Directed Evolution of Pyruvate Decarboxylase-Negative Saccharomyces cerevisiae, Yielding a C2-Independent, Glucose-Tolerant, and Pyruvate-Hyperproducing Yeast

ABSTRACT The absence of alcoholic fermentation makes pyruvate decarboxylase-negative (Pdc−) strains of Saccharomyces cerevisiae an interesting platform for further metabolic engineering of central metabolism. However, Pdc−S. cerevisiae strains have two growth defects: (i) growth on synthetic medium in glucose-limited chemostat cultures requires the addition of small amounts of ethanol or acetate and (ii) even in the presence of a C2 compound, these strains cannot grow in batch cultures on synthetic medium with glucose. We used two subsequent phenotypic selection strategies to obtain a Pdc− strain without these growth defects. An acetate-independent Pdc− mutant was obtained via (otherwise) glucose-limited chemostat cultivation by progressively lowering the acetate content in the feed. Transcriptome analysis did not reveal the mechanisms behind the C2 independence. Further selection for glucose tolerance in shake flasks resulted in a Pdc−S. cerevisiae mutant (TAM) that could grow in batch cultures (μmax = 0.20 h−1) on synthetic medium, with glucose as the sole carbon source. Although the exact molecular mechanisms underlying the glucose-tolerant phenotype were not resolved, transcriptome analysis of the TAM strain revealed increased transcript levels of many glucose-repressible genes relative to the isogenic wild type in nitrogen-limited chemostat cultures with excess glucose. In pH-controlled aerobic batch cultures, the TAM strain produced large amounts of pyruvate. By repeated glucose feeding, a pyruvate concentration of 135 g liter−1 was obtained, with a specific pyruvate production rate of 6 to 7 mmol g of biomass−1 h−1 during the exponential-growth phase and an overall yield of 0.54 g of pyruvate g of glucose−1.

[1]  Duboc,et al.  An interlaboratory comparison of physiological and genetic properties of four Saccharomyces cerevisiae strains. , 2000, Enzyme and microbial technology.

[2]  A. Novick,et al.  Description of the chemostat. , 1950, Science.

[3]  W. A. Scheffers,et al.  Effect of benzoic acid on metabolic fluxes in yeasts: A continuous‐culture study on the regulation of respiration and alcoholic fermentation , 1992, Yeast.

[4]  Barbara M. Bakker,et al.  Modulating the distribution of fluxes among respiration and fermentation by overexpression of HAP4 in Saccharomyces cerevisiae. , 2001, FEMS yeast research.

[5]  M. Rosenkrantz,et al.  The HAP2,3,4 transcriptional activator is required for derepression of the yeast citrate synthase gene, CIT1 , 1994, Molecular microbiology.

[6]  S. Lun,et al.  Biotechnological production of pyruvic acid , 2001, Applied Microbiology and Biotechnology.

[7]  L. Grivell,et al.  Redirection of the Respiro-Fermentative Flux Distribution in Saccharomyces cerevisiae by Overexpression of the Transcription Factor Hap4p , 2000, Applied and Environmental Microbiology.

[8]  F. Messenguy,et al.  Regulation of compartmentation of amino acid pools in Saccharomyces cerevisiae and its effects on metabolic control. , 1980, European journal of biochemistry.

[9]  U. Sauer Evolutionary engineering of industrially important microbial phenotypes. , 2001, Advances in biochemical engineering/biotechnology.

[10]  R. Miyata,et al.  Breeding of high-pyruvate-producing Torulopsis glabrata with acquired reduced pyruvate decarboxylase. , 1999, Journal of bioscience and bioengineering.

[11]  E. Heinzle,et al.  Quantification of intracellular amino acids in batch cultures of Saccharomyces cerevisiae , 2001, Applied Microbiology and Biotechnology.

[12]  P. Foster Adaptive mutation: implications for evolution , 2000, BioEssays : news and reviews in molecular, cellular and developmental biology.

[13]  Barbara M. Bakker,et al.  Metabolic Engineering of Glycerol Production in Saccharomyces cerevisiae , 2002, Applied and Environmental Microbiology.

[14]  R. Tibshirani,et al.  Significance analysis of microarrays applied to the ionizing radiation response , 2001, Proceedings of the National Academy of Sciences of the United States of America.

[15]  H. Y. Steensma,et al.  Effects of Pyruvate Decarboxylase Overproduction on Flux Distribution at the Pyruvate Branch Point inSaccharomyces cerevisiae , 1998, Applied and Environmental Microbiology.

[16]  M. J. Teixeira de Mattos,et al.  Effects of a hexokinase II deletion on the dynamics of glycolysis in continuous cultures of Saccharomyces cerevisiae. , 2002, FEMS yeast research.

[17]  J. Pronk,et al.  Growth requirements of pyruvate-decarboxylase-negative Saccharomyces cerevisiae. , 1999, FEMS microbiology letters.

[18]  Filip Rolland,et al.  Glucose-sensing and -signalling mechanisms in yeast. , 2002, FEMS yeast research.

[19]  J. Nielsen,et al.  Metabolic Engineering of Saccharomyces cerevisiae , 2000, Microbiology and Molecular Biology Reviews.

[20]  J. Pronk Auxotrophic Yeast Strains in Fundamental and Applied Research , 2002, Applied and Environmental Microbiology.

[21]  S. Hohmann,et al.  Characterization of PDC6, a third structural gene for pyruvate decarboxylase in Saccharomyces cerevisiae , 1991, Journal of bacteriology.

[22]  A. Matin,et al.  Microbial Selection in Continuous Culture , 1977 .

[23]  L. Alberghina,et al.  Development of Metabolically Engineered Saccharomyces cerevisiae Cells for the Production of Lactic Acid , 1995, Biotechnology progress.

[24]  Steven Hahn,et al.  Yeast HAP2 and HAP3 activators both bind to the CYC1 upstream activation site, UAS2, in an interdependent manner , 1987, Cell.

[25]  U von Stockar,et al.  Growth and metabolism of Saccharomyces cerevisiae in chemostat cultures under carbon-, nitrogen-, or carbon- and nitrogen-limiting conditions , 1993, Journal of bacteriology.

[26]  H. Yamada,et al.  The GLY1 gene of Saccharomyces cerevisiae encodes a low-specific L-threonine aldolase that catalyzes cleavage of L-allo-threonine and L-threonine to glycine--expression of the gene in Escherichia coli and purification and characterization of the enzyme. , 1997, European journal of biochemistry.

[27]  F. Gamo,et al.  The mutation DGT1-1 decreases glucose transport and alleviates carbon catabolite repression in Saccharomyces cerevisiae , 1994, Journal of bacteriology.

[28]  R. D. Gietz,et al.  New yeast-Escherichia coli shuttle vectors constructed with in vitro mutagenized yeast genes lacking six-base pair restriction sites. , 1988, Gene.

[29]  J. Diderich,et al.  Physiological Properties of Saccharomyces cerevisiae from Which Hexokinase II Has Been Deleted , 2001, Applied and Environmental Microbiology.

[30]  J. Pronk,et al.  Reproducibility of Oligonucleotide Microarray Transcriptome Analyses , 2002, The Journal of Biological Chemistry.

[31]  J. Nielsen,et al.  Investigation of the impact of MIG1 and MIG2 on the physiology of Saccharomyces cerevisiae. , 1999, Journal of biotechnology.

[32]  D. Hartl,et al.  Selection in chemostats. , 1983, Microbiological reviews.

[33]  J. Gancedo Yeast Carbon Catabolite Repression , 1998, Microbiology and Molecular Biology Reviews.

[34]  J. Pronk,et al.  Pyruvate Metabolism in Saccharomyces cerevisiae , 1996, Yeast.

[35]  H. Tabak,et al.  Molecular characterization of carnitine‐dependent transport of acetyl‐CoA from peroxisomes to mitochondria in Saccharomyces cerevisiae and identification of a plasma membrane carnitine transporter, Agp2p , 1999, The EMBO journal.

[36]  L. Guarente,et al.  Regulation of the yeast CYT1 gene encoding cytochrome c1 by HAP1 and HAP2/3/4. , 1991, Molecular and cellular biology.

[37]  W. A. Scheffers,et al.  Physiology of Saccharomyces cerevisiae in anaerobic glucose-limited chemostat cultures. , 1990, Journal of general microbiology.

[38]  A. Novick,et al.  Experiments with the Chemostat on spontaneous mutations of bacteria. , 1950, Proceedings of the National Academy of Sciences of the United States of America.

[39]  C. Hollenberg,et al.  Heterologous protein production in yeast , 1992, Antonie van Leeuwenhoek.

[40]  Pronk,et al.  Regulation of fermentative capacity and levels of glycolytic enzymes in chemostat cultures of Saccharomyces cerevisiae. , 2000, Enzyme and microbial technology.

[41]  J. Pronk,et al.  The Genome-wide Transcriptional Responses of Saccharomyces cerevisiae Grown on Glucose in Aerobic Chemostat Cultures Limited for Carbon, Nitrogen, Phosphorus, or Sulfur* , 2003, The Journal of Biological Chemistry.

[42]  J. Pronk,et al.  Metabolic responses of pyruvate decarboxylase-negative Saccharomyces cerevisiae to glucose excess , 1997, Applied and environmental microbiology.

[43]  J. Pronk,et al.  Steady-state and transient-state analysis of growth and metabolite production in a Saccharomyces cerevisiae strain with reduced pyruvate-decarboxylase activity. , 1999, Biotechnology and bioengineering.

[44]  J. Pronk,et al.  Pyruvate decarboxylase: An indispensable enzyme for growth of Saccharomyces cerevisiae on glucose , 1996, Yeast.

[45]  R. D. Gietz,et al.  Transformation of yeast by lithium acetate/single-stranded carrier DNA/polyethylene glycol method. , 2002, Methods in enzymology.

[46]  Jack T. Pronk,et al.  Overproduction of Threonine Aldolase Circumvents the Biosynthetic Role of Pyruvate Decarboxylase in Glucose-Limited Chemostat Cultures of Saccharomyces cerevisiae , 2003, Applied and Environmental Microbiology.

[47]  F. Zimmermann,et al.  Genetic analysis of the pyruvate decarboxylase reaction in yeast glycolysis , 1982, Journal of bacteriology.

[48]  J. Collado-Vides,et al.  A web site for the computational analysis of yeast regulatory sequences , 2000, Yeast.