Alginate Microcapsules Loaded with - ACS Publications

Oct 7, 2016 - chitosan/alginate microcapsules simultaneously loaded with biological and chemical agents in plant nutrition and protection. KEYWORDS: ...
0 downloads 0 Views 1MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Encapsulation of Biological and Chemical Agents for Plant Nutrition and Protection: Chitosan/Alginate Microcapsules Loaded with Copper Cations and Trichoderma viride Marko Vincekovic, Nenad Jalšenjak, Snježana TopolovecPintari#, Edyta #ermi#, Marija Bujan, and Slaven Juri# J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b02879 • Publication Date (Web): 07 Oct 2016 Downloaded from http://pubs.acs.org on October 8, 2016

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

Encapsulation of Biological and Chemical Agents for Plant Nutrition and Protection: Chitosan/Alginate Microcapsules Loaded with Copper Cations and Trichoderma viride Marko Vinceković1,*, Nenad Jalšenjak,1 Snježana Topolovec-Pintarić,2 Edyta Đermić,2 Marija Bujan,1 Slaven Jurić1 Department of Chemistry1, University of Zagreb Faculty of Agriculture, 10000 Zagreb Croatia Department of Plant Pathology2, University of Zagreb Faculty of Agriculture, 10000 Zagreb Croatia

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

1

ABSTRACT: Novel chitosan/alginate microcapsules simultaneously loaded with copper

2

cations and Trichoderma viride have been prepared and characterized. Information about

3

intermolecular interactions between biopolymers and bioactive agents were obtained by

4

Fourier transform infrared spectroscopy. Encapsulation of Trichoderma viride spores and the

5

presence of copper cations in the same compartment do not inhibit their activity. Dependence

6

of microcapsule loading capacity and efficiency, swelling behavior and releasing depend on

7

both, the size of the microcapsule and bioactive agents. The in vitro copper cations release

8

profile was fitted to Korsmeyer–Peppas empirical model. Fickian diffusion was found to be a

9

rate-controlling mechanism of release from smaller microcapsules, whereas anomalous

10

transport kinetics controlled release from larger microcapsules. Trichoderma viride spores

11

releasing profile exhibited exponential release over the initial lag time. Results obtained

12

opened up perspectives for the future use of chitosan/alginate microcapsules simultaneously

13

loaded with biological and chemical agents in the plant nutrition and protection.

14 15

KEYWORDS: chitosan/alginate microcapsules, encapsulation, copper cations, Trichoderma

16

viride spores, sustainability

ACS Paragon Plus Environment

Page 2 of 39

Page 3 of 39

Journal of Agricultural and Food Chemistry

17

1. INTRODUCTION

18

The encapsulation of bioactive agents has been developed in recent years as a new potential

19

tool for ecological and sustainable plant production. Encapsulation in biopolymer matrices

20

has been recognized as an effective method for controlled release of a bioactive agent used for

21

plant protection.1 Biopolymer based microcapsules with a single bioactive agent have

22

extensive applications in agriculture and became one of the standard capsule formulations.2 and

23

references therein

24

chemical agents, there are no data in the literature about simultaneous encapsulation and

25

delivery of biological and chemical agents.

26

Polysaccharides such as chitosan (abbreviation, CS) and alginate (abbreviation, ALG) are

27

biopolymers that easily create capsules in which an active ingredient can be incorporated

28

using an aqueous system at ambient temperature. Alginate is an anionic polysaccharide

29

composed of two repeating carboxylated monosaccharide units (manuronic and guluronic

30

acids) the ratio of which influences the properties of the biopolymer. Chitosan is a partially

31

deacetylated polymer of N-acetylglucosamine obtained after alkaline deacetylation of chitin.

32

The N-deacetylation is almost never complete and the fraction of N-acetylglucosamine

33

determines the degree of acetylation which serves as a base to classify the biopolymer as

34

chitin or chitosan. When the degree of N-acetylation (defined as the average number of N-

35

acetylglucosamine units per 100 monomers expressed as a percentage) is less than 50%, chitin

36

becomes soluble in aqueous acidic solutions (pH < 6.0) and is called chitosan. The

37

electrostatic attraction between the cationic amino groups of chitosan and the carboxylic

38

groups of the alginate lead to the formation of the polyelectrolyte complexes of various

39

structures. The structure and physicochemical properties of these complexes may be tailored

40

by controlling the degree of association among the functional groups.3 There are many

Despite the array of methodologies for simultaneous encapsulation of two

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

41

possible applications for CS and ALG complexes in the microcapsule form such as for

42

immobilization and the controlled release of various chemical or biologically active agents.4

43

It is well known that chemical elements (primary, secondary and micronutrients) are

44

important for plant growth and survival. Copper cations are an essential micronutrient for

45

plants and it is used for the management of a wide range of fungal and bacterial diseases in

46

various crops. Plants require copper cations for normal growth and development, and when it

47

is not available, plants develop specific deficiency symptoms, most of which affect young

48

leaves and reproductive organs.5 Thus, a deficiency in the copper cations supply can alter the

49

essential functions in plant metabolism. Copper cations have traditionally been used in

50

agriculture as plant pathogens control, and it is also extensively released into the environment

51

by other human activities that often cause environmental pollution. Therefore, it is important

52

to control the level of copper cations added to the plant. It is worth to notice that positive

53

charge of chitosan microcapsules loaded with copper cations facilitate better bioadhesion on

54

leaves that enables prolonged presence of copper cations on the leaf surface.6,7 The prolonged

55

release of copper cations from microcapsules and its prolonged presence on the leaves may

56

lead to reduced levels of copper cations needed for sufficient crop protection showing

57

microencapsulation is a better alternative to traditional applying of copper cations formulation

58

in plant protection.

59

Plants at all stages of their development, as well harvested products in storage, are susceptible

60

to attack by many pathogens that cause severe damage to plants. Biological agents

61

(biopesticides) may represent a new approach to plant protection.8,9 Biopesticides are based

62

on microorganisms such as bacteria, viruses, fungi, nematodes or natural substances,

63

including plant extracts and semi chemicals (e.g. insect pheromones). However, applications

64

of biopesticides are still limited to only a few percent of all active agents used for plant

65

protection.

ACS Paragon Plus Environment

Page 4 of 39

Page 5 of 39

Journal of Agricultural and Food Chemistry

66

Filamentous fungi Trichoderma viride (abbrevation, TV) is an opportunistic avirulent plant

67

symbiont as well as mycoparasite of plant pathogenic fungi. Its agricultural importance is

68

good antagonistic abilities against soil born plant pathogenic fungi thanks to different

69

mechanisms of antagonism, the production of antifungal metabolites (antibiosis), competition

70

for space and nutrients, induction of defense responses in plant and mycoparasitism.

71

Investigations revealed also the ability of T. viride to promote plant growth that was first

72

treated as a side effect of suppression of plant pathogenic fungi.10 Today is considered that the

73

direct effects of these fungi on plant growth and development are crucially important for

74

agricultural uses and for understanding the roles of T. viride in natural and managed

75

ecosystems.10,11 The dual roles of antagonistic activity against plant pathogens and promotion

76

of soil fertility make Trichoderma species a promising alternative to standard plant protection

77

and nutrition technologies.

78

The benefits of biopolymer microcapsules simultaneously loaded with copper cations and T.

79

viride are delayed and controlled release of bioactive agents, as well as prolonged fertilization

80

effect in order to supply plant with the protecting/nutrient agents during the whole vegetation.

81

Finally, only one application would be enough to cover plant active agent needs in the

82

vegetation period. There is not important economic aspect only, but also agroecological

83

because a reduced number of mechanization passes will result with lower soil compaction,

84

and thus the root would have much better conditions for growth and development.

85

In the light of the importance of T. viride as a biocontrol agent, showing the high ability to

86

accumulate and compatibility with macro- and micronutrients,12,13 simultaneous encapsulation

87

of copper cations and T. viride could give wider opportunities in both, the plant nutrition and

88

protection. The aim of this work was to prepare chitosan/alginate microcapsules

89

simultaneously loaded with copper cations and T. viride (abbreviation, CS/(ALG/(Cu+TV))

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

90

for plant protection and nutrition. Copper cation concentrations and microcapsule size were

91

considered as variables with possible influence on the essential microcapsule parameters.

92

2. MATERIALS AND METHODS

93

2.1 Materials. Low viscosity sodium alginate (CAS Number: 9005-38-3; Brookfield viscosity

94

4 - 12 cps (1% in H2O at 25oC) was purchased from Sigma Aldrich (USA). Low molecular

95

weight chitosan (CAS RN: 9012-76-4, molecular weight: 100 000–300 000) was obtained

96

from Acros Organic (USA). A commercially available product copper sulphate pentahydrate,

97

CuSO4 · 5H2O was used as a copper donating substance (Kemika Croatia). All other

98

chemicals were of analytical grade and used as received without further purification.

99

Indigenous isolate of T. viride (abbreviation, STP) originated from parasited sclerotia of

100

Athelia rolfsii was used in all experiments.

101

2.2. Preparation of Trichoderma viride isolates and spore suspension. The culture of the

102

isolate was grown in Petri dishes of 10 cm in diameter containing 20 ml of potato dextrose

103

agar (PDA, Biolife, Italy) plates and incubated in an incubator at 25 °C for 7 days until

104

conidiation occurred. To obtain spore suspensions, the STP was grown in Erlenmeyer flasks

105

containing potato dextrose broth (PDB). Flasks were inoculated with 5 PDA mycelial plugs.

106

Incubation took place at 22 ºC under the constant aeration for 10 days under illumination.

107

After this incubation period the liquid cultures with fungal biomass, consisting of hyphal

108

segments, chlamydiospores and conidia were filtrated by suction through the filter paper (595

109

Schleicher & Schuell; Whatman International, Ltd., Kent, England) so that the major parts of

110

mycelium was removed. Nebulisation of microencapsulated spores was checked in vitro using

111

an ultrasonic nebuliser (Omron Healthcare Europe, Netherlands).

112

2.3. Methods of microencapsulation. 2.3.1. Preparation of microcapsules viscous

113

dispersion. Copper cations loaded microcapsules (abbrevation, CS/(ALG/Cu)) viscous

ACS Paragon Plus Environment

Page 6 of 39

Page 7 of 39

Journal of Agricultural and Food Chemistry

114

dispersions (abbreviation, CCVD) were prepared by modifying known procedure7 as

115

described in Supporting Information (S2.3.1.).

116

2.3.2. Preparation of CS/(ALG/Cu) and CS/(ALG/(Cu+TV) microcapsules. Microcapsules

117

were prepared by the ionic gelation technique at ambient temperature. Preparation is rapid and

118

reliable, and microcapsules were obtained spontaneously under very mild conditions in two

119

stages. The first stage comprises the formation of core microcapsules loaded with copper

120

cations (abbrevation, ALG/Cu) or loaded with T. viride (abbrevation, ALG/(Cu+TV)), and the

121

second stage includes the coating of core microcapsules by chitosan. Details of preparation

122

are presented in Supporting Information (S2.3.2.1. and S2.3.2.2.).

123

2.4. Methods. 2.4.1. Influence of copper cations on Trichoderma viride growth and survival.

124

The cultures of fungus T. viride were maintained on potato dextrose agar (PDA, Biolife, Italy)

125

plates at 25 °C. Influence of copper cations concentration was examined by spraying CCVD

126

loaded with copper cations. Mycelial growth was observed visually.

127

The viability of spores was monitored according to ability of germination that is by measuring

128

changes in the concentration of T. viride expressed as the number of spores (NS) per 1 g of

129

dry microcapsule. Samples for measurements were prepared by dissolving 4 g of

130

microcapsules in 100 ml of a mixture (0.2 mol dm-3 NaHCO3 and 0.06 mol dm-3 Na2C6H5O7 x

131

2H2O). Dissolved microcapsules were mixed for 30 minutes with a magnetic stirrer at room

132

temperature. Five milliliters of a sample were filtered through the sterilized muslin cloth.14

133

Aseptic conditions were provided throughout the assay. The number of spores was

134

determined spectrophotometrically by a method of Waghunde et al.15 Stock solution of T.

135

viride spores (adjusted to 1.4 x 106 of spores/ml) used for calibration curve was diluted with

136

sterilized water. The number of spores in stock solution was determined with a

137

hemocytometer, using Neubauer counting chambers (Hirschmann EM Techcolor, Germany).

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

138

Absorbance was measured at 550 nm using UV-VIS spectrophotometer (UV-1700, Shimadzu,

139

Japan).

140

2.4.2. Determination of Trichoderma viride spores charge. The zeta potential (ζ/mV) of T.

141

viride spore was measured by Zetasizer Nano ZS (Malvern, UK). The zeta potential was

142

estimated from electrophoretic measurements using Henry equation: Ue = 2ε ζ (fκa/3η)

143

(1),

144

where ζ is the zeta potential, ε is the dielectric constant, Ue is the electrophoretic mobility and

145

η is the viscosity. fκa is in this case 1.5 and is referred to as the Smoluchowski

146

approximation.16 Deviations ranged within ±1 mV.

147

2.5. Characterization of delivery systems. 2.5.1. Microscopic observations.The size of

148

prepared microcapsules was measured using a light binocular. Diameters of about 100

149

microcapsules were measured. Microphotographs were taken by a Leica DFC295 digital

150

camera on a trinocular mount of a Leica MZ16a stereo-microscope (Leica Microsystems Ltd.,

151

Switzerland). Microcapsules were also observed by confocal laser scanning microscope

152

(CLSM, TCP SP2, Leica Lasertechnik, Germany). Samples for CLSM were stained with

153

eosin (0.01 % (w/v)). The microscope was operated in fluorescence and transmitted mode at

154

an acceleration voltage of 80 kV. All sample preparations for microscopic observation were

155

performed at room temperature.

156

2.5.2. Fourier Transform Infrared Spectroscopy. The Fourier transform infrared spectroscopy

157

(FTIR) spectra of the samples were recorded with the FTIR Instrument - Cary 660 FTIR

158

(MIR system) spectrometer (Agilent Technologies, USA). Dry microcapsules were crushed

159

with potassium bromide to get pellets. Spectral scanning was done in the range of 400-4000

160

cm-1.

ACS Paragon Plus Environment

Page 8 of 39

Page 9 of 39

Journal of Agricultural and Food Chemistry

161

2.5.3. Encapsulation efficiency, loading capacity, swelling degree and in vitro active agent

162

release. Encapsulation efficiency (EE) and bioactive agent loading capacity (LC)

163

determination were performed to obtain information on the yield and amount of copper

164

cations and T. viride spores encapsulated in microcapsules. The concentration of copper

165

cations was determined at λ = 795 nm and concentration of T. viride at 550 nm by a

166

spectrophotometer (Shimadzu, UV-1700).

167

The swelling degree (Sw) was determined in both CS/(ALG/Cu) and CS/(ALG/(Cu+TV))

168

microcapsules loaded with various amounts of copper cations to better understand

169

mechanisms of bioactive agents releasing from microcapsules. All measurements were made

170

in triplicate.

171

Detailed procedures of sample preparation for measurements and calculations are presented in

172

Supporting Information (from S2.5.3.1. to S2.5.3.4.).

173

3. RESULTS AND DISCUSSION

174

3.1. Copper cations effect on the growth and viability of Trichoderma viride spores. One

175

of the problems that should be solved in the simultaneous encapsulation of biological and

176

chemical agents is that the presence of active agents in the same compartment should not

177

diminish their activities. It is well known that T. viride belong to the group of

178

microorganisms, which can survive in high concentrations of different metals and have the

179

potential to bind them.12,13 However, this has been well explored only in the light of the

180

possible application of fungi in metal removal from solid and liquid substrates and industrial

181

wastes.

182

Visual observation after 10 days of CCVD application revealed high resistance of T. viride

183

spores to applicated copper cations concentrations. The presence of copper cations did not

184

inhibit the growth of fungi, i.e. none of the applied concentrations caused inhibition of

185

mycelial growth. On the contrary, the presence of copper cations promotes growth, e.g. at the

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

186

highest copper cation concentration the growth of T. viride was the most homogeneous (Fig.

187

1). It is in accordance with investigations showing that the copper cations concentration at

188

0.75 mmol dm-3 promotes the growth of Trichoderma species.13

189

Experiments performed on T. viride spores survival in CS/ALG/(Cu+TV) revealed that during

190

storage at the room temperature the number of spores was almost constant in the smaller

191

microcapsule, whereas in larger microcapsules somewhat increased (Table 1). It seems that

192

smaller microcapsules were fully loaded with T. viride spores and there was no space for their

193

germination. Smaller loading capacity (see later) and available internal volume in larger

194

microcapsules allowed spores to germinate indicating that copper does not inhibit their

195

activity.

196

Chemical analyses of the fungal cell walls have revealed a very complex chemical

197

composition that may be different in the several taxonomic groups.17 Cell walls have been

198

shown to be composed of polysaccharides (mainly glucan and chitin, up to 90%) and

199

glycoproteins, as well as of lipids and other minor components (pigments and inorganic salts).

200

Melanin, which in other fungi is associated with chitin, was shown to replace this polymer in

201

the spore wall of Trichoderma species and chitin was found only in mycelial cell walls. FTIR

202

spectrum of T. viride is presented in Fig. 2. Characteristics of T. viride spectrum are strong

203

and a broad peak at 3321 cm-1 assignable to amino (-NH) group superimposed on the side

204

hydroxyl (-OH) groups, frequencies at 2921 and 2854 cm-1 indicative of alkyl (>CH2) and

205

hydroxyl (-OH) groups, small peak at 2367 cm-1 assigned to the asymmetric stretching of the

206

isocyanate group (-N=C=O),18 the strong peak at 1625 cm-1 as well as several small peaks

207

between 1452 and 1200 cm-1 caused by stretching mode of carbonyl group (C=O) conjugated

208

to -NH deformation mode (indicative of amide bond formation).19 The small peak at 1545 cm-

209

1

210

deformation mode,20 and a strong peak at 1030 cm-1 represents C-F and C-Br vibrations. The

represents amine group stretching vibration resulting from -NH deformation mode to C=N

ACS Paragon Plus Environment

Page 10 of 39

Page 11 of 39

Journal of Agricultural and Food Chemistry

211

functional groups detected in the FTIR spectrum relate well with the the chemical structure of

212

the cell wall.

213

The characteristic peaks of copper sulfate pentahydrate at 3200, 1667, 1067 and 860 cm-1 are

214

presented in Fig. 2. The peaks over 3000 cm-1 represent the crystal water in the structure,

215

whereas peaks at lower band values correspond to the vibrations between O and nonmetal

216

atoms. A broad stretching frequency at 3200 cm-1 and a band of medium intensity at 1667 cm-

217

1

218

characteristic bands of inorganic sulfates ((SO4)2- stretching region).21

219

The spectrum of T. viride with bind copper cations shows much more intense and broad -OH

220

and -NH stretching vibrations band, the disappearance of T. viride bands at 2921, 2854 and

221

1545 cm-1, and the absence of small peaks between 1452 and 1200 cm-1. The disappearance of

222

bands and shifting of peaks towards the lower frequency (from 3321 to 3274 cm-1) or towards

223

higher frequency (1625 to 1635 cm-1, 1072 to 1087cm-1, and 887 to 981 cm-1) have suggested

224

that at least amine, hydroxyl, carbonyl and amide bonds are the major sites for binding of

225

copper cations.

226

Observation conducted by electron microscopy and cell fractionation studies revealed copper

227

cations location on the cell wall of T. viride spores indicating this is the place where the

228

interaction between T. viride and copper cations occurred.22 Our preliminary zeta potential

229

measurements performed on T. viride spores dispersed in water showed the broad charge

230

distribution (including negative and positive) with the average zeta potential of -35.1 mV. It

231

seems that the interaction between copper cations and T. viride spores primarily occurred due

232

to the electrostatic attractions.

233

3.2. Guidelines for the preparation of microcapsules simultaneously loaded with

234

biological and chemical agents. When a solution of sodium alginate comes into contact with

represent the bending modes of the hydroxyl group. Frequencies at 1067 and 860 cm-1 are

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

235

divalent or trivalent cations, a rapid, strong and irreversible formation of a gel takes place, e.g.

236

cations cooperatively interact with alginate blocks of guluronic acid residues forming a gel

237

network. Due to very rapid and irreversible binding reaction of gelling cations to alginate

238

chains, they should be mixed under controlled conditions.23,24 There are wide variations in the

239

procedure of alginate microcapsules preparation. The essential properties of all microcapsules

240

include high loading capacity, high mechanical and chemical stability, controllable releasing

241

and swelling properties, low toxicity, etc. The ideal microcapsules which meet all the

242

requirements do not exist, but in accordance with intended use, procedures of preparation

243

must be adapted. To obtain the well-designed microcapsules efficient for copper cations and

244

T. viride encapsulation, and prolonged release, it is important to choose an alginate with high

245

content of guluronic acid residues (a high guluronic acid content develop stiffer, more porous

246

gels, which maintain their integrity for longer periods of time) as well as the proper

247

concentration of alginate and gelling cation.23-25 The ratio between gelling cations and

248

alginate quantities in the gelation system determines the kinetics of gelation and the

249

characteristics of the gel formed.25 It is shown that the decrease of alginate concentration from

250

18 mmol dm-3 to ~ 7 mmol dm-3 increase the mechanical stability of alginate microcapsule,

251

but when the alginate concentration was further lowered to 4.5 mmol dm-3 mechanical

252

stability of capsules decreased.24,26 Another important parameter is the viscosity of alginate

253

solution. Goosen et al.27 showed that the minimum viscosity of alginate solution in order to

254

form a spherical microcapsule is 30 cPs. Spherical microcapsules are obtained over a wide

255

range of viscosity of the sodium alginate solution. Usually, the viscosity of the aqueous

256

sodium alginate solution does not exceed about 1000 cps. Having in mind that both amounts

257

of alginate and gelling cations concentration in solution affect markedly the cation binding

258

and kinetics of gel formation,26 we have investigated the impact of copper cations

ACS Paragon Plus Environment

Page 12 of 39

Page 13 of 39

Journal of Agricultural and Food Chemistry

259

concentration on essential microcapsule properties at an adjusted constant alginate

260

concentration convenient for uniform spherical microcapsules formation.

261

Chosen procedure took place in two stages by method, relying on ionic gelation and

262

polyelectrolyte complexation.28 The first stage involves the preparation of alginate core

263

microcapsules loaded with copper cations or simultaneously loaded with copper cations and

264

T. viride, and the second stage includes a coating of alginate core microcapsules by chitosan

265

(Scheme 1).

266

The aim of the second step was to reduce the porosity, improve stability and encapsulation

267

efficiency, and delay the release behavior.28,29 By dispersing core microcapsules in chitosan

268

acidic solution the chitosan rapidly bind onto their surface by electrostatic interactions

269

between protonated amino groups on chitosan and ionized carboxylic acid groups on

270

alginate.30 The electrostatic interaction between chitosan and alginate tightens and stabilizes

271

the surface of the microcapsules.31 Once the chitosan bind to the core microcapsules

272

competing ions (H+, Na+) have minor influence on the stability of the polyelectrolyte

273

complex,32 and the chitosan diffusion into the inner core is limited.33

274

3.3. Characterization of microcapsules loaded with bioactive agents. 3.3.1. Microscopic

275

observation. Examination of prepared microcapsules under a light binocular revealed

276

formation of spherical microcapsules which size was determined by the diameter of the funnel

277

or needle, respectively (Figs. 3a,b,c). They were colored due to the copper cations presence.

278

Microphotographs of two closely spaced spherical microcapsules taken under CLSM in

279

fluorescence mode and transmitted light clearly showed the existence of the chitosan layer on

280

the surface of microcapsules (Figs. 4a,b). Chitosan layer thickness became visible by staining

281

with eosin, which binds to the amino groups of chitosan. The thickness of coating layer is ~ 7

282

µm on microcapsules loaded only with copper cations (Fig. 4c) and ~ 11 µm loaded with

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

283

copper cations and T. viride (Fig. 4d), respectively. Obviously, the different core content

284

affected the thickness of coating layer, e.g. microcapsules loaded only with copper cations

285

exhibited somewhat thinner coating layer indicating the lower extent of reaction between

286

chitosan and sodium alginate molecules. It seems that electrostatic interaction between

287

protonated amino groups on chitosan and ionized carboxylic acid groups on alginate is

288

diminished due to the binding of T. viride to sodium alginate (see later in the section 3.3.2.).

289

Microphotographs of the core matrix in transmitted mode revealed almost homogenous

290

matrix texture of CS/(ALG/Cu) (Fig. 4e) and distributed T. viride spores throughout the core

291

matrix in CS/(ALG/(Cu+TV)) microcapsule (Figs. 4f,g). On air drying, the sphericity of

292

microcapsule was lost and its size somewhat decreased (Fig. 4h). The surface of the

293

microcapsules becomes rough containing many wrinkles. The occurrence of wrinkles on the

294

surface of the microcapsules can be explained by the syneresis.

295

3.3.2. Fourier transform infrared spectroscopy (FTIR). Information on intermolecular

296

interactions between alginate and active agents as well as in microcapsules coated with

297

chitosan was obtained using FTIR. FTIR spectra of sodium alginate, the mixture of sodium

298

alginate and T. spores, and core microcapsules ALG/Cu and ALG/(Cu + TV) are presented in

299

Fig. 5a, whereas spectra of chitosan and microcapsules coated with chitosan, CS/ALG/Cu and

300

CS/(ALG/(Cu TV), are presented in Fig. 5b.

301

Characteristic FTIR bands of sodium alginate with assignements listed in Table 2 are in

302

accordance with literature data.34 The spectrum of T. viride and alginate shows most intense

303

change of broad band from 3700 to 3000 as well as shifting of the other characteristic bands

304

of sodium alginate at 1405, 1295, 1125, 1081 and 1025 cm-1 to 1415, 1146, 1094 and 1038

305

cm-1. The disappearance of some bands (at 3198, 2925, 1595 cm-1 as well as bands attributed

306

to the ALG saccharide structure) is an indication of at least interactions with amine,

307

carboxylate and CO groups. Broader and somewhat shifted to a lower wave number as well

ACS Paragon Plus Environment

Page 14 of 39

Page 15 of 39

Journal of Agricultural and Food Chemistry

308

more intense (almost five times higher intensity) the -OH stretching vibrations band (around

309

3400 cm-1) suggested enhanced intermolecular hydrogen bonds.

310

The spectrum of ALG/Cu showed the most significant changes in the alginate functional

311

groups region: carboxylate (COO-), ether (COC) and hydroxyl (OH). It can be seen from the

312

Fig. 5a that the intensities of main alginate peaks become reduced and shifted to lower or

313

somewhat higher wavenumbers. Shifting of alginate broad band around 3400 cm-1 to lower

314

frequencies could be ascribed to copper cations interaction with hydroxyl groups. The

315

characteristic peak for asymmetric COO- stretching vibrations shifted from 1595 to 1590 cm-1

316

and the peak for symmetric COO- stretching vibrations shifted to higher wavenumber from

317

1405 to 1411 cm-1. Additional observed shifts are at 2925, 1081, 950 and 813 cm-1 to 2926,

318

1068, 944 and 811 cm-1, respectively. In agreement with FTIR data, X-ray diffraction

319

evidence and the egg-box model, Papageorgiou et al.35 proposed in the metal-alginate

320

complexes a "pseudo-bridged" unidentate coordination with intermolecular hydrogen bond in

321

polyguluronic regions and the bidentate bridging coordination the polymannuronic region.

322

In comparison with ALG/Cu spectrum, all peaks in ALG/(Cu+TV) spectrum are somewhat

323

more intense. Small shifts of alginate peaks (from 3400, 2925, 1405, 1295, 1125, 1025, 950

324

and 606 cm-1 to 3328, 2927, 1411, 1301, and 1026, 947 and 599 cm-1, respectively) and

325

disappearance of alginate peaks at 1125 and 1081 cm-1 indicated complex interactions

326

between all components in the microcapsule.

327

Characteristic FTIR bands of chitosan with assignements listed in Table 3 are in accordance

328

with literature data.36,37 Spectra of microcapsules with added chitosan (Fig. 5b) show some

329

peaks disappeared or become weaker due to the interaction between or superposition of the

330

groups of chitosan and alginate (at 2875, 1648, 1582, 1373 and 1150 cm-1).36 In comparison

331

with ALG/Cu spectrum (Fig. 5a), both spectra, CS/(ALG/Cu) and CS/(ALG/Cu+TV)) show a

332

little broader band around 3265 cm-1 due to enhanced hydrogen bonding. In comparison with

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

333

peaks in ALG/Cu spectrum all peaks in CS/(ALG/Cu) spectrum are less intense, but more

334

intense in CS/(ALG/Cu+TV)) spectrum. The absence of chitosan band at 1582 cm-1 (-NH

335

bending vibration) and shifting of alginate carboxylate region bands, from 1595 to 1585 cm-1

336

for CS/(ALG/Cu) and to 1590 cm-1 for CS/(ALG/Cu+TV), and from 1405 to 1409 cm-1 for

337

CS/(ALG/Cu) and to 1411 cm-1 for CS/(ALG/Cu+TV), respectively) indicated electrostatic

338

interactions between two oppositely charged polyelectrolytes.

339

3.3.3. Encapsulation efficiency and loading capacity. Results on copper cations encapsulation

340

efficiency and loading capacity are presented in Figs. 6a and 6b, respectively. Apparently, the

341

increase of copper cations concentration decreases encapsulation efficiency for CS/(ALG/Cu)

342

microcapsules, whereas microcapsules loaded with both active agents exhibits the increase in

343

encapsulation efficiency (Fig. 6a). Results are in accordance with the values of microcapsules

344

loading capacity (Fig. 6b). The decrease in encapsulation efficiency of microcapsules loaded

345

with copper cations is a consequence of loading capacity constancy, that is, microcapsules

346

reached maximum loading capacity even at the lowest copper cations concentration. On the

347

contrary, the loading capacity and consequently encapsulation efficiency by simultaneously

348

loaded microcapsules increased with copper cations concentration reaching values of

349

microcapsules without TV only at the highest copper cations concentration. The smaller

350

loading capacity of CS/(ALG/(Cu + TV)) in comparison with CS/(ALG/(Cu) microcapsules

351

can be explained by copper cations binding to T. viride, that is, the decrease in active copper

352

cations available due to binding to T. viride in solution. It is in accordance with results of

353

Bespalova et al.38 who have found that T. viride addition leads to a decrease in the amount of

354

copper cations bound to organic matter (by about 1.1-1.2-fold). The size of microcapsules

355

does not significantly affect the mode of changes in EE and LC with copper cations

356

concentration.

ACS Paragon Plus Environment

Page 16 of 39

Page 17 of 39

Journal of Agricultural and Food Chemistry

357

Spores of T. viride were not detected in filtrate after core microcapsule separation showing

358

the encapsulation efficiency was almost 100%. Data on the loading capacity expressed as the

359

number of spores per 1 g of dry microcapsule are listed in Table 1.

360

3.3.4. Swelling degree. On in vivo applying, alginate microcapsules coated with chitosan

361

usually swell thus influencing the release of active agents from them. In comparison with

362

uncoated alginate microcapsules, those coated with chitosan exhibit higher swelling degree

363

due to highly hydrophilic nature of polyelectrolyte complex formed.31,39-42 The swelling

364

degree of alginate-chitosan complex depends on the extent of electrostatic interaction between

365

the alginate carboxylate groups and protonated amine groups of chitosan. Both, sodium

366

alginate and chitosan are weak polyelectrolytes and the degree of dissociation of their

367

functional groups strongly depends on the pH of the solution. Sodium alginate chains with

368

pKa values of 3.38 and 3.65 for mannuronic and guluronic acids, respectively, are negatively

369

charged across a wide range of pH values, whereas chitosan is positively charged (protonated)

370

below its pKa of 6.5.43 Swelling degree of coating layer is low in acidic medium due to the

371

dense polyelectrolyte complex structure as a result of strong interactions between protonated

372

amino groups of chitosan and carboxylate anions of alginate. With pH increasing the

373

carboxylic acid groups become more ionized. Close to pH ~ 6.5 chitosan become

374

deprotonated and consequently the extent of complexation is reduced forming a less dense

375

structure of coating layer which alows higher swelling.43

376

When chitosan/alginate microcapsules loaded with copper, and with copper and T. viride were

377

dispersed in deionized water (pH ~ 6) they started to swell like a microcapsule consisting of a

378

chitosan wall and Ca-alginate core.40,43 During three hours of exposure in deionized water the

379

pH of solutions decreased from pH 6 to ~ 5. The decrease of pH may be attributed to the

380

small quantity of H+ ions released due to ionization of carboxylic groups.31

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

381

The effect of increasing copper cations concentration on swelling degree is presented in Fig.

382

7. It can be clearly seen that higher copper cations concentration promoted swelling, but there

383

existed obvious differences regarding loaded agents and microcapsule size. The higher

384

swelling extent of microcapsules loaded with copper cations than those simultaneously loaded

385

with copper cations and T. viride may be attributed to the difference in chitosan layer

386

thickness (Figs. 4c,d), e.g. the increase of coating layer thickness resulted in a decrease of

387

swelling extent. Bartkowiak et al.29 have shown that the chitosan coating layer thins with an

388

increase in microcapsule size prepared under the same conditions. Accordingly, the

389

differences in swelling behavior between differently sized microcapsules can be also

390

explained by differences in chitosan layer thickness.

391

3.3.5. In vitro release of active agents. The possible use of biopolymer microcapsules in the

392

plant imposes research regarding their release capacity in certain physicochemical conditions.

393

In this direction, the kinetics of the copper cations ion release was studied on microcapsules

394

prepared at the highest copper cations concentration. This concentration was used because of

395

its positive affect on the growth and behavior of T. viride spores. The release profiles of

396

copper cations from different types of sodium alginate/chitosan microcapsules are presented

397

in Fig. 8. A set of curves for copper cations release exhibit rapid initial release followed by

398

slower release obeying power law equation.

399

It can be clearly seen that the amount of copper cations released depends on microcapsule size

400

and loaded active agents. The microcapsules of different sizes prepared under the same

401

conditions differ in surface layer thickness and it should be possible to correlate most

402

microcapsule properties by taking into the account the changes in microcapsule

403

surface/volume ratio.29 It was observed that chitosan layer thins with the increase of

404

microcapsule diameter causing some changes in mechanical and permeability properties.

405

Faster releasing from the large microcapsules may be attributed to the thinner surface layer

ACS Paragon Plus Environment

Page 18 of 39

Page 19 of 39

Journal of Agricultural and Food Chemistry

406

allowing easier penetration of both the dissolution medium (swelling) and the copper through

407

the layer. Large microcapsules released a larger amount of copper cations, because there were

408

more copper cations loaded per microcapsule than in smaller microcapsules (Fig. 6b). Despite

409

to the somewhat thinner surface layer (Fig. 4e), both small and large microcapsules loaded

410

with copper and T. viride released smaller amount of copper. This may be attributed to the

411

binding of copper cations to T. viride.

412

Various mechanisms such as desorption from the surface, diffusion through the pore of the

413

core and wall, microcapsule disintegration, dissolution or erosion of the structure, or on their

414

combination, may be included in the release of active agents from microcapsules. Mechanism

415

of active agents releasing primarily depends on the characteristics of core material and active

416

agents as well as on the microcapsule size.44

417

To identify the type of mechanism involved in copper cations releasing a semi-empirical

418

Korsmeyer–Peppas model was applied.45,46 According to Korsmeyer–Peppas, the release

419

exponent n can be characterized by three different mechanisms (Fickian diffusion, (n)

420

anomalous (non Fickian diffusion), or Type II transport). Values of n < 0.43 indicates the

421

release is controlled by classical Fickian diffusion, n > 0.85 is controlled by Type II transport,

422

involving polymer swelling and relaxation of the polymeric matrix, whereas values of n

423

between 0.43 and 0.85 shows the anomalous transport kinetics determined by a combination

424

of the two diffusion mechanisms and Type II transport.

425

All curves presented in Fig. 8 can be described by the equation:

426

ோ೟

݂ ሺ‫ݑܥ‬ሻ = ோ

= ݇‫ ݐ‬௡

(2),

೟೚೟

427

where f(Cu) represents the fraction of released copper cations, Rt is the amount of copper

428

cations released at time t, Rtot is the total amount of Cu loaded in capsules, k is a constant

429

characteristic of the active agents/polymer system that considers structural and geometrical

430

aspects of the system, and the value of the exponent (n) is an exponent which characterize the

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

431

transport mechanism of active agents through microcapsule. The values of the release

432

constants and exponents are listed in Table 4.

433

Lower n values than 0.43 for larger microcapsules indicated that the release mechanism

434

involved is controlled by a classical Fickian diffusion, whereas higher n values for smaller

435

microcapsules indicated copper cations release followed non-Fickian kinetics, due to probable

436

rapid swelling and partial dissolution of microcapsules.

437

Change of released T. viride amount (expressed as spore number, NS/g) over time presented

438

in Fig. 9. shows that after initial lag time the amount of spores in the medium increases

439

exponentially. The processes involved during the lag phase are penetration of water and

440

filling of the microcapsule surface pore with water, as well as transport of T. viride through

441

the alginate matrix and the water-filled chitosan pores, and finally diffusion in the

442

surrounding media. Slower T. viride diffusion through alginate matrix and water filled pores

443

can be ascribed to higher size of T. viride spores in comparison with the size of copper cations

444

as well to their intermolecular interactions with alginate and copper cations. As the swelling

445

process progressed, transport of T. viride toward water phase occurred through the swollen

446

layer. The exponential increase in the amount of T. viride and much higher concentration

447

detected in water than concentration loaded in microcapsule indicated that spores germinated

448

and germ tube biomass was formed in the surrounding medium. Actually, copper cations

449

released from microcapsule promote germination in water. The increasing amount of T. viride

450

in the dispersing medium is closely related with two processes, one is the release of T. viride

451

from microcapsules and the other is germination. It can be seen that smaller microcapsules

452

deliver larger amount of T. viride to the surrounding medium and consequently exhibited

453

greater yield of biomass.

454

3.3.6. Concluding remarks

ACS Paragon Plus Environment

Page 20 of 39

Page 21 of 39

Journal of Agricultural and Food Chemistry

455

We reported, for the first time, the possibility to encapsulate both, biological and chemical

456

bioactive agents in a chitosan/alginate microcapsule. Encapsulation of T. viride spores and the

457

presence of copper cations in the same compartment do not inhibit their activity and even it

458

promoted T. viride germination. Investigation of intermolecular interactions between

459

oppositely charged biopolymers and bioactive agents using FTIR spectroscopy revealed an

460

interaction between copper cations and T. viride spores functional groups as well between

461

alginate and bioactive agents.

462

Loading capacity and efficiency, swelling behavior and releasing of active agent depend on

463

both, the kind of encapsulated agents and microcapsule size. The increase in copper cations

464

concentration promoted swelling, but there exist obvious differences regarding to the loaded

465

agents and microcapsule size. Higher swelling extent and greater amount of released copper

466

cations from microcapsules loaded only with copper cations than those simultaneously loaded

467

with copper cations and T. viride spores can be attributed to the differences in chitosan layer

468

thickness.

469

The in vitro copper cations release profile was fitted to Korsmeyer–Peppas empirical model.

470

Fickian diffusion was found to be the rate-controlling mechanism at smaller microcapsules,

471

whereas anomalous transport kinetics (combination of the diffusion mechanisms and Type II

472

transport) controlled release from larger microcapsules. The copper cations release exhibited

473

initial burst followed by a slower release, but Trichoderma viride spores releasing profile

474

showed exponential increasing over initial lag time. Much slower release of T. viride spores at

475

the early stage may be ascribed to their higher size in comparison with copper cations, and

476

intermolecular interactions with alginate and copper cations.

477

Results showed chitosan/alginate microcapsules can simultaneously incorporate T.viride

478

spores and chemical bioactive agent without inhibiting their activities. With all these results it

479

seems that copper cations and T. viride spores loaded in chitosan/alginate microcapsules can

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

480

be suitable for plant nutrition and protection and that is worth to preform further research. Our

481

future investigations are directed to test CS/ALG microcapsules simultaneously loaded with

482

chemical and biological agents on plants under greenhouse conditions as well as in an open

483

field.

484 485 486 487 488 489 490

ASSOCIATED CONTENT Supporting Information. Additional information on microcapsule preparation and methods of characterization as noted in the text. This material is available free of charge via the Internet at http://pubs.acs.org. AUTHOR INFORMATION * Corresponding author

491 492 493 494

Marko Vinceković Department of Chemistry, University of Zagreb Faculty of Agriculture Svetošimunska cesta 25, 10000 Zagreb, Croatia E-mail: [email protected] ; Tel: + 385 1 239 3953

495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512

Authors’ contributions The corresponding author designed the research work and wrote the manuscript. All the authors carried out this research work under the guidance of the corresponding author. The final manuscript was read, corrected, improved and approved by all coauthors. Funding We acknowledge the financial support of Croatian Science Foundation (Project: UIP-201409-6462). Notes The authors declare no competing financial interest. ABBREVIATIONS USED ALG, alginate; CS, chitosan; TV, Trichoderma viride (T. viride); ALG/Cu, alginate core microcapsule loaded with copper cations; CS/(ALG/Cu), chitosan/alginate microcapsule loaded with copper cations, CS/(ALG/(Cu+TV)), chitosan/alginate microcapsule loaded with copper cations and T. viride; PDA, potato dextrose agar; PDB, potato dextrose broth; STP, indigenous isolate of T. viride; EE, encapsulation efficiency, LC, loading capacity, Sw, swelling degree; Rt, the amount of Cu released at time t; Rtot, the total amount of Cu loaded in capsules.

ACS Paragon Plus Environment

Page 22 of 39

Page 23 of 39

Journal of Agricultural and Food Chemistry

513

REFERENCES

514

1. Rodkham, D.K. Colloid and interface science in formulation research for crop protection

515

products, Curr. Opin. Colloid Interface Sci. 2000, 5, 280–287).

516

Nuruzzaman, M.; Rahman, M.M.; Liu, Y.; Naidu R. Nanoencapsulation, Nano-Guard for

517

Pesticides: A New Window for Safe Application, J. Agric. Food Chem. 2016, 64, 1447–1483

518

3. Lawrie, G.; Keen, I.; Drew,B.; Chandler-Temple, A.; Rintoul, L.; Fredericks, P.; Grøndahl,

519

L. Interactions between alginate and chitosan biopolymers characterized using FTIR and

520

XPS. Biomacromolecules 2007, 8, 2533–2541.

521

4. Sundar, S.; Kundu, J.; Kundu, S.C. Biopolymeric nanoparticles. Sci. Technol. Adv. Mat.

522

2010, 11, 014104–014115.

523

5. Yruela, I. Copper cations in plants: acquisition, transport and interactions. Functl. Plant

524

Biol.. 2009, 36, 409–430.

525

6. Hadwiger, L.A.; Mc Bride, P.O. Low-Level copper cations plus chitosan applications

526

provide protection against late blight of potato. Online. Plant Health Prog. 6 April 2006, 1-7.

527

DOI:10.1094/PHP-2006-0406-01-RS.

528

(http://www.plantmanagementnetwork.org/pub/php/research/2006/chitosan/)

529

7. Jalšenjak, N. Preparation of copper cations-loaded microcapsule formulations. Agri. Consp.

530

Sci. 2011, 76, 115–119.

531

8. Chandler, D.; Bailey, A.; Tatchell, G.M.; Davidson, G.; Greaves, J.; Grant, W.P. The

532

development, regulation and use of biopesticides for integrated pest management. Philos.

533

Trans. R. Soc. Lond. B 2011, 386, 1987–1998.

534

9. Seiber, J.N.; Coats,J.; Duke,S.O.; Gross, A.D. Biopesticides: State of the art and future

535

opportunities. J. Agric. Food Chem.2014, 62, 11613–11619.

536

10. Topolovec-Pintarić, S.; Žutić, I.; Đermić, E. Enhanced growth of cabbage and red beet by

537

Trichoderma Viride, Acta Agr. Slovenica 2013, 101, 87–92.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

538

11. Celar, F.; Valič, N. Effects of Trichoderma Spp. and Gliocladium Roseum culture filtrates

539

on seed germination of vegetables and maize. J. Plant Dis. Prot. 2005, 112, 343–350.

540

12. Valarmathi, P.; Pareek, S. K.; Priya, V.; R. Rabindran, Chandrasekar, G. Compatibility of

541

copper cations hydroxide (Kocide 3000) with biocontrol agents. IOSR J. Agr. Veter. Sci.

542

(IOSR-JAVS) 2013, 3, 28–31.

543

13. Jovičić Petrović, J.; Danilović, G.; Ćurčić, N.; Milinković, M.; Stošić, N.; Panković, D.;

544

Raičević, V. Copper cations tolerance of Trichoderma species. Arch. Biol. Sci. 2014, 66, 137–

545

142.

546

14. Trinci, A. P. J. Culture turbidity as a measure of mould growth. Trans. British

547

Mycological Soc. 1972, 58, 467–474.

548

15. Waghunde, R.R.; Priya, J.; Naik, B.M.; Solanky, K.U.; Sabalpara, A.N. Optical density -

549

A tool for the estimation of spore count of Trichoderma Viride. J. Biopesticides, 2010, 3,

550

624–626.

551

16. Delgado, A.V.; Gonzales-Caballero, F.; Hunter, R.J.; Koopal, L.K.: Lyklema, J.

552

Measurement and interpretation of elektrokinetic phenomena. Pure Appl. Chem., 2005, 77,

553

1753–1805.

554

17. Gruber, S.; Seidl-Seiboth, V.

555

Trichoderma. Microbiology. 2012, 158, 26-34.

556

18. Singh, R.; Chadetrik, R.; Kumar, R.; Bishnoi, K.; Bhatia, D.; Kumar, A.; Bsihnoi, N.R.;

557

Singh, N. Biosorption optimization of lead(II), cadmium(II) and copper cations(II) using

558

response surface methodology and applicability in isotherms and thermodynamics modeling.

559

J. Hazard. Mater. 2010, 174, 623–634.

560

19. Barbotin, J.N.; Nava Saucedo, J.E. Bioencapsulation of living cells by entrapment in

561

polysaccharide gels. In S. Dimitriu, Ed., Polysaccharides: Structural Diversity and

562

Functional Versatility, Marcel Dekker: New York, 1998; 749–774.

Self versus non-self: fungal cell wall degradation in

ACS Paragon Plus Environment

Page 24 of 39

Page 25 of 39

Journal of Agricultural and Food Chemistry

563

20. Saeed, A.; Iqbal, M.; Zafar, I.S. Immobilization of Trichoderma viride for enhanced

564

methylene blue biosorption: batch and column studies. J. Hazar. Mater. 2009, 168, 406–415.

565

21. Gamo, I. Infrared Absorption spectra of water crystallization in copper sulfate penta- and

566

monohydrate crystals. Bull. Chem. Soc. Jpn. 1961, 34, 764–766.

567

22. Anand, P.; Isar, J.; Saran, S.; Saxena, R. K. Bioaccumulation of copper cations by

568

Trichoderma Viride. Bioresour. Technol. 2006, 97, 1018–1025.

569

23. Irmanida, B.; Devi, R.; Kusdiantoro, M.; Wahono Esthi, P. Leydig cells encapsulation

570

with alginate-chitosan: optimization of microcapsule formation“. J. Encaps. Adsorp. Sci.

571

2012, 2, 15–20.

572

24. Daly, M.M.; Knorr, D. Chitosan-alginate complex coacervate capsules: effects of calcium

573

chloride, plasticizers, and polyelectrolytes on mechanical stability. Biotechnol. Prog. 1988, 4,

574

76–81

575

25. Selimoglu, S.M.; Elib, M. Alginate as an immobilization material for MAb production via

576

encapsulated hybridoma cells. Crit. Rev. Biotechnol. 2010, 30, 145–159.

577

26. Rodrigues, J.R.; Lagoa, R. Copper cations ions binding in Cu‐alginate gelation. J.

578

Carbohydrate Chem. 2006, 25,219–232.

579

27. Goosen, M.F.A.; O’Shea G.M.; Sun, M.F. Micro-encapsulation of living tissue and cells.

580

US Patent 1987, No. 4, 806, 355; http:// www.google.com/patents /US4806355).

581

28. Gasseröd, O.; Sannes, A.; Skjἅk-Bræk, G. Microcapsules of alginate-chitosan-I. A

582

Quantitative study of the interaction between alginate and chitosan, Biomaterials 1998, 19,

583

1815–1825.

584

29. Bartkowiak, A.; Hunkeler, D. Alginate-Oligochitosan Microcapsules. II. Chem. Mater.

585

2000, 12, 206–212.

586

30. Fwu, L. Shin, M. S. S.; Hsing, W. S. Drug release from chitosan–alginate complex beads

587

reinforced by a naturally occurring cross-linking agent. Carbohyd.Polym. 2002, 48, 61–72.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

588

31. Shi, J.; Alves, N.M.; Mano, J.F. Chitosan coated alginate beads containing

589

Poly(Nisopropylacrylamide) for dual-stimuli-responsive drug release. J. Biomed. Mater. Res.

590

Part B: Appl. Biomater. 2008, 84, 595–603.

591

32. Gasseröd, O.; Sannes, A.; Skjåk-Bræk, G. Microcapsules of alginate-chitosan. II. A study

592

of capsule stability and permeability. Biomaterials 1999, 20, 773–783.

593

33. Draget, K. I.; Smidsrød, O.; Skjåk-Bræk, G. Alginates from algae. In Polysaccharides and

594

polyamides in the food Industry: properties, production, and patents, Steinbüchel, A. S.;

595

Rhee, K. Eds.;Wiley-VCH: Weinheim, 2005, Vol. 1, 1-30.

596

34. Sartori, C.; Finch, D.S.; Ralph, B. Determination of the cation content of alginate thin

597

films by FTIR Spectroscopy. Polymer 1997, 38, 43–51.

598

35. Papageorgiou, S.K.; Kouvelos, E.P.; Favvas, E.P.; Sapalidis, A.A.; Romanos, G.E.;

599

Katsaros, F.K. Metal–carboxylate interactions in metal–alginate complexes studied with FTIR

600

Spectroscopy. Carbohyd. Res. 2010, 345, 469–473.

601

36. Sankalia, M.G.; Mashru, R.C.; Sankalia, J.M.; Sutariya, V.B. Reversed chitosan-alginate

602

polyelectrolyte complex for stability improvement of alpha-amylase: optimization and

603

physicochemical characterization. Eur. J. Pharm. Biopharm. 2007, 65, 215–232.

604

37. Khor, E. The structural properties of chitin as it are known today. Chitin: fulfilling a

605

biomaterials promise, Elsevier, New York, 2001, pp. 73–82.

606

38. Bespalova, A.Yu.; Motuzova, G.V.; Marfenina, O.E. Secondary mobilization of heavy

607

metals in polluted soils under microbial Influence (model experiment). Develop. Soil Sci.

608

2002, 28B, 187–193.

609

39. Silva, R.M.; Silva, G.A.; Coutinho, O.P.; Mano, J.F. Reis, R.L. Preparation and

610

characterisation in simulated body conditions of glutaraldehyde crosslinked chitosan

611

membranes. J. Mater Sci. Mater. In Med. 2004, 15, 1105–1112.

ACS Paragon Plus Environment

Page 26 of 39

Page 27 of 39

Journal of Agricultural and Food Chemistry

612

40. Lee, O. S.; Ha, B.J.; Parka, S.N.; Lee, Y.S. Studies on the pH-dependent swelling

613

properties and morphologies of chitosan/calcium/alginate complexed beads. Macromol.

614

Chem. Phys. 1997, 198, 2971–2976.

615

41. Rajendran, A.; Basu, S. K. Alginate-chitosan particulate system for sustained release of

616

nimodipine. Trop. J. Pharm. Res. 2009, 8, 433–440.

617

42. Zhang, L.; Guo, J.; Peng, X.; Jin, Y. Preparation and release behavior of

618

carboxymethylated chitosan/alginate microspheres encapsulating bovine serum albumin. J.

619

App. Polymer Sci. 2004, 92 878–882.

620

43. Liu, X.; Xue, W.; Liu, Q.; Yu, W.; Fu, Y.; Xiong, X.; Ma, X.; Yuan, Q. Swelling

621

behaviour of alginate–chitosan microcapsules prepared by external gelation or internal

622

gelation technology. Carbohyd. Polym. 2004, 56, 459–464.

623

44. Polaković, M.; Gӧrner, T.; Gref, R.; Dellacherie, E. Lidocaine loaded biodegradable

624

nanospheres, II Modelling of drug release. J. Control. Release. 1999, 60, 169–177.

625

45. Korsmeyer, R. W.; Gurny, R.; Doelker, E.; Buri, P.; Peppas, N.A. Mechanisms of solute

626

release from porous hydrophilic polymers. Int. J. Pharm. 1983, 15, 25–35.

627

46. Grillo, R.; De Melo, N.F.S.; De Araujo, D.R.; De Paula, E.; Rosa, A.H.; Fraceto, L.F.

628

Polymeric alginate nanoparticles containing the local anesthetic bupivacaine. J. Drug Target.

629

2010, 18, 688–699.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure Caption Scheme 1. Schematic diagram of the experimental setup corresponding to ionic gelation (the first stage - I) and polyelectrolyte complexation (the second stage - II). Figure 1. Microphotographs of the mycelial growth of Trichoderma viride spores sprayed with CCVD prepared at increasing initial copper cations concentration (ci) = 4.5, 9 and 18 mmol dm-3 (from left to right side) taken after spraying. Figure 2. FTIR spectra of Trichoderma viride spores (black line), copper sulfate pentahydrate (cyan line) and their mixture (red line). Figure 3. Microcapsules prepared with (a) needle (0.45 mm) and (b) funnel (2 mm), (c) enlarged picture of a microcapsule prepared with funnel (black bar in the background = 0.5 mm). Figure 4. CLSM microphotographs: parts of two closely spaced microcapsules in fluorescence (a) and transmitted mode (b); part of the CS/(ALG/Cu) microcapsule in fluorescence (c) and transmitted mode (d); part of the CS/(ALG/(Cu + TV)) microcapsule in fluorescence (e) and transmitted mode (f); cross-section of CS/(ALG/(Cu + TV) microcapsule matrix obtained by optical microscopy (g); image of dried CS/(ALG/(Cu + TV)) microcapsule obtained by optical microscopy (h). Microcapsules are prepared at the initial copper cations concentration, ci = 18 mmol dm-3. Bars are indicated. Figure 5. FTIR spectra of (a) sodium alginate (ALG - black line), alginate and Trichoderma viride (ALG/TV - red line), core microcapsule (ALG/Cu - green line), core microcapsule with copper cations and Trichoderma viride (ALG/(Cu+TV) - blue line); (b) chitosan (CS - black line), core microcapsule with copper cations ((CS/ALG/Cu) - blue line), and core microcapsule with copper cations and Trichoderma viride coated with chitosan (CS/ALG/(Cu+TV) - red line).

ACS Paragon Plus Environment

Page 28 of 39

Page 29 of 39

Journal of Agricultural and Food Chemistry

Figure 6. Variation of copper cations (a) encapsulation efficiency (EE) and (b) loading capacity (LC) of CS/(ALG/Cu) (open signs) and CS/(ALG/(Cu+TV)) (full signs) microcapsules with initial copper cations concentration (ci). Microcapsule diameters are denoted in brackets. The error bars indicate the standard deviation of the means. Figure 7. Variation of the swelling degree (Sw) of CS/(ALG/Cu) (open signs) and CS/(ALG/(Cu + TV)) (full signs) microcapsules in deionized water with initial copper cations concentration (ci). Microcapsule diameters are denoted in brackets. The error bars indicate the standard deviation of the means. Figure 8. Fraction of released copper cations, f(Cu), from CS/(ALG/Cu) (open signs) and CS/((ALG/(Cu+TV)) (full signs) microcapsules at initial copper cations concentration c(Cu)i = 18 mmol dm-3 with time (t). Microcapsule diameters are denoted in brackets. The error bars indicate the standard deviation of the means. Figure 9. Variation of a number of Trichoderma viride spores (NS g-1) with time (t). Microcapsule diameters are denoted in brackets. The error bars indicate the standard deviation of the means.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Table 1. Number of Trichoderma Viride Spores (NS) per 1 g of CS/(ALG/(Cu+TV)) after 1 and 10 Days of Storage at Room Temperature. size / mm

NS/g (1 day)

NS/g (10 days)

0.45

14444

14188

2

6500

7750

Table 2. FTIR Bands of Sodium Alginate with Assignments. sodium alginate vibration (cm-1)

assignment

3700 - 3000

O-H stretching

3198 2925

O-H stretching (intermolecular hydrogen bond) C-H stretching

1595

COO- stretching (asymmetric)

1405

COO- stretching (symmetric)

1295

C-O stretching

1125

C-C stretching

1081- 1027

assymmetric C-O-C stretching

1033

C-O stretching

950 cm-1

C-O stretching of uronic acid

ACS Paragon Plus Environment

Page 30 of 39

Page 31 of 39

Journal of Agricultural and Food Chemistry

Table 3. FTIR Bands of Chitosan with Assignments. Reversed chitosan-alginate chitosan vibration (cm-1)

assignment

3750-3000

O-H and N-H stretching

2925

C-H in CH2 stretching

2875

C-H in CH3 stretching

1648 1582

C=O stretching of the secondary amide (amide band I) N-H bending of amine and amide II

1425

CH2 bending

1373

CH3 smmetrical deformation

1150

assymmetric C-O-C and C-N stretching

1026

skeletal vibration of C-O stretching

Table. 4. The Values of the Release Constant (k) and Exponent (n) of Copper Cations Encapsulated in CS/(ALG/Cu) and CS/(ALG/(Cu+TV)) microcapsules. microcapsule

size (mm)

k (day-1)

n

CS/(ALG/Cu)

0.45

0.167

0.45

CS/(ALG/Cu)

2.0

0.551

0.23

CS/(ALG/(Cu + TV))

0.45

0.081

0.68

CS/(ALG/(Cu + TV))

2.0

0.436

0.27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Scheme 1.

Figure. 1.

ACS Paragon Plus Environment

Page 32 of 39

Page 33 of 39

Journal of Agricultural and Food Chemistry

1,1 1,0

Transmittance

0,9 0,8 0,7 0,6 0,5 0,4

Trichoderma viride CuSO4 x 5H2O

0,3

Trichoderma viride + CuSO4 x 5H2O

0,2 3500

3000

2500

2000

1500

1000

500

-1

Wavenumber (cm )

Figure 2.

(a)

(b) Figure 3.

ACS Paragon Plus Environment

(c)

Journal of Agricultural and Food Chemistry

(a)

(b)

(c)

(d)

(e)

(f)

(g)

(h)

Figure 4.

ACS Paragon Plus Environment

Page 34 of 39

Page 35 of 39

Journal of Agricultural and Food Chemistry

1,1

(a)

1,0

Transmittance

0,9 0,8 0,7 0,6 ALG ALG/TV ALG/Cu ALG/(Cu+TV)

0,5 0,4 0,3 0,2 3500

3000

2500

2000

1500

1000

500

-1

Wavenumber (cm )

Figure 5a.

1,00

(b)

0,95

Transmittance

0,90 0,85 0,80 0,75

CS/(ALG/Cu) CS/(ALG/(Cu+TV)) CS

0,70 0,65 0,60 3500

3000

2500

2000

1500

1000

-1

Wavenumber (cm )

Figure 5b.

ACS Paragon Plus Environment

500

Journal of Agricultural and Food Chemistry

CS/(ALG/(Cu) (0.45 mm)

(a)

CS/(ALG/Cu ) (2 mm)

80

CS/(ALG/(Cu + TV)) (0.45 mm)

EE / %

CS/(ALG/(Cu + TV)) (2 mm)

60

40

8

10

12

14

16

ci / mmol dm

18

-3

Figure 6a.

3,0

(b)

CS/(ALG/Cu) (0.45 mm)

2,5

CS/(ALG/Cu) (2 mm)

LC / mmol g-1

CS/(ALG/(Cu + TV)) (0.45 mm) CS/(ALG/(Cu + TV)) (2 mm)

2,0

1,5

1,0

0,5

8

10

12

14

ci / mmol dm

16

-3

Figure 6b.

ACS Paragon Plus Environment

18

Page 36 of 39

Page 37 of 39

Journal of Agricultural and Food Chemistry

CS/(ALG/(Cu) (0.45 mm)

90

CS/(ALG/Cu ) (2 mm) CS/(ALG/(Cu + TV)) (0.45 mm)

80

CS/(ALG/(Cu + TV)) (2 mm)

Sw / %

70 60 50 40 30 8

10

12

14

16

18

-3

ci / mmol dm

Figure 7.

1,0

0,8

f(Cu)

0,6

0,4 CS/(ALG/(Cu) (0.45 mm)

0,2

CS/(ALG/Cu ) (2 mm) CS/(ALG/(Cu + TV)) (0.45 mm)

0,0

CS/(ALG/(Cu + TV)) (2 mm)

0

2

4

6

8

10

12

t / days

Figure 8.

ACS Paragon Plus Environment

14

Journal of Agricultural and Food Chemistry

2x105 CTS/(ALG/(Cu + TV)) (2.0 mm)

NS g-1

CTS/(ALG/(Cu + TV)) (0.45 mm)

1x105

0 0,0

0,5

1,0

1,5

2,0

2,5

t / week

Figure 9.

ACS Paragon Plus Environment

3,0

Page 38 of 39

Page 39 of 39

Journal of Agricultural and Food Chemistry

For Table of Contents Only

Figure caption

ACS Paragon Plus Environment