Preparation of Fucoxanthin-Loaded Nanoparticles Composed of

Nov 23, 2016 - File failed to load: https://cdn.mathjax.org/mathjax/contrib/a11y/accessibility-menu.js .... To facilitate the utilization of fucoxanth...
2 downloads 9 Views 1MB Size
Subscriber access provided by UNIV NEW ORLEANS

Article

Preparation of Fucoxanthin-Loaded Nanoparticles Composed of Casein and Chitosan with Improved Fucoxanthin Bioavailability Song Yi Koo, Il-Kyoon Mok, Cheol-Ho Pan, and SANG MIN KIM J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b04376 • Publication Date (Web): 23 Nov 2016 Downloaded from http://pubs.acs.org on December 6, 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 33

Journal of Agricultural and Food Chemistry

Preparation of Fucoxanthin-Loaded Nanoparticles Composed of Casein and Chitosan with Improved Fucoxanthin Bioavailability

Song Yi Koo†, Il-Kyoon Mok†,§, Cheol-Ho Pan†, Sang Min Kim†,*



Systems Biotechnology Research Center, KIST Gangneung Institute of Natural Products,

Gangneung, Gangwon-do 25451, Republic of Korea

§

Department of Food Processing and Distribution, Gangneung-Wonju National University,

Gangneung, Gangwon-do 210-702, Republic of Korea

*Convergence Research Center for Smart Farm Solution, KIST Gangneung Institute of Natural Products, Gangneung, Gangwon-do 25451, Republic of Korea

*Corresponding author: Sang Min Kim, Tel: +82-33-650-3640, Fax: +82-33-650-3679 E-mail: [email protected]

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

1

 ABSTRACT

2 3

To facilitate the utilization of fucoxanthin (FX), a valuable marine carotenoid, in the food

4

industry, FX-loaded casein nanoparticles (FX-CN) and chitosan-coated FX-CN (FX-CS-CN)

5

were developed using the FX-enriched fraction from Phaeodactylum tricornutum. Two

6

nanoscale particles (237 ± 13 nm for FX-CN and 277 ± 26 nm for FX-CN-CN) with spherical

7

and smooth surface showed over 71% encapsulation efficiency and polydispersity index (PDI)

8

value of 0.31~0.39 in water. Owing to the chitosan coating, FX-CS-CN showed a positive

9

zeta potential (24.00 mV), whereas that of FX-CN was negative (−12.87 mV). In vitro-

10

simulated digestion demonstrated better FX bioaccessibility from the nanoparticles versus P.

11

tricornutum powder (Pt-powder), and from FX-CN versus FX-CS-CN. However, in C57BL/6

12

mice, fucoxanthinol absorption to the blood circulation was two times higher for FX-CS-CN

13

versus FX-CN, possibly due to increased retention or adsorption to mucin by the cationic

14

biopolymer in the chitosan-coated particles. These results demonstrate that FX-CS-CN can

15

enable the application of FX, with improved bioavailability and water dispersibility, in the

16

food industry.

17 18

KEYWORDS: bioavailability; casein; chitosan; fucoxanthin; nanoparticles; Phaeodactylum

19

tricornutum

20 21

2 ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33

22

Journal of Agricultural and Food Chemistry

 INTRODUCTION

23 24

Fucoxanthin (FX), a marine xanthophyll carotenoid, is abundantly found in macroalgae such

25

as Laminaria japonica, Undaria pinnatifida, and Ecklonia cava.1 Until now, multiple health-

26

promoting properties, such as antioxidant, anticancer, and anti-inflammatory activities, have

27

been reported for FX.1 Among these properties, the antiobesity activity is the most promising

28

effect, and this has been supported by multiple in vitro and in vivo studies.1 Recently, some

29

microalgae, including Phaeodactylum tricornutum, Odontella aurita, and Isochrysis galbana,

30

were suggested as new sources for FX production.2-4 However, the global production and

31

distribution of FX from macro- or microalgae is very limited compared to those of lutein and

32

astaxanthin, which are two popular carotenoids from microalgae, owing to the difficulty of

33

achieving cost-effectiveness for the biomass production and extraction of FX. Most FX-

34

related products advertising an antiobesity effect, which are currently available on the market,

35

contain simple ethanolic extracts from macroalgae as the FX source. However, simple algal

36

extracts are not suitable for application in the food industry owing to their unique odor, sticky

37

consistency, and low concentration of FX. Thus, improved purification and processing

38

techniques are required to facilitate the wider utilization of FX in the food industry and other

39

fields.

40

Like other carotenoids, FX is sensitive to light, oxygen, and pH.5 To seek a solution for this

41

problem of low stability, several emulsion and nanoparticle techniques have been evaluated

42

and FX has been demonstrated to have increased stability and bioavailability inside these new

43

materials. Liquid systems comprising FX emulsions from micro to nano scales have been

44

produced based on triglycerols and canola oil, and these have been shown to reduce the

45

degradation of FX by heat, oxygen, and light.5,6 The bioaccessibility of FX under in vitro

46

digestion and animal feeding studies was also found to be improved to different extents 3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

47

according to the type of carrier oil used in the following order : long chain triacylglycerols ˃

48

medium chain triacylglycerols ˃ indigestible orange/mineral oils.7 A solid system composed

49

of chitosan, sodium tripolyphosphate, and glycolipid was used for FX encapsulation by an

50

ionic gelation method, whereby FX demonstrated increased stability and bioavailability.8

51

Sodium caseinate, produced by neutralization and spray-drying of the precipitated caseins

52

from skimmed milk, is commonly used as an emulsifying, stabilizing, and delivering agent in

53

the food and pharmaceutical industries.9 It has excellent surface activity and shows unique

54

self-assembly, as well as abundance and low cost.10 Four major components of caseins (αs1-,

55

αs2-, β-, and κ-casein) show amphiphilic properties with high proportions of hydrophobic and

56

hydrophilic amino acid residues.11 Therefore, casein can self-assemble easily to generate

57

stable spherical micelles with an average diameter of 150 nm and form complexes with

58

bioactive compounds to improve their stability and bioavailability. Another beneficial aspect

59

of sodium caseinate is the ease of preparing it as a powdered form by drying techniques such

60

as spray-dying and freeze-drying for the removal of solvents and convenience of

61

transportation, storage, and application in the food and pharmaceutical industries.12 Bioactive

62

compounds with a low aqueous solubility, stability, and bioavailability, such as flutamide (an

63

anti-androgenic agent), bixin (the major colorant component of annatto), curcumin (a natural

64

polyphenol from the rhizome of turmeric), and β-carotene (a carotenoid precursor of vitamin

65

A) have been used to make mixed micelles with sodium caseinate, and their properties were

66

characterized with respect to the stability and bioavailability of bioactive compounds.10-13

67

Meanwhile, chitosan is a cationic polysaccharide under neutral or basic pH conditions and

68

can be obtained by deacetylation of chitin which is the most widely distributed biopolymer in

69

nature and produced industrially from the crustacean wastes. By controlling the degree of

70

deacetylation and molecular weight of chitosan, this biopolymer can be useful as an

71

encapsulation material for micro/nanoparticles. The main advantages of using chitosan in 4 ACS Paragon Plus Environment

Page 4 of 33

Page 5 of 33

Journal of Agricultural and Food Chemistry

72

micro/nanoparticles include the controlled release of active agents, cross-linking ability,

73

cationic nature for a mucoadhesive character, and solubility in aqueous acidic conditions.14 A

74

number of methods have been developed to prepare micro/nanoparticles of chitosan. These

75

include emulsion cross-linking, coacervation/precipitation, spray-drying, emulsion-droplet

76

coalescence, ionic gelation, reverse micellar extraction, and sieving methods.6, 14

77

The purpose of this study was to develop and characterize FX-loaded nanoparticles with

78

casein (FX-CN) and FX-CN coated with chitosan (FX-CS-CN) from the microalga P.

79

tricornutum, to facilitate the wider utilization of FX in the food industry. The characteristics

80

of the nanoparticles were investigated, including their size, morphology, structure, zeta

81

potential, polydispersity index (PDI), encapsulation efficiency, and adsorption to mucin. In

82

addition, the release of FX from the matrix in the gastrointestinal (GI) tract and the

83

absorption of fucoxanthinol (FXOH), a primary metabolite of FX by lipase (Figure 1A), into

84

the mouse blood circulation system were tested, with P. tricornutum dry powder as a

85

comparator, by an in vitro simulated digestion assay and an in vivo pharmacokinetic study,

86

respectively, to evaluate the bioaccessibility and bioavailability of FX from FX-CN and FX-

87

CS-CN.

88 89

 Materials and Methods

90 91

Materials. The powder of P. tricornutum (Pt-powder) and an FX-enriched fraction (FX-fr)

92

containing over 30% (w/w) of FX (as shown in Figure 1B) were provided from BioOne Co.

93

Ltd. (Gangneung-si, Gangwon-do, Korea). FX-fr was produced by partial purification of FX

94

from the ethanolic extract of Pt-powder through the silica gel chromatography method

95

modified from our previous report. 2 Casein sodium salt from bovine milk, chitosan (medium

96

molecular weight), Tween® 80, xanthan gum, glycerol, and calcium chloride were purchased 5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

97

from Sigma-Aldrich (St. Louis, MO, USA). All solvents were of HPLC grade and were

98

purchased from Sigma-Aldrich and Fisher Scientific Korea Ltd (Seoul, Korea).

99 100

Preparation of FX-CN and FX-CS-CN. Two types of FX-loaded nanoparticles were

101

prepared based on the method described by Pan et al.12 Three grams of casein sodium salt

102

was dissolved in 150 mL of 40% (v/v) aqueous ethanol by moderate stirring with slight

103

heating for 120 min. After the sodium casein became totally hydrated, an ethanolic solution

104

(2 mL) of 12 mg mL-1 FX-fr was added to 150 mL of 2% sodium caseinate solution while

105

stirring for 60 min. Then, 35 mL of 0.1 M K2HPO4 and 3 mL of 0.4 M tripotassium citrate

106

were added, followed by a dropwise addition of 110 mL 0.03 M CaCl2 and blending at 15000

107

rpm for 5 min using a high speed homogenizer (ULTRA-TURRAX® T18, IKA, Stauffen,

108

Germany). After homogenization, the mixture was injected into the electrospray system,

109

using a syringe pump. Then, the mixture solution was sprayed through the nozzle while

110

applying a fixed voltage of 5 kV and a flow rate of 130 µL/min to obtain FX-CN. FX-CS-CN

111

was prepared by a chitosan coating process with the FX-CN solution. Chitosan solution

112

(0.1%, w/v) was prepared by dissolving chitosan in distilled water containing 0.1% (w/v)

113

acetic acid. To remove the impurities, chitosan solution was filtered using a 0.45-µm syringe

114

filter (Minisart®, Sartorius AG, Göttingen, Germany). The prepared chitosan solution was

115

added dropwise to FX-CN solution and stirred mildly for 1 h to obtain FX-CS-CN. Two

116

freshly prepared nanoparticle dispersions were used for the measurement of various

117

characteristics and then the nanoparticles were freeze-dried at −120 ℃ for 48 h using a

118

freeze-drier (FDCF-12003, OPERON, Kimpo, Korea), and the dried samples were stored in

119

glass vials at −20 ℃ until use.

120

6 ACS Paragon Plus Environment

Page 6 of 33

Page 7 of 33

Journal of Agricultural and Food Chemistry

121

Characterization of FX-CN and FX-CS-CN. The mean particle size, PDI, and zeta

122

potential of FX-CS-CN and FX-CN were measured using dynamic light scattering with a

123

NanoZS (Malvern Instruments, Worcestershire, UK) at 25 ℃, with a detector angle of 90°A

124

minimum of three parallel measurements were performed and the average value was

125

calculated.

126

Structural characterization of the two types of nanoparticles was performed by Fourier

127

transform infrared (FT-IR), using a model V430 apparatus (Jasco, Tokyo, Japan) at 20 ℃. The

128

FT-IR spectra of FX-fr, chitosan, casein, FX-CN, and FX-CS-CN were obtained to estimate

129

the interactions between chitosan and casein. All samples were mixed with potassium

130

bromide (KBr) at a ratio of 1:10 and formed into pellets by compression under a force of 5

131

tons in a hydraulic press. These pellets were scanned in transmission mode. Each spectrum

132

was obtained from the average of 64 scans at a resolution of 4 cm-1 in the wavelength range

133

of 600−4000 cm-1.

134

The morphologies of FX-CN and FS-CS-CN were examined by field emission scanning

135

electron microscopy (FE-SEM). To prepare SEM samples, the samples were mounted on FE-

136

SEM stubs with double-sided adhesive tape and coated under vacuum with platinum to make

137

the samples conductive. Samples were observed using an FE-SEM instrument (SU70, Hitachi,

138

Tokyo, Japan) at an accelerating voltage of 15 kV. The photographs were taken at selected

139

magnifications and representative images were reported.

140

To evaluate FX stability in the nanoparticles, two nanoparticles were dissolved in water

141

and adjusted to the final FX concentration of 8 µg/mL. After then, these aqueous solutions

142

were stored at three different temperatures (2 ℃, 10 ℃, 26 ℃) for 4 weeks under darkness. FX

143

contents were analyzed every week by HPLC.

144

Accurately weighed amounts (0.1 g) of FX-CS-CN and FX-CN were dispersed and

145

dissolved in ethanol and sonicated for 1 h. The encapsulation efficiency of the nanoparticles 7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

146

was determined using the equation below:

147

Encapsulation efficiency (%) = Fa/Fth × 100

148

where Fa is the actual amount of FX encapsulated in FX-CS-CN and FX-CN and Fth is the

149

theoretical amount, assuming that all of the FX added in this experiment was encapsulated

150

within the nanoparticles without any losses during the preparation procedure.

151

A mucin adsorption test was performed using a slight modification of a previously

152

reported method by Filipović-Grčić et al. 15 and the amount of free mucin was calculated by

153

the Bradford method.16 Mucin solution (1 mL of 1 mg/mL concentration) was blended with 1

154

mL of FX-CN or FX-CS-CN at 37 ℃ for 1 h. Then, the suspension was centrifuged at 15000

155

rpm at 4 ℃ for 30 min to isolate the free mucin (supernatant) and adsorbed mucin (pellet). In

156

the Bradford method, Bradford reagent was diluted 5-fold and mixed with the supernatant.

157

Then, the mixture was incubated in a 37 ℃ water bath with shaking at 180 rpm for 10 min.

158

The quantity of mucin in the supernatant was calculated by measuring the absorbance at 595

159

nm. To make the standard curve, mucin solutions of 31.25, 62.5, 125, 250, 500, 1000, and

160

2000 µg/mL were prepared and the absorbance of each solution was measured. Mucin

161

adsorption (%) was calculated using the equation below:

162 163

Mucin adsorption (%) = [(Ct-Cf)/Ct] × 100 where Ct is the total amount of mucin and Cf is the free mucin in the supernatant.

164 165

In vitro simulated digestion. Simulated digestion was performed according to the method

166

described by Garrett et al.,17 with slight modifications. To test in vitro simulated digestion,

167

digestive steps were performed sequentially. Briefly, FX-CS-CN or FX-CN powder (25 mg)

168

was homogenized in 10 mL saline solution containing 120 mM NaCl, 5 mM KCl, and 6 mM

169

CaCl2 (pH 5.5). Then, 1000 units of α-amylase was added, and the pH was adjusted to 6.5.

170

After the saline solution had been added to a volume of 12.5 mL, the samples were incubated 8 ACS Paragon Plus Environment

Page 8 of 33

Page 9 of 33

Journal of Agricultural and Food Chemistry

171

at 37 ℃ for 5 min in a shaking water bath (Lab Companion, Jeio Tech, Seoul, Korea) at 95

172

rpm to simulate the oral phase of digestion. To mimic the gastric phase of human digestion,

173

the pH of the sample was acidified to 2.2 with HCl and 0.5 mL of a porcine pepsin solution

174

(0.075 g/mL in 0.1 N HCl) was added. The samples were suspended in a saline solution to a

175

volume of 15 mL and incubated at 37 ℃ for 2 h. Three intestinal parts (duodenum, jejunum,

176

and ileum) were separated and simulated sequentially. To simulate the duodenum stage, 250

177

mg of bile extract, 0.5 mL of pancreatic lipase (0.01 g/mL), and 0.5 mL of pancreatin (0.08

178

g/mL) were added and the pH was increased to 5.5 by adding 1 M sodium bicarbonate. The

179

samples were incubated for 30 min at 37 ℃ (final volume of 20 mL). To mimic the jejunum

180

stage, the pH was adjusted to 6.0 and the samples were incubated for 90 min at 37 ℃ (final

181

volume of 22.5 mL). Finally, the pH was adjusted to 7.0 and the samples were incubated for 5

182

h at 37 ℃ (final volume of 25 mL) to simulate the ileum stage of digestion.

183 184

Pharmacokinetic study. In vivo animal study was adhered to the Guide for the Care and Use

185

of Laboratory Animals developed by the Institute of Laboratory Animal Resources of the

186

National Research Council. The protocol for the pharmacokinetic study was approved by the

187

Institutional Animal Care and Use Committee of KIST (2015-018) in Seoul, Korea. For a

188

single dose administration corresponding to 7 mg FX/kg mouse body weight, the mice (male

189

C57BL/6, 7 weeks old, 25~50 g) were administered with the powder type samples (Pt-

190

powder, FX-CN, and FX-CS-CN) dissolved in distilled water by intragastric gavage. The

191

animals were sacrificed at 0.5, 1, 2, 4, 8, and 24 h after administration. The control mice at

192

the 0 h time point were not orally administered anything, but just anaesthetized. The blood

193

was collected by cardiac puncture in a heparinized syringe and was centrifuged at 1,000 × g

194

for 10 min at 4 ℃. The supernatant fraction was used as the plasma samples. The plasma 9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 33

195

concentrations of FXOH were determined by extracting FXOH from plasma samples with

196

ethyl acetate solvent and performing liquid chromatography/tandem mass spectrometry (LC-

197

MS/MS) analysis. As an internal standard, carbamazepine was used. The pharmacokinetic

198

study of each treatment group was measured with three mice (n = 3). Non-compartmental

199

pharmacokinetic parameters of FXOH were calculated from the time course of plasma

200

concentration by WinNonlin® program (Pharsight, Mountain View, CA, USA).

201 202

High pressure liquid chromatography and LC-MS/MS conditions. For the in vitro

203

simulated digestion assay, the target compounds (FX and FXOH) were quantified using a

204

high pressure liquid chromatography system coupled to a diode array detector (1200 series,

205

Agilent, Santa Clara, CA, USA). A YMC C-30 carotenoid column (150 × 4.6 mm internal

206

diameter, 3 µm particle size, Waters, Milford, MA, USA) was used for the separation. A

207

methanol and water solvent system was used as the mobile phase at a flow rate of 0.7 mL/min

208

with a column temperature of 35 ℃. The solvent gradient program was as follows: the

209

methanol/water ratio was increased from 90:10 to 100:0 over 20 min, and then held at 100%

210

methanol for the next 5 min. The chromatogram obtained at 450 nm was used for the

211

quantitative analysis of FX and FXOH. For the pharmacokinetic study, FXOH in mouse

212

plasma samples was identified and quantified using an LC-MS/MS system (1200 series

213

instrument coupled with a 6410 series instrument, Agilent). Multiple reaction monitoring

214

modes were used for the quantification: m/z 617.1 > 109.2 for FXOH and m/z 237 > 194 for

215

carbamazepine (used as a standard reagent).

216 217

Statistical analysis. All experiments were repeated in triplicate. All data are presented as

218

mean ± standard deviation. The significant differences among means were analyzed by t-test,

219

one-way ANOVA (P < 0.05) and Duncan’s multiple range test (DRT). 10 ACS Paragon Plus Environment

The t-test and one-

Page 11 of 33

Journal of Agricultural and Food Chemistry

220

way ANOVA analysis were performed by Microsoft Office Professional Plus 2010 Excel and

221

DRT was conducted by the free statistical software R program (version 3.3.2).

222 223

 RESULTS AND DISCUSSION

224 225

Characterization of FX-CN and FX-CS-CN. The particle size, zeta potential, and PDI of

226

FX-CN and FX-CS-CN are listed in Table 1. The mean particle diameters of FX-CN and FX-

227

CS-CN were both in the nanoscale range, at 237 ± 13 nm and 277 ± 26 nm, respectively. The

228

size of FX-CS-CN was significantly larger than that of FX-CN owing to the chitosan coating

229

process. The thickness of the chitosan layer could be calculated from the mean particle

230

diameter data showing around 40 nm. Similar observations have been reported in many

231

papers, showing that applying a chitosan coating onto nanoparticles increases their diameter

232

compared with uncoated particles.18,19 The PDI values of FX-CN and FX-CS-CN were 0.31 ±

233

0.03 and 0.39 ± 0.02, respectively. The PDI value represents a heterogeneity index. A higher

234

PDI indicates that the dispersion state is more heterogeneous, whereas a lower PDI indicates

235

that the dispersion state is more homogeneous. Since FX-CN showed a lower PDI value than

236

FX-CS-CN did, it could be suggested that FX-CN represents a more homogeneous and stable

237

state than FX-CS-CN. Zeta potential is another significant characteristic of nanoparticles and

238

indicates the stability of a colloidal suspension. A zeta potential of particles close to an

239

absolute value of 30 mV (± 30 mV) has been reported to be a stable state.20,21 In our results,

240

the zeta potential of FX-CN was negatively charged (−12.9 ± 1.6 mV) and that of FX-CS-CN

241

was positively charged (+24.0± 2.8 mV). Therefore, in terms of colloidal stability, FX-CS-

242

CN is more stable than FX-CN. As expected, FX-CN possessed a negative charge because of

243

the predominance of ionized carboxylic groups of sodium casein at pH 7.4.21 However,

244

chitosan coating induced a change in the surface charge of FX-CS-CN because of the 11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

245

positively charged amine groups of chitosan.22-24

246

Different variants of the nanoparticles prepared with varying concentrations of casein in

247

the range of 0.5~3% (w/v) were examined to determine the optimal concentration of the

248

casein solution to maximize the FX encapsulation efficiency (data not shown). Among them,

249

the 2% casein solution demonstrated the best performance for FX encapsulation, and the

250

encapsulation efficiencies of FX-CN and FX-CS-CN prepared with the 2% casein solution

251

were 73.7% and 71.8%, respectively. Although mechanical processing steps such as

252

ultrasonic homogenization and electrospraying can negatively affect the encapsulation

253

efficiency, these results showed that the use of casein and chitosan yielded a good

254

encapsulation efficiency and a high stability of the nanoparticles.

255

After freeze-drying, fine powders of FX-CN and FX-CS-CN with a yellow color were

256

obtained, as shown in Figure 1B. The surface morphologies of these two nanoparticles were

257

visualized by FE-SEM and presented in Figure 2A and 2B, respectively. The morphologies of

258

FX-CN and FX-CS-CN were observed as almost spherical and uniform. The surface of FX-

259

CS-CN was smoother than that of FX-CN, which is due to the effect of coating. One of the

260

advantages of preparing nanoparticles with casein and chitosan is that the water solubility of

261

encapsulated fat-soluble components is improved by these hydrophilic coating materials.24

262

Figure 2C presents a comparison of the visual appearances of the two FX-loaded

263

nanoparticles and FX-fr added to water. FX-CN and FX-CS-CN showed no sedimentation,

264

indicating that FX encapsulated by casein and chitosan was well dispersed in water. However,

265

FX-fr was not dispersed and several aggregates were observed on the wall and bottom of the

266

tube, indicating the insolubility of FX-fr in water. Figure 2D and 2E showed FX stability in

267

the FX-CN and FX-CS-CN during 4 weeks of storage in water with different temperatures. In

268

our previous data,25 stability data of free FX in water was significantly decreased according to

269

temperature increase. Free FX in water was degraded to almost 30% of initial amount at 26 ℃ 12 ACS Paragon Plus Environment

Page 12 of 33

Page 13 of 33

Journal of Agricultural and Food Chemistry

270

after 4 weeks indicating unstable characteristic of free FX under aqueous condition. However,

271

more than 80% of initial FX in two nanoparticles was maintained under same aqueous

272

condition. Thus, we could confirm the increased stability of FX in the nanoparticles under

273

aqueous condition.

274

FT-IR analysis was performed to evaluate the interaction between chitosan and the FX-

275

loaded casein nanoparticles. Figure 3 shows FT-IR spectra in the wavelength range of

276

600−4000 cm-1 for FX-fr, casein, FX-CN, chitosan, and FX-CS-CN. Figure 3A shows the

277

spectrum of FX-fr. The peak at 1929.99 cm-1 was assigned as an allenic bond (C=C=C),

278

which is considered a functional group of FX.26,27 The peak at 3367.68 cm-1 was assigned as

279

hydrogen bonded O–H stretching vibrations. Peaks at 2924.06 cm-1 and 2854.62 cm-1 showed

280

the presence of alkanes with C–H bonds. Absorption at 1732.06 cm-1 means the presence of

281

ketones with –C=O bonds. Absorption peaks at 1456.24, 1433.80, 1406.09, and 1367.52 cm-1

282

were assigned as scissoring and bending alkanes with –C–H bonds. The peak at 1031.91 cm-1

283

was characteristic of esters with –C–O bonds. All of these peaks are generally found in FX.27

284

Figure 3B shows the spectrum of the casein sodium salt. The peaks at 1637.55 cm-1 and

285

1539.18 cm-1 showed the carbonyl group (C=O). Figure 3C shows the spectrum of FX-CN.

286

The major absorption peaks shown in Figures 3A and 3B were also found in this spectrum,

287

indicating that the FX in FX-fr was well encapsulated within casein in FX-CN. Figure 3D

288

shows the spectrum of chitosan. For native chitosan, two peaks at 1649.05 cm-1 and 1556.18

289

cm-1 were assigned to the carbonyl bonds (C=O) of the amide group and protonated amine

290

group (NH3+).28,29 Figure 3E shows the spectrum of FX-CS-CN, and the combination

291

absorption peaks of Figures 3C and 3D could be found in Figure 3E. Taking all of the results

292

obtained by FT-IR analysis together, it was concluded that FX from FX-fr has been

293

successfully encapsulated in casein and chitosan.

294

13 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

295

Stability and bioaccessibility of FX under in vitro digestion assay. Two FX-loaded

296

nanoparticles (FX-CN and FX-CS-CN) and Pt-powder were used for an in vitro digestion

297

assay to evaluate their digestive stability and bioaccessibility (release from the food matrix)

298

during the simulated digestion process of the human GI tract from the mouth to the small

299

intestine. In this assay system, each sample was serially treated with the proper enzymes,

300

temperature, time, and pH to simulate the five digestion stages, comprising those occurring in

301

the mouth, stomach, and three stages of the small intestine (duodenum, jejunum, and ileum)

302

in a single test tube. The collected whole digestion solutions from each stage were analyzed

303

to estimate the digestive stability of total FX (FX and FXOH), and the micelle phase obtained

304

by centrifugation of the solution was analyzed to evaluate the bioaccessibility of total FX

305

(Figure 4).

306

As shown in Figure 1A, a number of studies have reported that FX is metabolized to

307

FXOH by lipase and cholesterol esterase in the small intestine.30 This phenomenon was also

308

found with the two types of FX-loaded nanoparticles and Pt-powder during the in vitro

309

digestion assay (Figure 4). Since FXOH has an antiobesity activity and FX is primarily

310

metabolized to FXOH,30 total FX (FX + FXOH) was measured to determine the stability of

311

FX in the in vitro digestion assay. FXOH started to appear after adding pancreatic lipase (and

312

cholesterol esterase) at the beginning stage of the small intestine (duodenum) and the level of

313

transformation from FX to FXOH was slightly increased during the small-intestine stages

314

(duodenum to ileum) according to the increasing treatment time and pH. However, no FXOH

315

was produced before the small-intestine stages of digestion, indicating that this

316

transformation was exclusively mediated by an enzymatic reaction.

317

As a carotenoid, FX can be easily degraded by low pH and oxygen.1 During the digestion

318

process in the GI tract, FX is exposed to these factors by gastric acid and oxidative chemicals.

319

Thus, the stability of FX during the digestion process is an important concern when 14 ACS Paragon Plus Environment

Page 14 of 33

Page 15 of 33

Journal of Agricultural and Food Chemistry

320

developing new food materials. The analysis of the digestive stability of FX in the two types

321

of nanoparticles and Pt-powder demonstrated that total FX was slightly degraded during the

322

simulated digestion process (Figures 4A–4C). However, if the amount of FX in food is

323

denoted as 100%, the total FX levels were over 80% at the end stage of the small intestine

324

(ileum). This result indicates that most of the FX in the two types of nanoparticles can be

325

stable in either the FX or FXOH form, and had similar stability to that of Pt-powder during

326

digestion in the GI tract.

327

During the digestion process in vivo, mixed micelles and vesicles can solubilize lipids and

328

lipid digestion products from the surface of lipid droplets and are transported to the intestinal

329

epithelium cells coated with the mucous layer for absorption.31 In the in vitro digestion assay,

330

the micelle phase contains these mixed micelles and vesicles, and thus, the FX

331

bioaccessibility of the two types of nanoparticles and Pt-powder can be deduced by analyzing

332

the amount of total FX in this phase. The results presented in Figures 4D–4F showed that the

333

two types of nanoparticles yielded mixed micelles and vesicles containing FX in the micelle

334

phases isolated before the small intestine stages of digestion, while Pt-powder yielded almost

335

no mixed micelles and vesicles containing FX. In the small intestine stages of digestion, the

336

micelle phases of FX-CN and FX-CS-CN contained total FX levels in the ranges of 64–73%

337

and 36–61% of the initial level, respectively. Those levels were higher than that of Pt-powder

338

(26–48%), indicating that the two types of nanoparticles both yield a better bioaccessibility of

339

FX than Pt-powder does and FX can be released more easily from their matrices. In addition,

340

a higher content of FXOH was detected in the micelle phases of the two types of

341

nanoparticles than in Pt-powder. When comparing between FX-CN and FX-CS-CN, both the

342

total FX and FXOH contents in the micelle phase of FX-CN were higher than those in the

343

same phase of FX-CS-CN.

344

A higher bioaccessibility (i.e., a higher release from the food matrix during digestion) 15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 33

345

could be expected to result in a higher bioavailability (a higher absorption and entry into the

346

systemic blood circulation). Besides bioaccessibility, however, two other factors that can

347

influence bioavailability should also be considered. One of these factors is the transport of the

348

released lipophilic components to the intestinal epithelium cells through the mucous layer,

349

and the other factor is the metabolism of the lipophilic components during their absorption

350

and entry into the systemic circulation.31 Consequently, it is too early to conclude that a

351

higher amount of FX or FXOH in the micelle phase of FX-CN indicates more absorption of

352

these molecules into the blood circulation. Nevertheless, it is certain that a higher

353

bioaccessibility is beneficial for achieving a higher bioavailability.

354 355

Bioavailability of FX under in vivo pharmacokinetic and mucin adsorption studies. To

356

evaluate the in vivo absorption rate of FX or FXOH, an in vivo pharmacokinetic study was

357

conducted by feeding a single dose of each sample powder (FX-CN, FX-CS-CN, and Pt-

358

powder) corresponding to 7 mg FX/kg body weight by oral administration to C57BL/6 mice.

359

Among the three metabolites (FX, FXOH, and amarouciaxanthin A) derived from FX in

360

mouse plasma samples, only FXOH was analyzed by LC-MS/MS because it is the primary

361

metabolite

362

amarouciaxanthin A, the metabolite of FXOH in mouse liver, has not been reported in human

363

plasma, and thus, it remains uncertain whether this compound is a metabolite of FX in human.

364

After administration, plasma samples were collected at 1, 2, 4, 8, and 24 h, and

365

pharmacokinetic parameters such as Tmax, Cmax, T1/2, AUCt, and AUC∞ were calculated.

of

FX

detected

most

abundantly

in

mouse

plasma30.

Furthermore,

366

Figure 5A shows the FXOH concentration in plasma and Table 2 shows the

367

pharmacokinetic data of FXOH for the two types of nanoparticles (FX-CN and FX-CS-CN)

368

and Pt-powder. The time (Tmax) at which the maximum plasma concentration (Cmax) was 16 ACS Paragon Plus Environment

Page 17 of 33

Journal of Agricultural and Food Chemistry

369

detected was 2.7 h for Pt-powder and over 4.7 h for the two types of nanoparticles. This result

370

indicates that the two types of nanoparticles reached the blood circulation later than Pt-

371

powder through the intestinal epithelium cells even though bioaccessibility (release from the

372

food matrix) of each type of nanoparticle was higher than that of Pt-powder. The

373

bioavailability of FXOH can be deduced from the area under the plasma concentration-time

374

curve to infinity (AUC∞). As shown in Table 2, the AUC∞ of FX-CS-CN was 531 nmol·h/L,

375

which is almost two times higher than that of FX-CN (276 nmol·h/L) and Pt-powder (280

376

nmol·h/L). Therefore, it can be concluded that FX-CS-CN has a higher bioavailability than

377

the other two samples do and FX-CN has a similar bioavailability to that of Pt-powder.

378

The pharmacokinetic bioavailability data shown in Figure 5A and Table 2 showed quite

379

different results from those of the bioaccessibility data from the previous in vitro digestion

380

assay. There might be several factors influencing the differences between the results of these

381

experiments, and one of them could be the chemical characteristics of two nanoparticles.

382

Among the tested characteristics in Table 1 and Figures 2 and 3, the most significant

383

difference between these two nanoparticles was their opposite zeta potentials due to the

384

chitosan coating process. This factor may affect the absorption ratios of these two types of

385

nanoparticles through the intestinal epithelium cells. Since the epithelium cells in the small

386

intestine are coated with a mucous layer that is the first barrier for nutrients, and this layer

387

possesses a negative charge,31 FX-CS-CN with a positive charge is likely to be adsorbed for

388

longer on this mucous layer by electrostatic force than FX-CN with a negative charge. In

389

order to investigate this property, a mucin adsorption study was carried out with the two types

390

of nanoparticles, and the results showed that FS-CS-CN had a higher mucin adsorption level

391

(61.2%) than FX-CN (34.7%), as expected (Figure 5B). Interactions between FX-CS-CN and

392

mucin may be caused by electrostatic interactions between the amine group (NH3+) of 17 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

393

chitosan and the carboxyl group (COO−) of mucin. Consequently, FX-CS-CN demonstrated a

394

higher bioavailability than FX-CN did, even though the bioaccessibility of FX-CS-CN was

395

lower than that of FX-CN, and the positive zeta potential induced by chitosan coating was

396

suggested as one of the main factors underlying that difference.

397

In conclusion, two types of FX-loaded nanoparticles (FX-CN and FX-CS-CN) were

398

developed using casein and chitosan as encapsulation materials with an FX-enriched fraction

399

extracted from the microalga P. tricornutum. These two new materials showed typical

400

characteristics of such nanoparticles, including size, zeta potential, PDI, morphology, and the

401

improved water solubility. In addition, FX-CS-CN yielded a better bioavailability of FXOH

402

than FX-CN and Pt-powder. Chitosan may also affect the tight junction integrity and cell

403

permeability. These results indicate that FS-CS-CN can be a good material for the application

404

of FX with improved water solubility and bioavailability in the food industry.

405 406

Financial support

407

This research was supported by a grant from Marine Biotechnology Program (2MP0360)

408

funded by Ministry of Oceans and Fisheries, Korea and an intramural grant (2Z04690) from

409

KIST Gangneung Institute of Natural Products

410 411

Note

412

The authors declare no competing financial interest.

413 414

 References

415 416

(1) D′Orazio, N.; Gemello, E.; Gammone, M. A.; Girolamo, M.; Ficoneri, C.; Riccioni, G.

417

Fucoxanthin: A treasure from the sea. Mar. Drugs. 2012, 10, 604-616. 18 ACS Paragon Plus Environment

Page 18 of 33

Page 19 of 33

Journal of Agricultural and Food Chemistry

418 419

(2) Kim, S. M.; Jung, Y. J.; Kwon, O. N.; Cha, K. H.; Um, B. H.; Chung, D. H.; Pan, C. H.

420

A potential Commercial Source of Fucoxanthin Extracted from the Microalga

421

Phaeodactylum tricornutum. Appl. Biochem. Biotechnol. 2012, 166, 1843-1855.

422 423

(3) Kim, S. M.; Kang, S. W.; Kwon O. N.; Chung, D. W.; Pan, C. H. Fucoxanthin as a major

424

carotenoid in Isochrysis aff. galbana: Characterization of extraction for commercial

425

application. J. Korean Soc. Appl. Biol. Chem. 2012, 55, 477-483.

426 427

(4) Xia, S.; Wang, K.; Wan, L.; Li, A.; Hu, Q.; Zhang, C. Production, characterization, and

428

antioxidant activity of fucoxanthin from the marine diatom Odontella aurita. Mar. Drugs.

429

2013, 11, 2667-2681.

430 431

(5) Zhao, D.; Kim, S. M.; Pan, C. H.; Chung, D. Effects of heating, aerial exposure and

432

illumination on stability of fucoxanthin in canola oil. Food Chem. 2014, 145, 505-513.

433 434

(6) Dai, J.; Kim, S. M.; Shin, I. S.; Kim, J. D.; Lee, H. Y.; Shin, W. C.; Kim, J. C. Preparation

435

and stability of fucoxanthin-loaded microemulsions. J. Ind. Eng. Chem. 2014, 20, 2103-2110.

436 437

(7) Salvia-Trujillo, L.; Sun, Q.; Um, B. H.; Park, Y.; McClements, D. J. In vitro and in vivo

438

study of fucoxanthin bioavailability from nanoemulsion-based delivery systems: Impact of

439

lipid carrier type. J. Funct. Foods. 2015, 17, 293-304.

440 441

(8) Ravi, H.; Baskaran, V. Biodegradable chitosan-glycolipid hybrid nanogels: A novel

442

approach to encapsulate fucoxanthin for improved stability and bioavailability. Food 19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

443

Hydrocolloids 2015, 43, 717-725.

444 445

(9) Luo, Y.; Teng, Z; Wang, T. T. Y.; Wang, Q. Cellular uptake and transport of zein

446

nanoparticles: effects of sodium caseinate. J. Agric. Food Chem. 2013, 61, 7621-7629.

447 448

(10) Zhang, Y.; Zhong, Q. Encapsulation of bixin in sodium caseinate to deliver the colorant

449

in transparent dispersions. Food Hydrocolloids 2013, 33, 1-9.

450 451

(11) Elzoghby, A. O.; Helmy, M. W.; Samy, W. M.; Elgindy, N. A. Novel ionically

452

crosslinked casein nanoparticles for flutamide delivery: formulation, characterization, and in

453

vivo pharmacokinetics. Int. J. Nanomedicine. 2013, 8, 1721-1732.

454 455

(12) Pan, K.; Zhong, Q.; Baek, S. J. Enhanced dispersibility and bioactivity of curcumin by

456

encapsulation in casein nanocapsules. J. Agric. Food Chem. 2013, 61, 6036-6043.

457 458

(13) Yi, J.; Li, Y.; Zhong, F.; Yokoyama, W. The physicochemical stability and in vitro

459

bioaccessibility of beta-carotene in oil-in-water sodium caseinate emulsions. Food

460

Hydrocolloids 2014, 35, 19-27.

461 462

(14) Agnihorti, S. A.; Mallikarjuna, N. N.; Aminabhavi, T. M. Recent advances on chitosan-

463

based micro- and nanoparticles in drug delivery. J. Control Release. 2004, 100, 5-28.

464 465

(15) Filipović-Grčić, J.; Škalko-Basnet, N.; Jalšenjak, I. Mucoadhesive chitosan-coated

466

liposomes: characteristics and stability. J. Microencapsul. 2001, 18, 3-12.

467

20 ACS Paragon Plus Environment

Page 20 of 33

Page 21 of 33

Journal of Agricultural and Food Chemistry

468

(16) Bradford, M. M. A rapid and sensitive method for the quantitation of microgram

469

quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 1976, 72,

470

248-254.

471 472

(17) Garrett, D. A.; Failla, M. L.; Sarama, R. J. Development of an in vitro digestion method

473

to assess carotenoid bioavailability from meals. J. Agri. Food Chem. 1999, 47, 4301-4309.

474 475

(18) Liu, Y.; Liu, D.; Zhu, L.; Gan, Q.; Le, W. Temperature-dependent structure stability and

476

in vitro release of chitosan-coated curcumin liposome. Food Res. Int. 2015, 74, 97-105.

477 478

(19) Shin, G. H.; Chung, S. K.; Kim, J. T.; Joung, H. J.; Park, H. J. Preparation of chitosan-

479

coated nanoliposomes for improving the mucoadhesive property of curcumin using the

480

ethanol injection method. J. Agri. Food Chem. 2013, 61, 11119-11126.

481 482

(20) Muller, R. H.; Jacobs, C.; Kayser, O. Nanosuspensions as particulate drug formulations

483

in therapy rationale for development and what we can expect for the future. Adv. Drug Deliv.

484

Rev. 2001, 47, 3-19.

485 486

(21) Elzoghby, A. O.; Helmy, M. W.; Samy, W. M.; Elgindy, N. A. Spray-dried casein-based

487

micelles as a vehicle for solubilixation and controlled delivery of flutamide: Formulation,

488

characterization, and in vivo pharmacokinetics. Eur. J. Pharm. Biopharm. 2013, 84, 487-496.

489 490

(22) Feng, C.; Li, J.; Kong, M.; Liu, Y.; Cheng, X. J.; Li, Y.; Park, H. J.; Chen, X. G. Surface

491

charge effect on mucoadhesion of chitosan based nanogels forlocal anti-colorectal cancer

492

drug delivery. Colloids Surf. B Biointerfaces. 2015, 128, 439-447. 21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

493 494

(23) Hasan, M.; Ben Nessaoud, G.; Michaux, F.; Tamayol, A.; Kahn, C. J. F.; Belhaj, N.;

495

Linder, M., Arab-Tehrany, E. Chitosan-coated liposomes encapsulating curcumin; study of

496

lipid-polysaccharide interactions and nanovesicle behavior. RSC Adv. 2016, 6, 45290-45304.

497 498

(24) Dhawan, S.; Singla, A. K.; Sinha, V. R. Evaluation of mucoadhesive properties of

499

chitosan microspheres prepared by different methods. AAPS Pharmscitech. 2004, 5, 122–128.

500

(25) Mok, I. L.; Yoon, J. R.; Pan, C. H.; Kim, S. M. Development, quantification, method

501

validation, and stability study of a novel fucoxanthin-fortified milk. J. Agri. Food Chem.

502

2016, 64, 6196-6202.

503 504

(26) Hasan, M.; Ben Nessaoud, G.; Michaux, F.; Tamayol, A.; Kahn, C. J. F.; Belhaj, N.;

505

Linder, M., Arab-Tehrany, E. Chitosan-coated liposomes encapsulating curcumin; study of

506

lipid-polysaccharide interactions and nanovesicle behavior. RSC Adv. 2016, 6, 45290-45304.

507 508

(27) Pangestuti, R.; Kim, S. K. Biological activities and health benefit effects of natural

509

pigments derived from marine algae. J. Funct. Foods. 2011, 3, 255-266.

510 511

(28) Cho, Y.; Kim, J. T.; Park, H. J. Size-controlled self-aggregated N-acyl chitosan

512

nanoparticles as a vitamin C carrier. Carbohydr. Polym. 2012, 88, 1087-1092.

513 514

(29) Mady, M. M.; Darwish, M. M.; Khalil, S.; Khalil, W. M. Biophysical studies on

515

chitosan-coated liposomes. Eur. Biophys J. 2009, 38, 1127-1133.

516 517

(30) Miyashita K.; Nishikawa S.; Beppu F.; Tsukui T.; Abe M.; Hosokawa M. The allenic 22 ACS Paragon Plus Environment

Page 22 of 33

Page 23 of 33

Journal of Agricultural and Food Chemistry

518

carotenoid fucoxanthin, a novel marine nutraceutical from brown seaweeds. J. Sci. Food

519

Agric. 2011, 91, 1166-1174.

520 521

(31) McClements D. J.; Li Y. Review of in vitro digestion models for rapid screening of

522

emulsion-based systems. Food Funct. 2010, 1, 32-59.

523 524 525

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

 Figure captions

Figure 1. Fucoxanthin (Fx) metabolism during digestion process (A) and the preparation process of FX-loaded nanoparticle from P. tricornutum powder (Pt-powder) based on casein and chitosan (B).

Figure 2. Field emission scanning electron microscope (FE-SEM) images of FX-CN (A) and FX-CS-CN (B) with a scale bar of 500 nm. At the concentration of 3 mg/mL, these nanoparticles dissolved in water showed good dispersion, whereas FX-fr was not fully dissolved in water (C), and fucoxanthin in FX-CN (D) and FX-CS-CN (E) was maintained over 80% of initial amount in the aqueous solution during 4 weeks storage period.

Figure 3. Fourier transform-infrared (FT-IR) spectrums of a fucoxanthin (FX)-enriched fraction (FX-fr) from P. tricornutum (A), casein (B), FX-casein nanoparticles (FX-CN) (C), chitosan (D), and chitosan-coated FX-CN (FX-CS-CN) (E). Each FT-IR spectrum was analyzed in the wavelength range of 4000–600 cm-1.

Figure 4. Digestive stability (A–C) and bioaccessibility (D–F) of fucoxanthin (FX) in FX-CN, FX-CS-CN, and Pt-powder during an in vitro simulated digestion assay. From the duodenum stage of digestion, FX started to be transformed to FXOH by lipase and cholesterol esterase. The difference numbers and letters represent significant differences of means.

Figure 5. (A) Pharmacokinetic data of two types of fucoxanthin (FX)-loaded nanoparticles and P. tricornutum powder (Pt-powder). A single dose of each sample powder was orally 24 ACS Paragon Plus Environment

Page 24 of 33

Page 25 of 33

Journal of Agricultural and Food Chemistry

administrated to C57BL/6 mice, and fucoxanthinol (FXOH), which is a primary metabolite of FX, was analyzed in plasma samples collected at several time points during 24 h after the administration. Pt-powder was used as a control. (B) Mucin adsorption characteristics of the two types of FX-loaded nanoparticles. Asterisk indicates significant difference at P < 0.001.

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 33

 Tables

Table 1. Characteristics of FX-CN and FX-CS-CN. Different Asterisks Indicate Significant Differences (*, P < 0.5; **, P < 0.1).

Nanoparticle

Size

Zeta potential

PDI

Encapsulation

(nm)

(mV)

FX-CN

237 ± 13

–12.9 ± 1.6**

0.31 ± 0.03*

73.7 ± 9.0

FX-CS-CN

277 ± 26

24.0 ± 2.8

0.39 ± 0.03

71.8 ± 2.2

efficiency (%)

26 ACS Paragon Plus Environment

Page 27 of 33

Journal of Agricultural and Food Chemistry

Table 2. Pharmacokinetic Parameters for Fucoxanthinol (FXOH) from the Plasma Samples of C57BL/6 Mice Orally Administered with FX-CN, FX-CS-CN, and Pt-powder in a Single Dose Mode. The Difference Letters Represent Significant Differences of Means.

Pharmacokinetic parameters Sample

Tmax

Cmax

T1/2

AUCt

AUC∞

(h)1

(nmol/L)2

(h)3

(nmol·h/L)4

(nmol·h/L)5

a

FX-CN

5.3 ± 2.3

FX-CS-CN

4.7 ± 3.1

Pt-powder

2.7 ± 1.2

ab

b

23.7 ± 9.1

b

48.6 ± 16.2 16.2 ± 8.1

a

ab

b

4.4 ± 0.9 5.1 ± 1.1

ab

6.8 ± 3.3

a

b

324 ± 37

509 ± 134 219 ± 125

1

Cmax: Maximum plasma concentration

2

Tmax: Time for Cmax

3

T1/2: Terminal half-life

4

AUCt,: Area under the plasma concentration-time curve

5

AUC∞: Area under the plasma concentration-time curve to infinity

27 ACS Paragon Plus Environment

276 ± 45 a

b

b

532 ± 121 280 ± 53

a

b

Journal of Agricultural and Food Chemistry

Figure 1

28 ACS Paragon Plus Environment

Page 28 of 33

Page 29 of 33

Journal of Agricultural and Food Chemistry

Figure 2

29 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 3

30 ACS Paragon Plus Environment

Page 30 of 33

Page 31 of 33

Journal of Agricultural and Food Chemistry

Figure 4

31 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 5

32 ACS Paragon Plus Environment

Page 32 of 33

Page 33 of 33

Journal of Agricultural and Food Chemistry

The Table of Contents Graphic

33 ACS Paragon Plus Environment