You are on page 1of 81

International Journal of Biological Macromolecules

Effects of pectin structure on carotenoid bioaccessibility in simulated juice model


--Manuscript Draft--

Manuscript Number: IJBIOMAC-D-23-13059R1

Article Type: Research Paper

Section/Category: Carbohydrates, Natural Polyacids and Lignins

Keywords: β-carotene bioaccessibility; Pectic polymers; homogalacturonan

Abstract: The impact of pectin structure on carotenoid bioaccessibility is still uncertain. This
study aims to investigate how the different pectic polymers affected the bioaccessibility
of carotenoids in a simulated juice model during static in vitro digestion. This study
includes homogalacturonan (HG), which is a linear pectic polymer,
rhamnogalacturonan-I (RG-I), which is a branched pectic polymer, and
rhamnogalacturonan (RG), which is a diverse pectic polymer rich in RG-I,
rhamnogalacturonan-II (RG-II), and xylogalacturonan domains. Juice models without
pectin had the highest carotenoid bioaccessibility, suggesting pectin has negative
effects on carotenoid bioaccessibility. During the intestinal phase, systems with HG
showed the highest viscosity, followed by systems with RG and systems with RG-I.
Systems with RG-I had lower carotenoid bioaccessibility than systems with HG and
RG-II. Both the percentage of RG-I and the average side chain length of RG-I had
negative correlations with carotenoid bioaccessibility. RG-I side chains
with more arabinose and/or galactose might cause lower carotenoid bioaccessibility in
this juice model system.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Conflict of interest

Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:
Response to Reviewers

Response to Reviewer Comments

Manuscript Number: IJBIOMAC-D-23-13059

Title: Effects of pectin structure on carotenoid bioaccessibility in simulated juice

model

Journal: International Journal of Biological Macromolecules

Dear Prof. Corsaro,

Thank you for your time and effort in handling our manuscript. We would also like to

express our gratitude to the reviewers for their insightful comments on our manuscript.

Their feedback has been invaluable in improving the quality of our work. In response

to their comments, we have made revisions to the manuscript, and the changes have

been highlighted in red. Point-by-point responses to the reviewers’ comments are

listed below this letter.

We believe that the revisions have greatly strengthened the manuscript, and we are

confident that the revised version meets the journal's standards and will be

well-received by the readers. Thank you once again and we look forward to hearing

from you.

Best regards,

Jinfeng Bi
The responses to the reviewers’ comments are as follows:

Reviewer 1:

Comment 1.

Highlights: in the final highlight is it possible to indicate how RG-I affects

bioavailability? Positively/ negatively?

Response:

Thanks for your kind comments. The side chain length of rhamnogalacturonan-I

negatively affects carotenoid bioaccessibility. Due to the word limitation of each

bullet point, it was rewritten as: Longer side chains of rhamnogalacturonan-I reduce

carotenoid bioaccessibility.

Comment 2.

Abstract - can you be more clear on what your mean by RG? It is it not clear whether

this is the same as RG-II or different, try to be more clear in the nomenclature. Could

you suggest - how RG-I causes lower bioavailability in the abstract?

Response:

The RG was clarified in the Abstract (Line 28-29). This study includes

homogalacturonan (HG), which is a linear pectic polymer, rhamnogalacturonan-I

(RG-I), which is a branched pectic polymer, and rhamnogalacturonan (RG), which is

a diverse pectic polymer rich in RG-I, rhamnogalacturonan-II (RG-II), and

xylogalacturonan domains.

We have suggested how RG-I causes lower bioavailability in the Abstract (Line
35-37). RG-I side chains with more arabinose and/or galactose might cause lower

carotenoid bioaccessibility in this juice model system.

Comment 3.

Line 50 - what is the proportion of XGA?

Response:

The proportion of XGA has been added in the Introduction (Line 51-52).

Line 51-52: XGA is an HG substituted at O-3 with a β-linked xylose, and it is found

in pectin in relatively small proportions (<7%).

Comment 4.

Line 75 - could you add a reference?

Response:

References have been added in Line 78.

Comment 5.

Lines 88 - 89 - the pectins are from different sources, could this influence your

findings?

Response:

We acknowledge that pectin structure can vary depending on the sources. The aim of

our study is to investigate the different pectin subdomains on carotenoid

bioaccessibility. Different sources contribute to the different pectin subdomains,


which makes it possible for further exploration of their effects on carotenoid

bioaccessibility. If pectins are from the same sources, this may be challenging to

provide distinct pectin subdomains.

Comment 6.

Line 90 - do you know DE or DM?

Response:

This polygalacturonic acid product contains approximately 5% methylation according

to the description from Megazyme. We have added this information in the Materials

(Line 90).

Comment 7.

Line 102 - mixer rather than machine

Response:

“Machine” has been modified into “mixer” (Line 103).

Comment 8.

Line 109 - could you add a table with the sample information or add this to Table 1?

Response:

We have added a table with sample information (Table 1).

Comment 9.
Line 145 - 146 - number average molar mass (Mn).

Response:

This sentence has been modified according to your comment (Line 147).

Comment 10.

Line 161 - did you measure the zeta-potential of the pectin solutions?

Response:

Yes, we measured the zeta-potential of the pectin solutions with β-carotene and these

values are presented in Fig. 3A as initial zeta-potential values. We have added this in

the Materials & Methods (Line 163-164).

Comment 11.

Line 202 - can you estimate DE from the IR spectra? Are the pectins methylated,

acetylated or both?

Response:

Yes, we can estimate the degree of methyl-esterification (DM) from IR spectra, and

we have added this estimated DM values in Line 205-208.

Line 205-208: The ratio of peak area at 1745 cm-1 (related to methyl ester carbonyl

group stretching) to the sum of the peak areas at 1745 cm-1 and 1608 cm-1 (related to

carboxylate group stretching) was calculated as the degree of methyl-esterification

(DM) [24,25]. The calculated DM of HG and RG-I were 8% and 4%, respectively.
Comment 12.

Lines 188 - 189 - some repetition

Response:

We have modified the sentence according to your comment (Line 190-191).

Line 190-191: The experiments were conducted in triplicate, and each sample was

analyzed three times to ensure accuracy and reproducibility.

Comment 13.

Line 205 - is it that simple? Do different functional groups have the same sensitivity?

Could you add a reference?

Response:

Yes, we have added references in Line 205-208, and this method has been verified as

an effective means to estimate the degree of methyl-esterification (DM) of pectin.

Line 205-208: The ratio of peak area at 1745 cm-1 (related to ester carbonyl group

stretching) to the sum of the peak areas at 1745 cm-1 and 1608 cm-1 (related to

carboxylate group stretching) was calculated as DM [24,25].

[24] G.D. Manrique, F.M. Lajolo, FT-IR spectroscopy as a tool for measuring degree of methyl

esterification in pectins isolated from ripening papaya fruit. Postharvest Biol. Technol. 25(1) (2002)

99-107.

[25] C. Kyomugasho, S. Christiaens, A. Shpigelman, A.M. Van Loey, M.E. Hendrickx, FT-IR

spectroscopy, a reliable method for routine analysis of the degree of methylesterification of pectin in

different fruit-and vegetable-based matrices. Food Chem. 176 (2015) 82-90.


Comment 14.

Figure 2 - Could you redraw this using the conventional symbols for the

monosaccharides?

Response:

Please find the updated Fig. 2 with conventional symbols for the monosaccharides.

Comment 15.

Line 242 - should it be rhamnification?

Response:

Yes, it is. We have now corrected the word (Line 245).

Comment 16.

Line 243 - this is the third figure you have mentioned so it should be Figure 3.

Response:

We have modified this into Fig. 3 (Line 246).

Comment 17.

Line 245 - do you have any information about Rg and hence conformation from your

MALS data? This might help to explain some of your viscosity data.

Line 280 - this depends on conformation see my comment for line 245.

Response:

It’s a pity that we don’t have the data regarding Rg. In our forthcoming study, we will
conduct Rg measurements and establish a correlation with viscosity values.

Comment 18.

Line 255 - check your wording

Response:

We have modified this sentence (Line 258-259).

Line 258-259: Different physical indicators were estimated to investigate the effects

of pectic polymers on digestion process.

Comment 19.

Line 259 - what about the GalA content? This will impart negative charge.

Response:

We have added more discussion regarding GalA content (Line 264-266).

Line 264-266: Typically, an increased GalA content in pectin results in a higher

negative charge [33]. Additionally, more demethylesterified pectin can carry more

negative charges [34].

Comment 20.

Line 269 - "...pectin is close to zero.."

Response:

This sentence has been modified (Line 274).


Comment 21.

Figure 4 - could you redraw this as Figures 4A and 4B as it is not very clear as it is

drawn.

Response:

Yes, it is more clear to present the data in two separated figures. We have redrawn this

figure as Fig. 5A and Fig. 5B according to your comment.

Comment 22.

Line 320 - "... bioavailability compared to other systems..."

Response:

This sentence has been modified (Line 325).

Comment 23.

Line 321 - it is not clear whether this is good or bad, could you be more explicit.

Response:

We have added more explanation and make it more explicit (Line 325-327).

Line 325-327: SS had the highest β-carotene bioaccessibility compared to other

systems (SS-HG, SS-RG-I and SS-RG), which signified that the presence of pectin

could reduce the bioaccessibility carotenoids, thus limiting their potential health

effects.

Comment 24.
Line 342 - did you measure ferulic acid content or protein content?

Response:

We did not measure the ferulic acid content or protein content in this study. Therefore,

we have removed this statement from the manuscript.

Comment 25.

Line 354 - what are the units/ %?

Response:

We have added the unit (%) in Line 357.

Comment 26.

Lines 357 - 358 - "...helpful in increasing ..."

Line 370 - "... SS-RG-I had lower ..."

Response:

These sentences have been modified (Line 361 and 374).

Comment 27.

Line 373 - you might need to be careful with unique as the RG-I structure is an

average of a very broad range of polymers which are very different. For example, PDI

= 2.11, but this is only the heterogeneity in terms of molar mass, there are

heterogeneities in terms of composition and perhaps even conformation. Therefore, it

is perhaps not appropriate to call the structure "unique".


Response:

Yes, you are correct. We have removed the term "unique" from the entire manuscript.

Reviewer 2:

In this manuscript, three pectin-rich polysaccharides which were rich in HG, RG-I and

RG were used to analyze how the different pectic polymers affected the

bioaccessibility of carotenoids in a simulated juice model during static in vitro

digestion. Some issues were listed below.

Comment 1.

To detect how different pectic polymers affected the bioaccessibility of carotenoids, it

is suggested to purify and prepare pectin domains which are composed of one major

type of pectin, e.g. HG, RG-I, RG-II and XGA. In this study, the RG-I and RG

samples are both complex, which were not suitable for the study purpose.

Response:

Thanks for your kind comments. Complex pectin samples, which closely mimic the

natural composition of fruits and vegetables, can offer valuable insights into the

interactions and effects of multiple pectic polymers on carotenoid bioaccessibility.

Furthermore, the technical challenges and limited availability of pure RG-I and RG

samples make them less suitable for this study. As you suggested, the purification and

preparation of pure pectin subdomains might be our further study.

Comment 2.
If RG-II domain also existed in RG-I (potato)? Based on monosaccharide composition,

this pectin seemed to contain more HG than RG-I.

Response:

According to the monosaccharide composition (70.9% GalA, 19.8% Gal, 6.3%Rha,

1.8% Ara, 1.05% Xyl and 0.14% Fuc), RG-II domain seems not existed or existed in a

very small amount in RG-I rich pectin polymers.

The structural framework of RG-I is composed of repeating disaccharides of GalA

and rhamnose. By analyzing the sugar ratios, it has been determined that the RG-I rich

pectin polymer constitutes 34% of RG-I, whereas the HG rich pectin polymer

contains only 7% RG-I. Therefore, in our study, the RG-I derived from potatoes can

be classified as a pectin polymer with a relatively higher abundance of RG-I.

Comment 3.

How to prove RG-II domain were present in RG (soybean)?

Response:

As fucose exist in the pectin polymers, so it is assumed the RG-II domain is present in

RG (soybean).

Comment 4.

In line 228, how to determine the ratios of different types of pectin?

Response:

The composition ratios were calculated based on the mol% quantifiable of GalA and
neutral sugars as follows:
2×Rha+Ara+Gal
The percentage of RG-I (% RG-I) = GalA+Rha+Ara+Gal+Xyl+Fuc

The percentage of XGA is calculated as the mol% of Xyl, and the percentage of RG-II

is estimated as the mol% of the remaining monosaccharides.


Abstract

Abstract

The impact of pectin structure on carotenoid bioaccessibility is still uncertain. This

study aims to investigate how the different pectic polymers affected the

bioaccessibility of carotenoids in a simulated juice model during static in vitro

digestion. This study includes homogalacturonan (HG), which is a linear pectic

polymer, rhamnogalacturonan-I (RG-I), which is a branched pectic polymer, and

rhamnogalacturonan (RG), which is a diverse pectic polymer rich in RG-I,

rhamnogalacturonan-II (RG-II), and xylogalacturonan domains. Juice models without

pectin had the highest carotenoid bioaccessibility, suggesting pectin has negative

effects on carotenoid bioaccessibility. During the intestinal phase, systems with HG

showed the highest viscosity, followed by systems with RG and systems with RG-I.

Systems with RG-I had lower carotenoid bioaccessibility than systems with HG and

RG-II. Both the percentage of RG-I and the average side chain length of RG-I had

negative correlations with carotenoid bioaccessibility. RG-I side chains with more

arabinose and/or galactose might cause lower carotenoid bioaccessibility in this juice

model system.

Keywords: β-carotene bioaccessibility; Pectic polymers; Homogalacturonan;


Rhamnogalacturonan-I; Rhamnogalacturonan-II; Digestive fluids
Revised manuscript (clean version) Click here to view linked References

1 Effects of pectin structure on carotenoid bioaccessibility in


2 simulated juice model
3

4 Jianing Liua,b, Jinfeng Bia,*, Xuan Liua,*, Dazhi Liua,c, Vincenzo Foglianob, Matthijs

5 Dekkerb, Ruud Verkerkb

a
7 Institute of Food Science and Technology, Chinese Academy of Agricultural Sciences

8 (CAAS), Key Laboratory of Agro-Products Processing, Ministry of Agriculture and

9 Rural Affairs, Beijing 100193, China

b
10 Food Quality and Design Group, Wageningen University & Research, Bornse

11 Weilanden 9, 6708 WG, Wageningen, the Netherlands

c
12 Laboratory of Food Chemistry, Wageningen University & Research, Bornse

13 Weilanden 9, 6708 WG, Wageningen, the Netherlands

14

15

16

*
17 Corresponding author.

18 Tel.:+86-1062812584.

19 E-mail address: bijinfeng2010@163.com and liuxuancaas@126.com.

20 ADD: No. 2, Yuanmingyuan West Road, Haidian District, Beijing, 100193, China.

21

1
22 Abstract

23 The impact of pectin structure on carotenoid bioaccessibility is still uncertain. This

24 study aims to investigate how the different pectic polymers affected the bioaccessibility

25 of carotenoids in a simulated juice model during static in vitro digestion. This study

26 includes homogalacturonan (HG), which is a linear pectic polymer,

27 rhamnogalacturonan-I (RG-I), which is a branched pectic polymer, and

28 rhamnogalacturonan (RG), which is a diverse pectic polymer rich in RG-I,

29 rhamnogalacturonan-II (RG-II), and xylogalacturonan domains. Juice models without

30 pectin had the highest carotenoid bioaccessibility, suggesting pectin has negative

31 effects on carotenoid bioaccessibility. During the intestinal phase, systems with HG

32 showed the highest viscosity, followed by systems with RG and systems with RG-I.

33 Systems with RG-I had lower carotenoid bioaccessibility than systems with HG and

34 RG-II. Both the percentage of RG-I and the average side chain length of RG-I had

35 negative correlations with carotenoid bioaccessibility. RG-I side chains with more

36 arabinose and/or galactose might cause lower carotenoid bioaccessibility in this juice

37 model system.

38 Keywords: β-carotene bioaccessibility; Pectic polymers; Homogalacturonan;

39 Rhamnogalacturonan-I; Rhamnogalacturonan-II; Digestive fluids

2
40 Introduction

41 Pectin is a polysaccharide found in plant cell walls, predominantly consisting of

42 1,4-linked-α-D-galacturonic acid units. The primary structural domains of pectin are

43 homogalacturonan (HG), rhamnogalacturonan-I (RG-I), rhamnogalacturonan-II (RG-

44 II), and xylogalacturonan (XGA). HG makes up approximately 65% of the pectin

45 molecule, and it contains a linear chain of galacturonic acid (GalA) with methyl-

46 esterified carboxyl groups. RG-I is the second most abundant pectic polymer (i.e. 20-

47 35% of pectins), which comprises a backbone of repeating GalA and rhamnose

48 disaccharides with arabinan, galactan, and/or arabinogalactan side chains. The most

49 structurally complex pectin, RG-II, makes up about 10% of pectin. The backbone of

50 RG-II is made up of HG, with intricate side chains including simple sugars that are

51 connected to the GalA. XGA is an HG substituted at O-3 with a β-linked xylose, and it

52 is found in pectin in relatively small proportions (<7%) [1,2].

53 Different structural characteristics of pectic polymers contribute to their different

54 functional properties and bioactivities [3,4]. Many of pectin's functional characteristics

55 are influenced by the quantity of esterified GalA residues and their distribution in HG.

56 When methyl-ester groups are removed from HG, it becomes calcium cross-linkable,

57 allowing the formation of supramolecular assemblies and gels [5]. Furthermore, earlier

58 research has shown that cooperative cation binding to pectin requires at least 6 to 10

59 consecutive non-methyl-esterified GalA units [6,7]. RG-I pectin is heterogeneous due

60 to its variable glycosyl linkages, monosaccharide composition of side chains, chain

61 lengths, and distribution. Ferulic acid groups alter the molecular weight in the cell wall

3
62 and lead to poor pectin gelation properties by promoting the cross-linking of neutral

63 sugars in RG-I through ferulic bridges [8]. Recently, the research focused on the

64 importance of the pectin hairy region. Pectin RG-I portion showed improved

65 bioactivities, including inhibition of cell migration, and immunological and prebiotic

66 activities [9,10]. In vitro tests with human colon adenocarcinoma (Caco-2) cells showed

67 that RG-II had promisingly strong anti-cancer proliferation activities [11].

68 During the upper gastrointestinal stage, pectin may affect the digestion and

69 absorption of lipophilic carotenoids, by inhibiting lipid digestion [12]. Compared with

70 high pectin concentration, low pectin concentration favored the binding of bile salt and

71 carotenoid micellization [13]. The degree of lipid digestion decreased with an

72 increasing degree of pectin methyl-esterification [14]. Several studies have reported

73 carotenoid bioaccessibility and pectin structural and functional characteristics

74 (galacturonic acid concentration, degree of methyl-esterification and acetylation, and

75 emulsifying activity) of juices are related [15,16], however, the molecular mechanism

76 underpinning this relationship is still uncertain. It is possible that specific pectic

77 polymers are actually responsible for the reduction of carotenoid bioaccessibility

78 [12,16]. Therefore, the aim of this study was to investigate the influences of different

79 pectic polymers on carotenoid bioaccessibility of simulated juice models using static in

80 vitro digestion. In the present study, the pectic polymers include HG domain-rich pectic

81 polymer, RG-I region-rich pectic polymer, and rhamnogalacturonan (RG) pectic

82 polymer, which is more diverse with RG-I, RG-II and XGA domains. The simulated

83 juice models were applied to directly understand how pectin polymers influences β-

4
84 carotene digestion. Viscosity and ζ-potential of digestive fluids, β-carotene retention

85 and β-carotene bioaccessibility were recorded, to gain better knowledge of the digestive

86 process.

87

88 1. Materials and methods

89 2.1 Materials

90 Polygalacturonic acid (citrus pectin, ~5% methylation), rhamnogalacturonan I

91 (potato), and rhamnogalacturonan (soybean) were purchased from Megazyme (Bray,

92 Ireland). Polygalacturonic acid is seen as an HG domain-rich pectic polymer.

93 Rhamnogalacturonan I is considered to be an RG-I region-rich pectic polymer.

94 Rhamnogalacturonan is considered a more diverse pectic polymer rich in RG-I, RG-II

95 and XGA domains. Porcine gastric mucin, pepsin from porcine gastric mucosa, lipase

96 from porcine pancreas, β-carotene (≥ 93%), β-carotene (HPLC grade) and butylated

97 hydroxytoluene (BHT) were purchased from Sigma-Aldrich (St. Louis, MO, USA).

98

99 2.2 Methods

100 2.2.1 Juice model system preparation

101 The juice model system was prepared according to our previous study, with minor

102 modifications [17]. β-carotene (30 mg) in 100 mL acetic acid-sodium acetate buffer (0.2

103 M, pH=6) with 0.2% corn oil (w/w) was dispersed using a high-speed dispersion mixer

104 (T25 Ultra-TURRAX, IKA, Germany) at 14,000 rpm for 40 min. Subsequently, HG

105 (polygalacturonic acid), RG-I (rhamnogalacturonan I), and RG (rhamnogalacturonan),

5
106 at a proportion of 1% (w/v), were added to the β-carotene solution, and fully mixed

107 using the Ultra-TURRAX at 7,000 rpm for 10 min. The systems were then adjusted to

108 pH 6 and homogenized at 250 MPa (SPCH-EP-IC-16-30, Homogenising Systems Ltd,

109 London, UK). Simulated systems containing β-carotene, β-carotene with HG, RG-I, or

110 RG were abbreviated as SS, SS-HG, SS-RG-I, and SS-RG, respectively (Table 1).

111

112 2.2.2 Structural characterization of pectic polymers

113 2.2.2.1 Fourier transform infrared spectroscopy (FT-IR) analysis

114 The FT-IR spectra of pectic polymers were determined using an FT-IR

115 spectrometer (Tensor 27, Bruker, Germany) in the wavenumber region 4000~400 cm-1

116 at a resolution of 4 cm-1 and a cumulative scan of 64. Before FTIR analysis, pectic

117 polymers were pulverized and pressed into flake with KBr at a ratio of 1:100 (w/w).

118

119 2.2.2.2 Determination of galacturonic acid (GalA) and neutral sugar concentration

120 GalA and neutral sugar concentrations were performed according to the previous

121 study [16]. High-performance anion-exchange chromatography (HPAEC) was achieved

122 using ICS-3000 system (Dionex, Sunnyvale, CA, USA) equipped with a CarboPacTM

123 PA20 column and a pulsed amperometric detector. Monosaccharides including GalA,

124 galactose (Gal), arabinose (Ara), rhamnose (Rha), xylose (Xyl), and fucose (Fuc) were

125 used for identification and quantification.

126 To obtain information on the domains of the pectic polymers, five composition

127 ratios were calculated based on the mol% quantifiable of GalA and neutral sugars, with

6
128 some modifications [18,19]:
GalA+Rha
129 The linearity of pectic polymers = (1)
GalA+Rha+Ara+Gal+Xyl+Fuc

GalA−Rha
130 The percentage of HG (% HG) = (2)
GalA+Rha+Ara+Gal+Xyl+Fuc

GalA−Rha
131 The ratio of HG to RG-I backbone = (3)
2×Rha

2×Rha+Ara+Gal
132 The percentage of RG-I (% RG-I) = (4)
GalA+Rha+Ara+Gal+Xyl+Fuc

Ara+Gal
133 The average side chain length of RG-I = (5)
Rha

134

135 2.2.2.3 Determination of molar mass distribution

136 The molar mass distributions of pectins were determined using high-performance

137 size exclusion chromatography (HPSEC) coupled with multiangle laser light scattering

138 (MALLS) (Dawn Heleos II, Wyatt Technology, Santa Barbara, CA, USA), reflective

139 index (RI) detector (Optilab rEX, Wyatt Technology, Santa Barbara, CA, USA) and

140 ultraviolet (UV) detector (L-2400, Hitachi, Tokyo, Japan). The dn/dc value (refractive

141 index increment) was 0.146 mL/g. Pectins (10 mg) were dissolved in eluent (0.1 M

142 NaCl) overnight and filtrated through a 0.45 μm filter membrane before analysis.

143 Samples (100 μL) were injected into an OHpak SB-806M HQ column (8.0 × 300 mm,

144 Shodex, Tokyo, Japan). Elution was performed with 0.1 M NaCl solution at a flow rate

145 of 0.5 mL/min for 30 min and the column was kept at 35 °C. The polydispersity index

146 (PDI) was calculated as the ratio of the weight-average molar mass (Mw) to the number

147 average molar mass (Mn).

148

149 2.2.3 Static in vitro digestion

7
150 Juice model systems were subjected to an in vitro digestion that encompassed

151 simulated oral, gastric, and intestinal phases [20]. During intestine digestion stage, an

152 automatic titration facility (Metrohm, USA, Inc.) was utilized to maintain pH at 7.0 by

153 adding 0.25 M NaOH solution over 2 h at 37°C. To stop the enzymatic reaction, samples

154 were placed on ice after each digestion phase.

155

156 2.2.4 Digestive fluid characteristics

157 2.2.4.1 Viscosity property of the digestive fluids

158 The viscosity of the digestive fluids at a shear rate of 50 s-1 was measured by using

159 a Physica MCR 301 rheometer (Anton Paar, Graz, Austria). The parallel plate geometry

160 had a 40 mm diameter and the gap size was 1 mm.

161

162 2.2.4.2 ζ-potential of the digestive fluids

163 The ζ-potential values of the initial systems (pectin solutions with β-carotene) and

164 their digestive fluids during digestion were measured by using a Zetasizer Nano ZS

165 analyzer (Malvern Instruments Ltd., Malvern, UK).

166

167 2.2.5 β-carotene bioaccessibility and retention ratio

168 After the above digestion process, the raw intestinal digesta was centrifuged at

169 8000 rpm (8801×g) and 4 °C for 1 h. β-carotene bioaccessibility was calculated as

170 follows:

171 β-carotene bioaccessibility (%)=(Cmicelle/Craw digesta) ×100 (6)

8
172 where Cmicelle and Craw digesta are the contents of β-carotene in the micelle fraction and

173 the raw digesta, respectively.

174 The β-carotene retention ratio during each digestion phase was calculated as

175 follows:

176 β-carotene retention ratio (%)=Cdigestion/Cinitial×100 (7)

177 where Cdigestion and Cinitial are the β-carotene concentrations (μg/mL) in different systems

178 during each digestion phase and in the initial system, respectively.

179

180 2.2.6 Extraction and determination of β-carotene

181 A previous procedure was used to extract and quantify β-carotene [21]. Samples

182 were separated and analyzed on high-performance liquid chromatography (HPLC)

183 system with a reversed-phase C30-column (250 mm × 4.6 mm, 5 μm; YMC Europe,

184 Dinslaken, Germany), a binary HPLC pump (1525, Waters, Milford, MA, USA) and a

185 photodiode array detector (2998, Waters, Milford, MA, USA). The gradient program,

186 detection wavelength, flow rate of mobile phase were performed according to a

187 previous procedure [16].

188

189 2.2.7 Statistics

190 The experiments were conducted in triplicate, and each sample was analyzed three

191 times to ensure accuracy and reproducibility. All data were presented as mean ±

192 standard deviation. Analysis of variance (ANOVA) followed by a Tukey test was used

193 to determine significant differences using R software (Version 4.2.0). The Pearson

9
194 correlation analysis was carried out by the “corrplot” package in R software.

195

196 2. Results and discussion

197 2.1 Structural characterization of pectin polymers

198 3.1.1 Functional groups of pectic polymers

199 As shown in Fig. 1A, the following were the main characteristics of the absorption

200 peaks: (i) The absorption peak at ≈ 3364 cm-1 was attributed to O-H stretching vibration,

201 and the peak at ≈ 2936 cm-1 was caused by C-H stretching of CH2 groups; (ii) the peaks

202 at ≈ 1745 cm-1 and 1608 cm-1 corresponded to the C=O stretching vibration of methyl-

203 esterified carboxyl groups and C=O asymmetrical stretching vibration of free carboxyl,

204 respectively; (iii) the peak at ≈ 1418 cm-1 was ascribed to a symmetric carboxylate

205 stretch [22,23]. The ratio of peak area at 1745 cm-1 (related to methyl ester carbonyl

206 group stretching) to the sum of the peak areas at 1745 cm-1 and 1608 cm-1 (related to

207 carboxylate group stretching) was calculated as the degree of methyl-esterification (DM)

208 [24,25]. The calculated DM of HG and RG-I were 8% and 4%, respectively. However,

209 the absorption area of RG at 1745 cm-1 was not detected, which suggested that methyl-

210 esters were not present in the RG chain. The variation of the DM observed could be

211 attributed to the different extraction methods applied.

212

213 3.1.2 Monosaccharide content and composition ratios of pectic polymers

214 As shown in Fig. 1B, the GalA content of HG was the highest (94.1%), followed

215 by RG-I (70.9%) and RG (50.9%). HG is composed of 94% GalA and 6% other

10
216 monosaccharides, such as 3.9% Gal and 1.5% Rha. RG-I comprises 70.9% GalA,

217 followed by 19.8% Gal, 6.3% Rha, and 1.8% Ara that constitute the RG-I region. RG

218 is a more diverse pectic polymer compared to HG and RG-I, as evidenced by its

219 composition of 50.9% GalA, 16.9% Xyl, and 13.1% Fuc. Individual monosaccharide

220 contents are not sufficient to gain insight into the pectin chains. With the help of the

221 ratios between the composing monosaccharides, the overall structure of pectin chains

222 becomes more visible.

223 As shown in Table 2, compared to the other two pectic polymers, HG is composed

224 of more linear/less branched pectin evidenced by the higher linearity of pectic

225 polymers, % HG and the ratio of HG to RG-I backbone. The higher linearity of HG

226 makes it potential to have higher Ca2+-binding capability, due to the higher possibility

227 of consecutive non-methyl-esterified GalA residues. RG-I is composed of a more

228 branched RG-I domain according to higher % RG-I and average side chain length of

229 RG-I. Pectin polymers rich in RG-I side chains have better emulsion stability by

230 forming a hydrated layer [26]. RG is a pectic polymer with more diverse domains,

231 which consists of 26.87% RG-I and approximately 13.14% RG-II side chains and 16.94%

232 XGA side chains. The percentage of the RG-II and XGA backbones is unclear, though,

233 as both of their backbones are HG. Based on the obtained structural characteristics, the

234 structures of the pectic polymers are visualized as hypothetical model structures

235 presented in Fig. 2. These distinct monosaccharide compositions among different pectic

236 polymers make it possible for further investigation of the effects of pectic polymers on

237 carotenoid bioaccessibility.

11
238

239 3.1.3 Molar mass distribution of pectic polymers

240 Mw and PDI values are presented in Table 2, and a higher PDI represents a wider

241 molecular mass distribution. RG showed the highest Mw, followed by RG-I and HG.

242 HG showed the lowest PDI values compared to RG and RG-I. The rich side chains of

243 RG-I and RG might contribute to their higher Mw and wider molecular mass distribution.

244 According to correlation analysis, the linearity of pectin had a negative correlation with

245 Mw, indicating that a higher degree of rhamnification might correspond to a larger Mw

246 of the polysaccharide (Fig. 3). Many pectin properties are linked to its Mw, and high

247 molecular mass polymers generally contribute to promising functionalities, such as

248 emulsifying and emulsion stabilizing abilities [27]. In most cases, higher Mw

249 macromolecules obstruct fluid motion, resulting in higher viscosity solutions [28]. It

250 should be noted that other factors, such as pectin sources and pectin fractions, might

251 also affect the viscosity of pectins [29,30]. Decreasing Mw might reduce the maximum

252 amount of Ca2+ that pectin can bind [31], however, the calcium binding ability also

253 depends on the degree and distribution pattern of methyl-ester groups among HG region

254 [32].

255

256 3.2 Physical characteristics of digestive fluids

257 3.2.1 ζ-potential of digestive fluids

258 Different physical indicators were estimated to investigate the effects of pectic

259 polymers on digestion process. Generally, changes in the ζ-potential of particles can

12
260 influence their electrostatic repulsion or attraction, which in turn can affect the stability

261 and behavior of the systems during digestion. As seen in Fig. 4A, initial ζ-potential

262 values of SS-HG and SS-RG were lower than those of SS and SS-RG-I, which

263 suggested that SS-HG and SS-RG were enveloped by more negatively charged

264 substances. Typically, an increased GalA content in pectin results in a higher negative

265 charge [33]. Additionally, more demethylesterified pectin can carry more negative

266 charges [34]. During the oral phase, the ζ-potential value of SS was the highest,

267 followed by SS-RG, SS-RG-I and SS-HG. The ζ-potential values increased from the

268 initial phase to the gastric phase, while decreased in the intestinal phase. The highly

269 acidic conditions, which cause the protein-coated droplets to be positive, and the

270 presence of anionic mucin, which binds to the cationic protein surfaces, are both related

271 to the increase in ζ-potential during the gastric phase [35,36]. Specifically, GalA, the

272 degree of methyl-esterification and the distribution of methyl esters of pectin influence

273 their interactions with protein [37]. Moreover, the pectin surface charge decreased with

274 the decreasing pH, and at pH 2, the ζ-potential of pectin is close to zero [38]. The

275 amount of the negative charge on the backbone of pectin decreased as a result of

276 protonation (a drop in pH). The fact that the ζ-potential values changed from their initial

277 positive values to more negative ones during the intestinal phase could be attributed to

278 a number of negatively charged species, including free fatty acids released during

279 lipolysis, intricate colloidal structures, undigested oil, and anionic polymers [14].

280

281 3.2.2 Apparent viscosity of digestive fluids

13
282 The apparent viscosity of digestive fluids is given in Fig. 4B. The physical feature

283 of apparent viscosity is linked to soluble dietary fibers, particularly pectic

284 polysaccharides. Initial SS-HG and SS-RG showed higher viscosity than SS-RG-I.

285 Generally, pectin with a higher Mw has a higher viscosity [28]. Moreover, HG pectin

286 contributed greatly to the viscosity of pectin [39]. There was a significant negative

287 correlation between the initial viscosity of the systems and the RG-I percentage (Fig.

288 3). Due to the neutral sugar side chains, solutions containing high contents of RG-I

289 and/or RG-II pectin usually have lower viscosity values than solutions with high GalA

290 content [39]. From the initial phase to the gastric phase, the viscosity of the four systems

291 all decreased, which might be caused by the dilution effect of digestive fluids. These

292 results are consistent with those obtained with broccoli-digested fractions, where

293 viscosity decreased from the gastric phase to the intestinal phase [40]. During the

294 intestinal phase, SS-HG showed the highest viscosity, followed by SS-RG, SS-RG-I

295 and SS. A negative correlation has been found between the HG to RG-I ratio and the

296 viscosity of the systems during the intestinal phase. The interactions between pectic

297 polymers and the digestive fluids might also contribute to the results, especially the

298 crosslinking between pectin and digestive enzymes.

299

300 3.3 β-carotene retention ratio during digestion phases

301 As seen in Fig. 5, in the oral phase, the β-carotene retention ratio of SS-RG was

302 lower than that of SS, while in the gastric phase, the β-carotene retention ratio of SS-

303 RG was the highest compared with other systems. This suggested that SS-RG might be

14
304 more stable under acidic conditions. At the intestinal phase, the β-carotene retention

305 ratios of SS-HG, SS-RG-I, and SS-RG had no significant difference. Carotenoids are

306 isoprenoid pigments with a polyene backbone that contains a variable number of

307 conjugated double bonds, therefore, they are prone to be oxidized. A previous study

308 showed that the stability of carotenoids at different processing can be significantly

309 improved in an oil-based system, compared with a water-based juice system [41]. In

310 our study, the carotenoids were in a homogenized juice model system, which is similar

311 to a water-based juice system, therefore, carotenoid stability might be affected by high-

312 pressure homogenization to some extent. A negative correlation was observed between

313 the β-carotene retention ratio during the intestinal phase and the ζ-potential values of

314 initial systems. This suggests that a higher negative charge could potentially enhance

315 the delivery of carotenoids during the intestinal stage. Additionally, a negative

316 correlation was found between the β-carotene retention ratio and the viscosity of

317 digestive fluids during the intestinal phase. Therefore, the pectin-induced viscosity

318 might have effects on the β-carotene stability during the intestinal stage.

319

320 3.4 β-carotene bioaccessibility

321 Carotenoids need to be released from the food matrix, solubilized into the lipid

322 phase, and incorporated into mixed micelles in order to be absorbed via the human

323 intestinal epithelium [12]. The levels of carotenoids available for intestinal absorption,

324 specifically the carotenoids in mixed micelles, are referred to as bioaccessible

15
325 carotenoids. SS had the highest β-carotene bioaccessibility compared to other systems

326 (SS-HG, SS-RG-I and SS-RG), which signified that the presence of pectin could reduce

327 the bioaccessibility carotenoids, thus limiting their potential health effects. Similar

328 observations were also found by Cervantes-Paz et.al [13], in which high pectin

329 concentration decreased carotenoid micellization. Pectin can increase the viscosity of

330 digestive fluids, allowing digesta to last longer in the gastrointestinal phase while

331 simultaneously reducing the movement of substrates and digestive enzymes (eg., pepsin

332 and lipase), decreasing lipase activity and micelle generation [42]. Moreover, different

333 pectic polymers caused different influences on carotenoid bioaccessibility. SS-HG and

334 SS-RG had similar β-carotene bioaccessibility, while SS-RG-I had a smaller β-carotene

335 bioaccessibility than those of SS-HG and SS-RG. Furthermore, both % RG-I and the

336 average side chain length of RG-I had negative correlations with carotenoid

337 bioaccessibility. Therefore, it was deduced that more RG-I side chains with Ara and/or

338 Gal might cause lower carotenoid bioaccessibility in this juice model system. These

339 results are consistent with those obtained with carrot juice, where higher emulsifying

340 properties caused a lower value of total carotenoid bioaccessibility [16]. The proposed

341 mechanisms of the effects of pectic polymers on carotenoid bioaccessibility are shown

342 in Fig. 6. The lipid hydrolysis produces monoacylglycerols and free fatty acids, which

343 together with bile salts contribute to the formation of mixed micelles. In this context,

344 lipase can greatly increase lipolysis and carotenoid bioaccessibility [43]. Pectin

345 segments with enriched RG-I/neutral sugar side chains contribute to better emulsifying

346 properties [44,45]. Pectin with better emulsification properties might be able to form

16
347 thicker hydrated layers at the oil-water interface and inhibit lipase to the surface,

348 resulting in low carotenoid bioaccessibility. A recent study based on molecular

349 dynamics simulations also revealed that RG-I and RG-II domains were adsorbed on the

350 oil-water interface in a folded form, creating a thick viscoelastic film while the HG

351 domain was tiled on the interface [46].

352 It should also be mentioned that in real fruit and vegetable systems, RG-I side

353 chains could also be cross-linked to xylans, xyloglucans, lignins, and proteins [5]. The

354 above-mentioned cross-linking among pectic polymers can also influence carotenoid

355 bioaccessibility, but this is not considered in the juice model system in this study.

356 Despite the absence of explicit localization obstacles of carotenoids in model systems,

357 β-carotene bioaccessibility was still comparatively low, ranging from 7.97% to 14.55%.

358 It implies that lipids are important in enhancing β-carotene bioaccessibility since dietary

359 lipids could stimulate bile salt secretion and increase the number of carotenoids carried

360 in micelles. Furthermore, other consumption forms of food products might be helpful

361 in increasing carotenoid bioaccessibility. Making fruit and vegetable-based emulsions

362 has been noted as a promising method to increase carotenoid bioaccessibility by

363 researchers [47,48]. Therefore, further research might focus on the effects of pectic

364 polymers on carotenoid bioaccessibility in model systems with the addition of more

365 lipids.

366

367 4. Conclusions

368 The simulated juice models were helpful to understand the effects of different

17
369 pectic polymers on carotenoid bioaccessibility. The ζ-potential values of all the systems

370 increased from the initial phase to the gastric phase, while decreased in the intestinal

371 phase. When compared to the initial ζ-potential values, ζ-potential became more

372 negative during the intestinal phase. During the intestinal phase, SS-HG showed the

373 highest viscosity, followed by SS-RG, SS-RG-I and SS. SS-HG and SS-RG had similar

374 β-carotene bioaccessibility, while SS-RG-I had a lower β-carotene bioaccessibility than

375 those of SS-HG and SS-RG. Both % RG-I and the average side chain length of RG-I

376 had negative influences on carotenoid bioaccessibility.

377

378 Acknowledgements

379 This work was supported by the National Key Research and Development

380 Program of China (2022YFD1600505), the earmarked fund for China Agriculture

381 Research System (CARS-27), and the Joint PhD Program between Wageningen

382 University & Research (WUR) and Chinese Academy of Agricultural Sciences (CAAS)

383 (No. MOE11NL1A20151701N).

18
384 References

385 [1] D. Mohnen, Pectin structure and biosynthesis, Curr. Opin. Plant Biol. 11(3) (2008)

386 266-277.

387 [2] J. Zandleven, S.O. Sørensen, J. Harholt, G. Beldman, H.A. Schols, H.V. Scheller, &

388 A.J. Voragen, Xylogalacturonan exists in cell walls from various tissues of

389 Arabidopsis thaliana, Phytochemistry. 68(8) (2007) 1219-1226.

390 [3] B. Neckebroeck, S.H.E. Verkempinck, G. Vaes, K. Wouters, J. Magnee, M. E.

391 Hendrickx, et al. Advanced insight into the emulsifying and emulsion stabilizing

392 capacity of carrot pectin subdomains, Food Hydrocolloids, 102 (2020) 105594.

393 [4] P. M. Pieczywek, J. Ciesla, W. Płazinski, A. Zdunek, Aggregation and weak gel

394 formation by pectic polysaccharide homogalacturonan, Carbohydr. Polym. 256

395 (2021) 117566.

396 [5] W.G.T. Willats, L. Mccartney, W. Mackie, & J.P. Knox, Pectin: Cell biology and

397 prospects for functional analysis, Plant Mol. Biol. 47(1-2) (2001) 9-27.

398 [6] P.J.H. Daas, B. Boxma, A.M.C.P. Hopman, A.G.J. Voragen, H.A. Schols,

399 Nonesterified galacturonic acid sequence homology of pectins, Biopolymers. 58(1)

400 (2001) 1-8.

401 [7] G.A. Luzio, R.G. Cameron, Demethylation of a model homogalacturonan with the

402 salt-independent pectin methylesterase from citrus: Part II. structure–function

403 analysis, Carbohydr. Polym. 71(2) (2008) 300-309.

404 [8] A. Oosterveld, I.E. Pol, G. Beldman, A.G. Voragen, Isolation of feruloylated

405 arabinans and rhamnogalacturonans from sugar beet pulp and their gel forming

19
406 ability by oxidative cross-linking, Carbohydr. Polym. 44(1) (2001) 9-17.

407 [9] D. Wu, J. Zheng, G. Mao, W. Hu, X. Ye, R.J. Linhardt, et al. Rethinking the impact

408 of RG-I mainly from fruits and vegetables on dietary health, Crit. Rev. Food Sci.

409 Nutr. 60 (2020) 2938-2960.

410 [10] M.Y. Zhang, J. Cai, Preparation of branched RG-I-rich pectin from red dragon fruit

411 peel and the characterization of its probiotic properties, Carbohydr. Polym. (2023)

412 120144.

413 [11] L. Ai, Y. Chung, S. Lin, K. Lee, P. Lai, Y. Xia, et al. Active pectin fragments of

414 high in vitro antiproliferation activities toward human colon adenocarcinoma cells:

415 Rhamnogalacturonan II, Food Hydrocolloids. 83 (2018) 239-245.

416 [12] B. Cervantes-Paz, J.J. Ornelas-Paz, S. Ruiz-Cruz, C. Rios-Velasco, V. Ibarra-

417 Junquera, E.M. Yahia, et al. Effects of pectin on lipid digestion and possible

418 implications for carotenoid bioavailability during pre-absorptive stages: A review,

419 Food Res. Int. 99 (2017) 917-927.

420 [13] B. Cervantes-Paz, J.D.J. Ornelas-Paz, J.D. Pérez-Martínez, J. Reyes-Hernández,

421 P.B. Zamudio-Flores, C. Rios-Velasco, et al. Effect of pectin concentration and

422 properties on digestive events involved on micellarization of free and esterified

423 carotenoids, Food Hydrocolloids. 60 (2016) 580-588.

424 [14] M. Espinal-Ruiz, L.P. Restrepo-Sanchez, C.E. Narvaez-Cuenca, D.J. McClements,

425 Impact of pectin properties on lipid digestion under simulated gastrointestinal

426 conditions: Comparison of citrus and banana passion fruit (Passiflora tripartita var.

427 mollissima) pectins, Food Hydrocolloids. 52 (2016) 329-342.

20
428 [15] Y. Ding, X. Liu, J. Bi, X. Wu, X. Li, J. Liu, et al. Effects of pectin, sugar and pH

429 on the β-Carotene bioaccessibility in simulated juice systems, LWT--Food Sci.

430 Technol. 124 (2020) 109125.

431 [16] X. Liu, J. Liu, J. Bi, J. Yi, J. Peng, C. Ning, et al. Effects of high pressure

432 homogenization on pectin structural characteristics and carotenoid bioaccessibility

433 of carrot juice, Carbohydr. Polym. 203 (2019) 176-184.

434 [17] J. Liu, J. Bi, X. Liu, D. Liu, X. Wu, J. Lyu, et al. Effects of pectins and sugars on

435 β-carotene bioaccessibility in an in vitro simulated digestion model, J. Food

436 Compos. Anal. 91 (2020) 103537.

437 [18] S.E. Broxterman, P. Picouet, H.A. Schols, Acetylated pectins in raw and heat

438 processed carrots, Carbohydr. Polym. 177 (2017) 58-66.

439 [19] K. Houben, R.P. Jolie, I. Fraeye, A.M. Van Loey, M.E. Hendrickx, Comparative

440 study of the cell wall composition of broccoli, carrot, and tomato: Structural

441 characterization of the extractable pectins and hemicelluloses, Carbohydr. Res.

442 346(9) (2011) 1105-1111.

443 [20] J. Liu, D. Liu, J. Bi, X. Liu, Y. Lyu, R. Verkerk, et al. Micelle separation conditions

444 based on particle size strongly affect carotenoid bioaccessibility assessment from

445 juices after in vitro digestion, Food Res. Int. 151 (2022) 110891.

446 [21] G. Knockaert, L. Lemmens, S. Van Buggenhout, M. Hendrickx, A. Van Loey,

447 Changes in β-carotene bioaccessibility and concentration during processing of

448 carrot puree, Food Chem. 133(1) (2012) 60-67.

449 [22] M. Kazemi, F. Khodaiyan, S.S. Hosseini, Utilization of food processing wastes of

21
450 eggplant as a high potential pectin source and characterization of extracted pectin,

451 Food Chem. 294 (2019) 339-346.

452 [23] J.S. Yang, T.H. Mu, M.M. Ma, Extraction, structure, and emulsifying properties of

453 pectin from potato pulp, Food Chem. 244 (2018) 197-205.

454 [24] G.D. Manrique, F.M. Lajolo, FT-IR spectroscopy as a tool for measuring degree

455 of methyl esterification in pectins isolated from ripening papaya fruit. Postharvest

456 Biol. Technol. 25(1) (2002) 99-107.

457 [25] C. Kyomugasho, S. Christiaens, A. Shpigelman, A.M. Van Loey, M.E. Hendrickx,

458 FT-IR spectroscopy, a reliable method for routine analysis of the degree of

459 methylesterification of pectin in different fruit-and vegetable-based matrices. Food

460 Chem. 176 (2015) 82-90.

461 [26] T. Funami, M. Nakauma, S. Ishihara, R. Tanaka, T. Inoue, G.O. Phillips, Structural

462 modifications of sugar beet pectin and the relationship of structure to

463 functionality, Food Hydrocolloids. 25(2) (2011) 221-229.

464 [27] B. Neckebroeck, S.H.E. Verkempinck, T. Bernaerts, D. Verheyen, M.E. Hendrickx,

465 A. M. Van Loey, Investigating the role of the different molar mass fractions of a

466 pectin rich extract from onion towards its emulsifying and emulsion stabilizing

467 potential, Food Hydrocolloids. 117 (2021) 106735.

468 [28] L. Wan, Q. Chen, M. Huang, F. Liu, S. Pan, Physiochemical, rheological and

469 emulsifying properties of low methoxyl pectin prepared by high hydrostatic

470 pressure-assisted enzymatic, conventional enzymatic, and alkaline de-

471 esterification: A comparison study, Food Hydrocolloids. 93 (2019) 146-155.

22
472 [29] J. Mierczyńska, J. Cybulska, P.M. Pieczywek, A. Zdunek, Effect of storage on

473 rheology of water-soluble, chelate-soluble and diluted alkali-soluble pectin in

474 carrot cell walls, Food Bioprocess Technol. 8 (2015) 171-180.

475 [30] J. Mierczyńska, J. Cybulska, A. Zdunek, Rheological and chemical properties of

476 pectin enriched fractions from different sources extracted with citric

477 acid, Carbohydr. Polym. 156 (2017) 443-451.

478 [31] C. Kyomugasho, C. Munyensanga, M. Celus, K. Dewettinck, A.M. Van Loey, T.

479 Grauwet, M.E. Hendrickx, Molar mass influence on pectin-Ca2+ adsorption

480 capacity, interaction energy and associated functionality: Gel microstructure and

481 stiffness, Food Hydrocolloids. 85 (2018) 331-342.

482 [32] M.C. Ralet, V. Dronnet, H.C. Buchholt, J.F. Thibault, Enzymatically and

483 chemically de-esterified lime pectins: Characterisation, polyelectrolyte behaviour

484 and calcium binding properties, Carbohydr. Res. 336 (2001) 117-125.

485 [33] D.N. Sila, S. Van Buggenhout, T. Duvetter, I. Fraeye, A. De Roeck, A. Van Loey,

486 M. Hendrickx, Pectins in processed fruits and vegetables: Part II—Structure–

487 function relationships. Compr. Rev. Food Sci. Food Saf. 8(2) (2009) 86-104.

488 [34] S.H.E. Verkempinck, C. Kyomugasho, L. Salvia-Trujillo, S. Denis, M. Bourgeois,

489 A.M. Van Loey, M.E. Hendrickx, T. Grauwet, Emulsion stabilizing properties of

490 citrus pectin and its interactions with conventional emulsifiers in oil-in-water

491 emulsions, Food Hydrocolloids. 85 (2018) 144-157.

492 [35] Y. Tan, Z. Zhang, J.M. Mundo, D.J. McClements, Factors impacting lipid digestion

493 and nutraceutical bioaccessibility assessed by standardized gastrointestinal model

23
494 (INFOGEST): Emulsifier type, Food Res. Int. 137 (2020) 109739.

495 [36] K. Yao, D.J. McClements, C. Yan, J. Xiao, H. Liu, Z. Chen, X. Liu, In vitro and in

496 vivo study of the enhancement of carotenoid bioavailability in vegetables using

497 excipient nanoemulsions: Impact of lipid content, Food Res. Int. 141 (2021)

498 110162.

499 [37] S. Warnakulasuriya, P.K. Pillai, A.K. Stone, M.T. Nickerson, Effect of the degree

500 of esterification and blockiness on the complex coacervation of pea protein isolate

501 and commercial pectic polysaccharides, Food Chem. 264 (2018) 180-188.

502 [38] Y. Lan, J.B. Ohm, B. Chen, J. Rao, Phase behavior, thermodynamic and

503 microstructure of concentrated pea protein isolate-pectin mixture: Effect of pH,

504 biopolymer ratio and pectin charge density, Food Hydrocolloids. 101 (2020)

505 105556.

506 [39] J. Chen, R.H. Liang, L. Wei, S. J. Luo, C.M. Liu, S.S. Wu, et al. Extraction of

507 pectin from Premna microphylla turcz leaves and its physicochemical properties,

508 Carbohydr. Polym. 102(1) (2014) 376-384.

509 [40] Z. Zhang, M. Nie, Y. Xiao, L. Zhu, R. Gao, C. Zhou, Z. Dai, Positive effects of

510 ultrasound pretreatment on the bioaccessibility and cellular uptake of bioactive

511 compounds from broccoli: Effect on cell wall, cellular matrix and

512 digesta, LWT. 149 (2021) 112052.

513 [41] R. Zhang, G. Chen, B. Yang, Y. Wu, M. Du, J. Kan, Insights into the stability of

514 carotenoids and capsaicinoids in water-based or oil-based chili systems at different

515 processing treatments, Food Chem. 342 (2021) 128308.

24
516 [42] T.A.J. Verrijssen, S.H.E. Verkempinck, S. Christiaens, A.M. Van Loey, M. E.

517 Hendrickx, The effect of pectin on in vitro β-carotene bioaccessibility and lipid

518 digestion in low fat emulsions, Food Hydrocolloids. 49 (2015) 73-81.

519 [43] M. Iddir, J.F.P. Yaruro, Y. Larondelle, T. Bohn. Gastric lipase can significantly

520 increase lipolysis and carotenoid bioaccessibility from plant food matrices in the

521 harmonized INFOGEST static in vitro digestion model, Food Funct. 12(19) (2021)

522 9043-9053.

523 [44] K. Alba, V. Kontogiorgos, Pectin at the oil-water interface: Relationship of

524 molecular composition and structure to functionality, Food Hydrocolloids. 68

525 (2017) 211-218.

526 [45] D. A. Mendez, M. J. Fabra, A. Martínez-Abad, Μ. Μartínez-Sanz, M. Gorria, A.

527 López-Rubio, Understanding the different emulsification mechanisms of pectin:

528 Comparison between watermelon rind and two commercial pectin sources, Food

529 Hydrocolloids. 120 (2021) 106957.

530 [46] H. Niu, X. Chen, T. Luo, H. Chen, X. Fu, Relationships between the behavior of

531 three different sources of pectin at the oil-water interface and the stability of the

532 emulsion, Food Hydrocolloids. 128 (2022) 107566.

533 [47] L.M. de Souza Mesquita, B.V. Neves, L.P. Pisani, V. V. de Rosso, Mayonnaise as

534 a model food for improving the bioaccessibility of carotenoids from Bactris

535 gasipaes fruits, LWT. 122 (2020) 109022.

536 [48] L. Salvia-Trujillo, S.H.E. Verkempinck, X. Zhang, A.M. Van Loey, T. Grauwet,

537 M.E. Hendrickx, Comparative study on lipid digestion and carotenoid

25
538 bioaccessibility of emulsions, nanoemulsions and vegetable-based in situ

539 emulsions, Food Hydrocolloids. 87 (2019) 119-128.

26
540 Figure captions

541 Fig. 1 Fourier transform infrared spectroscopy (FT-IR) of pectic polymers (A).

542 Composition and content of monosaccharides of pectic polymers (B). HG, RG-I

543 and RG are the abbreviations of homogalacturonan, rhamnogalacturonan-I and

544 rhamnogalacturonan, respectively.

545 Fig. 2 Hypothetical depiction of the pectic polymers structure. HG, RG-I and RG are

546 the abbreviations of homogalacturonan, rhamnogalacturonan-I and

547 rhamnogalacturonan, respectively.

548 Fig. 3 Pearson correlation analysis of pectic polymer structure and carotenoid

549 bioaccessibility. A positive and negative correlation between these indicators is

550 denoted by the colors blue and red, respectively. Higher color intensity represents

551 the higher correlation coefficient.

552 Fig. 4 ζ-potential (A) and apparent viscosity (B) of different systems during digestion.

553 Simulated systems containing β-carotene, β-carotene with HG, RG-I or RG were

554 abbreviated as SS, SS-HG, SS-RG-I, SS-RG, respectively. Different letters

555 indicate significant differences (P<0.05) among different systems in the same

556 digestion phase.

557 Fig. 5 β-carotene retention ratio and β-carotene bioaccessibility of different systems.

558 Simulated systems containing β-carotene, β-carotene with HG, RG-I or RG were

559 abbreviated as SS, SS-HG, SS-RG-I, SS-RG, respectively. Different letters

560 indicate statistically significant variations (P<0.05) among the systems during the

561 same digestion phase.

27
562 Fig. 6 Proposed mechanisms explaining the effects of pectic polymers on carotenoid

563 bioaccessibility. HG, RG-I and RG are the abbreviations of homogalacturonan,

564 rhamnogalacturonan-I and rhamnogalacturonan, respectively.

28
Revised manuscript (with changes marked)

1 Effects of pectin structure on carotenoid bioaccessibility in


2 simulated juice model
3

4 Jianing Liua,b, Jinfeng Bia,*, Xuan Liua,*, Dazhi Liua,c, Vincenzo Foglianob, Matthijs

5 Dekkerb, Ruud Verkerkb

a
7 Institute of Food Science and Technology, Chinese Academy of Agricultural Sciences

8 (CAAS), Key Laboratory of Agro-Products Processing, Ministry of Agriculture and

9 Rural Affairs, Beijing 100193, China

b
10 Food Quality and Design Group, Wageningen University & Research, Bornse

11 Weilanden 9, 6708 WG, Wageningen, the Netherlands

c
12 Laboratory of Food Chemistry, Wageningen University & Research, Bornse

13 Weilanden 9, 6708 WG, Wageningen, the Netherlands

14

15

16

*
17 Corresponding author.

18 Tel.:+86-1062812584.

19 E-mail address: bijinfeng2010@163.com and liuxuancaas@126.com.

20 ADD: No. 2, Yuanmingyuan West Road, Haidian District, Beijing, 100193, China.

21

1
22 Abstract

23 The impact of pectin structure on carotenoid bioaccessibility is still uncertain. This

24 study aims to investigate how the different pectic polymers affected the bioaccessibility

25 of carotenoids in a simulated juice model during static in vitro digestion. This study

26 includes homogalacturonan (HG), which is a linear pectic polymer,

27 rhamnogalacturonan-I (RG-I), which is a branched pectic polymer, and

28 rhamnogalacturonan (RG), which is a diverse pectic polymer rich in RG-I,

29 rhamnogalacturonan-II (RG-II), and xylogalacturonan domains. Juice models without

30 pectin had the highest carotenoid bioaccessibility, suggesting pectin has negative

31 effects on carotenoid bioaccessibility. During the intestinal phase, systems with HG

32 showed the highest viscosity, followed by systems with RG and systems with RG-I.

33 Systems with RG-I had lower carotenoid bioaccessibility than systems with HG and

34 RG-II. Both the percentage of RG-I and the average side chain length of RG-I had

35 negative correlations with carotenoid bioaccessibility. RG-I side chains with more

36 arabinose and/or galactose might cause lower carotenoid bioaccessibility in this juice

37 model system.

38 Keywords: β-carotene bioaccessibility; Pectic polymers; Homogalacturonan;

39 Rhamnogalacturonan-I; Rhamnogalacturonan-II; Digestive fluids

2
40 Introduction

41 Pectin is a polysaccharide found in plant cell walls, predominantly consisting of

42 1,4-linked-α-D-galacturonic acid units. The primary structural domains of pectin are

43 homogalacturonan (HG), rhamnogalacturonan-I (RG-I), rhamnogalacturonan-II (RG-

44 II), and xylogalacturonan (XGA). HG makes up approximately 65% of the pectin

45 molecule, and it contains a linear chain of galacturonic acid (GalA) with methyl-

46 esterified carboxyl groups. RG-I is the second most abundant pectic polymer (i.e. 20-

47 35% of pectins), which comprises a backbone of repeating GalA and rhamnose

48 disaccharides with arabinan, galactan, and/or arabinogalactan side chains. The most

49 structurally complex pectin, RG-II, makes up about 10% of pectin. The backbone of

50 RG-II is made up of HG, with intricate side chains including simple sugars that are

51 connected to the GalA. XGA is an HG substituted at O-3 with a β-linked xylose, and it

52 is found in pectin in relatively small proportions (<7%) [1,2].

53 Different structural characteristics of pectic polymers contribute to their different

54 functional properties and bioactivities [3,4]. Many of pectin's functional characteristics

55 are influenced by the quantity of esterified GalA residues and their distribution in HG.

56 When methyl-ester groups are removed from HG, it becomes calcium cross-linkable,

57 allowing the formation of supramolecular assemblies and gels [5]. Furthermore, earlier

58 research has shown that cooperative cation binding to pectin requires at least 6 to 10

59 consecutive non-methyl-esterified GalA units [6,7]. RG-I pectin is heterogeneous due

60 to its variable glycosyl linkages, monosaccharide composition of side chains, chain

61 lengths, and distribution. Ferulic acid groups alter the molecular weight in the cell wall

3
62 and lead to poor pectin gelation properties by promoting the cross-linking of neutral

63 sugars in RG-I through ferulic bridges [8]. Recently, the research focused on the

64 importance of the pectin hairy region. Pectin RG-I portion showed improved

65 bioactivities, including inhibition of cell migration, and immunological and prebiotic

66 activities [9,10]. In vitro tests with human colon adenocarcinoma (Caco-2) cells showed

67 that RG-II had promisingly strong anti-cancer proliferation activities [11].

68 During the upper gastrointestinal stage, pectin may affect the digestion and

69 absorption of lipophilic carotenoids, by inhibiting lipid digestion [12]. Compared with

70 high pectin concentration, low pectin concentration favored the binding of bile salt and

71 carotenoid micellization [13]. The degree of lipid digestion decreased with an

72 increasing degree of pectin methyl-esterification [14]. Several studies have reported

73 carotenoid bioaccessibility and pectin structural and functional characteristics

74 (galacturonic acid concentration, degree of methyl-esterification and acetylation, and

75 emulsifying activity) of juices are related [15,16], however, the molecular mechanism

76 underpinning this relationship is still uncertain. It is possible that specific pectic

77 polymers are actually responsible for the reduction of carotenoid bioaccessibility

78 [12,16]. Therefore, the aim of this study was to investigate the influences of different

79 pectic polymers on carotenoid bioaccessibility of simulated juice models using static in

80 vitro digestion. In the present study, the pectic polymers include HG domain-rich pectic

81 polymer, RG-I region-rich pectic polymer, and rhamnogalacturonan (RG) pectic

82 polymer, which is more diverse with RG-I, RG-II and XGA domains. The simulated

83 juice models were applied to directly understand how pectin polymers influences β-

4
84 carotene digestion. Viscosity and ζ-potential of digestive fluids, β-carotene retention

85 and β-carotene bioaccessibility were recorded, to gain better knowledge of the digestive

86 process.

87

88 1. Materials and methods

89 2.1 Materials

90 Polygalacturonic acid (citrus pectin, ~5% methylation), rhamnogalacturonan I

91 (potato), and rhamnogalacturonan (soybean) were purchased from Megazyme (Bray,

92 Ireland). Polygalacturonic acid is seen as an HG domain-rich pectic polymer.

93 Rhamnogalacturonan I is considered to be an RG-I region-rich pectic polymer.

94 Rhamnogalacturonan is considered a more diverse pectic polymer rich in RG-I, RG-II

95 and XGA domains. Porcine gastric mucin, pepsin from porcine gastric mucosa, lipase

96 from porcine pancreas, β-carotene (≥ 93%), β-carotene (HPLC grade) and butylated

97 hydroxytoluene (BHT) were purchased from Sigma-Aldrich (St. Louis, MO, USA).

98

99 2.2 Methods

100 2.2.1 Juice model system preparation

101 The juice model system was prepared according to our previous study, with minor

102 modifications [17]. β-carotene (30 mg) in 100 mL acetic acid-sodium acetate buffer (0.2

103 M, pH=6) with 0.2% corn oil (w/w) was dispersed using a high-speed dispersion mixer

104 (T25 Ultra-TURRAX, IKA, Germany) at 14,000 rpm for 40 min. Subsequently, HG

105 (polygalacturonic acid), RG-I (rhamnogalacturonan I), and RG (rhamnogalacturonan),

5
106 at a proportion of 1% (w/v), were added to the β-carotene solution, and fully mixed

107 using the Ultra-TURRAX at 7,000 rpm for 10 min. The systems were then adjusted to

108 pH 6 and homogenized at 250 MPa (SPCH-EP-IC-16-30, Homogenising Systems Ltd,

109 London, UK). Simulated systems containing β-carotene, β-carotene with HG, RG-I, or

110 RG were abbreviated as SS, SS-HG, SS-RG-I, and SS-RG, respectively (Table 1).

111

112 2.2.2 Structural characterization of pectic polymers

113 2.2.2.1 Fourier transform infrared spectroscopy (FT-IR) analysis

114 The FT-IR spectra of pectic polymers were determined using an FT-IR

115 spectrometer (Tensor 27, Bruker, Germany) in the wavenumber region 4000~400 cm-1

116 at a resolution of 4 cm-1 and a cumulative scan of 64. Before FTIR analysis, pectic

117 polymers were pulverized and pressed into flake with KBr at a ratio of 1:100 (w/w).

118

119 2.2.2.2 Determination of galacturonic acid (GalA) and neutral sugar concentration

120 GalA and neutral sugar concentrations were performed according to the previous

121 study [16]. High-performance anion-exchange chromatography (HPAEC) was achieved

122 using ICS-3000 system (Dionex, Sunnyvale, CA, USA) equipped with a CarboPacTM

123 PA20 column and a pulsed amperometric detector. Monosaccharides including GalA,

124 galactose (Gal), arabinose (Ara), rhamnose (Rha), xylose (Xyl), and fucose (Fuc) were

125 used for identification and quantification.

126 To obtain information on the domains of the pectic polymers, five composition

127 ratios were calculated based on the mol% quantifiable of GalA and neutral sugars, with

6
128 some modifications [18,19]:
GalA+Rha
129 The linearity of pectic polymers = (1)
GalA+Rha+Ara+Gal+Xyl+Fuc

GalA−Rha
130 The percentage of HG (% HG) = (2)
GalA+Rha+Ara+Gal+Xyl+Fuc

GalA−Rha
131 The ratio of HG to RG-I backbone = (3)
2×Rha

2×Rha+Ara+Gal
132 The percentage of RG-I (% RG-I) = (4)
GalA+Rha+Ara+Gal+Xyl+Fuc

Ara+Gal
133 The average side chain length of RG-I = (5)
Rha

134

135 2.2.2.3 Determination of molar mass distribution

136 The molar mass distributions of pectins were determined using high-performance

137 size exclusion chromatography (HPSEC) coupled with multiangle laser light scattering

138 (MALLS) (Dawn Heleos II, Wyatt Technology, Santa Barbara, CA, USA), reflective

139 index (RI) detector (Optilab rEX, Wyatt Technology, Santa Barbara, CA, USA) and

140 ultraviolet (UV) detector (L-2400, Hitachi, Tokyo, Japan). The dn/dc value (refractive

141 index increment) was 0.146 mL/g. Pectins (10 mg) were dissolved in eluent (0.1 M

142 NaCl) overnight and filtrated through a 0.45 μm filter membrane before analysis.

143 Samples (100 μL) were injected into an OHpak SB-806M HQ column (8.0 × 300 mm,

144 Shodex, Tokyo, Japan). Elution was performed with 0.1 M NaCl solution at a flow rate

145 of 0.5 mL/min for 30 min and the column was kept at 35 °C. The polydispersity index

146 (PDI) was calculated as the ratio of the weight-average molar mass (Mw) to the number

147 average molar mass (Mn).

148

149 2.2.3 Static in vitro digestion

7
150 Juice model systems were subjected to an in vitro digestion that encompassed

151 simulated oral, gastric, and intestinal phases [20]. During intestine digestion stage, an

152 automatic titration facility (Metrohm, USA, Inc.) was utilized to maintain pH at 7.0 by

153 adding 0.25 M NaOH solution over 2 h at 37°C. To stop the enzymatic reaction, samples

154 were placed on ice after each digestion phase.

155

156 2.2.4 Digestive fluid characteristics

157 2.2.4.1 Viscosity property of the digestive fluids

158 The viscosity of the digestive fluids at a shear rate of 50 s-1 was measured by using

159 a Physica MCR 301 rheometer (Anton Paar, Graz, Austria). The parallel plate geometry

160 had a 40 mm diameter and the gap size was 1 mm.

161

162 2.2.4.2 ζ-potential of the digestive fluids

163 The ζ-potential values of the initial systems (pectin solutions with β-carotene) and

164 their digestive fluids during digestion were measured by using a Zetasizer Nano ZS

165 analyzer (Malvern Instruments Ltd., Malvern, UK).

166

167 2.2.5 β-carotene bioaccessibility and retention ratio

168 After the above digestion process, the raw intestinal digesta was centrifuged at

169 8000 rpm (8801×g) and 4 °C for 1 h. β-carotene bioaccessibility was calculated as

170 follows:

171 β-carotene bioaccessibility (%)=(Cmicelle/Craw digesta) ×100 (6)

8
172 where Cmicelle and Craw digesta are the contents of β-carotene in the micelle fraction and

173 the raw digesta, respectively.

174 The β-carotene retention ratio during each digestion phase was calculated as

175 follows:

176 β-carotene retention ratio (%)=Cdigestion/Cinitial×100 (7)

177 where Cdigestion and Cinitial are the β-carotene concentrations (μg/mL) in different systems

178 during each digestion phase and in the initial system, respectively.

179

180 2.2.6 Extraction and determination of β-carotene

181 A previous procedure was used to extract and quantify β-carotene [21]. Samples

182 were separated and analyzed on high-performance liquid chromatography (HPLC)

183 system with a reversed-phase C30-column (250 mm × 4.6 mm, 5 μm; YMC Europe,

184 Dinslaken, Germany), a binary HPLC pump (1525, Waters, Milford, MA, USA) and a

185 photodiode array detector (2998, Waters, Milford, MA, USA). The gradient program,

186 detection wavelength, flow rate of mobile phase were performed according to a

187 previous procedure [16].

188

189 2.2.7 Statistics

190 The experiments were conducted in triplicate, and each sample was analyzed three

191 times to ensure accuracy and reproducibility. All data were presented as mean ±

192 standard deviation. Analysis of variance (ANOVA) followed by a Tukey test was used

193 to determine significant differences using R software (Version 4.2.0). The Pearson

9
194 correlation analysis was carried out by the “corrplot” package in R software.

195

196 2. Results and discussion

197 2.1 Structural characterization of pectin polymers

198 3.1.1 Functional groups of pectic polymers

199 As shown in Fig. 1A, the following were the main characteristics of the absorption

200 peaks: (i) The absorption peak at ≈ 3364 cm-1 was attributed to O-H stretching vibration,

201 and the peak at ≈ 2936 cm-1 was caused by C-H stretching of CH2 groups; (ii) the peaks

202 at ≈ 1745 cm-1 and 1608 cm-1 corresponded to the C=O stretching vibration of methyl-

203 esterified carboxyl groups and C=O asymmetrical stretching vibration of free carboxyl,

204 respectively; (iii) the peak at ≈ 1418 cm-1 was ascribed to a symmetric carboxylate

205 stretch [22,23]. The ratio of peak area at 1745 cm-1 (related to methyl ester carbonyl

206 group stretching) to the sum of the peak areas at 1745 cm-1 and 1608 cm-1 (related to

207 carboxylate group stretching) was calculated as the degree of methyl-esterification (DM)

208 [24,25]. The calculated DM of HG and RG-I were 8% and 4%, respectively. However,

209 the absorption area of RG at 1745 cm-1 was not detected, which suggested that methyl-

210 esters were not present in the RG chain. The variation of the DM observed could be

211 attributed to the different extraction methods applied.

212

213 3.1.2 Monosaccharide content and composition ratios of pectic polymers

214 As shown in Fig. 1B, the GalA content of HG was the highest (94.1%), followed

215 by RG-I (70.9%) and RG (50.9%). HG is composed of 94% GalA and 6% other

10
216 monosaccharides, such as 3.9% Gal and 1.5% Rha. RG-I comprises 70.9% GalA,

217 followed by 19.8% Gal, 6.3% Rha, and 1.8% Ara that constitute the RG-I region. RG

218 is a more diverse pectic polymer compared to HG and RG-I, as evidenced by its

219 composition of 50.9% GalA, 16.9% Xyl, and 13.1% Fuc. Individual monosaccharide

220 contents are not sufficient to gain insight into the pectin chains. With the help of the

221 ratios between the composing monosaccharides, the overall structure of pectin chains

222 becomes more visible.

223 As shown in Table 2, compared to the other two pectic polymers, HG is composed

224 of more linear/less branched pectin evidenced by the higher linearity of pectic

225 polymers, % HG and the ratio of HG to RG-I backbone. The higher linearity of HG

226 makes it potential to have higher Ca2+-binding capability, due to the higher possibility

227 of consecutive non-methyl-esterified GalA residues. RG-I is composed of a more

228 branched RG-I domain according to higher % RG-I and average side chain length of

229 RG-I. Pectin polymers rich in RG-I side chains have better emulsion stability by

230 forming a hydrated layer [26]. RG is a pectic polymer with more diverse domains,

231 which consists of 26.87% RG-I and approximately 13.14% RG-II side chains and 16.94%

232 XGA side chains. The percentage of the RG-II and XGA backbones is unclear, though,

233 as both of their backbones are HG. Based on the obtained structural characteristics, the

234 structures of the pectic polymers are visualized as hypothetical model structures

235 presented in Fig. 2. These distinct monosaccharide compositions among different pectic

236 polymers make it possible for further investigation of the effects of pectic polymers on

237 carotenoid bioaccessibility.

11
238

239 3.1.3 Molar mass distribution of pectic polymers

240 Mw and PDI values are presented in Table 2, and a higher PDI represents a wider

241 molecular mass distribution. RG showed the highest Mw, followed by RG-I and HG.

242 HG showed the lowest PDI values compared to RG and RG-I. The rich side chains of

243 RG-I and RG might contribute to their higher Mw and wider molecular mass distribution.

244 According to correlation analysis, the linearity of pectin had a negative correlation with

245 Mw, indicating that a higher degree of rhamnification might correspond to a larger Mw

246 of the polysaccharide (Fig. 3). Many pectin properties are linked to its Mw, and high

247 molecular mass polymers generally contribute to promising functionalities, such as

248 emulsifying and emulsion stabilizing abilities [27]. In most cases, higher Mw

249 macromolecules obstruct fluid motion, resulting in higher viscosity solutions [28]. It

250 should be noted that other factors, such as pectin sources and pectin fractions, might

251 also affect the viscosity of pectins [29,30]. Decreasing Mw might reduce the maximum

252 amount of Ca2+ that pectin can bind [31], however, the calcium binding ability also

253 depends on the degree and distribution pattern of methyl-ester groups among HG region

254 [32].

255

256 3.2 Physical characteristics of digestive fluids

257 3.2.1 ζ-potential of digestive fluids

258 Different physical indicators were estimated to investigate the effects of pectic

259 polymers on digestion process. Generally, changes in the ζ-potential of particles can

12
260 influence their electrostatic repulsion or attraction, which in turn can affect the stability

261 and behavior of the systems during digestion. As seen in Fig. 4A, initial ζ-potential

262 values of SS-HG and SS-RG were lower than those of SS and SS-RG-I, which

263 suggested that SS-HG and SS-RG were enveloped by more negatively charged

264 substances. Typically, an increased GalA content in pectin results in a higher negative

265 charge [33]. Additionally, more demethylesterified pectin can carry more negative

266 charges [34]. During the oral phase, the ζ-potential value of SS was the highest,

267 followed by SS-RG, SS-RG-I and SS-HG. The ζ-potential values increased from the

268 initial phase to the gastric phase, while decreased in the intestinal phase. The highly

269 acidic conditions, which cause the protein-coated droplets to be positive, and the

270 presence of anionic mucin, which binds to the cationic protein surfaces, are both related

271 to the increase in ζ-potential during the gastric phase [35,36]. Specifically, GalA, the

272 degree of methyl-esterification and the distribution of methyl esters of pectin influence

273 their interactions with protein [37]. Moreover, the pectin surface charge decreased with

274 the decreasing pH, and at pH 2, the ζ-potential of pectin is close to zero [38]. The

275 amount of the negative charge on the backbone of pectin decreased as a result of

276 protonation (a drop in pH). The fact that the ζ-potential values changed from their initial

277 positive values to more negative ones during the intestinal phase could be attributed to

278 a number of negatively charged species, including free fatty acids released during

279 lipolysis, intricate colloidal structures, undigested oil, and anionic polymers [14].

280

281 3.2.2 Apparent viscosity of digestive fluids

13
282 The apparent viscosity of digestive fluids is given in Fig. 4B. The physical feature

283 of apparent viscosity is linked to soluble dietary fibers, particularly pectic

284 polysaccharides. Initial SS-HG and SS-RG showed higher viscosity than SS-RG-I.

285 Generally, pectin with a higher Mw has a higher viscosity [28]. Moreover, HG pectin

286 contributed greatly to the viscosity of pectin [39]. There was a significant negative

287 correlation between the initial viscosity of the systems and the RG-I percentage (Fig.

288 3). Due to the neutral sugar side chains, solutions containing high contents of RG-I

289 and/or RG-II pectin usually have lower viscosity values than solutions with high GalA

290 content [39]. From the initial phase to the gastric phase, the viscosity of the four systems

291 all decreased, which might be caused by the dilution effect of digestive fluids. These

292 results are consistent with those obtained with broccoli-digested fractions, where

293 viscosity decreased from the gastric phase to the intestinal phase [40]. During the

294 intestinal phase, SS-HG showed the highest viscosity, followed by SS-RG, SS-RG-I

295 and SS. A negative correlation has been found between the HG to RG-I ratio and the

296 viscosity of the systems during the intestinal phase. The interactions between pectic

297 polymers and the digestive fluids might also contribute to the results, especially the

298 crosslinking between pectin and digestive enzymes.

299

300 3.3 β-carotene retention ratio during digestion phases

301 As seen in Fig. 5, in the oral phase, the β-carotene retention ratio of SS-RG was

302 lower than that of SS, while in the gastric phase, the β-carotene retention ratio of SS-

303 RG was the highest compared with other systems. This suggested that SS-RG might be

14
304 more stable under acidic conditions. At the intestinal phase, the β-carotene retention

305 ratios of SS-HG, SS-RG-I, and SS-RG had no significant difference. Carotenoids are

306 isoprenoid pigments with a polyene backbone that contains a variable number of

307 conjugated double bonds, therefore, they are prone to be oxidized. A previous study

308 showed that the stability of carotenoids at different processing can be significantly

309 improved in an oil-based system, compared with a water-based juice system [41]. In

310 our study, the carotenoids were in a homogenized juice model system, which is similar

311 to a water-based juice system, therefore, carotenoid stability might be affected by high-

312 pressure homogenization to some extent. A negative correlation was observed between

313 the β-carotene retention ratio during the intestinal phase and the ζ-potential values of

314 initial systems. This suggests that a higher negative charge could potentially enhance

315 the delivery of carotenoids during the intestinal stage. Additionally, a negative

316 correlation was found between the β-carotene retention ratio and the viscosity of

317 digestive fluids during the intestinal phase. Therefore, the pectin-induced viscosity

318 might have effects on the β-carotene stability during the intestinal stage.

319

320 3.4 β-carotene bioaccessibility

321 Carotenoids need to be released from the food matrix, solubilized into the lipid

322 phase, and incorporated into mixed micelles in order to be absorbed via the human

323 intestinal epithelium [12]. The levels of carotenoids available for intestinal absorption,

324 specifically the carotenoids in mixed micelles, are referred to as bioaccessible

15
325 carotenoids. SS had the highest β-carotene bioaccessibility compared to other systems

326 (SS-HG, SS-RG-I and SS-RG), which signified that the presence of pectin could reduce

327 the bioaccessibility carotenoids, thus limiting their potential health effects. Similar

328 observations were also found by Cervantes-Paz et.al [13], in which high pectin

329 concentration decreased carotenoid micellization. Pectin can increase the viscosity of

330 digestive fluids, allowing digesta to last longer in the gastrointestinal phase while

331 simultaneously reducing the movement of substrates and digestive enzymes (eg., pepsin

332 and lipase), decreasing lipase activity and micelle generation [42]. Moreover, different

333 pectic polymers caused different influences on carotenoid bioaccessibility. SS-HG and

334 SS-RG had similar β-carotene bioaccessibility, while SS-RG-I had a smaller β-carotene

335 bioaccessibility than those of SS-HG and SS-RG. Furthermore, both % RG-I and the

336 average side chain length of RG-I had negative correlations with carotenoid

337 bioaccessibility. Therefore, it was deduced that more RG-I side chains with Ara and/or

338 Gal might cause lower carotenoid bioaccessibility in this juice model system. These

339 results are consistent with those obtained with carrot juice, where higher emulsifying

340 properties caused a lower value of total carotenoid bioaccessibility [16]. The proposed

341 mechanisms of the effects of pectic polymers on carotenoid bioaccessibility are shown

342 in Fig. 6. The lipid hydrolysis produces monoacylglycerols and free fatty acids, which

343 together with bile salts contribute to the formation of mixed micelles. In this context,

344 lipase can greatly increase lipolysis and carotenoid bioaccessibility [43]. Pectin

345 segments with enriched RG-I/neutral sugar side chains contribute to better emulsifying

346 properties [44,45]. Pectin with better emulsification properties might be able to form

16
347 thicker hydrated layers at the oil-water interface and inhibit lipase to the surface,

348 resulting in low carotenoid bioaccessibility. A recent study based on molecular

349 dynamics simulations also revealed that RG-I and RG-II domains were adsorbed on the

350 oil-water interface in a folded form, creating a thick viscoelastic film while the HG

351 domain was tiled on the interface [46].

352 It should also be mentioned that in real fruit and vegetable systems, RG-I side

353 chains could also be cross-linked to xylans, xyloglucans, lignins, and proteins [5]. The

354 above-mentioned cross-linking among pectic polymers can also influence carotenoid

355 bioaccessibility, but this is not considered in the juice model system in this study.

356 Despite the absence of explicit localization obstacles of carotenoids in model systems,

357 β-carotene bioaccessibility was still comparatively low, ranging from 7.97% to 14.55%.

358 It implies that lipids are important in enhancing β-carotene bioaccessibility since dietary

359 lipids could stimulate bile salt secretion and increase the number of carotenoids carried

360 in micelles. Furthermore, other consumption forms of food products might be helpful

361 in increasing carotenoid bioaccessibility. Making fruit and vegetable-based emulsions

362 has been noted as a promising method to increase carotenoid bioaccessibility by

363 researchers [47,48]. Therefore, further research might focus on the effects of pectic

364 polymers on carotenoid bioaccessibility in model systems with the addition of more

365 lipids.

366

367 4. Conclusions

368 The simulated juice models were helpful to understand the effects of different

17
369 pectic polymers on carotenoid bioaccessibility. The ζ-potential values of all the systems

370 increased from the initial phase to the gastric phase, while decreased in the intestinal

371 phase. When compared to the initial ζ-potential values, ζ-potential became more

372 negative during the intestinal phase. During the intestinal phase, SS-HG showed the

373 highest viscosity, followed by SS-RG, SS-RG-I and SS. SS-HG and SS-RG had similar

374 β-carotene bioaccessibility, while SS-RG-I had a lower β-carotene bioaccessibility than

375 those of SS-HG and SS-RG. Both % RG-I and the average side chain length of RG-I

376 had negative influences on carotenoid bioaccessibility.

377

378 Acknowledgements

379 This work was supported by the National Key Research and Development

380 Program of China (2022YFD1600505), the earmarked fund for China Agriculture

381 Research System (CARS-27), and the Joint PhD Program between Wageningen

382 University & Research (WUR) and Chinese Academy of Agricultural Sciences (CAAS)

383 (No. MOE11NL1A20151701N).

18
384 References

385 [1] D. Mohnen, Pectin structure and biosynthesis, Curr. Opin. Plant Biol. 11(3) (2008)

386 266-277.

387 [2] J. Zandleven, S.O. Sørensen, J. Harholt, G. Beldman, H.A. Schols, H.V. Scheller, &

388 A.J. Voragen, Xylogalacturonan exists in cell walls from various tissues of

389 Arabidopsis thaliana, Phytochemistry. 68(8) (2007) 1219-1226.

390 [3] B. Neckebroeck, S.H.E. Verkempinck, G. Vaes, K. Wouters, J. Magnee, M. E.

391 Hendrickx, et al. Advanced insight into the emulsifying and emulsion stabilizing

392 capacity of carrot pectin subdomains, Food Hydrocolloids, 102 (2020) 105594.

393 [4] P. M. Pieczywek, J. Ciesla, W. Płazinski, A. Zdunek, Aggregation and weak gel

394 formation by pectic polysaccharide homogalacturonan, Carbohydr. Polym. 256

395 (2021) 117566.

396 [5] W.G.T. Willats, L. Mccartney, W. Mackie, & J.P. Knox, Pectin: Cell biology and

397 prospects for functional analysis, Plant Mol. Biol. 47(1-2) (2001) 9-27.

398 [6] P.J.H. Daas, B. Boxma, A.M.C.P. Hopman, A.G.J. Voragen, H.A. Schols,

399 Nonesterified galacturonic acid sequence homology of pectins, Biopolymers. 58(1)

400 (2001) 1-8.

401 [7] G.A. Luzio, R.G. Cameron, Demethylation of a model homogalacturonan with the

402 salt-independent pectin methylesterase from citrus: Part II. structure–function

403 analysis, Carbohydr. Polym. 71(2) (2008) 300-309.

404 [8] A. Oosterveld, I.E. Pol, G. Beldman, A.G. Voragen, Isolation of feruloylated

405 arabinans and rhamnogalacturonans from sugar beet pulp and their gel forming

19
406 ability by oxidative cross-linking, Carbohydr. Polym. 44(1) (2001) 9-17.

407 [9] D. Wu, J. Zheng, G. Mao, W. Hu, X. Ye, R.J. Linhardt, et al. Rethinking the impact

408 of RG-I mainly from fruits and vegetables on dietary health, Crit. Rev. Food Sci.

409 Nutr. 60 (2020) 2938-2960.

410 [10] M.Y. Zhang, J. Cai, Preparation of branched RG-I-rich pectin from red dragon fruit

411 peel and the characterization of its probiotic properties, Carbohydr. Polym. (2023)

412 120144.

413 [11] L. Ai, Y. Chung, S. Lin, K. Lee, P. Lai, Y. Xia, et al. Active pectin fragments of

414 high in vitro antiproliferation activities toward human colon adenocarcinoma cells:

415 Rhamnogalacturonan II, Food Hydrocolloids. 83 (2018) 239-245.

416 [12] B. Cervantes-Paz, J.J. Ornelas-Paz, S. Ruiz-Cruz, C. Rios-Velasco, V. Ibarra-

417 Junquera, E.M. Yahia, et al. Effects of pectin on lipid digestion and possible

418 implications for carotenoid bioavailability during pre-absorptive stages: A review,

419 Food Res. Int. 99 (2017) 917-927.

420 [13] B. Cervantes-Paz, J.D.J. Ornelas-Paz, J.D. Pérez-Martínez, J. Reyes-Hernández,

421 P.B. Zamudio-Flores, C. Rios-Velasco, et al. Effect of pectin concentration and

422 properties on digestive events involved on micellarization of free and esterified

423 carotenoids, Food Hydrocolloids. 60 (2016) 580-588.

424 [14] M. Espinal-Ruiz, L.P. Restrepo-Sanchez, C.E. Narvaez-Cuenca, D.J. McClements,

425 Impact of pectin properties on lipid digestion under simulated gastrointestinal

426 conditions: Comparison of citrus and banana passion fruit (Passiflora tripartita var.

427 mollissima) pectins, Food Hydrocolloids. 52 (2016) 329-342.

20
428 [15] Y. Ding, X. Liu, J. Bi, X. Wu, X. Li, J. Liu, et al. Effects of pectin, sugar and pH

429 on the β-Carotene bioaccessibility in simulated juice systems, LWT--Food Sci.

430 Technol. 124 (2020) 109125.

431 [16] X. Liu, J. Liu, J. Bi, J. Yi, J. Peng, C. Ning, et al. Effects of high pressure

432 homogenization on pectin structural characteristics and carotenoid bioaccessibility

433 of carrot juice, Carbohydr. Polym. 203 (2019) 176-184.

434 [17] J. Liu, J. Bi, X. Liu, D. Liu, X. Wu, J. Lyu, et al. Effects of pectins and sugars on

435 β-carotene bioaccessibility in an in vitro simulated digestion model, J. Food

436 Compos. Anal. 91 (2020) 103537.

437 [18] S.E. Broxterman, P. Picouet, H.A. Schols, Acetylated pectins in raw and heat

438 processed carrots, Carbohydr. Polym. 177 (2017) 58-66.

439 [19] K. Houben, R.P. Jolie, I. Fraeye, A.M. Van Loey, M.E. Hendrickx, Comparative

440 study of the cell wall composition of broccoli, carrot, and tomato: Structural

441 characterization of the extractable pectins and hemicelluloses, Carbohydr. Res.

442 346(9) (2011) 1105-1111.

443 [20] J. Liu, D. Liu, J. Bi, X. Liu, Y. Lyu, R. Verkerk, et al. Micelle separation conditions

444 based on particle size strongly affect carotenoid bioaccessibility assessment from

445 juices after in vitro digestion, Food Res. Int. 151 (2022) 110891.

446 [21] G. Knockaert, L. Lemmens, S. Van Buggenhout, M. Hendrickx, A. Van Loey,

447 Changes in β-carotene bioaccessibility and concentration during processing of

448 carrot puree, Food Chem. 133(1) (2012) 60-67.

449 [22] M. Kazemi, F. Khodaiyan, S.S. Hosseini, Utilization of food processing wastes of

21
450 eggplant as a high potential pectin source and characterization of extracted pectin,

451 Food Chem. 294 (2019) 339-346.

452 [23] J.S. Yang, T.H. Mu, M.M. Ma, Extraction, structure, and emulsifying properties of

453 pectin from potato pulp, Food Chem. 244 (2018) 197-205.

454 [24] G.D. Manrique, F.M. Lajolo, FT-IR spectroscopy as a tool for measuring degree

455 of methyl esterification in pectins isolated from ripening papaya fruit. Postharvest

456 Biol. Technol. 25(1) (2002) 99-107.

457 [25] C. Kyomugasho, S. Christiaens, A. Shpigelman, A.M. Van Loey, M.E. Hendrickx,

458 FT-IR spectroscopy, a reliable method for routine analysis of the degree of

459 methylesterification of pectin in different fruit-and vegetable-based matrices. Food

460 Chem. 176 (2015) 82-90.

461 [26] T. Funami, M. Nakauma, S. Ishihara, R. Tanaka, T. Inoue, G.O. Phillips, Structural

462 modifications of sugar beet pectin and the relationship of structure to

463 functionality, Food Hydrocolloids. 25(2) (2011) 221-229.

464 [27] B. Neckebroeck, S.H.E. Verkempinck, T. Bernaerts, D. Verheyen, M.E. Hendrickx,

465 A. M. Van Loey, Investigating the role of the different molar mass fractions of a

466 pectin rich extract from onion towards its emulsifying and emulsion stabilizing

467 potential, Food Hydrocolloids. 117 (2021) 106735.

468 [28] L. Wan, Q. Chen, M. Huang, F. Liu, S. Pan, Physiochemical, rheological and

469 emulsifying properties of low methoxyl pectin prepared by high hydrostatic

470 pressure-assisted enzymatic, conventional enzymatic, and alkaline de-

471 esterification: A comparison study, Food Hydrocolloids. 93 (2019) 146-155.

22
472 [29] J. Mierczyńska, J. Cybulska, P.M. Pieczywek, A. Zdunek, Effect of storage on

473 rheology of water-soluble, chelate-soluble and diluted alkali-soluble pectin in

474 carrot cell walls, Food Bioprocess Technol. 8 (2015) 171-180.

475 [30] J. Mierczyńska, J. Cybulska, A. Zdunek, Rheological and chemical properties of

476 pectin enriched fractions from different sources extracted with citric

477 acid, Carbohydr. Polym. 156 (2017) 443-451.

478 [31] C. Kyomugasho, C. Munyensanga, M. Celus, K. Dewettinck, A.M. Van Loey, T.

479 Grauwet, M.E. Hendrickx, Molar mass influence on pectin-Ca2+ adsorption

480 capacity, interaction energy and associated functionality: Gel microstructure and

481 stiffness, Food Hydrocolloids. 85 (2018) 331-342.

482 [32] M.C. Ralet, V. Dronnet, H.C. Buchholt, J.F. Thibault, Enzymatically and

483 chemically de-esterified lime pectins: Characterisation, polyelectrolyte behaviour

484 and calcium binding properties, Carbohydr. Res. 336 (2001) 117-125.

485 [33] D.N. Sila, S. Van Buggenhout, T. Duvetter, I. Fraeye, A. De Roeck, A. Van Loey,

486 M. Hendrickx, Pectins in processed fruits and vegetables: Part II—Structure–

487 function relationships. Compr. Rev. Food Sci. Food Saf. 8(2) (2009) 86-104.

488 [34] S.H.E. Verkempinck, C. Kyomugasho, L. Salvia-Trujillo, S. Denis, M. Bourgeois,

489 A.M. Van Loey, M.E. Hendrickx, T. Grauwet, Emulsion stabilizing properties of

490 citrus pectin and its interactions with conventional emulsifiers in oil-in-water

491 emulsions, Food Hydrocolloids. 85 (2018) 144-157.

492 [35] Y. Tan, Z. Zhang, J.M. Mundo, D.J. McClements, Factors impacting lipid digestion

493 and nutraceutical bioaccessibility assessed by standardized gastrointestinal model

23
494 (INFOGEST): Emulsifier type, Food Res. Int. 137 (2020) 109739.

495 [36] K. Yao, D.J. McClements, C. Yan, J. Xiao, H. Liu, Z. Chen, X. Liu, In vitro and in

496 vivo study of the enhancement of carotenoid bioavailability in vegetables using

497 excipient nanoemulsions: Impact of lipid content, Food Res. Int. 141 (2021)

498 110162.

499 [37] S. Warnakulasuriya, P.K. Pillai, A.K. Stone, M.T. Nickerson, Effect of the degree

500 of esterification and blockiness on the complex coacervation of pea protein isolate

501 and commercial pectic polysaccharides, Food Chem. 264 (2018) 180-188.

502 [38] Y. Lan, J.B. Ohm, B. Chen, J. Rao, Phase behavior, thermodynamic and

503 microstructure of concentrated pea protein isolate-pectin mixture: Effect of pH,

504 biopolymer ratio and pectin charge density, Food Hydrocolloids. 101 (2020)

505 105556.

506 [39] J. Chen, R.H. Liang, L. Wei, S. J. Luo, C.M. Liu, S.S. Wu, et al. Extraction of

507 pectin from Premna microphylla turcz leaves and its physicochemical properties,

508 Carbohydr. Polym. 102(1) (2014) 376-384.

509 [40] Z. Zhang, M. Nie, Y. Xiao, L. Zhu, R. Gao, C. Zhou, Z. Dai, Positive effects of

510 ultrasound pretreatment on the bioaccessibility and cellular uptake of bioactive

511 compounds from broccoli: Effect on cell wall, cellular matrix and

512 digesta, LWT. 149 (2021) 112052.

513 [41] R. Zhang, G. Chen, B. Yang, Y. Wu, M. Du, J. Kan, Insights into the stability of

514 carotenoids and capsaicinoids in water-based or oil-based chili systems at different

515 processing treatments, Food Chem. 342 (2021) 128308.

24
516 [42] T.A.J. Verrijssen, S.H.E. Verkempinck, S. Christiaens, A.M. Van Loey, M. E.

517 Hendrickx, The effect of pectin on in vitro β-carotene bioaccessibility and lipid

518 digestion in low fat emulsions, Food Hydrocolloids. 49 (2015) 73-81.

519 [43] M. Iddir, J.F.P. Yaruro, Y. Larondelle, T. Bohn. Gastric lipase can significantly

520 increase lipolysis and carotenoid bioaccessibility from plant food matrices in the

521 harmonized INFOGEST static in vitro digestion model, Food Funct. 12(19) (2021)

522 9043-9053.

523 [44] K. Alba, V. Kontogiorgos, Pectin at the oil-water interface: Relationship of

524 molecular composition and structure to functionality, Food Hydrocolloids. 68

525 (2017) 211-218.

526 [45] D. A. Mendez, M. J. Fabra, A. Martínez-Abad, Μ. Μartínez-Sanz, M. Gorria, A.

527 López-Rubio, Understanding the different emulsification mechanisms of pectin:

528 Comparison between watermelon rind and two commercial pectin sources, Food

529 Hydrocolloids. 120 (2021) 106957.

530 [46] H. Niu, X. Chen, T. Luo, H. Chen, X. Fu, Relationships between the behavior of

531 three different sources of pectin at the oil-water interface and the stability of the

532 emulsion, Food Hydrocolloids. 128 (2022) 107566.

533 [47] L.M. de Souza Mesquita, B.V. Neves, L.P. Pisani, V. V. de Rosso, Mayonnaise as

534 a model food for improving the bioaccessibility of carotenoids from Bactris

535 gasipaes fruits, LWT. 122 (2020) 109022.

536 [48] L. Salvia-Trujillo, S.H.E. Verkempinck, X. Zhang, A.M. Van Loey, T. Grauwet,

537 M.E. Hendrickx, Comparative study on lipid digestion and carotenoid

25
538 bioaccessibility of emulsions, nanoemulsions and vegetable-based in situ

539 emulsions, Food Hydrocolloids. 87 (2019) 119-128.

26
540 Figure captions

541 Fig. 1 Fourier transform infrared spectroscopy (FT-IR) of pectic polymers (A).

542 Composition and content of monosaccharides of pectic polymers (B). HG, RG-I

543 and RG are the abbreviations of homogalacturonan, rhamnogalacturonan-I and

544 rhamnogalacturonan, respectively.

545 Fig. 2 Hypothetical depiction of the pectic polymers structure. HG, RG-I and RG are

546 the abbreviations of homogalacturonan, rhamnogalacturonan-I and

547 rhamnogalacturonan, respectively.

548 Fig. 3 Pearson correlation analysis of pectic polymer structure and carotenoid

549 bioaccessibility. A positive and negative correlation between these indicators is

550 denoted by the colors blue and red, respectively. Higher color intensity represents

551 the higher correlation coefficient.

552 Fig. 4 ζ-potential (A) and apparent viscosity (B) of different systems during digestion.

553 Simulated systems containing β-carotene, β-carotene with HG, RG-I or RG were

554 abbreviated as SS, SS-HG, SS-RG-I, SS-RG, respectively. Different letters

555 indicate significant differences (P<0.05) among different systems in the same

556 digestion phase.

557 Fig. 5 β-carotene retention ratio and β-carotene bioaccessibility of different systems.

558 Simulated systems containing β-carotene, β-carotene with HG, RG-I or RG were

559 abbreviated as SS, SS-HG, SS-RG-I, SS-RG, respectively. Different letters

560 indicate statistically significant variations (P<0.05) among the systems during the

561 same digestion phase.

27
562 Fig. 6 Proposed mechanisms explaining the effects of pectic polymers on carotenoid

563 bioaccessibility. HG, RG-I and RG are the abbreviations of homogalacturonan,

564 rhamnogalacturonan-I and rhamnogalacturonan, respectively.

28
Highlights

Highlights
 Simulated models reveal links between pectin and carotenoid bioaccessibility.

 Carotenoid bioaccessibility was lower in rhamnogalacturonan-I system.

 Rhamnogalacturonan-I ratio is negatively linked with carotenoid bioaccessibility.

 Longer side chains of rhamnogalacturonan-I reduce carotenoid bioaccessibility.


Figure Click here to access/download;Figure;Figures.docx

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Table (Editable version) Click here to access/download;Table (Editable version);Tables.doc

Table 1 Juice model systems in this study.

Juice systems Abbreviations

β-carotene solution SS
β-carotene solution+ 1% polygalacturonic acid SS-HG
β-carotene solution+ 1% rhamnogalacturonan I SS-RG-I
β-carotene solution+ 1% rhamnogalacturonan SS-RG
Table 2 Structural characteristics of different pectic polymers.

Homogalacturonan (HG) Rhamnogalacturonan-I (RG-I) Rhamnogalacturonan (RG)

The linearity of pectic polymers 95.59 ± 0.06 77.21 ± 0.32 58.73 ± 0.16
The percentage of HG (% HG) 92.56 ± 0.22 64.62 ± 0.77 43.06 ± 0.20
The ratio of HG to RG-I backbone 30.56 ± 1.61 5.13 ± 0.24 2.75 ± 0.08
The percentage of RG-I (% RG-I) 7.27 ± 0.23 34.18 ± 0.86 26.87 ± 0.27
The average side chain length of RG-I 2.80 ± 0.09 3.43 ± 0.06 1.43 ± 0.04
Weight-average molar mass (Mw) (g/mol) 9.63 × 104 4.53 × 105 1.43 × 106
The polydispersity index (PDI) 2.15 ± 0.36 2.11 ± 0.04 3.07 ± 0.11
The results were expressed as mean ± standard deviation.

You might also like