• No results found

Hog1 Mitogen-Activated Protein Kinase Plays Conserved and Distinct Roles in the Osmotolerant Yeast Torulaspora delbrueckii

N/A
N/A
Protected

Academic year: 2020

Share "Hog1 Mitogen-Activated Protein Kinase Plays Conserved and Distinct Roles in the Osmotolerant Yeast Torulaspora delbrueckii"

Copied!
10
0
0

Loading.... (view fulltext now)

Full text

(1)

1535-9778/06/$08.00⫹0 doi:10.1128/EC.00068-06

Copyright © 2006, American Society for Microbiology. All Rights Reserved.

Hog1 Mitogen-Activated Protein Kinase Plays Conserved and

Distinct Roles in the Osmotolerant Yeast

Torulaspora delbrueckii

Marı´a Jose

´ Hernandez-Lopez, Francisca Randez-Gil, and Jose

´ Antonio Prieto*

Department of Biotechnology, Instituto de Agroquı´mica y Tecnologı´a de los Alimentos, Consejo Superior de Investigaciones Cientı´ficas, P.O. Box 73, E-46100 Burjassot, Valencia, Spain

Received 7 March 2006/Accepted 5 June 2006

Torulaspora delbrueckiihas emerged during evolution as one of the most osmotolerant yeasts. However, the

molecular mechanisms underlying this unusual stress resistance are poorly understood. In this study, we have characterized the functional role of the high-osmolarity glycerol (HOG) mitogen-activated protein kinase pathway in mediating the osmotic stress response, among others, in T. delbrueckii. We show that the T.

delbrueckiiHog1p homologue TdHog1p is phosphorylated after cell transfer to NaCl- or sorbitol-containing

medium. However, TdHog1p plays a minor role in tolerance to conditions of moderate osmotic stress, a trait related mainly with the osmotic balance. In consonance with this, the absence of TdHog1p produced only a weak defect in the timing of the osmostress-induced glycerol andGPD1mRNA overaccumulation. Tdhog1

mutants also failed to display aberrant morphology changes in response to osmotic stress. Furthermore, our data indicate that theT. delbrueckiiHOG pathway has evolved to respond to specific environmental conditions and to play a pivotal role in the stress cross-protection mechanism.

The ability ofSaccharomyces cerevisiaeto withstand increas-ing osmotic pressures is critical for its survival in natural habitats and is an essential trait in many biotechnological processes. Variation in ambient osmolarity is a constant phe-nomenon occurring in almost all steps from yeast biomass production to downstream applications in which cells have to grow and perform optimal fermentation under hostile condi-tions (51). However, the success of this industrial microorgan-ism is based on its capacity to transform carbohydrates rapidly into ethanol and CO2 rather than its unusual resistance to

environmental stresses. Yeasts other thanS. cerevisiae, such as

Zygosaccharomyces rouxii,Debaryomyces hansenii, and Torulas-pora delbrueckii, display much higher resistance to adverse conditions and in particular to osmotic stress. This character-istic has made these yeasts potential models to explore and understand the mechanisms underlying osmotic tolerance in eukaryotic cells. However, little is known about the molecular basis of this phenotype, and, in most cases, the characterization of signaling pathways, transcription factors, and gene targets in these non-Saccharomycesspecies remains elusive.

Stress responses inS. cerevisiaehave been widely researched. Yeast cells exposed to an osmostressing environment show a particular transcription profile. Thus, over 250 to 400 genes, covering a wide variety of physiological functions, are up-reg-ulated after different conditions of osmotic shock (16, 24, 54). Stimulated expression of these genes appears to depend mainly on well-characterized molecular signaling pathways: the cyclic AMP-activated protein kinase A pathway (61), the high-osmo-larity glycerol (HOG) pathway (66), one of the five

mitogen-activated protein kinase (MAPK) pathways known inS. cerevisiae

(28), and the calcineurin/Crz1p pathway, which is specifically re-quired for adaptation to high-salt conditions (55).

The HOG pathway consists of two discrete signaling branches composed of a putative osmosensor coupled to a MAPK cascade, some of which can lead to the phosphorylation and activation of the core MAPK Hog1p, the orthologue to mammalian p38 and fission yeast Sty1p stress-activated protein kinase (22). Osmostress-induced phosphorylation of Hog1p triggers its nu-clear accumulation (21, 52) and the later induction of many osmostress-responsive genes (49, 54), among themGPD1(53), the gene encoding the main enzyme that produces the com-patible osmolyte glycerol in S. cerevisiae (2). Although the correlation between Hog1p activation, enhanced glycerol pro-duction, and osmotic resistance inS. cerevisiae is well estab-lished, it is unclear whether the MAPK pathway plays a similar role in other yeasts, particularly in highly osmotolerant species. MAPKs homologous toS. cerevisiaeHog1p have been iden-tified in different yeast species such asSchizosaccharomyces pombe(46),Candida albicans(57),Z. rouxii(37), andD. han-senii(9) as well as in multicellular fungi includingAspergillus nidulans(31),Neurospora crassa(70), and the human pathogen

Cryptococcus neoformans (8). Although only a few of them have been studied in detail, it appears that Hog1p orthologues share conserved but different roles in response to a variety of environmental cues. Whereas inactivation of MAPK inS. cer-evisiaehas dramatic effects on growth under hyperosmotic con-ditions,hog1⌬null mutants fromZ. rouxiican grow as well as the parental strain in the presence of 2 M NaCl (37). Similarly, disruption ofHOG1homologues fromC. neoformans(8) orA. nidulans(39) has a weak effect or no effects on growth at high osmolyte concentrations. Moreover, in some species, the HOG pathway appears to have evolved to respond to additional extracellular stimuli and carry out different cellular roles in a niche-dependent way. Examples include cell-to-cell signaling and virulence in C. albicans (5, 59) or C. neoformans (8), * Corresponding author. Mailing address: Department of

Biotech-nology, Instituto de Agroquı´mica y Tecnologı´a de los Alimentos, Con-sejo Superior de Investigaciones Cientı´ficas, P.O. Box 73, E-46100 Burjassot, Valencia, Spain. Phone: 34-963900022. Fax: 34-963636301. E-mail: prieto@iata.csic.es.

† Supplemental material for this article may be found at http://ec .asm.org/.

1410

on September 8, 2020 by guest

http://ec.asm.org/

(2)

anisms other than the HOG pathway. By contrast, the T. delbrueckiiHOG pathway has undergone a functional special-ization, being used as the central module of the stress cross-protection mechanism in this yeast.

MATERIALS AND METHODS

Yeast strains and culture conditions.T. delbrueckiistrain PYCC5321 (4),S. cerevisiaewild-type strain W303-1A (MAT␣ leu2-3112 ura3-1 trp1-1 his3-1,15 ade2-1 can1-100 GAL SUC2 mal0) (62) and ahog1⌬mutant strain (50) were used throughout this work. TheT. delbrueckiiTdhog1⌬mutant strain was con-structed as described below.Escherichia colistrain DH10B was used as a host for plasmid construction. Yeast cells were cultured at 30°C in defined media: yeast-peptone-dextrose (YPD) medium (1% yeast extract, 2% peptone, 2% glucose) or SD medium (0.17% yeast nitrogen base without amino acids [DIFCO], 0.5% ammonium sulfate, 2% glucose) supplemented with the appropriate auxotrophic requirements (58). Citric acid sensitivity was monitored in malt extract medium as reported previously (42).T. delbrueckiitransformants containing the nourseo-thricin (natMX4)-resistant module were selected on YPD agar plates containing 10 mg/liter of nourseothricin (clonNAT; WERNER Bioagents, Germany).E. coli

was grown in Luria Bertani (LB) medium (1% peptone, 0.5% yeast extract, 0.5% NaCl) supplemented with ampicillin (50 mg/liter).

Stress sensitivity tests.For stress experiments, cells were grown to mid-expo-nential phase at 30°C, collected, and transferred to fresh medium containing NaCl or sorbitol at the indicated concentrations. Plate phenotype experiments were done by diluting the cultures to an optical density at 600 nm (OD600) of 0.3

and streaking the cells onto SD or YPD agar solid medium containing the stressor to be tested. More-detailed sensitivity assays were performed by spotting 10-fold serial dilutions (3␮l) of the cell culture. Unless otherwise indicated, colony growth was inspected after 2 to 4 days of incubation at 30°C.

Strain and plasmid construction.Plasmid pMJH28, carrying a DNA fragment containing theHOG1 gene fromT. delbrueckii(TdHOG1) and the flanking regions around this gene, was isolated from a genomic library (32) by comple-mentation of the osmosensitivity phenotype of ahog1⌬mutant strain inS. cerevisiae(see Results). Multicopy and centromeric plasmids containing the isolated TdHOG1gene were constructed by cloning a 1,541-bp PCR fragment, which includes the complete open reading frame plus 143 bp upstream from the ATG and 140 bp of the 3⬘untranslated region. Amplification was carried out under standard conditions using synthetic oligonucleotides (see Table S1 in the supplemental material) and plasmid pMJH28 as a template. The amplified frag-ment was digested with XbaI/HindIII and ligated into the vectors YEplac195 and YCplac33 (26), which were previously digested with the same set of enzymes, resulting in plasmids YEpTdHOG1 and YCpTdHOG1, respectively.

The TdHOG1disruption cassette containing the nourseothricin resistance

natMX4module (27) was constructed by restriction. First, the PCR-amplified fragment of TdHOG1, obtained as described above, was cloned into the pGEM-T Easy vector (Promega), resulting in plasmid pGEM-TdHOG1. This was digested with EcoRV and BamHI, and the released fragment was replaced by thenatMX4module obtained from plasmid pAG25 (27) by digestion with the same enzymes. Finally, the resulting plasmid, pGEM-TdHOG1-natMX4, was treated with XbaI and HindIII to release the disruption cassette.

The correct gene disruption of TdHOG1was detected by diagnostic PCR using whole yeast cells (35) from isolated colonies and oligonucleotides designed to bind outside or inside the replaced TdHOG1sequences and within the marker modules (see Table S1 in the supplemental material). Yeasts were transformed by the lithium acetate method (36).E. coliwas

antibody yC-20 (catalog no. sc-9079; Santa Cruz Biotechnology, Santa Cruz, CA), respectively, according to the manufacturers’ instructions. The phosphory-lated forms of Slt2p and Fus3p-Kss1p MAPKs were detected with anti-phospho-p44/42 antibody (catalog no. 9101; Cell Signaling) at a 1:1,000 dilution. As a secondary antibody, we used horseradish peroxidase-conjugated goat anti-rabbit antibody(1:2,000, catalog no. 7074; Cell Signaling). Blots were developed using the ECL Western blotting detection kit from Amersham Biosciences.

Glycerol determination.Total glycerol content was determined colorimetri-cally as previously described (50). Values are expressed as micrograms of glycerol per milligram (dry weight) of yeast cells. Growth throughout the experiment was estimated by measuring the OD600(an OD600of 1 equals 0.21 and 0.30 mg [dry

weight] of cells/ml forT. delbrueckiiandS. cerevisiae, respectively).

Sequencing and sequence analysis.DNA sequencing was performed in both strands by the dideoxy-chain termination procedure (56). Similarity searches were performed using the BLAST server at the Munich Information Center for Protein Sequences. We searched for TdHog1p domains by scanning the se-quence against the protein profile databases PROSITE (available at http://www .expasy.org) and Pfam. Multiple sequence alignment was done using the ClustalW program (63).

Nucleotide sequence accession number.The nucleotide sequence for TdHOG1

has been deposited in the GenBank database (available at http://www.ncbi.nlm .nih.gov/GenBank/index.html) under accession number DQ020519.

RESULTS

Cloning and molecular characterization of theT. delbrueckii HOG1gene.Functional complementation for growth of anS.

cerevisiae hog1⌬mutant by aT. delbrueckiigenomic library (32) on 0.5 M NaCl was used to isolate DNA fragments encoding Hog1p. From more than 1,000 transformants, only one plas-mid, pMJH28, which hybridized under heterologous condi-tions with a probe of theS. cerevisiae HOG1gene (data not shown), was able to restore growth of the osmosensitive strain on selective medium. DNA sequencing of the fragment in pMJH28 revealed the presence of a 1,284-bp uninterrupted open reading frame encoding a putative 427-amino-acid pro-tein, TdHog1p. This peptide showed 94 and 88% overall iden-tity withZ. rouxiiZrHog2 and ZrHog1 proteins, respectively, 88% identity withS. cerevisiaeHog1p, and 85% identity withD. hansenii DHog1p. High percentages of identity between TdHog1p and other members of the Hog1/Spc/p38 MAPK subfamily, such asN. crassaOs-2 (83%),C. albicansCaHog1p (78%),A. nidulansHogA (76%),C. neoformansHog1p (76%), and mammalian p38 (47%), were also recorded.

Protein domain analysis showed that TdHog1p contains a typical catalytic protein kinase domain (positions 23 to 302), which is also present in other Hog1p and p38 MAPK homo-logues. Within this region, we identified a protein kinase ATP-binding region (positions 29 to 53) and a consensus Ser/Thr protein kinase active site (positions 140 to 152). Moreover, the

T. delbrueckiiHog1 protein contains a TGY dual

on September 8, 2020 by guest

http://ec.asm.org/

(3)

lation signature (amino acids 174 to 176) characteristic of MAPKs activated by hyperosmolarity (15), a common docking domain (60), and an alanine-rich region at the C terminus at positions 302 to 316 and 364 to 383, respectively, which is also found inS. cerevisiaeHog1p.

To further confirm the identity of the TdHOG1gene prod-uct, we cloned the entire open reading frame into the yeast expression vectors YEplac195 and YCplac33, and the resulting constructs, YEpTdHOG1 and YCpTdHOG1, were trans-formed into theS. cerevisiaeW303-1Ahog1⌬strain and exam-ined for salt tolerance. As shown in Fig. 1A, expression of TdHOG1at a high copy number complemented the osmosen-sitive phenotype of theS. cerevisiae hog1⌬mutant strain on 1 M NaCl. However, the replacement of the Hog1p MAPK by its counterpart TdHog1p in a single copy did not improve the intrinsic salt sensitivity ofS. cerevisiae. Indeed, YCpTdHOG1 and YCpHOG1 transformants, with the latter carrying the gene fromS. cerevisiae, showed similar growth on SD liquid medium supplemented with 1.5 M NaCl (Fig. 1B).

TdHog1p is phosphorylated under osmotic stress. We ex-amined the effect of a shift to NaCl-containing medium on the phosphorylation state of TdHog1p in cells of the wild-type strain. As shown in Fig. 2A, TdHog1p was clearly activated when cells grown in YPD medium (OD600of 1.0) were

sub-jected to mild NaCl stress conditions (0.5 M). The kinetics of phosphorylation were similar to those previously reported for

S. cerevisiaecells exposed to similar NaCl concentrations (65). Indeed, the phosphorylation level reached a maximum within 1 to 5 min of exposure to 0.5 M NaCl and decreased markedly after 10 min (Fig. 2A). As expected, TdHog1p was also sensi-tive to the presence of external sorbitol (Fig. 2B). Thus, both NaCl and sorbitol trigger the activation of the HOG pathway. Effect of TdHOG1on growth under NaCl and sorbitol stress. In order to confirm the functional role of the T. delbrueckii

HOG pathway, we constructed a Tdhog1⌬mutant. Detection of the correct disruption of TdHOG1 was done by PCR and Western blotting. From a total of 10 transformants verified, all showed the pattern of PCR-amplified fragments expected for the correct gene disruption of TdHOG1. In addition, they

lacked a specific band of hybridization with antiserum against total Hog1p (data not shown). Thus, it seems that a single TdHOG1wild-type allele is responsible for the MAPK activity of the HOG pathway in the haploid T. delbrueckii strain PYCC5321.

Figure 2C shows the results of the phenotypic characteriza-tion of Tdhog1⌬ mutant cells on YPD plates supplemented with 0.5 M NaCl or 1 M sorbitol, which both give approxi-mately the same water activity, around 0.98 (53). In contrast to the dramatic osmosensitivity of theS. cerevisiae hog1⌬mutant, the absence of TdHog1p had scarcely any effect on the growth ofT. delbrueckii cells under conditions of moderate osmotic stress. Similar results were obtained when the phenotype was tested using 1 M NaCl (Fig. 2C). However, Tdhog1⌬mutant cells were unable to proliferate at 1.5 M NaCl, while the wild-type strain showed noticeable growth (Fig. 2C).

Osmostress-induced glycerol overaccumulation and GPD1

expression. In S. cerevisiae, hyperosmotic shock triggers the Hog1p-dependent transcriptional induction ofGPD1, the gene for glycerol production (2) and the main osmoregulator in this yeast (11). Likewise, T. delbrueckiioverproduces glycerol in response to osmotic stress (33). Consequently, we analyzed the kinetics of glycerol production in YPD cultures of wild-type and Tdhog1⌬ mutant strains subjected to osmotic stress. As can be seen in Fig. 3A, the absence of MAPK slightly delayed the accumulation of total glycerol in response to 1 M NaCl. Similar results were obtained at different salt concentrations or in the presence of sorbitol (data not shown). However, after 3 h of incubation with 1 M NaCl, the glycerol content in the culture of the Tdhog1⌬mutant increased to about 80% of the wild-type levels (Fig. 3A). Under identical conditions, the total glycerol content ofS. cerevisiae hog1⌬ cultures treated with NaCl was found to be lower, around 30% of the wild-type levels (Fig. 3A, right panel).

We then used Northern blot analysis to examine the expres-sion profile ofGPD1in NaCl-stressedTorulasporacells. Since this gene has not been cloned, total RNA samples were hy-bridized with a probe of its S. cerevisiae counterpart under heterologous conditions. Previous work in our laboratory has FIG. 1. Expression of TdHOG1complements the osmosensitive phenotype of ahog1mutant but does not confer enhanced salt tolerance in

S. cerevisiae. (A) Mid-exponential-phase cultures of W303-1Ahog1⌬mutant cells transformed with the empty plasmid YEplac195 or YEpTdHOG1 were adjusted to an OD600of 0.3, diluted (1 ⫻10⫺3), and spotted (3 ␮l) onto SD plates lacking (control) or containing 1 M NaCl (final

concentration). Plates were inspected after 2 to 4 days at 30°C. (B) YCplac33 (empty plasmid) (䊐), YCpHOG1 (E), and YCpTdHOG1 (F) transformants of theS. cerevisiae hog1mutant strain were grown on SD liquid medium, collected, and transferred (initial OD600of 0.05) to the

same medium containing 1.5 M NaCl, and the OD600was measured at regular intervals. A representative experiment is shown. Independent

experiments revealed identical growth kinetics.

on September 8, 2020 by guest

http://ec.asm.org/

(4)

shown the reliability of this approach (G. Hernandez, F. Randez-Gil, J. A. Prieto, and F. Estruch, unpublished results). Moreover, Southern blot experiments appear to indicate that a singleGPD1allele is present in the genome ofT. delbrueckii

(data not shown). As expected, expression ofGPD1 was in-duced in wild-type cells exposed to 1 M NaCl, with an approx-imately 10-fold induction within 60 min (Fig. 3B). The absence of TdHog1p altered the osmostress-induced expression of

GPD1. However, the effects were not as important as those reported previously for the S. cerevisiae hog1⌬ strain (53). Although maximum induction was shifted to later time points, the mutant was able to accumulate aGPD1mRNA level of about 60% of the wild-type level (Fig. 3B). In addition, the absence of MAPK did not affect GPD1 expression in non-stressed control cells. Similar results were observed for cells exposed to 1 M sorbitol (data not shown).

Osmostress-induced morphology changes in Tdhog1 mu-tants.Exposure of wild-typeS. cerevisiaecells to osmotic stress has no effect on their morphology. By contrast, under the same conditions, ahog1⌬mutant displays morphological alterations (shmoo-like cells) that are characteristic of pheromone-treated cells (12). Figure 4 shows the budded morphology of wild-type cells ofS. cerevisiaestrain W303-1A exposed to 1 M sorbitol or NaCl, in contrast to the shmoos formed in theS. cerevisiae hog1⌬ mutant. However, wild-type and Tdhog1⌬ mutant strains ofT. delbrueckiidisplayed normal morphologies in both the presence and absence of 1 M sorbitol. Only exposure to 1 M NaCl (Fig. 4) or higher concentrations (data not shown) led to a clear morphological response, although shmoo-like cells were not observed in any case. Instead, cells showed a clear

inability to separate normally. Thus, the response appeared to depend on the chemical used. However, blocking in cell sepa-ration was also evident after exposure to 2 M sorbitol (data not shown).

Cross talk between the T. delbrueckii HOG pathway and parallel MAPK pathways under hyperosmotic conditions.The formation of shmoo-like cells inS. cerevisiae Hog1p-deficient strains occurs by the inappropriate activation of two MAPKs (17, 30, 48), Kss1p, in the MAPK of the filamentation/invasion pathway (23), and Fus3p in the pheromone response pathway (19). Therefore, we investigated the phosphorylation states of these MAPKs in wild-type and Tdhog1⌬cells exposed to high osmolarity. Protein extracts were examined with Western anal-ysis by using a monoclonal phospho-p44/42 MAPK anti-body, which recognizes the TEY activation sequence found in mammalian p42 and p44 MAPKs (Erk1/Erk2) and yeast MAPKs Kss1p and Fus3p. As shown in Fig. 5, a single band of the expected mobility for eitherT. delbrueckiiKss1p or Fus3p MAPK (⬃41 kDa) appeared in protein extracts from Tdhog1

cells 30 min after osmotic shock. In contrast, there was no evidence of this effect in control cells (Fig. 5).

Interestingly, inspection of the Western blot revealed the presence of an additional ⬃61-kDa strong immunoreactive band in both wild-type and Tdhog1⌬mutant cells (Fig. 5). This band might correspond to the phosphorylated form of Slt2p, the MAPK in the protein kinase C (PKC) pathway (28), which also contains the same TEY activation loop as Fus3p and Ssk1p and is therefore recognized by the antibody. Notably, we also observed that Slt2p exhibited sustained phosphorylation upon exposure to osmotic stress (Fig. 5). Indeed, only the 30-FIG. 2. TdHog1p phosphorylation and growth of the Tdhog1⌬ mutant under osmotic stress conditions. (A) Cells of the T. delbrueckii

PYCC5321 wild-type (wt) strain were grown in YPD medium at 30°C until early exponential phase (OD600of 0.4 to 0.6) and then transferred to

NaCl-containing medium (0.5 M). At the indicated times, cells were harvested and processed, and the crude protein extracts were analyzed by Western blot. Nonstressed cells (time zero) were used as a control. Total protein extracts were analyzed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and blotting with anti-phospho-p38 antibody (P-TdHog1p). The membrane was then reblotted with anti-Hog1p antibody as a loading control (TdHog1p). (B) Western blot analysis of YPD-grown cells of the wild-type strain transferred to YPD plus sorbitol at a 0.5, 1, or 1.5 M final concentration. After 5 min, cells were harvested and analyzed for phospho-TdHog1p levels as described above. (C) Growth of wild-type andhog1⌬mutant strains ofS. cerevisiaeW303-1A andT. delbrueckiiPYCC5321 (Tdhog1⌬) was examined on solid YPD medium (control) or YPD medium containing sorbitol or NaCl at the indicated concentrations. Cells pregrown in YPD medium were diluted, spotted, and incubated as described in the legend of Fig. 1A. Representative experiments are shown.

on September 8, 2020 by guest

http://ec.asm.org/

(5)

to 60-min samples showed a significant reduction of the phos-pho-Slt2p signal. In addition, a loss of Slt2p phosphorylation was not observed in the Tdhog1⌬ strain (Fig. 5). This is in sharp contrast to the situation reported previously forS. cer-evisiae, where Slt2p is rapidly dephosphorylated (within 1 min), in a Hog1p-independent way, upon osmotic stress (68). Hence, it seems that a functionalT. delbrueckiiHOG pathway is es-sential for the proper deactivation of the PKC signal in re-sponse to hyperosmolarity.

The functional role of T. delbrueckiiHog1p in response to several types of stress.Functions other than osmostress pro-tection have been identified for HOG homologue pathways in some yeast species. Based on this, we analyzed the need for TdHog1p in the response ofT. delbrueckiito diverse stimuli (Fig. 6A). Clearly, the MAPK was essential for tolerance to methylglyoxal and citric acid. However, it appeared to be dis-pensable for growth at a high temperature, 34°C (Fig. 6A). We also noted that theT. delbrueckiiMAPK was not essential for the response to H2O2. To further confirm this point, we

ana-lyzed the kinetics of TdHog1p phosphorylation in T. del-brueckiiwild-type cells following oxidative stress. As shown in Fig. 6B, the level of phospho-TdHog1p did not vary after exposure of yeast cells to 4 mM H2O2. Similar results were

FIG. 3. Glycerol production andGPD1induction in wild-type and Tdhog1⌬mutant cells ofT. delbrueckii. (A) Total glycerol levels after hyperosmotic shock in cultures of T. delbrueckii (left graph) PYCC5321 (wild type [wt]) (black bars) and Tdhog1⌬mutant (gray bars) strains andS. cerevisiae(right graph) W303-1A wild-type (black bars) andhog1⌬mutant (gray bars) strains. Cells were grown in YPD medium to mid-exponential phase (OD600of 1.5), collected, and

trans-ferred to the same medium containing 1 M NaCl. At the indicated times, aliquots of the cell cultures were withdrawn and analyzed for glycerol content as described in Materials and Methods. Data repre-sent the mean values of at least three independent experiments. Errors were calculated by using the formula (1.96⫻SD)/公n, wherenis the number of measurements (SD, standard deviation). dw, dry weight. (B) Cells ofT. delbrueckiiwild-type (■) and Tdhog1⌬mutant (䊐) strains grown in YPD medium were transferred to medium containing 1 M NaCl. Samples were taken at the indicated times and analyzed by Northern blotting as described in Materials and Methods. Filters were probed forGPD1mRNA. The graph in panel B represents the quan-tification of the mRNA levels ofGPD1relative to those ofACT1.

FIG. 4. Effect of TdHog1p on the osmostress-induced morphological changes ofT. delbrueckii. Wild-type (wt) strains ofS. cerevisiae(W303-1A) andT. delbrueckii(PYCC5321) and the correspondinghog1mutantshog1

and Tdhog1⌬were pregrown in YPD medium, collected, and then trans-ferred (initial OD600of 0.05) to the same medium (control) or YPD medium

plus 1 M NaCl or sorbitol. Microphotographs were taken using an optical microscope after 16 h of culture. A representative cluster of cells is shown.

FIG. 5. TheT. delbrueckiiHOG pathway modulates signaling through parallel MAPK pathways in response to osmotic stress. Cells of wild-type

T. delbrueckiistrain PYCC5321 (wt) and the isogenic Tdhog1mutant (Tdhog1⌬) were cultured in YPD medium to mid-log phase at 30°C and transferred to the same medium containing 1 M NaCl. At the indicated times, cells were collected and processed, and protein extracts were ana-lyzed by Western blotting with anti-phospho-p44/42 MAPK antibody, which allows the visualization of the phosphorylated (P) form of Slt2p, Kss1p, and Fus3p MAPKs. Molecular weight (Mw) markers are shown (in thousands). PutativeT. delbrueckiiproteins Kss1 and Fus3 are expected to comigrate under the conditions applied. A representative experiment is shown.

on September 8, 2020 by guest

http://ec.asm.org/

(6)

observed for cells shifted to 34°C (Fig. 6B). Hence, the T. delbrueckiiHOG pathway appears to be necessary under spe-cific environmental conditions.

Stress cross-protection in T. delbrueckii. S. cerevisiae cells exposed to nonlethal doses of one type of stress develop tol-erance to higher doses of the same stress as well as to unrelated stresses; these phenomena are described as induced tolerance and cross-protection, respectively (20, 43). It is commonly ac-cepted that yeasts other thanSaccharomyces display induced tolerance. However, the existence of cross-protection mecha-nisms is not always evident and has frequently been the subject of controversy (18). Therefore, we were interested in investi-gating this phenomenon in the osmotolerant yeast T. del-brueckii.

As shown in Fig. 7A, cells of the T. delbrueckii wild-type strain were unable to grow at 37°C on YPD plates lacking (control) or containing 0.5 M sorbitol. However, this pheno-type was suppressed when cells were challenged with 1 M (Fig. 7A) or 1.5 M (data not shown) sorbitol. Nevertheless, we reasoned that this phenomenon does not necessarily imply previous adaptation of the cells by osmotic stress. Reduction of water activity by sorbitol or salts is indeed an effective way to reduce thermal death, as demonstrated for osmotolerant yeasts likeD. hansenii(3). Consequently, we looked into how TdHog1p is involved in this phenomenon. As can be seen in Fig. 7A, the thermal protection afforded by 1 M sorbitol was retained in the Tdhog1⌬strain. Nevertheless, growth of mutant

cells was clearly impaired at 37°C compared to the wild type, whereas at 30°C, the differences were less pronounced. Hence, adaptation to thermal stress appears to be dependent upon TdHog1p. Presumably, growth at high temperatures is also improved by water activity reduction mechanisms, being more pronounced at high osmolyte concentrations.

To investigate the role of TdHog1p in the above-described findings in greater depth, we first tested the levels of phospho-TdHog1p in cells exposed to increased temperatures plus high osmolarity (Fig. 7B). Exposure of yeast cells at a temperature of 37°C induced a modest increase in the level of phospho-TdHog1p, around 1.5- to 2-fold over basal levels, compared with that observed after exposure to 1 M sorbitol, about nine-fold (Fig. 2B). Moreover, overphosphorylation of TdHog1p took place late and was only evident within 15 to 30 min after the thermal transfer, a temporary pattern often linked to an indirect mechanism of activation (34). In contrast, cells sub-jected to highly osmotic conditions at 37°C showed the same kinetics of TdHog1p phosphorylation that are characteristic for hyperosmolarity (Fig. 2A and 7B).

Next, we investigated the thermal protection conferred by pretreatment ofT. delbrueckiicells with sorbitol. Cells grown in YPD medium were exposed to 1 M sorbitol at 30°C for 3 h and then transferred to fresh prewarmed (37°C) YPD medium lacking sorbitol (Fig. 7C). As can be seen in Fig. 7C, pretreat-ment of wild-type cells with osmotic stress afforded higher growth ability at 37°C. However, this effect was lost in Tdhog1

FIG. 6. TdHog1p is essential for the response to different stresses. (A) Exponentially growing cells ofT. delbrueckiiwild-type strain PYCC5321 (wt) and the corresponding Tdhog1mutant were spotted onto malt extract (ME), SD, or YPD plates containing 200 mM citric acid, 20 mM methylglyoxal (MG), and 4 mM H2O2, respectively. Control plates lacking the above-cited compounds were tested in parallel. Growth at high

temperatures was inspected on YPD plates incubated at 34°C. For the rest, plates were incubated at 30°C for 2 to 4 days. (B) Cells of wild-type

T. delbrueckiistrain PYCC5321 were grown in YPD medium at 30°C until early exponential phase (OD600of 0.4 to 0.6) and then shifted to 34°C

or transferred to medium containing 4 mM H2O2. At the indicated times, cells were harvested and processed, and the crude protein extracts were

assayed for TdHog1p phosphorylation (P-TdHog1) levels as described in the legend of Fig. 2A. In all cases, representative experiments are shown.

on September 8, 2020 by guest

http://ec.asm.org/

(7)

cells. Hence, cross-protection exists inTorulasporaand is me-diated by the HOG pathway.

DISCUSSION

Phylogenetically, T. delbrueckiiandS. cerevisiaeare closely related, as previously demonstrated by sequence comparisons (10, 32). Despite this genealogical relationship,T. delbrueckiiis able to survive and proliferate at high osmolyte concentrations, a property not shared by S. cerevisiae (33). As a result, this nonconventional yeast is frequently found in different habitats compared to those ofS. cerevisiae, including cane molasses, syrups, honey, and foods with a high sugar content (41). Hence, they would be expected to differ in the regulatory mechanisms and molecular targets that enable adaptability to high-osmo-larity environments.

These differences were first evident in our analysis of the functional response of the T. delbrueckii HOG pathway to osmotic stress. Although the MAPK TdHog1p was rapidly phosphorylated after transfer to NaCl- or sorbitol-containing medium, this was not essential for the growth ofT. delbrueckii

cells under moderate stress conditions, 0.5 M NaCl or 1 M sorbitol. Only when severe stress conditions were tested, for example, 1.5 M NaCl, was the need for a functional HOG pathway clearly evident. This result fully agrees with previous observations ofZ. rouxii hog1⌬mutant cells, which show NaCl sensitivity only at concentrations above 2 M (37). Therefore, we predicted that this fact could reflect a distinct role of the HOG pathway in the osmoadaptation mechanisms employed byT. delbrueckiiandS. cerevisiae. Consistent with this, neither osmostress-induced glycerol accumulation nor the expression ofGPD1 was strongly affected in Tdhog1⌬ mutant cells ex-posed to moderate hyperosmolarity. This is in sharp contrast with the situation in S. cerevisiae, where Hog1p is the main regulator of these responses (2, 53). Nevertheless, we noted that glycerol production and GPD1 mRNA levels were up-regulated upon osmotic shock. Therefore, a sudden change in the water activity of the environment is perceived byT. del-brueckiias a stressful condition. However,T. delbrueckii, likeS. cerevisiae, must be frequently exposed to significant osmolarity fluctuations in natural habitats. Therefore, adaptation to these FIG. 7. The HOG pathway mediates stress cross-protection inT. delbrueckii. (A) YPD-grown cells ofT. delbrueckiiPYCC5321 wild-type (wt) and Tdhog1⌬mutant strains were spotted onto YPD plates (control) and YPD plates containing 0.5 or 1 M sorbitol. Plates were incubated at 30 or 37°C for 2 to 4 days. (B) Cells of theT. delbrueckiiwild-type strain were grown in YPD medium at 30°C and then subjected to heat stress alone (37°C) or in combination with high osmolarity, 1 M sorbitol (37°C⫹1 M sorbitol). At the indicated times, cells were harvested and processed, and the crude protein extracts were assayed for TdHog1p phosphorylation (P-TdHog1) levels as described in the legend of Fig. 2A. (C) Expo-nentially growing cells of the wild-type (squares) andTdhog1⌬mutant (circles) strains in YPD medium were either left unstressed (open symbols) or exposed to 1 M sorbitol (closed symbols) at 30°C. After 3 h, cells were washed and shifted to prewarmed (37°C) YPD medium lacking sorbitol, and the OD600was measured at regular intervals. In all cases, representative experiments are shown.

on September 8, 2020 by guest

http://ec.asm.org/

(8)

cerevisiae hog1⌬mutants, resembling that of cells exposed to pheromones, shmoos, or pear-shaped cells (13). As we dem-onstrated, osmotic stress induced the overphosphorylation of Fus3p and Kss1p, the MAPKs in the pheromone response and filamentation/invasion pathways, respectively, in the TdHog1p-deficient strain. Thus, the MAPK TdHog1p prevents osmolar-ity-induced cross talk between the HOG and parallel MAPK pathways inT. delbrueckii. However, shmoos were not formed by Tdhog1⌬mutant cells exposed to a high level of sorbitol or NaCl. Therefore, S. cerevisiae and T. delbrueckii appear to control a different set of genes by the Fus3p-Kss1p MAPKs. It is also possible thatT. delbrueckiihas evolved additional mech-anisms to ensure the specificity of the osmostress signal. Nev-ertheless, further work is required to identify upstream ele-ments of the T. delbrueckii HOG pathway and clarify the activation mechanisms operating in this yeast.

Regarding the morphological effects of osmotic stress, we also noted that Tdhog1⌬mutant cells remain attached after budding at high osmolyte concentrations, a phenotype that was missing in the wild-type strain. This apparent cell division de-fect led to large branched aggregates, similar to those observed for chitinase-negativeS. cerevisiaestrains (40). This phenotype has also been reported forhog1⌬mutant cells ofC. albicans

subjected to 1 M NaCl, suggesting a link between cell wall metabolism and the activity of CaHog1p in this pathogenic yeast (5). Such a relationship has also been suggested forS. cerevisiae, since components of the HOG pathway appear to be involved in cell wall maintenance (6).

A connection between cell wall metabolism and theT. del-brueckiiHOG pathway was evidenced by the repressing effects that this pathway exerts on the activity of the PKC pathway following osmotic stress. The PKC signaling pathway is in-duced during budding and mating (69) as well as in response to environmental conditions such as high temperature, hypoos-motic stress, and cell wall-damaging conditions (38, 44). As we have shown, overphosphorylation of Slt2p, the MAPK of the PKC pathway, was maintained in Tdhog1⌬mutant cells after a shift to conditions of high osmolarity. This lack of repression might thus account for the cellular aggregation displayed by the Tdhog1⌬mutant strain upon osmotic stress. On the other hand, the level of phospho-Slt2p was notably high in cells grown at 30°C. This observation indicates thatT. delbrueckii

perceives stressful temperature fluctuations that do not stim-ulate a stress response inS. cerevisiae. Moreover, the kinetics of Slt2p dephosphorylation after osmotic stress were unusually slow inT. delbrueckiicompared to that exhibited byS. cerevi-siae(68). Thus, this result suggests that there has been a

di-provides protection against this stressful condition inS. cerevi-siae(29) as well as in other yeasts and fungi species such asC. albicans(7),Schizosaccharomyces pombe(14), andC. neofor-mans(8). Although we have no obvious explanation for this result, it could reflect a wider ability of this yeast to cope with high levels of reactive oxygen species. Consistent with this, we found thatT. delbrueckiiwas able to grow at high relative levels of H2O2(4 mM) compared with other yeasts and in particular

withS. cerevisiae. Thus, the T. delbrueckiiHOG pathway ap-pears to have evolved not to respond to oxidative stress.

In the present work, we have also observed stress cross-protection in T. delbrueckii, suggesting that a general stress response (20) exists in this yeast. InS. cerevisiae, the general stress response is mediated by a common pathway, the cyclic AMP-protein kinase A pathway (61), which controls the activ-ity and nuclear localization of the transcriptional factors

MSN2/MSN4(45). In addition,S. cerevisiaeuses strategies of coinduction of defense mechanisms. Instead of a common pathway, different stresses control a common set of genes via different signaling pathways and transcription factors (16, 24). Far from the classical concept of a general stress response, we observed that growth ofT. delbrueckiiat 37°C, a temperature that inhibits the proliferation of this yeast, was dependent on the activity of TdHog1p coupled with the thermal protection provided by sorbitol. Furthermore, we find that pretreatment with sorbitol provides a growth advantage at 37°C for wild-type cells but not for Tdhog1⌬mutant cells. Therefore, stress cross-protection exists in T. delbrueckii, but this is mediated by TdHog1p. Similarly, Smith et al. (59) previously reported Hog1p-dependent stress cross-protection inC. albicans, a re-sult that fits well with the lack of a functional role of homo-logues ofS. cerevisiae MSN2/MSN4in this yeast (47).

Overall, the results presented in this study emphasize the divergence of a classical stress signaling pathway between dif-ferent yeasts. Our current knowledge indicates that the HOG pathway has evolved in different yeasts in a niche-specific way (8, 59). We have also demonstrated that changes have taken place in the Hog1p-GPD1relationship during the evolutionary divergence ofS. cerevisiaeandT. delbrueckii. While it is easy to imagine the different mechanisms through which gene expres-sion regulation evolves (25), it is much more difficult to under-stand how these events determine the adaptability of yeasts to changing environments.

ACKNOWLEDGMENTS

We thank A. Blasco and J. Panadero for technical assistance and J. Aguilera and Y.-S. Bahn for critical reading of the manuscript.

on September 8, 2020 by guest

http://ec.asm.org/

(9)

This research was funded by the Comisio´n Interministerial de Cien-cia y Tecnologı´a projects (PACTI, COO1999AX173; AGL2001-1203; and AGL2004-0462) from the Ministry of Education and Science of Spain. M.J.H.-L. was supported by a CSIC-EPO fellowship.

REFERENCES

1.Aguilera, J., S. Rodriguez-Vargas, and J. A. Prieto.2005. The HOG MAP kinase pathway is required for the induction of methylglyoxal-responsive genes and determines methylglyoxal resistance inSaccharomyces cerevisiae. Mol. Microbiol.56:228–239.

2.Albertyn, J., S. Hohmann, J. M. Thevelein, and B. A. Prior.1994.GPD1, which encodes glycerol-3-phosphate dehydrogenase, is essential for growth under osmotic stress inSaccharomyces cerevisiae, and its expression is regu-lated by the high-osmolarity glycerol response pathway. Mol. Cell. Biol.

14:4135–4144.

3.Almagro, A., C. Prista, S. Castro, C. Quintas, A. Madeira-Lopes, J. Ramos, and M. C. Loureiro-Dias.2000. Effects of salts onDebaryomyces hanseniiand

Saccharomyces cerevisiae under stress conditions. Int. J. Food Microbiol.

56:191–197.

4.Almeida, M. J., and C. Pais.1996. Leavening ability and freeze tolerance of yeasts isolated from traditional corn and rye bread doughs. Appl. Environ. Microbiol.62:4401–4404.

5.Alonso-Monge, R., F. Navarro-Garcia, G. Molero, R. Diez-Orejas, M. Gustin, J. Pla, M. Sanchez, and C. Nombela. 1999. Role of the mitogen-activated protein kinase Hog1p in morphogenesis and virulence ofCandida albicans. J. Bacteriol.181:3058–3068.

6.Alonso-Monge, R., E. Real, I. Wojda, J. P. Bebelman, W. H. Mager, and M. Siderius.2001. Hyperosmotic stress response and regulation of cell wall integrity inSaccharomyces cerevisiaeshare common functional aspects. Mol. Microbiol.41:717–730.

7.Alonso-Monge, R., F. Navarro-Garcia, E. Roman, A. I. Negredo, B. Eximan, C. Nombela, and J. Pla.2003. The Hog1 mitogen-activated protein kinase is essential in the oxidative stress response and chlamydospore formation in

Candida albicans. Eukaryot. Cell2:351–361.

8.Bahn, Y. S., K. Kojima, G. M. Cox, and J. Heitman.2005. Specialization of the HOG pathway and its impact on differentiation and virulence of Cryp-tococcus neoformans. Mol. Biol. Cell16:2285–2300.

9.Bansal, P. K., and A. K. Mondal.2000. Isolation and sequence of theHOG1

homologue fromDebaryomyces hanseniiby complementation of thehog1

strain ofSaccharomyces cerevisiae. Yeast16:81–88.

10.Belloch, C., A. Querol, M. D. Garcı´a, and E. Barrio.2000. Phylogeny of the genusKluyveromycesinferred from the mitochondrial cytochrome-c oxidase II gene. Int. J. Syst. Evol. Microbiol.50:405–416.

11.Blomberg, A., and L. Adler.1992. Physiology of osmotolerance in fungi. Adv. Microb. Physiol.33:145–212.

12.Brewster, J. L., T. de Valoir, N. D. Dwyer, E. Winter, and M. C. Gustin.1993. An osmosensing signal transduction pathway in yeast. Science259:1760– 1763.

13.Brewster, J. L., and M. C. Gustin.1994. Positioning of cell growth and division after osmotic stress requires a MAP kinase pathway. Yeast10:425– 439.

14.Buck, V., J. Quinn, T. Soto Pino, H. Martin, J. Saldanha, K. Makino, B. A. Morgan, and J. B. Millar.2001. Peroxide sensors for the fission yeast stress-activated mitogen-stress-activated protein kinase pathway. Mol. Biol. Cell12:407– 419.

15.Cano, E., and L. C. Mahadevan.1995. Parallel signal processing among mammalian MAPKs. Trends Biochem. Sci.20:117–122.

16.Causton, H. C., B. Ren, S. S. Koh, C. T. Harbison, E. Kanin, E. G. Jennings, T. I. Lee, H. L. True, E. S. Lander, and R. A. Young.2001. Remodeling of yeast genome expression in response to environmental changes. Mol. Biol. Cell12:323–337.

17.Davenport, K. D., K. E. Williams, B. D. Ullmann, and M. C. Gustin.1999. Activation of theSaccharomyces cerevisiaefilamentation/invasion pathway by osmotic stress in high-osmolarity glycogen pathway mutants. Genetics153:

1091–1103.

18.Enjalbert, B., A. Nantel, and M. Whiteway.2003. Stress-induced gene ex-pression inCandida albicans: absence of a general stress response. Mol. Biol. Cell14:1460–1467.

19.Errede, B., A. Gartner, Z. Zhou, K. Nasmyth, and G. Ammerer.1993. MAP kinase-related FUS3 fromS. cerevisiaeis activated by STE7 in vitro. Nature

362:261–264.

20.Estruch, F. 2000. Stress-controlled transcription factors, stress-induced genes and stress tolerance in budding yeast. FEMS Microbiol. Rev.24:469– 486.

21.Ferrigno, P., F. Posas, D. Koepp, H. Saito, and P. A. Silver.1998. Regulated nucleo/cytoplasmatic exchange of HOG1 MAPK requires the importin␤

homologs NMD5 and XPO1. EMBO J.17:5606–5614.

22.Gacto, M., T. Soto, T. Vicente-Soler, T. G. Villa, and J. Cansado.2003. Learning from yeasts: intracellular sensing of stress conditions. Int. Micro-biol.6:211–219.

23.Gartner, A., K. Nasmyth, and G. Ammerer.1992. Signal transduction in

Saccharomyces cerevisiaerequires tyrosine and threonine phosphorylation of FUS3 and KSS1. Genes Dev.6:1280–1292.

24.Gasch, A. P., P. T. Spellman, C. M. Kao, O. Carmel-Harel, M. B. Eisen, G. Storz, D. Botstein, and P. O. Brown.2000. Genomic expression programs in the response of yeast cells to environmental changes. Mol. Biol. Cell11:

4241–4257.

25.Gasch, A. P., A. M. Moses, D. Y. Chiang, H. B. Fraser, M. Berardini, and M. B. Eisen.2004. Conservation and evolution of cis-regulatory systems in ascomycete fungi. PLoS Biol.2:e398.

26.Gietz, R. D., and A. Sugino.1988. New yeast-Escherichia colishuttle vectors constructed with in vitro mutagenized yeast genes lacking six-base pair re-striction sites. Gene74:527–534.

27.Goldstein, A. L., and J. H. McCusker.1999. Three new dominant drug resistance cassettes for gene disruption inSaccharomyces cerevisiae. Yeast

15:1541–1553.

28.Gustin, M. C., J. Albertyn, M. Alexander, and K. Davenport.1998. MAP kinase pathways in the yeastSaccharomyces cerevisiae. Microbiol. Mol. Biol. Rev.62:1264–1300.

29.Haghnazari, E., and W. D. Heyer.2004. The Hog1 MAP kinase pathway and the Mec1 DNA damage checkpoint pathway independently control the cel-lular responses to hydrogen peroxide. DNA Repair3:769–776.

30.Hall, J. P., V. Cherkasova, E. Elion, M. C. Gustin, and E. Winter.1996. The osmoregulatory pathway represses mating pathway activity inSaccharomyces cerevisiae: isolation of aFUS3mutant that is insensitive to the repression mechanism. Mol. Cell. Biol.16:6715–6723.

31.Han, K. H., and R. A. Prade.2002. Osmotic stress-coupled maintenance of polar growth inAspergillus nidulans. Mol. Microbiol.43:1065–1078. 32.Hernandez-Lopez, M. J., J. A. Prieto, and F. Randez-Gil.2002. Isolation and

characterization of the geneURA3 encoding the orotidine-5⬘-phosphate decarboxylase fromTorulaspora delbrueckii. Yeast19:1431–1435. 33.Hernandez-Lopez, M. J., J. A. Prieto, and F. Randez-Gil.2003.

Osmotoler-ance and leavening ability in sweet and frozen sweet dough. Comparative analysis betweenTorulaspora delbrueckiiandSaccharomyces cerevisiaebaker’s yeast strains. Antonie Leeuwenhoek84:125–134.

34.Hohmann, S.2002. Osmotic stress signaling and osmoadaptation in yeasts. Microbiol. Mol. Biol. Rev.66:300–372.

35.Huxley, C., E. D. Green, and I. Dunham.1990. Rapid assessment ofS. cerevisiaemating type by PCR. Trends Genet.6:236.

36.Ito, H., Y. Fukuda, K. Murata, and A. Kimura.1983. Transformation of intact yeast cells treated with alkali cations. J. Bacteriol.153:163–168. 37.Iwaki, T., Y. Tamai, and Y. Watanabe.1999. Two putative MAP kinase

genes,ZrHOG1andZrHOG2, cloned from the salt-tolerant yeast Zygosac-charomyces rouxiiare functionally homologous to theSaccharomyces cerevi-siae HOG1gene. Microbiology145:241–248.

38.Kamada, Y., U. S. Jung, J. Piotrowski, and D. E. Levin.1995. The protein kinase C-activated MAP kinase pathway ofSaccharomyces cerevisiae medi-ates a novel aspect of the heat shock response. Genes Dev.9:1559–1571. 39.Kawasaki, L., O. Sanchez, K. Shiozaki, and J. Aguirre.2002. SakA MAP

kinase is involved in stress signal transduction, sexual development and spore viability inAspergillus nidulans. Mol. Microbiol.45:1153–1163.

40.Kuranda, M. J., and P. W. Robbins.1991. Chitinase is required for cell separation during growth ofSaccharomyces cerevisiae. J. Biol. Chem.266:

19758–19767.

41.Kurtzman, C. P.1998.TorulasporaLindner, p. 404–408.InC. P. Kurtzman and J. W. Fell (ed.), The yeasts, a taxonomic study, 4th ed. Elsevier Science B.V., Amsterdam, The Netherlands.

42.Lawrence, C. L., C. H. Botting, R. Antrobus, and P. J. Coote.2004. Evidence of a new role for the high-osmolarity glycerol mitogen-activated protein kinase pathway in yeast: regulating adaptation to citric acid stress. Mol. Cell. Biol.24:3307–3323.

43.Lewis, J. G., R. P. Learmonth, and K. Watson.1995. Induction of heat, freezing and salt tolerance by heat and salt shock inSaccharomyces cerevisiae. Microbiology141:687–694.

44.Martin, H., J. M. Rodriguez-Pachon, C. Ruiz, C. Nombela, and M. Molina.

2000. Regulatory mechanisms for modulation of signaling through the cell integrity Slt2-mediated pathway inSaccharomyces cerevisiae. J. Biol. Chem.

275:1511–1519.

45.Martinez-Pastor, M. T., G. Marchler, C. Schuller, A. Marchler-Bauer, H. Ruis, and F. Estruch.1996. TheSaccharomyces cerevisiaezinc finger proteins Msn2p and Msn4p are required for transcriptional induction through the stress response element (STRE). EMBO J.15:2227–2235.

46.Millar, J. B., V. Buck, and M. G. Wilkinson.1995. Pyp1 and Pyp2 PTPases dephosphorylate an osmosensing MAP kinase controlling cell size at division in fission yeast. Genes Dev.9:2117–2130.

47.Nicholls, S., M. Straffon, B. Enjalbert, A. Nantel, S. Macaskill, M. Whiteway, and A. J. Brown.2004. Msn2- and Msn4-like transcription factors play no obvious roles in the stress responses of the fungal pathogenCandida albi-cans. Eukaryot. Cell3:1111–1123.

48.O’Rourke, S. M., and I. Herskowitz.1998. The Hog1 MAPK prevents cross talk between the HOG and pheromone response MAPK pathways in Sac-charomyces cerevisiae. Genes Dev.12:2874–2886.

49.O’Rourke, S. M., and I. Herskowitz.2004. Unique and redundant roles for

on September 8, 2020 by guest

http://ec.asm.org/

(10)

715–727.

54.Rep, M., M. Krantz, J. M. Thevelein, and S. Hohmann.2000. The transcrip-tional response ofSaccharomyces cerevisiaeto osmotic shock Hot1p and Msn2p/Msn4p are required for the induction of subsets of high osmolarity glycerol pathway-dependent genes. J. Biol. Chem.275:8290–8300. 55.Rusnak, F., and P. Mertz.2000. Calcineurin: form and function. Physiol.

Rev.80:1483–1521.

56.Sanger, F., S. Nicklen, and A. R. Coulson.1977. DNA sequencing with chain terminating inhibitors. Proc. Natl. Acad. Sci. USA74:5463–5467. 57.San Jose, C., R. A. Monge, R. Perez-Diaz, J. Pla, and C. Nombela.1996. The

mitogen-activated protein kinase homologHOG1 gene controls glycerol accumulation in the pathogenic fungusCandida albicans. J. Bacteriol.178:

5850–5852.

58.Sherman, F., G. R. Fink, and J. B. Hicks.1986. Methods in yeast genetics. Cold Spring Harbor Laboratory, Cold Spring Harbor, N.Y.

59.Smith, D. A., S. Nicholls, B. A. Morgan, A. J. Brown, and J. Quinn.2004. A conserved stress-activated protein kinase regulates a core stress response in the human pathogenCandida albicans. Mol. Biol. Cell15:4179–4190.

MAP kinase pathway. Mol. Microbiol.37:382–397.

66.Westfall, P. J., D. R. Ballon, and J. Thorner.2004. When the stress of your environment makes you go HOG wild. Science306:1511–1512.

67.Winkler, A., C. Arkind, C. P. Mattison, A. Burkholder, K. Knoche, and I. Ota.2002. Heat stress activates the yeast high-osmolarity glycerol mitogen-activated protein kinase pathway, and protein tyrosine phosphatases are essential under heat stress. Eukaryot. Cell1:163–173.

68.Wojda, I., R. Alonso-Monge, J. P. Bebelman, W. H. Mager, and M. Siderius.

2003. Response to high osmotic conditions and elevated temperature in

Saccharomyces cerevisiaeis controlled by intracellular glycerol and involves coordinate activity of MAP kinase pathways. Microbiology149:1193–1204. 69.Zarzov, P., C. Mazzoni, and C. Mann.1996. The SLT2(MPK1) MAP kinase

is activated during periods of polarized cell growth in yeast. EMBO J.

15:83–91.

70.Zhang, Y., R. Lamm, C. Pillonel, S. Lam, and J. R. Xu.2002. Osmoregula-tion and fungicide resistance: theNeurospora crassa os-2gene encodes a

HOG1mitogen-activated protein kinase homologue. Appl. Environ. Micro-biol.68:532–538.

on September 8, 2020 by guest

http://ec.asm.org/

Figure

FIG. 1. Expression of TdHOG1same medium containing 1.5 M NaCl, and the ODS. cerevisiaewere adjusted to an ODconcentration)
FIG. 2. TdHog1p phosphorylation and growth of the Tdhog1NaCl-containing medium (0.5 M)
Fig. 6B, the level of phospho-TdHog1p did not vary after
FIG. 6. TdHog1p is essential for the response to different stresses. (A) Exponentially growing cells of T
+2

References

Related documents

[9] applied descriptive statistical measure and a partial correlation analysis to groundwater quality data set monitored Groundwater Quality Investigation Using Multivariate

Comparison of cooling system done on the basis of power consumption, surface temperature of the motor, Capital Cost , Maintenance cost, Defects due to cooling

A modified periodic current reference is tracked to obtain smooth torque.In order to improve current tracking in the presence of periodic reference signals and disturbances in

Concerning application of industrial robots in European countries, increasing trend of industrial robot application is evident for Germany and Spain, while France

This system is proposed on the RFID technology where RFID tags are joined into the books, & on the user card and RFID reader are used to read RFID tags properly, and

Similar archaeological structures have been discovered by USM Centre for Global Archaeological Research (USM CGAR) in 2010 at Kampung Gading, Kampung Chemara and Kampung

energy technologies are widely deployed in microgrids. The deployment of these new technologies in the form of a microgrid is preferred due to several

ANSYS HFSS software is the industry standard for simulating 3-D full-wave electromagnetic fields. Its gold-standard accuracy, advanced solver and compute technology