Enhancement of Non-heme Iron Absorption by Anchovy (Engraulis

Jul 29, 2014 - College of Food Science and Engineering, Ocean University of China, 5 Yushan Road, Qingdao, Shandong Province 266003, China. § School ...
0 downloads 9 Views 897KB Size
Subscriber access provided by MEMORIAL UNIV

Article

Enhancement of non-heme iron absorption by anchovy (Engraulis japonicus) muscle protein hydrolysate involves a nanoparticle-mediated mechanism Haohao Wu, Suqin Zhu, Mingyong Zeng, Zunying Liu, shiyuan dong, Yuanhui Zhao, Hai Huang, and Martin Lo J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/jf5018719 • Publication Date (Web): 29 Jul 2014 Downloaded from http://pubs.acs.org on July 31, 2014

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



Enhancement of non-heme iron absorption by anchovy (Engraulis japonicus) muscle



protein hydrolysate involves a nanoparticle-mediated mechanism





Haohao Wu 1, Suqin Zhu 1, Mingyong Zeng 1,*, Zunying Liu 1, Shiyuan Dong 1, Yuanhui



Zhao 1, Hai Huang 2, Y. Martin Lo 3

6  1



College of Food Science and Engineering, Ocean University of China, 5 Yushan Road, Qingdao, Shandong Province, 266003, China

8  2



School of Aquatic Products, Rizhao Polytechnic, No.16 north of Yantai Road, Rizhao, Shandong Province, 276826, China

10  3

11 

Department of Nutrition and Food Science, University of Maryland, College Park, MD 20742, USA

12  13  *

14 

Corresponding author. Fax: +86 0532 82032400; E-mail address: [email protected]

15 

1   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

16 

ABSTRACT

17 

The mechanisms by which meat enhances human absorption of non-heme iron remain

18 

unknown. Recently, anchovy (Engraulis japonicus) muscle protein hydrolysate (AMPH) was

19 

found to mediate the formation of nano-sized ferric hydrolysis products in vitro. The current

20 

paper evaluates the effects of AMPH on the bioavailability and the intestinal speciation of

21 

non-heme iron in rats, followed by an investigation of cellular uptake pathways of in

22 

vitro-formed AMPH-stabilized nano-sized ferric hydrolysis products (ANPs) by polarized

23 

human intestinal epithelial (Caco-2) cells. The hemoglobin regeneration efficiencies in

24 

anemic rats followed the order of ferric citrate (9.79 ± 2.02 %) < commercial bare α-Fe2O3

25 

nanoparticles (16.37 ± 6.65 %) < mixture of ferric citrate and AMPH (40.33 ± 6.36 %) ≈

26 

ferrous sulfate (40.88 ± 7.67 %) < ANPs (56.25 ± 11.35 %). Percentage contents of intestinal

27 

low-molecular-weight iron in the groups of FC+AMPH, FeSO4 and ANPs were significantly

28 

lower than the corresponding hemoglobin regeneration efficiencies (P < 0.05), providing

29 

strong evidence for the involvement of nano-sized iron in intestinal iron absorption from

30 

FC+AMPH, FeSO4 and ANPs. Calcein-fluorescence measurements of the labile iron pool of

31 

polarized Caco-2 cells revealed the involvement of both divalent transporter 1 and

32 

endocytosis in apical uptake of ANPs, with endocytosis dominating at acidic extracellular pH.

33 

Overall, AMPH enhancement of non-heme iron absorption involves a nanoparticle-mediated

34 

mechanism.

35 

KEYWORDS

36 

Iron bioavailability; Muscle protein; Anchovy; Rats; Caco-2 cell; Nanoparticle. 2   

ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33

Journal of Agricultural and Food Chemistry

37 

ABBREVIATIONS USED

38 

AMPH, anchovy muscle protein hydrolysate; ANPs, in vitro-formed nano-sized ferric

39 

hydrolysis products stabilized by anchovy muscle protein hydrolysate; BPDS,

40 

bathophenanthrolinedisulfonic acid; CDTA, trans-1,2-cyclohexanediaminetetraacetic acid;

41 

DFB, deferoxamine mesylate; DMT1, the divalent metal transporter-1; DTPA, diethylene

42 

triamine pentaacetic acid; EDTA, ethylenediaminetetraacetic acid; EDX, energy dispersive

43 

X-ray spectroscopy; FC, ferric citrate; FC+AMPH, mixture of ferric citrate and anchovy

44 

muscle protein hydrolysate; FeSO4, ferrous sulfate; HRTEM, high resolution transmission

45 

electron microscopy; LMW, low-molecular-weight; Nano-α-Fe2O3, commercial bare

46 

α-Fe2O3 nanoparticles; NTA, nitrilotriacetic acid; TEM, transmission electron microscopy.

47 

3   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

48 

INTRODUCTION

49 

After birth, humans have to acquire iron from the diet to support growth and

50 

compensate iron loss; over two billion people worldwide suffer from iron deficiency. While

51 

dietary iron can be found in the form of heme or non-heme iron, interventions to reduce iron

52 

deficiency usually focus on increasing non-heme iron absorption, which is greatly influenced

53 

by various dietary components. Inhibitors of non-heme iron absorption in humans include

54 

phytates, oxalates, polyphenols, fibers, calcium, copper and zinc; ascorbic acid and muscle

55 

foods (i.e. meat, poultry, and fish) enhance nonheme iron absorption.1 The enhancing effect

56 

of muscle foods on non-heme iron absorption is usually attributed to certain “meat factors”

57 

(e.g. peptides, acidic polysaccharides, and L-α-glycerophosphocholine), which may react

58 

with non-heme iron in the intestinal lumen,  thereby increasing iron solubilization and

59 

absorption.2-4 However, the chemical nature of such soluble non-heme iron has still not been

60 

fully elucidated.

61 

One economical meat source that could be used to combat iron deficiency in

62 

developing countries is the anchovy, a small marine pelagic fish harvested in huge quantity

63 

worldwide, mainly for non-food uses.5 Recently, we demonstrated that anchovy (Engraulis

64 

japonicus) muscle protein hydrolysate (AMPH) could mediate the formation of nano-sized

65 

ferric hydrolysis products within a pH range similar to that of human digestive lumen;6 thus,

66 

AMPH might enhance non-heme iron absorption in humans via a nanoparticle-mediated

67 

mechanism. The current research examined the effects of AMPH on the bioavailability and

68 

the intestinal speciation of non-heme iron in rats, followed by an investigation of cellular 4   

ACS Paragon Plus Environment

Page 4 of 33

Page 5 of 33

Journal of Agricultural and Food Chemistry

69 

uptake pathways of in vitro-formed AMPH-stabilized nano-sized ferric hydrolysis products

70 

(ANPs) in polarized human intestinal epithelial (Caco-2) cells.

71 

MATERIALS AND METHODS

72 

Materials. Japanese anchovy, Engraulis japonicus (5−8 cm total length), caught in the

73 

Yellow Sea by Allen Ship Service Co. Ltd. (Shandong, China) were frozen immediately at

74 

−35 °C on board and transported in a Styrofoam-insulated box to the laboratory refrigerator

75 

(−20 °C) within 4 h. Trypsin from bovine pancreas with a declared activity of 250 U/mg was

76 

purchased from Sinopharm Chemical Co. Ltd. (Shanghai, China). Nano-α-Fe2O3 powder with

77 

declared diameters of 20-30 nm was purchased from Beijing Nachen Co. Ltd. (Beijing,

78 

China). Bathophenanthrolinedisulfonic acid (BPDS), deferoxamine mesylate (DFB),

79 

ethylenediaminetetraacetic acid (EDTA), nitrilotriacetic acid (NTA), phytic acid sodium salt,

80 

diethylene triamine pentaacetic acid (DTPA) and trans-1,2-cyclohexanediaminetetraacetic

81 

acid (CDTA) sodium selenite, hydrocortisone, and triiodothyronine were obtained from

82 

Sigma-Aldrich Co. (St. Louis, USA). Calcein acetoxymethylester, cell culture mediums,

83 

insulin, and epidermal growth factor were purchased from Gibco (Grand Island, USA). All

84 

other reagents used in this study were commercially available and of analytical grade.

85 

Preparation of AMPH and ANPs. AMPH was obtained as described previously.7

86 

Briefly, 25% (w/v) homogenates of anchovy meat was hydrolyzed by trypsin with a 25:1

87 

substrate to enzyme ratio (w/w) at 37 °C and pH 8 for 4 h with continuous magnetic stirring,

88 

followed by heat inactivation of the enzyme in a boiling water bath for 10 min, and the

89 

reaction mixture was then centrifuged at 10000g for 20 min. The supernatant was stored at 5   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 33

90 

4 °C for no more than 48 h and filtered through 0.22 μm cellulose acetate filters before use.

91 

Total protein concentration in the supernatant was determined according to Lowry’s method

92 

using bovine serum albumin as the standard protein.8 ANPs were prepared according to the procedure described by Wu et al

93 

6

with slight

94 

modification. Briefly, one volume of freshly prepared ferric chloride (20 mM) was added

95 

dropwise into ten volumes of magnetic stirred AMPH solution (20 g peptide/L) with the

96 

solution pH maintained at 7.0 ± 0.2 using 1 M sodium hydroxide. The obtained sample

97 

solution was stored at 4 °C for no more than 24 h and filtered through 0.22 μm cellulose

98 

acetate filters before use. For the structural characterization of ANPs, sample solutions were

99 

dropped onto carbon-coated copper grids, allowed to air-dry, and then examined in a JEM

100 

1200 EX (JEOL, Tokyo, Japan) electron microscope. Iron content in the sample solution was

101 

determined by flame atomic absorption spectroscopy using a Shimadzu AA-6800

102 

spectrometer (Kyoto, Japan).

103 

Animal Experiments. All animal handling was conducted according to the principles

104 

and guidelines outlined in National Institute of Health (NIH) Guide for the Care and Use of

105 

Laboratory Animals, and approved by the Ethical Committee of Animal Care and Use at

106 

Ocean University of China (Permit No: 20001013). Sprague-Dawley rats were obtained from

107 

Shandong Lukang Pharmaceutical Group Co., Ltd (Shandong, China). Pelletized purified

108 

AIN-93G-based diets were prepared by TROPHIC Animal Feed High-tech Co., Ltd.

109 

(Nantong, China). The Fe-deficient diet contained ~10 mg/kg of iron derived from raw

110 

materials. The Fe-sufficient diet contained ~25 mg/kg of iron in the form of ferric citrate (FC) 6   

ACS Paragon Plus Environment

Page 7 of 33

Journal of Agricultural and Food Chemistry

111 

and ~10 mg/kg of iron derived from raw materials. The Fe-repletion diets contained ~10

112 

mg/kg of iron derived from raw materials, and ~20 mg/kg of iron in forms of ferrous sulfate

113 

(FeSO4), ANPs, FC, mixture of FC and AMPH (FC+AMPH), and commercial bare α-Fe2O3

114 

nanoparticles (Nano-α-Fe2O3). The iron compounds were incorporated into the diets during

115 

preparation of the wet ingredient mixtures. FeSO4, FC or Nano-α-Fe2O3 equivalent to 20 mg

116 

iron was mixed with 1 kg of the other dry ingredients and 200 mL of Milli-Q water to obtain

117 

the wet ingredient mixture. When incorporating the solution form of AMPH (18 g peptide/L)

118 

and ANPs (1.8 mM iron and 18 g peptide/L), 200 mL of either of them was mixed with 1 kg

119 

of the other dry ingredients to make the wet ingredient mixture. After being pelletized, the

120 

wet ingredient mixture was dried overnight at 40°C in a drying oven. Iron contents in the

121 

dried pelletized diets were determined by flame atomic absorption spectroscopy  using a 

122 

Shimadzu AA-6800 spectrometer (Kyoto, Japan). Rats received Milli-Q water ad libitum

123 

throughout the experiments. Body weight was recorded at the beginning and end of each

124 

experimental period.

125 

The iron bioavailability was measured in anemic rats using a hemoglobin regeneration

126 

bioassay.9,10 Female Sprague-Dawley rats (n = 12) breastfeeding male pups (16 days old; n =

127 

52) were housed in individual stainless steel cages with four to five pups each and fed a

128 

Fe-deficient diet ad libitum for a week. On the weaning day, 52 male rats (23 days old) were

129 

transferred into individual stainless steel cages under controlled conditions at temperatures of

130 

22 ± 2 °C and relative humidities of 55 ± 10% with a light/dark cycle of 12 h/12 h. The

131 

just-weaned rats were stratified according to body weight and within each stratum were 7   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

132 

randomly allocated at random to two groups: the Fe-depleted group (n = 48) fed a

133 

Fe-deficient diet ad libitum and a control group (n = 4) fed a Fe-sufficient diet ad libitum.

134 

After a Fe-depletion period of 27 days, 20 µL of fresh rat blood sampled by tail vein incision

135 

was diluted with EDTA and subsequently analyzed using a Cell-Dyn 1600 hematology

136 

analyzer (Abbott Laboratories, Abbott Park, USA). The mean hemoglobin concentrations of

137 

the Fe-depleted group and the control group were 71.1 ± 13.8 g/L and 153.8 ± 13.9 g/L,

138 

respectively. Rats in the Fe-depleted group were stratified by the hemoglobin concentrations

139 

and randomized into six groups (eight rats each). The rats in each group consumed one of the

140 

five Fe-repletion diets containing FeSO4, ANPs, FC, FC+AMPH or Nano-α-Fe2O3, or the

141 

Fe-deficient diet for 13 days (the repletion period) ad libitum. Rats in the control group

142 

remained on the Fe-sufficient diet during the repletion period.

143 

Periodically, daily food intake of each rat was manually recorded. A known and

144 

adequate amount of a diet was given to a rat in the feeding container  between nine and ten

145 

o’clock in the morning. After 24 h, the diet remaining in the feeding container and at the

146 

bottom of the cage was reweighed; calculating the differences showed the rat’s daily food

147 

intake.

148 

At the end of the repletion period, blood samples were once again taken by tail vein

149 

incision. The rats were anaesthetized with ether before their livers were taken, rinsed with

150 

saline and stored at -20 °C. Iron concentration in the liver was measured by flame atomic

151 

absorption spectroscopy using a Shimadzu AA-6800 spectrometer (Kyoto, Japan). Hemoglobin iron pool (mg), assuming a total rat blood volume of 6.7 % body weight

152 

8   

ACS Paragon Plus Environment

Page 8 of 33

Page 9 of 33

153 

Journal of Agricultural and Food Chemistry

and an iron content in hemoglobin of 0.335 %, was calculated as follows:10 Hemoglobin iron pool (mg) = body weight (kg) × 0.067 × hemoglobin concentration

154  155 

(g/L) × 3.35.

156 

The hemoglobin regeneration efficiency was calculated based on the following equation:

157 

Hemoglobin regeneration efficiency (%) = [final hemoglobin iron pool (mg) – initial

158 

hemoglobin iron pool (mg)] / total iron consumed (mg) × 100. The relative biological value was calculated based on the hemoglobin regeneration

159  160 

efficiencies: Relative biological value (%) = hemoglobin regeneration efficiency of each animal /

161  162 

average hemoglobin regeneration efficiency of the FeSO4 group × 100.

163 

Iron speciation analysis of rat intestinal contents was carried out as previously described

164 

by Kapsokefalou and Miller11 and Simpson et al.12 with some modifications. Male

165 

Sprague-Dawley rats (53 days old; n = 40) weighing 290−310 g were stratified according to

166 

body weight and within each stratum were randomly distributed into five groups (eight rats

167 

each). The rats were fed the Fe-deficient diet for a week before they were daytime fasted for

168 

9 h. The fasted rats in each group then consumed one of the five Fe-repletion diets for 30 min

169 

ad libitum and were anesthetized with ether ninety minutes after consumption of the meal.

170 

The rats were killed by cervical dislocation before an abdominal midline incision was

171 

performed. After the small intestine was removed from the carcass, content in the proximal

172 

one-third segment was flushed with saline into a centrifuge tube and further fractionated into 9   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

173 

two fractions using Millipore filters (Göttingen, Germany): the 0.22-µm filtrate and the

174 

3-kDa (≈ 1 nm) filtrate. Iron in the intestinal content, 0.22-µm filtrate and 3-kDa filtrate were

175 

determined by flame atomic absorption spectroscopy using a Shimadzu AA-6800

176 

spectrometer (Kyoto, Japan).

177 

Cellular Uptake Experiment. Caco-2 cells were obtained from the Cell Bank of the

178 

Chinese Academy of Sciences (Shanghai, China), and seeded at a density of 5×104 cell/cm2

179 

in collagen-treated 24-well plates (BD Biosciences, San Jose, CA, USA). The cells were

180 

routinely maintained in high glucose (4.5 g/L) Dulbecco’s Modified Eagle Medium

181 

supplemented with 10% heat-inactivated fetal bovine serum, 25 mM HEPES, 4mM

182 

glutamine and 1 mM sodium pyruvate at 37 °C in an incubator with a 95% air/5% CO2

183 

atmosphere at constant humidity, and the medium was renewed every two days. After

184 

reaching confluence (2-3 days post-seeding), the cells were maintained in complete

185 

Dulbecco’s Modified Eagle Medium for another 12 days to differentiate completely, and

186 

further incubated for 24 h in serum-free minimum essential medium supplemented with 4

187 

mg/L hydrocortisone, 5 mg/L insulin, 5 µg/L selenium, 34 µg/L triiodothyronine, and 20 µg/L

188 

epidermal growth factor. Following this, the medium was aspirated, and the cells were stained

189 

with 1 µM calcein acetoxymethylester in Hanks’ balanced salt solution for 30 min at 37 °C

190 

before being washed three times with HBSS. The cells were then incubated in Tyrode

191 

solution (137 mM NaCl, 2.7 mM KCl, 1 mM MgCl2, 1.8 mM CaCl2, 5.5 mM D-glucose)

192 

buffered with 20 mM MES (pH 5.5) or MOPS (pH 7.0) containing FeSO4, ANPs, or any of

193 

1:1 ferric chelates of NTA, phytate, EDTA, DTPA, CDTA and DFB at the concentration 10   

ACS Paragon Plus Environment

Page 10 of 33

Page 11 of 33

Journal of Agricultural and Food Chemistry

194 

equivalent to 20 µM iron in a Synergy H4 Hybrid Microplate Reader (Bio-Tek, Winooski, VT,

195 

USA) at 37 °C for 2.5 h, and calcein fluorescence (485 nm excitation, 530 nm emission) was

196 

monitored every 5 min. To trap and detect ferrous iron, 100 µM BPDS was used, and the

197 

absorbance of the ferrous tris(BPDS) complex was measured at 536 nm (molar extinction

198 

coefficient, ε =1.927 × 104 L mol-1 cm-1). Stock solution of FeSO4 (1 mM) was freshly

199 

prepared before use. Stock solution of ferric chelates (1:1 molar ratio) of NTA, phytate,

200 

EDTA, DTPA, CDTA, and DFB was prepared by mixing of equal volume of chelators (2 mM)

201 

and freshly prepared FeCl3 (2 mM).

202 

Statistical Analysis. Data are expressed as the mean ± standard deviation (SD).

203 

Differences between groups were tested for significance by the least significant difference

204 

(LSD) mean comparison using the SPSS software program (SPSS Inc., Chicago, IL, USA).

205 

The relation between variables was determined by the Pearson correlation test. The

206 

probability level < 0.05 was considered statistically significant.

207 

RESULTS AND DISCUSSION

208 

To evaluate the structural properties of ANPs, i.e. the products of ferric hydrolysis in the

209 

presence of AMPH in vitro at room temperature, we used transmission electron microscopy

210 

(TEM), high resolution transmission electron microscopy (HRTEM), electron diffraction, and

211 

energy dispersive X-ray spectroscopy (EDX). TEM imaging (Figure 1A) showed that the

212 

average particle size ranged from 10 to 20 nm, and this size was greater than pure two-line

213 

ferrihydrite (2−4 nm).13 Whole area EDX analysis (Figure 1B) of one of the particles in

214 

Figure 1A revealed three peaks (0.7, 6.4, and 7.1 keV) that corresponded to the Lα, Kα, and 11   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

215 

Kβ lines of iron, respectively, so the particles in the TEM image were iron-containing. In

216 

addition, the EDX spectra also showed prominent peaks for copper, sodium, potassium,

217 

chloride, sulfur, carbon, oxygen, phosphorus and silicon. The copper obviously represented

218 

the supporting copper grid, and the silicon peak was presumably due to contamination.

219 

Spurious silicon peaks have been recognized as a problem in EDX analysis of biological

220 

specimens due to silicon contamination from oils or greases during the course of specimen

221 

preparation or from the vacuum system of the electron microscope.14 Any remaining peaks

222 

would presumably represent true compositions of AMPH or the specimen support film. The

223 

HRTEM image (Figure 1C) revealed no recognizable lattice fringe, suggesting the

224 

amorphous property of ANPs. The selected area electron diffraction (Figure 1D) showed a

225 

two-ring pattern similar to that of 2-line ferrihydrite with the obtained d-spacing values (2.90

226 

and 1.97 Å), a little higher than typical corresponding values of the latter (~2.5 and 1.5 Å).

227 

ANPs are likely to be a poorly crystalline ferric oxyhydroxide similar to 2-line ferrihydrite. It

228 

is known that poorly crystalline iron oxides are prone to electron beam damage,15 so the

229 

actual structure of ANPs might vary somewhat in size and crystalline phase from that

230 

observed under TEM, HRTEM and electron diffraction.

231 

Details and results of the hemoglobin regeneration bioassay are shown in Table 1. It is

232 

known that body iron is distributed in hemoglobin (65%), iron stores (ferritin and

233 

hemosiderin, 30%), myoglobin (3.5%) and other heme proteins (1.5%).16 Liver iron stores

234 

did not show significant increase in all Fe-repletion groups during the repletion period

235 

(Figure 2), so most of the iron absorbed should be used to generate new hemoglobin 12   

ACS Paragon Plus Environment

Page 12 of 33

Page 13 of 33

Journal of Agricultural and Food Chemistry

236 

molecules, amounts of which could be calculated from changes of blood volumes and

237 

hemoglobin concentrations. The negative changes of the hemoglobin concentrations in

238 

groups of Fe-deficient, FC and Fe-sufficient were the result of a marked increase in blood

239 

volume, which is proportional to body weight, during fast growth of the anemic rat. As

240 

shown in Table 1, body weights and hemoglobin concentrations varied independently among

241 

the groups, so to measure iron bioavailability by hemoglobin regeneration in anemic rats, it is

242 

necessary to calculate the hemoglobin regeneration efficiencies from body weight gain, diet

243 

intake, diet iron content, and hemoglobin concentration.

244 

As shown in Table 2, the hemoglobin regeneration efficiency of FC+AMPH was as

245 

high as that of FeSO4, a gold standard in iron bioavailability studies, and significantly higher

246 

than that of FC alone (P < 0.05). Apparently, AMPH effectively enhanced iron absorption

247 

from FC. ANPs displayed an even higher hemoglobin regeneration efficiency than

248 

FC+AMPH and FeSO4, although not statistically significant (P = 0.06 and 0.12, respectively).

249 

Since crystalized bare iron oxide nanoparticles feature simple production process and low

250 

cost, we examined the bioavailability of iron from Nano-α-Fe2O3, a commercial product with

251 

declared size similar to ANPs. The hemoglobin regeneration efficiency of Nano-α-Fe2O3 was

252 

found to be significantly lower than that of ANPs (P < 0.05). Interestingly, the relative

253 

biological value of Nano-α-Fe2O3 (39.0 ± 15.8%) was also found to be significantly lower

254 

than those of previously reported ZnO-, MgO-, CaO- or FePO4-doped iron oxide

255 

nanopowders, namely Fe2O3/ZnO/MgO (81.8−100.1%), Fe2O3/ZnO/CaO (72.4−92.6%),

256 

Fe2O3/ZnO (68.0−87.5%), and Fe2O3/FePO4 (69.0−87.4%),9 which might be explained by the 13   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

257 

higher iron dissolution rates of these latter iron oxide nanostructures in gastric acid due to

258 

doping with compounds more soluble in dilute acid than iron oxide.17

259 

To elucidate whether nano-sized iron involved iron absorption in vivo, we analyzed

260 

iron speciation in the luminal content of the proximal one-third segment of small intestine,

261 

where dietary iron is mainly absorbed.18 Iron in the 3-kDa filtrates were regarded as the truly

262 

dissolved iron species or low-molecular-weight (LMW) iron, whereas that in the 0.22-µm

263 

filtrate was supposed as total soluble iron including both LMW iron and nano-sized colloidal

264 

iron. As shown in Table 2, percentage contents of intestinal LMW iron in the groups of

265 

FC+AMPH, FeSO4 and ANPs were significantly lower than the corresponding hemoglobin

266 

regeneration efficiencies (P < 0.05), so LMW iron alone was unlikely to account for total iron

267 

absorption in these groups. This provided strong evidence for the involvement of nano-sized

268 

iron in intestinal iron absorption from FC+AMPH, FeSO4 and ANPs.

269 

LMW iron in the intestinal tract might be ferrous iron or ferric chelates. Except for

270 

FeSO4, the other four iron compounds in the present study are all in the form of ferric iron.

271 

The reducing power of tert-butylhydroquinone, the fat-soluble antioxidant added in the

272 

AIN-93G diet, might not contribute to the formation of intestinal LMW iron due to its very

273 

low solubility in the aqueous phase. Although meat digestion products are capable of

274 

reducing non-heme ferric iron to the ferrous form in vitro,19 AMPH and ANPs did not

275 

increase the percentage content of LMW iron (Table 2), which might be explained by the

276 

oxidation of protein sulfhydryl groups during preparation and storage of AMPH and

277 

pelletized diets.20 There are usually considerable amounts of residual phytates in the 14   

ACS Paragon Plus Environment

Page 14 of 33

Page 15 of 33

Journal of Agricultural and Food Chemistry

278 

commercial starch that comprises an important part of the AIN-93G diet,21 so iron phytate

279 

chelates might account for part of the intestinal LMW iron in the present study. It is also

280 

possible that some vitamins added in the AIN-93G diet could form LMW ferric complexes in

281 

the intestinal lumen.22

282 

Ferric hydrolytic polymers could be stabilized at the nanoscale by biomacromolecules

283 

(e.g. polysaccharides and peptides) derived from some common food components (e.g. starch,

284 

carrageenan, pectin, alginate, casein, whey, and muscle protein) or gastrointestinal secretions

285 

(e.g. mucins).6,23-27 The percentage contents of nano-sized iron in the rat proximal small

286 

intestinal tracts were calculated by subtraction of LMW iron from total soluble iron (Table 2),

287 

and were found to be over 30% in all five groups, suggesting that AIN-93G diet could induce

288 

a high basal level of intestinal nano-sized iron. This might be attributed to the digestion

289 

products of maltodextrin, starch and casein in the basal formulation of AIN-93G diet or

290 

gastrointestinal secretions. Interestingly, the percentage of nano-sized iron for the FeSO4

291 

group was also significantly higher than that of its LMW counterpart (P < 0.05), which might

292 

be explained by the formation of nano-sized iron via rapid in situ oxidation and hydrolysis of

293 

ferrous iron bound to soluble macromolecules in the presence of swallowed oxygen in the

294 

intestinal tract. Due to their tendency to aggregate and sediment in aqueous solutions, bare

295 

iron oxide nanoparticles are difficult to disperse into stable suspensions within the nanoscale

296 

range unless very efficient dispersing techniques (e.g. probe ultrasonication) are employed,28

297 

so it is unsurprising that Nano-α-Fe2O3 induced a significantly lower level of intestinal

298 

nano-sized iron than ANPs (P < 0.05). 15   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

299 

The involvement of intestinal nano-sized iron in iron absorption from FC+AMPH,

300 

FeSO4 and ANPs raises questions about how gut epithelial cells absorb intraluminal

301 

nano-sized iron. Usually, exogenous iron enters the cellular labile iron pool before it is stored

302 

in ferritin, exported from the cell, or used to synthesize iron proteins.29 Calcein, a

303 

fluorescence probe for the labile iron pool, has been frequently used to monitor cellular iron

304 

uptake in real time.30,31 We thus used calcein-loaded polarized Caco-2 cell monolayers as a

305 

gut epithelial model to investigate the intestinal iron absorption from ANPs, a mimetic for

306 

intraluminal nano-sized iron. As shown in Figure 3, iron uptake from FeSO4 and ANPs at

307 

concentrations equivalent to 20 µM iron was monitored by the quenching of calcein

308 

fluorescence, i.e. the incorporation of exogenous iron into the cellular labile iron pool, under

309 

two pH conditions (5.5 and 7.0) simulating those in the proximal and distal small intestine.32

310 

BPDS, a strong ferrous chelator, was used to block iron absorption via the divalent metal

311 

transporter 1 (DMT1). As expected, cellular iron uptake from FeSO4 was completely blocked

312 

by BPDS at pH 5.5 and 7.0 (data for the latter pH not shown), while iron uptake from ANPs

313 

was only partially blocked at both pH values, suggesting that both DMT1-dependent and

314 

DMT1-independent mechanisms are involved in iron absorption from ANPs.

315 

The two best known DMT1-independent iron uptake mechanisms are by means of

316 

heme carrier protein 1 and endocytosis. Heme carrier protein 1 is specific to heme iron

317 

uptake,33 so endocytosis, a very common mechanism for cellular entry of nanoparticles,

318 

attracted much attention in recent studies on intestinal absorption of nano-sized iron. Work by

319 

Theil and colleagues demonstrated that iron absorption from ferritin (i.e., ferrihydrite-like 16   

ACS Paragon Plus Environment

Page 16 of 33

Page 17 of 33

Journal of Agricultural and Food Chemistry

320 

nanostructures encapsulated by a protein cage) is independent of heme iron and ferrous salts

321 

in both women and rat intestinal segments,34 and a paper published later by Antileo et al.

322 

revealed an endocytic pathway of ferritin in Caco-2 cells using TEM and fluorescence

323 

confocal microscopy.35 The elegant studies of Pereira et al. and Powell et al.,36,37 utilizing 

324 

Caco-2 cells and mice, proved that intestinal uptake of luminally hydrolyzed nano-sized iron

325 

invokes a nanoparticulate endocytic pathway. Thus, endocytosis could be the mechanism for

326 

DMT1-independent apical uptake of ANPs.

327 

The calcein fluorescence changes in Figure 3 also showed a more noticeable uptake of

328 

iron from ANPs at pH 5.5 in the presence of BPDS than that observed at pH 7.0, suggesting a

329 

lower intraluminal pH in the proximal small intestine favors the endocytic uptake of ANPs.

330 

Acidic extracellular pH enhancement of endocytosis due to the proton-induced change of

331 

cellular membrane properties has been well-documented,38 so it is not surprising that in the

332 

present study an extracellular pH of 5.5 was more favorable for the endocytic uptake of ANPs

333 

compared to pH 7.0. The marginal inhibition of iron absorption from ANPs by BPDS at pH

334 

5.5 indicated the dominant role of the endocytic pathway at acidic extracellular pH. Iron

335 

uptake from FeSO4 at pH 7.0 without the presence of BPDS was less evident than that at pH

336 

5.5 after about 20 min of incubation, which could be explained by the relatively fast

337 

oxidation rate of ferrous iron at pH 7.0.39

338 

Ferric iron must be reduced to the ferrous form by duodenal cytochrome-b in the

339 

apical membrane of enterocytes before being absorbed by the DMT1 pathway, so the part

340 

played by the DMT1-dependent pathway in apical uptake of ANPs depends on the 17   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 18 of 33

341 

reducibility of ANPs. As shown in Figure 5, the reducibility of ferric chelates of NTA,

342 

phytate, EDTA, DTPA, CDTA, and DFB to polarized Caco-2 cells followed the reversed

343 

order of ferric stability constants of the chelators,40 and ANPs were as reducible as ferric

344 

NTA, which is the most reducible of these ferric chelates.

345 

Some previous studies have revealed that reduction rates of ferric chelates by plant

346 

roots (Arachis hypogeae L.) and yeast (Saccharomyces cerevisiae) were inversely related to

347 

the ferric stability constants,41,42 and our data confirmed the previous observations in human

348 

intestinal cells. The reducibility of iron oxides is greatly dependent on their crystalline phases

349 

(e.g. ferrihydrite, lepidocrocite, maghemite, goethite, and hematite) and surrounding organic

350 

substances. Ferrihydrite is the most reducible phase of iron oxides. The phase of intestinal

351 

nano-sized iron in the present study should be similar to ANPs, a poorly crystalline ferric

352 

oxyhydroxide like 2-line ferrihydrite. Iron-binding organic substances such as AMPH in the

353 

present study might enhance reduction dissolution of iron oxides via solubilizing ferric iron

354 

from the oxide surface or complexation of ferrous iron.43 The high reducibility of ANPs to

355 

polarized Caco-2 cell monolayers could thus be easily understood. In

356 

summary,

our

data

suggest

that

intestinal

nano-sized

iron

involves

357 

AMPH-enhancement of non-heme iron absorption in rats. While both DMT1 and endocytosis

358 

play parts in the intestinal absorption of luminal nano-sized iron, endocytosis dominates at

359 

acidic extracellular pH levels.

360 

18   

ACS Paragon Plus Environment

Page 19 of 33

361 

Journal of Agricultural and Food Chemistry

ACKNOWLEDGEMENTS

362 

Financially supported by  the National Natural Science Foundation of China (No.

363 

31371758), National Key Technology Research and Development Program of China (No.

364 

2012BAD28B05), and China Scholarship Council (No. 201306330015).

365 

REFERENCES

366 

(1) WHO/UNICEF/UNU. Iron Deficiency Anemia Assessment, Prevention, and Control; World Health Organization: Geneva, Switzerland, 2001.

367 

368 

(2) Storcksdieck genannt Bonsmann, S.; Hurrell, R. F. Iron-binding properties, amino acid

369 

composition, and structure of muscle tissue peptides from in vitro digestion of different

370 

meat sources. J. Food. Sci. 2007, 72, S019–029.

371 

(3) Laparra, J. M.; Tako, E.; Glahn, R. P.; Miller, D. D. Isolated glycosaminoglycans from

372 

cooked haddock enhance non-heme iron uptake by Caco-2 cells. J. Agric. Food. Chem.

373 

2008, 56, 10346–10351.

374 

(4) Armah, C. N.; Sharp, P.; Mellon, F. A.; Pariagh, S.; Lund, E. K.; Dainty, J. R.; Teucher,

375 

B.;

376 

enhancement of non-heme iron absorption. J. Nutr. 2008, 138, 873–877.

377 

Fairweather-Tait,

S.

J.

L-a-glycerophosphocholine

contributes

to

meat’s

(5) Food and Agriculture Organization of the United Nations. Global Capture Production 1950–2010;

378 

19   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

379 

http://www.fao.org.proxy-um.researchport.umd.edu/fishery/statistics/global-capture-pro

380 

duction/query/en (accessed Mar 2014).

381 

(6) Wu, H.; Liu, Z.; Dong, S.; Zhao, Y.; Huang, H.; Zeng, M. Formation of ferric

382 

oxyhydroxide nanoparticles mediated by peptides in anchovy (Engraulis japonicus)

383 

muscle protein hydrolysate. J. Agric. Food. Chem. 2013, 61, 219–224.

384 

(7) Wu, H.; Liu, Z.; Zhao, Y.; Zeng, M. Enzymatic preparation and characterization of

385 

iron-chelating peptides from anchovy (Engraulis japonicus) muscle protein. Food Res.

386 

Int. 2012, 48, 435–441.

387 

(8) Lowry, O. H.; Rosebrough, N. J.; Farr, A. L.; Randall, R. J. Protein measurement with the Folin phenol reagent. J. Biol. Chem. 1951, 193, 265–275.

388 

389 

(9) Hilty, F. M.; Arnold, M.; Hilbe, M.; Teleki, A;  Knijnenburg, J. T.; Ehrensperger, F.;

390 

Hurrell, R. F.; Pratsinis, S. E.; Langhans, W.; Zimmermann, M. B. Iron from

391 

nanocompounds containing iron and zinc is highly bioavailable in rats without tissue

392 

accumulation. Nat. Nanotechnol. 2010, 5, 374–380.

393 

(10) Mahoney, A. W., Van Orden, C. C., Hendricks, D. G. Efficiency of converting food iron into hemoglobin by the anemic rat. Nutr Metabolism. 1974, 17, 223–230.

394 

395 

(11) Kapsokefalou, M.; Miller, D. D.  Iron speciation in intestinal contents of rats fed meals

396 

composed of meat and nonmeat sources of protein and fat. Food. Chem. 1995, 52, 47–56.

397 

(12) Simpson, R. J.; Sidhar, S.; Peters, T. J. Application of selective extraction to the study of

398 

iron species present in diet and rat gastrointestinal tract contents. Br J Nutr. 1992, 67, 20   

ACS Paragon Plus Environment

Page 20 of 33

Page 21 of 33

Journal of Agricultural and Food Chemistry

437-444.

399 

400 

(13) Janney, D. E.; Cowley, J. M.; Buseck, P. R. Transmission electron microscopy of synthetic 2-and 6-line ferrihydrite. Clay. Clay. Miner. 2000, 48, 111–119.

401 

402 

(14) Smith, N. K. A review of sources of spurious silicon peaks in electron microprobe X-ray spectra of biological specimens. Anal Biochem. 1979, 94, 100-104.

403 

404 

(15) Pan, Y. H.; Vaughan, G.; Brydson, R.; Bleloch, A.; Gass, M.; Sader, K.; Brown, A.

405 

Electron-beam-induced reduction of Fe(3+) in iron phosphate dihydrate, ferrihydrite,

406 

haemosiderin and ferritin as revealed by electron energy-loss spectroscopy.

407 

Ultramicroscopy. 2010, 110, 1020-1032.

408 

(16) Hoffbrand, V.; Moss, P. Essential haematology, edition 6th. John Wiley & Sons: West Sussex, UK, 2011; pp. 34-36.

409 

410 

(17) Hilty, F. M.; Knijnenburg, J. T. N.; Teleki, A.; Krumeich, F.; Hurrell, R. F.; Pratsinis, S.

411 

E.; Zimmermann, M. B. Incorporation of Mg and Ca into nanostructured Fe2O3

412 

improves Fe solubility in dilute acid and sensory characteristics in foods. J. Food. Sci.

413 

2011, 76, N2–N10.

414 

(18) McKie, A.; Simpson, R. Intestinal iron absorption. In Iron Physiology and

415 

Pathophysiology in Humans; Anderson, G.; McLaren, G. D., Eds.; Humana Press:  New

416 

York, NY, 2012; pp. 101-116.

417 

(19) Kapsokefalou, M.; Miller, D. D. Effects of meat and selected food components on the

418 

valence of non-heme iron during in vitro digestion. J. Food. Sci. 1991, 56, 352–355. 21   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

419 

(20) Karava, N. B.; Mahoney, R. R. Lyophilization decreases the formation of dialyzable iron

420 

by extraction and digestion of chicken breast muscle. Int. J. Food. Sci. Nutr. 2011, 62,

421 

397–403.

422 

(21) Ma, G.; Jin, Y.; Piao, J.; Kok, F.; Guusje, B.; Jacobsen, E. Phytate, calcium, iron, and

423 

zinc contents and their molar ratios in foods commonly consumed in China. J. Agric.

424 

Food. Chem. 2005, 53, 10285–10290.

425 

(22) Layrisse, M.; García-Casal, M. N.; Solano, L.; Barón, M. A.; Arguello, F.; Llovera, D.;

426 

Ramírez, J.; Leets, I.; Tropper, E. New property of vitamin A and beta-carotene on

427 

human iron absorption: effect on phytate and polyphenols as inhibitors of iron

428 

absorption. Arch. Latinoam. Nutr. 2000, 50, 243–248.

429 

(23) Somsook, E.; Hinsin, D.; Buakhrong, P.; Teanchai, R.; Mophan, N.; Pohmakotr, M.;

430 

Shiowatana, J. Interactions between iron(III) and sucrose, dextran, or starch in

431 

complexes. Carbohydr. Polym. 2005, 61, 281–287.

432 

(24) Chaud, M. V.; Izumi, C.; Nahaal, Z.; Shuhama, T.; Bianchi Mde, L.; de Freitas, O. Iron

433 

derivatives from casein hydrolysates as a potential source in the treatment of iron

434 

deficiency. J. Agric. Food. Chem. 2002, 50, 871–877.

435 

(25) Jones, F.; Cölfen, H.,; Antoneitti, M. Iron oxyhydroxide colloids stabilized with polysaccharides. Colloid. Polym. Sci. 2000, 278, 491–501.

436 

22   

ACS Paragon Plus Environment

Page 22 of 33

Page 23 of 33

Journal of Agricultural and Food Chemistry

437 

(26) Hurrell, R. F.; Lynch, S. R.; Trinidad, T. P.; Dassenko, S. A.; Cook, J. D. Iron

438 

absorption in humans as influenced by bovine milk proteins. Am J Clin Nutr. 1989, 49,

439 

546-–552.

440 

(27) Conrad, M. E.; Umbreit, J. N. A concise review: Iron absorption—The

441 

mucin-mobilferrin-integrin pathway. A competitive pathway for metal absorption. Am. J.

442 

Hematol. 1993, 42, 67–73.

443 

(28) Dickson, D.; Liu, G.; Li, C.; Tachiev, G.; Cai, Y. Dispersion and stability of bare

444 

hematite nanoparticles: effect of dispersion tools, nanoparticle concentration, humic acid

445 

and ionic strength. Sci. Total. Environ. 2012, 419, 170–177.

446 

(29) Kakhlon, O.; Cabantchik, Z. I. The labile iron pool: characterization, measurement, and participation in cellular processes. Free Radic Biol Med. 2002, 33, 1037–1046.

447 

448 

(30) San Martin, C. D.; Garri, C.; Pizarro, F.; Walter, T.; Theil, E. C.; Núñez, M. T. Caco-2

449 

intestinal epithelial cells absorb soybean ferritin by µ2 (AP2)-dependent endocytosis. J

450 

Nutr. 2008, 138, 659-666.

451 

(31) Ma, Y.; Yeh, M.; Yeh, K. Y.; Glass, J. Iron Imports. V. Transport of iron through the

452 

intestinal epithelium. Am J Physiol Gastrointest Liver Physiol. 2006, 290, G417-422.

453 

(32) Dressman, J. B.; Berardi, R. R.; Dermentzoglou, L. C.; Russell, T. L.; Schmaltz, S. P.;

454 

Barnett, J. L.; Jarvenpaa, K. M. Upper gastrointestinal (GI) pH in young, healthy men

455 

and women. Pharm Res. 1990, 7, 756–761.

23   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

456 

(33) Latunde-Dada, G. O.; Takeuchi, K.; Simpson, R.J.; McKie, A. T. Haem carrier protein 1

457 

(HCP1): Expression and functional studies in cultured cells. FEBS Lett. 2006, 580,

458 

6865-6870.

459 

(34) Theil, E. C.; Chen, H.; Miranda, C.; Janser, H.; Elsenhans, B.; Núñez, M. T.; Pizarro, F.;

460 

Schümann, K. Absorption of iron from ferritin is independent of heme iron and ferrous

461 

salts in women and rat intestinal segments. J Nutr. 2012, 142, 478-483.

462 

(35) Antileo, E.; Garri, C.; Tapia, V.; Muñoz, J. P.; Chiong, M.; Nualart, F.; Lavandero, S.;

463 

Fernández, J.; Núñez, M. T. Endocytic pathway of exogenous iron-loaded ferritin in

464 

intestinal epithelial (Caco-2) cells. Am. J. Physiol. Gastrointest. Liver. Physiol. 2013,

465 

304, G655–661.

466 

(36) Pereira, D. I. A.; Mergler, B. I.; Faria, N.; Bruggraber, S. F. A.;  Aslam, M. F.; Poots, L.

467 

K.; Prassmayer, L.; Lönnerdal, B.; Brown, A. P.; Powell, J. J. Caco-2 cell acquisition of

468 

dietary iron(III) invokes a nanoparticulate endocytic pathway. PLoS. ONE. 2013, 8,

469 

e81250.

470 

(37) Powell, J. J., Bruggraber, S. F., Faria, N., Poots, L. K., Hondow, N,, Pennycook, T. J.,

471 

Latunde-Dada, G. O., Simpson, R. J., Brown, A. P., Pereira, D. I. A nano-disperse

472 

ferritin-core mimetic that efficiently corrects anemia without luminal iron redox activity.

473 

Nanomedicine. 2014. DOI: 10.1016/j.nano.2013.12.011.

24   

ACS Paragon Plus Environment

Page 24 of 33

Page 25 of 33

Journal of Agricultural and Food Chemistry

474 

(38) Ben-Dov, N.; Korenstein, R. Proton-induced endocytosis is dependent on cell membrane

475 

fluidity, lipid-phase order and the membrane resting potential. Biochim. Biophys Acta.

476 

2013, 1828, 2672-2681.

477 

(39) Morgan, B.; Lahav, O. The effect of pH on the kinetics of spontaneous Fe(II) oxidation

478 

by O2 in aqueous solution--basic principles and a simple heuristic description.

479 

Chemosphere. 2007, 68, 2080–2084.

480 

(40) Martell, A. E.; Smith, R. M. (Eds.), NIST Standard Reference Database 46: NIST

481 

critically selected stability constants of metal complexes: version 8, National Istitute of

482 

Standards and Technology, Gaithersburg, 2004.

483 

(41) Römheld, V.; Marschner, H. Mechanism of iron uptake by peanut plants. I. FeIII reduction, chelate splitting, and release of phenolics. Plant. Physiol. 1983, 71, 949–954.

484 

485 

(42) Lesuisse, E.; Raguzzi, F.; Crichton, R. R. Iron uptake by the yeast Saccharomyces cerevisiae: involvement of a reduction step. J. Gen. Microbiol. 1987, 133, 3229–3236.

486 

487 

(43) Wu, H.; Yin, J. J.; Wamer, W. G.; Zeng, M.; Lo, Y. M. Reactive oxygen species-related

488 

activities of nano-iron metal and nano-iron oxides. J. Food. Drug. Anal. 2014, 22, 86–

489 

94.

490 

491 

25   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 33

492 

Figure captions

493 

Figure 1. Structural characterization of in vitro-formed nano-sized ferric hydrolysis products

494 

stabilized by anchovy muscle protein hydrolysate (ANPs). A) Transmission electron

495 

microscopy images collected from a drop of ANPs. B) Whole area energy dispersive X-ray

496 

analysis of one of the particles in “A” showing elemental compositions.  Peaks for sodium

497 

(Na), copper (Cu), potassium (K), silicon (Si), phosphorus (P), sulphur (S), chlorine (Cl),

498 

carbon (C), and oxygen (O) are identified. C) High resolution transmission electron

499 

microscopy. D) Selected area electron diffraction.

500 

Figure 2. Liver iron concentrations at the end of the iron-repletion period. Abbreviations: FC,

501 

ferric citrate; FC+AMPH, the mixture of ferric citrate and anchovy muscle protein

502 

hydrolysate, FeSO4, ferrous sulfate; ANPs,

503 

products stabilized by anchovy muscle protein hydrolysate; Nano-α-Fe2O3, commercial bare

504 

α-Fe2O3 nanoparticles. Values are means ± standard deviations of a representative experiment

505 

(n = 4 for the Fe-sufficient group; n = 8 for the other groups). Groups not sharing a common

506 

letter are significantly different (P < 0.05).

507 

Figure 3. Iron uptake by polarized Caco-2 cell monolayers in the presence of 20 µM iron at

508 

extracellular pH values of 5.5 and 7.0. Abbreviations: FeSO4, ferrous sulfate; ANPs, in

509 

vitro-formed

510 

bathophenanthrolinedisulfonic acid. Cellular uptake of exogenous iron into the labile iron

511 

pool was measured as the quenching of calcein fluorescence. BPDS is a strong ferrous

512 

chelator to block the  divalent metal transportor-1 pathway. Data are means of three

513 

independent experiments with the standard deviations less than 1.6.

514 

Figure 4. Ferrous concentrations after 2-h incubation of 20 µM different forms of ferric iron

515 

with polarized Caco-2 cell monolayers at an extracellular pH of 5.5. Abbreviations: ANPs, in

516 

vitro-formed nano-sized ferric hydrolysis products stabilized by anchovy muscle protein

517 

hydrolysate; CDTA, trans-1,2-cyclohexanediaminetetraacetic acid; DFB, deferoxamine

518 

mesylate; DTPA, diethylene triamine pentaacetic acid; EDTA, ethylenediaminetetraacetic

519 

acid; NTA, nitrilotriacetic acid. Ferric stability constants were obtained from the literature.40 

520 

Data are presented as means ± standard deviations of three independent experiments. Groups

521  522 

not sharing a common letter are significantly different (P < 0.05).

AMPH-stabilized

in vitro-formed nano-sized ferric hydrolysis

nano-sized

ferric

hydrolysis

26   

ACS Paragon Plus Environment

products;

BPDS,

Page 27 of 33

Journal of Agricultural and Food Chemistry

523 

Table 1. Details and results of the hemoglobin regeneration bioassay. Group*

Fe-deficient

FeSO4

ANPs

FC

FC+AMPH

Nano-α-Fe2O3

Fe-sufficient

N

8

8

8

8

8

8

4

Diet iron (mg/kg)

10.5±1.5

32.4±2.3

30.5±0.7

29.8±2.0

28.8±0.2

31.8±2.4

36.8±2.5

Diet intake (g)

265±30.5a

419.5±70.3c

332.1±32.7ab

268.2±43.7a

380.6±70.0bc

267.2±16.9a

383.9±37.9b

Baseline BW (g)

222.7±21.5

221.0±20.1

221.4±39.2

222.4±38.8

215.0±17.2

219.9±11.8

297.5±33.9

c

ab

ab

ab

85.0±3.6ab

*

BW gain (g)

a

62.4±27.8

101.7±32.9

Baseline Hb (g/L)

71.1±11.9

71.4±16.8



Hb change (g/L)

524  525  526  527 

-14.4±11.1

a

b

50±19.8

bc

109.7±37.2

79.7±25.9

88.7±28.9

73.7±10.5

70.5±15.3

71.9±12.7

b

a

b

52.8±19.8

-6.9±21.7

42.8±21.0

75.6±26.7

71.1±19.0

153.8±13.9 a

0.125±17.2

-25.3±23.8



FeSO4, ferrous sulfate; ANPs, in vitro-formed AMPH-stabilized nano-sized ferric hydrolysis products; FC, ferric citrate; FC+AMPH, the mixture of ferric citrate and anchovy muscle protein hydrolysate; Nano-α-Fe2O3, commercial bare α-Fe2O3 nanoparticles. Values are means ± SD. Means in a column without a common superscript letter (a, b and c) are statistically different, P < 0.05. *BW = body weight; Hb = hemoglobin. †

528 

27   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 33

529  530 

Table 2. Hemoglobin regeneration efficiencies and percentage contents of intestinal iron species in different groups.

531  532  533  534  535  536  537 

Percentage contents of intestinal iron species Hemoglobin Group regeneration Nano-sized Total soluble Low-molecular-weight 1 2 efficiency (%) iron iron3 iron FC 9.79±2.01a 43.18±4.49ef 12.44±3.47ab 30.74±6.49d Nano-α-Fe2O3 16.37±6.65abc 56.57±11.69fg 24.35±4.67cd 32.22±12.24de ef g bc 60.74±6.84 18.31±4.16 42.43±3.40ef FC+AMPH 40.33±6.36 40.88±7.67ef 59.01±9.02g 19.21±5.26bc 39.80±8.87def FeSO4 ANPs 56.25±11.35fg 61.24±11.01g 14.47±2.63ab 46.77±8.89f 1 2 3 Iron in the 0.22-µm filtrate. Iron in the 3-kDa filtrate. Calculated by subtraction of low-molecular-weight iron from total soluble iron. Abbreviations: FC, ferric citrate; Nano-α-Fe2O3, commercial bare α-Fe2O3 nanoparticles; FeSO4, ferrous sulfate; ANPs, in vitro-formed nano-sized ferric hydrolysis products stabilized by anchovy muscle protein hydrolysate; FC+AMPH, the mixture of ferric citrate and anchovy muscle protein hydrolysate. Values are means ± SD of a representative experiment (n = 8). Means in a column without a common superscript letter (a, b and c) are statistically different, P < 0.05. 

538 

28   

ACS Paragon Plus Environment

Page 29 of 33

539 

Journal of Agricultural and Food Chemistry

Figure 1

540 

 

541 

29   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

542 

Figure 2

543  544 

Page 30 of 33

   

 

30   

ACS Paragon Plus Environment

Page 31 of 33

545 

Journal of Agricultural and Food Chemistry

Figure 3

546 

 

547 

31   

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

548 

Page 32 of 33

Figure 4

549 

 

550 

32   

ACS Paragon Plus Environment

Page 33 of 33

551 

Journal of Agricultural and Food Chemistry

Table of Contents (TOC) Graphic

552 

 

553 

TOC. Enhancement of non-heme iron absorption in humans by fish possibly involves the

554 

formation of intestinal nano-sized iron, which is likely to be apically absorbed by means of

555 

divalent metal transporter-1 and endocytosis.

33   

ACS Paragon Plus Environment