covid
Buscar en
Brazilian Journal of Microbiology
Toda la web
Inicio Brazilian Journal of Microbiology Comparative proteomic analyses of Hyphozyma roseonigra ATCC 20624 in response to...
Journal Information
Vol. 49. Issue S1.
Pages 160-165 (November 2018)
Share
Share
Download PDF
More article options
Visits
1152
Vol. 49. Issue S1.
Pages 160-165 (November 2018)
Biotechnology and Industrial Microbiology
Open Access
Comparative proteomic analyses of Hyphozyma roseonigra ATCC 20624 in response to sclareol
Visits
1152
Xiuwen Wanga, Xiaohua Zhanga, Qingshou Yaoa, Dongliang Huab, Jiayang Qina,
Corresponding author
qinjy@bzmc.edu.cn

Corresponding author at: College of Pharmacy, Binzhou Medical University, Yantai 264003, China.
a Binzhou Medical University, Yantai, China
b Shandong Academy of Sciences, Energy Research Institute, Key Laboratory for Biomass Gasification Technology of Shandong Province, Jinan, China
This item has received

Under a Creative Commons license
Article information
Abstract
Full Text
Bibliography
Download PDF
Statistics
Figures (4)
Show moreShow less
Tables (1)
Table 1. Proteins identified by mass spectrometry.
Additional material (1)
Abstract

Sclareol is an important intermediate for ambroxide synthesis industries. Hyphozyma roseonigra ATCC 20624 was the only reported strain capable of degrading sclareol to the main product of sclareol glycol, which is the precursor of ambroxide. To date, knowledge is lacking about the effects of sclareol on cells and the proteins involved in sclareol metabolism. Comparative proteomic analyses were conducted on the strain H. roseonigra ATCC 20624 by using sclareol or glucose as the sole carbon source. A total of 79 up-regulated protein spots with a >2.0-fold difference in abundance on 2-D gels under sclareol stress conditions were collected for further identification. Seventy spots were successfully identified and finally integrated into 30 proteins. The up-regulated proteins under sclareol stress are involved in carbon metabolism; and nitrogen metabolism; and replication, transcription, and translation processes. Eighteen up-regulated spots were identified as aldehyde dehydrogenases, which indicating that aldehyde dehydrogenases might play an important role in sclareol metabolism. Overall, this study may lay the fundamentals for further cell engineering to improve sclareol glycol production.

Keywords:
Comparative proteomic
Hyphozyma roseonigra
Sclareol
Sclareol glycol
Ambroxide
Full Text
Introduction

Sclareol, a diterpene alcohol isolated from Salvia sclarea (Clary sage), can be applied to medicine, cosmetics, health products, flavors and fragrances, and pesticides.1–3 The main application of sclareol is for synthesizing high-end substitutes of ambergris in the perfume industry.4 Given the supply shortage and price inflation of ambergris, a number of substitutes have been developed. Ambroxide is the most appreciated substitute of ambergris and is obtained from the semi-synthesis of sclareol.5–7 Sclareol can be transformed to ambroxide through redox and cyclization, during which sclareol glycol and sclareolide are important intermediates (Fig. 1). Thus, the biological production of sclareol glycol or sclareolide to complete the biosynthesis pathway has gained much interest due to the increasing demand of natural fragrances.

Fig. 1.

Diagram of conversion from sclareol to ambroxide.

(0.15MB).

Numerous strains reportedly transform sclareol, whereas most strains merely introduce hydroxyl at the main ring rather than modify the branched chain. Only few strains can transform sclareol to sclareol glycol or sclareolide. Cryptococcus albidus ATCC 20918 converts sclareol to sclareolide at a high yield of more than 100g/L and has been exploited for industrialization.8 To date, Hyphozyma roseonigra ATCC 20624, which can exist in both yeast-like and filamentous forms, is the only reported strain capable of degrading sclareol to the main product of sclareol glycol.9 However, the explicit conversion mechanism from sclareol to sclareol glycol and other products has not been reported. Knowledge is also lacking about the effects of sclareol on cell growth, considering that sclareol is a labdane diterpene with a high antimicrobial activity.10 These problems hinder the further improvement of conversion rate and the expanded application of the biotechnology. H. roseonigra ATCC 20624 can survive in basic inorganic salt medium with sclareol as the sole carbon source and is therefore a good candidate for exploring the mechanism to some extent.

In the present study, comparative proteomic analyses were conducted on the strain H. roseonigra ATCC 20624, with sclareol or glucose as the sole carbon source. The difference in protein expression between the two conditions is expected to enrich our knowledge on sclareol stress in microorganisms.

Material and methodsMicroorganism and cultivation conditions

H. roseonigra ATCC 20624 was purchased from ATCC (MD, USA) and grown on yeast-malt medium (per liter: 3g yeast extract, 3g malt extract, 10g glucose, and 5g tryptone at pH 6.2) at 24°C and 180rpm. The seed culture was incubated in two kinds of basic inorganic salt medium (BSM) containing sclareol (BSMS) or glucose (BSMG) to monitor the cell growth rates.

The modified BSM medium contained (per liter) 2.44g KH2PO4, 14.04g Na2HPO4·12H2O, 2g NH4Cl, 0.2g MgCl2·6H2O, 1mg CaCl2·2H2O, 0.01g yeast extract, 5mL metal ion mixture, and 0.2mL vitamin mixture. Meanwhile, the metal ion mixture included (per liter) 0.5g FeCl2·4H2O, 0.5g MnCl2, 0.1g NaMoO4·2H2O, 0.05g CuCl2, and 120mM HCl. The vitamin mixture contained (per liter) 2g calcium pantothenate, 1g creatine, 2g nicotinic acid, 2g pyridoxamine, 2.5mg cobalamin, and 1g para-aminobenzoic acid.

Tween 80 (0.8%, v/v) and sclareol (2mM) were added to BSM medium to prepare the BSMS medium, whereas glucose (1.33mM) was added to BSM medium to produce the BSMG medium. Cells grown in the BSMG medium were run in parallel with the samples in the BSMS medium.

Preparation of protein extracts and comparative proteomic analysis

Cells in the exponential growth phase were harvested from 50mL cultures by centrifugation at 4000×g for 10min at 4°C. The cell pellets were frozen in liquid nitrogen and stored at −80°C until proteome analysis.

Comparative proteomic analysis by 2-D gel was carried out as described in literature.11 Cells were suspended in lysis buffer of 2DE (Sangon Co., Shanghai, China) and ultrasonically disrupted at 80–100W for 3min. The debris was removed by centrifugation and the supernatant was incubated overnight with ice-cold acetone. After centrifugation at 12,000×g and 4°C for 30min, the precipitated proteins were solubilized in sample lysis buffer. Total protein concentrations were determined using the Non-interference Protein Assay Kit (Sangon Co., Shanghai, China). Protein spots on the gels were visualized by silver staining in accordance with the protocol.12 Each set of samples was independently analyzed in triplicate.

Analyses of 2-D gels

2-D gels were scanned at a resolution of 300 dots per inch (dpi) by an Image Scanner III (GE healthcare). Statistically significant differences were discerned by performing Student's t test. Protein spots showing significant changes in up-regulation and down-regulation were extracted, digested, rehydrated, and identified in a 5800 Plus MALDI TOF/TOFTM analyzer (Applied Biosystems). Data were acquired in a positive mass spectrometry (MS) reflect or by using a CalMix5 standard to calibrate the instrument (ABI 5800 Calibration Mixture).

ResultsGrowth characteristics of H. roseonigra ATCC 20624

The growth rates of H. roseonigra ATCC 20624 on media containing different sclareol concentrations (2, 6, and 10mM) were investigated. The growth curves of H. roseonigra ATCC 20624 on media containing 2 and 6mM sclareol were similar, and the optical density (OD600) of 2.23±0.11 for 2mM sclareol and 2.41±0.05 for 6mM sclareol were obtained at 120h. During the first 50h incubation, H. roseonigra ATCC 20624 grew well on the medium containing 10mM sclareol, and prolonged cultivation times resulted in retarded growth rates (Fig. 2A). Glucose is a carbon source preferred by most microorganisms. Hence, the cell growth on the medium containing 1.33mM glucose was also investigated; this concentration was set to provide the same number of moles of carbon as that lost if 2mM sclareol was completely transformed to sclareol glycol. The cell mass of sclareol cultures reached the same value observed as those in the glucose cultures during the first 12h cultivation. H. roseonigra ATCC 20624 also grew much faster on the medium with glucose than in that with sclareol from 12h to 15h (Fig. 2B). However, further incubation in glucose decreased cell growth, and the OD600 value of H. roseonigra ATCC 20624 on the glucose-containing medium reached 0.53±0.04 at 24h. Interestingly, the growth of H. roseonigra ATCC 20624 continued to increase after 15h of cultivation in BSMS medium. This result indicates the predominant sclareol-tolerant capacity of H. roseonigra ATCC 20624. Cells cultured in BSMG medium and BSMS medium were collected at 12h and used for subsequent proteomic analyses. At this time point, sclareol had not been completely converted to sclareol glycol (Supplementary data, Fig. I).

Fig. 2.

Effects of sclareol stress on the cell growth of Hyphozyma roseonigra ATCC 20624. (A) Effects of sclareol on cell growth. (B) Growth curves of H. roseonigra ATCC 20624 in BSMS and BSMG media. Experiments were performed in triplicate. Error bars represent the standard deviations of the means of three independent experiments.

(0.13MB).
Protein expression alterations in response to sclareol stress

The global proteomic response of H. roseonigra ATCC 20624 to sclareol was profiled by 2-D gel electrophoresis. Glucose cultures were then used as the control group, whereas sclareol cultures were adopted as the experimental group. In three independent experiments, protein abundances were analyzed using the PDQuest software. Protein abundances were significantly altered (p<0.05) in 117 spots (Table S1 and Fig. 3) from the BSMS and BSMG media. This change suggests that sclareol affected the fungus's cellular physiology. Among these spots, 79 up-regulated spots with a >2.0-fold difference in abundance on 2-D gels were collected for further mass spectrometric identification. Seventy spots were successfully recognized, whereas several spots were identified as the same protein after Omics Bean analysis. Finally, 70 protein spots were integrated into 30 proteins (Table 1).

Fig. 3.

Comparative proteomic profiling of H. roseonigra ATCC 20624 exposed to (A) 1.33mM glucose and (B) 2mM sclareol. The differentially expressed proteins were labeled with spot numbers. Experiments were performed in triplicate.

(1.05MB).
Table 1.

Proteins identified by mass spectrometry.

Spot numbera  Protein identification  Fold changeb 
2007, 5722, 6117, 7208, 7209, 7210, 7315, 4720, 5122, 7113, 2017, 2019, 7712, 5719, 7608, 3111, 5817, 6109  Aldehyde dehydrogenase (ALD5)  11.62, 3.37, 33.63, 2.40, 2.91, 2.97, 5.00, 2.85, 3.23, 3.67, 12.66, 9.35, 4.30, 8.81, 2.70, 6.52, 4.82, 3.86 
5806, 6119, 7108  Nicotinic acid transporter (TNA1)  2.91, 8.25, 3.50 
8114  Meiotic recombination protein (REC114)  4.43 
3109  Polyamine oxidase (FMS1)  2.33 
6014  Histone-lysine N-methyltransferase (SET1)  3.76 
6212  Peroxisomal 2,4-dienoyl-CoA reductase (SPS19)  4.89 
6213  Enolase 1 (ENO1)  3.54 
6215, 6306  Transcriptional regulator (URE2)  4.10, 4.16 
6313  Citrate synthase, mitochondrial (CIT1)  3.35 
6319, 7114, 7118, 6320  DNA replication licensing factor (MCM3)  11.15, 4.81, 8.72, 10.35 
7014  Mitochondrial peroxiredoxin Prx1 (PRX1)  6.12 
119, 6118  Actin (ACT1)  2.89, 2.89 
5115  Heat shock protein (SSB2)  2.58 
3003, 122, 313, 121  Fructose-bisphosphate aldolase (FBA1)  3.59, 3.74, 3.16, 3.91 
134  Saccharopepsin (PEP4)  3.16 
1215  78kDa Glucose-regulated protein homolog (KAR2)  6.97 
2218  Structural maintenance of chromosomes protein (SMC4)  2.60 
3816  GTPase (MTG2)  6.52 
4005  Heat shock protein (SSA2)  7.96 
1312, 4409, 4818, 5010, 7110, 6312  Malate dehydrogenase (MDH1)  3.10, 4.58, 6.37, 110.69, 7.80, 6.54 
5117  Flocculation protein (FLO11)  3.40 
7812  Peroxisomal catalase A (CTA1)  3.89 
8112  Acetyl-coenzyme A synthetase 2 (ACS2)  5.39 
2006, 6118, 8314, 1223  NADP-dependent 3-hydroxy acid dehydrogenase (YMR226C)  2.40, 2.40, 16.42, 4.60 
1312  Succinate dehydrogenase (SDH1)  6.19 
2117  Eukaryotic translation initiation factor (HYP2)  3.48 
5716  Glutamine synthetase (GLN1)  6.13 
6009  Cerevisin (PRB1)  17.27 
28, 2712, 6012, 7012, 7013  Phosphoenolpyruvate carboxykinase (PEPCK1)  13.54, 5.45, 15.45, 13.45, 9.74 
7108, 7112  ATP synthase subunit (ATP1)  5.59, 16.42 
a

Consistent with the protein number as shown in Fig. 3.

b

Fold changes in protein expression were calculated using the cells from the BSMG medium as reference.

Discussions

Many of the differentially expressed proteins in Table 1 are common to several categories. Approximately 27% of the differentially expressed proteins belonged to carbon metabolism, followed by 10% belonging to the metabolic pathway and 10% belonging to protein processing in the endoplasmic reticulum. Meanwhile, 3% of the proteins each belonged to the cell cycle, phagosome formation, DNA replication, glutathione metabolism, peroxisome formation, lysine degradation, β-alanine metabolism, antibioticbiosynthesis, and glyoxylate metabolism.

The proteins involved in carbon metabolism (including glycolysis and the citric acid cycle), such as fructose bisphosphate aldolase (FBA), enolase (ENO), citrate synthase (CIT), malate dehydrogenase (MDH), succinate dehydrogenase (SDH) and phosphoenolpyruvate carboxykinase, were all up-regulated in response to sclareol stress. Notably, the fold change of MDH, which catalyzes the conversion of malate and oxaloacetate,13 was the largest, at approximately 111 folds. The observed up-regulation of these proteins indicated that the H. roseonigra ATCC 20624 cells expressed additional enzymes for carbon metabolism to metabolize sclareol more efficiently than usual. Additionally, sclareol stress induced significant changes in the expression of aldehyde dehydrogenases (ALDHs). ALDHs, generally considered as a superfamily of NADP-dependent enzymes and participating in aldehyde metabolism, catalyze the oxidation of exogenous aldehydes into corresponding carboxylic acids.14 The up-regulated ALDHs may protect H. roseonigra ATCC 20624 from damage induced by active aldehydes. Eighteen up-regulated spots were identified as ALDHs, which indicated that ALDHs may play an important role in sclareol metabolism.

Besides the above-mentioned protein narrative, two other proteins, that is, heatshock proteins SSB2 and SSA2, were found up-regulated (2.58 and 7.96 fold increases, respectively). SSB2 and SSA2 are molecular chaperones belonging to the 70kDa heat-shock proteins (Hsp70s). The two enzymes exhibited low intrinsic ATPase activities in the native protein, whereas the isolated ATPase domains manifested much greater ATPase activities.15 Both proteins participate in the transport of polypeptides but function distinctly. Furthermore, proteins acting on cellular processes, such as the nicotinic acid transporter (TNA), meiotic recombination protein (REC), DNA replication licensing factor (MCM), and eukaryotic translation initiation factor (HYP) were up-regulated. This finding indicates that sclareol stress also forced H. roseonigra ATCC 20624 to strengthen its life activities.

H. roseonigra ATCC 20624 also expressed additional structural proteins, such as actins, structural maintenance of chromosomes protein 4, and flocculation protein FLO11, to survive under sclareol stress. Three proteins, that is, 78kDa glucose-regulated protein homolog, GTPase, and ATP synthase subunit α, were also up-regulated. This protein homolog is active as a homodimer and exhibits ATPase activity. Meanwhile, GTPase is required for mitochondrial protein synthesis.16 The up-regulation of these proteins implies that H. roseonigra ATCC 20624 can generate additional ATP or GTP by expressing additional enzymes related to energy metabolism when exposed to sclareol stress.

Conclusions

In summary, the results indicated that proteins involved in carbon metabolism; nitrogen metabolism; and replication, transcription, and translation processes are up-regulated when sclareol is the sole carbon source relative to that when glucose is the sole carbon source. Some of the identified proteins are known to serve heat-inductive functions and play general roles in cross-protecting cells against diverse stress conditions. These cross-protecting proteins may play important roles in the sclareol tolerance of H. roseonigra ATCC 20624. Overall, the present study may lay the fundamentals for further cell engineering to improve sclareol glycol production.

Conflicts of interest

The authors declare no conflicts of interest.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (31200078), the Shandong Province Science and Technology Project (No. ZR2017MC023 and 2015GSF117034), and the Jinan Youth Science and Technology Star Project (No. 201406021).

Appendix A
Supplementary data

The following are the supplementary data to this article:

Relative abundance of the 117 differentially displayed protein spots.

References
[1]
S.P. Bhatia, D. Mcginty, C.S. Letizia, A.M. Api.
Fragrance material review on sclareol.
Food Chem Toxicol, 46 (2008), pp. S270-S274
[2]
F. Li, F. Tao.
Recent advances in biotechnological production of sclareol.
Guangzhou Chem Ind, 43 (2015), pp. 5-7
[3]
L. Wang, H.S. He, H.L. Yu, et al.
Sclareol, a plant diterpene, exhibits potent antiproliferative effects via the induction of apoptosis and mitochondrial membrane potential loss in osteosarcoma cancer cells.
Mol Med Rep, 11 (2015), pp. 4273-4278
[4]
X.W. Pan, L. Han, Y.H. Zhang, D.F. Chen, H.T. Simonsen.
Sclareol production in the moss physcomitrella patens, and observations on growth and terpenoid biosynthesis.
Plant Biotechnol Rep, 9 (2015), pp. 149-159
[5]
G. Aranda, J.Y. Lallemand, A. Mammoumi, R. Azerad.
Microbial hydroxylation of sclareol by Mucor plumbeus.
Tetrahedron Lett, 32 (1991), pp. 1783-1786
[6]
A.F. Barrero, E.J. Alvarez-Manzaneda, J. Altarejos, S. Salido, J.M. Ramos.
Synthesis of ambrox® from (−)-sclareol and (+)-cis-abienol.
Tetrahedron, 49 (1993), pp. 10405-10412
[7]
M. Schalk, L. Pastore, M.A. Mirata, et al.
Toward a biosynthetic route to sclareol and amber odorants.
J Am Chem Soc, 134 (2012), pp. 18900-18903
[8]
Farbood MI, Morris JA, Downey AE. Process for producing diol and lactone and microorganisms capable of same. 1990; US Patent 4,970,163.
[9]
Farbood MI, Willis BJ. Process for producing diol and furan and microorganism capable of same. 1986; US Patent 4,798,799.
[10]
F.A. Trikka, A. Nikolaidis, A. Athanasakoglou, et al.
Iterative carotenogenic screens identify combinations of yeast gene deletions that enhance sclareol production.
Microb Cell Fact, 14 (2015), pp. 60-79
[11]
H. Wang, L. Wang, H. Yang, et al.
Comparative proteomic insights into the lactate responses of halophilic Salinicoccus roseus W12.
Sci Rep, 5 (2015), pp. 13776
[12]
A. Shevchenko, M. Wilm, O. Vorm, M. Mann.
Mass spectrometric sequencing of proteins from silver-stained polyacrylamide gels.
Anal Chem, 68 (1996), pp. 850-858
[13]
J.S. Steffan, L. McAlister-Henn.
Structural and functional effects of mutations altering the subunit interface of mitochondrial malate dehydrogenase.
Arch Biochem Biophys, 287 (1991), pp. 276-282
[14]
S.A. Marchitti, C. Brocker, D. Stagos, V. Vasiliou.
Non-P450 aldehyde oxidizing enzymes: the aldehyde dehydrogenase superfamily.
Expert Opin Drug Metab Toxicol, 4 (2008), pp. 697-720
[15]
P. Lopez-Buesa, C. Pfund, E.A. Craig.
The biochemical properties of the ATPase activity of a 70-kDa heat shock protein (Hsp70) are governed by the C-terminal domains.
Proc Natl Acad Sci U S A, 95 (1998), pp. 15253-15258
[16]
K. Datta, J.L. Fuentes, J.R. Maddock.
The yeast GTPase Mtg2p is required for mitochondrial translation and partially suppresses an rRNA methyltransferase mutant, mrm2.
Mol Biol Cell, 16 (2005), pp. 954-963
Article options