was ultrasonicated for 5 min

180

to prevent any aggregation.

181

X-Ray photoelectron spectroscopy (XPS) was conducted on a photoelectron

182

spectrometer (ESCALab 250Xi, Thermo Fisher Scientific, USA) using 150 W

183

monochromatic Al Kα radiation (1486.6 eV, 500 µm spot size) as the excitation source;

184

all binding energies were calibrated by the C1s peak of the surface adventitious

185

carbon at 284.8 eV.

186

Loading of 2,4-D Sodium Salt into MSN-TA Samples. A typical procedure for

187

loading 2,4-D sodium salt into MSN-TA samples followed the procedure reported by

188

Lee et al with minor modification.32 Specifically, about 30 mg of MSN-TA were

189

dispersed in aqueous solution of 2,4-D sodium salt with different concentrations (1.0

190

mL) followed by another 4.0 mL of water. The suspension was stirred at room

191

temperature for 4 h and centrifuged at 10,000 rpm for 10 min. The nanoparticles were

192

collected and washed one time with 1 mL of water, and were freeze-dried with

193

vacuum freeze dryer under -40°C. The 2,4-D sodium salt-loaded MSN-TA sample

194

was denoted as 2,4-D sodium salt@MSN-TA. The amount of unloaded 2,4-D sodium

195

salt in the supernatant and washes were determined using high performance liquid

196

chromatography (HPLC, 1200-DAD (Diode Array Detector), Agilent, Santa Clara,

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

197

CA, USA). The loading content (LC) of 2,4-D sodium salt was determined by an

198

indirect method. The difference between the amount of 2,4-D sodium salt initially

199

employed and its content in the combined supernatant and washes was defined as the

200

amount of pesticide entrapped. The LC (%) of 2,4-D sodium salt was calculated as

201

follows: LC (%) = (weight of pesticide entrapped in nanoparticles/weight of

202

nanoparticles) × 100%. The encapsulation efficiency (EE) of 2,4-D sodium salt was

203

calculated as follows: EE(%) = (weight of pesticide entrapped in nanoparticles/initial

204

weight of pesticide) × 100%.

205

The HPLC operating parameters were as follows: Eclipse Plus C18

206

reversed-phase column (5 µm × 4.6 mm × 150 mm); column temperature: 30°C;

207

mobile phase: (acetonitrile: 0.2% formic acid aqueous solution (V/V) = 75:25); flow

208

rate: 1.0 mL/min; and DAD signals: 284 nm.

209

In vitro Release of 2,4-D Sodium Salt. About 20 mg of 2,4-D sodium

210

salt@MSN-TA nanoparticles were dispersed in 2.0 mL of release medium in dialysis

211

bags (MWCO: 2,000 g/mol). The dialysis bag was placed into 200 mL of release

212

medium in a D-800LS dissolution tester (Tianjin University, Tianjin, China) at a

213

stirring speed of 100 rpm. The release medium was water at different pH values via

214

diluted HCl and NaOH. To study the effects of ionic strength on the release profiles, a

215

NaCl aqueous solution (0.1 M with pH of 6.8) was used as a release medium. To

216

study the temperature effect, pure water at different temperatures ( 20, 30 and 40°C)

217

was adopted as release medium. The accumulative release profile of 2,4-D sodium

218

salt was calculated by measuring the concentrations of 2,4-D sodium salt dissolved in

ACS Paragon Plus Environment

Page 10 of 39

Page 11 of 39

Journal of Agricultural and Food Chemistry

219

the release medium at different times. To measure the concentration, 1.0 mL of release

220

medium was withdrawn at a given time intervals for HPLC analysis followed by

221

supplying the same volume of fresh release medium to ensure the same total solution

222

volume. The accumulative 2,4-D sodium salt released was calculated according to the

223

following equation: n-1

Ve ∑ Ci +V0Cn Er =

224

i= 0

mpesticide

× 100%

225

where Er is the accumulative 2,4-D sodium salt released (%) in regard to loaded

226

pesticide; Ve is the sampled volume taken at a predetermined time interval (Ve = 1.0

227

mL); Cn (mg/mL) is the 2,4-D sodium salt concentration in release medium at time n;

228

V0 is the volume of release solution (200 mL); The mpesticide (mg) is the total amount of

229

pesticide entrapped in the nanoparticles. The measurements were performed in

230

triplicate.

231

Soli Leaching Experiment. The performance of nano-carrier of MSN-TA in

232

reducing the leaching of 2,4-D sodium salt was evaluated with leaching experiments

233

through a soil column. The soil was collected locally from a depth of 0 to 15 cm, dried

234

in the shade, ground to pass through a 2-mm sieve, and stored in polythene bags at

235

room temperature. It has a pH of 7.2 and organic matter content of 4.0%. The

236

composition was sand (85.3%), clay (1.7%), and silt (13%). Glass columns (20 cm

237

length, 2.5 cm diameter) were packaged uniformly with air-dried soil, which occupied

238

about 12 cm of the column.38 The top 2 cm were filled with quartz sand, and the

239

bottom 2 cm with other quartz sand plus glass wool to minimize losses of soil and

240

contamination of leachates with soil particles. Before pesticide application, columns

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

241

were conditioned by passing them with water using a flow rate of 1 mL min-1

242

controlled by peristaltic pump. This was followed by pesticide soil leaching

243

experiments. The 2,4-D sodium salt@MSN-TA or 2,4-D sodium salt were evenly

244

dispersed in 500 mg of silica gel (corresponding to ~20 mg active ingredient), and

245

then were placed on the top of the columns. Another 1 cm of quartz sand was

246

uniformly placed on top of the pesticide. Every 10 mL of column leachate in each

247

experiment was collected, and 2,4-D sodium salt was determined using HPLC. The

248

accumulative 2,4-D sodium salt leached was calculated according to the following

249

equation: n

Ve∑ Ci 250

El =

i=1

mpesticide

× 100%

251

where El is the accumulative 2,4-D sodium salt leached (%) in regard to pesticide

252

applied; Ve is the leached volume at every sample time (Ve = 10 mL); Cn (mg/mL) is

253

the 2,4-D sodium salt concentration in leachate at sample time n; The mpesticide (mg) is

254

the total amount of pesticide applied. The measurements were performed in triplicate.

255

Bioactivity of 2,4-D Sodium Salt@MSN-TA Nanoparticles. The bioactivity of

256

2,4-D sodium salt@MSN-TA nanoparticles were assessed by means of a laboratory

257

bioassay with one target dicot plant cucumber (Cucumis sativus L.) and one monocot

258

non-target plant wheat (Triticum aestivum L.). For cucumber bioassay experiments, 9

259

cm Petri dishes with a filter paper were used, according to the laboratory bioassay of

260

fluorescent 2,4-D derivatives against Vigna radiate reported by Atta et al.14 Fifteen

261

similarly sized germinated seeds were placed in each Petri dish moistened with 10 mL

262

of 2,4-D sodium salt@MSN-TA samples at a concentration equal to the field

ACS Paragon Plus Environment

Page 12 of 39

Page 13 of 39

Journal of Agricultural and Food Chemistry

263

application rate (0.6 kg/ha). Control was similarly performed with the same amount of

264

2,4-D sodium salt free technical and pure distilled water. Each treatment was

265

performed in triplicate. The Petri dishes were incubated in a light growth chamber

266

with a 12 h photoperiod with a light intensity of 80 µmol photon m-2s-1 provided by

267

fluorescent lamps. Day and night temperatures were set at 26 and 15°C, respectively,

268

and the humidity was kept at 80%. Each Petri dish was moistened with an equal

269

volume of distilled water for daily watering. After 10 days of incubation, the root

270

length and fresh weight were recorded to evaluate the bioactivity on the target plant.

271

For the non-target plant wheat, the bioactivity was tested on pots (10 cm high

272

with diameter of 9.0 cm) filled with 240 g of soil, according to the bioassay of 2,4-D

273

nanoformulations against Zea mays reported by Abigail et al.39 Six seeds of wheat were

274

sown in each pot and grown in the greenhouse. The pots were watered daily with 10

275

mL of distilled water. One week after germination, the solution of 2,4-D sodium

276

salt@MSN-TA samples was applied post-emergence at an application rate of 2.5

277

kg/ha corresponding to the maximum application dose recommended for field

278

application of 2,4-D.6 The same amount of free 2,4-D sodium salt technical and

279

treatment without herbicide were used as controls. Each treatment was replicated

280

three times. One week after herbicide application, the plant height and fresh weight of

281

the aerial part of the wheat were determined to monitor non-target plant responses to

282

the nano-formulation.

283

Statistical Analysis. Statistical analysis of the values was conducted using SPSS

284

10.0 (SPSS Inc., Chicago, IL, USA) software. Statistical analysis was performed

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

285

using one way analysis of variance (ANOVA) followed by Duncan’s multiple range

286

test (DMRT). The values are mean ± SD for three determinations in each group. P

287

values ≤0.05 were considered as significant.

288 289

RESULTS AND DISCUSSION

290

Preparation and Characterization of MSN-TA Nanoparticles. P-MSN was

291

synthesized via a liquid crystal templating mechanism using CTAB as the

292

structure-directing agent and TEOS as the silica source under basic conditions.

293

Positive-charge functionalized MSNs (MSN-TA) were fabricated through a

294

post-grafting strategy by incorporating TA group onto P-MSN. Figure 1 shows the

295

schematic illustration of the synthesis of MSN-TA nanoparticles. The morphology of

296

MSN-TA nanoparticles was observed using SEM and TEM (Figure 2). The SEM and

297

TEM micrographs show that the as-synthesized MSN-TA nanoparticles exhibited

298

spherical morphology with a relatively smooth surface and an average particle size of

299

about 423 nm. Particle diameters were estimated by statistical analysis of the SEM

300

images of randomly selected 300 nanoparticles. The histograms of particle size

301

distributions of MSN-TA are shown in Supporting Information (Figures S1). Well

302

ordered mesoporous structures with hexagonal arrays (Figure 2D) and straight lattice

303

fringes (Figure 2E) can be seen when the electron beam is parallel and perpendicular

304

to the pore axis, which is the characteristic of MCM-41-type MSN.40 Good

305

monodispersity was achieved due to the electrostatic repulsion between positively

306

charged particles from the introduction of quaternary ammonium groups.

ACS Paragon Plus Environment

Page 14 of 39

Page 15 of 39

Journal of Agricultural and Food Chemistry

307

Incorporation of TA functional groups onto the P-MSN samples did not obviously

308

affect the P-MSN’s morphology and size (data not shown).

309

The FTIR spectra of P-MSN and MSN-TA are shown in Figure 3. The 1087 cm−1

310

broad absorption band found in P-MSN and MSN-TA was attributed to characteristic

311

Si–O–Si (siloxane) stretching vibrations. The absorption band at 1478 cm-1 in

312

MSN-TA could be assigned to the C–H bending vibration of –N(CH3)3+, confirming

313

the conjugation of TA groups on P-MSN.

314

TGA is frequently used to study the thermal stability and decomposition pattern

315

of chemicals and materials. Figure 4 displays TG analysis of P-MSN and MSN-TA.

316

Calcined P-MSN is thermo-stable and maintains a constant weight in the temperature

317

ranges studied here. About 15% weight loss is clearly seen in the TGA curve of

318

MSN-TA, and this is mainly due to the decomposition of organic TA groups, which is

319

further evidence that P-MSN was successfully grafted with positively charged TA

320

groups.

321

XPS provides valuable information about the elements on the MSN surface. The

322

XPS spectra given in Figure 5 display bands assigned to elements in P-MSN and

323

MSN-TA. In Figure 5A, binding energies of about 104.1 and 533.4 eV are assigned to

324

the Si2p and O1s in P-MSN, respectively. The weak signal at 284.8 eV corresponds to

325

the C1s originated from the residual carbon after calcination for removing the

326

template. Figure 5B shows the signals at 402.7, 286.2, and 197.6 eV corresponding to

327

the N1s, C1s and Cl2p, respectively, which confirmed successful incorporation of TA

328

functional groups onto P-MSN. The XPS results show that the atomic percentage of N

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

329

Page 16 of 39

is 3.27%.

330

BET specific surface area and BJH pore size and volume analysis were used to

331

explore the nanoparticles’ mesoporous structure. The values for the BET specific

332

surface area (SBET), the total pore volume (Vt), and the BJH pore diameter (DBJH) are

333

summarized in Table 1. Nitrogen adsorption–desorption isotherms and pore-size

334

distribution of P-MSN and MSN-TA are shown in Figure 6. The P-MSN samples

335

display a type IV isotherm curve with a sharp increase in volume adsorbed between

336

0.3 and 0.4 of P/P0 shown in Figure 6A, which is characteristic of a well-defined

337

mesoporous structure. The BET specific surface area reduced from 1356.0 to 956.4

338

m2/g after the TA groups were grafted, while the pore volume decreased from 1.65 to

339

0.59 cm3/g suggesting that part of the pores were blocked with the TA groups.

340

However, the MSN-TA nanoparticles also have a narrow pore-size distribution with

341

an average pore diameter of 2.54 nm according to Figure 6B. The adsorption of 2,4-D

342

sodium salt into MSN-TA samples was related to extreme decreases in pore volume

343

(0.59 to 0.03 cm3/g). This decrease occurred because the nanochannels were almost

344

fully occupied by the pesticide molecules, which left little space for nitrogen

345

adsorption.

346

The zeta potential measures residual charges on the surface of nanoparticles. It is

347

an indicator for the quaternary ammonium groups on the surface of MSN. The

348

magnitude of the zeta potential is very important in determining the stability of

349

nanoparticle systems. Generally, nanoparticles having the zeta potential values higher

350

than +30 mV or lower than -30 mV are considered as stable systems.41,

ACS Paragon Plus Environment

42

The

Page 17 of 39

Journal of Agricultural and Food Chemistry

351

presence of quaternary ammonium may produce higher zeta potential values. The zeta

352

potential of P-MSN and MSN-TA samples in solutions at pH 3, 7, and 9 was

353

measured. At pH 3.0, both the P-MSN and MSN-TA showed the positive zeta

354

potential values of +7.2 and +67.8 mV, respectively. At pH 7.0, the zeta potential

355

values for P-MSN and MSN-TA were –22.9 and +19.6 mV, respectively. At this

356

neutral solution condition, the silanol groups (Si-OH) on the surface of MSN became

357

deprotonated, and thus the P-MSN exhibited negative zeta potential. The MSN-TA

358

retained their positive zeta potential due to the high density of positively charged TA

359

groups. At pH 9.0, all the samples showed negative zeta potential: –42.4 mV for

360

P-MSN and –23.1 mV for MSN-TA. This pH-dependent zeta potential could explain

361

the pH-responsive controlled release.

362

Loading of 2,4-D Sodium Salt into MSN-TA Nanoparticles. With this

363

MSN-TA carrier in hand, the LC and EE of 2,4-D sodium salt were next optimized

364

including the solvent and ratio of carrier-to-pesticide. The LC and EE results of 2,4-D

365

sodium salt under various conditions are presented in Table 2. The LC increased with

366

increasing 2,4-D sodium salt. This is possibly due to the higher concentrations of

367

2,4-D sodium salt, which generated a strong gradient to facilitate the diffusion of

368

cargo molecules into the MSN pores. The amount of 2,4-D sodium salt adsorbed onto

369

the MSN-TA nanoparticles plateaued (up to 21.7%) when the mass ratio of pesticide

370

to carrier reached 1:0, which implied a saturation absorption of 2,4-D sodium salt

371

(entry 4, Table 2). The EE decreased with the increase of the mass ratio of cargo to

372

carrier. When methanol was used as the solvent, the LC decreased from 21.7% to 14.5%

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

373

(entry 6, Table 2) due to the higher solubility of 2,4-D sodium salt in water. It is very

374

interesting that when P-MSN instead of MSN-TA was used, the LC decreased sharply

375

to 1.5% (entry 7, Table 2). At this neutral solution condition, the determined zeta

376

potential value for P-MSN was –22.9 mV. As a consequence, the electrostatic

377

repulsions between the negative charges and 2,4-D sodium salt lead to the prominent

378

decrease of loading content. On the other hand, these results clearly demonstrated the

379

vital importance of electrostatic attractions for enhanced loading content. Considering

380

the LC and EE together, the conditions of the cargo/carrier ratio of 1:1, pure water as

381

solvent and room temperature (entry 4, Table 2) were adopted for scale preparation of

382

2,4-D sodium salt-loaded MSN-TA samples, which were used for the in vitro release,

383

soil leaching and bioactivity assay. The successful loading of 2,4-D sodium salt into

384

MSN-TA was also confirmed by FTIR and TGA analysis (Figures 3 and 4).

385

The effects of solution pH, ionic strength, and temperature on the LC were also

386

studied. When inorganic salt NaCl or Na2SO4 was added to the solution, the LC of

387

2,4-D sodium salt decreased with increasing ionic strength, pH, and temperature

388

(Figures S2-S4). The electrostatic interactions dominated the loading of 2,4-D sodium

389

salt into MSN-TA nanoparticles under various conditions. Loading and release are

390

opposite process—higher LC means lower release. Thus, we next discussed the

391

possible reasons underlying the stimuli-responsive loading and release patterns.

392

Controlled Release of 2,4-D Sodium Salt. Pesticide carriers with controllable

393

release in response to environmental stimuli are highly desirable for better efficacy

394

and fewer side effects. The pesticide release profiles of the as-prepared positively

ACS Paragon Plus Environment

Page 18 of 39

Page 19 of 39

Journal of Agricultural and Food Chemistry

395

charged MSNs were studied to reveal their potential in pesticide delivery system. The

396

release profiles of 2,4-D sodium salt@MSN-TA samples under three different pH

397

values of 3, 7 and 10 were investigated. Figure 7A shows that the release behaviors of

398

2,4-D sodium salt were obviously pH-sensitive. The samples showed the lowest initial

399

release in the first 2 h (16%) under pH 3.0, while the corresponding release rate was

400

43% under pH 10.0. The electrostatic interactions dominated the release profiles when

401

the pH of the environment changed. The schematic illustration of the

402

controlled-release mechanism of an anionic pesticide 2,4-D sodium salt is shown in

403

Figure 8. In a weak acid solution, the MSN-TA nanoparticles carry more positive

404

charge (zeta potential: +67.8 mV at pH 3.0), and a strong electrostatic attraction

405

impeded the release of negatively charged 2,4-D sodium salt. At higher pH values, the

406

silanol groups (Si-OH) in the MSN-TA nanoparticles are deprotonated, and a strong

407

electrostatic repulsion between the negative charges of SiO– groups and negative

408

charges of 2,4-D sodium salt would increase the release rate (Mechanism A, Figure

409

8).

410

The effect of ionic strength on the release profiles was also studied. Figure 7A

411

shows that when a NaCl solution (0.1 M) was used as a release medium, the release

412

rate was faster than in pure water (red line in Figure 7A). The release profile has an

413

obvious burst release. About 90% of the 2,4-D sodium salt was released after only 5 h;

414

the release was only 40% in pure water at the same time interval. This ionic

415

strength-triggering release is mainly due to the ion-exchange mechanism. The

416

competitive electrostatic attraction between negatively charged chloride ions and

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

417

positively charged TA groups would compel 2,4-D sodium salt to be far away from

418

the TA group centers (Mechanism B, Figure 8). This definitely facilitated payload

419

release.

420

We also studied thermo-responsive cargo molecule delivery because it is one of

421

the most common stimuli-responsive strategies. Figure 7B shows that the release of

422

2,4-D sodium salt was temperature-dependent. More 2,4-D sodium salt was released

423

at higher temperatures. At 40, 30 and 20°C, the accumulative releases of 2,4-D

424

sodium salt were 96, 75 and 52%, respectively, after 900 min. These

425

temperature-controllable release patterns could possibly occur via the well-known

426

temperature-dependent, diffusion-controlled process. High temperature may facilitate

427

diffusion of payloads from the pores of MSN to the release medium.

428

Retarded 2,4-D Sodium Salt Leaching in Soil. The results of the soil column

429

experiments are seen in breakthrough curves (BTCs) in which the amount of 2,4-D

430

sodium salt leached (mg) in each collected fraction is shown as the ordinate in relation

431

to the cumulative volume of eluent applied presented as the abscissa. The BTCs of the

432

2,4-D sodium salt@MSN-TA samples and free 2,4-D sodium salt technical as control

433

are shown in Figure 9A. Although the breakthrough of 2,4-D sodium salt occurs under

434

the same cumulative volume of eluent for 2,4-D CRFs and free technical, the

435

maximum leaching amount was greatly reduced from 3.7 to 1.7 mg. When the

436

cumulative volume of 340 mL eluent was applied, the total amount of 2,4-D sodium

437

salt leached was clearly lower in the CRFs (48.4%) than in the free system (97.3%)

438

(Figure 9B). The soil column leaching test confirmed controlled release of 2,4-D

ACS Paragon Plus Environment

Page 20 of 39

Page 21 of 39

Journal of Agricultural and Food Chemistry

439

sodium salt from formulations based on MSN-TA nanoparticles. This retarded the

440

vertical movement of the herbicide through soil and reduced the leaching potential.

441

Bioactivity of 2,4-D sodium salt@MSN-TA samples. The bioactivity of the

442

nano-formulations was compared to free herbicide and pure water using one dicot

443

target plant cucumber (Cucumis sativus L.) and one monocot non-target plant wheat

444

(Triticum aestivum L.). The 2,4-D sodium salt nano-formulations were statistically as

445

effective as the free herbicide in controlling of the test plant cucumber (Figure 10),

446

when 2,4-D sodium salt at a concentration equal to the field application rate (0.6

447

kg/ha) was applied. There was obvious root length inhibition. Ten days after

448

application, the fresh weight was reduced to about 50% compared to treatment

449

without herbicide, demonstrating good herbicidal bioactivity. Slight lower inhibition

450

effect of 2,4-D sodium salt nano-formulation than free herbicide indicated the

451

controlled release at the first stage.

452

2,4-D is a selective, systemic herbicide used for control of broad-leaved weeds.

453

It is safe for monocot plant under recommended dosage. In the present study, wheat

454

was selected as model plant to evaluate the safety of 2,4-D sodium salt@MSN-TA

455

samples toward monocot plant. At an application rate of 2.5 kg/ha corresponding to

456

the maximum application dose recommended for field application of 2,4-D, both the

457

nano-formulation and free 2,4-D sodium salt applied post-emergency did not affect

458

the plant height and free weight (Figure 11). Abigail et al also reported that the

459

nano-formulation of 2,4-D based on rice husk nanosorbents as carriers does not affect

460

the development of non-target plant (Zea mays).39 Therefore, the prepared CRFs of

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

461

2,4-D sodium salt can be satisfactorily used to control dicot plant without injuring the

462

monocot plant.

463

In summary, a stimuli-responsive controllable anionic pesticide release system

464

has been designed by incorporating positive charges on the surface of MSN.

465

Electrostatic interactions are the driving forces that facilitate pesticide loading. This

466

regulates release and decreases soil leaching potential. Good bioactivity was seen on

467

the target plant with no impact on the non-target plant. Hence, positively charged

468

MSNs used as nanocarriers for 2,4-D sodium salt could reduce environmental

469

pollution without affecting bioactivity. The strategy based on electrostatic interactions

470

could be widely applied to charge-carrying agrochemicals using carriers bearing

471

opposite charges to alleviate the potential adverse effects on the environment.

472

473

ASSOCIATED CONTENT

474

Supporting Information. This material is available free of charge via the Internet at

475

http://pubs.acs.org.

476

The distribution of particle size of MSN-TA nanoparticles (Figure S1); The effects of

477

solution pH, ionic strength and temperature on the loading content of 2,4-D sodium

478

salt into MSN-TA nanoparticles (Figures S2-S4).

479 480

AUTHOR INFORMATION

481

Corresponding Author

482

Qiliang Huang, [email protected]; Tel./Fax: +86-10-6281-6909.

ACS Paragon Plus Environment

Page 22 of 39

Page 23 of 39

Journal of Agricultural and Food Chemistry

483

ORCID

484

Lidong Cao: 0000-0001-7217-7102

485

Qiliang Huang: 0000-0001-9820-7218

486

Author Contributions

487

ǁ

488

Funding

489

This work was supported by the State Key Development Program for Basic Research

490

of China (No. 2014CB932204) and the National Natural Science Foundation of China

491

(NSFC) (No. 31471805).

492

Notes

493

The authors declare no competing financial interest.

L.C. and Z.Z. contributed equally to this work.

494 495

REFERENCES

496

[1] Mogul, M. G.; Akin, H.; Hasirci, N.; Trantolo, D. J. ; Gresser, J. D. ; Wise, D. L.

497

Controlled release of biologically active agents for purpose of agricultural crop

498

management. Resour. Conserv. Recy. 1996, 16, 289–320.

499

[2] Garabrant, D. H.; Philbert, M. A. Review of 2,4-dichlorophenoxyacetic acid

500

(2,4-D) epidemiology and toxicology. Crit. Rev. Toxicol. 2002, 32, 233–257.

501

[3] Chao, Y.-F.; Chen, P.-C.; Wang, S.-L. Adsorption of 2, 4-D on Mg/Al–NO3 layered

502

double hydroxides with varying layer charge density. Appl. Clay Sci. 2008, 40,

503

193–200.

504

[4] Hyun, S.; Lee, L. S. Quantifying the contribution of different sorption mechanisms

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

505

for 2,4-dichlorophenoxyacetic acid sorption by several variable-charge soils. Environ.

506

Sci. Technol. 2005, 39, 2522–2528.

507

[5] Nuruzzaman, M.; Rahman, M. M.; Liu, Y.; Naidu, R. Nanoencapsulation,

508

nano-guard for pesticides: A new window for safe application. J. Agric. Food Chem.

509

2016, 64, 1447–1483.

510

[6] Hermosín, M. C.; Celis R.; Facenda G.; Carrizosa M. J.; Ortega-Calvo J. J.;

511

Cornejo J. Bioavailability of the herbicide 2,4-D formulated with organoclays. Soil

512

Boil. Biochem. 2006, 38, 2117–2124.

513

[7] Li, J.; Jiang, M.; Wu, H.; Li, Y. Addition of modified bentonites in polymer gel

514

formulation of 2,4-D for its controlled release in water and soil. J. Agric. Food Chem.

515

2009, 57, 2868–2874.

516

[8] Bakhtiary, S.; Shirvani, M.; Shariatmadari, H. Adsorption–desorption behavior of

517

2,4-D on NCP-modified bentonite and zeolite: Implications for slow-release herbicide

518

formulations. Chemosphere 2013, 90, 699–705.

519

[9] Xi, Y., Mallavarapu, M., Naidu, R. Adsorption of the herbicide 2,4-D on

520

organo-palygorskite. Appl. Clay Sci. 2010, 49, 255–261.

521

[10] Ferraz, A.; Souza, J. A.; Silva, F. T.; Goncalves, A. R.; Bruns, R. E.; Cotrim, A.

522

R.; Wilkins, R. M. Controlled release of 2,4-D from granule matrix formulations

523

based on six lignins. J. Agric. Food Chem. 1997, 45, 1001–1005.

524

[11] Aksu, Z.; Kabasakal, E. Batch adsorption of 2,4-dichlorophenoxy-acetic acid

525

(2,4-D) from aqueous solution by granular activated carbon. Sep. Purif. Technol. 2004,

526

35, 223–240.

ACS Paragon Plus Environment

Page 24 of 39

Page 25 of 39

Journal of Agricultural and Food Chemistry

527

[12] Kearns, J. P.; Wellborn, L. S.; Summers, R. S.; Knappe, D. R. U. 2,4-D

528

adsorption to biochars: Effect of preparation conditions on equilibrium adsorption

529

capacity and comparison with commercial activated carbon literature data. Water Res.

530

2014, 62, 20–28.

531

[13] El bahri Z., Taverdet J.-L. Elaboration of microspheres and coated microspheres

532

for the controlled release of the herbicide 2,4-D. Polym. Bull. 2007, 59, 709–719.

533

[14] Atta, S.; Jana, A.; Ananthakrishnan R.; Singh, N. D. P.

534

compounds of 2,4-dichlorophenoxyacetic acid (2,4-D): Photorelease technology for

535

controlled release of 2,4-D. J. Agric. Food Chem. 2010, 58, 11844–11851.

536

[15] Atta, S.; Paul, A.; Banerjee, R.; Bera, M.; Ikbal, M.; Dhara, D.; Singh, N. D. P.

537

Photoresponsive polymers based on a coumarin moiety for the controlled release of

538

pesticide 2,4-D. RSC Adv. 2015, 5, 99968–99975.

539

[16] Atta, S.; Bera, M.; Chattopadhyay, T.; Paul, A.; Ikbal, M.; Maiti, M. K.; Singh, N.

540

D. P. Nano-pesticide formulation based on fluorescent organic photoresponsive

541

nanoparticles: for controlled release of 2,4-D and real time monitoring of

542

morphological changes induced by 2,4-D in plant systems. RSC Adv. 2015, 5,

543

86990–86996.

544

[17] Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.; Beck, J. S. Ordered

545

mesoporous molecular sieves synthesized by a liquid-crystal template mechanism.

546

Nature 1992, 359, 710–712.

547

[18] Tarn, D.; Ashley, C. E.; Xue, M.; Carnes, E. C.; Zink, J. I.; Brinker, C. J.

548

Mesoporous silica nanoparticle nanocarriers: Biofunctionality and biocompatibility.

ACS Paragon Plus Environment

Fluorescent caged

Journal of Agricultural and Food Chemistry

549

Acc. Chem. Res. 2013, 46, 792–801.

550

[19] Aznar, E.; Oroval, M.; Pascual, L.; Murguía, J. R.; Martínez-Máñez, R.;

551

Sancenón, F. Gated materials for on-command release of guest molecules. Chem. Rev.

552

2016, 116, 561–718.

553

[20] Song, Y.; Li, Y.; Xu, Q.; Liu, Z. Mesoporous silica nanoparticles for

554

stimuliresponsive controlled drug delivery: advances, challenges, and outlook. Int. J.

555

Nanomed. 2017, 12, 87–110.

556

[21] Chen, J.; Wang, W.; Yu, Y.; Zhang, X.J. Slow-release formulation of a new

557

biological pesticide, pyoluteorin, with mesoporous silica. J. Agric. Food Chem. 2011,

558

59, 307–311.

559

[22] Popat, A.; Liu, J.; Hu, Q.; Kennedy, M.; Peters, B.; Lu, G. Q.; Qiao, S. Z.

560

Adsorption and release of biocides with mesoporous silica nanoparticles. Nanoscale

561

2012, 4, 970–975.

562

[23] Prado, A. G. S.; Moura, A. O.; Nunes, A. R. Nanosized silica modified with

563

carboxylic acid as support for controlled release of herbicides. J. Agric. Food Chem.

564

2011, 59, 8847–8852.

565

[24] Yi, Z.; Hussain, H. I.; Feng, C.; Sun, D.; She, F.; Rookes, J. E.; Cahill, D. M.;

566

Kong, L. Functionalized mesoporous silica nanoparticles with redox-responsive

567

short-chain gatekeepers for agrochemical delivery. ACS Appl. Mater. Interfaces, 2015,

568

7, 9937–9946.

569

[25] Wanyika, H. Sustained release of fungicide metalaxyl by mesoporous silica

570

nanospheres. J. Nanopart. Res. 2013, 15, 1831.

ACS Paragon Plus Environment

Page 26 of 39

Page 27 of 39

Journal of Agricultural and Food Chemistry

571

[26] Mas, N.; Galiana, I.; Hurtado, S.; Mondragón, L.; Bernardos, A.; Sancenón, F.;

572

Marcos, M. D.; Amorós, P.; Abril-Utrillas, N.; Martínez-Máñez, R.; Murguía, J. R.

573

Enhanced antifungal effcacy of tebuconazole using gated pH-driven mesoporous

574

nanoparticles. Int. J. Nanomed. 2014, 9, 2597–2606.

575

[27] Qi, G.; Li, L.; Yu, F.; Wang, H. Vancomycin-modified mesoporous silica

576

nanoparticles for selective recognition and killing of pathogenic gram-positive

577

bacteria over macrophage-like cells. ACS Appl. Mater. Interfaces 2013, 5,

578

10874-10881.

579

[28] Bernardos, A.; Marina, T.; Žáček, P.; Pérez-Esteve, É.; Martínez-Máñez, R.;

580

Lhotka, M.; Kouřimská, L.; Pulkrábek, J.; Klouček, P. Antifungal effect of essential

581

oil components against Aspergillus niger when loaded into silica mesoporous supports.

582

J. Sci. Food Agric. 2015, 95, 2824–2831.

583

[29] Cao, L.; Zhang, H.; Cao, C.; Zhang, J.; Li, F.; Huang, Q. Quaternized

584

chitosan-capped mesoporous silica nanoparticles as nanocarriers for controlled

585

pesticide release. Nanomaterials 2016, 6, 126.

586

[30] Ke, X.; Lin, V. W.; Ono, R. J.; Chan, J. M. W.; Krishnamurthy, S.; Wang, Y.;

587

Hedrick, J. L.; Yang, Y. Y. Role of non-covalent and covalent interactions in cargo

588

loading capacity and stability of polymeric micelles. J. Control. Release 2014, 193,

589

9–26.

590

[31] De Robertis, S.; Bonferoni, M. C.; Elviri, L.; Sandri, G.; Caramella, C.; Bettini,

591

R. Advances in oral controlled drug delivery: the role of drug-polymer and

592

interpolymer non-covalent interactions. Expert Opin. Drug Deliv. 2015, 12, 441–453.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

593

[32] Lee, C.-H.; Lo, L.-W.; Mo, C.-Y.; Yang, C.-S. Synthesis and characterization of

594

positive-charge functionalized mesoporous silica nanoparticles for oral drug delivery

595

of an anti-inflammatory drug. Adv. Funct. Mater. 2008, 18, 3283–3292.

596

[33] Li, Y.; Hei, M.; Xu, Y.; Qian, X. Zhu, W. Ammonium salt modified mesoporous

597

silica nanoparticles for dual intracellular-responsive gene delivery. Int. J. Pharm. 2016,

598

511, 689–702.

599

[34] Radu, D. R.; Lai, C.-Y.; Jeftinija, K.; Rowe, E.W.; Jeftinija, S.; Lin, V.S.-Y. A

600

polyamidoamine dendrimer-capped mesoporous silica nanosphere-based gene

601

transfection reagent. J. Am. Chem. Soc. 2004, 126, 13216–13217.

602

[35] Hikosaka, R.; Nagata, F.; Tomita, M.; Kato, K. Adsorption and desorption

603

characteristics of DNA onto the surface of amino functional mesoporous silica with

604

various particle morphologies. Colloids Surf. B: Biointerfaces, 2016, 140, 262-268.

605

[36] Walton, K. S.; Snurr, R. Q. Applicability of the BET method for determining

606

surface areas of microporous metal-organic frameworks. J. Am. Chem. Soc. 2007, 129,

607

8552–8556.

608

[37] Choma, J.; Jaroniec, M.; Burakiewicz-Mortka, W.; Kloske, M. Critical appraisal

609

of classical methods for determination of mesopore size distributions of MCM-41

610

materials. Appl. Surf. Sci. 2002, 196, 216–223.

611

[38] García-Jaramillo, M.; Cox, L.; Cornejo, J.; Hermosín, M. C. Effect of soil

612

organic amendments on the behavior of bentazone and tricyclazole. Sci. Total Environ.

613

2014, 466–467, 906–913.

614

[39] Abigail, E. A. M.; Samuel, M. S.; Chidambaram, R. Application of rice husk

ACS Paragon Plus Environment

Page 28 of 39

Page 29 of 39

Journal of Agricultural and Food Chemistry

615

nanosorbents containing 2,4-dichlorophenoxyacetic acid herbicide to control weeds

616

and reduce leaching from soil. J. Taiwan Inst. Chem. E. 2016, 63, 318–326.

617

[40] Chenite, A.; Le Page, Y.; Sayari, A. Direct TEM imaging of tubules in calcined

618

MCM-41 type mesoporous materials. Chem. Mater. 1995, 7, 1015–1019.

619

[41] Duman, O.; Tunç, S. Electrokinetic and rheological properties of Na-bentonite

620

in some electrolyte solutions. Microporous Mesoporous Mater. 2009, 117, 331–338.

621

[42] Tunç, S.; Duman, O.; Kancı, B. Rheological measurements of Na-bentonite and

622

sepiolite particles in the presence of tetradecyltrimethylammonium bromide, sodium

623

tetradecyl sulfonate and Brij 30 surfactants. Colloids Surf. A: Physicochem. Eng.

624

Aspects, 2012, 398, 37–47.

625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

642

FIGURES CAPTIONS

643

Figure 1. Schematic illustration of the synthesis of MSN-TA nanoparticles.

644

Figure 2. SEM (A, B) and TEM (C, D, E) images of the TA groups functionalized

645

MSN (MSN-TA). TEM images D and E were viewed along and perpendicular to the

646

channel direction. Scale bars: (A) 5.0 µm; (B) 1.0 µm; (C) 0.5 µm; and (D, E) 50 nm.

647

Figure 3. FTIR spectra of P-MSN, MSN-TA, 2,4-D sodium salt and 2,4-D sodium

648

salt@MSN-TA.

649

Figure 4. TGA of P-MSN, MSN-TA, 2,4-D sodium salt and 2,4-D sodium

650

salt@MSN-TA.

651

Figure 5. XPS spectra of P-MSN (A) and MSN-TA (B).

652

Figure 6. Nitrogen adsorption–desorption isotherms of P-MSN, MSN-TA and 2,4-D

653

sodium salt@MSN-TA (A), and pore size distributions of P-MSN and MSN-TA (B).

654

Figure 7. 2,4-D sodium salt released at different pH value and ionic strength (A) and

655

temperature (B). Error bars correspond to standard errors of triplicate measurements.

656

Figure 8. Schematic illustration of the controlled-release mechanism of an anionic

657

pesticide 2,4-D sodium salt loaded in positive-charge functionalized MSNs.

658

Mechanism A: pesticide released by electrostatic repulsion under neutral or basic

659

solution; Mechanism B: pesticide released through ion-exchange by increasing the

660

ionic strength.

661

Figure 9. Breakthrough curves (A) and accumulative leaching profiles (B) for 2,4-D

662

sodium salt applied for soil columns as MSN-TA formulations and free technical.

663

Error bars correspond to standard errors of triplicate measurements.

664

Figure 10. Bioactivity for target plant cucumber (Cucumis sativus L.) determined in

665

terms of root length and fresh weight. (A) Control without treatment; (B) Free 2,4-D

ACS Paragon Plus Environment

Page 30 of 39

Page 31 of 39

Journal of Agricultural and Food Chemistry

666

sodium salt technical; (C) 2,4-D sodium salt@MSN-TA samples. Bars marked with

667

different letters are statistically different at P ≤0.05 as determined by Duncan’s

668

multiple range test.

669

Figure 11. Bioactivity for non-target plant wheat (Triticum aestivum L.) determined in

670

terms of plant height and fresh weight. (A) Control without treatment; (B) Free 2,4-D

671

sodium salt technical; (C) 2,4-D sodium salt@MSN-TA samples. Bars marked with

672

different letters are statistically different at P ≤0.05 as determined by Duncan’s

673

multiple range test.

674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 32 of 39

Table 1. Mesoporous structure characterization of nanoparticles.a

691

a

692

Sample

SBET (m2/g)

Vt (cm3/g)

DBJH (nm)

P-MSN

1356.0

1.65

3.75

MSN-TA

956.4

0.59

2.31

2,4-D sodium salt@MSN-TA

454.8

0.03



SBET: BET specific surface area; Vt: total pore volume; DBJH: BJH pore diameter.

693 694

Table 2. The loading content (LC) and encapsulation efficiency (EE) of

695

2,4-D sodium salt into MSN-TA nanoparticles. a Entry

Solvent

Mass ratiob

LC (%)

EE (%)

1

H2O

0.4

15.0±0.3d

44.0±1.2a

2

H2O

0.6

17.5±0.3c

35.3±0.3b

3

H2O

0.8

19.2±0.2b

29.7±0.4c

4

H2O

1.0

21.7±0.3a

27.7±0.4d

5

H2O

1.2

22.0±0.4a

23.5±0.6e

6

MeOH

1.0

14.5±0.3

17.0±0.4

7c

H2O

1.0

1.5±0.1

1.6±0.1

696

a

MSN-TA (30 mg), H2O (5.0 mL), room temperature; b Mass ratio of 2,4-D sodium

697

salt to MSN-TA;

698

three replicates. Values in each column followed by different letters are statistically

699

different at P ≤0.05 as determined by Duncan’s multiple range test.

c

P-MSN instead of MSN-TA was used. Values are mean±SD of

700 701 702

ACS Paragon Plus Environment

Page 33 of 39

Journal of Agricultural and Food Chemistry

703 704

Figure 1. Schematic illustration of the synthesis of MSN-TA nanoparticles.

705 706

707 708

Figure 2. SEM (A, B) and TEM (C, D, E) images of the TA groups functionalized

709

MSN (MSN-TA). TEM images D and E were viewed along and perpendicular to the

710

channel direction. Scale bars: (A) 5.0 µm; (B) 1.0 µm; (C) 0.5 µm; and (D, E) 50 nm.

711 712

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

713 714

Figure 3. FTIR spectra of P-MSN, MSN-TA, 2,4-D sodium salt and 2,4-D sodium

715

salt@MSN-TA.

716

717 718

Figure 4. TGA of P-MSN, MSN-TA, 2,4-D sodium salt and 2,4-D sodium

719

salt@MSN-TA.

720 721

ACS Paragon Plus Environment

Page 34 of 39

Page 35 of 39

Journal of Agricultural and Food Chemistry

722 723

Figure 5. XPS spectra of P-MSN (A) and MSN-TA (B).

724

725 726 727

Figure 6. Nitrogen adsorption–desorption isotherms of P-MSN, MSN-TA and 2,4-D

728

sodium salt@MSN-TA (A), and pore size distributions of P-MSN and MSN-TA (B).

729

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

730 731

Figure 7. 2,4-D sodium salt released at different pH value and ionic strength (A) and

732

temperature (B). Error bars correspond to standard errors of triplicate measurements.

733 734

Figure 8. Schematic illustration of the controlled-release mechanism of an anionic

735

pesticide 2,4-D sodium salt loaded in positive-charge functionalized MSNs.

736

Mechanism A: pesticide released by electrostatic repulsion under neutral or basic

737

solution; Mechanism B: pesticide released through ion-exchange by increasing the

738

ionic strength.

739 740 741

ACS Paragon Plus Environment

Page 36 of 39

Page 37 of 39

Journal of Agricultural and Food Chemistry

742

743

Figure 9. Breakthrough curves (A) and accumulative leaching profiles (B) for 2,4-D

744

sodium salt applied for soil columns as MSN-TA formulations and free technical.

745

Error bars correspond to standard errors of triplicate measurements.

746

747 748

Figure 10. Bioactivity for target plant cucumber (Cucumis sativus L.) determined in

749

terms of root length and fresh weight. (A) Control without treatment; (B) Free 2,4-D

750

sodium salt technical; (C) 2,4-D sodium salt@MSN-TA samples. Bars marked with

751

different letters are statistically different at P ≤0.05 as determined by Duncan’s

752

multiple range test.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

753 754

Figure 11. Bioactivity for non-target plant wheat (Triticum aestivum L.) determined in

755

terms of plant height and fresh weight. (A) Control without treatment; (B) Free 2,4-D

756

sodium salt technical; (C) 2,4-D sodium salt@MSN-TA samples. Bars marked with

757

different letters are statistically different at P ≤0.05 as determined by Duncan’s

758

multiple range test.

759 760

761

762

763

764

765

766

767

768

ACS Paragon Plus Environment

Page 38 of 39

Page 39 of 39

Journal of Agricultural and Food Chemistry

769

770

For Table of Contents use only

771

772

ACS Paragon Plus Environment