Protein kinase C, a highly conserved signaling molecule among eukaryotes, has been implicated in the regulation of cellular processes such as cell proliferation and polarized growth. In Saccharomyces cerevisiae, the unique protein kinase C Pkc1p is thought to have multiple functions, including the activation of the Mpk1p (Slt2p) MAP kinase pathway, which is essential for cell wall construction and bud emergence. However, little is known about the other functions of Pkc1p. In the course of screening for the mutants that suppress the Ca2+-sensitivity phenotype of the Ca2+-sensitive strain zdsΔ, we isolated a novel mutant allele (scz6/pkc1-834) of PKC1. Unlike the previously characterized PKC1 allele stt1-1, heat-shock-induced Mpk1p activation and cell-wall integrity were not impaired in the pkc1-834 mutant. By contrast, the mutant was defective in the maintenance of Ca2+-induced F-actin polarization in a manner independent of Mpk1p activation. This phenotype was caused by a decreased expression level of the G1 cyclin Cln2p. The Rho1 small G protein molecular switch was suggested to be involved in the novel Pkc1p function. The Pkc1p novel function was required for posttranscriptional upregulation of Cln2p and appeared to be important for the coordinated regulation of polar bud growth and the cell cycle.

Cell polarity is essential for morphogenesis and development (Drubin and Nelson, 1996). In the budding yeast Saccharomyces cerevisiae, a number of stage-specific events direct the position and the morphology of the bud. Environmental stimuli activate a bifurcated signal transduction pathway to promote a coordinated response involving a specific pattern of gene expression, G2/M cell-cycle delay, apically polarized actin distribution and unipolar distal bud-site selection.

As yeast cells begin the G1/S transition, Cln1p and Cln2p G1 cyclins accumulate, activate the cyclin-dependent kinase Cdc28p, and induce clustering of actin at the presumptive bud site, thus focusing secretion of new cell-wall components onto the nascent bud (Pruyne and Bretscher, 2000). Continued Cdc28p-Cln1/2p activity maintains this clustered distribution of actin in the bud tip and polarized growth (Lew and Reed, 1993). Cells over-expressing Cln1p or Cln2p form a highly elongated bud because of sustained polarized growth (Lew and Reed, 1995). Polarized growth continues in G2, as the mitotic cyclin Clb1/2p-Cdc28p activity remains sequestered via inhibitory phosphorylation by Swe1p, a protein kinase that inhibits Cdc28p (Booher et al., 1993; Lew and Reed, 1995). Recently, we reported that Ca2+-signaling pathways induced the formation of highly polarized bud and delayed the cell-cycle progression in G2 (Mizunuma et al., 1998; Mizunuma et al., 2001). The effect of Ca2+ on cell-cycle regulation and polarized bud growth are pronounced on a zds1Δ background lacking the negative regulator for SWE1 and CLN2 transcription (Bi and Pringle, 1996; Ma et al., 1996; Yu et al., 1996; Mizunuma et al., 1998; Mizunuma et al., 2001; Mizunuma et al., 2004). These phenotypes are peculiar to calcium, because a similar effect was not observed with sorbitol (300 mM) or MgCl2 (100 mM; our unpublished data).

In S. cerevisiae, a unique protein kinase C, Pkc1p, functions in a multiplicity of pathways, including those related to cell wall integrity, bud emergence, stretching of the plasma membrane (Mazzoni et al., 1993; Levin and Errede, 1995; Gustin et al., 1998; Heinisch et al., 1999) and organization of the actin cytoskeleton (Heinisch et al., 1999). One well-known pathway for Pkc1p involves the sequentially activated protein kinases Bck1p, the redundant Mkk1p/Mkk2p and Mpk1p (Slt2p), which ultimately activates by phosphorylation of transcription factors, including the Rlm1p and SBF (Swi4p-Swi6p) complex (Dodou and Treisman, 1997; Madden et al., 1997; Baetz et al., 2001). This pathway is essential for cell wall integrity, bud emergence, responses to hypotonic shock, stretching of the plasma membrane, and repolarization of actin upon cell-wall stress (Levin et al., 1990; Levin and Bartlett-Heubusch, 1992; Paravicini et al., 1992; Yoshida et al., 1992; Lew and Reed, 1995; Davenport et al., 1995; Kamada et al., 1995; Igual et al., 1996; Marini et al., 1996; Zarzov et al., 1996; Gray et al., 1997; Gustin et al., 1998; Helliwell et al., 1998; Delley and Hall, 1999). In addition, Pkc1p is involved in controlling the depolarization of actin during adaptation to growth at higher temperatures in a manner independent of the Mpk1p MAP kinase pathway (Delley and Hall, 1999). However, little is known about this Pkc1p effecter branch.

Much of our understanding of Pkc1p functions has come from the phenotypes caused by MPK1 deletion and/or severe temperature-sensitive alleles of PKC1, all of which involve a defect in the cell wall. The identification of a Mpk1p-independent role of Pkc1p will be facilitated by the isolation and characterization of a PKC1 allele specifically defective in the novel function.

Here, we describe the isolation and characterization of a pkc1 mutant allele (pkc1-834) defective in a previously unknown pathway. The pkc1-834 allele was defective in the maintenance of Ca2+-induced F-actin polarization in a manner independent of Mpk1p activation. This phenotype appeared to be caused by decreased expression of Cln2p. Pkc1p was required for posttranscriptional upregulation of Cln2p. The Rho1 small G protein molecular switch was suggested to be involved in the novel Pkc1p function.

Strains and media

Yeast strains used in this study are listed in Table 1. The stt1-1 (SYT11-12A) mutant was backcrossed to the W303 strain at least five times to homogenize the background. Yeast cells were grown in YPD medium (1% yeast extract, 2% Bacto-peptone, 2% glucose supplemented with 400 μg/ml adenine and 200 μg/ml uracil).

Table 1.

Yeast strains used in this study

Strain Genotype Source or reference
W303 background: trp1 leu2 ade2 ura3 his3 can1-100    
DHT22-1b  MATa trp1 leu2 ade2 ura3 his3 can1-100  Our stock  
W303-1D  MATa/α  trp1/trp1 leu2/leu2 ade2/ade2 ura3/ura3 his3/his3 can1-100/can1-100  Our stock  
YAT1  MATa zds1::TRP1 Mizunuma et al., 1998  
TNP46  MATa mpk1::HIS3  T. Nakamura  
YMM28  MATa scz6/pkc1-834  This study  
YMM53  MATa hsl1::ura3 3xHA-HSL1 Mizunuma et al., 2001  
YMM70  MATa mpk1::HIS3 hsl1::ura3 3xHA-HSL1 Mizunuma et al., 2001  
YMM114  MATa stt1-1 Yoshida et al., 1992  
YMM125  MATa/α  zds1::TRP1/zds1::TRP1  This study  
YMM134  MATa zds1::TRP1 stt1-1  This study  
YMM143  MATa pkc1-834 hsl1::ura3 3xHA-HSL1  This study  
YMM144  MATa stt1-1 hsl1::ura3 3xHA-HSL1  This study  
YMM170-1  MATa zds1::TRP1 pkc1-834 cln2::GAL1-CLN2-3xHA-LEU2  This study  
YMM178  MATa zds1::TRP1 pkc1-834 swe1::HIS3::SWE1-9xMyc CLN2-3xHA  This study  
YMM179  MATa pkc1-834 swe1::HIS3::SWE1-9xMyc CLN2-3xHA  This study  
YMM180  MATa swe1::HIS3::SWE1-9xMyc CLN2-3xHA Mizunuma et al., 2004  
YMM187  MATa zds1::TRP1 swe1::HIS3::SWE1-9xMyc CLN2-3xHA Mizunuma et al., 2004  
YMM196  MATa stt1-1 swe1::HIS3::SWE1-9xMyc CLN2-3xHA  This study  
YMM218  MATa/α  zds1::TRP1 stt1-1/zds1::TRP1 stt1-1  This study  
YMM219  MATa/α  zds1::TRP1 pkc1-834/zds1::TRP1 pkc1-834  This study  
YMM220  MATa/α  zds1::TRP1 pkc1-834/zds1::TRP1 stt1-1  This study  
YMM235  MATa/α  zds1::TRP1/ZDS1  This study  
YMM236  MATa/α  zds1::TRP1/ZDS1, pkc1::LEU2/PKC1  This study  
YMM4020  MATa zds1::TRP1 pkc1-834  This study  
YPH500 background: trp1 leu2 ade2 ura3 his3 lys2    
YOC729  MATα  rho1::LYS2 ade3::rho1-3::HIS3  Y. Ohya  
YOC752  MATα  rho1::HIS3 ade3::rho1-2::LEU2  Y. Ohya  
YOC754  MATα  rho1::HIS3 ade3::rho1-4::HIS3  Y. Ohya  
YOC755  MATα  rho1::HIS3 ade3::rho1-5::LEU2  Y. Ohya  
YOC764  MATα  rho1::HIS3 ade3::RHO1::LEU2  Y. Ohya  
YMM147  MATα  rho1::LYS2 ade3::rho1-3::HIS3 zds1::URA3  This study  
YMM148  MATα  rho1::HIS3 ade3::rho1-4::HIS3 zds1::URA3  This study  
YMM149  MATα  rho1::HIS3 ade3::rho1-5::LEU2 zds1::URA3  This study  
YMM230  MATα  rho1::HIS3 ade3::RHO1::LEU2 zds1::URA3  This study  
YMM232  MATα  rho1::HIS3 ade3::rho1-2::LEU2 zds1::URA3  This study  
EG123 background: leu2-3, 112 trp1-1 ura3-52 his4 can1r    
DL376  MATa pkc1::LEU2 Levin and Bartlett-Heubusch, 1992  
Strain Genotype Source or reference
W303 background: trp1 leu2 ade2 ura3 his3 can1-100    
DHT22-1b  MATa trp1 leu2 ade2 ura3 his3 can1-100  Our stock  
W303-1D  MATa/α  trp1/trp1 leu2/leu2 ade2/ade2 ura3/ura3 his3/his3 can1-100/can1-100  Our stock  
YAT1  MATa zds1::TRP1 Mizunuma et al., 1998  
TNP46  MATa mpk1::HIS3  T. Nakamura  
YMM28  MATa scz6/pkc1-834  This study  
YMM53  MATa hsl1::ura3 3xHA-HSL1 Mizunuma et al., 2001  
YMM70  MATa mpk1::HIS3 hsl1::ura3 3xHA-HSL1 Mizunuma et al., 2001  
YMM114  MATa stt1-1 Yoshida et al., 1992  
YMM125  MATa/α  zds1::TRP1/zds1::TRP1  This study  
YMM134  MATa zds1::TRP1 stt1-1  This study  
YMM143  MATa pkc1-834 hsl1::ura3 3xHA-HSL1  This study  
YMM144  MATa stt1-1 hsl1::ura3 3xHA-HSL1  This study  
YMM170-1  MATa zds1::TRP1 pkc1-834 cln2::GAL1-CLN2-3xHA-LEU2  This study  
YMM178  MATa zds1::TRP1 pkc1-834 swe1::HIS3::SWE1-9xMyc CLN2-3xHA  This study  
YMM179  MATa pkc1-834 swe1::HIS3::SWE1-9xMyc CLN2-3xHA  This study  
YMM180  MATa swe1::HIS3::SWE1-9xMyc CLN2-3xHA Mizunuma et al., 2004  
YMM187  MATa zds1::TRP1 swe1::HIS3::SWE1-9xMyc CLN2-3xHA Mizunuma et al., 2004  
YMM196  MATa stt1-1 swe1::HIS3::SWE1-9xMyc CLN2-3xHA  This study  
YMM218  MATa/α  zds1::TRP1 stt1-1/zds1::TRP1 stt1-1  This study  
YMM219  MATa/α  zds1::TRP1 pkc1-834/zds1::TRP1 pkc1-834  This study  
YMM220  MATa/α  zds1::TRP1 pkc1-834/zds1::TRP1 stt1-1  This study  
YMM235  MATa/α  zds1::TRP1/ZDS1  This study  
YMM236  MATa/α  zds1::TRP1/ZDS1, pkc1::LEU2/PKC1  This study  
YMM4020  MATa zds1::TRP1 pkc1-834  This study  
YPH500 background: trp1 leu2 ade2 ura3 his3 lys2    
YOC729  MATα  rho1::LYS2 ade3::rho1-3::HIS3  Y. Ohya  
YOC752  MATα  rho1::HIS3 ade3::rho1-2::LEU2  Y. Ohya  
YOC754  MATα  rho1::HIS3 ade3::rho1-4::HIS3  Y. Ohya  
YOC755  MATα  rho1::HIS3 ade3::rho1-5::LEU2  Y. Ohya  
YOC764  MATα  rho1::HIS3 ade3::RHO1::LEU2  Y. Ohya  
YMM147  MATα  rho1::LYS2 ade3::rho1-3::HIS3 zds1::URA3  This study  
YMM148  MATα  rho1::HIS3 ade3::rho1-4::HIS3 zds1::URA3  This study  
YMM149  MATα  rho1::HIS3 ade3::rho1-5::LEU2 zds1::URA3  This study  
YMM230  MATα  rho1::HIS3 ade3::RHO1::LEU2 zds1::URA3  This study  
YMM232  MATα  rho1::HIS3 ade3::rho1-2::LEU2 zds1::URA3  This study  
EG123 background: leu2-3, 112 trp1-1 ura3-52 his4 can1r    
DL376  MATa pkc1::LEU2 Levin and Bartlett-Heubusch, 1992  

DNA sequence

Sequencing of double strand DNA was done by the cycle sequencing method, using a ALFred DNA sequencer (Amersham Biosciences). For PCR amplification of the genomic DNA, a Thermo Sequenase fluorescently labeled primer cycle sequencing kit with 7-deaza-dGTP (Amersham Biosciences) was used. M13 universal and reverse primers from Cy5™ AutoRead Sequencing Kit (Amersham Biosciences) were also used. The entire open reading frame of PKC1 was divided into four regions. PCR products were synthesized using primers, region I: 5′-AAAGAGCTCATGAGTTTTTCACAAT-3′ (sense strand, the SacI site is underlined) and 5′-TGCACCTGCAGATTCTTTGCTGCTGAAGTAGAC-3′ (antisense strand, the PstI site is underlined). Region II: 5′-AAACTGCAGGTGCAAGGCGAAAATA-3′ (sense strand, the PstI site is underlined), and 5′-AACTTCTAGATCCTCTTGGATTTCC-3′ (antisense strand, the XbaI site is underlined). Region III: 5′-AAATCTAGAAGTTGATCACATTGAT-3′ (sense strand, the XbaI site is underlined), and 5′-GATAGGTCGACTCTTTGGTTTTGAA-3′ (antisense strand, the SalI site is underlined). Region IV: 5′-AAAGTCGACCTATCTGTAAGAAGGG-3′ (sense strand, the SalI site is underlined), and 5′-TTTCTAAGCTTTCATAAATCCAAATCATCT-3′ (antisense strand, the HindIII site is underlined).

Site-directed mutagenesis

To construct plasmids containing the PKC1 gene with single mutations, we employed site-directed mutagenesis. The mutagenic primers for PKC1 (P1102S) were 5′-CCACCCTACATCTCAGAAATTAAATCTCCG-3′ and 5′-CGGAGATTTAATTTCTGAGATGTAGGGTGG-3′; and those for the PKC1 (N834D), 5′-GTTCTTGGTAAAGGTGATTTTGGTAAAG-3′ and 5′-CTTTACCAAAATCACCTTTACCAAGAAC-3′. PCR was performed with PfuTurbo DNA polymerase (Stratagene) and pUC119-PKC1 as a PCR template. Successful mutagenesis was confirmed by DNA sequencing. The low-copy plasmid YCp50-PKC1 (P1102S) or YCp50-PKC1 (N834D) construct was generated as follows. The SalI-PvuII fragment of pUC119-PKC1 containing PKC1 (P1102S) or PKC1 (N834D) was ligated into the SalI-NruI site of YCp50. YCp50-PKC1 (P1102S) or YCp50-PKC1 (N834D) was used for complementation analysis for pkc1Δ or zds1Δ pkc1Δ.

Gene disruption and Strain construction

The pkc1Δ strain in the W303 background was constructed by gene replacement. Genomic DNA was isolated from the pkc1Δ::LEU2 (DL376) strain (Levin et al., 1990). The PKC1 locus was amplified by PCR using primers 5′-ATGTTTGTCCCACTCCAGGTTGCAC-3′ and 5′-TTTGAGACGTCATGAACTCTCGCGG-3′. The amplified fragment was used to transform a diploid W303 strain [ZDS1/ZDS1 (W303-1D) or zds1Δ::TRP1/ZDS1(YMM235)]. Replacement was confirmed by PCR. The diploid was sporulated and the tetrads were dissected and analyzed. When the spores were grown at room temperature, all of the tetrads produced two colonies that grew at the wild-type rate and two colonies that grew very poorly even in the presence of 1 M sorbitol. The fast growing colonies all had a Leu phenotype, indicating that the very poor growth results from the integration of the knockout construct. Deletion of the PKC1 gene in the Leu+ strains was further confirmed by complementation of the very poor growth by a centromere plasmid carrying the PKC1 gene.

The pkc1Δ zds1Δ (pkc1Δ) strain bearing the N834D, P1102S or wild-type PKC1 allele was constructed as follows. The YCp50-PKC1 (P1102S), YCp50-PKC1 (N834D) or YCp50-PKC1 plasmid was transformed into a diploid strain (YMM236 or YMM235). The diploid was sporulated and the tetrads were dissected and analyzed. The spores that were grown minus uracil (Ura), tryptophan (Trp) and leucine (Leu) (which selects for the zds1Δ pkc1Δ strain bearing the N834D, P1102S or wild-type PKC1 allele) or minus Ura and Leu (which selects for the pkc1Δ strain bearing N834D, P1102S or the wild-type PKC1 allele) were isolated and their phenotypes examined.

The zds1Δ strain of the YPH500 background was constructed by gene replacement. Disruption of the ZDS1 gene was performed using a zds1 disruption plasmid pUC119-zds1::URA3. The zds1::URA3 fragment was used to transform the YPH500 background.

Actin staining

Actin was visualized in formaldehyde-fixed cells by using Rhodamine-conjugated phalloidin as described previously (Adams and Pringle, 1991).

Assessment of cell lysis

Qualitative assessment of cell lysis in colonies was done using the alkaline phosphatase assay described previously (Saka et al., 2001). Colonies were formed on YPD agar plates at 25°C for 2 days, and the plates were then shifted to 35°C and overlaid with an alkaline phosphatase assay solution. Colonies that contain lysed cells stain blue, whereas intact colonies remain white.

Cell culture synchronization, RNA isolation and northern blot analysis

Cell culture synchronization was done by the procedure described previously (Mizunuma et al., 1998). Total RNA was isolated by the hot phenol method. The isolated RNA was separated on a 1% agarose gel, transferred to a nylon membrane, and then subjected to northern blot analysis. The SWE1, CLN2, and ACT1 probes were generated by random-primed labeling of a 0.7 kb BglII fragment of SWE1, a 1.3 kb NcoI-XhoI fragment of CLN2 and a 1.1 kb XhoI-KpnI fragment of ACT1, respectively, with [α-32P]dCTP by use of a multiprime DNA labeling kit (Amersham Biosciences).

Western blot analysis

Preparation of cell extracts and immunoprecipitations were performed basically as described previously (Mizunuma et al., 2001; Mizunuma et al., 2004). For detection of the HA-tagged proteins, Myc-tagged proteins, Pkc1p, phosphorylated Mpk1, Mpk1p and Cdc28p, monoclonal antibodies 12CA5 (BAbCO) against the HA epitope, 9E10 (BAbCO) against the Myc epitope, polyclonal anti-PKC1 (Santa Cruz Biotech.), anti-phospho-p44/42 MAP kinase (New England Biolabs Inc.), anti-MAPK antibody (Santa Cruz Biotech.), and anti-PSTAIRE (Santa Cruz Biotech.), respectively, were used.

Identification of scz6 mutation, a mutant allele of PKC1 gene, as a suppressor of calcium sensitivity of zds1Δ strain

The growth of the zds1Δ strain of yeast in medium containing 300 mM CaCl2 is severely inhibited, showing G2 arrest and highly polarized bud growth. In our previous study, suppressor mutants of the calcium-sensitive phenotypes were isolated and classified into 14 complementation groups (Mizunuma et al., 2001). One of these mutants, scz6, is the subject of this report. As shown in Fig. 1A, the inhibition of the growth of zds1Δ cells by exogenous calcium was suppressed by an additional mutation, scz6. Among the physiological effects of calcium on the zds1Δ strain, the formation of a highly polarized bud and G2 delay were elicited by exogenous CaCl2 at concentrations (∼100 mM) lower than those (∼300 mM) required for the growth inhibition. We therefore used 100 mM CaCl2 in most of the subsequent experiments to examine the effect of calcium on the cell cycle and morphology. The polarized bud growth induced by calcium was also suppressed by the scz6 mutation (Fig. 1B), whereas the other calcium phenotype (G2 delay) was not suppressed by it (Fig. 1C). These results suggested that the suppression of growth inhibition and polarized bud growth, but not of the cell-cycle delay, was the consequence of the scz6 mutation. In addition to these suppressor phenotypes, the zds1Δ scz6 mutant exhibited a growth defect at 37°C (Fig. 1A). We further found that the scz6 mutation in itself exhibited a slow growth phenotype at temperatures between 25°C and 37°C (Fig. 1A). To identify the mutant gene, we introduced a centromeric genomic library into the scz6 strain and recovered a plasmid that suppressed the slow growth at 37°C. The PKC1 gene, encoding a unique protein kinase C, was identified as the gene that complemented the scz6 mutation. To demonstrate that the mutation was situated at the PKC1 locus, we constructed a YIp5-PKC1 plasmid for integration and linearized the plasmid for use in transforming the wild-type strain. The transformants were crossed with the scz6 strain, and the resulting diploid was sporulated and subjected to tetrad analysis. In all 12 tetrads the Ura+ or Ura phenotype cosegregated with normal or slow growth, respectively, indicating that the mutation was situated at, or closely linked to, the PKC1 locus (Levin et al., 1990; Levin and Bartlett-Heubusch, 1992). As will be shown later, because the scz6 mutation causes a change of amino acid 834 of Pkc1p from asparagine to aspartic acid (Fig. 5), we shall hereafter refer to scz6 mutation as pkc1-834.

The pkc1-834 mutant is defective in sustaining calcium-induced Cln2p elevation in zds1Δ strain

Why does the pkc1-834 allele of PKC1 suppress only the calcium-induced hyperpolarized bud growth, but not that of G2 delay? Previously, we showed that exogenous calcium induces the sustained transcription of the SWE1 and CLN2 genes (Mizunuma et al., 1998; Mizunuma et al., 2001; Mizunuma et al., 2004). We therefore compared the effect of exogenous calcium on SWE1 and CLN2 mRNA levels in synchronized cell cultures of wild-type and pkc1-834 strains by northern blot hybridization (Fig. 2). However, no effect of calcium with respect to either timing or expression levels of these genes was detected.

Fig. 1.

The scz6 mutation, a mutant allele of the PKC1 gene, suppresses various phenotypes of the zds1Δ strain. (A) Effect of scz6/pkc1-834 and stt1-1, mutant alleles of PKC1, on the growth of the zds1Δ mutant strain on solid medium. Wild-type (DHT22-1b), zds1Δ (YAT1), zds1Δ scz6/pkc1-834 (YMM4020), scz6/pkc1-834 (YMM28), zds1Δ stt1-1 (YMM134), and stt1-1 (YMM114) cells were spotted on YPD plates and grown at 25°C (2 days), or at 33°C or 37°C with or without 300 mM CaCl2 (2 days). (B,C) Cell morphology (B) and flow cytometry analysis of PI-stained cells (C) of various strains after 6 hours incubation with 100 mM CaCl2 at 25°C (1C, haploid; 2C, diploid).

Fig. 1.

The scz6 mutation, a mutant allele of the PKC1 gene, suppresses various phenotypes of the zds1Δ strain. (A) Effect of scz6/pkc1-834 and stt1-1, mutant alleles of PKC1, on the growth of the zds1Δ mutant strain on solid medium. Wild-type (DHT22-1b), zds1Δ (YAT1), zds1Δ scz6/pkc1-834 (YMM4020), scz6/pkc1-834 (YMM28), zds1Δ stt1-1 (YMM134), and stt1-1 (YMM114) cells were spotted on YPD plates and grown at 25°C (2 days), or at 33°C or 37°C with or without 300 mM CaCl2 (2 days). (B,C) Cell morphology (B) and flow cytometry analysis of PI-stained cells (C) of various strains after 6 hours incubation with 100 mM CaCl2 at 25°C (1C, haploid; 2C, diploid).

We next investigated the effect of exogenous calcium on Swe1p and Cln2p levels by western blot analysis of synchronized cell cultures of strains with chromosomally integrated constructs for both Myc-tagged Swe1p and HA-tagged Cln2p (Fig. 3A). The Swe1p and Cln2p levels in the cells released from G1 arrest were analyzed by western blotting. As previously reported, Swe1p and Cln2p levels in wild-type cells oscillated during the cell cycle, peaking at the time of bud emergence and declining before nuclear division (McMillan et al., 1998; Sia et al., 1998). Exogenous calcium caused a 20-minute lag in the increase in Swe1p and Cln2p levels, reflecting the delays in the elevation of the respective mRNA (Fig. 2); and the increased protein levels were sustained longer by calcium (Fig. 3A). There was no significant difference among the strains when cultivated in YPD or YPD plus CaCl2 medium with regard to the periodic patterns of the Swe1p level. In the presence of exogenous calcium, Swe1p in the pkc1-834 and zds1Δ pkc1-834 cells accumulated, consistent with the observation that the scz6 mutation failed to suppress the G2 delay caused by calcium (Fig. 1C and Fig. 3A). The G2 delay was partially suppressed by the deletion of the SWE1 gene, suggesting that the delay was indeed mediated at least partially by the activation of Swe1p (data not shown). However, Cln2p levels in the pkc1-834 cells were significantly lower than those in the wild-type strain. Moreover, Cln2p levels in the pkc1-834 strain, in contrast to those in the wild-type strain, were not elevated by exogenous calcium in either ZDS1 or zds1Δ backgrounds, suggesting that Pkc1p is important for the elevation and maintenance of Cln2p levels in response to exogenous calcium (Fig. 3A). The defect in Cln2p accumulation in the pkc1-834 cells could be due to a decreased rate of Cln2p synthesis, the destabilization of Cln2p, or both.

To examine the stability of Cln2p, we determined the decay rate of Cln2p in the cells treated with cycloheximide at a restricted temperature (Fig. 3B). In random culture of pkc1-834 cells, it was noted that the Cln2p abundance at time zero was similar to that of wild-type cells, possibly reflecting the prolonged G2 phase in the pkc1-834 cells in random culture. The abundance of Cln2p in wild-type and stt1-1 cells decreased rapidly, reflecting the rapid turnover of this protein (Lanker et al., 1996). On the other hand, the Cln2p level in pkc1-834 mutant cells was maintained more stably. This could be due to a defect in Cln2p destabilization in the pkc1-834 strain (see Discussion). These results, together with the observation that the pkc1-834 mutation suppressed the Ca2+-induced hyper-polarized bud growth, further suggested that the Pkc1p-mediated Cln2p accumulation was important for polarizing bud growth. Consistent with this notion, the over-expression of CLN2 from a GAL1 promoter in the zds1Δ pkc1-834 strain led to hyper-polarized bud growth (Fig. 3D).

Fig. 2.

Pkc1p is not involved in the regulation of oscillation of SWE1 and CLN2 mRNAs by calcium during cell cycle. (A) SWE1 or CLN2 mRNA levels in synchronized cell cultures of various strains were determined by Northern blotting, using the SWE1, CLN2, and ACT1 probes. Cells of wild-type (YMM180), pkc1-834 (YMM179) and stt1-1 (YMM196) at early log-phase (OD600 of 0.2∼0.3) were synchronized with α-factor in G1, and resuspended in YPD or YPD plus 100 mM CaCl2. Samples were taken at an interval of 20 minutes after release from arrest. (B) Changes in the relative intensity of the SWE1, CLN2 and ACT1 mRNA. The amount of the SWE1 and CLN2 mRNA in A was normalized to a constant level of ACT1 mRNA. In each case, the maximum value for the first cell cycle was referred to as 1. Time 0 is the point of release from the G1-phase arrest.

Fig. 2.

Pkc1p is not involved in the regulation of oscillation of SWE1 and CLN2 mRNAs by calcium during cell cycle. (A) SWE1 or CLN2 mRNA levels in synchronized cell cultures of various strains were determined by Northern blotting, using the SWE1, CLN2, and ACT1 probes. Cells of wild-type (YMM180), pkc1-834 (YMM179) and stt1-1 (YMM196) at early log-phase (OD600 of 0.2∼0.3) were synchronized with α-factor in G1, and resuspended in YPD or YPD plus 100 mM CaCl2. Samples were taken at an interval of 20 minutes after release from arrest. (B) Changes in the relative intensity of the SWE1, CLN2 and ACT1 mRNA. The amount of the SWE1 and CLN2 mRNA in A was normalized to a constant level of ACT1 mRNA. In each case, the maximum value for the first cell cycle was referred to as 1. Time 0 is the point of release from the G1-phase arrest.

Using the same synchronous cell cultures as used in the experiment for Fig. 3A, we determined the cellular DNA content (Fig. 3C). As previously reported (Mizunuma et al., 1998; Mizunuma et al., 2001), a delay in G1, as well as in G2, was induced by calcium in all the strains examined (compare YPD and YPD + Ca2+ of Fig. 3C). The G1 delay seemed to be due to the downregulation of CLN2 mRNA (Fig. 2). The pkc1-834 cells showed a G1 delay after release from G1 arrest, even in the absence of exogenous calcium, indicating that Pkc1p is required for the G1-S transition (Fig. 3C).

Pkc1p is required for the maintenance of calcium-induced F-actin polarization

Both actin polarization and hyper-polarized bud growth is dependent on the G1 cyclins (Lew and Reed, 1993). Cell polarization can be estimated by the distribution of F-actin. We therefore examined the ability of the pkc1-834 cells to polarize their F-actin cytoskeleton in response to calcium. Using the same synchronous cell cultures described in the above section, we examined the effects of exogenous calcium on bud emergence and F-actin polarization at the bud site (Fig. 4). After a 2-hour treatment of the cells with α-factor, about 80% of wild-type cells yielded a mating projection, whereas in the pkc1-834 cells it was about 30%, although the treatment resulted in a G1 block in the mutant cells (data not shown and Fig. 3C). However, pkc1-834 cells on the zds1Δ background formed a mating projection at a ratio comparable to that of the wild-type cells. Therefore, we used the zds1Δ mutant background in this study (Fig. 4). Upon release of the α-factor-treated cells to YPD medium, bud emergence of the zds1Δ pkc1-834 cells was delayed by 20 minutes compared with that of the zds1Δ cells. Bud emergence of both zds1Δ pkc1-834 and pkc1-834 cells was delayed by exogenous calcium similarly (Fig. 4B). These results suggested that the suppression of the calcium-induced polarized bud growth by the pkc1-834 mutation is not due to the defect in bud emergence.

In YPD medium, the F-actin organization of the zds1Δ and zds1Δ pkc1-834 cells, as determined by rhodamine-phalloidin staining, was essentially normal (Fig. 4A,B, compare 40∼80-minute samples for the zds1Δ and zds1Δ pkc1-834 cells in YPD medium). Patches of F-actin were restricted to the growing buds, and F-actin cables were oriented toward the bud. Large budded cells were capable of polarizing F-actin to the bud neck (Fig. 4A, 100∼120 minute samples for the zds1Δ and zds1Δ pkc1-834 cells in YPD medium). However, in the presence of calcium, F-actin polarization in the zds1Δ pkc1-834 cells was strikingly different from that in the zds1Δ cells (Fig. 4). In zds1Δ cells, the accumulation of F-actin patches in the bud after release from G1 arrest was transiently disturbed by exogenous calcium (compare with and without (±) Ca2+samples of zds1Δ cells at 40 minutes), and then the patches became rapidly localized to the bud (see zds1Δ cells +Ca2+ at 40 and 80 minutes). The patches were maintained in the growing buds during polarized growth (compare ±Ca2+ samples of zds1Δ cells at 100 and 120 minutes). The timing of F-actin polarization was coincident with the elevation of Cln2p levels during the cell cycle (Fig. 3A), suggesting that Cln2p elevation is important for the polarization of F-actin patches. In the zds1Δ pkc1-834 cells, F-actin patches were transiently dispersed throughout the cell and then rapidly became localized at the bud, as in the zds1Δ cells (compare ±Ca2+ samples of zds1Δ pkc1-834 cells at 40 and 80 minutes). However, the maintenance of the localized F-actin patches in growing buds of the zds1Δ pkc1-834 cells was abolished faster than that in the zds1Δ cell (compare +Ca2+ samples of the zds1Δ and zds1Δ pkc1-834 cells at 120 minutes). F-actin patches of these cells varied, being concentrated in the bud or the bud neck or dispersed throughout the cells. The failure of maintenance of F-actin polarization in the zds1Δ pkc1-834 cells might be due to the low level of Cln2p expression (Fig. 3A). These results suggested that Pkc1p is important for the maintenance of F-actin polarization, but not for the de- and repolarization of it, in response to exogenous calcium.

Activation of Mpk1p and cell wall integrity are not impaired in the pkc1-834 mutant

Yeast bearing the stt1-1 mutation, a previously characterized temperature-sensitive PKC1 allele, exhibits a defect in cell wall maintenance and in G2-M transition (Yoshida et al., 1992). So we examined whether the stt1-1 mutation could suppress the calcium sensitivity phenotypes of the zds1Δ strain. However, the stt1-1 allele only partially suppressed the phenotype, suggesting that the defect of the pkc1-834 allele may be functionally distinct from that of the stt1-1 allele (Fig. 1A,B).

Fig. 3.

Pkc1p is required for the maintenance of Cln2p. (A) Oscillation of Swe1 or Cln2 protein levels during cell-cycle progression in various strains that were synchronized with α-factor was determined by western blot analysis. Cells in early log-phase (OD600 of 0.2∼0.3) of wild-type (YMM180), pkc1-834 (YMM179), zds1Δ (YMM187), zds1Δ pkc1-834 (YMM178) and stt1-1 (YMM196) strains, with a chromosomally integrated construct encoding genomic copies of Swe1p-9xMyc and Cln2p-3xHA, were synchronized with α-factor in G1, and resuspended in YPD or YPD plus 100 mM CaCl2 at 25°C. Samples were taken at of 20-minute intervals after release from arrest. Cdc28 protein was used as an internal loading control. (B) Stability of Cln2p. Wild-type (YMM180), pkc1-834 (YMM179) and stt1-1 (YMM196) cells were cultured at 37°C for 2 hours; and then samples were taken at the indicated times after the addition of 100 μg/ml cycloheximide, and subjected to western bolt analysis. The asterisk indicates a non-specific band. (C) DNA content for the same samples as in A. (D) Effect of over-expression of CLN2 on the bud growth of the zds1Δ pkc1-834 strain. A strain containing a chromosomally integrated construct for GAL-regulated Cln2p-3xHA (YMM170-1; cln2::GAL1–CLN2HA–LEU2) were grown in the medium containing 2% raffinose (GAL1 promoter off) or 2% galactose (GAL1 promoter on) for 8 hours at 25°C.

Fig. 3.

Pkc1p is required for the maintenance of Cln2p. (A) Oscillation of Swe1 or Cln2 protein levels during cell-cycle progression in various strains that were synchronized with α-factor was determined by western blot analysis. Cells in early log-phase (OD600 of 0.2∼0.3) of wild-type (YMM180), pkc1-834 (YMM179), zds1Δ (YMM187), zds1Δ pkc1-834 (YMM178) and stt1-1 (YMM196) strains, with a chromosomally integrated construct encoding genomic copies of Swe1p-9xMyc and Cln2p-3xHA, were synchronized with α-factor in G1, and resuspended in YPD or YPD plus 100 mM CaCl2 at 25°C. Samples were taken at of 20-minute intervals after release from arrest. Cdc28 protein was used as an internal loading control. (B) Stability of Cln2p. Wild-type (YMM180), pkc1-834 (YMM179) and stt1-1 (YMM196) cells were cultured at 37°C for 2 hours; and then samples were taken at the indicated times after the addition of 100 μg/ml cycloheximide, and subjected to western bolt analysis. The asterisk indicates a non-specific band. (C) DNA content for the same samples as in A. (D) Effect of over-expression of CLN2 on the bud growth of the zds1Δ pkc1-834 strain. A strain containing a chromosomally integrated construct for GAL-regulated Cln2p-3xHA (YMM170-1; cln2::GAL1–CLN2HA–LEU2) were grown in the medium containing 2% raffinose (GAL1 promoter off) or 2% galactose (GAL1 promoter on) for 8 hours at 25°C.

In characterization of the pkc1-834 and stt1-1 mutations, the former mutation was found to involve an asparagine to aspartic acid substitution at amino acid position 834, and the latter, a proline to serine substitution at amino acid position 1102 of Pkc1p (Fig. 5A). The stt1-1 mutation (P1102S) was within the kinase domain in a highly conserved proline residue, whereas the pkc1-834 mutation (N834D) was within the predicted ATP-binding site of Pkc1p in an evolutionarily non-conserved residue. We constructed N834D and P1102S mutant versions of the PKC1 gene on a centromeric plasmid. The N834D and P1102S alleles could suppress any of the phenotypes of the pkc1-834 or stt1-1 mutation, respectively (data not shown). Furthermore, the pkc1Δ zds1Δ strain bearing the N834D or P1102S allele exhibited calcium phenotypes similar to those of the pkc1-834 zds1Δ or stt1-1 zds1Δ strain, respectively, confirming that each mutant allele is responsible for the observed phenotype (Fig. 5B).

It was previously shown that the growth of the pkc1-1 strain (on EG123 background) was dependent on the presence of exogenous calcium in the medium (Levin and Bartlett-Heubusch, 1992). Interestingly, the mutation site of pkc1-834 (N834D) coincided with that of this pkc1-1 mutation (N834K). The pkc1-834 mutant was able to grow on a YPD plate (without added CaCl2) at a rate slower than that of the wild-type strain, and the growth rate could be restored to an apparently normal level by the addition of CaCl2, but not MgCl2 (Fig. 1A and Fig. 5B; data not shown). Thus, the N834 residue seems to be involved somehow in the regulation of Pkc1p by calcium. Western blot analysis of Pkc1p in cell extracts of wild-type, pkc1-834 and stt1-1 strains, using an anti-Pkc1p antibody, demonstrated that the expression level of Pkc1-834p and of Stt1-1p was comparable to that of the wild-type Pkc1p (data not shown).

Fig. 4.

Pkc1p is required for the maintenance of Ca2+-induced polarized bud growth. The zds1Δ (YMM187) and zds1Δ pkc1-834 (YMM178) strains was grown in YPD medium to early-log phase at 25°C. The cells were synchronized in G1 with α-factor and then released into YPD medium with or without 100 mM CaCl2 at 25°C. Samples were taken at 20-minute intervals after removal of α-factor and fixed with formaldehyde. (A) F-actin staining with Rhodamine-phalloidin. (B) The cumulative percentage of cells that initiated bud formation (left) or the percentage of F-actin patches localized at the incipient bud site (right; actin patches localized at the second-cycle buds were excluded) is plotted for the same samples shown in A.

Fig. 4.

Pkc1p is required for the maintenance of Ca2+-induced polarized bud growth. The zds1Δ (YMM187) and zds1Δ pkc1-834 (YMM178) strains was grown in YPD medium to early-log phase at 25°C. The cells were synchronized in G1 with α-factor and then released into YPD medium with or without 100 mM CaCl2 at 25°C. Samples were taken at 20-minute intervals after removal of α-factor and fixed with formaldehyde. (A) F-actin staining with Rhodamine-phalloidin. (B) The cumulative percentage of cells that initiated bud formation (left) or the percentage of F-actin patches localized at the incipient bud site (right; actin patches localized at the second-cycle buds were excluded) is plotted for the same samples shown in A.

We further genetically characterized the defects of the two PKC1 mutant alleles. Homo- and hetero-allelic diploids were constructed and examined for their temperature sensitivities and resistance to calcium on the zds1Δ background. In the presence of 100 mM CaCl2 at 37°C, the homo-allelic diploids pkc1-834/pkc1-834 and stt1-1/stt1-1 grew poorly, whereas the hetero-allelic diploid (pkc1-834/stt1-1) was able to grow (all on the zds1Δ background) showing an intragenic complementation (Fig. 6A). This result indicated that the defects of the stt1-1 and pkc1-834 mutations were functionally different. In addition, various diploid strains, with respect to the PKC1 allele, exhibited a range of calcium resistance: (in decreasing order) pkc1-834/pkc1-834, pkc1-834/stt1-1, stt1-1/stt1-1 and PKC1/PKC1 (all on the zds1Δ background), suggesting that the wild-type Pkc1p inhibits growth in the presence of calcium and that the mutation pkc1-834 reduces the inhibitory effect more than stt1-1 (Fig. 6A).

We further compared the phenotypic differences of the pkc1-834 and stt1-1 alleles. To see if the pkc1-834 mutant exhibited similar defects to the stt1-1 mutant, we examined the cell-lysis phenotype by a simple plate-overlay assay using a plate containing BCIP (Saka et al., 2001). Although the stt1-1 colonies exhibited a cell-lysis phenotype at 35°C, the pkc1-834 colonies remained apparently normal, suggesting that the Pkc1p-Mpk1p pathway functioned adequately in pkc1-834 cells (Fig. 6B). To evaluate the activation of Mpk1p in Pkc1p mutants upon shift of the growth temperature from 25°C to 37°C, we determined the level of phospho-Mpk1p using a monoclonal antibody that specifically recognizes phospho-Mpk1p. Although, Mpk1p phosphorylation in stt1-1 cells was very low, Mpk1p was similarly phosphorylated in pkc1-834 and wild-type cells (Fig. 6C). Moreover, over-expression of MPK1 partially suppressed the temperature sensitivity of the stt1-1 strain, but not that of the pkc1-834 strain (Fig. 6D). Furthermore, both pkc1-834 mpk1Δ and stt1-1 mpk1Δ double mutations were lethal, and only the lethality caused by the pkc1-834 mpk1Δ double mutation was suppressed by 1 M sorbitol, suggesting that the pkc1-834 mutation has additional defect(s) in essential growth function(s) (data not shown). Based on these results, we concluded that pkc1-834 cells have an important defect in their Mpk1p-independent function(s).

Fig. 5.

Mutation sites in Pkc1p. (A) The pkc1-834 mutant allele carries a single nucleotide change (A2501G) that causes replacement of asparagine 834 with aspartic acid. The stt1-1 mutant allele carries a single nucleotide change (C3304T) that causes replacement of proline 1102 with serine. (B) The amino acids P1102 and N834 of Pkc1p are important for its function. pkc1Δ or zds1Δ pkc1Δ cells with low-copy plasmids, carrying either PKC1 (YCp50-PKC1), stt1-1 (YCp50-PKC1P1102S) or pkc1-834 (YCp50-PKC1N834D), were spotted on YPD plate with or without 300 mM CaCl2 and grown at the indicated temperatures for 2∼3 days.

Fig. 5.

Mutation sites in Pkc1p. (A) The pkc1-834 mutant allele carries a single nucleotide change (A2501G) that causes replacement of asparagine 834 with aspartic acid. The stt1-1 mutant allele carries a single nucleotide change (C3304T) that causes replacement of proline 1102 with serine. (B) The amino acids P1102 and N834 of Pkc1p are important for its function. pkc1Δ or zds1Δ pkc1Δ cells with low-copy plasmids, carrying either PKC1 (YCp50-PKC1), stt1-1 (YCp50-PKC1P1102S) or pkc1-834 (YCp50-PKC1N834D), were spotted on YPD plate with or without 300 mM CaCl2 and grown at the indicated temperatures for 2∼3 days.

Fig. 6.

Intragenic complementation between the pkc1-834 and stt1-1 mutations. (A) Diploid strains of ZDS1/ZDS1 (W303-1D), zds1Δ PKC1/zds1Δ PKC1 (YMM125), zds1Δ pkc1-834/zds1Δ pkc1-834 (YMM219), zds1Δ stt1-1/zds1Δ stt1-1 (YMM218), and zds1Δ pkc1-834/zds1Δ stt1-1 (YMM220) were spotted onto YPD plates with the indicated concentration of CaCl2 and incubated at the indicated temperatures. (B) Alkaline phosphatase colony assay of the various strains. Cells were spotted on YPD plates and cultured at 25°C for 2 days. The plates were further incubated overnight at 25°C or 35°C and alkaline phosphatase released from the cells as a result of cell lysis was detected using BCIP. The blue color (seen here as darker disks) indicates defective cell walls. The mpk1Δ (TNP46) strain was used as a positive control. (C) Mpk1 activation by heat-shock in various strains. Wild-type (DHT22-1b), pkc1-834 (YMM28) and stt1-1 (YMM114) cells were shifted from 25°C to 37°C, and incubated for 2 hours. The Mpk1p phosphorylation was monitored by western blotting. Immunoblot analysis was carried out using anti-phospho-p44/42 MAPK antibody (top panel) or anti-Mpk1p antibody to detect Mpk1p (bottom panel). (D) The temperature sensitivity of stt1-1 is partially suppressed by overexpression of MPK1. scz6 (YMM28) or stt1-1 (YMM114) cells with high-copy plasmids: control (YEp24), PKC1 (YEp24-PKC1) or MPK1 (YEp24-MPK1), were spotted on YPD plates and grown at the indicated temperatures for 2 days.

Fig. 6.

Intragenic complementation between the pkc1-834 and stt1-1 mutations. (A) Diploid strains of ZDS1/ZDS1 (W303-1D), zds1Δ PKC1/zds1Δ PKC1 (YMM125), zds1Δ pkc1-834/zds1Δ pkc1-834 (YMM219), zds1Δ stt1-1/zds1Δ stt1-1 (YMM218), and zds1Δ pkc1-834/zds1Δ stt1-1 (YMM220) were spotted onto YPD plates with the indicated concentration of CaCl2 and incubated at the indicated temperatures. (B) Alkaline phosphatase colony assay of the various strains. Cells were spotted on YPD plates and cultured at 25°C for 2 days. The plates were further incubated overnight at 25°C or 35°C and alkaline phosphatase released from the cells as a result of cell lysis was detected using BCIP. The blue color (seen here as darker disks) indicates defective cell walls. The mpk1Δ (TNP46) strain was used as a positive control. (C) Mpk1 activation by heat-shock in various strains. Wild-type (DHT22-1b), pkc1-834 (YMM28) and stt1-1 (YMM114) cells were shifted from 25°C to 37°C, and incubated for 2 hours. The Mpk1p phosphorylation was monitored by western blotting. Immunoblot analysis was carried out using anti-phospho-p44/42 MAPK antibody (top panel) or anti-Mpk1p antibody to detect Mpk1p (bottom panel). (D) The temperature sensitivity of stt1-1 is partially suppressed by overexpression of MPK1. scz6 (YMM28) or stt1-1 (YMM114) cells with high-copy plasmids: control (YEp24), PKC1 (YEp24-PKC1) or MPK1 (YEp24-MPK1), were spotted on YPD plates and grown at the indicated temperatures for 2 days.

The degradation of Hsl1p, a negative regulator of Swe1p, via the ubiquitin-mediated 26S proteasome pathway is triggered by the activation of calcineurin and the Mpk1p-Mck1p pathway (Mizunuma et al., 2001). Pkc1p is known as an upstream regulator of the Mpk1p MAPK pathway in cell wall construction. We expected that calcium-induced Hsl1p degradation, which is dependent on the activation of Mpk1p occurs normally in the pkc1-834 strain, but not in the stt1-1 strain. However, the level of Hsl1p in these strains, in contrast to that in the mpk1Δ strain, decreased more rapidly, in a similar manner as in wild-type strain (Fig. 7). Consistent with this observation, both pkc1-834 and stt1-1 alleles failed to suppress the Ca2+-induced G2 delay of the zds1Δ strain (Fig. 1C), suggesting that this effect is due to the activation of Swe1p. Thus, it was suggested that the Pkc1-834p and Stt1-1p mutant proteins are functional in regulating the Hsl1p degradation. Alternatively, it was also possible that lower activity of Mpk1p in the stt1-1 cells is still sufficient to trigger Hsl1p degradation.

The stt1-1 allele of PKC1 is functional in Cln2p elevation by Ca2+

It was previously indicated that the stt1-1 mutation causes a defect in the activation of Mpk1p MAPK by heat shock. Our observation that the stt1-1 mutation failed to suppress the Ca2+-induced hyper-polarized bud growth of the zds1Δ strain suggested that Mpk1p MAPK function may not be required for sustaining a high level of Cln2p in the presence of exogenous calcium (Fig. 1A,B). Therefore, we presumed that the stt1-1 mutation does not affect the maintenance of high Cln2p levels in the presence of calcium. As expected, no significant difference was detected in the expression levels of Cln2p between wild-type and stt1-1 strains, using western blot analysis (Fig. 3A). There was also no significant difference in the levels of CLN2 mRNA between stt1-1 and wild-type strains (Fig. 2). These results indicated that Pkc1p is required for the maintenance of the calcium-induced elevation of Cln2p levels in a manner independent of Mpk1p MAPK activation.

Fig. 7.

pkc1-834 and stt1-1 are not required for the regulation of Hsl1p abundance. Comparison of Hsl1p abundance in various strains. Wild-type (YMM53), pkc1-834 (YMM143), stt1-1 (YMM144) and mpk1Δ (YMM70) strains with a chromosomally integrated construct encoding genomic copies of 3xHA-Hsl1p were grown in YPD at 25°C until early log phase. CaCl2 was added to the cell cultures at the indicated time, to a final concentration of 100 mM, and samples were taken for western blot analysis.

Fig. 7.

pkc1-834 and stt1-1 are not required for the regulation of Hsl1p abundance. Comparison of Hsl1p abundance in various strains. Wild-type (YMM53), pkc1-834 (YMM143), stt1-1 (YMM144) and mpk1Δ (YMM70) strains with a chromosomally integrated construct encoding genomic copies of 3xHA-Hsl1p were grown in YPD at 25°C until early log phase. CaCl2 was added to the cell cultures at the indicated time, to a final concentration of 100 mM, and samples were taken for western blot analysis.

Rho1p GTPase is involved in the Ca2+-induced polarized bud growth

In the activation of the Mpk1p MAPK pathway, Pkc1p is a down-stream component of Rho1p, a small G protein molecular switch (Nonaka et al., 1995; Kamada et al., 1996; Helliwell et al., 1998; Delley and Hall, 1999). Because Rho1p is an essential protein, several conditional, lethal (high temperature-sensitive) mutations in the RHO1 gene have been isolated and characterized (Helliwell et al., 1998; Saka et al., 2001). Of these, rho1-2 and rho1-5 strains are defective in the activation of Mpk1p, whereas rho1-3 and rho1-4 mutants are not (Saka et al., 2001). We examined allele specificity of various rho1 mutations on the Ca2+-induced polarized bud growth of the zds1Δ strain. Consistent with the distinct role of Pkc1p in Ca2+-induced polarized bud growth, the rho1-3 and rho1-4 mutations, but not the rho1-2 and rho1-5 mutations, suppressed the calcium phenotypes of the zds1Δ strain, suggesting that the defect caused by the rho1-3 and rho1-4 alleles, rather than the rho1-2 and rho1-5 alleles, are related in having the novel Pkc1p function (Fig. 8). Consistent with the notion that the functional defects of the rho1-3 and rho1-4 alleles are related with that of the pkc1-834 allele (Fig. 3A), the levels of Cln2p in rho1-3 and rho1-4 strains and the pkc1-834 strain were similarly low compared with those of wild-type and rho1-2 and rho1-5 strains (data not shown). The typical effects of calcium, observed in zds1Δ mutants of W303 strain background, namely Ca2+-induced growth arrest, hyper-polarized bud growth and G2 delay, were also observed in the YPH500 strain background at calcium concentrations higher than those for the W303 strain. The pkc1-834 mutation, but not that of stt1-1, suppressed the Ca2+-induced hyper-polarized bud growth on this strain background (data not shown). Taken together, these data implicated the Mpk1p-independent novel pathway of Pkc1p in Ca2+-induced polarized bud growth in a manner dependent on Rho1p.

Fig. 8.

Rho1p GTPase is involved in the Ca2+-induced polarized bud growth. Effect of several rho1 mutant alleles on the cell morphology of the zds1Δ mutant strain. Wild-type (YOC729), zds1Δ (YMM230), zds1Δ rho1-2 (YMM232), zds1Δ rho1-3 (YMM147), zds1Δ rho1-4 (YMM148) and zds1Δ rho1-5 (YMM149) cells were used. Cell morphology after 6 hours of incubation with 300 mM CaCl2 at 25°C was examined.

Fig. 8.

Rho1p GTPase is involved in the Ca2+-induced polarized bud growth. Effect of several rho1 mutant alleles on the cell morphology of the zds1Δ mutant strain. Wild-type (YOC729), zds1Δ (YMM230), zds1Δ rho1-2 (YMM232), zds1Δ rho1-3 (YMM147), zds1Δ rho1-4 (YMM148) and zds1Δ rho1-5 (YMM149) cells were used. Cell morphology after 6 hours of incubation with 300 mM CaCl2 at 25°C was examined.

We have demonstrated a novel important role of S. cerevisiae protein kinase C, i.e. a role in the maintenance of polarized bud growth in a manner independent of Mpk1p MAPK activation. Thus, the strain with the pkc1-834 allele, which is defective in this function, is apparently normal in polarity establishment, but defective in the maintenance of polarized growth during budding and formation of the mating projection. The effect of calcium on cell-cycle progression and polarized bud growth are pronounced on a zds1Δ background lacking the negative regulator for SWE1 and CLN2 transcription, leading to the elevation of Cln2p and Swe1p (Mizunuma et al., 1998; Mizunuma et al., 2001; Mizunuma et al., 2004). Pkc1p was important for the elevation and maintenance of the Cln2p level, but not for that of Swe1p, indicating that the regulatory mechanisms for Swe1p and Cln2p are different.

The polarized bud growth and defect in cell proliferation, but not the G2 delay, caused by high calcium in the zds1Δ strain were abolished by the pkc1-834 mutation (Fig. 1). This suggested the possibility that the growth inhibition of zds1Δ cells by high calcium is due to the hyperpolarization of bud growth induced by high Cln2p levels. However, we consider this possibility unlikely for the following reasons. The growth of the cln2Δ zds1Δ strains was more sensitive to calcium than growth of the zds1Δ strain, although cln2Δ zds1Δ cells failed to show hyperpolarized bud growth in response to calcium (data not shown). Moreover, the rho1-3 and rho1-4 mutant strains (like the pkc1-834 strain) that expressed lower levels of Cln2p than the wild-type strain could not suppress calcium sensitivity of zds1Δ strain (data not shown). Furthermore, double mutant strains having the pkc1-834 mutation in combination with either of the rho1-2, -3, -4 or -5 mutations, exhibited more severe growth defect on YPD medium at a permissive temperature (25°C) than either of the single mutants, suggesting that rho1-2, 3, 4 and 5 mutations may have additional defect besides the defect in the activation of the Pkc1p (data not shown).

We have shown that the regulation of Cln2p expression (and probably Cln1p, as well) is mediated by Pkc1p at the posttranscriptional level (Fig. 3A,B). However, its molecular mechanism still remains to be elucidated. Thus, the level of Cln2p in the pkc1-834 strain, defective in the regulatory mechanism, was lower that in the wild-type strain, and Cln2p was more stable in the pkc1-834 strain (Fig. 3A,B). Since the phosphorylation of G1 cyclin, which is reported to be due to the Cdc28-dependent autophosphorylation of the Cln subunit, is required for their rapid degradation (Lanker et al., 1996), the stable nature of Cln2p in the pkc1-834 strain may be the result of the lower activity of Cln2p-Cdc28p in this strain. Alternatively, it is also possible that Pkc1p may regulate Cln2p at the posttranscriptional level.

An earlier study suggested that the activation of Cdc28p-Cln1/2p in G2 leads to hyperpolarization of F-actin and formation of elongated buds (Lew and Reed, 1993). Thus, the effect of the pkc1-834 mutation on the calcium-induced hyperpolarized bud growth of the zds1Δ strain can be explained by the reduced activity of the Cdc28p-Cln1/2p kinase, which is insufficient to fully promote F-actin polarization. During adaptation to higher temperature conditions, Pkc1p is required for both de- and repolarization of the actin cytoskeleton. The former is independent of the Mpk1p MAP kinase pathway, while the latter is dependent on it (Delley and Hall, 1999). By contrast, in the response to calcium, Pkc1p was not required for the de- and repolarization of F-actin (Fig. 4). Instead, Pkc1p was required for the maintenance of F-actin polarization (Fig. 4). We also found that Rho1p Rho GTPase is involved in Ca2+-induced polarized bud growth in a manner independent of Mpk1p. Pkc1p interacts with Rho1p in vivo and in vitro (Nonaka et al., 1995; Kamada et al., 1996). Pkc1p mutant protein (Pkc1-834p) of the pkc1-834 strain can physically interact with Rho1p, as revealed by the two-hybrid assay using LexA-Rho1p (Q68L, GTP-bound form) with GAD-Pkc1-834p (data not shown). Co-immunoprecipitation of Pkc1-834p and Rho1p in vivo was also observed (data not shown). The interaction of Rho1p and Pkc1-834p is consistent with the ability of Pkc1-834p to activate the Mpk1p pathway, leading to cell wall construction.

How does Pkc1p regulate the establishment and maintenance of polarized bud growth in zds1Δ strain? Earlier studies showed that the Pkc1p-Mpk1p pathway is activated at the G1/S transition, concomitant with bud emergence, and that its activation is partly dependent on the activation of the Cdc28p-Cln1/2p complex (Marini et al., 1996; Zarzov et al., 1996; Gray et al., 1997). The pkc1-834 zds1Δ cells were able to initiate bud formation, but were unable to sustain polarized bud growth by exogenous calcium (Fig. 4). Even the reduced levels of Cln1p and Cln2p in pkc1-834 zds1Δ cells appeared to be sufficient to establish polarized bud growth, but insufficient to support Ca2+-induced hyper-polarized bud growth. Our data indicated that Pkc1p is important for maintaining the Cln2p levels, placing the Pkc1p function upstream of Cln1p and Cln2p. Our results and those of others indicate that the Pkc1p-Mpk1p pathway is required for the establishment of polarized bud growth, while the Mpk1p-independent Pkc1p pathway is required for its maintenance.

How does Ca2+ signaling induce polarized bud growth and G2 cell-cycle delay? These may be achieved by maintaining a high level of the G1 cyclins and by triggering the Swe1p-mediated inhibition of Cdc28p-Clb1/2p. Of these, the former is dependent on the novel Pkc1p function, whereas the latter is mediated by a coordinated action of the calcineurin and Mpk1p-Mck1p pathways to activate Swe1p, which is achieved by elevating its abundance and downregulating Hsl1p (a Swe1p negative regulator) leading to the accumulation of the hyperphosphorylated, active form of Swe1p (Mizunuma et al., 2001). However, the possibility that Pkc1p is involved in the downregulation of Hsl1p could not be excluded, because the calcium-induced Hsl1p degradation in pkc1Δ strain could not be determined because the growth of this strain was extremely poor even in the presence of an osmo-stabilizer. Significantly, the swe1Δ pkc1-834 double mutant exhibits a severe growth defect even at the temperatures permissive for the pkc1-834 strain (data not shown), suggesting that Ca2+-mediated coordinated regulation of polar bud growth and cell-cycle is important for normal cell growth. It will be of considerable interest to study whether a similar mechanism operates in higher eukaryotes. A model summarizing our results is shown in Fig. 9 

Fig. 9.

Summary and model of novel Pkc1p pathway for polarized bud growth in budding yeast. The arrows and the blunt-ended line indicate positive and negative controls, respectively.

Fig. 9.

Summary and model of novel Pkc1p pathway for polarized bud growth in budding yeast. The arrows and the blunt-ended line indicate positive and negative controls, respectively.

We thank Yoshimi Takai, Kazuma Tanaka, Yoshikazu Ohya, and Tsutomu Kishi for gifts of plasmids or strains, as well as Hisako Nonaka, Keiko Ehara and Koichi Kirizawa for their technical help. This work was supported in part by Grant-in-Aid for Scientific Research from the Ministry of Education, Science, Culture, Arts and Sports of Japan to M.M. (14780546, 16770151) and T.M. (14206012, 17208009), and by a Research Grant of the Naito Foundation to M.M.

Adams, A. E. M. and Pringle, J. R. (
1991
). Staining of actin with fluorochrome-conjugated phalloidin.
Methods Enzymol.
194
,
729
-731.
Baetz, K., Moffat, J., Haynes, J., Chang, M. and Andrews, B. (
2001
). Transcriptional coregulation by the cell integrity mitogen-activated protein kinase Slt2 and the cell cycle regulator Swi4.
Mol. Cell. Biol.
21
,
6515
-6528.
Bi, E. and Pringle, J. R. (
1996
). ZDS1 and ZDS2, genes whose products may regulate Cdc42p in Saccharomyces cerevisiae.
Mol. Cell. Biol.
16
,
5264
-5275.
Booher, R. N., Deshaies, R. J. and Kirschner, M. W. (
1993
). Properties of Saccharomyces cerevisiae wee1 and its differential regulation of p34 CDC28 in response to G1 and G2 cyclins.
EMBO J.
12
,
3417
-3426.
Davenport, K. R., Sohaskey, M., Kamada, Y., Levin, D. E. and Gustin, M. C. (
1995
). A second osmosensing signal transduction pathway in yeast. Hypotonic shock activates the PKC1 protein kinase-regulated cell integrity pathway.
J. Biol. Chem.
50
,
30157
-30161.
Delley, P. A. and Hall, M. N. (
1999
). Cell wall stress depolarizes cell growth via hyperactivation of RHO1.
J. Cell Biol.
147
,
163
-174.
Dodou, E. and Treisman, R. (
1997
). The Saccharomyces cerevisiae MADS-box transcription factor Rlm1 is a target for the Mpk1 mitogen-activated protein kinase pathway.
Mol. Cell. Biol.
17
,
1848
-1859.
Drubin, D. G. and Nelson, W. J. (
1996
). Origins of cell polarity.
Cell
84
,
335
-344.
Gray, J. V., Ogas, J. P., Kamada, Y., Stone, M., Levin, D. E. and Herskowitz, I. (
1997
). A role for the Pkc1 MAP kinase pathway of Saccharomyces cerevisiae in bud emergence and identification of a putative upstream regulator.
EMBO J.
16
,
4924
-4937.
Gustin, M. C., Albertyn, J., Alexander, M. and Davenport, K. (
1998
). MAP kinase pathways in the yeast Saccharomyces cerevisiae. Microbiol.
Mol. Biol. Rev.
62
,
1264
-1300.
Heinisch, J. J., Lorberg, A., Schmitz, H. P. and Jacoby, J. J. (
1999
). The protein kinase C-mediated MAP kinase pathway involved in the maintenance of cellular integrity in Saccharomyces cerevisiae.
Mol. Microbiol.
32
,
671
-680.
Helliwell, S. B., Schmidt, A., Ohya, Y. and Hall, M. N. (
1998
). The Rho1 effector Pkc1, but not Bni1, mediates signalling from Tor2 to the actin cytoskeleton.
Curr. Biol.
8
,
1211
-1214.
Igual, J. C., Johnson, A. L. and Johnston, L. H. (
1996
). Coordinated regulation of gene expression by the cell cycle transcription factor SWI4 and the protein kinase C MAP kinase pathway for yeast cell integrity.
EMBO J.
15
,
5001
-5013.
Kamada, Y., Sung Jung, U., Piotrowski, J. and Levin, D. E. (
1995
). The protein kinase C-activated MAP kinase pathway of Saccharomyces cerevisiae mediates a novel aspect of the heat shock response.
Genes Dev.
9
,
1559
-1571.
Kamada, Y., Qadota, H., Python, C. P., Anraku, Y. and Levin, D. E. (
1996
). Activation of yeast protein kinase C by Rho1 GTPase.
J. Biol. Chem.
271
,
9193
-9196.
Lanker, S., Valdivieso, M. H. and Wittenberg, C. (
1996
). Rapid degradation of the G1 cyclin Cln2 induced by CDK-dependent phosphorylation.
Science
271
,
1597
-1601.
Levin, D. E. and Bartlett-Heubusch, E. (
1992
). Mutants in the S. cerevisiae PKC1 gene display a cell cycle-specific osmotic stability defect.
J. Cell Biol.
116
,
1221
-1229.
Levin, D. E. and Errede, B. (
1995
). The proliferation of MAP kinase signaling pathways in yeast.
Curr. Opin. Cell Biol.
7
,
197
-202.
Levin, D. E., Fields, F. P., Kunisawa, R., Bishop, J. M. and Thorner, J. (
1990
). A candidate protein kinase C gene, PKC1, is required for the S. cerevisiae cell cycle.
Cell
62
,
213
-224.
Lew, D. J. and Reed, S. I. (
1993
). Morphogenesis in the yeast cell cycle: Regulation by Cdc28 and cyclins.
J. Cell Biol.
120
,
1305
-1320.
Lew, D. J. and Reed, S. I. (
1995
). A cell cycle checkpoint monitors cell morphogenesis in budding yeast.
J. Cell Biol.
129
,
739
-749.
Ma, X. J., Lu, Q. and Grunstein, M. (
1996
). A search for proteins that interact genetically with histone H3 and H4 amino termini uncovers novel regulators of the Swe1 kinase in Saccharomyces cerevisiae.
Genes Dev.
10
,
1327
-1340.
Madden, K., Sheu, Y. J., Baetz, K., Andrews, B. and Snyder, M. (
1997
). SBF cell cycle regulator as a target of the yeast PKC-MAP kinase pathway.
Science
275
,
1781
-1784.
Marini, N. J., Meldrum, E., Buehrer, B., Hubberstey, A. V., Stone, D. E., Traynor-Kaplan, A. and Reed, S. I. (
1996
). A pathway in the yeast cell division cycle linking protein kinase C (Pkc1) to activation of Cdc28 at START.
EMBO J.
15
,
3040
-3052.
Mazzoni, C., Zarzov, P., Rambourg, A. and Mann, C. (
1993
). The Slt2 (Mpk1) MAP kinase homolog is involved in polarized cell growth in Saccharomyces cerevisiae.
J. Cell Biol.
123
,
1821
-1833.
McMillan, J. N., Sia, R. A. and Lew, D. J. (
1998
). A morphogenesis checkpoint monitors the actin cytoskeleton in yeast.
J. Cell Biol.
142
,
1487
-1499.
Mizunuma, M., Hirata, D., Miyahara, K., Tsuchiya, E. and Miyakawa, T. (
1998
). Role of calcineurin and Mpk1 in regulating the onset of mitosis in budding yeast.
Nature
392
,
303
-306.
Mizunuma, M., Hirata, D., Miyaoka, R. and Miyakawa, T. (
2001
). GSK-3 kinase Mck1 and calcineurin coordinately mediate Hsl1 down-regulation by Ca2+ in budding yeast.
EMBO J.
20
,
1074
-1085.
Mizunuma, M., Miyamura, K., Hirata, D., Yokoyama, H. and Miyakawa, T. (
2004
). Involvement of S-adenosylmethionine in G1 cell-cycle regulation in Saccharomyces cerevisiae.
Proc. Natl. Acad. Sci. USA
101
,
6086
-6091.
Nonaka, H., Tanaka, K., Hirano, H., Fujiwara, T., Kohno, H., Umikawa, M., Mino, A. and Takai, Y. (
1995
). A downstream target of RHO1 small GTP-binding protein is PKC1, a homolog of protein kinase C, which leads to activation of the MAP kinase cascade in Saccharomyces cerevisiae.
EMBO J.
14
,
5931
-5938.
Paravicini, G., Cooper, M., Friedli, L., Smith, D. J., Carpentier, J. L., Klig, L. S. and Payton, M. A. (
1992
). The osmotic integrity of the yeast cell requires a functional PKC1 gene product.
Mol. Cell. Biol.
12
,
4896
-4905.
Pruyne, D. and Bretscher, A. (
2000
). Polarization of cell growth in yeast.
J. Cell Sci.
113
,
571
-585.
Saka, A., Abe, M., Okano, H., Minemura, M., Qadota, H., Utsugi, T., Mino, A., Tanaka, K., Takai, Y. and Ohya, Y. (
2001
). Complementing yeast rho1 mutation groups with distinct functional defects.
J. Biol. Chem.
276
,
46165
-46171.
Sia, R. A., Bardes, E. S. and Lew, D. J. (
1998
). Control of Swe1p degradation by the morphogenesis checkpoint.
EMBO J.
17
,
6678
-6688.
Yoshida, S., Ikeda, E., Uno, I. and Mitsuzawa, H. (
1992
). Characterization of a staurosporine- and temperature-sensitive mutant, stt1, of Saccharomyces cerevisiae: STT1 is allelic to PKC1.
Mol. Gen. Genet.
231
,
337
-344.
Yu, Y., Jiang, Y. W., Wellinger, R. J., Carlson, K., Roberts, J. M. and Stillman, D. J. (
1996
). Mutation in the homologous ZDS1 and ZDS2 genes affect cell cycle progression.
Mol. Cell. Biol.
16
,
5254
-5263.
Zarzov, P., Mazzoni, C. and Mann, C. (
1996
). The Slt2 (Mpk1) MAP kinase is activated during periods of polarized cell growth in yeast.
EMBO J.
15
,
83
-91.