Structural analysis, in vitro antioxidant and lipid-lowering activities of purified Tremella fuciformis polysaccharide fractions

DOI: https://doi.org/10.21203/rs.3.rs-2327340/v1

Abstract

In order to study the composition, structure and physicochemical properties of Tremella fuciformis polysaccharide (TFPS), four new polysaccharide fractions (TFPS1, TFPS2, TFPS3 and TFPS4) were successfully isolated from the hot-water extracted crude polysaccharides by DEAE-52 cellulose column. Their structures were analyzed by FT-IR, SEM, HPGPC, monosaccharide composition, Congo red and I2-KI experiments. TFPS1 was a heteropolysaccharide with a molecular weight of 742 Da for the major polysaccharide fragment. TFPS2, TFPS3 and TFPS4 were single component homogeneous polysaccharides with 350774 Da, 622675 Da and 624724 Da, respectively. Glucose was the major monosaccharide of TFPS1 and TFPS2, while Mannose was the major monosaccharide of TFPS3 and TFPS4. TFPS1 and TFPS2 displayed a stable triple helix structure at aqueous solution, while TFPS3 and TFPS4 presented longer side chains and more branched chains. TFPS1, TFPS2, TFPS3 and TFPS4 all possessed in vitro antioxidant activities. Moreover, the four polysaccharide fractions all demonstrated significant cholesterol binding ability, as well as obvious sodium taurocholate binding ability. The results of this study may provide some guidance on the medicinal value of TFPS.

1. Introduction

As the fruiting body of Basidiomycota fungus [1], Tremella fuciformis is not only an edible fungus with economic value, but also a long-established medicine in Chinese medicine [2]. It is rich in protein, polysaccharides, dietary fiber, many minerals, trace elements and vitamins [3, 4]. Among them, the main active ingredient of Tremella fuciformis is Tremella fuciformis polysaccharide (TFPS) [5]. Related studies have demonstrated that TFPS has anti-tumor [6], antioxidant [7], moisturizing [810], hypolipidemic [11], anti-inflammatory [12, 13] and antibacterial functions [11]. TFPS can be used in a variety of applications such as pharmaceuticals [14], cosmetics [15] and food [16, 17]. It exhibited inhibitory effect on cholesterol micelles in a concentration-dependent manner in vitro [18]. Moreover, Dextran sodium sulfate-induced colitis can be alleviated by high-dose TFPS immunomodulation. TFPS can reduce colon tissue damage, promote the production of anti-inflammatory cytokines, and restore the intestinal microbiota and microbial metabolites [19, 20]. Furthermore, TFPS solution at certain concentration can effectively reduce UVA damage to skin fibroblasts [21]. In addition, Hepatic fatty acid synthesis can be regulated by TFPS, which suppresses NAFLD mainly by decreasing the expression of fatty acid synthesis regulators [22].

TFPS is a kind of heterosaccharide, different fractions of polysaccharides should have different biological activities [23]. Although bioactivities of the heterosaccharide have been extensively studied, it is not very clear about the relationship between structure, conformation and bioactivity of their isolated and purified fractions. In this study, TFPS were separated by ion exchange column chromatography. And four polysaccharide fractions (TFPS1, TFPS2, TFPS3 and TFPS4) were isolated and purified from the crude TFPS. They were physically characterized by FT-IR and SEM, and their structural features were tested in combination with monosaccharide composition, Congo red and I2-KI experiments. The in vitro antioxidant activity, cholesterol binding and sodium taurocholate binding assays were also tested for the four polysaccharide fractions. The information obtained can provide a theoretical basis for better development and utilization of TFPS.

2. Materials And Methods

2.1 Materials and Reagents

Tremella fuciformis were commercially available from Gutian, Fujian (China). DEAE-52 cellulose was purchased from Shanghai Dipper Biotechnology Co. Sulfuric acid (H2SO4) and sodium hydroxide (NaOH) were bought from Tianjin Windship Chemical Reagent Technology Co. Congo red and sodium chloride (NaCl) were obtained from Aladdin Biochemical Technology Co. Phenol (C6H6O) was acquired from Tianjin Yongda Chemical Reagent Co. Salicylic acid (C7H6O3) was procured from Zhengzhou Jinbei Chemical Co. Anhydrous ethanol (C2H6O) was received from Shandong Bosheng Chemical Co. Standard monosaccharides (including Mannose ( Man), ribose (Rib), rhamnose (Rha), glucuronic acid (Glc-UA), galacturonic acid (Gal-UA), N-acetyl-aminoglucose (Glc-NAc), glucose (Glc), N-acetyl-aminogalactose (Gal-NAc), galactose (Gal), xylose (Xyl), arabinose (Ara) and fucose (Fuc)) were purchased from Shanghai Yuanye Biotechnology Co. The water used in the experiments was distilled water. All chemicals were of analytical grade and used straight away without further purification.

2.2 Experimental Methods

2.2.1 Extraction of TFPS

Extraction of TFPS was carried out by hot water extraction method. The dried Tremella fuciformis (3 g) was weighed for crushing and placed in a beaker. The obtained powder was stirred in 180 mL distilled water and refluxed for 7 h. After filtering through a double layer of filter cloth, 200 ml of distilled water was added to the filter residue, heated and stirred all the time for 2 h. After filtration, the two filtrates were combined and concentrated under vacuum to 40 mL. To the filtrate, four times the amount of ethanol (90%) was added. The filtrate was left overnight at 4°C in a refrigerator and separated by centrifugation (8000 r/min, 8 min). The obtained precipitate was dissolved with of water. Proteins were removed by static adsorption with AB-8 macroporous resin. Then, the TFPS solution was dialyzed (8000 Da) to remove small molecules and inorganic ions. Finally, the dialysate was vacuum concentrated to 30 mL and lyophilized using an lyophilizer (FD-1A-50, Shanghai, China), then the TFPS solid was obtained.

2.2.2 Separation and purification

The flow chart for the preparation of TFPS1, TFPS2, TFPS3 and TFPS4 were shown in Fig. S1. Separation was performed by DEAE-52 cellulose column. The TFPS was prepared as a 10 mg/mL aqueous solution sample with a loading volume of 100 mL. The sample was eluted with distilled water, 0.2, 0.4, 0.6 and 2 mol/L NaCl solution in a gradient at a flow rate of 4 mL/min, and one tube was collected every 10 mL. The purified fractions corresponding to the peaks were collected under the monitoring of phenol-sulfuric acid method [24]. The four purified polysaccharide fractions were dialyzed for 72 h and then lyophilized before further study.

2.3 Analysis of total sugar content and yield

The total sugar content was measured by the sulfuric acid-phenol method using glucose as the standard [25].

$$\text{Y}\text{i}\text{e}\text{l}\text{d}(\text{%})={\text{M}}_{1}/{\text{M}}_{2}\times 100\text{%}$$
1
$$\text{T}\text{o}\text{t}\text{a}\text{l} \text{s}\text{u}\text{g}\text{a}\text{r} (\text{%})=(\text{A}\times \text{B}\times \text{D})/{\text{M}}_{3}\times 100\text{%}$$
2

Where: M1 indicates the mass of each fraction after separation and purification, mg; M2 represents the mass of polysaccharide before separation and purification, mg; A is the mass concentration of the solution calculated from the absorbance value, mg/mL; B shows the dilution of the sample; D means the volume of the test solution, mL; M3 expresses the mass of the sample, mg.

2.4 UV analysis

The UV spectra of TFPS1, TFPS2, TFPS3 and TFPS4 were recorded in the range of 200 ~ 400 nm using an Ultraviolet spectrophotometer (UV, TU-1810, Beijing, China) [26].

2.5 Structural Analysis

2.5.1 Molecular weight determination

The average molecular weights of TFPS1, TFPS2, TFPS3 and TFPS4 were determined by high performance gel permeation chromatography (HPGPC) on a gel chromatograph (GPC, Model 515, Waters, USA) [27]. The analytical columns of this instrument were Waters Ultrahydrogel 500 and 120 in two tandem, mobile phase: 0.1 mol/L aqueous NaNO3 solution at a flow rate of 1 mL/min.

2.5.2 Monosaccharide composition analysis

The monosaccharide compositions of TFPS1, TFPS2, TFPS3 and TFPS4 were examined by 1-phenyl-3-methyl-5-pyrazolone (PMP) derivatization [28]. First, weigh 20 mg of polysaccharide sample in a 10 ml ampoule, added 3 ml of trifluoroacetic acid (TFA, 2 mol/L), acid digestion at 120 ℃ for 4 h. Added methanol nitrogen blowing to evaporate the TFA, then added water to redissolve to get 1 mg/ml acid hydrolysis sample solution. Mix 250 uL of sample solution, 250 uL of 0.6 mol/L NaOH and 500 uL of 0.4 mol/L LPMP-methanol, and reacted at 70°C for 1 h. After cooling for 10 min, added 500 uL of 0.3 mol/L HCl to neutralize, and then added 1 mL of chloroform to extract 3 times. The supernatant was taken and used for monosaccharide composition analysis. The analysis was performed at 30°C using a high performance liquid chromatograph (HPLC, Shimadzu LC-20AD, Japan) with an Xtimate C18 (4.6*200 mm 5 um). The control standard monosaccharides were Man, Rib, Rha, Glc-UA, Gal-UA, Glc-NAc, Glc, Gal-NAc, Gal, Xyl, Ara and Fuc.

2.5.3 FT-IR spectral analysis

The functional groups of TFPS1, TFPS2, TFPS3 and TFPS4 were examined using FT-IR [29]. A sample of 5 mg polysaccharide was weighed, mixed with 100 mg KBr and pressed. A spectrometer (FT-IR, Nicolet iS50, ThermoFisher, China) was used in the wavenumber range of 4000 ~ 600 cm− 1.

2.5.4 SEM text

The surface morphology of polysaccharides is frequently analyzed by SEM. TFPS1, TFPS2, TFPS3 and TFPS4 were attached to the sample stage using conductive adhesive followed by spraying a thin gold layer [30]. The gold sprayed samples were placed in a scanning electron microscope (SEM, SU3500, HITACHI, Japan) and the images were magnified at 1000×.

2.5.5 Iodine-potassium iodide (I2-KI) experiments

The side chain structures of TFPS1, TFPS2, TFPS3 and TFPS4 were determined by I2-KI experiments [31]. The sample solution (1 mg/mL) was mixed with iodine reagent (0.2% KI solution containing 0.02% I2), which was scanned in the UV-vis light range from 300 to 700 nm.

2.5.6 Congo red experiment

The triple helix structures of TFPS1, TFPS2, TFPS3 and TFPS4 were identified using Congo red experiments [32]. The sample solution (2 mg/mL) was mixed with Congo red solution dissolved in different concentrations of NaOH with UV-vis light scanning in the range of 300 ~ 700 nm.

2.6 Antioxidant studies in vitro

TFPS1, TFPS2, TFPS3 and TFPS4 samples were prepared as 2 mg/mL solutions as well as diluted into 0.1 mg/mL, 0.2 mg/mL, 0.5 mg/mL, 1 mg/mL and 1.5 mg/mL sample solutions, respectively.

2.6.1 Determination of ferric ion reducing power

The reduction power of TFPS1, TFPS2, TFPS3 and TFPS4 was tested according to the previous reported method [33], but with minor modifications. Mix 1 mL of sample solution at different concentrations, 2.5 mL of phosphate buffer at pH 6.6 (0.2 mol/L) and 2.5 mL of 1% potassium ferricyanide, and react in a water bath at 50°C for 20 min. Afterwards, 2.5 mL of 10% (W/V) trichloroacetic acid was added and centrifuged at 4000 r/min for 10 min. Then, 2.5 mL of supernatant was taken, 2.5 mL of distilled water and 0.5 mL of 0.1% FeCl3 were added and mixed, and the sample was left for 10 min. Finally, the absorbance of the reaction solution was measured at 700 nm and zeroed using distilled water instead of FeCl3 solution as a blank. The higher is the value, the higher is the reducing power of the sample. The reducing power was calculated by Eq. (3):

$${\text{A}}_{1}={\text{A}}_{\text{X}}-{\text{A}}_{0}$$
3

Where: A1 indicates reducing power of Fe3+; AX denotes the absorbance value of the sample group; A0 means the absorbance of the blank control group.

2.6.2 Scavenging capacity of DPPH free radicals

The DPPH radical scavenging ability of TFPS1, TFPS2, TFPS3 and TFPS4 was examined based on previously reported methods with appropriate modifications [34, 35]. The samples were mixed and shaken with 3 mL of polysaccharide solution and 1 mL of DPPH (0.1 mmol/mL) in anhydrous ethanol, and then reacted for 1 h at room temperature and protected from light. The absorbance value of the supernatant was measured at 517 nm and recorded as ASample, followed by an equal volume of distilled instead of the sample to be measured and recorded as ABlank. As well as replacing the DPPH solution with an equal volume of anhydrous ethanol, the absorbance value was measured as AControl. DPPH scavenging activity was calculated by Eq. (4):

\({\text{A}}_{2} (\text{%})=[1-({\text{A}}_{\text{S}\text{a}\text{m}\text{p}\text{l}\text{e}}-{\text{A}}_{\text{C}\text{o}\text{n}\text{t}\text{r}\text{o}\text{l}})/{\text{A}}_{\text{B}\text{l}\text{a}\text{n}\text{k}}]\times 100\text{%}\) ()

Where: A2 denotes DPPH radical scavenging rate; ASample represents the absorbance value of sample group; AControl indicates the absorbance value of sample control group; ABlank means the absorbance value of blank control group.

2.6.3 Scavenging capacity of hydroxyl radicals

The hydroxyl radical scavenging ability of TFPS1, TFPS2, TFPS3 and TFPS4 was determined on the basis of the reported method with slight modifications [36]. Different concentrations of 1 mL of sample solution, 1 ml (9 mmol/L) salicylic acid-ethanol solution, 1 mL (9 mmol/L) FeSO4 and 1 ml (8.8 mmol/L) H2O2 solution were mixed and reacted in a water bath at 37°C for 1 h. After the reaction solution was cooled to room temperature, its absorbance value was measured at a wavelength of 510 nm. Under the same conditions the absorbance values of the blank control solution were determined using distilled water instead of the polysaccharide solution. The absorption values of the background solution were recorded without the addition of H2O2 solution. The hydroxyl radical scavenging activity was calculated by Eq. (5).

\({\text{A}}_{3} (\text{%})={[\text{A}}_{\text{c}}-{(\text{A}}_{\text{i}}-{\text{A}}_{\text{j}}\left)\right]/{\text{A}}_{\text{c}}\times 100\text{%}\) ()

Where: A3 denotes the hydroxyl radical scavenging rate; Ai indicates the absorbance value of sample solution; Aj represents the absorbance value of the background solution; Ac refers to the absorbance value of blank control solution.

2.7 In vitro hypolipidemic activity assay

2.7.1 Determination of cholesterol binding amount

TFPS1, TFPS2, TFPS3 and TFPS4 samples were prepared as 10 mg/mL solutions. 5 mL of different sample solutions were mixed with 10 mL of cholesterol solution (1 mg/mL). In order to simulate the gastrointestinal environment, the pH values were adjusted to 2 and 7, respectively, and shaken for 2 h at a constant speed of 80 r/min at 37°C. Then, 0.4 mL of the mixed solution was taken and mixed thoroughly with 0.2 mL of o-phthalaldehyde solution and 4 mL of mixed acid solution (Vsulfuric acid : Vhydrochloric acid = 1:1) in a 37°C water bath for 10 minutes. Afterwards, the supernatant was centrifuged at 4000 r/min for 20 min and the absorbance value of the solution was measured at 550 nm. The values were determined by the cholesterol standard curve, which was developed at the concentrations of 0.05, 0.075, 0.1, 0.15 and 0.2 mg/mL of standard cholesterol solution [37]. The measured cholesterol standard curve was: Y = 4.6X + 0.045, R2 = 0.988. The cholesterol binding was calculated by Eq. (6):

$$\text{B}\text{i}\text{n}\text{d}\text{i}\text{n}\text{g} \text{a}\text{m}\text{o}\text{u}\text{n}\text{t} (\text{m}\text{g}/\text{g})={(\text{C}}_{1}-{\text{C}}_{2})/{\text{M}}_{4}$$
6

Where: C1 indicates the amount of cholesterol added, mg; C2 shows the amount of cholesterol remaining, mg; M4 indicates the sample mass, g.

2.7.2 Determination of the binding rate of sodium taurocholate

TFPS1, TFPS2, TFPS3 and TFPS4 samples were prepared as 2 mg/mL solutions. The sample solution (2 mg/mL) was mixed with 1 ml of HCl solution (0.01 mol/L) and allowed to shake at 37°C for 1 h. Subsequently, 4 mL of trypsin (10 mg/mL) was added and shaken for 1 h. Followed by 4 mL of sodium taurocholate solution (2 mg/mL) was added and shaken again for 1 h. Then, 2.5 mL of supernatant was centrifuged at 5 000 r/min for 20 min, and 2.5 mL of supernatant was mixed with 7.5 mL of sulfuric acid solution (60%), incubated at 70°C for 30 min and then was cooled to room temperature. Finally, the absorbance value of each sample was measured at a wavelength of 387 nm. The values were determined by the standard curve of sodium taurocholate, which was developed at the concentrations of 0.05, 0.075, 0.1, 0.15, and 0.2 mg/mL of standard sodium taurocholate solution [38]. The measured standard curve of sodium taurocholate was: Y = 3.7738X + 0.028, R2 = 0.9823. The binding rate of sodium taurocholate was calculated by Eq. (7):

$$\text{B}\text{i}\text{n}\text{d}\text{i}\text{n}\text{g} \text{r}\text{a}\text{t}\text{e} \left(\text{%}\right)=[1-(\text{C}\times \text{V})/4]\times 100\text{%}$$
7

Where: C denotes the concentration of sodium taurocholate in the supernatant obtained by the standard curve, mg/ml; V stands for the volume of the supernatant, ml.

3. Results And Discussion

3.1 Isolation and purification of TFPS

100ml TFPS solution (10 mg/mL) was added to the DEAE-52 cellulose column, and five elution peaks were obtained by sequential elution with deionized water and different concentrations of NaCl aqueous solution, as shown in Fig. 1. The four purified polysaccharide fractions (TFPS1, TFPS2, TFPS3 and TFPS4) eluted with 0, 0.2, 0.4 and 0.6 mol/L NaCl solutions were collected, respectively. The fifth fraction eluted with 2 mol/L NaCl solution was not collected because of its very low polysaccharide content and the feasibility of subsequent operations. It was observed that TFPS1 was a light yellow loose flocculent polysaccharide, while TFPS2, TFPS3 and TFPS4 were all white loose flocculent polysaccharides. The solubility analysis revealed that the four purified polysaccharide fractions were soluble in hot water, but insoluble in cold water and in organic solvents such as ethanol, methanol and acetone.

3.2 Analysis of total sugar content and yield

The total sugar content of the four purified polysaccharide fractions were determined by sulfuric acid-phenol method. The glucose standard curve was shown in Fig. S2 (a). The yield and total sugar content were calculated according to Equations (1) and (2). The yields of TFPS1, TFPS2, TFPS3 and TFPS4 were 5.7%, 6.19%, 66.68% and 16.54%, respectively. And the total sugar contents of the four purified polysaccharide fractions were 88.79%, 83.43%, 98.88% and 94.33%, respectively.

3.3. UV analysis

The UV scan showed that TFPS1, TFPS2, TFPS3 and TFPS4 had no obvious absorption peaks at 260 nm and 280 nm (As shown in Fig. S2 (b)). The results indicated that none of the four purified polysaccharide fractions contained protein and nucleic acid-like substances [29].

3.4 Structural Analysis

3.4.1 Analysis of molecular weight

The homogeneity and molecular weights of TFPS1, TFPS2, TFPS3 and TFPS4 were determined by HPGPC. Figure 2 showed the molecular weight results for the four purified polysaccharide fractions. TFPS1 had a significant peak and two small peaks. The main component of TFPS1 was a small-sized polysaccharide fragment with a molecular weight of 742 Da. In addition, there were two small amounts of large fragments with molecular weights of 18938 Da and 200468 Da. TFPS2, TFPS3 and TFPS4 were single component homogeneous polysaccharides with 350774 Da, 622675 Da and 624724 Da, respectively.

3.4.2 Monosaccharide composition analysis

According to the relevant literature [39], TFPS1, TFPS2, TFPS3, and TFPS4 were hydrolyzed to monosaccharides by trifluoroacetic acid and NaOH, and subsequently derivatized by LPMP methanol. Then their monosaccharide compositions were known by HPLC analysis (as shown in Fig. S3). The compositions and molar ratios were shown in Table 1. The monosaccharide compositions of the four purified polysaccharides were slightly different. Glc was the main monosaccharide of TFPS1 and TFPS2, while Man was the major monosaccharide of TFPS3 and TFPS4.

Table 1

Composition and molar ratio of polysaccharide monosaccharides

 

Man

Rib

Rha

Glc-UA

Gal-UA

Glc

Gal

Xyl

Ara

Fuc

TFPS1

3.407

0.01

0

0.399

0.014

6.71

0.194

1.298

0.056

0.546

TFPS2

2.122

0.048

0

0.446

0.021

25.741

0.609

0.565

0.072

0.186

TFPS3

10.553

0.01

0.02

1.819

0.056

2.962

0.093

3.136

0.072

2.281

TFPS4

9.18

0.1

0.016

1.785

0.035

0.994

0.057

2.553

0.068

2.021

3.4.3 FT-IR spectral analysis

FT-IR can be used to analyze the type of functional groups in polysaccharides because of their high characterization. Figure 3 displayed the scanned FT-IR spectra of TFPS1, TFPS2, TFPS3 and TFPS4. The broad peak near the wave number 3320 cm− 1 was caused by the O-H stretching vibration in the sugar molecule, indicating the existence of inter- or intramolecular hydrogen bonds in the polysaccharide. The characteristic peaks at wave numbers 2931 cm− 1 represented the C-H stretching vibrations of methyl or methylene groups [40]. The characteristic peaks at wave number 1415 cm− 1 showed the C-H stretching vibrations of carboxyl groups and variable angle vibrations of C-H. While the characteristic peak at wave number 1248 cm− 1 indicated the contraction vibration of S = O, suggesting that the polysaccharide contains sulfate radicals [41, 42]. Furthermore, the characteristic peaks near the wave number 1603 cm− 1 were expressed as C = O asymmetric stretching vibration of COO- and symmetric stretching vibration of C = O, indicating the presence of uronic acids in polysaccharides [43]. Meanwhile, the absorption peak at 1019 cm− 1 corresponds to the absorption peak of the stretching vibration formed by the C-O-H and C-O-C of the pyranose ring. What’s more, the absorption peaks between 947 and 865 cm− 1 may represent the presence of β-glycosidic bonds in all four purified polysaccharides [44]. In addition, the characteristic absorption peak of the α-glycosidic bond was demonstrated at 844 cm. The presence of α-D-mannose was demonstrated by an absorption peak near 800 cm− 1 [45]. Therefore, it can be inferred that TFPS1 was a neutral polysaccharide with α-D-mannose as the main chain. TFPS2 was an acidic polysaccharide with α-mannose as the main chain. TFPS3 and TFPS4 were both acidic polysaccharides with α-D-mannose as the main chain. Moreover, all four purified polysaccharide fractions had both α-type and β-type glycosidic linkages.

3.4.5 SEM analysis

The microstructures of TFPS1, TFPS2, TFPS3 and TFPS4 were investigated by SEM. As showed in Fig. 4, there revealed differenct surface morphologies in size and shape. The TFPS1 exhibited powder accumulation, whose surface morphology showed a dense spherical shape. The surface of TFPS2 demonstrated interlaced extensions in the form of reticular aggregates. The surface of TFPS3 manifested staggered extensions of filaments and curled aggregates. As for TFPS4, there was smooth interlaced extension in the form of lamellar aggregates. The TFPS1 showed poor film-forming properties because it was a small molecular weight polysaccharide fraction. However, TFPS2, TFPS3 and TFPS4 were large molecular weight polysaccharide fractions and thus had good film-forming properties.

3.4.6 Analysis of I2-KI experiments

The conformations of the polysaccharides could be elucidated by the absorption spectra of their solution reacted with I2-KI. If the polysaccharide had longer side chains and more branched chains, the spectrum should have an absorption peak at 350 nm. If there was an absorption peak at 565 nm, the polysaccharide should have shorter side chains and fewer branched chains [46, 47]. The UV-vis spectra of the four purified polysaccharide fractions reacted with I2-KI solution were shown in Fig. 5. TFPS3 and TFPS4 exhibited absorption peaks at 350 nm, indicating that the two purified polysaccharide fractions possessed longer side chains and more branched chains. In contrast, TFPS1 showed small absorption peaks at 565 nm, suggesting that the purified polysaccharide fraction may manifest shorter side chains and fewer branched chains. However, TFPS2 exhibited no absorption peak at 350 nm and 565 nm, which could only indicate that its conformation is different from the other three purified polysaccharide fractions.

3.4.7 Analysis of Congo red experiment

Polysaccharides with triple helix structure could complex with Congo red solution. For the Congo red-polysaccharide complex solution, their maximum absorption wavelength of the complex should shift to the long-wave direction at a certain range of NaOH concentration, compared with that of the pure Congo red solution. When the concentration of NaOH is greater than 0.3 mol/L, the triple helix conformation of the complex may be disrupted and its maximum absorption wavelength should be gradually reduced [48]. The interactions of the four purified polysaccharide fractions with Congo red solution were shown in Fig. 6. For the solutions of the purified polysaccharide fractions and Congo red, the maximum absorption wavelengths gradually shifted to the long-wave direction with the increase of NaOH concentration, indicating that the complexes had been formed. Compared with the maximum absorption wavelengths of the pure Congo red solution, those of Congo red-TFPS1 and Congo red-TFPS2 solutions decreased significantly when the NaOH concentration was greater than 0.3 mol/L. The phenomenon indicated that TFPS1 and TFPS2 possessed the triple helix structure. However, the maximum absorption wavelengths of Congo red-TFPS3 and Congo red-TFPS4 solutions were not significantly reduced, suggesting that the triple helix structure may not be existed at the aqueous solution of TFPS3 and TFPS4.

3.5 In vitro antioxidant activity analysis of TFPS1, TFPS2, TFPS3 and TFPS4

3.5.1 The reducing power of ferric ion

TFPS can reduce Fe3+ of K3[Fe(CN)6] to Fe2+, and Fe2+ further reacts with FeCl3 to form Fe3[Fe(CN)6]2, which has the maximum absorbance at 700 nm. So, the absorbance was usually measured at 700 nm for the TFPS-Fe3+ reacting solutions [49]. The higher was the absorbance, the stronger did the TFPS display reduction ability. The results of the four TFPS reducing power test were illustrated in Fig. 7 (a). All the four purified polysaccharide fractions displayed certain reducing power. In a certain concentration range, the reducing power exhibited a quantitative influence relationship with concentration, and the reducing power increased with the increase of concentration. When the polysaccharide concentration reached 2 mg/mL, the four polysaccharides had the strongest reducing ability, and the reducing power was ranked from largest to smallest as TFPS1, TFPS4, TFPS3, and TFPS2. Therefore, TFPS1 has the best reducing ability. Its stronger reducing power may be related to the sulfate content [5052].

3.5.2 Scavenging capacity of DPPH free radicals

The anhydrous ethanol solution of DPPH• has a maximum absorption wavelength at 517 nm. Polysaccharides can scavenge DPPH• and decrease the absorbance of DPPH•, from which the DPPH• radical scavenging rate can be evaluated [53, 54]. The results of the four purified polysaccharide fractions scavenging DPPH• radical were summarized in Fig. 7 (b). It is obviously that TFPS1, TFPS2, TFPS3 and TFPS4 all possessed certain DPPH• radical scavenging ability. The DPPH• scavenging ability of TFPS4, TFPS3 and TFPS2 increased with the increase of polysaccharide concentrations. When the polysaccharide concentration reached 2 mg/mL, their scavenging rate were 48.07%, 46.53% and 31.1%, respectively. However, TFPS1 displayed the best DPPH• radical scavenging ability when the concentration was 1.5 mg/ml. The scavenging rate was 46.71%. In addition, TFPS2 displayed significantly smaller DPPH• scavenging ability than the other three TFPS fractions. The variation in DPPH• radical scavenging activity may be attributed to the differences in their sulfate content [32, 55, 56].

3.5.3 Scavenging capacity of hydroxyl radicals

H2O2 and Fe2+ gave rise to hydroxyl radicals, which could react with salicylic acid and produce 2,3-dihydroxybenzoic acid and 2,5-dihydroxybenzoic acid. The formed two acids have the maximum absorption wavelength at 510 nm [57]. When polysaccharides scavenged the hydroxyl radicals, the absorbance at 510 nm should decrease, from which the hydroxyl radical scavenging ability can be evaluated [58]. As shown in Fig. 7 (c), all the four purified polysaccharide fractions exhibited certain hydroxyl radical scavenging ability. And the hydroxyl radical scavenging rate increased with the increasing polysaccharide concentrations. When the polysaccharide concentration reached 2 mg/mL, the maximum hydroxyl radical scavenging rates for TFPS1, TFPS2, TFPS3 and TFPS4 were 64.13%, 38.76%, 37.66% and 41.18%, respectively. Thus, the results indicated that the four purified TFPS demonstrated significant free hydroxyl radical scavenging ability. The hydroxyl radical scavenging ability of TFPS1 was obviously higher than those of TFPS2, TFPS3 and TFPS4, probably owing to the better hydroxyl radical scavenging ability of polysaccharides with small molecular weight [59, 60].

3.6 Analysis of the in vitro hypolipidemic activity of TFPS1, TFPS2, TFPS3 and TFPS4

3.6.1 Cholesterol binding capacity in vitro

Many polysaccharides showed the potential to lower blood cholesterol [6163]. So, cholesterol binding tests were also performed for the four purified polysaccharide fractions in this paper. The total cholesterol content was determined by the o-phthalaldehyde method [64]. Cholesterol exists in the body as both free cholesterol and cholesterol esters, which can produce a purplish red substance in the presence of sulfuric acid and o-phthalaldehyde. The purplish red substance has a maximum absorption at 550 nm wavelength. So, the cholesterol binding capacity of the four TFPS were tested by detecting the absorption intensity at 550 nm of their mixed solutions, in the presence of sulfuric acid and o-phthalaldehyde. Since the pH value of the human stomach environment is 2 and that of the intestinal environment is 7, the pH values of the mixed solutions were adjusted to 2 and 7 to simulate the human environmental. As shown in Fig. 8 (a, b), pH had a great influence on the cholesterol binding amounts for the four purified polysaccharide fractions. The cholesterol binding amounts at pH = 7 was significantly higher than those at pH = 2. Among the four TFPS, TFPS2 showed the highest potential to lower cholesterol. At pH = 7, the cholesterol binding amount of TFPS2 reached 113.52 mg/g, followed by 97.28 mg/g for TFPS4 and 94.74 mg/g for TFPS1, and lowest in 91.74 mg/g for TFPS3. However, at pH = 2, the order of cholesterol binding capacity is TFPS2, TFPS1, TFPS4 and TFPS3.

3.6.2 Sodium taurocholate binding capacity in vitro

In the human body, the removal of cholesterol is also dependent on the synthesis bile acids by cholesterol and the secretion of bile acids. Moreover, bile acids are primarily present in the form of bile salts in the body. As bile salts combines with other substances and the free bile salts amount decreases, cholesterol can be further turned into bile acids. Therefore, cholesterol depletion can also be determined based on the binding ability of bile salts. Among the many bile salts, sodium taurocholate is the most difficult to bind, so the binding capacity with sodium taurocholate is usually used to determine the level of cholesterol-lowering capacity [6567]. Consequently, sodium taurocholate binding tests were also performed on TFPS1, TFPS2, TFPS3, and TFPS4 in this paper. The in vitro sodium taurocholate binding rates of the four purified polysaccharide fractions were presented in Fig. 8 (c). TFPS1, TFPS2, TFPS3 and TFPS4 expressed obvious binding ability to sodium taurocholate, and their sodium taurocholate binding rates were similar. The binding ability of the four purified polysaccharide fractions were ranked from largest to smallest as TFPS4, TFPS3, TFPS2, and TFPS1, with binding rates of 46.58%, 46.55%, 45.93%, and 45.86%, respectively.

4. Conclusion

The four purified polysaccharide fractions (TFPS1, TFPS2, TFPS3 and TFPS4) were isolated from the crude TFPS by DEAE-52 cellulose column. The monosaccharide composition analysis indicates that their monosaccharide compositions are slightly different. Glc was the major monosaccharide of TFPS1 and TFPS2, while Man was the major monosaccharide of TFPS3 and TFPS4. FI-IR analysis showed that the main chain of all four purified polysaccharides was pyranose. TFPS1 and TFPS2 exhibited stable triple helix structures, while TFPS3 and TFPS4 showed longer side chains and more branched chains. TFPS1, TFPS2, TFPS3 and TFPS4 all possessed significant DPPH• radical and hydroxyl radical scavenging effects and Fe3+ reducing power. Among them, TFPS1 had the strongest Fe3+ reduction and hydroxyl radical scavenging ability, while TFPS4 presented the strongest DPPH• radical scavenging ability. In addition, the four polysaccharide fractions all conferred obvious cholesterol and sodium taurocholate binding capacity, suggesting that they all exerted certain hypolipidemic activity. The cholesterol binding capacity at pH = 7 was higher than that at pH = 2. TFPS2 had the highest cholesterol binding rate, while TFPS4 expressed the highest sodium taurocholate binding rate. These may provide a theoretical basis and technical support for the purified TFPS’s application in the food and pharmaceutical industries.

Declarations

Author contributions

Xige Li: Methodolog, Data curation, Writing and Editing.

Yanhui Hou: Methodolog, Writing-Reviewing and Editing, Funding acquisition.

Fengyi Zhang: Data curation. 

Chaoxian Jiang: Data curation. 

Jun Jiang: Data curation. 

Funding

This work was supported by the small pilot project of PetroChina. LTD (PRlKY17031) and the Scientific research and development project of Zhengzhou Kema Biotechnology Co. LTD (ST22059).

Data availability All data generated or analysed during this study are included in this published article.

Compliance with ethical standards

Conflict of interests There are no conflicts of interest to declare.

Ethical approval This article does not contain any studies with human participants or animals performed by any of the authors

References

  1. S. Tsai, F. Huang, H. Juan, C. Lin.: Evaluation of food-processing conditions of various particle sizes of tremella fuciformis powder via dsc and tg analyses, J. Therm. Anal. Calorim. 134 (2) 857-864 (2018). https://doi.org/10.1007/s10973-018-7466-6.
  2. J.X. Xu, D.N. Wang, Y.P. Lei, L.J. Cheng, W.J. Zhuang, Y.T. Tian.: Effects of combined ultrasonic and microwave vacuum drying on drying characteristics and physicochemical properties of tremella fuciformis, Ultrason. Sonochem. 84 (2022). https://doi.org/10.1016/j.ultsonch.2022.105963.
  3. L. Wang, Q. Wu, J.Y. Zhao, X.Y. Lan, K. Yao, D.Y. Jia.: Physicochemical and rheological properties of crude polysaccharides extracted from tremella fuciformis with different methods, Cyta-J. Food 19 (1) 247-256 (2021). https://doi.org/10.1080/19476337.2021.1884607.
  4. C.P. Lin, S.Y. Tsai.: Differences in the moisture capacity and thermal stability of tremella fuciformis polysaccharides obtained by various drying processes, Molecules 24 (15) (2019). https://doi.org/10.3390/molecules24152856.
  5. L.D. Chen, J.Q. Chen, J.H. Li, J.C. Xie, M.M. Yu, M.L. Zhou, M.J. Xi, S.J. Sun.: Physicochemical properties and in vitro digestion behavior of a new bioactive tremella fuciformis gum, Int. J. Biol. Macromol. 207 611-621 (2022). https://doi.org/10.1016/j.ijbiomac.2022.02.166.
  6. L.L. Qin, G.Q. Su, C. Wu, Q. Zhou, X. Peng, L. Hu, Y. Liu, R. Wang, Q. Xu, Z.F. Fang, Y. Lin, S.Y. Xu, B. Feng, J. Li, D. Wu, L.Q. Che.: Effects of tremella fuciformis extract on growth performance, biochemical and immunological parameters of weaned piglets challenged with lipopolysaccharide, Anim. Prod. Sci. 62 (5) 462-469 (2022). https://doi.org/10.1071/AN20425.
  7. Y.L. Zhang, X.Y. Chen, P. Hu, Q.W. Liao, Y. Luo, J. Li, D.D. Feng, J. Zhang, Z.Q. Wu, H.X. Xu.: Extraction, purification, and antioxidant activity of exopolysaccharides produced by lactobacillus kimchi sr8 from sour meat in vitro and in vivo, Cyta-J. Food 19 (1) 228-237 (2021). https://doi.org/10.1080/19476337.2021.1883117.
  8. M.H. Jo, B. Kim, J.H. Ju, S.Y. Heo, K.H. Ahn, H.J. Lee, H.S. Yeom, H. Jang, M.S. Kim, C.H. Kim, B.R. Oh.: Tremella fuciformis tfcuv5 mycelial culture-derived exopolysaccharide production and its anti-aging effects on skin cells, Biotechnol. Bioproc. E. 26 (5) 738-748 (2021). https://doi.org/10.1007/s12257-020-0361-6.
  9. M. Yang, Z.L. Zhang, Y. He, C.L. Li, J.M. Wang, X. Ma.: Study on the structure characterization and moisturizing effect of tremella polysaccharide fermented from gcmcc5.39, Food Science And Human Wellness 10 (4) 471-479 (2021). https://doi.org/10.1016/j.fshw.2021.04.009.
  10. X.Q. Xu, A.J. Chen, X.Y. Ge, S. Li, T. Zhang, H. Xu.: Chain conformation and physicochemical properties of polysaccharide (glucuronoxylomannan) from fruit bodies of tremella fuciformis, Carbohyd. Polym. 245 (2020). https://doi.org/10.1016/j.carbpol.2020.116354.
  11. B. Niu, S.M. Feng, S.Q. Xuan, P. Shao.: Moisture and caking resistant tremella fuciformis polysaccharides microcapsules with hypoglycemic activity, Food Res. Int. 146 (2021). https://doi.org/10.1016/j.foodres.2021.110420.
  12. X. Ma, M. Yang, Y. He, C.T. Zhai, C.L. Li.: A review on the production, structure, bioactivities and applications of tremella polysaccharides, Int. J. Immunopath. Ph. 35 (2021). https://doi.org/10.1177/20587384211000541.
  13. Y. Ruan, H. Li, L.M. Pu, T. Shen, Z.N. Jin.: Tremella fuciformis polysaccharides attenuate oxidative stress and inflammation in macrophages through mir-155, Anal. Cell. Pathol. 2018 (2018). https://doi.org/10.1155/2018/5762371.
  14. Y. Wu, Z. Wei, F. Zhang, R.J. Linhardt, P. Sun, A. Zhang.: Structure, bioactivities and applications of the polysaccharides from tremella fuciformis mushroom: a review, Int. J. Biol. Macromol. 121 1005-1010 (2019). https://doi.org/https://doi.org/10.1016/j.ijbiomac.2018.10.117.
  15. J. Zhang, Y.K. Zhang, Y. Liu, J.H. Wang.: Emulsifying properties of tremella fuciformis: a novel promising food emulsifier, Int. J. Food Eng. 15 (3-4) (2019). https://doi.org/10.1515/ijfe-2018-0217.
  16. F. Yang, Q.L. Du, T. Miao, X. Zhang, W. Xu, D.Y. Jia.: Interaction between potato starch and tremella fuciformis polysaccharide, Food Hydrocolloid. 127 (2022). https://doi.org/10.1016/j.foodhyd.2022.107509.
  17. Y. Lin, Q.L. Xu, X.M. Li, P. Shao.: Tremella fuciformis polysaccharides as a fat substitute on the rheological, texture and sensory attributes of low-fat yogurt, CURRENT RESEARCH IN FOOD SCIENCE 5 1061-1070 (2022). https://doi.org/10.1016/j.crfs.2022.06.007.
  18. C.H. Chiu, K.C. Chiu, L.C. Yang.: Amelioration of obesity in mice fed a high-fat diet with uronic acid-rich polysaccharides derived from tremella fuciformis, Polymers-Basel 14 (8) (2022). https://doi.org/10.3390/polym14081514.
  19. Y.Y. Xu, L.Y. Xie, Z.Y. Zhang, W.W. Zhang, J. Tang, X.L. He, J. Zhou, W.H. Peng.: Tremella fuciformis polysaccharides inhibited colonic inflammation in dextran sulfate sodium-treated mice via foxp3+t cells, gut microbiota, and bacterial metabolites, Frontiers In Immunology 12 (2021). https://doi.org/10.3389/fimmu.2021.648162.
  20. H.Y. Xiao, H.L. Li, Y.F. Wen, D.X. Jiang, S.M. Zhu, X.L. He, Q.P. Xiong, J. Gao, S.Z. Hou, S. Huang, L. He, J. Liang.: Tremella fuciformis polysaccharides ameliorated ulcerative colitis via inhibiting inflammation and enhancing intestinal epithelial barrier function, Int. J. Biol. Macromol. 180 633-642 (2021). https://doi.org/10.1016/j.ijbiomac.2021.03.083.
  21. H. Fu, S.Q. You, D. Zhao, Q. An, J.C. Zhang, C.T. Wang, D.D. Wang, M. Li.: Tremella fuciformis polysaccharides inhibit uva-induced photodamage of human dermal fibroblast cells by activating up-regulating nrf2/keap1 pathways, J. Cosmet. Dermatol.-Us 20 (12) 4052-4059 (2021). https://doi.org/10.1111/jocd.14051.
  22. T.J. Khan, X.F. Xu, X.L. Xie, X.M. Dai, P.N. Sun, Q.D. Xie, X.L. Zhou.: Tremella fuciformis crude polysaccharides attenuates steatosis and suppresses inflammation in diet-induced nafld mice, Curr. Issues Mol. Biol. 44 (3) 1224-1234 (2022). https://doi.org/10.3390/cimb44030081.
  23. P. Li, Z. Jiang, T. Sun, C. Wang, Y. Chen, Z. Yang, B. Du, C. Liu.: Comparison of structural, antioxidant and immuno‐stimulating activities of polysaccharides from tremella fuciformis in two different regions of china, Int. J. Food Sci. Tech. 53 (8) 1942-1953 (2018). https://doi.org/10.1111/ijfs.13782.
  24. X.Y. Mei, W.J. Yang, G. Huang, H.L. Huang.: The antioxidant activities of balsam pear polysaccharide, Int. J. Biol. Macromol. 142 232-236 (2020). https://doi.org/10.1016/j.ijbiomac.2019.09.168.
  25. M.H. Fan, J.X. Zhu, Y.L. Qian, W. Yue, Y. Xu, D.D. Zhang, Y.Q. Yang, X.Y. Gao, H.Y. He, D.F. Wang.: Effect of purity of tea polysaccharides on its antioxidant and hypoglycemic activities, J. Food Biochem. 44 (8) (2020). https://doi.org/10.1111/jfbc.13277.
  26. T.W. Guan, X.Y. Wei, P. Xu, K.B. Chen, Y.F. Zou, M.S. Chen, Z.Y. Zhu.: Comparison of structural and antioxidant activity of polysaccharide extracted from truffles, J. Food Sci. 87 (7) 2999-3012 (2022). https://doi.org/10.1111/1750-3841.16207.
  27. X.L. Ji, C.Y. Hou, Y.Z. Yan, M.M. Shi, Y.Q. Liu.: Comparison of structural characterization and antioxidant activity of polysaccharides from jujube (ziziphus jujuba mill.) Fruit, Int. J. Biol. Macromol. 149 1008-1018 (2020). https://doi.org/10.1016/j.ijbiomac.2020.02.018.
  28. X. Pang, Y. Jing, P. Li, X. Qiu, Y. Zheng, Q. Wang, L. Wu.: Structural characterization and antioxidant activities of polysaccharides from angelica dahurica as extracted by optimized ultrasonic-assisted method, Int. J. Food Prop. 25 (1) 1635-1649 (2022). https://doi.org/10.1080/10942912.2022.2096066.
  29. J.X.C.Z. Zhu.: Comparison and structural characterization of polysaccharides from natural and artificial se-enriched green tea, Int. J. Biol. Macromol. 130 388-398 (2019). https://doi.org/10.1016/j.ijbiomac.2019.02.102.
  30. Y.B. Li, J.C. Chen, P.F. Lai, B.S. Tang, L. Wu.: Influence of drying methods on the physicochemical properties and nutritional composition of instant tremella fuciformis, Food Sci. Tech.-Brazil 40 (3) 741-748 (2020). https://doi.org/10.1590/fst.20519.
  31. Y.F. Gong, Y.X. Ma, E. Cheung, L.J. You, L. Liao, S. Pedisic, V. Kulikouskaya.: Structural characteristics and anti-inflammatory activity of uv/h2o2-treated algal sulfated polysaccharide from gracilaria lemaneiformis, Food Chem. Toxicol. 152 (2021). https://doi.org/10.1016/j.fct.2021.112157.
  32. Y. Liu, C.F. Hu, X. Feng, L. Cheng, S.A. Ibrahim, C.T. Wang, W. Huang.: Isolation, characterization and antioxidant of polysaccharides from stropharia rugosoannulata, Int. J. Biol. Macromol. 155 883-889 (2020). https://doi.org/10.1016/j.ijbiomac.2019.11.045.
  33. Z.P. Zhu, J. Chen, Y. Chen, Y.T. Ma, Q.S. Yang, Y.Q. Fan, C.M. Fu, B. Limsila, R. Li, W. Liao.: Extraction, structural characterization and antioxidant activity of turmeric polysaccharides, Lwt-Food Sci. Technol. 154 (2022). https://doi.org/10.1016/j.lwt.2021.112805.
  34. Y. Su, L. Li.: Structural characterization and antioxidant activity of polysaccharide from four auriculariales, Carbohyd. Polym. 229 (2020). https://doi.org/10.1016/j.carbpol.2019.115407.
  35. G. Teniente-Martinez, A. Bernardino-Nicanor, M.D. Valadez-Vega, J.L. Montanez-Soto, J. Juarez-Goiz, L. Gonzalez-Cruz.: In vitro study of the antihypertensive, antioxidant and antiproliferative activities of peptides obtained from two varieties of phaseolus coccineus l, Cyta-J. Food 20 (1) 102-110 (2022). https://doi.org/10.1080/19476337.2022.2090611.
  36. Y.F. Zou, M.S. Chen, Y.P. Fu, Z.K. Zhu, Y.Y. Zhang, B.S. Paulsen, F. Rise, Y.L. Chen, Y.Z. Yang, R.Y. Jia, L.X. Li, X. Song, H.Q. Tang, B. Feng, C. Lv, G. Ye, D.T. Wu, Z.Q. Yin, C. Huang.: Characterization of an antioxidant pectic polysaccharide from platycodon grandiflorus, Int. J. Biol. Macromol. 175 473-480 (2021). https://doi.org/10.1016/j.ijbiomac.2021.02.041.
  37. M.Y. Jia, J.J. Chen, X.Z. Liu, M.Y. Xie, S.P. Nie, Y. Chen, J.H. Xie, Q. Yu.: Structural characteristics and functional properties of soluble dietary fiber from defatted rice bran obtained through trichoderma viride fermentation, Food Hydrocolloid. 94 468-474 (2019). https://doi.org/10.1016/j.foodhyd.2019.03.047.
  38. C.Y. Zhao, F. Yang, F. Lin, Q.S. Qu, Z.X. Li, X. Liu, L. Han, X.Y. Shi.: Process optimization in ginseng fermentation by monascus ruber and study on bile acid-binding ability of fermentation products in vitro, Prep. Biochem. Biotech. 51 (2) 120-126 (2021). https://doi.org/10.1080/10826068.2020.1799389.
  39. X.C. Liu, Z.C. Zhang, L. Xu, W.J. Shi, Q.S. Sun, C.Y. Ding, H.H. Zheng, F. Wang.: Dioscorea saponin transforms the structure of truffle exo-polysaccharide and enhances its antioxidant activity, Lwt-Food Sci. Technol. 127 (2020). https://doi.org/10.1016/j.lwt.2020.109417.
  40. X.L. Ji, Y.Z. Yan, C.Y. Hou, M.M. Shi, Y.Q. Liu.: Structural characterization of a galacturonic acid-rich polysaccharide from ziziphus jujuba cv. Muzao, Int. J. Biol. Macromol. 147 844-852 (2020). https://doi.org/10.1016/j.ijbiomac.2019.09.244.
  41. C. Cui, J.H. Lu, D. Sun-Waterhouse, L.X. Mu, W.Z. Sun, M.M. Zhao, H.F. Zhao.: Polysaccharides from laminaria japonica: structural characteristics and antioxidant activity, Lwt-Food Sci. Technol. 73 602-608 (2016). https://doi.org/10.1016/j.lwt.2016.07.005.
  42. S.H. Chen, H.L. Huang, G.L. Huang.: Extraction, derivatization and antioxidant activity of cucumber polysaccharide, Int. J. Biol. Macromol. 140 1047-1053 (2019). https://doi.org/10.1016/j.ijbiomac.2019.08.203.
  43. H.L. Huang, F. Chen, R. Long, G.L. Huang.: The antioxidant activities in vivo of bitter gourd polysaccharide, Int. J. Biol. Macromol. 145 141-144 (2020). https://doi.org/10.1016/j.ijbiomac.2019.12.165.
  44. F. Chen, S.Y. Huang, G.L. Huang.: Preparation, activity, and antioxidant mechanism of rice bran polysaccharide, Food Funct. 12 (2) 834-839 (2021). https://doi.org/10.1039/d0fo02498h.
  45. X.Y. Lan, Y.H. Wang, S. Deng, J.Y. Zhao, L. Wang, K. Yao, D.Y. Jia.: Physicochemical and rheological properties of tremella fuciformis polysaccharide fractions by ethanol precipitation, Cyta-J. Food 19 (1) 645-655 (2021). https://doi.org/10.1080/19476337.2021.1950212.
  46. D.Q. Wang, D.G. Wang, T.X. Yan, W.F. Jiang, X.Y. Han, J.F. Yan, Y.R. Guo.: Nanostructures assembly and the property of polysaccharide extracted from tremella fuciformis fruiting body, Int. J. Biol. Macromol. 137 751-760 (2019). https://doi.org/10.1016/j.ijbiomac.2019.06.198.
  47. C.Z. Wu, M.M. Zhao, X.Y. Bu, Z.Y. Qing, L.B. Wang, Y.Q. Xu, Y. Yang, J.W. Bai.: Preparation, characterization, antioxidant and antiglycation activities of selenized polysaccharides from blackcurrant, Rsc Adv. 10 (54) 32616-32627 (2020). https://doi.org/10.1039/d0ra06462a.
  48. J.Y. Gu, H.H. Zhang, C.T. Wen, J.X. Zhang, Y.Q. He, H.L. Ma, Y.Q. Duan.: Purification, characterization, antioxidant and immunological activity of polysaccharide from sagittaria sagittifolia l., Food Res. Int. 136 (2020). https://doi.org/10.1016/j.foodres.2020.109345.
  49. S.Y. Zhou, G.L. Huang, G.Y. Chen.: Extraction, structural analysis, derivatization and antioxidant activity of polysaccharide from chinese yam, Food Chem. 361 (2021). https://doi.org/10.1016/j.foodchem.2021.130089.
  50. S.Y. Zhou, G.L. Huang.: Preparation, structure and activity of polysaccharide phosphate esters, Biomed. Pharmacother. 144 (2021). https://doi.org/10.1016/j.biopha.2021.112332.
  51. D.T. Wu, Y.X. Zhao, H. Guo, R.Y. Gan, L.X. Peng, G. Zhao, L. Zou.: Physicochemical and biological properties of polysaccharides from dictyophora indusiata prepared by different extraction techniques, Polymers-Basel 13 (14) (2021). https://doi.org/10.3390/polym13142357.
  52. Y.T. Wu, Y.F. Huo, L. Xu, Y.Y. Xu, X.L. Wang, T. Zhou.: Purification, characterization and antioxidant activity of polysaccharides from porphyra haitanensis, Int. J. Biol. Macromol. 165 2116-2125 (2020). https://doi.org/10.1016/j.ijbiomac.2020.10.053.
  53. S.Y. Zhou, G.L. Huang.: Extraction, derivatization, and antioxidant activity of morinda citrifolia polysaccharide, Chem. Biol. Drug Des. 99 (4) 603-608 (2022). https://doi.org/10.1111/cbdd.14023.
  54. H.B. Hu, H.P. Liang, H.M. Li, R.N. Yuan, J. Sun, L.L. Zhang, M.H. Han, Y. Wu.: Isolation, purification, characterization and antioxidant activity of polysaccharides from the stem barks of acanthopanax leucorrhizus, Carbohyd. Polym. 196 359-367 (2018). https://doi.org/10.1016/j.carbpol.2018.05.028.
  55. F. Wang, L.M. Kong, Y.Y. Xie, C. Wang, X.L. Wang, Y.B. Wang, L.L. Fu, T. Zhou.: Purification, structural characterization, and biological activities of degraded polysaccharides from porphyra yezoensis, J. Food Biochem. 45 (4) (2021). https://doi.org/10.1111/jfbc.13661.
  56. J.Y. Gu, H.H. Zhang, J.X. Zhang, C.T. Wen, H.L. Ma, Y.Q. Duan, Y.Q. He.: Preparation, characterization and bioactivity of polysaccharide fractions from sagittaria sagittifolia l., Carbohyd. Polym. 229 (2020). https://doi.org/10.1016/j.carbpol.2019.115355.
  57. Q.R. Meng, Z.H. Chen, F. Chen, Z.G. Zhang, W.J. Gao.: Optimization of ultrasonic-assisted extraction of polysaccharides from hemerocallis citrina and the antioxidant activity study, J. Food Sci. (2021). https://doi.org/10.1111/1750-3841.15806.
  58. X. Ge, W. Huang, X. Xu, P. Lei, D. Sun, H. Xu, S. Li.: Production, structure, and bioactivity of polysaccharide isolated from tremella fuciformis xy, Int. J. Biol. Macromol. 148 173-181 (2020). https://doi.org/10.1016/j.ijbiomac.2020.01.021.
  59. Z.S. Zhang, X.M. Wang, M.X. Zhao, H.M. Qi.: Free-radical degradation by fe2+/vc/h2o2 and antioxidant activity of polysaccharide from tremella fuciformis, Carbohyd. Polym. 112 578-582 (2014). https://doi.org/10.1016/j.carbpol.2014.06.030.
  60. M. Li, F.M. Ma, R. Li, G.Y. Ren, D.L. Yan, H.N. Zhang, X.Y. Zhu, R.N. Wu, J.R. Wu.: Degradation of tremella fuciformis polysaccharide by a combined ultrasound and hydrogen peroxide treatment: process parameters, structural characteristics, and antioxidant activities, Int. J. Biol. Macromol. 160 979-990 (2020). https://doi.org/10.1016/j.ijbiomac.2020.05.216.
  61. Q. Yuan, Y. He, P.Y. Xiang, Y.J. Huang, Z.W. Cao, S.W. Shen, L. Zhao, Q. Zhang, W. Qin, D.T. Wu.: Influences of different drying methods on the structural characteristics and multiple bioactivities of polysaccharides from okra (abelmoschus esculentus), Int. J. Biol. Macromol. 147 1053-1063 (2020). https://doi.org/10.1016/j.ijbiomac.2019.10.073.
  62. Y. Fu, Q. Yuan, S. Lin, W. Liu, G. Du, L. Zhao, Q. Zhang, D.R. Lin, Y.T. Liu, W. Qin, D.Q. Li, D.T. Wu.: Physicochemical characteristics and biological activities of polysaccharides from the leaves of different loquat (eriobotrya japonica) cultivars, Int. J. Biol. Macromol. 135 274-281 (2019). https://doi.org/10.1016/j.ijbiomac.2019.05.157.
  63. D.D. Guo, J.Y. Lei, L.J. Xu, Y.F. Cheng, C.P. Feng, J.L. Meng, M.C. Chang, X.R. Geng.: Two novel polysaccharides from clitocybe squamulosa: their isolation, structures, and bioactivities, FRONTIERS IN NUTRITION 9 (2022). https://doi.org/10.3389/fnut.2022.934769.
  64. Z.H. Xue, X.D. Gao, Y.N. Jia, Y.J. Wang, Y.P. Lu, M. Zhang, P. Panichayupakaranant, H.X. Chen.: Structure characterization of high molecular weight soluble dietary fiber from mushroom lentinula edodes (berk.) Pegler and its interaction mechanism with pancreatic lipase and bile salts, Int. J. Biol. Macromol. 153 1281-1290 (2020). https://doi.org/10.1016/j.ijbiomac.2019.10.263.
  65. C. Lopez-Pena, I.J. Arroyo-Maya, D.J. Mcclements.: Interaction of a bile salt (sodium taurocholate) with cationic (epsilon-polylysine) and anionic (pectin) biopolymers under simulated gastrointestinal conditions, Food Hydrocolloid. 87 352-359 (2019). https://doi.org/10.1016/j.foodhyd.2018.08.029.
  66. I. Silva, F. Machado, M.J. Moreno, C. Nunes, M.A. Coimbra, F. Coreta-Gomes.: Polysaccharide structures and their hypocholesterolemic potential, Molecules 26 (15) (2021). https://doi.org/10.3390/molecules26154559.
  67. G.P. Yu, C.H. Yue, X.D. Zang, C. Chen, L.W. Dong, Y.Q. Liu.: Purification, characterization and in vitro bile salt-binding capacity of polysaccharides from armillaria mellea mushroom, Czech J. Food Sci. 37 (1) 51-56 (2019). https://doi.org/10.17221/182/2018-CJFS.