Next Article in Journal
Fast Quantum Approach for Evaluating the Energy of Non-Covalent Interactions in Molecular Crystals: The Case Study of Intermolecular H-Bonds in Crystalline Peroxosolvates
Next Article in Special Issue
Production and Purification of Pectinase from Bacillus subtilis 15A-B92 and Its Biotechnological Applications
Previous Article in Journal
An In Silico Investigation to Explore Anti-Cancer Potential of Foeniculum vulgare Mill. Phytoconstituents for the Management of Human Breast Cancer
Previous Article in Special Issue
Bioactive Properties of a Novel Antibacterial Dye Obtained from Laccase-Mediated Oxidation of 8-Anilino-1-naphthalenesulfonic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enzymatic Synthesis of Novel and Highly Soluble Puerarin Glucoside by Deinococcus geothermalis Amylosucrase

1
Department of Cosmetic Science, Chia Nan University of Pharmacy and Science, No. 60 Erh-Jen Rd., Sec. 1, Jen-Te District, Tainan 71710, Taiwan
2
Biodiversity Research Center, Academia Sinica, Taipei 11529, Taiwan
3
Department of Food Science, National Quemoy University, Kinmen County 892, Taiwan
4
Department of Biological Sciences and Technology, National University of Tainan, Tainan 70005, Taiwan
*
Author to whom correspondence should be addressed.
Authors contributed equally to this manuscript.
Molecules 2022, 27(13), 4074; https://doi.org/10.3390/molecules27134074
Submission received: 13 May 2022 / Revised: 16 June 2022 / Accepted: 22 June 2022 / Published: 24 June 2022

Abstract

:
Puerarin (daidzein-8-C-glucoside) is an isoflavone isolated from several leguminous plants of the genus Pueraria. Puerarin possesses several pharmacological properties; however, the poor solubility of puerarin limits its applications. To resolve this poor solubility, Deinococcus geothermalis amylosucrase (DgAS) was used to modify puerarin into more soluble derivatives. The results showed that DgAS could biotransform puerarin into a novel compound: puerarin-4′-O-α-glucoside. The biotransformation reaction was manipulated at different temperatures, pH values, sucrose concentrations, reaction times, and enzyme concentrations. The results showed that the optimal reaction condition was biotransformed by 200 μg/mL DgAS with 20% (w/v) sucrose at pH 6 and incubated at 40 °C for 48 h, and the optimal production yield was 35.1%. Puerarin-4′-O-α-glucoside showed 129-fold higher solubility than that of puerarin and, thus, could be further applied for pharmacological use in the future.

1. Introduction

Puerarin (daidzein-8-C-glucoside) is an isoflavone isolated from several leguminous plants of the genus Pueraria. Puerarin possesses several pharmacological properties and was approved for clinical trials in diabetes mellitus (DM) by the Chinese Ministry of Health in 1993 [1]. Recently, puerarin has proved to be a potential anticancer drug [2]. However, its water solubility remains low although puerarin contains a glucosyl residue attached to the parent isoflavone [3]. Thus, its low solubility makes puerarin difficult to apply in food processing and results in poor absorption after oral administration. Only two dosage forms could be found on the market until now: puerarin injections and eye drops [4]. Therefore, improving the solubility of puerarin could strengthen its applications.
Molecular modifications could be achieved by either chemical or enzymatic approaches [5]. However, chemical modifications sometimes involve multiple steps of protection/deprotection to control regioselectivity. Multiple-step reactions usually reduce the final product yield. In contrast, one-step enzymatic modification could promote regio- and enantioselective catalytic activity. For example, glycosylation can significantly improve the solubility of molecules [6]. Glycosylation modulates the solubility, bioavailability, and chemical properties of many natural products, such as flavonoids and steroids [7]. Many novel bioactive flavonoid glycosides have been obtained by enzymatic glycosylation for developing new and potential drugs [8].
To develop highly soluble puerarin glycosides, two major forms—α-(1→6″)-puerarin and α-(1→4″)-puerarin—biotransformed with different enzymes have been previously reported. Li et al., (2004) used the Bacillus stearothermophilus maltogenic amylase (BsMA), a glycoside hydrolase family 13 (GH13) (EC 3.2.1.133) enzyme, to glycosylate puerarin with β-cyclodextrin (CD) and to produce glucosyl-α-(1→6″)-puerarin and maltosyl-α-(1→6″)-puerarin, which possessed 14-fold and 168-fold higher aqueous solubility than that of puerarin, respectively [9,10]. Similarly, Li et al., (2011) used the archaeon Thermofilum pendens MA (TfMA) to produce the same derivative [11]. Ko et al., (2012) used Leuconostoc lactis dextransucrase (GH70; EC 2.4.1.5) to produce glucosyl-α-(1→6″)-puerarin and maltosyl-α-(1→6″)-puerarin, which possessed 15-fold and 202-fold higher solubility than that of puerarin, respectively [12]. Huang et al., (2020) used another GH13 enzyme, Bacillus licheniformis cyclodextrin glucanotransferase (EC 2.4.1.19), to produce glucosyl-α-(1→4″)-puerarin, maltosyl-α-(1→4″)-puerarin, and maltotriosyl-α-(1→4″)-puerarin, which showed 15-fold, 100-fold, and 179-fold higher solubility than that of puerarin, respectively [13]. On the other hand, Wang et al., (2014) and Wu et al., (2013) used Arthrobacter nicotianae β-fructosidase (GH32; EC 3.2.1.153) to produce fructosyl-β-(2→6″)-puerarin and difructosyl-β-(2→6″)-puerarin [14,15], although the authors did not determine the solubility of the two puerarin fructosides. Recently, Nunez-Lopez et al., (2019, 2020) used Gluconacetobacter diazotrophicus levansucrase (GH68; EC 2.4.1.10) to produce the same fructosyl-β-(2→6″)-puerarin, which showed 23-fold higher solubility than that of puerarin [16,17].
Amylosucrase (AS, EC 2.4.1.4) is another GH13 versatile sucrose-hydrolyzing enzyme [18,19] which uses sucrose as the sole substrate to catalyze the α-1,4-glucans derivatives. AS has been proven to glycosylate many flavonoids, such as piceid [20], catechin [21], baicalein [22], isoquercetrin [23,24], rutin [25], phloretin [26], taxifolin, aseculetin, and luteolin [27]. For example, previous studies used Deinococcus geothermalis amylosucrase (DgAS) to glycosylate soyisoflavone 8-hydroxydaidzein (8-OHDe) and 8-OHDe-7-O-glucoside into highly soluble and stable 8-OHDe glucoside and diglucosides [28,29]. Because the chemical structure of puerarin is similar to that of 8-OHDe-7-O-glucoside, we expect that puerarin could be glycosylated by DgAS to produce more soluble puerarin analogs. Thus, DgAS with sucrose was used to modify puerarin, and the produced puerarin analog was purified, chemically identified, and characterized in terms of its solubility.

2. Results and Discussion

2.1. Biotransformation of Puerarin by DgAS

Puerarin was biotransformed by DgAS to modify puerarin. The biotransformation products were analyzed by HPLC. The results showed that DgAS could biotransform puerarin into a major product, compound (1) (Figure S1, see Supplementary Materials). Fifteen point six percent of puerarin was converted to compound (1) by DgAS under the original testing conditions.
To improve the efficiency of the biotransformation by DgAS, the reaction conditions were optimized with different sucrose concentrations, temperatures, pH values, and times. The results showed that the optimal reaction conditions were 20% (w/v) sucrose and pH 6 at 40 °C for 48 h (Figure 1). Under the optimal reaction conditions, the yield of compound (1) from the biotransformation of puerarin by DgAS was increased to 24.1%. To improve the production yield in advance, the DgAS enzyme concentration was increased from 25 μg/mL to 200 μg/mL in reactions. The results showed that the highest yield of compound (1) can reach 35.1% when DgAS was increased to 200 μg/mL (Figure 2).

2.2. Identification of the Biotransformation Product

To reveal the chemical structures of compound (1), the biotransformation was scaled up to 40 mL. The biotransformation product was purified by preparative HPLC. The chemical structure of the purified compound was then resolved using mass and nucleic magnetic resonance (NMR) spectral analyses. The molecular formula of compound (1) was established as C27H30O14 by the electrospray ionization mass spectrometry (ESI-MS) at m/z 577.3 [M-H], indicating the molecular weight of 578 and the presence of a glucoside to puerarin (Figure S2). The functional groups of compound (1) were analyzed by infrared (IR) spectroscopy. The results revealed significant absorption at 3273 cm1, representing the typical hydroxyl groups, and another at 1623.9 cm1, representing the typical carbonyl groups in compound (1) (Figure S3). The compound (1) characteristic 1H and 13C NMR sugar signals have been assigned to C-glucosyl and O-glucosyl moieties by 1D and 2D NMR experiments. The 1H spectrum of compound (1) in DMSO-d6 showed one singlet at 8.42 ppm; eight doublets at 4.81, 5.42, 6.99, 7.13, 7.13, 7.52, 7.52, and 7.94 ppm; and a complex 10-spin system between 3.0 and 5.0 ppm. An analysis of this second-order system revealed coupling constants typical of two glucose moieties. The compound (1) glycosidic linkage of the C-glucosyl moiety on puerarin C-8 has been revealed by the presence of heteronuclear multiple bond connectivity (HMBC) correlations between C-8/H-1″ (112.7/4.81 ppm) and anomeric proton H-1″ at 4.81 (d, J = 9.1 Hz), indicating a C-β-configuration of puerarin, supported by data from previous literature [30]. The puerarin O-glucosyl moiety was a doublet signal at H-1‴ (5.42 ppm, d, J = 3.5 Hz), with the corresponding carbon atom at C-1‴ (98.0 ppm) assigned to the anomeric proton, and indicated an O-α-configuration by heteronuclear single quantum coherence (HSQC), which is in the O-α-configuration, and the H-1‴ (δ = 5.42 ppm) of puerarin, as well as the HMBC cross signals H-1‴/C-4′ (5.42/156.9 ppm). The significant downfield shift in the 13C signal of C-4′ indicated the connection of the second glucosyl moiety. The NMR signals were fully identified, as shown in Table S1. Compound (1) was thus confirmed to be puerarin-4′-O-α-glucoside (Table S1 and Figures S4–S10). Figure 3 illustrates the biotransformation process of puerarin by DgAS.
Previously studied GH enzymes catalyzed two major glycosylations—α-(1→6″)-puerarin and α-(1→4″)-puerarin—on the C-glucoside residue of puerarin [9,10,11,12,13,14,15,16,17]. In contrast, DgAS preferred catalyzing glycosylation on the 4′-hydroxyl group of puerarin and produced a novel derivative. Our previous study also showed that DgAS catalyzed glycosylation on the 4′-hydroxyl group of 8-OHDe-7-O-glucoside and produced 8-OHDe-7,4′-O-diglucoside [29]. Both studies suggest that DgAS is a good enzyme for producing isoflavones-4′-glucoside.

2.3. Aqueous Solubility of Puerarin and Its Derivatives

The solubility of puerarin and its derivatives was determined by HPLC analysis. The results showed that puerarin-4′-O-α-glucoside had a solubility 129 folds higher than that of puerarin (Table 1). The well soluble puerarin-4′-O-α-glucoside could present an alternative application to the industry.
It is known that the more sugars are attached, the higher the solubility of the modified molecules. Table 2 summarizes various puerarin glycosides produced by different GH enzymes. All reported puerarin-monoglycosides linked to the C8-glucoside of puerarin possessed glycosyl attached to one site (C8). Most puerarin-monoglycosides possessed 14-fold to 23-fold higher aqueous solubility than that of puerarin. Furthermore, puerarin-diglycosides or puerarin-triglycosides possessed over 100-fold higher aqueous solubility than that of puerarin. On the other hand, the solubility of puerarin-4′-O-α-glucoside (puerarin-monoglycoside) was 129 folds higher than that of puerarin. A possible reason for such higher aqueous solubility might be the mono-sugar linked to the two glucosyl sites (C8 and C4′) of puerarin-4′-O-α-glucoside. Higher solubility of a similar linkage was also found in the 8-OHDe-7,4′-O-α-diglucoside, which was linked to the two glucosyl sites (C7 and C4′); it was seven folds higher than that of daidzin-4″-O-α-glucoside [29]. This study revealed that glycosylation at the 4′-hydroxyl group on the isoflavone skeleton would yield better solubility than glycosylation at the hydroxyl groups on the C-glucoside of puerarin. This study highlighted a novel and unique action of DgAS on the 4′-glycosylation of isoflavones.
Previous studies revealed that O-glycosylated flavonoids (vitexin, isovitexin, and isoorientin) could be deglycosylated to recover the bioactivities of the parental flavonoids via in vitro fecal fermentation [31] or in vivo intestinal microbes [32]. In contrast, it has been reported that C-glucosides (e.g., puerarin) are more resistant to acidic and enzymatic hydrolysis [31]. Thus, the novel puerarin glucoside (O-glucoside) might also be in vivo digested to the parental puerarin as previous studies has shown [31,32]. Nevertheless, we expect that the high solubility of puerarin glucosides is an advantage for clinical therapy, in which the puerarin glucosides could be absorbed more easily and deglycosylated into the human body with a wider range of pharmaceutical dosages. Further clinical experiments are needed to confirm the bioactivities in the future.

3. Materials and Methods

3.1. Microorganism and Chemicals

Puerarin was purchased from Baoji Herbest Bio-Tech (Xi-An, Shaanxi, China). Recombinant DgAS was obtained from our previous studies [28,29]. One unit of DgAS activity was defined as the amount of the enzyme that hydrolyzed sucrose into 1 μmol of fructose per minute. The specific sucrose hydrolysis activity of the purified recombinant DgAS was determined to be 6.6 U/mg. The other reagents and solvents used were commercially available.

3.2. Biotransformation Using DgAS

The reaction mixture (0.1 mL) comprised 25 μg/mL of DgAS, 1 mg/mL of the tested substrate compound (diluted from a stock of 20 mg/mL in DMSO), 50% (w/v) sucrose, and 50 mM of phosphate buffer (PB) at pH 7 and was incubated at 40 °C for 24 h. The reaction was stopped by adding an equal volume of methanol and was analyzed using high-performance liquid chromatography (HPLC). To optimize the reaction conditions, different temperatures, pH values, sucrose concentrations, and reaction times were used. The buffers used were 50 mM of acetate buffer (pH 5), PB (pH 6 and 7), and Tris buffer (pH 8).

3.3. HPLC Analysis

HPLC was performed with the Agilent 1100 series HPLC system (Santa Clara, CA, USA) equipped with a gradient pump (Waters 600, Waters, Milford, MA, USA). The stationary phase was a C18 column (5 μm, 4.6 i.d. × 250 mm; Sharpsil H-C18, Sharpsil, Beijing, China), and the mobile phase was 1% acetic acid in water (A) and methanol (B). The elution condition was a linear gradient from 0 min with 40% B to 20 min with 70% B, an isocratic elution from 20 min to 25 min with 70% B, a linear gradient from 25 min with 70% B to 28 min with 40% B, and an isocratic elution from 28 min to 35 min with 40% B. All eluants were eluted at a flow rate of 1 mL/min. The sample volume was 10 μL. The detection condition was set at 254 nm.

3.4. Purification and Identification of the Biotransformation Metabolite

The purification process followed a previously described method [33]. To purify compound (1), the biotransformation reaction was scaled up to 40 mL (1 mL per tube), and the 40-vial reactions were incubated with 180 rpm of shaking at 40 °C for 24 h. After the reaction, compound (1) was purified by a preparative YoungLin HPLC system. The fraction with the compound (1) was collected, condensed under a vacuum, and then dehydrated by freeze drying. In total, five batches of the 40 mL reactions were purified for 60.1 mg of compound (1). The production yield was 60.1 mg/277.8 mg = 21.6%, which is the purified product (60.1 mg) divided by that theoretical value of 100% conversion rate [(40 mL × 1 mg/mL puerarin)/(416 molecular weight of puerarin) × [578 molecular weight of compound (1)] × 5 batches] = 277.8 mg]. The structures of the compound were confirmed with NMR and mass spectral analyses. Mass analyses were performed using the Finnigan LCQ Duo mass spectrometer (ThermoQuest Corp., San Jose, CA, USA) with electrospray ionization (ESI). 1H- and 13C-NMR, distortionless enhancement by polarization transfer (DEPT), heteronuclear single quantum coherence (HSQC), heteronuclear multiple bond connectivity (HMBC), correlation spectroscopy (COSY), and nuclear Overhauser effect spectroscopy (NOESY) spectra were recorded on a Bruker AV-700 NMR spectrometer at ambient temperature. Standard pulse sequences and parameters were used for the NMR experiments, and all chemical shifts were reported in parts per million (ppm, δ).
Puerarin-4′-O-α-glucoside (1): light gray powder; ESI/MS m/z: 577.3 [M-H], 457.3, 439.1, 429.3, 414.1, 309.1, 295.2, 294.2, 266.0; IR (HBr): νmax = 3273 (OH), 1623.9 (C = O), 1507.4 (C – C), 1010.2 (C – O) cm−1; 1H-NMR (DMSO-d6, 700 MHz) Hδ: 3.13 (2H, m, H-4″, 4‴), 3.24 (1H, m, H-5″), 3.26 (1H, t, J = 9.1 Hz, H-3″), 3.35 (1H, m, H-2‴), 3.43 (1H, m, H-6″a), 3.47 (1H, m, H-6‴a), 3.48 (1H, m, H-5‴), 3.64 (1H, m, H-3‴), 3.57 (1H, m, H-6‴b), 3.72 (1H, m, H-6″b), 4.03 (1H, t, J = 9.1 Hz, H-2″), 4.81 (1H, d, J = 9.1 Hz, H-1″), 5.42 (1H, J = 3.5 Hz, H-1‴), 6.99 (1H, d, J = 8.8 Hz, H-6), 7.13 (2H, d, J = 8.8 Hz, H-3′, 5′), 7.52 (2H, d, J = 8.8 Hz, H-2′, 6′), 7.94 1H, d, J = 8.8 Hz, H-5), 8.42 (1H, s, H-2). 13C-NMR (DMSO-d6, 175 MHz) Cδ: 60.7 (C-6‴), 61.4 (C-6″), 69.9 (C-4″, C-4‴), 70.7 (C-2″), 71.6 (C-2‴), 73.1 (C-3‴), 73.4 (C-1″), 73.8 (C-5‴), 78.8 (C-3″), 81,9 (C-5″), 98.0 (C-1‴), 112.7 (C-8), 116.6 (C-4a, 6, 3′, 5′), 122.7 (C-3), 125.6 (C-1′), 126.6 (C-5), 129.9 (C-2′, 6′), 153.1 (C-2), 156.9 (C-8a, 4′), 161.4 (C-7), 174.8 (C-4).

3.5. Determination of Solubility

The aqueous solubility of puerarin and its glucoside were determined according to a previous method with slight modification [34]. Twenty milligrams of the tested compound were resuspended in 50 μL of double-deionized H2O with 180 rpm of shaking at 25 °C for 1 h. The mixture was then centrifuged at 10,000× g at 25 °C for 30 min. The supernatant was filtrated with 0.2 μm of nylon membrane for HPLC analysis. For HPLC analysis, 10 μL of the puerarin filtrate or 2 μL of the puerarin-4′-O-α-glucoside filtrate were 100-fold or 500-fold diluted with 50% methanol, respectively. The concentrations of the tested compounds were determined with calibration curves of authentic samples. The final concentrations of the standard solutions were prepared with 4, 8, 12, 16, or 20 mg/L of puerarin (from 20 mg/mL of stock in DMSO) or with 100, 200, 300, or 400 mg/L of puerarin-4′-O-α-glucoside (from 20 mg/mL stock in DMSO) in 50% methanol.

4. Conclusions

Puerarin possesses important pharmacology activities for clinical usage. However, the poor solubility property of puerarin limits its dosage. In the present study, puerarin could be glycosylated by DgAS to produce a new compound, puerarin-4′-O-α-glucoside. Puerarin-4′-O-α-glucoside showed 129-fold higher solubility than that of puerarin. Such higher solubility could be further applied for pharmacological therapy in the future.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules27134074/s1. Table S1: 1H and 13C NMR assignments in DMSO-d6 at 700 and 175 MHz for compounds (1); Figure S1: High-performance liquid chromatography (HPLC) analysis of the biotransformation products of puerarin using Deinococcus geothermalis amylosucrase (DgAS); Figure S2: The mass-mass analysis of puerarin-4′-O-α-glucoside (1) at the negative mode; Figure S3: The infrared (IR) analysis of puerarin-4′-O-α-glucoside (1); Figure S4: 1D NMR spectrum (1H-NMR, 700 MHz, DMSO-d6) of the puerarin-4′-O-α-glucoside (1); Figure S5: 1D NMR spectrum (13C-NMR, 175 MHz, DMSO- d6) of the puerarin-4′-O-α-glucoside (1); Figure S6: 1D NMR spectrum (DEPT-135, 175 MHz, DMSO-d6) of the puerarin-4′-O-α-glucoside (1); Figure S7: 2D NMR spectrum (1H-13C HSQC, 700 MHz, DMSO-d6) of the puerarin-4′-O-α-glucoside (1); Figure S8: 2D NMR spectrum (1H-13C HMBC, 700 MHz, DMSO-d6) of the puerarin-4′-O-α-glucoside (1); Figure S9: 2D NMR spectrum (1H-1H COSY, 700 MHz, DMSO-d6) of the puerarin-4′-O-α-glucoside (1); Figure S10: 2D NMR spectrum (1H-1H NOESY, 700 MHz, DMSO-d6) of the puerarin-4′-O-α-glucoside (1).

Author Contributions

Conceptualization: T.-S.C.; data curation and methodology: Y.-L.T., T.-S.C. and H.-Y.D.; project administration: T.-S.C. and J.-Y.W.; writing—original draft, review, and editing: T.-S.C., T.-Y.W., J.-Y.W. and H.-Y.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology of Taiwan under grant number MOST 110-2221-E-024-002 to T.-S.C. and grant number MOST 110-2221-E-507-002 to J.-Y.W.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of compound (1) are available from the authors.

References

  1. Xie, B.; Wang, Q.; Zhou, C.; Wu, J.; Xu, D. Efficacy and Safety of the Injection of the Traditional Chinese Medicine Puerarin for the Treatment of Diabetic Peripheral Neuropathy: A Systematic Review and Meta-Analysis of 53 Randomized Controlled Trials. Evid.-Based Complementary Altern. Med. 2018, 2018, 2834650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Ahmad, B.; Khan, S.; Liu, Y.; Xue, M.; Nabi, G.; Kumar, S.; Alshwmi, M.; Qarluq, A.W. Molecular Mechanisms of Anticancer Activities of Puerarin. Cancer Manag. Res. 2020, 12, 79–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Zhang, L. Pharmacokinetics and drug delivery systems for puerarin, a bioactive flavone from traditional Chinese medicine. Drug Deliv. 2019, 26, 860–869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Chen, X.; Yu, J.; Shi, J. Management of diabetes mellitus with puerarin, a natural isoflavone from Pueraria lobata. Am. J. Chin. Med. 2018, 46, 1771–1789. [Google Scholar] [CrossRef]
  5. Kren, V.; Martínková, L. Glycosides in medicine: “The role of glycosidic residue in biological activity”. Curr. Med. Chem. 2001, 8, 1303–1328. [Google Scholar] [CrossRef]
  6. Desmet, T.; Soetaert, W.; Bojarova, P.; Kren, V.; Dijkhuizen, L.; Eastwick-Field, V.; Schiller, A. Enzymatic glycosylation of small molecules: Challenging substrates require tailored catalysts. Chemistry 2012, 18, 10786–10801. [Google Scholar] [CrossRef]
  7. Mestrom, L.; Przypis, M.; Kowalczykiewicz, D.; Pollender, A.; Kumpf, A.; Marsden, S.R.; Bento, I.; Jarzebski, A.B.; Szymanska, K.; Chrusciel, A.; et al. Leloir Glycosyltransferases in Applied Biocatalysis: A Multidisciplinary Approach. Int. J. Mol. Sci. 2019, 20, 5263. [Google Scholar] [CrossRef] [Green Version]
  8. Overwin, H.; Wray, V.; Seeger, M.; Sepulveda-Boza, S.; Hofer, B. Flavanone and isoflavone glucosylation by non-Leloir glycosyltransferases. J. Biotechnol. 2016, 233, 121–128. [Google Scholar] [CrossRef]
  9. Li, D.; Park, S.H.; Shim, J.H.; Lee, H.S.; Tang, S.Y.; Park, C.S.; Park, K.H. In vitro enzymatic modification of puerarin to puerarin glycosides by maltogenic amylase. Carbohydr. Res. 2004, 339, 2789–2797. [Google Scholar] [CrossRef]
  10. Choi, C.H.; Kim, S.H.; Jang, J.H.; Park, J.T.; Shim, J.H.; Kim, Y.W.; Park, K.H. Enzymatic synthesis of glycosylated puerarin using maltogenic amylase from Bacillus stearothermophilus expressed in Bacillus subtilis. J. Sci. Food Agric. 2010, 90, 1179–1184. [Google Scholar] [CrossRef]
  11. Li, X.; Li, D.; Park, S.H.; Gao, C.; Park, K.H.; Gu, L. Identification and antioxidative properties of transglycosylated puerarins synthesised by an archaeal maltogenic amylase. Food Chem. 2011, 124, 603–608. [Google Scholar] [CrossRef]
  12. Ko, J.A.; Ryu, Y.B.; Park, T.S.; Jeong, H.J.; Kim, J.H.; Park, S.J.; Kim, J.S.; Kim, D.; Kim, Y.M.; Lee, W.S. Enzymatic synthesis of puerarin glucosides using Leuconostoc dextransucrase. J. Microbiol. Biotechnol. 2012, 22, 1224–1229. [Google Scholar] [CrossRef] [Green Version]
  13. Huang, W.; He, Q.; Zhou, Z.R.; He, H.B.; Jiang, R.W. Enzymatic Synthesis of Puerarin Glucosides Using Cyclodextrin Glucanotransferase with Enhanced Antiosteoporosis Activity. ACS Omega 2020, 5, 12251–12258. [Google Scholar] [CrossRef]
  14. Wang, S.; Liu, G.; Zhang, W.; Cai, N.; Cheng, C.; Ji, Y.; Sun, L.; Zhan, J.; Yuan, S. Efficient glycosylation of puerarin by an organic solvent-tolerant strain of Lysinibacillus fusiformis. Enzym. Microb. Technol. 2014, 57, 42–47. [Google Scholar] [CrossRef]
  15. Wu, X.; Chu, J.; Wu, B.; Zhang, S.; He, B. An efficient novel glycosylation of flavonoid by beta-fructosidase resistant to hydrophilic organic solvents. Bioresour. Technol. 2013, 129, 659–662. [Google Scholar] [CrossRef]
  16. Nunez-Lopez, G.; Herrera-Gonzalez, A.; Hernandez, L.; Amaya-Delgado, L.; Sandoval, G.; Gschaedler, A.; Arrizon, J.; Remaud-Simeon, M.; Morel, S. Fructosylation of phenolic compounds by levansucrase from Gluconacetobacter diazotrophicus. Enzym. Microb. Technol. 2019, 122, 19–25. [Google Scholar] [CrossRef]
  17. Nunez-Lopez, G.; Morel, S.; Hernandez, L.; Musacchio, A.; Amaya-Delgado, L.; Gschaedler, A.; Remaud-Simeon, M.; Arrizon, J. One-pot bi-enzymatic cascade synthesis of puerarin polyfructosides. Carbohydr. Polym. 2020, 247, 116710. [Google Scholar] [CrossRef]
  18. Tian, Y.; Xu, W.; Zhang, W.; Zhang, T.; Guang, C.; Mu, W. Amylosucrase as a transglucosylation tool: From molecular features to bioengineering applications. Biotechnol. Adv. 2018, 36, 1540–1552. [Google Scholar] [CrossRef]
  19. Seo, D.H.; Yoo, S.H.; Choi, S.J.; Kim, Y.R.; Park, C.S. Versatile biotechnological applications of amylosucrase, a novel glucosyltransferase. Food Sci. Biotechnol. 2020, 29, 1–16. [Google Scholar] [CrossRef]
  20. Hyunsu, P.; Jieun, K.; Ji-Hae, P.; Nam-In, B.; Cheon-Seok, P.; Hee-Seob, L. Bioconversion of Piceid to Piceid Glucoside Using Amylosucrase from Alteromonas macleodii Deep Ecotype. J. Microbiol. Biotechnol. 2012, 22, 1698–1704. [Google Scholar]
  21. Cho, H.K.; Kim, H.H.; Seo, D.H.; Jung, J.H.; Park, J.H.; Baek, N.I.; Kim, M.J.; Yoo, S.H.; Cha, J.; Kim, Y.R.; et al. Biosynthesis of (+)-catechin glycosides using recombinant amylosucrase from Deinococcus geothermalis DSM 11300. Enzym. Microb. Technol. 2011, 49, 246–253. [Google Scholar] [CrossRef]
  22. Kim, K.H.; Park, Y.D.; Park, H.; Moon, K.O.; Ha, K.T.; Baek, N.I.; Park, C.S.; Joo, M.; Cha, J. Synthesis and biological evaluation of a novel baicalein glycoside as an anti-inflammatory agent. Eur. J. Pharmacol. 2014, 744, 147–156. [Google Scholar] [CrossRef]
  23. Rha, C.S.; Choi, J.M.; Jung, Y.S.; Kim, E.R.; Ko, M.J.; Seo, D.H.; Kim, D.O.; Park, C.S. High-efficiency enzymatic production of α-isoquercitrin glucosides by amylosucrase from Deinococcus geothermalis. Enzym. Microb. Technol. 2019, 120, 84–90. [Google Scholar] [CrossRef]
  24. Rha, C.S.; Kim, H.G.; Baek, N.I.; Kim, D.O.; Park, C.S. Using Amylosucrase for the Controlled Synthesis of Novel Isoquercitrin Glycosides with Different Glycosidic Linkages. J. Agric. Food Chem. 2020, 68, 13798–13805. [Google Scholar] [CrossRef]
  25. Kim, M.D.; Jung, D.H.; Seo, D.H.; Jung, J.H.; Seo, E.J.; Baek, N.I.; Yoo, S.H.; Park, C.S. Acceptor Specificity of Amylosucrase from Deinococcus radiopugnans and Its Application for Synthesis of Rutin Derivatives. J. Microbiol. Biotechnol. 2016, 26, 1845–1854. [Google Scholar] [CrossRef] [Green Version]
  26. Overwin, H.; Wray, V.; Hofer, B. Biotransformation of phloretin by amylosucrase yields three novel dihydrochalcone glucosides. J. Biotechnol. 2015, 211, 103–106. [Google Scholar] [CrossRef]
  27. Jang, S.W.; Cho, C.H.; Jung, Y.S.; Rha, C.; Nam, T.G.; Kim, D.O.; Lee, Y.G.; Baek, N.I.; Park, C.S.; Lee, B.H.; et al. Enzymatic synthesis of alpha-flavone glucoside via regioselective transglucosylation by amylosucrase from Deinococcus geothermalis. PLoS ONE 2018, 13, e0207466. [Google Scholar] [CrossRef]
  28. Chang, T.S.; Wang, T.Y.; Yann, S.Y.; Kao, Y.H.; Wu, J.Y.; Chiang, C.M. Potential Industrial Production of a Well-Soluble, Alkaline-Stable, and Anti-Inflammatory Isoflavone Glucoside from 8-Hydroxydaidzein Glucosylated by Recombinant Amylosucrase of Deinococcus geothermalis. Molecules 2019, 24, 2236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Chiang, C.M.; Wang, T.Y.; Wu, J.Y.; Zhang, Y.R.; Lin, S.Y.; Chang, T.S. Production of New Isoflavone Diglucosides from Glycosylation of 8-Hydroxydaidzein by Deinococcus geothermalis Amylosucrase. Fermentation 2021, 7, 232. [Google Scholar] [CrossRef]
  30. Prasain, J.K.; Jones, K.; Brissie, N.; Moore, R.; Wyss, J.M.; Barnes, S. Identification of Puerarin and Its Metabolites in Rats by Liquid Chromatography−Tandem Mass Spectrometry. J. Agric. Food Chem. 2004, 52, 3708–3712. [Google Scholar] [PubMed]
  31. Hein, E.M.; Rose, K.; Slot, G.V.; Friedrich, A.W.; Humpf, H.U. Deconjugation and degradation of flavonol glycosides by pig cecal microbiota characterized by fluorescence in situ hybridization (FISH). J. Agric. Food Chem. 2008, 56, 2281–2290. [Google Scholar] [CrossRef]
  32. Jang, D.; Jung, Y.S.; Seong, H.; Kim, M.S.; Rha, C.S.; Nam, T.G.; Han, N.S.; Kim, D.O. Stability of enzyme-modified flavonoid C- and O-glycosides from common buckwheat sprout extracts during in vitro digestion and colonic fermentation. J. Agric. Food Chem. 2021, 69, 5764–5773. [Google Scholar] [CrossRef]
  33. Wu, J.Y.; Ding, H.Y.; Wang, T.Y.; Tsai, Y.L.; Ting, H.J.; Chang, T.S. Improving Aqueous Solubility of Natural Antioxidant Mangiferin through Glycosylation by Maltogenic Amylase from Parageobacillus galactosidasius DSM 18751. Antioxidants 2021, 10, 1817. [Google Scholar] [CrossRef]
  34. Rha, C.S.; Kim, E.R.; Kim, Y.J.; Jung, Y.S.; Kim, D.O.; Park, C.S. Simple and efficient production of highly soluble daidzin glycosides by amylosucrase from Deinococcus geotherFmalis. J. Agri. Food Chem. 2019, 67, 12824–12832. [Google Scholar] [CrossRef]
Figure 1. Effects of sucrose concentration (a), pH (b), temperature (c), and time (d) on the production of compound (1) from biotransformation of puerarin by DgAS. The standard reaction condition was 1 mg/mL of puerarin, 25 μg/mL of DgAS, 50% (w/v) sucrose, and 50 mM of PB (pH 7) at 40 °C for 24 h. To determine suitable reaction conditions, different sucrose concentrations (w/v), pH values, temperatures, and reaction times were tested. After the reaction, the reaction product was analyzed by HPLC, and the yield of compound (1) was calculated by dividing the HPLC area of compound (1) by that of the sum of compound (1) and the residual puerarin in the HPLC analysis. The detailed reaction conditions and the HPLC procedure are described in the Materials and Methods section.
Figure 1. Effects of sucrose concentration (a), pH (b), temperature (c), and time (d) on the production of compound (1) from biotransformation of puerarin by DgAS. The standard reaction condition was 1 mg/mL of puerarin, 25 μg/mL of DgAS, 50% (w/v) sucrose, and 50 mM of PB (pH 7) at 40 °C for 24 h. To determine suitable reaction conditions, different sucrose concentrations (w/v), pH values, temperatures, and reaction times were tested. After the reaction, the reaction product was analyzed by HPLC, and the yield of compound (1) was calculated by dividing the HPLC area of compound (1) by that of the sum of compound (1) and the residual puerarin in the HPLC analysis. The detailed reaction conditions and the HPLC procedure are described in the Materials and Methods section.
Molecules 27 04074 g001
Figure 2. Effects of DgAS concentration on the yields of compound (1) biotransformed from puerarin by DgAS. The reaction condition was 1 mg/mL of puerarin, 25 to 200 μg/mL of DgAS, 20% (w/v) sucrose, and 50 mM of PB (pH 6) at 40 °C for 48 h. After the reaction, the product was analyzed by HPLC, and the yield of compound (1) was estimated by dividing the HPLC area of compound (1) by that of the sum of compound (1) and the residual puerarin. The detailed reaction conditions and the HPLC procedure are described in Section 3.3.
Figure 2. Effects of DgAS concentration on the yields of compound (1) biotransformed from puerarin by DgAS. The reaction condition was 1 mg/mL of puerarin, 25 to 200 μg/mL of DgAS, 20% (w/v) sucrose, and 50 mM of PB (pH 6) at 40 °C for 48 h. After the reaction, the product was analyzed by HPLC, and the yield of compound (1) was estimated by dividing the HPLC area of compound (1) by that of the sum of compound (1) and the residual puerarin. The detailed reaction conditions and the HPLC procedure are described in Section 3.3.
Molecules 27 04074 g002
Figure 3. The biotransformation process of puerarin by DgAS.
Figure 3. The biotransformation process of puerarin by DgAS.
Molecules 27 04074 g003
Table 1. Aqueous solubility of puerarin and its glucoside.
Table 1. Aqueous solubility of puerarin and its glucoside.
CompoundAqueous Solubility (mg/L)Fold 1
Puerarin2.02 × 103 ± 3.37 × 1021.0
Puerarin-4′-O-α-glucoside (1)2.60 × 105 ± 2.86 × 103128.7
1 The fold of aqueous solubility of puerarin glucoside derivatives is expressed relative to that of puerarin, normalized to 1.
Table 2. Aqueous solubility of puerarin glycosides reported in the literature.
Table 2. Aqueous solubility of puerarin glycosides reported in the literature.
Puerarin GlycosideCatalyzed EnzymesSugar DonorAdded SugarRelative Solubility 1Reference
Puerarin--01[9,12,13,16], this study
Puerarin-4′-α-glucosideAmylosucrase (DgAS)Sucrose1129This study
Glucosyl-α-(1→6′′)-puerarinMaltogenic amylase (BsMA);
Dextransucrase (LlDexT)
Maltotriose114–15[9,12]
Maltosyl-α-(1→6′′)-puerarinSucrose2168–202[9,12]
Glucosyl-α-(1→4′′)-puerarinGlucanotransferase (BlCGT)
Glucanotransferase (BlCGT)
Glucanotransferase (BlCGT)
Cyclodextrin115[13]
Maltosyl-α-(1→4′′)-puerarinCyclodextrin2100[13]
Maltotriosyl-α-(1→4′′)-puerarinCyclodextrin3179[13]
Fructosyl-β-(2→6′′)-puerarinLevansucrase (LsdA)Sucrose123[16]
1 Relative solubility was expressed relative to the solubility of puerarin normalized to 1.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ding, H.-Y.; Wang, T.-Y.; Wu, J.-Y.; Tsai, Y.-L.; Chang, T.-S. Enzymatic Synthesis of Novel and Highly Soluble Puerarin Glucoside by Deinococcus geothermalis Amylosucrase. Molecules 2022, 27, 4074. https://doi.org/10.3390/molecules27134074

AMA Style

Ding H-Y, Wang T-Y, Wu J-Y, Tsai Y-L, Chang T-S. Enzymatic Synthesis of Novel and Highly Soluble Puerarin Glucoside by Deinococcus geothermalis Amylosucrase. Molecules. 2022; 27(13):4074. https://doi.org/10.3390/molecules27134074

Chicago/Turabian Style

Ding, Hsiou-Yu, Tzi-Yuan Wang, Jiumn-Yih Wu, Yu-Li Tsai, and Te-Sheng Chang. 2022. "Enzymatic Synthesis of Novel and Highly Soluble Puerarin Glucoside by Deinococcus geothermalis Amylosucrase" Molecules 27, no. 13: 4074. https://doi.org/10.3390/molecules27134074

Article Metrics

Back to TopTop