You are on page 1of 8

Industrial Crops & Products 159 (2021) 113079

Contents lists available at ScienceDirect

Industrial Crops & Products


journal homepage: www.elsevier.com/locate/indcrop

Extraction, structure characterization, carboxymethylation and antioxidant


activity of acidic polysaccharides from Craterellus cornucopioides
Yuntao Liu a, *, 1, Xiaoyu Duan a, 1, Mingyue Zhang a, Cheng Li a, Zhiqing Zhang a, Bin Hu a,
Aiping Liu a, Qin Li a, Hong Chen a, Zizhong Tang b, Wenjuan Wu c, Daiwen Chen d
a
College of Food Science, Sichuan Agricultural University, Yaan 625014, China
b
College of Life Science, Sichuan Agricultural University, Yaan 625014, China
c
College of Science, Sichuan Agricultural University, Yaan 625014, China
d
Key Laboratory of Animal Disease-Resistance Nutrition, Animal Nutrition Institute, Sichuan Agricultural University, Chengdu 611130, China

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

Keywords: Herein, we isolated and purified polysaccharides from Craterellus cornucopioides. Chemical composition analysis
Craterellus cornucopioides revealed that both polysaccharides (CCPs-1 amd CCPs-2) were composed of different molar ratios of rhamnose,
Polysaccharides fucose, arabinose, xylose, mannose, glucose, and galactose. The structure of polysaccharides was determined
Structure characterization
using FT-IR, periodate oxidation, smith degradation, methylation analysis, and NMR, which demonstrated that
Carboxymethylation
Antioxidant activity
the two polysaccharides were primarily connected by mannose with (1→3)-linked. Notably, the configuration of
the two polysaccharides exhibited random coil with pyranoid polysaccharide containing α or β glycosidic bond.
Furthermore, to increase the antioxidant activity of polysaccharides (CCPs-1 amd CCPs-2), two carboxymethy­
lated polysaccharides were obtained as cmCCPs-1 and cmCCPs-2 with the degree of substitution values of 0.34
and 0.52, respectively. Structural analysis elucidated that nonselective carboxymethylation occurred in cmCCPs-
1 and cmCCPs-2, where C-2, C-4 or C-6 were partially substituted. Of note, cmCCPs-2 exhibited the most sig­
nificant antioxidant activity.

1. Introduction solubility and antioxidant abilities of polysaccharides (Xu et al., 2009;


Wang et al., 2016a; Liu et al., 2017a).
Functional polysaccharides are naturally abundant materials that In the past few years, there has been a growing body of literature,
possess essential pharmacological activities including immunoregula­ emphasizing separation, characterization, modification, and utilization
tory, anti-inflammatory, antioxidant, and antimicrobial. Therefore, they of different polysaccharides from various edible fungus (Xu et al., 2009;
are considered as one of the most promising new resources for the Guo et al., 2019; Yang et al., 2011). For instance, polysaccharides
development of drugs and nutraceuticals (Xu et al., 2009; Wang et al., extracted from C. cornucopioides have been outlined to exhibit various
2016a; Liu et al., 2020f). Specifically, previous works have demon­ bioactive effects. In our previous study, we obtained C. cornucopioides
strated that bioactivity of polysaccharides primarily depended on its polysaccharides (CCPs) with a specific antioxidant activity (Liu et al.,
physicochemical property and molecular structure, such as water solu­ 2020c). In addition, studies on the structure-activity relationship of
bility, chemical composition, main-chain, and branched-chain, etc. polysaccharides have shown that the bioactivities of most of the natural
(Ferreira et al., 2015; Liu et al., 2017a, 2017b). Chemical modifications polysaccharides were directly or indirectly affected by their structures
such as sulfated modification, acetylation, and selenylation (Liu et al., (Duan et al., 2020). However, the detailed structures and types of
2017a, 2017b) are deemed to be an effective strategy to introduce the polysaccharides from C. cornucopioides remains elusive, thus hindering
functional groups, improve the waters solubility of those poorly the investigation of their structure and activity relationships. Therefore,
water-soluble fractions and increase the bioactivity of polysaccharides it is necessary to further purify the CCPs to obtain high purity, and to
(Xu et al., 2009; Wang et al., 2016a). Furthermore, it has been widely clarify its structural characteristics, so as to understand its
accepted that carboxymethyl modification might enhance the water structure-activity relationship. Notably, compared with the positive

* Corresponding author at: College of Food Science, Sichuan Agricultural University, 46# Xinkang Road, Yaan, Sichuan 625014, China.
E-mail addresses: lyt_taotao@163.com, nutritionlab@sicau.edu.cn (Y. Liu).
1
The authors contributed equally to this article.

https://doi.org/10.1016/j.indcrop.2020.113079
Received 2 September 2020; Received in revised form 24 October 2020; Accepted 25 October 2020
Available online 4 November 2020
0926-6690/© 2020 Elsevier B.V. All rights reserved.
Y. Liu et al. Industrial Crops & Products 159 (2021) 113079

control, CCPs and their purified components depicted insufficient po­ CCPs-2 was measured using HPGPC with an Agilent 1100 HPLC system.
tential to improve its biological properties and pharmacological appli­ Setting procedure is as follows: the column was maintained at 30 ◦ C and
cations. The structural modification of polysaccharides may change its eluted with ultrapure water at a flow rate of 0.8 mL/min. According to a
spatial configuration and biological activities. Besides, to the best of our technique described in our previous study (Liu et al., 2020c).
knowledge, there were no references to the antioxidant activity of car­
boxymethylated CCPs as well as the structural changes caused by car­ 2.3.2.2. Fourier transform infrared (FT-IR) spectra. The infrared spec­
boxymethyl. In this context, we synthesized carboxymethylated trum was acquired using the FT-IR spectrometer at room temperature
polysaccharides and then investigated their structure and antioxidant (25 ◦ C). In particular, the polysaccharides were blended with KBr
activities compared with the native one. This was executed for further powder (1:100, w/w), and pressed into pellets for the collection of
understanding of the structure-function relationships of CCPs and ex­ infrared spectra in a range of 400–4000 cm− 1 at 4 cm− 1 resolution using
pands the application of C. cornucopioides resources. 32 scans for three or more times.

2. Materials and methods 2.3.2.3. Analysis of glycosidic bond connection mode. Periodate oxida­
tion and Smith degradation analysis were performed based on a previ­
2.1. Materials and chemicals ously established method (Liu et al., 2020g), whereas methylation
analysis was assessed according to a reported method with some mod­
Craterellus cornucopioides was purchased and collected from Kunm­ ifications (Liu et al., 2020g).
ing, Yunnan, China, while DPPH (1, 1-diphenyl-2-picrylhydrazyl), BHT Fifty mg polysaccharides were proton-exchanged in deuterium oxide
(2, 6-di-tert-butyl-4-methylphenol), EDTA (ethylene diamine tetraacetic (D2O, 99.9 %) three times, lyophilized, and then redissolved in 1 mL of
acid), and standard monosaccharide samples were obtained from San­ D2O. Chemical shifts were expressed in ppm using tetramethylsilane
gon Biotech (China). All other analytical reagent grade chemicals used (TMS). Thereafter, 1H nuclear magnetic resonance (1H NMR), 13C nu­
in this study were purchased from Sinopharm Chemical Reagent Co. Ltd, clear magnetic resonance (13C NMR), 1H-1H correlated spectroscopy
China. (1H-1H COSY), and heteronuclear single-quantum coherence (13C-1H
HSQC) were recorded using an NMR apparatus (AVANCE II 400 M,
2.2. Isolation and purification of the polysaccharide Bruker, Germany). The NMR test parameters were given in Table S1.

To obtain CCPs from crude polysaccharides of C. cornucopioides, they 2.3.3. Conformation studies of C. cornucopioides polysaccharides (CCPs)
were extracted according to the method we previously described (Liu
et al., 2020c), through hot water extraction, Sevag to remove protein 2.3.3.1. Solubility determination. The water-solubility test of poly­
and ethanol for precipitation. saccharides from C. cornucopioides was performed based on a previously
Here, CCPs were dissolved with ultra-pure water (100 mg/mL) and reported method with slight modification (Sharifian-Nejad and Shek­
then purified using a DEAE-52 column (3.5 cm × 20 cm). After loading archizadeh, 2019). Specifically, polysaccharide samples (100 mg) were
the CCPs solution, the column was eluted with twice the column volume placed into a centrifuge tube and weighed as W0. Then, distilled water
of ultra-pure water. Afterward, the column was also eluted with a (10 mL) was added and mixed for a certain time (0, 5, 10, 15, 20, 25, 30,
stepwise gradient of 0.1 M to 2.0 M NaCl at a flow rate of 1 mL/min. The 40, 50, 60, 120, 180, 240, 300 min) at room temperature. After centri­
eluent was automatically collected (10 mL per tube), while the sugar fugation (8000 rpm, 30 min), we removed supernatant and dried the
distribution was monitored using the phenol-sulfuric acid method. centrifuged tubes at 55 ◦ C for 30 min. Consequently, the centrifuged
Notably, the CCPs were divided into 2 peaks. The ratios based on crude tubes were weighed as W1. Lastly, the water-solubility of polysaccharide
polysaccharide content were 32.36 % and 67.64 %. The obtained eluent samples was computed using the following formula:
containing the carbohydrate was collected, concentrated, and lyophi­ Water-solubility (%) = (W0-W1)/ W0 × 100 %
lized. Sephadex G-200 chromatography column (2.6 cm × 80 cm) was
used to purify further the two fractions, which were then dissolved in
distilled water (30 mg/mL). Subsequently, the column was washed with
ultra-pure water at a flow rate of 0.3 mL/min. Finally, the eluent was 2.3.3.2. Congo red analysis. Congo red staining was used to test the
detected using the phenol-sulfuric acid method, then collected and interaction of polysaccharides by examining the visible absorption
concentrated. The resultant solution was freeze-dried to obtain CCPs-1 maximum wavelength (range of 400–600 nm) according to a previous
and CCPs-2. method (Liu et al., 2017a).

2.3. Characterization 2.3.3.3. The I2-KI analysis. To determine the degree of branching of
polysaccharides and their derivatives, the I2-KI experiment was used,
2.3.1. Chemical composition analysis while dextran T10 was applied as a positive control to detect the
The CCPss’ (CCPs-1 and CCPs-2) chemical compositions were absorbance in the wavelength range of 300–700 nm after reacting with
analyzed as previously described (Liu et al., 2020c). Afterward, the I2-KI reagent (0.02 % I2 and 0.2 % KI, w/v) at 20 ◦ C for 15 min (Guo
meta-hydroxybiphenyl method was used to assess the content of uronic et al., 2019).
acid, while the monosaccharide compositions of polysaccharides were
evaluated using gas chromatography (GC, Shimadzu GC-14A) equipped 2.4. The carboxymethylation of polysaccharides
with flame ionization detector (250 ◦ C) and PEG20 M silica capillary
column (30 m ×0.53 mm I.D.). Setting procedure is as follows: initial 2.4.1. The cmCCPs preparation
column temperature was maintained at 150 ◦ C for 3 min, increased to The carboxymethylation modification of polysaccharide from
240 ◦ C at a rate of 3 ◦ C/min and maintained for 20 min. C. cornucopioides (cmCCPs) was performed as previously described with
slight modification (Liu et al., 2017a). Briefly, 200 mg CCPs, 8 mL iso­
2.3.2. Structure determination of C. cornucopioides polysaccharides propyl alcohol, 4 mL 20 % (v/v) dropwise sodium hydroxide solution
(CCPs) and carboxymethyl agent (including 4 g chloroacetic acid, 2 mL 20 %
NaOH and 4 mL isopropanol) were added one at a time to form a mixed
2.3.2.1. Molecular weight (Mw). The molecular weight of CCPs-1 and solution and allowed to react at 60 ◦ C for 3 h. After the reaction was

2
Y. Liu et al. Industrial Crops & Products 159 (2021) 113079

complete, the solution was cooled and the pH adjusted to 7. Then, the Table 1
mixture was dialyzed (3 kDa) against ultra-pure water for 48 h and Molecular weight (Mw), uronic acid, and monosaccharides composition (%) of
precipitated with 95 % ethanol. After centrifugation (5000 rpm, CCPs-1 and CCPs-2.
10 min), the precipitates were lyophilized for 48 h to yield cmCCPs-1 Process CCPs-1 CCPs-2
and cmCCPs-2. Mw (Da) 1.38 × 105 2.73 × 105
Uronic acid content (%) 5.50 ± 0.03a 11.36 ± 0.25b
2.4.2. Characterization of cmCCPs Monosaccharides composition (%)
After the yields of cmCCPs-1 s were obtained, the DS (degree of Rhamnose 0.19 0.19
Fucose 8.28 6.05
substitution) was examined according to our previous study (Liu et al.,
Arabinose 0.33 0.40
2017a). Next, the structure of cmCCPs-1 and cmCCPs-2 was character­ Xylose 28.05 31.78
ized (FT-IR and NMR) as elucidated in section 2.3.2.2 and 2.3.2.3, while Mannose 42.75 43.76
their conformation (Water-solubility, the interaction of polysaccharides, Glucose 6.36 9.38
and branching of polysaccharides) was determined as outlined in section Galactose 14.05 8.44

2.3.3. CCPs, polysaccharides from C. cornucopioides; CCPs-1, eluted with ultra-pure


water obtained polysaccharides by DEAE-52, and purified further by Sephadex
2.5. Determination of the antioxidant activity in vitro G-200; CCPs-2, eluted with 0.1 M NaCl obtained polysaccharides by DEAE-52,
and purified further by Sephadex G-200.
Antioxidant activities (DPPH free radical scavenging activity,
reducing power, and metal chelating activity) were determined as pre­ However, the composition of the monosaccharides differed from those
viously described (Zhang et al., 2018). of C. cornucopioides polysaccharides noted in other previous studies (Guo
et al., 2019; Yang et al., 2018), and the differences could be attributed to
2.6. Statistical analysis the enhanced release of more C. cornucopioides polysaccharides due to
change in the extraction time or the concentration of acid, temperature
All numerical variables are presented as the mean plus standard and time of hydrolysis (Shi et al., 2020).
deviation of the mean. The statistical significance was assessed using
Tukey’s, while its significance level was set at p < 0.05. 3.3. Fourier transform infrared (FT-IR) analysis

3. Results and discussion The features of functional groups in CCPs-1 and CCPs-2 were
examined using FT-IR spectroscopy (Fig. 1). The FT-IR spectra of CCPs-1
3.1. Extraction and purification of CCPs and CCPs-2 exhibited absorption in bands at 3382, 2925, 1601, 922, and
893 cm− 1, which are typical bands for carbohydrates (range
Herein, CCPs were obtained using hot water extraction and ethanol 4000− 400 cm− 1). Moreover, the FT-IR spectra of CCPs (CCPs-1 and
sedimentation method resulting in a yield of 3.89 % based on the dry CCPs-2) were similar, thereby indicating that purification induced no
weight of raw material. After deproteinization and decolorization, the significant effects on the structure feature of CCPs (Liu et al., 2020c). In
DEAE-52 column and Sephadex G-200 chromatography column were
employed to fractionate and purify CCPs. As depicted in Figure S1, the
fractions obtained after the purification using the DEAE-52 cellulose
column and Sephadex G-200 chromatography column, yielded a single
and symmetrical peak named CCPs-1 and CCPs-2. The calculated rela­
tive weight average molecular weights of CCPs-1 and CCPs-2 were
1.38 × 105 and 2.73 × 105 Da, respectively.
Based on the previous reports, the DEAE 52-cellulose and Sepharose
CL-4B column was adopted to purify polysaccharide extracted from
CCPs, and the relative weight average molecular weight of CCPs ob­
tained was 9.2 × 105 Da (Yang et al., 2018). In another study by Guo
et al. (2019) used a DEAE-sepharose CL-6B column, and the relative
weight average molecular weight of the water-soluble polysaccharide
was 1.97 × 103 kDa. A major reason for this discrepancy in the molec­
ular weight of polysaccharides between these two studies and our
research might be the source or extraction method (Zhang et al., 2020).

3.2. Chemical components of CCPs-1 and CCPs-2

The contents of uronic acid and monosaccharides composition are


outlined in Table 1. These findings suggest that the two polysaccharides
are acidic, where CCPs-2 contained the highest amount of uronic acid
residues.
The GC results of monosaccharides composition analysis revealed
that CCPs-1 was mainly composed of rhamnose, fucose, arabinose,
xylose, mannose, glucose, and galactose in the ratio of 0.19 %, 8.28 %,
0.33 %, 28.05 %, 42.75 %, 6.36 %, and 14.05 %, respectively, whereas
CCPs-2 was primarily comprised of rhamnose, fucose, arabinose, xylose,
mannose, glucose, and galactose in the ratio of 0.19 %, 6.05 %, 0.40 %,
31.78 %, 43.76 %, 9.38 %, and 8.44 %, respectively. Notably, mannose
may be the backbone of the main chain of CCPs-1 and CCPs-2. This Fig. 1. Fourier-transforms infrared spectroscopy analysis of the CCPs-1,
finding was consistent with the previous research by Guo et al. (2019). cmCCPs-1 (A) and CCPs-2, cmCCPs-2 (B).

3
Y. Liu et al. Industrial Crops & Products 159 (2021) 113079

summary, CCPs (CCPs-1 and CCPs-2) possessed the characteristic ab­ 3.5. Methylation analysis
sorption patterns of polysaccharides, which primarily composed of py­
ranose rings with different molecular vibrations and structures. Results of methylation analysis (Table S3) revealed that the largest
residue in the CCPs-1 structure was the glycosidic bond of mannose
3.4. Periodate oxidation and Smith degradation analysis which included Manp-(1→ (18.38 %) and →3,6) Manp-(1→ (29.56 %).
The second-largest residue in the structure was the glycosidic bond of
Based on the calculation, the ratio of consumed NaIO4 and produced xylose containing 15.39 % of →3)-Xylp-(1→ and 10.21 % of Xylp-(1→.
HCOOH during oxidation was 7.95:1 of CCPs-1 (consumed NaIO4 of These were followed by the glycosidic bond of galactose containing →3)-
0.735 mmol and the production of HCOOH was 0.092 mmol), which Galp-(1→ (9.67 %), Galp-(1→ (3.82 %) and →2,4)-Galp-(1→ (1.24 %),
indicated the existence of (1→)-linked or (1 → 6)-linked or little branch. glycosidic bond of fucose including →3,4)-Fucp-(1→ (6.24 %), glyco­
In addition, the consumption of periodate was much higher than the sidic bond of glucose, consisting of Glcp-(1→ (2.57 %) and →6)-Glcp-
release of HCOOH, which implied the presence of (1→2), (1→2,6), (1→ (2.61 %), and glycosidic bond of arabinose, →3)-Araf-(1→ (0.31
(1→4), or (1→4,6) glycosidic bonds in CCPs-1. Besides, periodate was %). The inferences were consistent with the results of the periodate
not completely consumed, which also indicates that linkage types which oxidation and Smith degradation.
cannot be oxidized by periodate such as (1→3), (1→2,3), (1→3,4), Similarly, the following 12 components →3,4)-Fucp-(1→, →3)-Araf-
(1→3,6) and (1→2,4), might either exist or extensive in CCPs-1 (Guo (1→, →3)-Xylp-(1→, Xylp-(1→, →4)-Xylp-(1→, Manp-(1→, →3,6)-
et al., 2019; Liu et al., 2020e). Manp-(1→, Glcp-(1→, →6)-Glcp-(1→, →3)-Galp-(1→, Galp-(1→, →2,4)-
After dialysis, reduction and acetylation reaction, GC–MS analysis Galp-(1→ —were found at a molar ratio of 5.13: 0.34: 6.26: 8.15: 17.56:
was conducted (Table S2). The presence of rhamnose, fucose, xylose, 19.86: 25.37: 2.79: 4.23: 7.84: 1.68: 0.79.
mannose, glucose, and galactose implied that their partial residues
contained (1→3)-linked, (1→3,4)-linked, (1→2,4)-linked or (1→2,3,4)-
3.6. NMR spectroscopy analysis
linked, which were not oxidized. Additionally, the mannose content was
88.37 %, which indicated that it is the primary main chain of CCPs-1.
The 1H, 13C and 13C-135 NMR spectra for CCPs-1 and CCPs-2 are
Besides, a few glycerols were detected in the products, which reflected
shown in Figs. 2 and 3 . The signals of protons derived from − CH3 (C-6)
that there were either (1→)-linked, (1→2)-linked or (1→6)-linked in
of deoxy-sugar were generally seen in the region of 0.8–1.4 ppm (here is
CCPs-1. However, there was no erythritol found in the products, thus
fucose) (Li et al., 2020), and the signals for − CH3 protons of O-acetyl
suggesting that (1→4)-linkage or (1→4,6)-linkages residues were ab­
groups (− OCOCH3) were in the region of 1.8–2.2 ppm (Liu et al.,
sent. Of note, the amount of ethylene glycol produced was higher than
2020a). Generally, residues with anomeric protons chemical shifts
that of glycerol, and there was no erythritol, which might be the
exceeding δH 4.95 ppm were in α-configuration, because chemical shift
oxidative degradation of xylose (five-carbon sugar). Therefore, there
below δH 4.95 ppm indicated the existence of β-configuration (Li et al.,
were contain (1→2)-linked or (1→4)-linked xylose in CCPs-1. Hence, the
2015). In the 13C NMR spectrum, the resonances in the region of δC
linkages of the backbone of the polysaccharide of CCPs-1 were inferred
95.0–110.0 ppm were assigned to the anomeric carbon atoms, and the
to be (1→)-linked, (1 → 3)-linked, (1 → 6)-linked, (1→2)-linked, (1→4)-
chemical shift range of the α-configuration is usually δC 90− 102 ppm
linked or (1 → 3,6)-linked.
(Liu et al., 2020a). Thus, the residues in CCPs-1 and CCPs-2 contained
Similarly, the periodate oxidation findings and the changes in smith
both α-configuration and β-configuration. The absence of 13C NMR
degradation products of CCPs-2 were close to CCPs-1. In this context, the
spectrum of signals in the region δ 83–88 reflected that all sugar residues
linkages of the backbone of polysaccharide of CCPs-2 were also inferred
of CCPs-1 and CCPs-2 were in the pyranose form (Velichko et al., 2020),
to be (1→)-linked, (1 → 3)-linked, (1 → 6)-linked, (1→2)-linked, (1→4)-
which consistent with the results of FT-IR.
linked or (1 → 3,6)-linked.
Since the 13C NMR signals of the two polysaccharide samples were
crowded and the content of some monosaccharide residues was low, the

Fig. 2. The NMR spectra of CCPs-1 and cmCCPs-1, (A) 1H NMR of CCPs-1; (B) 13C HMR of CCPs-1; (C) DEPT 135 of CCPs-1; (D) 1H NMR of cmCCPs-1; and (E) DEPT
135 of cmCCPs-1.

4
Y. Liu et al. Industrial Crops & Products 159 (2021) 113079

Fig. 3. The NMR spectra of CCPs-2 and cmCCPs-2, (A) 1H NMR of CCPs-2; (B) 13C HMR of CCPs-2; (C) DEPT 135 of CCPs-2; (D) 1H NMR of cmCCPs-2; and (E) DEPT
135 of cmCCPs-2.

number of sugar residues cannot be determined from 13C NMR. There­ 160–180 ppm (Fig. 3B), which illustrated that CCPs-2 contained uronic
fore, it was necessary to determine the type of the primary glycosidic acid. This was consistent with the result of the determination of uronic
bond and the specific position of the corresponding H and C atoms by acid test in section 3.2 (Wang et al., 2016b). Similarly, these structural
combining 2D NMR correlation spectra (Figure S2, S3, S4, S5). residuals were determined individually as follow: residues A2:
Analysis of the 1H NMR spectrum (Fig. 2A) indicated that CCPs-1 had →1)-α-L-Fuc- (5.27/99.52) (Li et al., 2020), residues B2:
eight main anomeric proton signals, indicating that CCPs-1 contained →2)-α-L-Xylp-(1→ (5.20/99.58) (Shakhmatov et al., 2020), residues C2:
many types of sugars. The eight major anomeric proton signals at 5.27, α-D-Manp-(1→ (5.03/101.35) (Huo et al., 2020), residues D2:
5.20, 5.03, 4.98, 4.89, 4.83, 4.67 and 4.31 ppm, exhibited relative in­ →3)-α-D-GalpA-(1→ (4.97/101.51) (Kostalova and Hromadkova, 2019),
tegrals of 1.0: 1.37: 4.89: 2.11: 2.62: 0.48: 0.45: 5.14. The 1H NMR spin residues E2: →3)-α-D-Xylp(1→ (4.89/97.82) (Huo et al., 2020; Hu et al.,
systems of the polysaccharide were assigned based on the 1H-1H COSY 2020), residues F2: →3)-α-D-Manp-(1→ (4.82/98.39) (Kokoulin et al.,
spectrum (Figure S2) and the direct CH coupling was determined by the 2020), residues G2: →1)-β-D-Glcp-(6→ (4.66/102.87) (Perepelov et al.,
1 13
H- C HSQC spectrum (Figure S3). Eight pairs of 1H signals of anomeric 2018; Liu et al., 2020g), residues H2: →4)-β-D-Xylp-(1→ (4.42/102.66)
CH attributed to eight different types of glycosidic bonds were assigned (Sigida et al., 2020), residues I2: →6)-β-D-Manp-(1→ (4.31/103.46)
to δ5.27/99.48, 5.20/99.71, 5.03/101.32, 4.98/101.55, 4.89/97.79, (Wang and Guo, 2020; Liu et al., 2020b). The entire assignment of
4.83/98.26, 4.83/98.26, 4.65/102.91 and 4.31/103.48 based on their chemical shifts is summarized in Table 2. Monosaccharide composition
chemical shifts in the 1H and 13C analysis and corresponding HSQC analysis confirmed that CCPs-2 was mainly composed of fucose, xylose,
spectra, labeled A1-H1. Comparing the chemical shift data of similarly mannose, glucose and galactose, which is consistent with glycoside
substituted sugar residues with the above monosaccharide composition, residues.
periodate oxidation and smith degradation, methylation and infrared
spectrum results indicated that the nine types of glycosidic bonds
3.7. Carboxymethylation modification
confirmed the existence of nine organic residues in CCPs-1. For example,
the signal at δC 99.48 ppm that correlated with the anomeric proton
3.7.1. Characterization of cmCCPs
signal at δH 5.27 was found in the HSQC spectrum for CCPs-1. Combined
In the past, we have done much research to improve the activities of
with the COSY spectrum and published data, H-2 (3.57 ppm), H-3
polysaccharides (Liu et al., 2020c, 2020d). Afterward, the findings
(3.79 ppm), H-4 (3.97 ppm), H-5 (4.14 ppm) and H-6 (1.16 ppm) were
indicated that carboxymethylation is an effective method to modify the
unambiguously assigned (Table 2). The values of signals for C-1 to C-6
polysaccharides. This method of modifying polysaccharides included a
were assigned based on the HSQC spectra. The field signals of C-2
two-step etherification reaction. Herein, the calculated DS values for
(72.45 ppm), C-3 (66.53 ppm), C-4 (69.82 ppm), C-5 (66.44 ppm) and
CCPs-1 and CCPs-2 were 0.34 ± 0.04 and 0.52 ± 0.06, respectively,
C-6 (15.42 ppm) suggested that residue was →1)-α-L-Fuc- (Li et al.,
which deemed that substitution occurred. Additionally, the yields of the
2020). Similarly, the other residues were assigned as follows: residual
native polysaccharides for CCPs-1 and CCPs-2 were 71.85 ± 1.04 % and
B1: →2)-α-L-Xylp-(1→ (Shakhmatov et al., 2020), residual C1:
80.45 ± 0.69 %, respectively (data not shown).
α-D-Manp-(1→ (Huo et al., 2020), residual D1: →3)-α-D-GalpA-(1→
Fourier-transforms infrared (FT-IR) spectrum of cmCCPs-1 and
(Kostalova and Hromadkova, 2019), residual E1: →3)-α-D-Xylp(1→
cmCCPs-2 (Fig. 1) revealed three new strong peaks at 1601 cm− 1,
(Huo et al., 2020; Hu et al., 2020), residual F1: →3)-α-D-Manp-(1→
1419 cm− 1, and 1313 cm− 1 after carboxymethylation modification.
(Kokoulin et al., 2020), residual G1: →1)-β-D-Glcp-(6→ (Perepelov et al.,
Thereafter, these peaks were assigned to asymmetrical and symmetrical
2018; Liu et al., 2020g), residual H1: →6)-β-D-Manp-(1→ (Wang and
COO- stretching vibrations (Liu et al., 2017a). Remarkably, these find­
Guo, 2020; Liu et al., 2020b).
ings sufficiently indicated that the native polysaccharide was carbox­
The 1H NMR spectrum of fraction CCPs-2 contained nine main peaks
ymethylated successfully.
in the anomeric region, which were labeled A2 to I2. The 13C NMR
The 1H and 13C NMR spectrum (DEPT-135) of cmCCPs-1 and
spectrum of CCPs-2 had a signal at a low field ranging from δC
cmCCPs-2 provided more detailed structural information (see Figs. 2

5
Y. Liu et al. Industrial Crops & Products 159 (2021) 113079

Table 2 activity value. Additionally, its solubility and dissolution kinetics are
1
H and 13C NMR chemical shift data (δ, ppm) for CCPs-1 and CCPs-2. thus pivotal specifications in a large number of technological application
Residue H/ 1.00 2.00 3.00 4.00 5.00 6.00 processes. Figure S6A illustrates the solubility of CCPs, CCPs-1, CCPs-2,
C cmCCPs-1, and cmCCPs-2 as a function of dissolution time. Notably,
CCPs-1 CCPs were steadily soluble in water and reached their maximum degree
→1)-α-L- 1
H 5.27 3.57 3.79 3.97 4.14 1.16 of solubility (around 95 %) after 5 h. In particular, its two fractions
A1
Fuc- 13
C 99.48 72.45 66.53 69.82 66.44 15.42 (CCPs-1 and CCPs-2) were almost completely dissolved in water after
1
→2)-α-L- H 5.20 3.69 3.88 4.06 40 min. The water solubility of cmCCPs-1 and cmCCPs-2 was enhanced
B1 13
Xylp-(1→ C 99.71 71.77 71.27
α-D- 1
H 5.03 3.81 4.12 4.35
compared with original polysaccharide, and this perhaps may be due to
C1 Manp- 13
the increase in hydrophilic carboxyl groups (Liu et al., 2017a).
C 101.32 78.73 75.19
(1→
1
→3)-α-D- H 4.98 4.00 4.14 4.31 3.7.3. Congo red test
D1 GalpA- 13
C 101.55 71.81 66.44 Congo red (CR) test is a useful, simple, and fast method that has been
(1→
→3)-α-D- 1
H 4.89 3.74 4.04 4.14 extensively applied to examine the helical conformation of poly­
E1 13 saccharides. Furthermore, the order-disorder transitions in a helical
Xylp(1→ C 97.79 77.40 70.09 66.44
→3)-α-D- 1
H 4.83 3.69 3.87 4.05 polymer are usually accompanied by changes in the visible absorption
F1 Manp- 13 spectra of the complexes formed with congo red (Liu et al., 2017a).
C 98.26 71.77 78.46 70.09
(1→
→1)-β-D- 1
H 4.65 3.44 3.78 3.97 4.14
Elsewhere, compared with the congo red (control group), the λmax of the
G1 13 complexes with a polysaccharide, which possessed a triple helix struc­
Glcp-(6→ C 102.91 71.70 69.49 69.82 66.44
→6)-β-D- 1
H 4.31 3.21 3.55 3.86 4.04 ture, showed a bathochromic shift and thereby forming a metastable
H1 Manp- 13
C 103.48 65.14 76.26 78.46 70.09 region in a certain concentration range. Here, the comparison results of
(1→
the congo red experiment of C. cornucopioides polysaccharides samples
CCPs-2
→1)-α-L- 1
H 5.27 3.56 are depicted in Figure S6B. The λmax of the CCPss and the cmCCPss,
A2 13 when mixed with congo red, gradually decreased and indicated irregular
Fuc- C 99.52

B2
→2)-α-L- 1
H 5.20 3.70 changes with the increase in NaOH concentration (0− 0.8 mol/L).
13
Xylp-(1→ C 99.58 Collectively, the above arguments indicated that CCPs-1, CCPs-2,
1
α-D- H 5.03 3.81 4.10
cmCCPs-1, and cmCCPs-2 exhibited no triple helix structure. Simulta­
C2 Manp- 13
C 101.35 66.55 68.62 neously, the descending behavior of the λmax for the CCPss-CR and
(1→
→3)-α-D- 1
H 4.97 3.71 4.00 cmCCPss-CR complex was elucidated in low NaOH concentration
D2 GalpA- 13
C 101.51 71.61 66.80 (0− 0.3 mol/L). These findings enumerated that CCPss posses a rigid
(1→
1 conformation, whereas their highly ordered conformation may convert
→3)-α-D- H 4.89 3.75
E2
Xylp(1→ 13
C 97.82 69.47 into a flexible conformation under alkaline conditions (Wang et al.,
→3)-α-D- 1
H 4.82 3.71 2016c).
F2 Manp- 13
C 98.39 71.61
(1→ 3.7.4. The I2-KI analysis
1
H 4.66 3.47
The result (Figure S6C) showed that the maximum absorption of
→1)-β-D-
G2 13
Glcp-(6→ C 102.87 71.67
→4)-β-D- 1
H 4.42 3.25 3.87 polysaccharides reacting with I2-KI reagent was 350 nm, whereas there
H2
Xylp-(1→ 13
C 102.66 73.36 65.10 was no absorption peak at around 565 nm for each sample. Thus, indi­
1
→6)-β-D- H 4.31 3.21 3.36 3.52 3.85 4.12 cating that there were many branches on the CCPss and cmCCPss mo­
I2 Manp- 13
C 103.46 73.36 69.43 69.33 65.10 68.62 lecular backbone. This finding corresponded with the study of Guo et al.
(1→
(2019).

and 3). The 5 major anomeric proton signals at 5.20, 5.06, 4.91, 4.73, 3.8. The antioxidant activity in vivo
and 4.64 ppm for cmCCPs-1, compared with those of CCPs-1, revealed
that it was difficult to identify the carbon signals in 13C NMR spectrum of The antioxidant activities of CCPss (CCPs-1 and CCPs-2) and its
cmCCPs-1, which severely overlapped with one another (Fig. 2E). This carboxymethylation fractions (cmCCPs-1 and cmCCPs-2) are depicted in
may due to the impact of the carboxymethyl group that may have Table 3. Herein, the polysaccharide of cmCCPs-2 exhibited the greatest
changed the chemical environment around every group thus weakening activity to scavenge DPPH radicals (EC50 = 136.0 μg/mL), while CCPs-1
the 13C NMR resonance signals (He et al., 2015). Compared with the possessed the lowest scavenging activity (EC50 = 233.2 μg/mL). These
spectrum of CCPs-1, the new negative peaks that appeared around δC findings implied that chemically modified polysaccharides have
71.26, 70.16, 69.59, 68.12 and 67.20 ppm were assigned to C-6 (the five
major C-6 signals at 66.43, 65.17, 61.00, 60.72 and 60.21 ppm for Table 3
CCPs-1), and this down-field shift may due to the carboxymethylation of Antioxidant properties of polysaccharides and its carboxymethylation fractions.
C-6 (Liu et al., 2017a). Furthermore, the enhancement in peak in­
Samples DPPH (EC50) Reducing power Metal chelating activity
tensities 69.42 ppm, and the new peak appeared at around 76.01, 75.24 (EC0.5) (EC50)
were possibly due to the substitution of − CH2COOH group at C-4 or C-2
CCPs-1 233.2 ± 14.4c 210.5 ± 13.4c 535.7 ± 45.7d
position. Based on the above information, we proposed that the sub­ CCPs-2 191.8 ± 19.5bc 190.1 ± 11.2c 480.6 ± 17.8c
stitutions occurred at C-6, C-4, or C-2 (Yang et al., 2011). Similarly, as cmCCPs-1 180.6 ± 16.4bc 191.5 ± 16.3bc 475.9 ± 24.1c
illustrated in Fig. 3D, E, it revealed that nonselective carbox­ cmCCPs-2 136.0 ± 9.9ab 148.9 ± 14.1ab 418.3 ± 25.9b
ymethylation has occurred for cmCCPs-2, that is, C-2, C-4 or C-6 were Positive 89.0 ± 13.9a 138.5 ± 14.8a 304.0 ± 22.6a
control
being partially substituted.
Each value is expressed as the mean ± standard deviation (n = 3). Means with
3.7.2. Solubility determination different letters within a column are significantly different (p < 0.05). Positive
Polysaccharides are commonly prepared in a powdery form and must control, BHT was used as a positive control for DPPH radical scavenging activity
be dissolved in an aqueous solution before being of any significant and reducing power activity, while EDTA was used as a positive control for metal
chelating activity.

6
Y. Liu et al. Industrial Crops & Products 159 (2021) 113079

favorable DPPH radical scavenging activity. However, the discrepancy from C. cornucopioides and demonstrated that they have antioxidant
in scavenging activity between the two polysaccharides might be due to activities themselves through in vitro studies (Yang et al., 2018; Liu et al.,
the difference in the degree of substitution, molecular weight, uronic 2020d). However, the relationship between their antioxidant property
acid content, and monosaccharide composition of polysaccharides. and the Nrf2 signaling pathway remains enigmatic. Hence, this neces­
Moreover, the activity of polysaccharide to provide hydrogen is an sitates more research to explore the relationship between the structure
important factor in antioxidant activity. In another study, it was and activity of C. cornucopioides polysaccharides and their role in the
demonstrated that the carboxymethylation components of poly­ signaling pathway, particularly through in vivo experiments.
saccharides elucidated a stronger antioxidant effect than native poly­
saccharides, which was consistent with the results of our study (Liu 4. Conclusions
et al., 2017a). Their carboxyl groups behave like electrophiles to facil­
itate the intramolecular hydrogen abstraction. In summary, CCPs-1 (1.38 × 105 kDa) and CCPs-2 (2.73 × 105 kDa)
Based on Table 3, the reducing power of the polysaccharides from obtained from C. cornucopioides, are all pyranoid polysaccharides con­
caps ranked as cmCCPs-2 > CCPs-2 > cmCCPs-1 > CCPs-1. In particular, taining α or β glycosidic bond formed by (1→3)-links with mainly
the reducing power of CCPs-2 was greater compared to that of the CCPs- monosaccharides (mannose, xylose). Structural analysis revealed that
1, which may correlate with the structural properties of the nonselective carboxymethylation occurred in cmCCPs-1 and cmCCPs-2.
polysaccharides. Besides, the relationships between structure and activities of CCPss and
Similar to DPPH free radical scavenging activity and reducing power, its derivatives were established based on antioxidant activities in vitro.
carboxymethylated polysaccharides exhibited higher metal chelating Results indicated that cmCCPss had more significant effects than CCPss
activity. The findings are also similar to the reports of the antioxidant at the same dosage, showing that carboxymethylation modification is an
activities of the carboxymethylated derivatives from blackcurrant fruits effective method for enhancing the antioxidant activities of natural
which improved scavenging OH activity compared to native poly­ polysaccharides. Introduction of functional groups improved the water-
saccharides (Daun et al., 2020). solubility and surface area of the polysaccharides, making them interact
with free radicals. Collectively, our results show that derivatization of
3.9. Correlation relationship analysis between structure and activity CCPss improves its bioactivities, hence makes it applicable in functional
food and pharmaceutical materials.
Notably, it is well known that the antioxidant activity of poly­
saccharide largely depends on its structure (molecular weight, mono­ CRediT authorship contribution statement
saccharide composition, amount of hydroxyl group, type of glycosidic
linkage, degree of branching, the changes of advanced stereotactic Yuntao Liu: Validation, Conceptualization, Visualization, Writing -
structure, and chain conformation). However, in this case, it was noted review & editing, Funding acquisition, Supervision. Xiaoyu Duan:
that the antioxidant activity of CCPss (CCPs-1 and CCPs-2) was not a Conceptualization, Data curation, Formal analysis, Methodology,
function of a single factor but rather a combination of several factors. In Writing - original draft. Mingyue Zhang: Methodology, Writing - orig­
this study, the polysaccharides that exhibited higher Mw possessed inal draft. Cheng Li: Supervision. Zhiqing Zhang: Resources. Bin Hu:
stronger antioxidant properties, which agreed with our previous Methodology. Aiping Liu: Visualization. Qin Li: Software. Hong Chen:
research (Zhang et al., 2018). On the other hand, Gu et al. (2020) re­ Project administration. Zizhong Tang: Validation. Wenjuan Wu:
ported that the polysaccharides from Sagittaria sagittifolia L. with lower Investigation. Daiwen Chen: Supervision.
Mw possessed stronger antioxidant activity because the polysaccharide
with low Mw can present a higher content of reducing end and accept
more free radicals. Therefore, the inconsistency in findings between Declaration of Competing Interest
these studies suggests that the polysaccharide Mw is not the primary
factor affecting its antioxidant activity. For instance, the effects of There are no conflicts to declare.
glucose and mannose on the activity of antioxidants were significant
according to the result of Zhang et al. (2016). Also, galactose, mannose, Acknowledgments
arabinose, rhamnose, xylose, etc. are important compositions of the
polysaccharide, which have antioxidant activity (An et al., 2020). Be­ This work was financially supported by the Sichuan Science and
sides, studies have outlined that high acidic monosaccharide contents Technology Program (2020JDRC0085); the Scientific Research Project
are favorable for the antioxidant activities of polysaccharides (Miao of Sichuan Cuisine Development Research Center (CC20Z09); the
et al., 2020), this is consistent with our research results. Polysaccharides Technology Innovation R&D Project of Chengdu City (2019-YFYF-
(CCPs-1 and CCPs-2) possessed better performances in antioxidant as­ 00045-SN).
says, acidic polysaccharides were more likely act as secondary antioxi­
dants because the carboxyl groups of the uronic acid of the
Appendix A. Supplementary data
polysaccharides might play the role of hydrogen-donating and
electron-transfer agent (Zhang et al., 2013).
Supplementary material related to this article can be found, in the
In addition, the DPPH radical scavenging activity and ferrous ion
online version, at doi:https://doi.org/10.1016/j.indcrop.2020.113079.
reducing power of cmCCPs-2 were close to BHT, which indicated that
cmCCPs-2 possessed high antioxidant activity. Studies have shown that
the presence of polyelectron materials can enhance the free radical References
scavenging activity of antioxidants, a major reason for this finding may An, Q., Ye, X., Han, Y., Zhao, M., Chen, S., Liu, X., Li, X., Zhao, Z., Zhang, Y., Ouyang, K.,
be attributed to the fact that the carboxymethylated samples were more Wang, W., 2020. Structure analysis of polysaccharides purified from Cyclocarya
water-soluble and had a higher surface area, which facilitated their paliurus with DEAE-Cellulose and its antioxidant activity in RAW264.7 cells. Int. J.
Biol. Macromol. 157, 604–615.
contact with the radical (Liu et al., 2017a). According to the previous
Cao, Y.Y., Ji, Y.H., Liao, A.M., Huang, J.H., Thakur, K., Li, X.L., Hu, F., Zhang, J.G.,
study, some functional groups such as OH, COOH, PO3H2, and CO can Wei, Z.J., 2020. Effects of sulfated, phosphorylated and carboxymethylated
combine with Fe2+. Therefore, this perhaps may explain the high Fe2+ modifications on the antioxidant activities in-vitro of polysaccharides sequentially
chelation activity of cmCCPs associated with these functional groups extracted from Amana edulis. Int. J. Biol. Macromol. 146, 887–896.
Duan, S., Zhao, M., Wu, B., Wang, S., Yang, Y., Xu, Y., Wang, L., 2020. Preparation,
(Cao et al., 2020). characteristics, and antioxidant activities of carboxymethylated polysaccharides
Numerous previous studies have isolated different polysaccharides from blackcurrant fruits. Int. J. Biol. Macromol. 155, 1114–1122.

7
Y. Liu et al. Industrial Crops & Products 159 (2021) 113079

Ferreira, S.S., Passos, C.P., Madureira, P., Vilanova, M., Coimbra, M.A., 2015. Structure- Liu, Y., Tang, T., Duan, S., Li, C., Lin, Q., Wu, H., Liu, A., Hu, B., Wu, D., Li, S., Shen, L.,
function relationships of immunostimulatory polysaccharides: a review. Carbohydr. Wu, W., 2020g. The purification, structural characterization and antidiabetic
Polym. 132, 378–396. activity of a polysaccharide from Anoectochilus roxburghii. Food Funct. 11,
Gu, J., Zhang, H., Wen, C., Zhang, J., He, Y., Ma, H., Duan, Y., 2020. Purification, 3730–3740.
characterization, antioxidant and immunological activity of polysaccharide from Miao, J., Regenstein, J.M., Qiu, J., Zhang, J., Zhang, X., Li, H., Zhang, H., Wang, Z., 2020.
Sagittaria sagittifolia L. Food Res. Int. 136, 109345. Isolation, structural characterization and bioactivities of polysaccharides and its
Guo, M.Z., Meng, M., Duan, S.Q., Feng, C.C., Wang, C.L., 2019. Structure derivatives from Auricularia-A review. Int. J. Biol. Macromol. 150, 102–113.
characterization, physicochemical property and immunomodulatory activity on Perepelov, A.V., Chen, T., Senchenkova, S.N., Filatov, A.V., Song, J., Shashkov, A.S.,
RAW264.7 cells of a novel triple-helix polysaccharide from Craterellus cornucopioides. Liu, B., Knirel, Y.A., 2018. Structure and genetics of the O-specific polysaccharide of
Int. J. Biol. Macromol. 126, 796–804. Escherichia coli O27. Carbohydr. Res. 456, 1–4.
He, Y., Ye, M., Jing, L., Du, Z., Surahio, M., Xu, H., Li, J., 2015. Preparation, Shakhmatov, E.G., Toukach, P.V., Makarova, E.N., 2020. Structural studies of the pectic
characterization and bioactivities of derivatives of an exopolysaccharide from polysaccharide from fruits of Punica granatum. Carbohydr. Polym. 235, 115978.
Lachnum. Carbohydr. Polym. 117, 788–796. Sharifian-Nejad, M.S., Shekarchizadeh, H., 2019. Physicochemical and functional
Hu, L., Liu, R., Wu, T., Sui, W., Zhang, M., 2020. Structural properties of homogeneous properties of oleaster (Elaeagnus angustifolia L.) polysaccharides extracted under
polysaccharide fraction released from wheat germ by hydrothermal treatment. optimal conditions. Int. J. Biol. Macromol. 124, 946–954.
Carbohydr. Polym. 240, 116238. Shi, H., Wan, Y., Li, O., Zhang, X., Xie, M., Nie, S., Yin, J., 2020. Two-step hydrolysis
Huo, J., Wu, J., Huang, M., Zhao, M., Sun, W., Sun, X., Zheng, F., 2020. Structural method for monosaccharide composition analysis of natural polysaccharides rich in
characterization and immuno-stimulating activities of a novel polysaccharide from uronic acids. Food Hydrocoll. 101, 105524.
Huangshui, a byproduct of Chinese Baijiu. Food Res. Int. 136, 109493. Sigida, E.N., Shashkov, A.S., Zdorovenko, E.L., Konnova, S.A., Fedonenko, Y.P., 2020.
Kokoulin, M.S., Lizanov, I.N., Romanenko, L.A., Chikalovets, I.V., 2020. Structure of Structure of the O-specific polysaccharide from Azospirillum formosense CC-Nfb-7(T).
phosphorylated and sulfated polysaccharides from lipopolysaccharide of marine Carbohydr. Res. 494, 108060.
bacterium Marinicella litoralis KMM 3900T. Carbohydr. Res. 490, 107961. Velichko, N.S., Kokoulin, M.S., Sigida, E.N., Kuchur, P.D., Komissarov, A.S., Kovtunov, E.
Kostalova, Z., Hromadkova, Z., 2019. Structural characterisation of polysaccharides from A., Fedonenko, Y.P., 2020. Structural and genetic characterization of the colitose-
roasted hazelnut skins. Food Chem. 286, 179–184. containing O-specific polysaccharide from the lipopolysaccharide of Herbaspirillum
Li, N., Mao, W., Yan, M., Liu, X., Xia, Z., Wang, S., Xiao, B., Chen, C., Zhang, L., Cao, S., frisingense GSF30(T). Int. J. Biol. Macromol. 161, 891–897.
2015. Structural characterization and anticoagulant activity of a sulfated Wang, Y., Guo, M., 2020. Purification and structural characterization of polysaccharides
polysaccharide from the green alga Codium divaricatum. Carbohydr. Polym. 121, isolated from Auricularia cornea var. Li. Carbohydr. Polym. 230, 115680.
175–182. Wang, Z.J., Xie, J.H., Shen, M.Y., Tang, W., Wang, H., Nie, S.P., Xie, M.Y., 2016a.
Li, S., Xia, H., Xie, A., Wang, Z., Ling, K., Zhang, Q., Zou, X., 2020. Structure of a fucose- Carboxymethylation of polysaccharide from Cyclocarya paliurus and their
rich polysaccharide derived from EPS produced by Kosakonia sp. CCTCC M2018092 characterization and antioxidant properties evaluation. Carbohydr. Polym. 136,
and its application in antibacterial film. Int. J. Biol. Macromol. 159, 295–303. 988–994.
Liu, Y., You, Y., Li, Y., Zhang, L., Tang, T., Duan, X., Li, C., Liu, A., Hu, B., Chen, D., Wang, X.T., Zhu, Z.Y., Zhao, L., Sun, H.Q., Meng, M., Zhang, J.Y., Zhang, Y.M., 2016b.
2017a. Characterization of carboxymethylated polysaccharides from Catathelasma Structural characterization and inhibition on alpha-d-glucosidase activity of non-
ventricosum and their antioxidant and antibacterial activities. J. Funct. Foods 38, starch polysaccharides from Fagopyrum tartaricum. Carbohydr. Polym. 153, 679–685.
355–362. Wang, J., Yang, W., Tang, Y., Xu, Q., Huang, S., Yao, J., Zhang, J., Lei, Z., 2016c.
Liu, Y., You, Y., Li, Y., Zhang, L., Yin, L., Shen, Y., Li, C., Chen, H., Chen, S., Hu, B., Regioselective sulfation of Artemisia sphaerocephala polysaccharide: solution
Chen, D., 2017b. The characterization, selenylation and antidiabetic activity of conformation and antioxidant activities in vitro. Carbohydr. Polym. 136, 527–536.
mycelial polysaccharides from Catathelasma ventricosum. Carbohydr. Polym. 174, Xu, J., Liu, W., Yao, W., Pang, X., Yin, D., Gao, X., 2009. Carboxymethylation of a
72–81. polysaccharide extracted from Ganoderma lucidum enhances its antioxidant activities
Liu, H., Fan, H., Zhang, J., Zhang, S., Zhao, W., Liu, T., Wang, D., 2020a. Isolation, in vitro. Carbohydr. Polym. 78, 227–234.
purification, structural characteristic and antioxidative property of polysaccharides Yang, L., Zhao, T., Wei, H., Zhang, M., Zou, Y., Mao, G., Wu, X., 2011.
from A. Cepa L. Var. Agrogatum Don. Food Sci. Hum. Wellness 9, 71–79. Carboxymethylation of polysaccharides from Auricularia auricula and their
Liu, X., Xu, S., Ding, X., Yue, D., Bian, J., Zhang, X., Zhang, G., Gao, P., 2020b. Structural antioxidant activities in vitro. Int. J. Biol. Macromol. 49, 1124–1130.
characteristics of Medicago sativa L. Polysaccharides and Se-modified Yang, W.W., Wang, L.M., Gong, L.L., Lu, Y.M., Pan, W.J., Wang, Y., Zhang, W., Chen, Y.,
polysaccharides as well as their antioxidant and neuroprotective activities. Int. J. 2018. Structural characterization and antioxidant activities of a novel
Biol. Macromol. 147, 1099–1106. polysaccharide fraction from the fruiting bodies of Craterellus cornucopioides. Int. J.
Liu, Y., Duan, X., Duan, S., Li, C., Hu, B., Liu, A., Wu, Y., Wu, H., Chen, H., Wu, W., Biol. Macromol. 117, 473–482.
2020c. Effects of in vitro digestion and fecal fermentation on stability and metabolic Zhang, H., Li, J., Xia, J., Lin, S., 2013. Antioxidant activity and physicochemical
behavior of polysaccharides from Craterellus cornucopioides. Food Funct. 11, properties of an acidic polysaccharide from Morinda officinalis. Int. J. Biol.
6899–6910. Macromol. 58, 7–12.
Liu, Y., Duan, X., Zhang, M., Li, C., Zhang, Z., Liu, A., Hu, B., He, J., Wu, D., Chen, H., Zhang, H., Zhao, H., Zhou, X., Yang, X., Shen, S., Wang, J., Wang, Z., Geng, L., 2016.
Wu, W., 2020d. Cooking methods effect on the nutrients, bioaccessibility and Isolation and characterization of antioxidant polysaccharides (PKCP-D70-2-a and
antioxidant activity of Craterellus cornucopioides. LWT. 131, 109768. PKCP-D70-2-b) from the Pinus koraiensis pinecone. RSC Adv. 6, 110706–110721.
Liu, Y., Li, Y., Zhang, H., Li, C., Zhang, Z., Liu, A., Chen, H., Hu, B., Luo, Q., Lin, B., Zhang, L., Hu, Y., Duan, X., Tang, T., Shen, Y., Hu, B., Liu, A., Chen, H., Li, C., Liu, Y.,
Wu, W., 2020e. Polysaccharides from Cordyceps miltaris cultured at different pH: 2018. Characterization and antioxidant activities of polysaccharides from thirteen
sugar composition and antioxidant activity. Int. J. Biol. Macromol. 162, 349–358. boletus mushrooms. Int. J. Biol. Macromol. 113, 1–7.
Liu, Y., Liu, Y., Zhang, M., Li, C., Zhang, Z., Liu, A., Wu, Y., Wu, H., Chen, H., Hu, X., Zhang, Z., Guo, L., Yan, A., Feng, L., Wan, Y., 2020. Fractionation, structure and
Lin, B., Wu, W., 2020f. Structural characterization of a polysaccharide from Suillellus conformation characterization of polysaccharides from Anoectochilus roxburghii.
luridus and its antidiabetic activity via Nrf2/HO-1 and NF-κB pathways. Int. J. Biol. Carbohydr. Polym. 231, 115688.
Macromol. 162, 935–945.

You might also like