Next Article in Journal
Evaluation of the Effect of Photodynamic Therapy on CAM-Grown Sarcomas
Previous Article in Journal
Assessment of the Remineralizing Potential of Biomimetic Materials on Early Artificial Caries Lesions after 28 Days: An In Vitro Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Production of a Novel Protopanaxatriol-Type Ginsenoside by Yeast Cell Factories

State Key Laboratory of Bioactive Substance and Function of Natural Medicines, NHC Key Laboratory of Biosynthesis of Natural Products, CAMS Key Laboratory of Enzyme and Biocatalysis of Natural Drugs, Institute of Materia Medica, Chinese Academy of Medical Sciences, Peking Union Medical College, Beijing 100050, China
*
Authors to whom correspondence should be addressed.
Bioengineering 2023, 10(4), 463; https://doi.org/10.3390/bioengineering10040463
Submission received: 9 March 2023 / Revised: 24 March 2023 / Accepted: 1 April 2023 / Published: 11 April 2023
(This article belongs to the Section Biochemical Engineering)

Abstract

:
Ginsenosides, the main active compounds in Panax species, are glycosides of protopanaxadiol (PPD) or protopanaxatriol (PPT). PPT-type ginsenosides have unique pharmacological activities on the central nervous system and cardiovascular system. As an unnatural ginsenoside, 3,12-Di-O-β-D-glucopyranosyl-dammar-24-ene-3β,6α,12β,20S-tetraol (3β,12β-Di-O-Glc-PPT) can be synthesized through enzymatic reactions but is limited by the expensive substrates and low catalytic efficiency. In the present study, we successfully produced 3β,12β-Di-O-Glc-PPT in Saccharomyces cerevisiae with a titer of 7.0 mg/L by expressing protopanaxatriol synthase (PPTS) from Panax ginseng and UGT109A1 from Bacillus subtilis in PPD-producing yeast. Then, we modified this engineered strain by replacing UGT109A1 with its mutant UGT109A1-K73A, overexpressing the cytochrome P450 reductase ATR2 from Arabidopsis thaliana and the key enzymes of UDP-glucose biosynthesis to increase the production of 3β,12β-Di-O-Glc-PPT, although these strategies did not show any positive effect on the yield of 3β,12β-Di-O-Glc-PPT. However, the unnatural ginsenoside 3β,12β-Di-O-Glc-PPT was produced in this study by constructing its biosynthetic pathway in yeast. To the best of our knowledge, this is the first report of producing 3β,12β-Di-O-Glc-PPT through yeast cell factories. Our work provides a viable route for the production of 3β,12β-Di-O-Glc-PPT, which lays a foundation for drug research and development.

1. Introduction

Panax species have been extensively utilized as traditional medicines in Asia for about 5000 years. Ginsenosides are major components of Panax species, which have a wide range of pharmacological effects. Over 150 kinds of natural ginsenosides have been isolated and identified [1,2]. Natural ginsenosides contain glycosyl groups at the C3-OH and/or C20-OH positions of protopanaxadiol (PPD) or at the C6-OH and/or C20-OH positions of protopanaxatriol (PPT), while the glycosylation of dammarenediol-II (DM) at the C3-OH and/or C20-OH positions and the glycosylation of PPD and/or PPT at the C12-OH position lead to the formation of a variety of unnatural ginsenosides. The positions and numbers of hydroxyls and/or glycosyls determine the difference in pharmacological activity among various ginsenosides [3]. PPD-type ginsenosides affect cell cycle distribution and produce cell death signals, showing excellent anti-tumor activity [4,5,6]. Ginsenoside Rg3 has been used in clinical applications for twenty years and shows inhibitory effects on lung cancer and liver cancer [7]. Ginsenoside compound K has been proven to have a range of anti-cancer activities against lung cancer, leukemia, breast cancer, and colon cancer [8,9,10,11]. Ginsenoside Rh2 shows anti-cancer activity against lung cancer and liver cancer by mediating the apoptosis of cancer cells and suppressing cell proliferation and migration [12,13]. PPT-type ginsenosides have unique activities on the central nervous system and cardiovascular system. PPT, ginsenoside Rh1, and Rg1 can enhance the excitability of the hippocampus, which is conducive to the acquisition, consolidation, and reproduction of learning and memory [14,15,16,17]. Rg1 is also the most promising lead compound for anti-aging due to its effect on clearing free radicals. Ginsenoside Rb1 and Rg2 can effectively prevent or slow the development of Alzheimer’s disease [18]. Rg2 directly acts on human neuron nicotinic acetylcholine receptors (nAChRs) and regulates their desensitization [19]. Ginsenoside Re exhibits anti-peripheral nerve injury, anti-cerebral ischemia, and anti-neurotoxicity activities on nervous system diseases [20,21,22].
Unnatural ginsenosides may show unexpected activities compared with the natural ones produced in Panax species. Ginsenoside Ia was transformed from F1 by BSGT1 of Bacillus subtilis and showed an inhibitory effect on melanogenesis in B16BL6 cells [23]. In our previous study, a new UDP-glycosyltransferase (UGT) named UGT109A1 from B. subtilis was demonstrated to catalyze the glycosylation of DM at the C3-OH and C20-OH positions and PPD at the C3-OH and C12-OH positions, respectively, leading to the formation of a series of unnatural ginsenosides [24]. It has been confirmed that the anti-lung cancer activity of 3,12-Di-O-β-D-glucopyranosyl-dammar-24-ene-3β,12β,20S-triol (3β,12β-Di-O-Glc-PPD) is superior to that of the natural ginsenoside Rg3 [24]; 3,20-Di-O-β-D-glucopyranosyl-dammar-24-ene-3β,20S-diol (3β,20S-Di-O-Glc-DM) displays better anti-pancreatic activity than both of the natural ginsenosides Rg3 and F2 [25]; and 3-O-β-D-glucopyranosyl-dammar-24-ene-3β,20S-diol (3β-O-Glc-DM) exhibits a higher activity on colon cancer than that of Rg3 and compound K [26]. UGT109A1 also can glycosylate the C3-OH and C12-OH positions of PPT simultaneously to produce another unnatural ginsenoside, 3β,12β-Di-O-Glc-PPT (Figure 1). Due to the presence of the glucosyl moiety at the C12 position, it is speculated that 3β,12β-Di-O-Glc-PPT may have unique pharmacological activity different from natural PPT-type ginsenosides. Thus, it is necessary to accumulate a large amount of 3β,12β-Di-O-Glc-PPT for research on its pharmacological activity.
At present, there are many ways to produce ginsenosides, including chemical synthesis, enzymatic catalysis, biotransformation, and hetero-biosynthesis. Chemical synthesis has disadvantages in terms of sustainability, selectivity, and renewability. The methods of enzymatic catalysis and biotransformation are becoming more and more mature, but they still need a large amount of expensive substrates, which is a bottleneck for their development and utilization. The development of synthetic biology has made it possible to achieve the de novo biosynthesis of ginsenosides, which has become a new research hotspot. Saccharomyces cerevisiae has an intrinsic mevalonate biosynthesis pathway, providing the basic precursor of 2,3-oxidosqualene for triterpene synthesis. It also produces UDP-glucose (UDPG) and is therefore considered as an ideal chassis for constructing cell factories of triterpenoid saponins. Great progress has been made in the synthesis of PPD-type ginsenosides from cheap carbon sources in S. cerevisiae, such as Rg3 (1.3 mg/L) [27], Rh2 (2.25 g/L) [28], and compound K (5.74 g/L) [29]. The biosynthetic pathways of PPT-type ginsenosides are longer than those of PPD-type ginsenosides. The low efficiency of protopanaxatriol synthase (PPTS) makes the heterologous biosynthesis of PPT-type ginsenosides more difficult. Dai et al. [30] constructed the biosynthetic pathway of PPT in yeast by introducing Panax ginseng dammarenediol-II synthase (DS), protopanaxadiol synthase (PPDS), PPTS, and Arabidopsis thaliana ATR1 and overexpressing the key enzymes involved in the upstream pathway to improve the precursor supply. The PPT titer of the engineered yeast was 15.9 mg/L. Li et al. [31] constructed a high-yield PPT-producing yeast strain through protein engineering with a titer of over 5 g/L in a 1.3 L bioreactor. By introducing PgUGT71A54, the cell factory producing ginsenoside Rg1 was built with a titer of 1.95 g/L.
In this study, we utilized the glycosyltransferase UGT109A1 from B. subtilis to achieve the production of 3β,12β-Di-O-Glc-PPT in the engineered yeast. First, we designed the co-expressed gene module harboring PPTS and UGT109A1, which was integrated into the chromosomes of the PPD-producing yeast, resulting in the 3β,12β-Di-O-Glc-PPT-producing strain with a titer of 7.0 mg/L. Then, we replaced UGT109A1 with its mutant UGT109A1-K73A to enhance the glycosylation efficiency. We also modified the engineered strain by overexpressing the cytochrome P450 reductase ATR2 and/or overexpressing the key enzymes of UDPG biosynthesis to improve the supply of the glycosyl acceptor and donor. Unexpectedly, the application of these optimization strategies did not increase the 3β,12β-Di-O-Glc-PPT production. However, an unnatural PPT-type ginsenoside, 3β,12β-Di-O-Glc-PPT, was obtained by the yeast cell factories, providing a raw material for innovative drug research.

2. Materials and Methods

2.1. Strains

Escherichia coli Trans1-T1 (TransGen Biotech, Beijing, China) was used for plasmid amplification. The PPD-producing strain YPU [32] was chosen as the yeast chassis for constructing the engineered yeasts.

2.2. The Engineered Yeast Construction

The gene expression cassettes harboring ATR2, UGT109A1, UGT109A1-K73A, PGM1, PGM2, and UGP1 were constructed according to the method described previously [24,25,26,32]. The gene of PPTS was synthesized (GenScript, Nanjing, China) according to the codon bias of S. cerevisiae and was inserted into gene expression cassette with promoter TEF1 and terminator CYC1.
The integration modules containing PPTS, UGT109A1, and HIS3 were transformed into the δ1 site of the PPD-producing strain YPU to generate strain YPUT. The integration modules containing PPTS, UGT109A1-K73A, and HIS3 were transformed into the δ1 site of strain YPU to generate strain YPUK. The integration modules containing PPTS, ATR2, UGT109A1, and HIS3 were transformed into the δ1 site of strain YPU to generate strain YPUT-B. The integration modules containing PGM1, PGM2, UGP1, PPTS, UGT109A1, and HIS3 were transformed into the δ1 site of strain YPU to construct strain YPUT-C. The integration modules harboring PGM1, PGM2, UGP1, PPTS, ATR2, UGT109A1, and HIS3 were transformed into the δ1 site of strain YPU to construct strain YPUT-D. Colonies of each group were randomly picked from SD-HIS-LEU-TRP-URA plates and validated by PCR amplification. Eight correct colonies were cultivated in YPD medium for 6 days and the titers of 3β,12β-Di-O-Glc-PPT were detected.
All the constructed strains are listed in Table 1.

2.3. Fermentation and Product Extraction of the Engineered Yeasts

The shake flask fermentation of all the engineered strains was performed as described in our previous work [32]. When the fermentation was finished, the cultured cells were harvested by centrifugation (10,000× g, 3 min). The extraction and analysis methods of ginsenosides and aglycones have been described previously [32]. All data are presented as means ± SD of three independent repeat experiments (n = 3).

2.4. Product Analysis by HPLC, HPLC-ESI-MS and NMR

The methods of HPLC, HPLC-ESI-MS, and NMR were used as previously reported [24]. The detailed conditions of analytic and semi-preparative HPLC for extracts of yeast fermentation are described in Table S1.

3. Results

3.1. Constructing 3β,12β-Di-O-Glc-PPT-Producing Yeast

We have demonstrated that UGT109A1 from B. subtilis can transfer a glucosyl moiety to the C3-OH and C12-OH positions of PPT simultaneously to yield 3β,12β-Di-O-Glc-PPT [24]. The recombinant enzyme can catalyze the formation of 3β,12β-Di-O-Glc-PPT by using UDPG as a sugar donor and PPT as a sugar acceptor in a reaction system. Since PPT is a very expensive chemical, we sought to construct engineered yeasts to achieve the de novo biosynthesis of 3β,12β-Di-O-Glc-PPT, which omits the external supplementary of PPT (Figure 1).
Previously, we constructed a PPD-producing strain YPU that could produce 93.1 mg/L PPD [32]. For producing 3β,12β-Di-O-Glc-PPT, the gene expression module containing PPTS from P. ginseng and UGT109A1 from B. subtilis was inserted into the δ1 site of strain YPU, generating strain YPUT. After cultivation for six days, the target product 3β,12β-Di-O-Glc-PPT in the intracellular extracts was identified by an HPLC analysis (Figure 2A), and its structure was elucidated by HPLC-ESI-MS, 1H NMR, and 13C NMR (Figure 2B and Figure S1). The titer of 3β,12β-Di-O-Glc-PPT in strain YPUT was 7.0 mg/L. There was no PPT detected from the extract of YPUT, but there was a large amount of 3β,12β-Di-O-Glc-PPD and 3β,20S-Di-O-Glc-DM (Figure 2A). It suggested that the low-efficiency PPTS was a key limiting factor in the biosynthetic pathway of 3β,12β-Di-O-Glc-PPT. In addition, neither 3β-O-Glc-PPT nor 12β-O-Glc-PPT was detected in the extract. A possible reason is that UGT109A1 can effectively and rapidly transform 3β-O-Glc-PPT or 12β-O-Glc-PPT into 3β,12β-Di-O-Glc-PPT. Further experiments are needed to determine the order in which the C3-OH and C12-OH of PPT are glycosylated. To the best of our knowledge, this is the first report of the de novo biosynthesis of 3β,12β-Di-O-Glc-PPT in yeast through the strategy of metabolic engineering.

3.2. Improving the Catalytic Efficiency of UGT109A1 to Promote the Glycosylation of PPT

Although we obtained the 3β,12β-Di-O-Glc-PPT-producing yeast, the yield was low. The catalytic efficiency and expression level of biosynthetic enzymes have important effects on the yield of the target products. The K73A mutation of UGT109A1 has been proven to markedly improve 3β,12β-Di-O-Glc-PPD production because the mutation expands the active pocket and increases its hydrophobicity [32]. Since the structures of 3β,12β-Di-O-Glc-PPD and 3β,12β-Di-O-Glc-PPT differ only at the C6 position, at which 3β,12β-Di-O-Glc-PPT possesses an extra hydroxyl group, we supposed that the mutant UGT109A1-K73A might also improve the glycosylation efficiency towards PPT and promote the yield of 3β,12β-Di-O-Glc-PPT. Strain YPUK was therefore constructed by replacing the wild-type UGT109A1 with its mutant UGT109A1-K73A.
After cultivation for six days, the titer of 3β,12β-Di-O-Glc-PPT was 6.8 mg/L in strain YPUK, which was even slightly lower than that in strain YPUT (Figure 3). There were two possible reasons for this result: one is that the K73A mutation did not improve the glycosylation efficiency of UGT109A1 towards PPT and even inhibited it to a certain extent; the other is that the catalytic efficiency of UGT109A1 was improved, but the low efficiency of PPTS limited the supply of precursor PPT. Notably, PPT was undetected in either strain YPUT or strain YPUK, indicating that our second conjecture was more likely. Thus, we speculated that the key was to ensure the sufficient supply of the precursor in order to boost the production of 3β,12β-Di-O-Glc-PPT.

3.3. Improving Precursor Supply for the Production of 3β,12β-Di-O-Glc-PPT

The supply of both glycosyl receptor PPT and glycosyl donor UDPG is important to the production of PPT-type ginsenosides. It has been reported that the balance between CYP450s and their reductases has a significant role in the formation of natural products [33]. In the host strain YPU, the reductase ATR2 from A. thaliana was shared by PPDS and PPTS [30]. Herein, we hypothesize that the low 3β,12β-Di-O-Glc-PPT production may be due to the mismatch of the expression levels of PPTS and ATR2. Thus, we optimized the oxidation–reduction system by overexpressing ATR2 in S. cerevisiae. The gene module harboring PPTS, ATR2, and UGT109A1 was inserted into the δ1 site of strain YPU, resulting in strain YPUT-B.
UDPG is an important sugar donor, and its natural supply is very limited in S. cerevisiae, and thus it probably cannot meet the demand for the synthesis of ginsenosides. Wang et al. [34] overexpressed S. cerevisiae phosphoglucomutase 1 (PGM1), phosphoglucomutase 2 (PGM2), and UDPG pyrophosphorylase 1 (UGP1) in the engineered yeast, increasing the production of UDPG more than eight times. Improving the supply of UDPG in yeast has effectively promoted the production of ginsenosides F1 and compound K in recombinant strains [34,35]. Here, we generated the expression cassettes of the three enzymes and integrated them into the δ1 site of strain YPU together with those of PPTS and UGT109A1, resulting in strain YPUT-C. Then, we overexpressed ATR2 and the key enzymes involved in UDPG synthesis to increase the supply of both the glycosyl donor and receptor at the same time, resulting in strain YPUT-D.
After cultivation for six days, the production of 3β,12β-Di-O-Glc-PPT in these engineered strains was identified and quantified. The 3β,12β-Di-O-Glc-PPT titers of the three strains were all lower than that of YPUT (Figure 3). The lowest titer of 3β,12β-Di-O-Glc-PPT appeared in strain YPUT-D (1.9 mg/L) (Table 2), which was only 27% that in strain YPUT. According to the results, the optimization of the glycosyl donor and receptor supply did not improve the production of 3β,12β-Di-O-Glc-PPT in the engineered strain. What is more, the overexpression of enzymes related to the synthesis of UDPG showed an inexplicable inhibitory impact on the production of 3β,12β-Di-O-Glc-PPT.

4. Discussion

As the major active components of Panax species, ginsenosides exhibit many pharmacological activities [36,37,38,39,40,41,42,43,44]. The different positions, numbers, and types of sugar moieties result in the different medicinal properties of ginsenosides. UGT109A1 from B. subtilis can glycosylate triterpenoid aglycones DM, PPD, and PPT to yield a variety of unnatural ginsenosides [24]. Among them, 3β,12β-Di-O-Glc-PPT is a novel compound reported by our lab. However, it remains ambiguous whether UGT109A1 can synthesize 3β,12β-Di-O-Glc-PPT in vivo in S. cerevisiae. In the present work, we integrated PPTS and UGT109A1 genes into the chromosome of the PPD-producing strain, and 3β,12β-Di-O-Glc-PPT was detected as expected, indicating that PPT was synthesized by PPTS and then glycosylated by UGT109A1 in vivo.
The catalytic efficiency of UGT109A1-K73A towards PPD was improved markedly compared with that of the wild-type UGT109A1 [32]. Considering the similarity of the structures of PPD and PPT, strain YPUK was constructed by expressing the mutant UGT109A1-K73A to improve the glycosylation efficiency. However, the titer of 3β,12β-Di-O-Glc-PPT in strain YPUK was not improved compared with that in strain YPUT. Since the distance between the C6-OH and C3-OH of 3β,12β-Di-O-Glc-PPT was close, the C6-OH may change the relative position of the substrate in the binding pocket of UGT109A1-K73A and hinder the glycosylation process. Moreover, the aglycone PPT was undetectable in the two engineered strains. It can be speculated that the catalytic efficiency of PPTS is far less than the glycosylation efficiency of UGT109A1, so the rate of PPT biosynthesis cannot keep up with the rate of glycosylation to form 3β,12β-Di-O-Glc-PPT.
The precursor supply level is known to play a crucial role in improving the production of the target products in the engineered strains. Since ATR2 derived from A. thaliana was considered to be a universal NADPH-cytochrome P450 reductase (CPR) shared by PPDS and PPTS [30], PPTS was effectively expressed in strain YPUT. The 3β,12β-Di-O-Glc-PPT titer of strain YPUT-B was lower than that of strain YPUT, implying that the overexpression of ATR2 in strain YPUT-B may destroy the balance between PPTS and its reductase partner. In the meantime, PPT was undetected in either strain YPUT or YPUT-B. PPTS turned out to be a rate-limiting enzyme for the production of PPT-type ginsenosides. Redox partners play an important role for CYP450s. Different CPRs have diverse effects on the catalytic activity of CYP450s. An appropriate CPR can promote the efficient synthesis of compounds. CYP450/CPR pairs excavated from the same plant species generally show better effects than unnatural combinations [45,46,47,48]. RoCPR1 from Rosmarinus officinalis stood out in a group of plant-derived CPRs when co-expressed with the RoCYP01 from R. officinalis to promote the production of betulinic acid (BA). The titer of BA obtained from the engineered strain increased to more than 1 g/L [49]. Zhu et al. [50] compared the coupling efficiency of six CPRs from different plants, including A. thaliana, Lotus japonicus, G. uralensis, and Medicago sativa, with Uni25647 and CYP72A154 from Glycyrrhiza uralensis. The results showed that GuCPR1 from G. uralensis exhibited the best coupling efficiency with the two CYP450s among the six CPRs. However, Wang et al. [35] found that the CPR from Vitis vinifera (VvCPR) combined with PgPPDS boosted the PPD production better than those from A. thaliana and P. ginseng. Since the CPR for PPTS with a high efficiency has not yet been identified, more CPRs from different sources should be investigated to improve the production of PPT.
In addition, there is a UDPG biosynthetic pathway composed of PGM1, PGM2, and UGP1 in S. cerevisiae, but its natural supply cannot meet the demand for the synthesis of ginsenosides. Previously, we improved the production of rare ginsenoside F2 and unnatural ginsenoside 3β,20S-Di-O-Glc-DM in yeast successfully by overexpressing the key enzymes related to the biosynthesis of UDPG [25]. In the present study, these enzymes were also overexpressed in an attempt to improve the yield of 3β,12β-Di-O-Glc-PPT. As shown in Table 2, the titer of 3β,12β-Di-O-Glc-PPT in strain YPUT-C and strain YPUT-D decreased markedly compared with that of strain YPUT, while the titer of 3β,12β-Di-O-Glc-PPD increased from 4.8 to 7.2 mg/L. These results suggested that there was competition between the glycosylation of PPD and PPT. Similar studies showed that amplifying the supply of UDPG apparently increased the production PPD-type ginsenosides [34,35]. Since the catalytic activity of UGT109A1 towards PPD was higher than that towards PPT, the optimization of the UDPG supply was more effective in improving the glycosylation of PPD. The production of 3β,12β-Di-O-Glc-PPT inevitably decreased in the engineered strains.
Other strategies need to be further developed to increase the yield of 3β,12β-Di-O-Glc-PPT in the engineered yeast, including the protein engineering of PgPPTS and UGT109A1, regulation of key enzymes’ expression, utilization of subcellular compartments, and cofactor engineering. Moreover, the yield of 3β,12β-Di-O-Glc-PPT can be further increased by optimizing the fermentation conditions of the engineered strain.

5. Conclusions

In this study, we achieved the de novo biosynthesis of 3β,12β-Di-O-Glc-PPT in S. cerevisiae by expressing PPTS and UGT109A1 in the PPD-producing strain, and the titer was 7.0 mg/L. Then, we sought to improve the efficiency of glycosyltransferase by replacing the wild-type UGT109A1 with its mutant UGT109A1-K73A and to increase the supply of the glycosyl receptor and donor by overexpressing ATR2 and the enzymes involved in the biosynthetic pathway of UDPG. Unexpectedly, these strategies did not boost the titer of 3β,12β-Di-O-Glc-PPT. However, to the best of our knowledge, this is the first report of developing a yeast cell factory for the de novo biosynthesis of unnatural ginsenoside 3β,12β-Di-O-Glc-PPT. This study has explored a green and sustainable approach to producing 3β,12β-Di-O-Glc-PPT based on synthetic biology, which lays a foundation for research on its pharmacological activity.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/bioengineering10040463/s1, Figure S1: The 1H-NMR (A) and 13C-NMR (B) spectra of 3,12-Di-O-β-D-glucopyranosyl-dammar-24-ene-3β,6α,12β,20S-tetraol (3β,12β-Di-O-Glc-PPT) extracted from strain YPUT; Table S1: The HPLC conditions for analysis and preparation of products from the engineered strains.

Author Contributions

Conceptualization, J.Y. and P.Z.; Data curation, T.G., J.C. and T.C.; Methodology, Supervision, J.Y. and P.Z.; Writing—original draft, C.Z.; Writing—review and editing, Funding Acquisition, J.Y. and P.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the grants from the Beijing Natural Science Foundation (7212158), the CAMS Innovation Fund for Medical Sciences (CIFMS) (2021-I2M-1-029), and the National Key Research and Development Program of China (2022YFF1100300).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kim, Y.J.; Zhang, D.; Yang, D.C. Biosynthesis and biotechnological production of ginsenosides. Biotechnol. Adv. 2015, 33, 717–735. [Google Scholar]
  2. Mohanan, P.; Yang, T.J.; Song, Y.H. Genes and regulatory mechanisms for ginsenoside biosynthesis. J. Plant Biol. 2023, 66, 87–97. [Google Scholar]
  3. Ali, M.Y.; Jannat, S.; Rahman, M.M. Ginsenoside derivatives inhibit advanced glycation end-product formation and glucose–fructose mediated protein glycation in vitro via a specific structure–activity relationship. Bioorg. Chem. 2021, 111, 104844. [Google Scholar]
  4. Ota, T.; Maeda, M.; Odashima, S. Mechanism of action of ginsenoside Rh2: Uptake and metabolism of ginsenoside Rh2 by cultured B16 melanoma cells. J. Pharm. Sci. 1991, 80, 1141–1146. [Google Scholar]
  5. Chen, Y.; Xu, Y.; Zhu, Y.; Li, X. Anti-cancer effects of ginsenoside compound K on pediatric acute myeloid leukemia cells. Cancer Cell Int. 2013, 13, 24. [Google Scholar]
  6. Kim, A.D.; Kang, K.A.; Kim, H.S.; Kim, D.H.; Choi, Y.H.; Lee, S.J.; Kim, H.S.; Hyun, J.W. A ginseng metabolite, compound K, induces autophagy and apoptosis via generation of reactive oxygen species and activation of JNK in human colon cancer cells. Cell Death Dis. 2013, 4, e750. [Google Scholar]
  7. Sun, M.; Ye, Y.; Xiao, L.; Duan, X.; Zhang, Y.; Zhang, H. Anticancer effects of ginsenoside Rg3. Int. J. Mol. Med. 2017, 39, 507–518. [Google Scholar]
  8. Yang, X.D.; Yang, Y.Y.; Ouyang, D.S.; Yang, G.P. A review of biotransformation and pharmacology of ginsenoside compound K. Fitoterapia. 2015, 100, 208–220. [Google Scholar]
  9. Metwaly, A.M.; Lianlian, Z.; Luqi, H.; Deqiang, D. Black ginseng and its saponins: Preparation, phytochemistry and pharmacological effects. Molecules. 2019, 24, 1856. [Google Scholar]
  10. Liu, J.; Wang, Y.; Yu, Z.; Lv, G.; Huang, X.; Lin, H.; Ma, C.; Lin, Z.; Qu, P. Functional mechanism of ginsenoside compound K on tumor growth and metastasis. Integr. Cancer Ther. 2022, 21, 1–13. [Google Scholar]
  11. Liu, T.; Zhu, L.; Wang, L. A narrative review of the pharmacology of ginsenoside compound K. Ann. Transl. Med. 2022, 10, 234. [Google Scholar]
  12. Ge, G.; Yan, Y.; Cai, H. Ginsenoside Rh2 inhibited proliferation by inducing ROS mediated ER stress dependent apoptosis in lung cancer cells. Biol. Pharm. Bull. 2017, 40, 2117–2124. [Google Scholar]
  13. Li, Q.; Li, B.; Dong, C.; Wang, Y.; Li, Q. 20(S)-Ginsenoside Rh2 suppresses proliferation and migration of hepatocellular carcinoma cells by targeting EZH2 to regulate CDKN2A-2B gene cluster transcription. Eur. J. Pharmacol. 2017, 815, 173–180. [Google Scholar]
  14. Mook-Jung, I.; Hong, H.-S.; Boo, J.H.; Lee, K.H.; Yun, S.H.; Cheong, M.Y.; Joo, I.; Huh, K.; Jung, M.W. Ginsenoside Rb1 and Rg1 improve spatial learning and increase hippocampal synaptophysin level in mice. J. Neurosci. Res. 2001, 63, 509–515. [Google Scholar]
  15. Wang, X.Y.; Zhang, J.T. NO mediates ginsenoside Rg1-induced long-term potentiation in anesthetized rats. Acta Pharmacol. Sin. 2001, 22, 1099–1102. [Google Scholar]
  16. Wang, Y.Z.; Chen, J.; Chu, S.F.; Wang, Y.S.; Wang, X.Y.; Chen, N.H.; Zhang, J.T. Improvement of memory in mice and increase of hippocampal excitability in rats by ginsenoside Rg1’s metabolites ginsenoside Rh1 and protopanaxatriol. J. Pharmacol. Sci. 2009, 109, 504–510. [Google Scholar]
  17. Shang, D.; Li, Z.; Tan, X.; Liu, H.; Tu, Z. Inhibitory effects and molecular mechanisms of ginsenoside Rg1 on the senescence of hematopoietic stem cells. Fundam. Clin. Pharmacol. 2023, 1–9. [Google Scholar] [CrossRef]
  18. Lu, J.; Wang, X.; Wu, A.; Cao, Y.; Dai, X.; Liang, Y.; Li, X. Ginsenosides in central nervous system diseases: Pharmacological actions, mechanisms, and therapeutics. Phytother. Res. 2022, 36, 1523–1544. [Google Scholar]
  19. Sala, F.; Mulet, J.; Choi, S.; Jung, S.-Y.; Nah, S.-Y.; Rhim, H.; Valor, L.M.; Criado, M.; Sala, S. Effects of ginsenoside Rg2 on human neuronal nicotinic acetylcholine receptors. J. Pharmacol. Exp. Ther. 2002, 301, 1052–1059. [Google Scholar]
  20. Wang, L.; Yuan, D.; Zhang, D.; Zhang, W.; Liu, C.; Cheng, H.; Song, Y.; Tan, Q. Ginsenoside Re promotes nerve regeneration by facilitating the proliferation, differentiation and migration of schwann cells via the ERK- and JNK-dependent pathway in rat model of sciatic nerve crush injury. Cell. Mol. Neurobiol. 2015, 35, 827–840. [Google Scholar]
  21. Chen, L.M.; Zhou, X.M.; Cao, Y.L.; Hu, W.X. Neuroprotection of ginsenoside Re in cerebral ischemia-reperfusion injury in rats. J. Asian Nat. Prod. Res. 2008, 10, 439–445. [Google Scholar]
  22. Tu, T.-H.T.; Sharma, N.; Shin, E.-J.; Tran, H.-Q.; Lee, Y.J.; Jeong, J.H.; Jeong, J.H.; Nah, S.Y.; Tran, H.-Y.P.; Byun, J.K.; et al. Ginsenoside Re protects trimethyltin-induced neurotoxicity via activation of IL-6-mediated phosphoinositol 3-kinase/Akt signaling in mice. Neurochem. Res. 2017, 42, 3125–3139. [Google Scholar]
  23. Wang, D.D.; Jin, Y.; Wang, C.; Kim, Y.J.; Perez, Z.E.J.; Baek, N.I.; Mathiyalagan, R.; Markus, J.; Yang, D.C. Rare ginsenoside Ia synthesized from F1 by cloning and overexpression of the UDP-glycosyltransferase gene from Bacillus subtilis: Synthesis, characterization, and in vitro melanogenesis inhibition activity in BL6B16 cells. J. Ginseng Res. 2018, 42, 42–49. [Google Scholar]
  24. Liang, H.; Hu, Z.; Zhang, T.; Gong, T.; Chen, J.; Zhu, P.; Li, Y.; Yang, J. Production of a bioactive unnatural ginsenoside by metabolically engineered yeasts based on a new UDP-glycosyltransferase from Bacillus subtilis. Metab. Eng. 2017, 44, 60–69. [Google Scholar]
  25. Jiang, F.; Zhou, C.; Li, Y.; Deng, H.; Gong, T.; Chen, J.; Chen, T.; Yang, J.; Zhu, P. Metabolic engineering of yeasts for green and sustainable production of bioactive ginsenosides F2 and 3β,20S-Di-O-Glc-DM. Acta Pharm. Sin. B. 2022, 12, 3167–3176. [Google Scholar]
  26. Hu, Z.-F.; Gu, A.-D.; Liang, L.; Li, Y.; Gong, T.; Chen, J.-J.; Chen, T.-J.; Yang, J.-L.; Zhu, P. Construction and optimization of microbial cell factories for sustainable production of bioactive dammarenediol-II glucosides. Green Chem. 2019, 21, 3286–3299. [Google Scholar]
  27. Jung, S.C.; Kim, W.; Park, S.C.; Jeong, J.; Park, M.K.; Lim, S.; Lee, Y.; Im, W.T.; Lee, J.H.; Choi, G.; et al. Two ginseng UDP-glycosyltransferases synthesize ginsenoside Rg3 and Rd. Plant Cell Physiol. 2014, 55, 2177–2188. [Google Scholar]
  28. Wang, P.; Wei, W.; Ye, W.; Li, X.; Zhao, W.; Yang, C.; Li, C.; Yan, X.; Zhou, Z. Synthesizing ginsenoside Rh2 in Saccharomyces cerevisiae cell factory at high-efficiency. Cell Discov. 2019, 5, 5. [Google Scholar]
  29. Wang, P.; Wang, J.; Zhao, G.; Yan, X.; Zhou, Z. Systematic optimization of the yeast cell factory for sustainable and high efficiency production of bioactive ginsenoside compound K. Synth. Syst. Biotechnol. 2021, 6, 69–76. [Google Scholar]
  30. Dai, Z.; Wang, B.; Liu, Y.; Shi, M.; Wang, D.; Zhang, X.; Liu, T.; Huang, L.; Zhang, X. Producing aglycons of ginsenosides in bakers’ yeast. Sci. Rep. 2014, 4, 3698. [Google Scholar]
  31. Li, X.; Wang, Y.; Fan, Z.; Wang, Y.; Wang, P.; Yan, X.; Zhou, Z. High-level sustainable production of the characteristic protopanaxatriol-type saponins from Panax species in engineered Saccharomyces cerevisiae. Metab. Eng. 2021, 66, 87–97. [Google Scholar]
  32. Zhou, C.; Chen, T.; Gu, A.; Hu, Z.; Li, Y.; Gong, T.; Chen, J.; Yang, J.; Zhu, P. Combining protein and metabolic engineering to achieve green biosynthesis of 12β-O-Glc-PPD in Saccharomyces cerevisiae. Green Chem. 2023, 25, 1356–1367. [Google Scholar]
  33. Zhao, Y.; Fan, J.; Wang, C.; Feng, X.; Li, C. Enhancing oleanolic acid production in engineered Saccharomyces cerevisiae. Bioresour. Technol. 2018, 257, 339–343. [Google Scholar]
  34. Wang, J.H.; Wang, D.; Li, W.X.; Huang, Y.; Dai, Z.B.; Zhang, X.L. Optimization of UDP-glucose supply module and production of ginsenoside F1 in Saccharomyces cerevisiae. Zhongguo Zhong Yao Za Zhi 2019, 44, 4596–4604. [Google Scholar]
  35. Wang, D.; Wang, J.; Shi, Y.; Li, R.; Fan, F.; Huang, Y.; Li, W.; Chen, N.; Huang, L.; Dai, Z.; et al. Elucidation of the complete biosynthetic pathway of the main triterpene glycosylation products of Panax notoginseng using a synthetic biology platform. Metab. Eng. 2020, 61, 131–140. [Google Scholar]
  36. Zhou, L.; Li, Z.K.; Li, C.Y.; Liang, Y.Q.; Yang, F. Anticancer properties and pharmaceutical applications of ginsenoside compound K: A review. Chem. Bio. Drug Des. 2021, 99, 286–300. [Google Scholar]
  37. Matsuda, H.; Samukawa, K.; Kubo, M. Anti-inflammatory activity of ginsenoside Ro1. Planta Med. 1990, 56, 19–23. [Google Scholar]
  38. Xie, J.T.; Mehendale, S.R.; Li, X.; Quigg, R.; Wang, X.; Wang, C.Z.; Wu, J.A.; Aung, H.H.; Rue, P.A.; Bell, G.I.; et al. Anti-diabetic effect of ginsenoside Re in ob/ob mice. Biochim. Biophys. Acta Mol. Basis Dis. 2005, 1740, 319–325. [Google Scholar]
  39. Sun, Y.; Liu, Y.; Chen, K. Roles and mechanisms of ginsenoside in cardiovascular diseases: Progress and perspectives. Sci. China Life Sci. 2016, 59, 292–298. [Google Scholar]
  40. Park, E.K.; Choo, M.K.; Kim, E.J.; Han, M.J.; Kim, D.H. Antiallergic activity of ginsenoside Rh2. Biol. Pharm. Bull. 2003, 26, 1581–1584. [Google Scholar]
  41. Liu, T.; Zuo, L.; Guo, D.; Chai, X.; Xu, J.; Cui, Z.; Wang, Z.; Hou, C. Ginsenoside Rg3 regulates DNA damage in non-small cell lung cancer cells by activating VRK1/P53BP1 pathway. Biomed. Pharmacother. 2019, 120, 109483. [Google Scholar]
  42. Chen, C.; Lv, Q.; Li, Y.; Jin, Y.H. The anti-tumor effect and underlying apoptotic mechanism of ginsenoside Rk1 and Rg5 in human liver cancer cells. Molecules. 2021, 26, 3926. [Google Scholar]
  43. Nakhjavani, M.; Palethorpe, H.M.; Tomita, Y.; Smith, E.; Price, T.J.; Yool, A.J.; Pei, J.V.; Townsend, A.R.; Hardingham, J.E. Stereoselective anti-cancer activities of ginsenoside Rg3 on triple negative breast cancer cell models. Pharmaceutics. 2019, 12, 117. [Google Scholar]
  44. Huang, Q.; Zhang, H.; Bai, L.P.; Law, B.Y.K.; Xiong, H.; Zhou, X.; Xiao, R.; Qu, Y.Q.; Mok, S.W.F.; Liu, L.; et al. Novel ginsenoside derivative 20 (S)-Rh2E2 suppresses tumor growth and metastasis in vivo and in vitro via intervention of cancer cell energy metabolism. Cell Death Dis. 2020, 11, 621. [Google Scholar]
  45. Guo, J.; Zhou, Y.J.; Hillwig, M.L.; Shen, Y.; Yang, L.; Wang, Y.; Zhang, X.; Liu, W.; Peters, R.J.; Chen, X.; et al. CYP76AH1 catalyzes turnover of miltiradiene in tanshinones biosynthesis and enables heterologous production of ferruginol in yeasts. Proc. Natl. Acad. Sci. USA. 2013, 110, 12108–12113. [Google Scholar]
  46. Girvan, H.M.; Munro, A.W. Applications of microbial cytochrome P450 enzymes in biotechnology and synthetic biology. Curr. Opin. Chem. Biol. 2016, 31, 136–145. [Google Scholar]
  47. Durairaj, P.; Li, S. Functional expression and regulation of eukaryotic cytochrome P450 enzymes in surrogate microbial cell factories. Eng. Microbiol. 2022, 2, 100011. [Google Scholar]
  48. Jin, Z.; Cong, Y.; Zhu, S.; Xing, R.; Zhang, D.; Yao, X.; Wan, R.; Wang, Y.; Yu, F. Two classes of cytochrome P450 reductase genes and their divergent functions in Camptotheca acuminata Decne. Int. J. Biol. Macromol. 2019, 138, 1098–1108. [Google Scholar]
  49. Huang, J.; Zha, W.; An, T.; Dong, H.; Huang, Y.; Wang, D.; Yu, R.; Duan, L.; Zhang, X.; Peters, R.J.; et al. Identification of RoCYP01 (CYP716A155) enables construction of engineered yeast for high-yield production of betulinic acid. Appl. Microbiol. Biotechnol. 2019, 103, 7029–7039. [Google Scholar]
  50. Zhu, M.; Wang, C.; Sun, W.; Zhou, A.; Wang, Y.; Zhang, G.; Zhou, X.; Huo, Y.; Li, C. Boosting 11-oxo-β-amyrin and glycyrrhetinic acid synthesis in Saccharomyces cerevisiae via pairing novel oxidation and reduction system from legume plants. Metab. Eng. 2018, 45, 43–50. [Google Scholar]
Figure 1. Construction of the biosynthetic pathway of 3β,12β-Di-O-Glc-PPT in the engineered yeast. The intrinsic Saccharomyces cerevisiae pathway is indicated by black and grey arrows. The double arrows show the multiple-step reactions. The Panax ginseng pathway is shown by blue arrows; the glycosylation reactions are shown by red arrows. The genes in black are up-regulated; the gene in grey is down-regulated; the genes in green are from P. ginseng; the gene in magenta is from Arabidopsis thaliana; and the gene in red is from Bacillus subtilis. The compounds in blue background are aglycones; the compounds in purple background are by-products; and the compound in orange background is target product.
Figure 1. Construction of the biosynthetic pathway of 3β,12β-Di-O-Glc-PPT in the engineered yeast. The intrinsic Saccharomyces cerevisiae pathway is indicated by black and grey arrows. The double arrows show the multiple-step reactions. The Panax ginseng pathway is shown by blue arrows; the glycosylation reactions are shown by red arrows. The genes in black are up-regulated; the gene in grey is down-regulated; the genes in green are from P. ginseng; the gene in magenta is from Arabidopsis thaliana; and the gene in red is from Bacillus subtilis. The compounds in blue background are aglycones; the compounds in purple background are by-products; and the compound in orange background is target product.
Bioengineering 10 00463 g001
Figure 2. (A) HPLC detection of the n-butanol extracts of strain YPUT. (B) ESI-MS spectrum of 3β,12β-Di-O-Glc-PPT.
Figure 2. (A) HPLC detection of the n-butanol extracts of strain YPUT. (B) ESI-MS spectrum of 3β,12β-Di-O-Glc-PPT.
Bioengineering 10 00463 g002
Figure 3. Production of 3β,12β-Di-O-Glc-PPT in the engineered strains. The gene in pink box is derived from P. ginseng. The gene in orange box is derived from A. thaliana. The gene in blue box is derived from B. subtilis. The genes in green boxes are derived from S. cerevisiae.
Figure 3. Production of 3β,12β-Di-O-Glc-PPT in the engineered strains. The gene in pink box is derived from P. ginseng. The gene in orange box is derived from A. thaliana. The gene in blue box is derived from B. subtilis. The genes in green boxes are derived from S. cerevisiae.
Bioengineering 10 00463 g003
Table 1. The strains used in this study.
Table 1. The strains used in this study.
StrainGenotype or CharacteristicTarget Product
YPUPTEF1-synDS-GFP-TCYC1, PTDH3-tPPDS-L18I-TADH2, PTEF2-ATR2-TTPI1, PPGK1-tHMG1-TADH1, PTEF1-HAC1-TCYC1 and TRP1 marker gene integrated into the rDNA site, and PTDH3-IDI1-TTPI1, PPGK1-ERG20-TADH1, PTEF1-ERG9-TCYC1, PPGK1-ERG1-TADH1, PTEF1-ERG7-TCYC1 and LEU2 marker gene integrated into the δ4 site of Y-ΔHXK2PPD
YPUTPTEF1-PPTS-TCYC1, PTDH3-UGT109A1-TADH2 and HIS3 marker gene integrated into the δ1 site of YPU3β,12β-Di-O-Glc-PPT
YPUKPTEF1-PPTS-TCYC1, PTDH3-UGT109A1-K73A-TADH2 and HIS3 marker gene integrated into the δ1 site of YPU3β,12β-Di-O-Glc-PPT
YPUT-BPTEF1-PPTS-TCYC1, PPGK1-ATR2-TTPI1, PTDH3-UGT109A1-TADH2 and HIS3 marker gene integrated into the δ1 site of YPU3β,12β-Di-O-Glc-PPT
YPUT-CPTEF1-PGM1-TCYC1, PPGK1-PGM2-TADH1, PTDH3-UGP1-TADH2, PTEF1-PPTS-TCYC1, PTDH3-UGT109A1-TADH2 and HIS3 marker gene integrated into the δ1 site of YPU3β,12β-Di-O-Glc-PPT
YPUT-DPTEF1-PGM1-TCYC1, PPGK1-PGM2-TADH1, PTDH3-UGP1-TADH2, PTEF1-PPTS-TCYC1, PPGK1-ATR2-TTPI1, PTDH3-UGT109A1-TADH2 and HIS3 marker gene integrated into the δ1 site of YPU3β,12β-Di-O-Glc-PPT
Table 2. Yields of 3β,12β-Di-O-Glc-PPT and other by-products produced in the engineered strains.
Table 2. Yields of 3β,12β-Di-O-Glc-PPT and other by-products produced in the engineered strains.
Yield a (mg/L)YPUTYPUT-BYPUT-CYPUT-D
3β,12β-Di-O-Glc-PPT7.0 ± 0.36.7 ± 0.42.6 ± 0.11.9 ± 0.4
3β,12β-Di-O-Glc-PPD4.8 ± 0.25.2 ± 0.56.3 ± 0.57.2 ± 0.4
3β,20S-Di-O-Glc-DM3.8 ± 0.23.6 ± 0.23.7 ± 0.23.1 ± 0.3
PPT0000
a Yield referred to the sum of intracellular and extracellular contents.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, C.; Gong, T.; Chen, J.; Chen, T.; Yang, J.; Zhu, P. Production of a Novel Protopanaxatriol-Type Ginsenoside by Yeast Cell Factories. Bioengineering 2023, 10, 463. https://doi.org/10.3390/bioengineering10040463

AMA Style

Zhou C, Gong T, Chen J, Chen T, Yang J, Zhu P. Production of a Novel Protopanaxatriol-Type Ginsenoside by Yeast Cell Factories. Bioengineering. 2023; 10(4):463. https://doi.org/10.3390/bioengineering10040463

Chicago/Turabian Style

Zhou, Chen, Ting Gong, Jingjing Chen, Tianjiao Chen, Jinling Yang, and Ping Zhu. 2023. "Production of a Novel Protopanaxatriol-Type Ginsenoside by Yeast Cell Factories" Bioengineering 10, no. 4: 463. https://doi.org/10.3390/bioengineering10040463

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop