Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Food Chemistry 377 (2022) 132001

Contents lists available at ScienceDirect

Food Chemistry
journal homepage: www.elsevier.com/locate/foodchem

A novel glucofucobiose with potential prebiotic activity prepared from the


exopolysaccharides of Clavibacter michiganensis M1
Mengshi Xiao a, Xinmiao Ren a, Jinzheng Cui d, Rong Li e, Zhemin Liu a, Lin Zhu a, c, Qing Kong a,
Xiaodan Fu b, *, Haijin Mou a, *
a
College of Food Science and Engineering, Ocean University of China, Qingdao 266003, Shandong, People’s Republic of China
b
State Key Laboratory of Food Science and Technology, China-Canada Joint Laboratory of Food Science and Technology (Nanchang), Nanchang University, 235
Nanjing East Road, Nanchang, Jiangxi 330047, People’s Republic of China
c
Weihai Deepsea Biotechnology Co., Ltd, Weihai 264300, Shandong, People’s Republic of China
d
School of Medicine and Pharmacy, Ocean University of China, Qingdao 266003, Shandong, People’s Republic of China
e
Qingdao Women and Children Hospital, Qingdao 266003, Shandong, People’s Republic of China

A R T I C L E I N F O A B S T R A C T

Keywords: Fucose and fucosylated oligosaccharides have important applications in various industries owing to their pre­
Fucosylated oligosaccharide biotic, anti-inflammatory, anticoagulant, and antiviral activities. Here, we aimed to obtain fucosylated oligo­
2′ -Fucosyllactose saccharides using the acidolysis method to depolymerize exopolysaccharides extracted from Clavibacter
Bacterial exopolysaccharides
michiganensis M1. Based on structural analysis, the prepared glucofucobiose was found to consist of D-glucose and
Glucofucobiose
L-fucose, with a molecular weight of 326 Da and a structure of D-Glcp-β-(1→4)-L-Fucp. The prebiotic activity of
Structure
Prebiotic activity glucofucobiose was compared with that of 2′ -fucosyllactose (2′ -FL), the most abundant oligosaccharide in human
milk. According to the results, glucofucobiose could significantly promote the proliferation of six probiotic
strains, and short-chain fatty acid production of five probiotic strains on glucofucobiose was substantially higher
than that on 2′ -FL at 48 h of fermentation. Overall, this study proposed a new technology for obtaining fuco­
sylated oligosaccharides. The prepared glucofucobiose was found to exhibit potential prebiotic activity and
should be further assessed.

1. Introduction capabilities of the brain have been well established (Matthies,


Schroeder, Smalla, & Krug, 2000). However, it is unrealistic to obtain
Fucose is an industrially useful sugar with high economic potential fucose via chemical synthesis to meet the needs of large-scale industrial
and diverse physiological activities, such as antibacterial, anticoagulant, production. Therefore, new sources of fucose need to be explored.
and anticancer activities (Torres et al., 2012). In addition to being an Although the intestinal microbiota of adults is complex, that of
important component in eukaryotic cells, fucose is also commonly found healthy full-term infants is relatively simple comprising primary colo­
in natural plants, including brown algae and tragacanth gum, and it is nizers of the genus Bifidobacterium, which can persist until early child­
widely used in the pharmaceutical and cosmetic industries (Silchenko hood (Bäckhed et al., 2015). In the first six months following birth, a
et al., 2018; Abbasi et al., 2019). Further, its importance in the regula­ decrease in bifidobacteria or an increase in other bacteria may consid­
tion of intestinal microbiota and ability to improve perception erably change the natural composition of the microbial community,

Abbreviations: EPS, exopolysaccharides; HMOs, human milk oligosaccharides; 2′ -FL, 2′ -fucosyllactose; HCl, hydrochloric acid; TFA, trifluoroacetic acid; GF,
glucofucobiose; TLC, thin-layer chromatography; SCFAs, short-chain fatty acids; HPLC, high-performance liquid chromatography; ESI-MS, electron spray ionization
mass spectrometry; ESI-CID-MS, electron spray ionization in tandem with collision-induced dissociation tandem mass spectrometry; NMR, nuclear magnetic reso­
nance; CMM1, Clavibacter michiganensis M1; BLI, Bifidobacterium longum subsp. infantis ATCC 15697; BB, Bifidobacterium breve ATCC 15700; BL, Bifidobacterium
longum ATCC BAA-999; LC, Lactobacillus casei ZX147; LB, Lactobacillus plantarum CGMCC 1.19; CB, Clostridium butyricum YX038.
* Corresponding authors at: State Key Laboratory of Food Science and Technology, China-Canada Joint Laboratory of Food Science and Technology (Nanchang),
Nanchang University, 235 Nanjing East Road, Nanchang 330047, Jiangxi, People’s Republic of China (X. Fu). College of Food Science and Engineering, Ocean
University of China, No. 5 Yushan Road, Qingdao 266003, Shandong, People’s Republic of China (H. Mou).
E-mail addresses: 18227591863@163.com (M. Xiao), m13992623179@163.com (X. Ren), 2970565907@qq.com (J. Cui), lirong81@163.com (R. Li),
ocean2013@126.com (Z. Liu), zhulin5566@126.com (L. Zhu), kongqing@ouc.edu.cn (Q. Kong), luna_9303@163.com (X. Fu), mousun@ouc.edu.cn (H. Mou).

https://doi.org/10.1016/j.foodchem.2021.132001
Received 18 September 2021; Received in revised form 14 November 2021; Accepted 18 November 2021
Available online 4 January 2022
0308-8146/© 2022 Elsevier Ltd. All rights reserved.
M. Xiao et al. Food Chemistry 377 (2022) 132001

leading to various negative effects on infant health, including allergy a reference strategy for exploring potential fucosylated oligosaccharides
and childhood obesity (Wampach et al., 2018). Therefore, discovering a with functional abilities similar to HMOs.
prebiotic that can promote the proliferation of probiotics in the infant
gut to reduce the occurrence of early infant diseases is of great signifi­ 2. Materials and methods
cance. Fucosylated oligosaccharides are common in the human intesti­
nal tract and mostly located at the end of mucin glycoproteins on the cell 2.1. Bacterial strains and chemicals
surface (Becker & Lowe, 2003). These fucosylated oligosaccharides
coated with mucins are considered to be adhesion targets for many in­ The strains used in the present study, CMM1 isolated from decaying
testinal bacteria (Stahl et al., 2011). Fucose and fucosylated oligosac­ tomato leaves, Lactobacillus casei ZX147 (LC), and Clostridium butyricum
charides also serve as chemoattractants and carbon sources for gut YX038 (CB) isolated from feces of infants, were provided by the Applied
bacteria. Human milk oligosaccharides (HMOs) are oligosaccharides Microbiology Laboratory, Ocean University of China. Bifidobacterium
that exist only in human milk (Stahl et al., 2011). It has been reported longum subsp. infantis ATCC 15697 (BLI), Bifidobacterium breve ATCC
that probiotic bacteria abundant in early life, such as bifidobacterial 15700 (BB), and Bifidobacterium longum ATCC BAA-999 (BL) were pur­
strains, are capable of recognizing and metabolizing fucosylated HMOs chased from the American Type Culture Collection (ATCC, Rockville,
(Chambers & Townsend, 2020). 2′ -Fucosyllactose (2′ -FL) is the most MD, USA). Lactobacillus plantarum CGMCC 1.19 (LP) was purchased
abundant oligosaccharide in human milk and it has been proved to from the China General Microbiological Culture Collection Center
regulate infant intestinal microbiota (Zabel et al., 2020). However, (CGMCC, Beijing, China). Bio-gel P4 was acquired from Bio-Rad (Her­
complex preparation methods have limited the wide application and cules, CA, USA). Standards for short-chain fatty acids (SCFAs) (lactic
further studies of 2′ -FL. Nevertheless, fucosylated oligosaccharides have acid, propionic acid, butyric acid, formic acid, acetic acid, and iso­
also been prepared via hydrolyzing fucoidan extracted from brown algae caproic acid) and monosaccharides (glucose, galactose, fucose,
by fucoidanase in recent years (Silchenko et al., 2018). mannose, rhamnose, glucuronic acid, and galacturonic acid) were pur­
Microbial exopolysaccharides (EPS) containing fucose are a pro­ chased from Sigma-Aldrich Co., Ltd. (St. Louis, MO, USA). Lactose,
spective source of fucose and fucosylated oligosaccharides. The ability mannotriose, mannotetraose, mannose pentasaccharide, and mannose
to produce EPS is an attribute widely found among various microbial hexasaccharide were purchased from Solarbio Technology Co., Ltd.
species, including bacteria, microalgae, and fungi (Vanhooren & Van­ (Beijing, China). Thin-layer chromatography (TLC) silica gel was pur­
damme, 1999). Microbial synthesis of EPS is amenable to control the chased from Merck (Darmstadt, Germany). All reagents were of
process conditions and it is unaffected by environmental factors (Alves analytical grade.
et al., 2010). However, only few bacterial EPS are commercially avail­
able, such as xanthan, gellan gum, and hyaluronic acid (Freitas, Alves, & 2.2. Extraction and purification of CMM1 EPS
Reis, 2011). Fucose-containing EPS have been extracted from several
bacterial strains (Vanhooren & Vandamme, 1999). Colanic acid is a type The CMM1 strain preserved in 30% (v/v) glycerin was inoculated
of EPS composed of fucose, glucose, galactose, and glucuronic acid, into a 250-mL shake flask containing 100 mL YPG medium (yeast extract
which is generally produced by members of Enterobacteriaceae, 5 g/L, peptone 10 g/L, glucose 5 g/L; pH 7.0–7.2) and cultivated for 24 h
including Escherichia and Klebsiella spp. (Rättö et al., 2006). Clavan, at 27 ◦ C. The fermentation medium used for CMM1 EPS production was
which is composed of glucose, galactose, fucose, and pyruvate, can be prepared according to the method described by Bulk et al. (1991), with
produced by Clavibacter michiganensis strains (Bulk, Zevenhuizen, Cor­ the following modifications: glucose 20 g/L, yeast extract 4 g/L, CaCO3
dewener, & Dons, 1991). Fucogel, a fucose-rich type of EPS extracted 5 g/L, (NH4)2SO4 0.3 g/L, KH2PO4 1.3 g/L, and K2HPO4 1.7 g/L; pH
from Klebsiella pneumoniae I-1507, is used in the cosmetic industry 7.0–7.2. The CMM1 strain was inoculated into 100 mL of the fermen­
(Guetta, Mazeau, Auzely, Milas, & Rinaudo, 2003). The EPS produced tation broth at a concentration of 1% (v/v) in a 250-mL conical flask.
by Enterobacter sp. A47, FucoPol, also have high fucose content The cultures were incubated at 27 ◦ C for 5 days. After the incubation
(Antunes, Freitas, Sevrin, Grandfils, & Reis, 2017). period, in order to remove the bacteria and insoluble substances, the
To obtain fucosylated oligosaccharides, fucose-containing EPS must fermentation broth was centrifuged at 6000 ×g for 10 min at 4 ◦ C. After
be further depolymerized. Over the last several years, hydrochloric acid centrifugation, the supernatant was evaporated to 1/4 of the original
(HCl) and trifluoroacetic acid (TFA) have been used to degrade EPS to volume at 50 ◦ C. Three volumes of 95% ethanol were then added, and
obtain oligosaccharides. The hydrolysis of EPS by HCl is usually carried the supernatant was allowed to stand overnight. The supernatant was
out using a high concentration of HCl at high temperature. The HCl centrifuged at 8000 ×g for 10 min at 4 ◦ C to collect the precipitate,
acidolysis process is simple and easy to perform; however, the produc­ which was washed with 95% ethanol. The precipitate was redissolved in
tion of oligosaccharides with molecular weights greater than those of 25 mL of ultrapure water, and the protein in the precipitate was removed
tetrasaccharides is not an easy task (Jang et al., 2005). Hydrolysis by using the Sevag method (Sevag, Lackman, & Smolens, 1938). The su­
TFA was found to be mild, and TFA could be removed via freeze-drying pernatant was dialyzed using a 10 kDa dialysis bag at 4 ◦ C for 72 h.
or vacuum distillation after hydrolysis, thereby avoiding salt production Finally, CMM1 EPS was obtained after vacuum freeze-drying.
during the traditional neutralization steps (Fels, Jakob, Vogel, & Wefers,
2018; Maalej et al., 2014). 2.3. Optimization of the acidolysis conditions for CMM1 EPS
Fucose-containing EPS extracted from C. michiganensis is a prospec­
tive source of fucose and fucosylated oligosaccharides. In the current Fucosylated oligosaccharides were prepared by degrading CMM1
study, C. michiganensis M1 (CMM1) was used to prepare fucose- EPS via acid hydrolysis (acidolysis). To select the degradation conditions
containing EPS, and fucosylated oligosaccharides were obtained by resulting in degradation products with single-component and low
acid hydrolysis of fucose-containing EPS. The acidolysis conditions for monosaccharide contents, the acidolysis conditions need to be opti­
CMM1 EPS were optimized to select the degradation conditions result­ mized. In the present study, the acidolysis reaction was carried out by
ing in degradation products with single-component and low mono­ treating CMM1 EPS with HCl (0.2 mol/L HCl or 0.1 mol/L HCl) or TFA
saccharide contents. The structure of the glucofucobiose (GF) obtained (0.2 mol/L TFA or 0.1 mol/L TFA) at 80 ◦ C or 90 ◦ C, respectively. The
by acidolysis was analyzed and the in vitro prebiotic activity of GF on six material to liquor ratio was set to 1:10, and the acidolysis reaction was
probiotic strains was compared with that of 2′ -FL and CMM1 EPS. By performed in a water bath shaker for 8 h. Samples were taken every 2 h
introducing a novel preparation method for fucosylated oligosaccha­ to characterize the variations of carbohydrate degradation and hydro­
rides and comparing the prebiotic activity of 2′ -FL and GF prepared lysate production via high-performance liquid chromatography (HPLC)
using the bacterial fermentation method, the present study may serve as and TLC. Before HPLC and TLC analyses, the pH of the acidolysis

2
M. Xiao et al. Food Chemistry 377 (2022) 132001

solution was adjusted to 7.0 using 1 mol/L NaOH. The hydrolysate was pyridine. The mixture was heated at 60 ◦ C for 1 h, and arylthiocyanate
then mixed with three volumes of ethanol to remove high molecular was added to the mixture. The reaction was allowed to proceed for 1 h
weight precipitates, and the supernatant was collected by centrifugation (Tanaka, Nakashima, Ueda, Tomii, & Kouno, 2007). Then, the deriva­
at 5000 ×g for 10 min and freeze-dried. tization of these samples was performed as described in Fu, Li, Zhang, Li,
The above lyophilized hydrolysate was solubilized at 5 mg/mL and and Mou (2018). The monosaccharide composition was determined
then separated on a Shodex OH pak SB 802.5 HQ column (8 mm × 300 using an HPLC system equipped with a ZORBAX 300 XDB-C18 column
mm, 6 μm, Shodex Technology) with a refractive index detector. The (4.6 mm × 250 mm, 5 μm, Agilent Technologies). The mobile phase
mobile phase was 100% water, the flow rate was 0.85 mL/min, and the consisted of mobile phase A (acetonitrile) and mobile phase B (0.05
temperature of the column oven was set at 50 ◦ C. mol/L KH2PO4 buffer solution) at a ratio of 17:83. The UV detection
The developing agents were N-butanol, acetic acid, and water at wavelength was 245 nm, and the flow rate was 1 mL/min. The sample
1:1:1. The chromogenic agent was composed of 1 mL HCl, 3 mL aniline, injection volume was 20 µL.
4 g diphenylamine, 10 mL 85% H3PO4, and 100 mL acetone. The loading
volume was 3 μL (5 mg/mL lyophilized hydrolysate solution). Glucose, 2.5.3. Nuclear magnetic resonance spectroscopy analysis
galactose, and fucose were used as the standards. Samples (15 mg) were exchanged twice in 99.9% D2O with inter­
mediate lyophilization, and finally dissolved in 500 µL D2O. One-
2.4. Preparation and purification of fucosylated oligosaccharides dimensional nuclear magnetic resonance (NMR) spectra (1H, 13C,
DEPT 90◦ and 135◦ ) and two-dimensional NMR spectra (1H–1H COSY,
1
The CMM1 EPS was degraded under the optimized conditions (0.2 H–1H TOCSY, 1H–13C HSQC, and 1H–13C HMBC) were recorded on an
mol/L TFA, 90 ◦ C, 2 h). Thereafter, the acidolyzed EPS solution was Agilent DD2-500 spectrometer (Agilent Technologies), with a frequency
cooled to 25 ◦ C. The remaining TFA in the solution was removed via a of 500 MHz and a temperature of 25 ◦ C. 1H and 13C chemical shifts are
dedicated electrodialysis (Sichuan Lvwo Environmental & Energy Co., expressed in ppm in reference to internal acetone (2.225 ppm for 1H and
Ltd) with the current and voltage set at 3 A and 18 V, respectively. The 31.07 ppm for 13C). All spectra were analyzed using MestReNova 6.1
acidolysis products were centrifuged at 5000 ×g for 10 min at 4 ◦ C. (Mestrelabs Research SL, Santiago de Compostela, Spain), with Whit­
Thereafter, 95% ethanol (three times the volume of supernatant) was taker Smoother baseline correction.
added to the supernatant to precipitate the high molecular weight
substances at 4 ◦ C overnight. Centrifugation was performed at 5000 ×g 2.6. In vitro prebiotic activity of fucosylated oligosaccharides
for 10 min at 4 ◦ C to collect the supernatant. Finally, crude fucosylated
oligosaccharides were obtained via vacuum lyophilization. MRS broth medium containing 0.50 g/L cysteine hydrochloride was
A Bio-gel P4 column was used for the purification of the acidolysis used to activate three bifidobacterial strains (BLI, BB, and BL) and two
products. The mobile phase was 0.2 mol/L NH4HCO3, the flow rate was lactobacillus strains (LC and LP). RCM medium, purchased from Qing­
0.15 mL/min, and the loading volume was 500 µL (50 mg/mL crude dao Hope Technology Co., Ltd. (Qingdao, China), was used to activate
oligosaccharides powder). The eluent was collected in 40 tubes with 1.5 CB. Each strain was activated three times at 37 ◦ C prior to fermentation
mL of eluent per tube. The total sugar content of each tube was then experiments. In vitro fermentation was performed in a simulated intes­
determined according to DuBois, Gilles, Hamilton, Rebers, and Smith tinal medium, which was prepared according to the methods described
(1956), with slight modifications. Briefly, 200 µL of sample, 400 µL of by Fu et al. (2020), with slight modifications. The amounts of yeast
6% phenol solution, and 2 mL of concentrated sulfuric acid were mixed extract and peptone were adjusted to 1 g/L and 2 g/L, respectively. EPS,
evenly and reacted for 30 min. Thereafter, their optical density at 490 2′ -FL, and GF were added to anaerobic bottles with 50 mL of simulated
nm (OD490) was determined, and the total sugar content was calculated intestinal medium at a final concentration of 10 mg/mL. In vitro
using a glucose standard curve (Y = 0.0052 X - 0.0706, R2 = 0.9999). fermentation was performed at 37 ◦ C for 48 h.
The elution curve was plotted with the elution time as the abscissa and Fermentation samples were collected at 0, 12, 24, 36, and 48 h. The
the total sugar content as the ordinate. After purification, acid hydro­ samples were centrifuged at 13000 ×g for 10 min at 4 ◦ C in a refriger­
lysates were collected for structural characterization. ated centrifuge (Thermo Fisher Scientific, Waltham, MA, USA). The
bacterial precipitation was eluted with phosphate-buffered saline buffer
2.5. Structural characterization of the fucosylated oligosaccharides (8 g/L NaCl, 0.2 g/L KCl, 1.44 g/L KH2PO4 and 0.24 g/L K2HPO4; pH
7.2) and the OD600 of this eluent was determined. The centrifuged su­
2.5.1. Mass spectrometry analysis pernatant was collected for further pH, SCFAs, and TLC analyses. The
The molecular weight of the oligosaccharide samples was identified SCFAs content was determined as described previously (Wei et al.,
by electron spray ionization mass spectrometry (ESI-MS) and electron 2020). The TLC analysis procedures were the same as those described in
spray ionization in tandem with collision-induced dissociation tandem Section 2.3.
MS (ESI-CID-MS), which was carried out using an Agilent 1290 Infinity
II system and 6460 triple quadrupole MS (Agilent Technologies, Santa 3. Results and discussion
Clara, CA, USA). The elution was performed using 50% acetonitrile and
50% water at a flow rate of 0.35 mL/min. The two mass spectrometers 3.1. Optimization of the acidolysis conditions on CMM1 EPS
were operated in positive ion mode with a capillary voltage of 3.5 kV
and a nebulizer pressure of 40 psi. The collision energy was set to 15–60 Different acidolysis conditions were used to degrade the CMM1 EPS,
eV, and the collision gas was nitrogen. All spectra were analyzed using and samples were taken every 2 h for TLC and HPLC analyses. For the
MassHunter Qualitative Analysis B.07.00 (Agilent Technologies). EPS samples treated with an equal amount of acid concentration, the
color of monosaccharide spots gradually deepened with an increase in
2.5.2. Monosaccharide composition analysis acidolysis temperature and time (Fig. S1). EPS were degraded more
Seven monosaccharide standards and one lactose internal standard completely by HCl than by TFA of the same concentration under the
were combined into a mixed standard solution with a concentration of 1 same acidolysis temperature and time conditions. Thus, more mono­
mg/mL, respectively. Ten milligrams of fucosylated oligosaccharides saccharides were produced during the degradation of EPS by HCl than
were dissolved in 2 mL of 2 mol/L TFA, reacted at 110 ◦ C for 4 h, and during the degradation of EPS by TFA.
then cooled to 25 ◦ C. After removing TFA, the pH was adjusted to 7.0 The effects of the acidolysis conditions on the molecular weight
using a solution of NaOH (6 mol/L). Thereafter, the hydrolysate and distribution of CMM1 EPS are shown in Fig. 1 and Table S1. The CMM1
monosaccharide standards were mixed with L-cysteine methyl ester and EPS was effectively hydrolyzed, generating substances with molecular

3
M. Xiao et al. Food Chemistry 377 (2022) 132001

Fig. 1. GPC spectra of the degradation of CMM1 EPS by hydrochloric acid (HCl) and trifluoroacetic acid (TFA). (A) The molecular weight distribution of CMM1 EPS
degraded by 0.2 mol/L TFA (A), 0.1 mol/L TFA (B), 0.2 mol/L HCl (C), and 0.1 mol/L HCl (D) at 80 or 90 ◦ C for 2, 4, 6, and 8 h, respectively. Mannose hex­
asaccharide (1), mannose hexasaccharide (2), mannotetraose (3), mannotriose (4), lactose (5), glucose (6), and fucose (7) were used as standards.

weight less than 1000 Da. However, different types and concentrations spectrum when the EPS was degraded by 0.2 mol/L TFA at 90 ◦ C for 2 h,
of acids and different acidolysis temperature and time led to different which was consistent with the TLC analysis (Fig. S1). The retention time
degrees of CMM1 EPS degradation. CMM1 EPS was degraded more of the peak was close to that of lactose, suggesting that the peak may
thoroughly under higher concentration of acid and acidolysis tempera­ correspond to a disaccharide. Considering the preparation of fucosylated
ture and longer acidolysis time. There was one main peak in the oligosaccharides with single-component and low monosaccharide

4
M. Xiao et al. Food Chemistry 377 (2022) 132001

contents, TFA (0.2 mol/L) was selected to degrade CMM1 EPS at 90 ◦ C which was expected to be [M + Na − 180]+. Altogether, the above re­
for 2 h. sults confirmed that peak P2 was a disaccharide with a molecular weight
of 326 Da.
3.2. Purification of the acidolysis products
3.3.2. Monosaccharide composition of peak P2
After acidolysis of CMM1 EPS, a Bio-gel P4 gel column was used to The mixed standard monosaccharides and peak P2 were analyzed by
separate and purify the resulting products. As shown in Fig. S2, three HPLC after derivatization using lactose as the internal standard (Fig. S3).
peaks, P1, P2, and P3, were obtained, with P2 as the main peak. The Based on the results, peak P2 comprised D-glucose and L-fucose with an
fractions corresponding to the three peaks were collected and purified approximate molar ratio of 1:1 according to the peak area. Based on the
for subsequent analyses. results of monosaccharide composition and MS, peak P2 was presumed
to be a disaccharide composed of D-glucose and L-fucose. 2′ -FL is the
most abundant oligosaccharide in human milk, and it improves immu­
3.3. Structural characterization
nity and promotes brain development in infants (Thongaram, Hoef­
linger, Chow, & Miller, 2017). The monosaccharide composition of 2′ -FL
3.3.1. Mass spectrometry analysis of the three peaks
was D-glucose and L-fucose, which is the same as that of the disaccharide
After several separation and purification steps, the three peak frac­
obtained in the present study. To identify the accurate structure of the
tions were analyzed by positive ion ESI-MS to determine the molecular
disaccharide, further analysis using NMR spectroscopy was performed,
weight. For P1, only one peak was observed at m/z 511.1, as shown in
which was beneficial for further assessments of its biological activities.
Fig. 2A, suggesting that peak P1 may be a trisaccharide (511.1 = 180 +
180 + 164 – 36 + Na). The MS result of peak P2 is shown in Fig. 2B; two
3.3.3. NMR analysis of peak P2
main peaks, at m/z 327.3 and m/z 349.1, which were [M + H]+ and [M
The signals of anomeric protons in the 1H NMR spectrum could be
+ Na]+, respectively, were observed, indicating that peak P2 may be a
easily identified between δ 4.3 ppm and 5.5 ppm (Hehemann, Boraston,
disaccharide. As shown in Fig. 2C, for P3, two peaks (m/z 165.5 and
& Czjzek, 2014). The 1H NMR spectrum of GF is displayed in Fig. 3A. The
181.1), which were assigned to [F + H]+ and [G + H]+, were observed,
peaks around 1.5 ppm (1.517 ppm and 1.482 ppm) were assigned to the
indicating that peak P3 was composed of monosaccharides with mo­
chemical shifts of C6 in L-fucose, suggesting that L-fucose may be at the
lecular weights of 180 and 164 Da. Owing to the low content of peaks P1
reducing end of the disaccharide. There were three terminal protons
and P3, peak P2 was mainly collected for structural characterization and
(5.407, 4.780, and 4.669 ppm) with relative integral ratios of
activity studies. Further, ESI-CID-MS was performed to verify the
0.37:0.64:1.02, which were assigned to the H-1 of α-L-fucose, β-L-fucose,
composition of peak P2, as shown in Fig. 2D. After ionization, the peak
and β-D-glucose, respectively. The coexistence of α- and β-anomers was
at m/z 349.1 was knocked out, resulting in a charge peak at m/z 169.2,

Fig. 2. Positive ion mode in the ESI-MS/CID-MS spectra of acidolysis products after purification with a Bio-gel P4 column. (A) ESI-MS spectrum of peak P1. (B) ESI-
MS spectrum of peak P2. (C) ESI-MS spectrum of peak P3. (D) ESI-CID-MS spectrum of peak P2 at m/z 349.1.

5
M. Xiao et al. Food Chemistry 377 (2022) 132001

Fig. 3. 1D NMR spectra of peak P2. (A) 1H spectrum; (B) 13


C spectrum; (C) DEPT 135◦ ; and (D) DEPT 90◦ .

6
M. Xiao et al. Food Chemistry 377 (2022) 132001

attributed to the random rearrangement of the anomeric protons during 77.010 ppm), and the C-1 of β-D-glucose was linked to the H-4 of α-L-
the degradation process (Hehemann et al., 2014; Ulaganathan et al., fucose (δ 3.566 ppm). In summary, the structure of the disaccharide, GF,
2017). was identified as D-Glcp-β-(1→4)-L-Fucp.
The resonance scope of anomeric carbon C-1 was distributed in the
range δ 90–110 ppm and three C-1 signals (δ 104.719 ppm, δ 97.760
ppm, and δ 93.817 ppm), assigned to the C-1 of β-D-glucose, β-L-fucose, 3.4. In vitro prebiotic activity
and α-L-fucose, respectively, were observed in Fig. 3B. In the DEPT 135◦
spectrum, the peaks of –CH and –CH3 were upward whereas those of 3.4.1. Utilization profile of GF, 2′ -FL, and EPS by six probiotic strains and
–CH2 were downward (Wu, Cui, Eskin, Goff, & Nikiforuk, 2011). As their growth
shown in Fig. 3C, 62.174 ppm was inferred to correspond to the C-6 of Intestinal microbiota commonly demonstrate various carbohydrate
β-D-glucose. According to previous studies, the chemical shifts of 17.128 utilization preferences and metabolic behaviors during the fermentation
ppm and 17.058 ppm were assigned to the C-6 of β-L-fucose and α-L- of non-digestible carbohydrates (Li et al., 2020). The growth charac­
fucose, respectively (Xiao et al., 2021). teristics of the six probiotic strains (BLI, BB, BL, LC, LP, and CB) on
As shown in Fig. 4B, the three cross-peaks of 5.407/3.952 ppm, different carbon sources (GF, 2′ -FL, or EPS) were evaluated by
5.407/4.074 ppm, and 5.407/3.566 ppm were considered to represent measuring the absorbance of cultures over 48 h (Fig. 5). Similar to the
the associations between H-1 and other H signals of α-L-fucose. The three blank controls, the proliferative effect of EPS on the six strains was not
cross-peaks of 4.780/3.631 ppm, 4.780/3.813 ppm, and 4.780/4.160 distinct. The addition of GF significantly promoted the growth of all
ppm represented the connections between H-1 and other H signals of β-L- strains compared to the blank controls. For the bifidobacterial strains,
fucose. Three other cross-peaks, 4.669/3.655 ppm, 4.669/3.924 ppm, GF exhibited a better promoting effect than 2′ -FL on the growth of BB
and 4.669/4.091 ppm, were assigned to the connections of H-1 with and BL. Although the proliferation effect of 2′ -FL on BLI was better than
other H signals of β-D-glucose. According to the 1H–1H COSY and 1H–1H that of GF during 24–36 h, the proliferation effect of 2′ -FL on BLI
TOCSY spectra, the H signals of β-L-fucose, β-D-glucose, and α-L-fucose decreased rapidly from 36 h to 48 h. In the present study, GF showed the
were determined. Thereafter, every C signal was identified using 1H–13C best prebiotic effect on LC, with the highest OD600 value of 0.668 at 24 h,
HSQC (Fig. 4C), and all 1H and 13C signals are summarized in Table S2. followed by CB, BL, BLI, BB, and LP. In addition, 2′ -FL was capable of
The HMBC spectrum commonly shows the correlation between C and promoting the proliferation of BLI, BL, and LPM, but it had no obvious
H separated by several chemical bonds (Cescutti, Cuzzi, Herasimenka, & effect on the proliferation of the other three strains.
Rizzo, 2013). According to the 1H–13C HMBC spectrum (Fig. 4D), the TLC was used to determine whether the observed growth patterns
cross peaks between H-1 of β-D-glucose (δ 4.669 ppm) and C-4 of β-L- matched the consumption of the three sugars (Fig. S4). The variations in
fucose (δ 81.575 ppm) indicated that the linkage of the disaccharide was spots revealed that GF was completely utilized by BL, BB, LC, and CB
β-D-Glcp-(1→4)-β-L-Fucp. In addition, the C-1 of β-D-glucose (δ 104.719 within 24 h, and utilized by BL and LP within 48 h, indicating that the
ppm) and H-4 of β-L-fucose (δ 4.160 ppm) were connected as indicated in three bifidobacteria, two lactobacilli, and Clostridium butyricum could
Fig. 4D. The H-1 of β-D-glucose was linked to the C-4 of α-L-fucose (δ produce β-(1→4) glucosidase to hydrolyze fucosylated disaccharide
during their metabolism processes. 2′ -FL was completely utilized by BLI

Fig. 4. 2D NMR spectra of peak P2. (A) 1H–1H COSY; (B) 1H–1H TOCSY; (C) 1H–13C HSQC; and (D) 1H–13C HMBC.

7
M. Xiao et al. Food Chemistry 377 (2022) 132001

Fig. 5. In vitro prebiotic activity of six selected bacterial strains (Bifidobacterium longum subsp. infantis ATCC 15697 (BLI), Bifidobacterium breve ATCC 15700 (BB),
Bifidobacterium longum ATCC BAA-999 (BL), Lactobacillus casei ZX147 (LC), Lactobacillus plantarum CGMCC 1.19 (LP), and Clostridium butyricum YX038 (CB)) on
different carbohydrate was assessed by cell proliferation in simulated intestinal mediums with glucofucobiose (GF), 2′ -fucosyllactose (2′ -FL), and exopolysaccharides
(EPS) as carbon sources or with no additive (control) at 0, 12, 24, 36 and 48 h. The results are presented as the mean values ± standard deviations (n = 3).

and BL within 24 h and 48 h, respectively. The points at 48 h were (Teixeira et al., 2013). In addition to their health benefits, SCFAs are also
significantly lower than those at 0 h during the fermentation processes considered to be associated with increased satiety caused by carbohy­
of LC and LP, suggesting that these strains were able to utilize 2′ -FL, drate consumption (Sleeth, Thompson, Ford, Zac-Varghese, & Frost,
although the effect was not obvious. However, 2′ -FL was not degraded 2010). The adhesion of bacteria to the human intestinal tract is closely
by BB and CB during the 48 h of in vitro fermentation. None of the six related to the pH of the intestinal environment, which leads to variations
selected strains had the ability to metabolize and utilize EPS within 48 h, in the structure and physiological function of bacteria (Wu et al., 2018).
which was consistent with the variations in the OD600 value. The content of SCFAs was measured as shown in Fig. 6. Not sur­
Bifidobacterial and lactobacilli strains have various genes encoding prisingly, there was no significant production of SCFAs by the six bac­
glycoside hydrolases, thus facilitating the metabolism and utilization of terial strains in the simulated intestinal medium with EPS or without an
different carbohydrates. The proliferation effect of 2′ -FL on BLI was additional carbon source (data not shown). 2′ -FL could promote the
better than that of GF in the first 24 h and then decreased rapidly from proliferation of bifidobacterial strains, and it has been shown to promote
36 h to 48 h, which was attributed to the low pH of the medium, leading the production of SCFAs, including formate, acetate, and lactate (Zabel
to bacterial death. According to the results, although BLI can proliferate et al., 2020). In the present study, the main SCFAs generated during the
better in short-term culture in the medium containing 2′ -FL, it can sur­ fermentation of 2′ -FL by BLI, BB, BL, LC, and LP were lactate and acetate,
vive more stably in the medium containing GF. A recent study evaluated with a small amount of formate. The total SCFAs production by BB, BL,
the ability of 16 bifidobacterial strains and 19 lactobacilli strains to LC, LP, and CB during the fermentation of GF was 2.2-, 1.8-, 2.9-, 1.6-,
ferment 2′ -FL, and the findings showed that 2′ -FL was only metabolized and 2.9-fold than that produced during the fermentation of 2′ -FL,
by four bifidobacterial strains and had no significant effect on the pro­ respectively. And GF stimulated CB to produce 3.7-fold butyrate than 2′ -
liferation of other bifidobacterial and lactobacilli strains (Salli et al., FL. Moreover, a low level of butyrate was produced during the
2021). Similarly, the present study also demonstrated that 2′ -FL is highly fermentation of GF by BB and LP, which was not observed during the
selective in promoting the proliferation of different probiotic strains fermentation of 2′ -FL. Recent studies have reported that butyrate plays
while GF can be utilized by all the examined strains, suggesting that its an important role in maintaining the stability of the intestinal environ­
effect is less selective and more conducive to the proliferation of these ment and providing an energy source for colon cells (Hamer, Jonkers,
strains. Venema, Vanhoutvin, & Brummer, 2008). GF, as a sugar capable of
promoting butyrate production by probiotics, has prospects as butyro­
3.4.2. Changes in the pH value and SCFAs content during fermentation genic prebiotics.
As shown in Fig. S5, GF was consumed by the six probiotic strains at
48 h, which decreased the pH of the medium. Only BLI, BL, and LP were 4. Conclusion
capable of decreasing the pH of the simulated intestinal medium con­
taining 2′ -FL. The pH values of the medium with EPS as a carbon source Fucosylated oligosaccharides in human milk, including 2′ -FL, 3-fuco­
or without an additional carbon source were not changed significantly. syllactose, and difucosyllactose, play an important role in promoting
The decrease in pH during fermentation is usually caused by the accu­ intestinal health and reducing allergic reactions. Bacterial EPS are raw
mulation of some metabolites during the growth and metabolism of materials for the exploration of diverse functional oligosaccharides and
bacteria, such as SCFAs, which have beneficial effects on human health polysaccharides. To date, the preparation of fucosylated oligosaccha­
(Wei et al., 2020; Fu, Liu, Zhu, Mou, & Kong, 2019). SCFAs are a class of rides using microbial fermentation methods has only been performed in
organic fatty acids with one to six carbon atoms. The most important a few studies. In the present study, CMM1 EPS were prepared by the
SCFAs for human metabolism include acetate, propionate, and butyrate bacterial fermentation method. Further, based on the diversity of the

8
M. Xiao et al. Food Chemistry 377 (2022) 132001

Fig. 6. Concentration of butyrate, acetate, formate, and lactate in simulated intestinal medium containing GF, 2′ -FL and EPS as carbon sources during the
fermentation of six selected bacterial strains at 0, 12, 24 and 48 h. Results are presented as the mean values ± SD (n = 3).

degraded components and monosaccharide content, the effects of physicochemical, texture, oxidative stability, and sensory property of reduced fat
emulsion type sausage. Meat Science, 147, 135–143.
different acidolysis conditions on CMM1 EPS were evaluated. Finally,
Alves, V. D., Freitas, F., Torres, C. A. V., Cruz, M., Marques, R., Grandfils, C., et al.
TFA (0.2 mol/L) was applied to degrade CMM1 EPS at 90 ◦ C for 2 h to (2010). Rheological and morphological characterization of the culture broth during
prepare fucosylated oligosaccharides. Under these conditions, GF, with exopolysaccharide production by Enterobacter sp. Carbohydrate Polymers, 81(4),
the structure D-Glcp-β-(1→4)-L-Fucp, was successfully obtained. In 758–764.
Antunes, S., Freitas, F., Sevrin, C., Grandfils, C., & Reis, M. A. M. (2017). Production of
addition, the present study evaluated the prebiotic effects of GF and 2′ - FucoPol by Enterobacter A47 using waste tomato paste by-product as sole carbon
FL on six probiotic strains. The results suggested that GF was capable of source. Bioresource Technology, 227, 66–73.
promoting the proliferation of three bifidobacterial, two lactobacilli, Bäckhed, F., Roswall, J., Peng, Y., Feng, Q., Jia, H., Kovatcheva-Datchary, P., et al.
(2015). Dynamics and stabilization of the human gut microbiome during the first
and one Clostridium butyricum strains, and of stimulating these probiotic year of life. Cell Host & Microbe, 17(5), 690–703.
strains to produce more SCFAs than 2′ -FL. Overall, GF, with a new Becker, D. J., & Lowe, J. B. (2003). Fucose: Biosynthesis and biological function in
structure, was successfully obtained by degrading bacterial EPS, and the mammals. Glycobiology, 13(7), 41R–53R.
Bulk, R., Zevenhuizen, L., Cordewener, J., & Dons, J. (1991). Characterization of the
present study may serve as a reference strategy for exploring potential extracellular polysaccharide produced by Clavibacter michiganensis subsp.
fucosylated oligosaccharides. Although well-designed studies with in michiganensis. Physiology and Biochemistry, 81(6), 619–623.
vitro and in vivo prebiotic activities are needed, our data point out a new Cescutti, P., Cuzzi, B., Herasimenka, Y., & Rizzo, R. (2013). Structure of a novel
exopolysaccharide produced by Burkholderia vietnamiensis, a cystic fibrosis
direction for the subsequent application of GF and fucosylated opportunistic pathogen. Carbohydrate Polymers, 94(1), 253–260.
oligosaccharides. Chambers, S. A., & Townsend, S. D. (2020). Like mother, like microbe: Human milk
oligosaccharide mediated microbiome symbiosis. Biochemical Society Transactions, 48
(3), 1139–1151.
Declaration of Competing Interest DuBois, M., Gilles, K. A., Hamilton, J. K., Rebers, P. A., & Smith, F. (1956). Colorimetric
method for determination of sugars and related substances. Analytical Chemistry, 28
(3), 350–356.
The authors declare that they have no known competing financial Fels, L., Jakob, F., Vogel, R. F., & Wefers, D. (2018). Structural characterization of the
interests or personal relationships that could have appeared to influence exopolysaccharides from water kefir. Carbohydrate Polymers, 189, 296–303.
Freitas, F., Alves, V. D., & Reis, M. A. M. (2011). Advances in bacterial
the work reported in this paper. exopolysaccharides: From production to biotechnological applications. Trends in
Biotechnology, 29(8), 388–398.
Fu, X. D., Li, R., Zhang, T., Li, M., & Mou, H. J. (2018). Study on the ability of partially
Acknowledgment
hydrolyzed guar gum to modulate the gut microbiota and relieve constipation.
Journal of Food Biochemistry, 43(2), 1–14.
This work was supported by the National Natural Science Foundation Fu, X., Liu, Z., Zhu, C., Mou, H., & Kong, Q. (2019). Nondigestible carbohydrates,
of China (31872893) and the China Postdoctoral Science Foundation butyrate, and butyrate-producing bacteria. Critical reviews in food science and
nutrition, 59(sup1), S130–S152.
(2021M701547). Fu, X., Wei, X., Xiao, M., Han, Z., Secundo, F., & Mou, H. (2020). Properties of
hydrolyzed guar gum fermented in vitro with pig fecal inocula and its favorable
impacts on microbiota. Carbohydrate Polymers, 237, 116116. https://doi.org/
Appendix A. Supplementary data 10.1016/j.carbpol.2020.116116
Guetta, O., Mazeau, K., Auzely, R., Milas, M., & Rinaudo, M. (2003). Structure and
Supplementary data to this article can be found online at https://doi. properties of a bacterial polysaccharide named Fucogel. Biomacromolecules, 4(5),
1362–1371.
org/10.1016/j.foodchem.2021.132001.
Hamer, H. M., Jonkers, D., Venema, K., Vanhoutvin, S., & Brummer, R. J. (2010). Review
article: the role of butyrate on colonic function. Alimentary Pharmacology &
References Therapeutics, 27(2), 104–119.
Hehemann, J.-H., Boraston, A. B., & Czjzek, M. (2014). A sweet new wave: Structures and
mechanisms of enzymes that digest polysaccharides from marine algae. Current
Abbasi, E., Amini Sarteshnizi, R., Ahmadi Gavlighi, H., Nikoo, M., Azizi, M. H., & Opinion in Structural Biology, 28, 77–86.
Sadeghinejad, N. (2019). Effect of partial replacement of fat with added water and
tragacanth gum (Astragalus gossypinus and Astragalus compactus) on the

9
M. Xiao et al. Food Chemistry 377 (2022) 132001

Jang, J.-H., Hia, H. C., Ike, M., Inoue, C., Fujita, M., & Yoshida, T. (2005). Acid hydrolysis Thongaram, T., Hoeflinger, J. L., Chow, JoMay, & Miller, M. J. (2017). Human milk
and quantitative determination of total hexosamines of an exopolysaccharide oligosaccharide consumption by probiotic and human-associated bifidobacteria and
produced by Citrobacter sp. Biotechnology Letters, 27(1), 13–18. lactobacilli. Journal of Dairy Science, 100(10), 7825–7833.
Maalej, H., Boisset, C., Hmidet, N., Buon, L., Heyraud, A., & Nasri, M. (2014). Torres, C. A. V., Antunes, S., Ricardo, A. R., Grandfils, C., Alves, V. D., Freitas, F., et al.
Purification and structural data of a highly substituted exopolysaccharide from (2012). Study of the interactive effect of temperature and pH on exopolysaccharide
Pseudomonas stutzeri AS22. Carbohydrate Polymers, 112, 404–411. production by Enterobacter A47 using multivariate statistical analysis. Bioresource
Matthies, H., Schroeder, H., Smalla, K.-H., & Krug, M. (2000). Enhancement of glutamate Technology, 119, 148–156.
release by L-Fucose changes effects of glutamate receptor antagonists on long-term Ulaganathan, T. S., Boniecki, M. T., Foran, E., Buravenkov, V., Mizrachi, N., Banin, E.,
potentiation in the rat hippocampus. Learning & Memory, 7(4), 227–234. et al. (2017). New ulvan-degrading polysaccharide lyase family: Structure and
Rättö, M., Verhoef, R., Suihko, M.-L., Blanco, A., Schols, H. A., Voragen, A. G. J., et al. catalytic mechanism suggests convergent evolution of active site architecture. ACS
(2006). Colanic acid is an exopolysaccharide common to many enterobacteria Chemical Biology, 12(5), 1269–1280.
isolated from paper-machine slimes. Journal of Industrial Microbiology and Vanhooren, P. T., & Vandamme, E. J. (1999). L-fucose: Occurrence, physiological role,
Biotechnology, 33(5), 359–367. chemical, enzymatic and microbial synthesis. Journal of Chemical Technology and
Salli, K., Hirvonen, J., Siitonen, J., Ahonen, I., Anglenius, H., & Maukonen, J. (2020). Biotechnology, 74(6), 479–497.
Selective utilization of the human milk oligosaccharides 2’-fucosyllactose, 3-fuco­ Wampach, L., Heintz-Buschart, A., Fritz, J. V., Ramiro-Garcia, J., Habier, J., Herold, M.,
syllactose, and difucosyllactose by various probiotic and pathogenic bacteria. et al. (2018). Birth mode is associated with earliest strain-conferred gut microbiome
Journal of Agricultural and Food Chemistry, 69(1), 170–182. functions and immunostimulatory potential. Nature Communication, 9(1). https://
Sevag, M. G., Lackman, D. B., & Smolens, J. (1938). The isolation of the components of doi.org/10.1038/s41467-018-07631-x
streptococcal nucleoproteins in serologically active form. Journal of Biological Wei, X., Fu, X., Xiao, M., Liu, Z., Zhang, L., & Mou, H. (2020). Dietary galactosyl and
Chemistry, 124(2), 425–436. mannosyl carbohydrates: In-vitro assessment of prebiotic effects. Food Chemistry,
Silchenko, A. S., Rasin, A. B., Kusaykin, M. I., Malyarenko, O. S., Shevchenko, N. M., 329, 127179. https://doi.org/10.1016/j.foodchem.2020.127179
Zueva, A. O., et al. (2018). Modification of native fucoidan from Fucus evanescens by Wu, Y., Cui, W., Eskin, N. A. M., Goff, H. D., & Nikiforuk, J. (2011). NMR analysis of a
recombinant fucoidanase from marine bacteria Formosa algae. Carbohydrate methylated non-pectic polysaccharide from water soluble yellow mustard mucilage.
Polymers, 193, 189–195. Carbohydrate Polymers, 84(1), 69–75.
Sleeth, M. L., Thompson, E. L., Ford, H. E., Zac-Varghese, S. E. K., & Frost, G. (2010). Free Wu, Z., Wang, G., Wang, W., Pan, D., Peng, L., & Lian, L. (2018). Proteomics analysis of
fatty acid receptor 2 and nutrient sensing: A proposed role for fibre, fermentable the adhesion activity of Lactobacillus acidophilus ATCC 4356 upon growth in an
carbohydrates and short-chain fatty acids in appetite regulation. Nutrition Research intestine-like pH environment. Proteomics, 18, 1700308.
Reviews, 23(1), 135–145. Xiao, M., Fu, X., Wei, X., Chi, Y., Gao, W., Yu, Y., et al. (2021). Structural
Stahl, M., Friis, L. M., Nothaft, H., Liu, X., Li, J., Szymanski, C. M., et al. (2011). L-Fucose characterization of fucose-containing disaccharides prepared from
utilization provides Campylobacter jejuni with a competitive advantage. PNAS, 108 exopolysaccharides of Enterobacter sakazakii. Carbohydrate Polymers, 252, 117139.
(17), 7194–7199. https://doi.org/10.1016/j.carbpol.2020.117139
Tanaka, T., Nakashima, T., Ueda, T., Tomii, K., & Kouno, I. (2007). Facile discrimination Li, Y.-X., Liu, H.-J., Shi, Y.-Q., Yan, Q.-J., You, X., & Jiang, Z.-Q. (2020). Preparation,
of aldose enantiomers by reversed-phase HPLC. Chemical and Pharmaceutical Bulletin, characterization, and prebiotic activity of manno-oligosaccharides produced from
55(6), 899–901. cassia gum by a glycoside hydrolase family 134 β-mannanase. Food Chemistry, 309,
Teixeira, T. F. S., Grześkowiak, Ł., Franceschini, S. C. C., Bressan, J., Ferreira, C. L. L. F., 125709. https://doi.org/10.1016/j.foodchem.2019.125709
& Peluzio, M. C. G. (2013). Higher level of faecal SCFA in women correlates with Zabel, B. E., Gerdes, S., Evans, K. C., Nedveck, D., Singles, S. K., Volk, B., et al. (2020).
metabolic syndrome risk factors. British Journal of Nutrition, 109(5), 914–919. Strain-specific strategies of 2’-fucosyllactose, 3-fucosyllactose, and difucosyllactose
assimilation by Bifidobacterium longum subsp. infantis Bi-26 and ATCC 15697.
Scientific Report, 10(1), 15919.

10

You might also like