15 March 2000

EFG1 Null Mutants of Candida albicansSwitch but Cannot Express the Complete Phenotype of White-Phase Budding Cells

ABSTRACT

The Candida albicans gene EFG1 encodes a putative trans-acting factor. In strain WO-1, which undergoes the white-opaque transition, EFG1 is transcribed as a 3.2-kb mRNA in white-phase cells and a less-abundant 2.2-kb mRNA in opaque-phase cells. cDNA sequencing and 5′ rapid amplification of cDNA ends analysis demonstrate that the major difference in molecular mass of the two transcripts is due to different transcription start sites. EFG1 null mutants form opaque-phase colonies and express the opaque-phase cell phenotype at 25°C. When shifted from 25 to 42°C, mutant opaque-phase cells undergo phenotypic commitment to the white phase, which includes deactivation of the opaque-phase-specific gene OP4 and activation of the white-phase-specific gene WH11, as do wild-type opaque-phase cells. After the commitment event, EFG1 null mutant cells form daughter cells which have the smooth (pimpleless) surface of white-phase cells but the elongate morphology of opaque-phase cells. Taken together, these results demonstrate thatEFG1 expression is not essential for the switch event per se, but is essential for a subset of phenotypic characteristics necessary for the full expression of the phenotype of white-phase cells. These results demonstrate that EFG1 is not the site of the switch event, but is, rather, downstream of the switch event.
Candida albicans and related species are capable of switching between a number of general phenotypes that can be distinguished by colony morphology (18, 29, 30, 31). Switching has been demonstrated at sites of commensalism (31) and infection (34, 35). In addition, infecting strains exhibit higher average switching frequencies than commensal strains (12), and isolates causing deep mycoses exhibit higher average switching frequencies than isolates causing superficial mycoses (14). Switching can affect a variety of virulence factors (1, 2, 13, 15, 24, 46, 47; K. Vargas and D. R. Soll, unpublished data). It was, therefore, no surprise to find that switching in C. albicans regulates expression of a number of phase-specific genes in a combinatorial fashion, including the white-phase-specific gene WH11(40), the opaque-phase-specific gene OP4(22, 23), the secreted aspartyl proteinase genesSAP1 and SAP3 (13, 22, 24, 47), the drug resistance gene CDR3 (5), and the two-component regulator gene CaNIK1 (41), and that switching in Candida glabrata regulates the expression of the metallothionein gene MT-II and the newly discovered hemolysin gene HLP (18). It has, therefore, been suggested that switching represents a mechanism for phenotypic plasticity that allows C. albicans and related species to rapidly adapt to environmental challenges in both the commensal and the pathogenic states (25, 31-33).
Using the white-opaque transition of C. albicans as a model experimental system, it was recently demonstrated that white-phase-specific expression of the gene WH11 was regulated through two unique upstream activation sequences and that white-phase-specific complexes formed between the two activation sequences and white-phase-cell extracts (37, 42). It was also demonstrated that opaque-phase-specific expression of the geneOP4 was regulated primarily through a MADS box consensus sequence (20). Therefore, phase-specific genes appear to be regulated by phase-specific transacting factors (32, 33). Recently, the gene EFG1 was cloned from C. albicans (19, 43). EFG1 encodes a protein homologous to a number of transcription factors that have been demonstrated to be involved in the regulation of morphogenesis inSaccharomyces cerevisiae, Aspergillus nidulans, and Neurospora crassa (4, 11, 21). Reduced levels of EFG1 expression suppressed hypha formation but not pseudohypha formation (43), and an efg1/efg1double mutant formed hyphae that were morphologically distinguishable from those of parental strains (19). In the white-opaque transition in strain WO-1, EFG1 was reported to be transcribed only in the white phase (36). Overexpression ofEFG1 in strain WO-1 stimulated opaque-phase cells to switch to the white phase and reduced expression of EFG1 in strain CAI8 resulted in a cell phenotype that was elongate like opaque-phase cells of strain WO-1, but lacked opaque-phase cell pimples (36). Taken together, these results suggested thatEFG1 played a role in the white-opaque transition. To directly assess the role of EFG1 in the white-opaque transition, we have reexamined the expression of this gene and have disrupted both alleles of the gene in strain WO-1 by using a urablast protocol (9) in a newly generatedura3 strain of WO-1.

MATERIALS AND METHODS

Maintenance of stock cultures.

C. albicans wild-type strain WO-1 (30) was maintained on agar containing modified Lee's medium (6). Strain Red 3/6, an ade2auxotroph (38), and strain TS3.3, a ura3auxotroph (Table 1), were maintained on agar containing modified Lee's medium supplemented with 0.6 mM adenine and 0.01 mM uridine, respectively. EFG1 mutant strains were maintained on agar containing modified Lee's medium.
Table 1.
Table 1. Genotypes of strains used in this study
Strain Relevant genotype Source or reference
WO-1 Wild type 29
Red 3/6 ade2/ade2 37
TS3 ade2/ade2 Δura3::ADE2/URA3 This study
TS3.3 ade2/ade2 Δura3::ADE2/Δura3::ADE2 This study
Ef3.1 ade2/ade2 Δura3::ADE2/Δura3::ADE2/Δefg1::hisG-URA3-hisG/EFG1 This study
Ef3.1.1 ade2/ade2 Δura3::ADE2/Δura3::ADE2/Δefg1::hisG/EFG1 This study
Efc20 ade2/ade2 Δura3::ADE2/Δura3::ADE2/Δefg1::hisG/Δefg1::CAT-URA3-CAT This study
Efc25 ade2/ade2 Δura3::ADE2/Δura3::ADE2/Δefg1::hisG/Δefg1::CAT-URA3-CAT This study
Efc30 ade2/ade2 Δura3::ADE2/Δura3::ADE2/Δefg1::hisG/Δefg1::CAT-URA3-CAT This study

Isolation of the EFG1 gene.

We originally set out to clone gene homologs in C. albicans of the APSES family of transcription factors (4) that included Phd1p (11), StuAp (21), and Sok2p (48). Two degenerate primers, P1 and P2, spanning common coding regions derived from Phd1p (11), StuAp (21), and Sok2p (48), were used to amplify a DNA fragment of approximately 380 bp encompassing the conserved region of these genes. The PCR-derived fragment was used to screen a λEMBL3A genomic library ofC. albicans WO-1 (40). Of approximately 50,000 plaques screened, 50 putative lambda clones were identified. Southern analysis with the DNA probe was used to select two lambda clones, λ14.1 and λ39.1, which contained approximately 10 and 12 kb of insert DNA, respectively. Partial sequence analysis demonstrated that both contained the EFG1 open reading frame (ORF) and flanking sequences. To isolate the 5′ flanking region ofEFG1, the λ14.1 and λ39.1 clones were screened by a gene-walking strategy by using the primer EC1 in combination with either the lambda left-arm-specific primer ELA or the lambda right-arm-specific primer ERA (Table 2). DNA fragments encompassing the 5′ upstream region of EFG1were obtained from λpH14.1 and λpH39.1 by PCR with the primers EC8, a sequence spanning −1 to −21 bp of EFG1, and ELA (Table2). The fragments were cloned into pGEM-5Z at the EcoRV site, generating the plasmids pTET14.1 and pB69.11, respectively. Both were sequenced in both directions by using an ABI model 373A automatic sequencing system and fluorescent Big Dye terminator chemistry (Perkin-Elmer–Applied Biosystems Inc., Foster City, Calif.).
Table 2.
Table 2. Primers used in this study
Name Sequence
   +239             +229
EC1 5′-CCAGTCTGTTGACCTGGTTGT-3′
           +1               +20
EC2 5′-TTTCTGCAGATGTCAACGTATTCTATACC-3′
        Pst1
ELA 5′-ATACTGTGATGCCATGGTGT-3′
ERA 5′-GCCAGTTATCTGGGCTTAAA-3′
               −1               −21
EC8 5′-ATGCATBBCTGCAGTAATATGGGTTATATTCTTGG-3′
     NsiI      PstI
EC3 5′-AAACTCGAGTGGATTTGGGAGAAGTTATG-3′
       XhoI
           +1            +18
FANEFG15′ 5′-GCGTCGCGAATGTCAACGTATTCTATA-3′
       NruI
           +1,162             +1,639
FANEFG3′m 5′-GCGCCGCGGCTTTTCTTCTTTGGCAACAGTCGT-3′
       SacII

Northern blot analysis.

Total RNA was extracted by methods previously described (41). Poly(A)+ mRNA was extracted by using the Oligotex Spin Column Kit according to manufacturer's specifications (Qiagen Inc., Santa Clarita, Calif.). RNA was separated in agarose formaldehyde gels, transferred to Zetabind nylon membranes, and hybridized with the appropriate probes (17). Prehybridization and hybridization procedures were performed by the methods of Church and Gilbert (8). Autoradiography was performed by exposing membranes at −70°C by using intensifying screens and Kodak XAR film (Eastman Kodak Co., Rochester, N.Y.). To measure the fold difference in transcript levels, a Northern blot containing a series of concentrations of total cell RNA ranging from 0.05 to 5.00 μg was hybridized with a radiolabeledEFIα2 gene probe (44), and the autoradiogram was digitized into the DENDRON program database (Solltech Inc., Oakdale, Iowa). Band intensities were measured and then used to generate a plot of measured intensity versus RNA concentration. Fold differences between Northern blot hybridization bands were then computed from the standards plot.

Southern blot analysis.

To confirm the configurations of either the EFG1 or URA3 locus in heterozygotes and null mutants by Southern blot analysis, DNA was extracted by methods previously described (24, 38). DNA (3 μg) was digested with the restriction enzymes described in Results, and the resulting fragments were separated in a 0.8% (wt/vol) agarose gel. The fragments were transferred to Hybond-N nylon membrane (Amersham International, Little Chalfont, Buckinghamshire, England) and hybridized with the DNA probes described in Results. Prehybridization and hybridization procedures were performed by the methods of Church and Gilbert (8), and autoradiography was performed as described above for Northern analyses.

Construction of cDNA libraries and cloning of phase-specificEFG1 transcripts.

Five micrograms of white- and opaque-phase poly(A)+ mRNA were converted into cDNA pools by using the Superscript Choice system and were directionally cloned into ZipLox lambda vector according to the manufacturer's specifications (Life Technologies, Gaithersburg, Md.). The titers of the unamplified white- and opaque-phase-specific libraries were 1 × 106 and 2 × 106, respectively. Approximately 105 plaques of each unamplified library were screened, using the PCR-derived 1.7-kb EFG1 ORF as a probe. Thirty independent clones were identified in each case. Each clone was subjected to a second screen. The DNA from each of 10 positive white- and opaque-phase lambda clones obtained in the second screen was digested with EcoRI, and Southern blots were probed with theEFG1 ORF to confirm that they encoded EFG1. The three largest white- and opaque-phase ZipLox clones were chosen and converted into plasmid derivatives according to the manufacturer's protocol (Life Technologies). The three white-phase-specific cDNAs were named EFW1.1, EFW2.1, and EFW4.1, and the three opaque-phase-specific cDNAs were named EFO1.1, EFO3.1, and EFO5.1. The six selected cDNAs, each larger than 2 kb, were sequenced in both directions as previously described.

5′ RACE analysis of white- and opaque-phase mRNAs.

To compare the 5′ untranslated sequences of white- and opaque-phaseEFG1 mRNAs, 1 μg of poly(A)+ mRNA from each phase was subjected to 5′ rapid amplification of cDNA ends (RACE) analysis using the 5′ RACE kit protocol (Life Technologies). For first-strand cDNA synthesis, two different EFG1-specific primers were used in order to confirm that the derived products were from the same mRNA species. The first-strand-specific primers were EC3, spanning the 3′ end of the EFG1 mRNA, 140 bp downstream from two tandem TAA stop codons, and EC1, spanning the 5′ end of theEFG1 mRNA 210 bp downstream of the ATG start codon (Table2). Following first-strand synthesis and dG tailing, double-stranded 5′ RACE products were generated by a high-fidelity PCR protocol (Roche Biochemicals, Indianapolis, Ind.) by using either the ECI primer for EC3-derived template, or the EC8 primer for EC1-derived template (Table 2). After confirming by Southern analysis that the 5′ RACE products were of predicted molecular sizes, we subcloned these products into pGEM-5Z (Promega Corp., Madison, Wis.) at theEcoRV site in order to determine their nucleotide sequences. Plasmid clones were named pB59W.11 and pB870.1 for white-phase- and opaque-phase-specific 5′ RACE inserts, respectively. Sequences of the two cloned 5′ RACE inserts were determined as described in a previous section.

Construction and analysis of RLUC EFG15′-untranslated region transcriptional fusions.

The 1.2-kb 5′-upstream region of the EFG1 ORF was inserted at the multiple cloning site immediately upstream of the Renillaluciferase (RLUC) ORF in the reporter plasmid pCRW3 (38). Integration was targeted to the ADE2 locus by linearizing the plasmid at an NsiI site in the ADE2 gene or to the EFG1 locus by linearizing at an HpaI site in the EFG1 promoter. Linearized plasmids were used to transform strain Red 3/6 by using the lithium acetate method (27). Five transformants of each construct were analyzed by Southern blot hybridization to confirm both the site of insertion and multiplicity of integration. Transformant clones harboring a single copy of the targeted plasmid in the correct location were chosen for further analysis. Measurements of RLUC activity were made according to methods previously described (38).

Construction of a ura3 derivative of strain WO-1.

The original plasmid p1164 containing the C. albicans URA3 gene was kindly provided by Stewart Scherer of Acacia Biosciences, Richmond, Calif. This plasmid contained a 4.2-kb DNA insert, which included at least 1 kb of DNA flanking theURA3 gene. The DNA fragment was subcloned at theEcoRV site of pGEM-5Z (Promega Corporation), and the resulting plasmid clone was designated p161. In order to construct aURA3 deletion cassette, p161 plasmid DNA was digested withEcoRV and XbaI to delete a central 2.0-kb fragment of the URA3 gene. Following digestion, the plasmid was end-repaired with T4 DNA polymerase and was treated with shrimp alkaline phosphatase. The deleted URA3 gene fragment of the plasmid was then replaced by a 2.4-kb EcoRV fragment of theC. albicans ADE2 gene. The ADE2 DNA was derived from pMC2 (16) as an EcoRV fragment and was subcloned into pGEM-5Z to derive pADE2-5Z. The plasmid containing the URA3 deletion cassetteURA3::ADE2 was designated p161ΔURA3:ADE2. Homozygous deletion of theURA3 gene was performed in strain Red 3/6, anade2 derivative of WO-1 (39). Approximately 50 μg of ApaI/SacI-digested p161ΔURA3:ADE2 DNA was used to transform Red 3/6 by the lithium acetate method (27).ADE2 prototrophic clones were chosen based on their capacity to grow on minimal medium containing no adenine sulfate.URA3 heterozygotes were identified by Southern analysis of genomic DNA probed successively with a 2.4-kb EcoRV fragment of the ADE2 gene (16) and the 0.34-kbXbaI-NsiI fragment of the URA3 gene from the p161 plasmid. Selected heterozygote(s) were subjected to a second round of transformation in order to increase the chances of obtaining homozygotes by integrative gene conversion rather than mitotic recombination. However, because the cassettes used in the first and second transformations were identical, we could not discriminate between the two mechanisms. To generate ura3homozygotes, heterozygous clones were grown to mid-log phase, then were inoculated into fresh yeast extract-peptone-dextrose broth and allowed to grow for one generation. Approximately 4 × 107cells were transformed with 50 μg of the URA3 deletion cassette DNA by the spheroplast method (38). Following transformation, 107 spheroplasts were spread on yeast extract-peptone-dextrose plates containing 1 M sorbitol for 16 h at 30°C. Cells were collected, and approximately 107cells were spread on minimal medium containing 1 mg of 5-fluororotic acid (FOA) per ml and 0.1 mM uridine. These plates were incubated for 4 to 5 days at 30°C for the appearance of FOA-resistant colonies. Thirty independent colonies were tested by Southern analysis for homozygosity of the URA3 locus by using the ADE2and URA3 probes employed in the analysis of heterozygotes. Of 24 clones exhibiting identical patterns and absence of theURA3 region spanning the XbaI-EcoRV restriction sites (9), two clones, TS3.3 and TS3.5, which underwent the white-opaque transition at normal frequencies and were incapable of growing in medium lacking uridine were selected.

Construction of homozygous EFG1 deletion strains.

A hisG-URA3-hisG-based cassette (9) was used to create an EFG1 heterozygote in the first round of transformation. To accomplish this, the expression plasmid containing the EFG1 ORF, pEF1α2:EFG1 (Fig.1B), was digested with DraIII to delete 740 bp of the EFG1 ORF (Fig. 1A), was end-repaired with T4 DNA polymerase, and was dephosphorylated by using shrimp alkaline phosphatase. The deleted fragment was substituted with the 4.0-kb hisG-URA3-hisG cassette from pMB7 (9) (Fig. 1B). The hisG-URA3-hisG cassette was prepared by digesting the pMB7 plasmid DNA with SalI/BglII, followed by end-repair using T4 DNA polymerase. Since no suitable restriction enzyme sites flanked the deletion cassette for isolation from the plasmid, a high-fidelity, long PCR protocol (Boehringer Mannheim, Indianapolis, Ind.) with the EFG1-specific primers EC2 and EC3 (Table 2) was used to isolate the cassette for transformation. Approximately 10 μg of PCR-generated cassette DNA was used to transform the ura3 strain TS3.3 (Table 1). All of the recovered clones from selection plates were tested for heterozygosity by digesting the total genomic DNA withBamHI, followed by Southern blot hybridization with theEFG1 ORF and a URA3 gene fragment. After confirming heterozygosity of the targeted EFG1 locus, the clones were subjected to 5-FOA treatment in order to induce “pop outs” of the URA3 gene. For the second round of transformation, a different disruption cassette was constructed. DNA of plasmid pB21.2, containing the EFG1 ORF (Fig. 1C), was digested with NsiI to remove 620 bp of DNA, end-repaired with T4 DNA polymerase, dephosphorylated with shrimp alkaline phosphatase, and substituted with approximately 3.5 kb of a CAT-URA3-CAT cassette from the plasmid pCUC (a gift from Paul T. Magee, University of Minnesota). CAT-URA3-CAT DNA was prepared by digesting pCUC plasmid DNA with BamHI and was end-repaired with T4 DNA polymerase. The derived plasmid was named pB45.1 (Fig. 1C). The disruption cassette used for the second round of transformation was obtained by digesting pB45.1 with the restriction enzyme DraIII, resulting in a digestion fragment with homologous ends that could only target to the functional undeleted chromosomal allele (Fig. 1C). Clones recovered on selection plates were tested for homozygosity by Southern analysis as described earlier.
Fig. 1.
Fig. 1. Map of the EFG1 locus and EFG1gene disruption cassettes. (A) Restriction map of the EFG1locus with the EFG1 ORF represented as a striped box. The start ATG codon is at 1 bp and the stop codon TAATAA is at 1,809 bp. (B) The EFG1 gene fragment used to create thehisG-URA3-hisG-based deletion cassette. ATG and TAATAA represent the start and stop codons, respectively, of Efg1p. EC2 and EC3 represent the two primers used to amplify the fragment to derive the pEF1α2:EFG1 plasmid and to amplify the EFG1disruption cassette for transformation. pA42.23 is the plasmid derivative of pEF1α2:EFG1 harboring thehisG-URA3-hisG module. The dark shaded region represents the deleted region of EFG1. (C) The EFG1 gene fragment used to create the CAT-URA3-CAT disruption cassette. In this cassette, a region spanning the NsiI restriction sites that is presented as a dark shaded region is deleted and replaced by the CAT-URA3-CAT module. In order to increase the efficiency of deleting the second copy of EFGI, the DraIII-digested fragment of pB45.1 was used for transformation.

Scanning electron microscopy.

Cells were washed in double-distilled H2O, fixed in 2.5% (wt/vol) gluteraldehyde in 0.1 M cacodylate buffer for 1 h, and postfixed in 1% osmium tetroxide in 0.1 M cacodylate buffer for 50 min. Cells were then washed three times in 0.1 M cacodylate buffer and treated with 6% thiocarbohydrazide at room temperature, followed by a second fixation in 1% osmium tetroxide to enhance surface architecture. Cells were then rinsed in double-distilled water, dehydrated through a graded series of ethanol solutions, dried, mounted on aluminum stubs, and sputter-coated with gold palladium. Samples were scanned with a Hitachi S-4000 scanning electron microscope (Hitachi Corp., San Diego, Calif.).

RESULTS

Northern analysis of EFG1 transcription in white- and opaque-phase cells.

Northern blots of total cellular RNA from white- and opaque-phase cells grown at 25°C were probed with a full-length EFG1 ORF DNA fragment. White-phase-cell RNA contained an EFG1 transcript of approximately 3.2 kb (Fig. 2A). Opaque-phase-cell RNA contained a single, less-abundant EFG1 transcript of approximately 2.2 kb (Fig. 3A). This result was obtained in every case with five additional RNA preparations from independent white- and opaque-phase clones. Because the 3.2-kb hybridization band possessed a hybridization tail in every one of the six northern blot hybridizations performed with white-phase RNA, the possibility existed that a faint 2.2-kb band may have been missed in Northern blots of white-phase RNA. This possibility was resolved when Northern blots of purified poly(A)+ mRNA were probed with the full-length EFG1 ORF DNA fragment. White-phase cell poly(A)+ RNA contained a 3.2-kb transcript and no 2.2-kb transcript, while opaque-phase cells contained a less-abundant 2.2-kb transcript and no 3.2-kb transcript (Fig. 2B). The ratio of the white-phase EFG1 3.2-kb transcript to the opaque-phase EFG1 2.2-kb transcript was estimated by densitometric analyses to be approximately 20 to 1 for both total RNA and poly(A)+ RNA.
Fig. 2.
Fig. 2. EFG1 transcript levels in white- and opaque-phase cells. Northern blots containing approximately 5 μg of either total RNA or 200 ng of poly(A)+ mRNA from white- (Wh) or opaque-phase (Op) cells were probed with the full-lengthEFG1 ORF derived with the primers FANEFG15′ and FANEFG3′m (Table 2). Following autoradiography, the blot was stripped and reprobed with the EF1α2 ORF.
Fig. 3.
Fig. 3. The nucleotide sequence of the 5′-untranslated transcribed region of EFGI. The sequences of both white- and opaque-phase EFGI 5′ RACE products of white- and opaque-phase cells were individually determined. The nucleotide sequences of the 5′ RACE products were compared with the sequences derived from the genomic clones. The 5′ ends of the white- and opaque-phase specific EFGI mRNAs are denoted as WEFGI and OEFGI, respectively. Two TATA box binding protein recognition motifs are shown as TBP-box1 and TBP-box2. The presence of a unique binding site for the MATα2 homeobox protein is also shown. Base pair position is shown on the left.

WO-1 contains one copy of EFG1.

Since Northern blots of poly(A)+ mRNA probed with EFG1revealed white- and opaque-phase EFG1 transcripts of different molecular sizes, the possibility was entertained thatC. albicans contained more than one EFG1 gene. The Southern blot hybridization patterns of DNA digested withBamHI, BglII, HindIII,NsiI, and XhoI (data not shown) were consistent with the physical map of the cloned EFG1 locus (Fig. 1A), suggesting that only one copy of EFGI existed in the WO-1 genome, as previously demonstrated for C. albicans strain CAI4 (19).

White- and opaque-phase EFG1 cDNA sequences and 5′ RACE products support one gene and two transcripts.

Two clones from a white-phase-specific cDNA library and two clones from an opaque-phase-specific cDNA library that were greater than 2 kb were isolated and sequenced. All four contained identical ORFs of 1,662 bp beginning with ATG, and two consecutive TAA stop codons. The ORF contained 554 amino acids, compared to 552 reported earlier (43, 44), and exhibited six additional amino acid differences. The two independent white-phase cDNA clones contained long untranslated 5′ regions of 370 and 380 bp, respectively. The length of the untranslated 3′ regions of both white-phase cDNA clones was 407 bp, and the lengths of the poly(A)+ stretches were 60 and 70 bp, respectively. The two independent opaque-phase cDNA clones contained short untranslated 5′ regions of 40 and 50 bp, respectively, which contrasted with the long 5′ regions of the white-phase cDNAs. The length of the untranslated 3′ region of both opaque-phase cDNA clones was 407 bp, identical to that of the white-phase cDNA clones. However, the poly(A)+ stretches were 15 and 20 bp, approximately one-third that of white-phase cDNA clones. Although this cDNA analysis demonstrated a moderate length difference at the 5′ untranslated ends of white- and opaque-phase EFG1 transcripts, the difference was not sufficient to account for the size differences of the transcripts demonstrated in Northern blots (Fig. 2).
To compare more precisely the 5′ ends of the white- and opaque-phaseEFG1 transcripts, 5′ RACE analysis (10) was performed with purified poly(A)+-containing white-phase and opaque-phase mRNA. Sequence analysis of the 5′ RACE products of two independent white-phase RNA preparations revealed 5′ untranslated regions of 1,126 and 1,173 bp in length. The comparable 1,126-bp sequences were identical. Based upon these lengths, the predicted size of the white-phase EFG1 mRNA was approximately 3.3 kb, close to the size estimated from the Northern blots in Fig. 2. Sequence analysis of the 5′ RACE products of two independent opaque-phase RNA preparations revealed 5′ untranslated regions of 145 and 162 bp in length. The comparable 145-bp sequences were identical. Based upon these lengths, the predicted size of the opaque-phaseEFG1 mRNA was approximately 2.1 kb, close to the size estimated from the Northern blots in Fig. 2. The comparable 5′ transcribed untranslated sequences of the white- and opaque-phaseEFG1 transcripts were identical. The sequence of the longest 5′ RACE product of a white-phase RNA preparation is presented with the predicted transcription start sites for both the white- and opaque-phase transcripts in Fig. 3.
To confirm that the sequences of the 5′ RACE products were included in the 5′ untranslated region of white- and opaque-phase EFG1mRNAs, Northern blots of total RNA of white- and opaque-phase cells were individually probed with a DNA fragment spanning the proximal (3′) portion (−1 to −209 bp relative to ATG) (Fig. 3) and the middle portion (−210 to −877 relative to ATG) (Fig. 3) of the 5′ transcribed untranslated EFG1 sequence. The proximal probe hybridized with both the white- and opaque-phase mRNAs (data not shown). The middle portion probe hybridized with the white-phase mRNA, but not the opaque-phase mRNA (data not shown). These results confirm that the 5′ RACE products of white- and opaque-phase cells accurately represented the 5′ upstream untranslated regions of the respective mRNAs.

The EFG1 promoter.

To sequence the EFG1promoter, a lambda clone, λEF39.1, was identified that contained a 6-kb sequence upstream of the EFG1 ORF. The 1.2-kb nucleotide sequence was compared both to the S. cerevisiaedatabase and the global eukaryotic promoter database (SCPD; Cold Spring Harbor Laboratory, Cold Spring Harbor, N.Y.). Two TATA box protein binding motifs (7) were identified, between nucleotides −1227 and −1220 (TBP-box2) and between nucleotides −1737 and −1731 (TBP-box1). A binding site for the repressor-activator protein RAPI (28) was identified between −1491 and −1480 bp. A J-chain variable-repeat sequence was identified between −1257 and −1249 bp. Seven TGANTN binding sites for the transcription factor GCN5 were identified in the 456-bp region immediately upstream of the untranslated region. Finally, a consensus binding site for a heat shock transcription factor (45) was identified between −2201 and −2191 bp. In the region upstream of the 5′ untranslated region of the opaque-phase EFG1 transcript, two additional TATA box protein binding motifs were identified between −245 and −253 bp (Fig.3).
To demonstrate that the differences in the level of white- and opaque-phase mRNA expression observed in Northern analyses (Fig. 2) were in fact regulated by 5′ upstream promoter sequences, the 1,369-bp region upstream of the ATG start site for EFG1 translation was inserted upstream of the RLUC gene in the reporter plasmid pCRW3 (38) to generate the plasmid pEPB26. pEPB26 integration was first targeted to the ADE2 locus (39), and integration at that locus was demonstrated by Southern analysis (data not shown). In this case, RLUC activity was approximately 33-fold higher in the opaque phase than in the white phase (Table3), the reverse of the result one might predict from Northern analysis (Fig. 2). Next, pEPB26 integration was targeted to the resident EFG1 locus, and integration at that locus was demonstrated by Southern analysis (data not shown). Integration at the EFG1 locus restored the complete EFG1 promoter upstream of the RLUC gene. In this case, RLUC activity was approximately 38-fold higher in the white phase than in the opaque phase (Table 3), which was the expected result. The level of luciferase activity for cells in the white phase was approximately 737 times higher in the strain in which pEPB26 was integrated at theEFG1 locus than in the strain in which it was integrated at the ADE2 locus (Table 3). However, the luciferase activity for cells in the opaque phase were similar in the two strains (Table3). These results demonstrate that in order to express the more abundant white-phase-specific transcript, sequences upstream of the 5′-untranslated region of the white-phase EFG1 transcript are essential. However, this region is sufficient for opaque-phase-specific expression.
Table 3.
Table 3. EFGI promoter function in the white and opaque phases at the ectopic ADE2 locus and the residentEFGI locusa
Target locus Plasmid No. of clones Switch phenotype RLUC activity (RLU/30 s/μg) Fold difference with pCRW3 Ratio of white- to opaque-phase RLUC activitiesb
ADE2 pEPB26 2 Opaque 1.3 × 106 ± 9.0 × 105 1,182
      White 4.3 × 104 ± 2.3 × 104 31 1:33
EFG1 pEPB26 2 Opaque 8.4 × 105 ± 6.1 × 105 764
      White 3.2 × 107 ± 1.3 × 107 22,858 38:1
ADE2 pCRW3 2 Opaque 1.1 × 103 ± 8.0 × 102
      White 1.4 × 103 ± 9.1 × 102
a
C. albicans WO-1 derivative Red 3/6 was transformed by the RLUC reporter plasmid either without the promoter (pCRW3) or with 1.2 kb of the EFG1 gene 5′ of the ATG codon (pEPB26). pEPB26 was targeted to either EFG1 orADE2 locus.
b
RLUC activities represent the mean ± standard deviations of three independent measurements for each clone.

Construction of an EFG1 null mutants of strain WO-1.

To directly assess the role of EFG1 in the white-to-opaque transition, we applied a urablast gene knockout strategy (9) for generating homozygous mutants in C. albicans WO-1. A ura3 strain, TS3.3, was first created as described in the Materials and Methods section. TS3.3 was transformed with the hisG-URA3-hisG-based EFG1Δ cassette, pA42.23 (Fig. 1B), and 24 transformant clones were obtained. One of the transformants, Ef3.1, was chosen and analyzed by Southern blot hybridization with the full-length EFG1 ORF probe. While Southern blots of TS3.3 digested with BamHI contained a single hybridization band at 4.5 kb, Southern blots of Ef3.1 digested with BamHI contained a 4.5-kb band and two additional bands of approximately 2.9 kb and 2.4 kb, resulting from the twoBamHI sites in the hisG cassette (Fig.4A). These two bands represent 5′ and 3′ flanking regions of the disrupted copy of the EFG1 allele. To induce URA3 auxotrophy, Ef3.1 was subjected to a 5-FOA pop out protocol (9) resulting in aura3 auxotroph, Ef3.1.1. Ef3.1.1 was transformed with a CAT-URA3-CAT-based EFG1disruption cassette from pB45.1 (Fig. 2C), and 52 transformant clones were obtained. Three of the transformants, Efc20, Efc25, and Efc30, were analyzed by Southern blot hybridization with the full-lengthEFG1 ORF probe. The 4.5-kb band was absent in all three transformants (Fig. 4A). Instead, Efc20 contained a 7.5-kb band, the expected size for the disruption cassette, while Efc25 and Efc30 contained 9.5- and 10.5-kb bands, respectively (Fig. 4A). To test whether the larger bands in the latter two were due to gross rearrangements of the flanking 4.5-kb BamHI fragment or to internal duplication of the CAT-URA3-CAT cassette, Southern analyses were performed of genomic DNA digested with BamHI,BglII, ApaLI, and HphI, and were probed with sequences that flank the 5′ or 3′ ends of theEFG1 transcript. The patterns of the three EFG1null mutants were identical to those of the wild-type WO-1 and the parental strain TS3.3 (data not shown). Since the flanking regions hybridizing to the probes spanned at least 10 kb, we conclude that the increased sizes of the inserts in Efc25 and Efc30 were due to internal duplication of the CAT cassettes rather than to reorganization of flanking regions.
Fig. 4.
Fig. 4. Southern blot analysis of the C. albicans EFGI null mutants. Approximately 3 μg of total genomic DNA from wild-type strain WO-1, the ura3 derivative TS3.3, the heterozygote prior to pop out, Ef3.1, and the three null mutants Efc20, Efc25, and Efc30 were individually digested with BamHI, resolved in an agarose gel, and transferred to Hybond-N nylon membrane. Southern blots were hybridized with either a full-lengthEFGI ORF (panel A) or the full-length URA3 ORF (panel B). The molecular weights of expected fragments are shown to the right of each panel.
All three mutants also contained the 2.4- and 2.7-kb bands representing the first disrupted allele (Fig. 4A). When probed with theURA3 ORF, TS3.3 showed no hybridization, as expected, while Ef3.1 showed hybridization of a 2.9-kb fragment, demonstrating that one of the two alleles contained the hisG-URA3-hisG cassette (Fig. 4A). The three mutants, Efc20, Efc25, and Efc30, showed no hybridization at 2.9 kb and hybridization at 7.5 to 10.5 kb, demonstrating that the first allele had lost the URA3 gene through a pop-out and the second allele contained the CAT-URA3-CAT cassette (Fig. 4A). Efc25 also contained a 3.4-kb band of unknown origin (Fig. 4A).

The EFG1 null mutant does not form white-phase colonies.

When cells of strains Efc20, Efc25, and Efc30 were plated at 25°C on agar containing modified Lee's medium, 100% of the colonies exhibited an opaque-phase colony morphology. When plated on agar containing phloxin B, all colonies (100%) were bright red (data not shown), an opaque-phase-specific characteristic (3). To test whether mutant cells could be induced to convert en masse to the white-phase phenotype by a temperature shift (23, 26, 30, 40), parental, heterozygous, and mutant cells in the opaque phase were first grown at 25°C to mid-log phase in liquid cultures of modified Lee's medium, then shifted to 42°C and incubated at the latter temperature in the same medium for an additional 16 h. Cells were then plated at low density on agar plates containing modified Lee's medium plus phloxin B, and white- and opaque-phase colony phenotypes were counted. The parental strain TS3.3 formed 94 and 92% white-phase colonies in two independent experiments (Table 4), demonstrating temperature-induced mass conversion from opaque to white phase. The heterozygous strain Ef3.1 formed 77 and 79% white-phase colonies in two independent experiments (Table 4), demonstrating mass conversion again, but at a slightly reduced level. In contrast, the three mutant strains formed no white-phase colonies in two independent experiments (Table 4), demonstrating that in the absence of EFG1expression, cells did not express the white-phase colony phenotype.
Table 4.
Table 4. Induced mass conversion from opaque to white phasea
Experiment no. Strain CFU % Conversion
Total White phase Opaque phase
1 TS3.3 1,550 1,457 93 94
  Ef3.1 1,725 1,335 390 77
  Efc20 1,595   1,595 0
  Efc25 1,610   1,610 0
  Efc30 1,474   1,474 0
2 TS3.3 1,294 1,191 103 92
  Ef3 1,316 1,039 277 79
  Efc20 1,505   1,505 0
  Efc25 1,762   1,762 0
  Efc30 1,425   1,425 0
a
Opaque-phase cells of each strain were grown in modified Lee's medium to mid-log phase at 25°C, then diluted into fresh medium at 42°C and incubated for 16 h. Cells were then spread on agar plates containing modified Lee's medium plus 5 μg of phloxine B per ml, which differentially stains opaque-phase colonies red. Plates were inoculated for 7 days at 25°C prior to counting the proportions of white and opaque CFUs.

The cellular phenotype of EFG1 null mutants.

The white-to-opaque transition in strain WO-1 involves a dramatic change in cellular phenotype (3, 30, 31). White-phase cells incubated at 25°C are round, produce round daughter cells, and exhibit a relatively homogeneous smooth surface. In contrast, opaque-phase cells incubated at 25°C are approximately twice the size of white-phase cells, are elongate or bean-shaped, and contain pimples on the cell membrane with central pores (3). When opaque-phase cells are transferred from 25 to 42°C and are incubated at the latter temperature for 16 h, they convert en masse to the white phase (23, 26, 30, 31, 40). The parent strain TS3.3 and the heterozygote Ef3.1 both formed smooth, round, budding cells in the white phase at 25°C, and pimpled, elongate, or bean-shaped cells in the opaque phase at 25°C (Fig. 5). When opaque-phase cells of strains TS3.3 and Ef3.1 were shifted from 25 to 42°C and were incubated at the latter temperature for 16 h, they converted en masse to the white phase, forming round, smooth cells characteristic of the white phase (Fig. 5). The three EFG1 null mutants Efc20, Efc25, and Efc30 all formed pimpled, elongate cells exclusively at 25°C, characteristic of the opaque phase (Fig. 5). These cells were indistinguishable from opaque-phase cells formed by strains WO-1, TS3.3, and Ef3.1 (Fig. 5). When opaque-phase cells of the three mutant strains Efc20, Efc25, and Efc30 were shifted from 25 to 42°C and were incubated at the latter temperature for 16 h, they formed cells with the smooth (i.e., devoid of pimples) surface of white-phase cells, but with the elongate shape of opaque-phase cells (Fig. 5).
Fig. 5.
Fig. 5. Scanning electron micrographs of representative cells of the parental strain TS3.3, the heterozygote Ef3.1 (Ef3), and the threeEFG1 null mutants Efc20, Efc25, and Efc30 at 25°C, and after a shift from 25 to 42°C. Wh, white phase; Op, opaque phase. Scale bars represent 2 μm.

Phase-specific gene expression in EFG1 null mutants.

The expression of several genes has been demonstrated to be phase specific in the white-to-opaque transition. While the gene WH11 is expressed exclusively by white-phase cells (40), the genes PEPI(SAPI) (24), OP4 (23), andCDR3 (5) are expressed exclusively in the opaque phase. Expression of WH11, PEPI(SAPI), and OP4 was examined in strain TS3.3, the heterozygote Ef3.1, and the three mutants Efc20, Efc25, and Efc30. In the white phase at 25°C, TS3.3 and Ef3.1 cells both expressedWH11, but neither expressed OP4 orSAPI. In the opaque phase at 25°C, cells of strains TS3.3, Ef3.1, Efc20, Efc25, and Efc30 expressed Op4 andPEP1, but did not express WH11.
Sixteen hours after opaque-phase cells of strains TS3.3 and Ef3.1 were transferred from 25 to 42°C to induce mass conversion, they no longer expressed OP4 and PEP1 (SAP1) but did express WH11 (Fig. 6). Sixteen hours after opaque-phase cells of mutant strain Efc20, Efc25, and Efc30 were transferred, they also no longer expressed OP4 andPEP1 (SAP1), but did express WH11(Fig. 6). However, the levels of WH11 transcript were severely reduced in all three mutants (Fig. 6).
Fig. 6.
Fig. 6. Northern blot analysis of phase-specific gene expression in EFGI null mutants at 25°C and after incubation for 16 h at 42°C. Opaque-phase cells were grown to mid-log phase at 25°C then diluted into fresh medium at 42°C and grown for 16 h. Northern blots of total cellular RNA from cells just prior to the shift (25°C) and 16 h after the shift (42°C) were probed with the phase-specific genes WH11, OP4, andPEPI. The length of autoradiographic exposure times is shown to the right of the hybridization patterns. Ethidium bromide-stained rRNA patterns are presented at the bottom of the hybridization patterns as a measure of loading.

Null mutants undergo switching at the level of gene regulation.

When opaque-phase cells of strain WO-1 are shifted from 25 to 42°C, they semisynchronously undergo three cell doublings, at approximately 2, 4, and 6 h (23, 40) (Fig.7A). When cells are returned to 25°C during the first 3 to 4 h at 42°C, they continue to multiply in the opaque phase, forming daughter cells with pimples. However, when cells are returned to 25°C after 4 h at 42°C, they multiply in the white phase, forming smooth, round, white-phase daughter cells (23, 40). Shift experiments have, therefore, identified a phenotypic commitment event to the white-phase between 3 and 4 h (23, 40) (Fig. 7A). When opaque-phase cells of strain WO-1 are incubated at 42°C and then transferred back to 25°C at intervals prior to the commitment event, they do not express the white-phase-specific gene WH11 (44) (Fig. 7A). However, when shifted back to 25°C after the commitment event,WH11 is reexpressed (40) (Fig. 7A). When opaque-phase cells are incubated at 42°C, they stop expressing bothOP4 and PEP1 (SAP1) (23) (Fig. 7A). When opaque-phase cells are then shifted from 42 to 25°C prior to the commitment event, they immediately reacquire transcripts of both genes within 1 h (Fig. 7A). These results demonstrate that although expression of the two phase-specific genes are temperature-sensitive, they can still be rapidly reactivated prior to the commitment event. However, when opaque-phase cells are shifted from 42 to 25°C after the commitment event, neitherOP4 nor PEP1 (SAP1) are reexpressed (23) (Fig. 7A). These results demonstrate that at the time of temperature-induced commitment to the white phase, a switch-associated event blocks subsequent temperature-induced (42-to-25°C) reactivation of opaque-phase-specific gene expression (23) (Fig. 7A).
Fig. 7.
Fig. 7. Analysis of the phenotypic commitment event inEFGI null mutants. (A) Expression of the white-phase-specific gene WH11 and the opaque-phase-specific gene OP4 prior to and after phenotypic commitment induced by a temperature shift from 25 to 42°C (23, 40). Cells were shifted from 25 to 42°C at 0 h and subfractions were then shifted down to 25°C each subsequent hour. Subsequent gene expression after 1 h at 25°C is indicated by a minus (no expression) or plus (expression) sign. The initiation of the first three rounds of semisynchronous cell division and the commitment event are indicated at the top of the figure. (B) Northern blot analysis of gene expression in the parental strain TS3.3 and the EFGI null mutant Efc20 1 h after shifts from 42 to 25°C prior to and after the commitment event. Experimental regimens are presented at the top of each lane. In the left lanes, Northern blot hybridization patterns are presented for cells at 25°C, 1 h after a shift to 42°C and 1 h after a shift from 42°C back to 25°C. In the right lanes, Northern blot hybridization patterns are presented for cells after 7 h at 42°C and 1 h after cells incubated at 42°C for 7 h are shifted back to 25°C. Ethidium bromide-stained rRNA patterns are presented at the bottom of the hybridization patterns as a measure of loading.
The phenotype observed after opaque-phase cells were shifted from 25 to 42°C (Fig. 5) suggested that the three EFG1 null mutants underwent at least a portion of the changes in cellular phenotype associated with the opaque-to-white-phase transition. If gene expression flipped in the normal fashion (Fig. 7A) in mutant cells after a shift to 42°C, this would add support to the conclusion that mutant cells undergo the switch event, but they cannot fully express the complete white-phase phenotype without EFG1 expression. When TS3.3 or Efc20 cells in the opaque phase were incubated at 42°C for 1 h, they stopped expressing OP4, and when subsequently returned to 25°C for 1 h, they reexpressedOP4 (Fig. 7B). In neither case did they expressWH11 (Fig. 7B). When opaque-phase cells of strains TS3.3 and Efc20 were incubated at 42°C for 7 h, they stopped expressingOP4, and when subsequently returned to 25°C for 1 h, they still did not reexpress OP4 (Fig. 7B). After 7 h of incubation at 42°C, WH11 was expressed, and when cells were transferred back to 25°C, expression was even greater. Therefore, both the parent TS3.3 and the EFG1 null mutant Efc20 conform to the regulation of WH11 and OP4gene expression previously described for the parental strain WO-1 (23, 40). These results demonstrate that cells lackingEFG1 undergo the changes in phase-specific gene expression that are associated with phenotypic commitment in wild-type cells.

DISCUSSION

EFG1 is expressed both in white- and opaque-phase cells.

Northern analysis of total cell RNA revealed a 3.2-kbEFG1 transcript in white-phase cells that was missing in opaque-phase cells. It also revealed a less-abundant 2.2-kbEFG1 transcript in opaque-phase cells. Densitometric measurements of the levels of the two transcripts revealed an approximately 20-fold difference in message levels, which may be the reason why Sonneborn et al. (36) did not detect the opaque-phase-specific transcript. Sequence analysis of 5′ RACE products demonstrated that the difference in the molecular sizes of the white- and opaque-phase EFG1 transcripts was due to differences in the transcription initiation sites. The transcription initiation site of the white-phase EFG1 transcript was approximately 1,173 bp upstream of the ATG translation start site, while that of the opaque-phase EFG1 transcript was approximately 162 bp upstream of the ATG translation start site. These results parallel those for the homologue StuA in A. nidulans (49). Two different promoters result in overlapping transcripts, designated StuAα and StuAβ, which encode the same protein (49).

The two EFG1 transcripts are regulated by different promoters.

The 1.2-kb 5′ sequence upstream of the ATG translation start site of EFG1 was tested for its ability to function as a promoter by placing it upstream of the Renilla reniformisluciferase gene in the plasmid pCRW3 (38). When integration of the resultant plasmid pEPB26 was targeted to the ADE2locus (39), luciferase expression was 33-fold higher in the opaque phase than in the white phase, which was opposite to expectations derived from Northern analysis. We previously reported that promoters of a variety of genes (OP4, WH11,GAL1, and EF1α2) functioned normally at this ectopic locus with luciferase activities correlating with transcript levels obtained by Northern analysis (38). However, when we targeted the reporter construct to the EFG1 locus, the promoter was reconstituted and luciferase expression was 38-fold higher in the white phase than in the opaque phase, which correlated with transcript levels revealed by Northern analysis. These results suggest that cis-acting regulatory sequences necessary for the synthesis of the higher-molecular-weight EFG1 transcript in the white phase are distal to the 1.2-kb sequence immediately upstream of the translation start site. However, the level of luciferase activity in the opaque phase was similar when integration was targeted to the ADE2 locus or the EFG1 locus, suggesting that the cis-acting regulatory sequences for opaque-phase expression of the low-molecular-weight EFG1 transcript reside in the 1.2-kb sequence immediately upstream of the translation start site. A detailed functional analysis of the EFG1promoter is now in progress.

EFG1 is necessary for the expression of the complete white-phase-cell phenotype.

A direct assessment of EFG1function in the white-opaque transition required the genesis of a homozygous disruptant. Three EFG1 null mutants uniformly formed phloxine-B-stained colonies at 25°C, which we interpreted to represent the opaque-phase phenotype (3). When cells from select colonies of the three mutants grown at 25°C were examined by scanning electron microscopy, in all cases they exhibited the elongate morphology of opaque-phase cells, and the majority of cells possessed pimples, which are a signature feature of the opaque-phase phenotype (3). Cells of all three mutants also expressed the opaque-phase-specific genes OP4 andPEP1(SAP1) and did not express the white-phase-specific gene WH11 at 25°C (32, 33). We initially interpreted these results to mean thatEFG1 null mutants were jammed in the opaque-phase phenotype. However, an analysis of phenotypic commitment during temperature-induced mass conversion from the opaque- to the white-phase phenotype (23, 26, 30, 40) proved otherwise.
When opaque-phase cells of strain WO-1, the ura3 parent strain TS3.3, and the heterozygote Ef3.1 were transferred from 25 to 42°C, they retained the capacity to multiply in the opaque phase for approximately 3 h then semisynchronously converted to white-phase growth (23, 40). The time at which 50% of cells underwent this conversion was 3.5 h, the average time of the commitment event for the differentiation from the opaque- to the white-phase phenotype. When cells of the three EFG1 null mutants were transferred from 25 to 42°C and examined after the expected time of phenotypic commitment, they formed daughter cells with the smooth wall of white-phase cells (i.e., no opaque-phase-cell pimples), but the elongate, bean-shaped morphology of opaque-phase cells. This result suggested that the temperature shift induced a switch, but mutant cells were unable to express the complete phenotype of white-phase budding cells. The cellular phenotype of temperature-shifted opaque phase cells of the three EFG1 null mutants was similar to that of a transformant of C. albicans strain CAI8 engineered to express very low levels of EFG1(36).
At the point of phenotypic commitment, wild-type cells also undergo dramatic changes in phase-specific gene expression (23, 40). When shifted to 42°C, transcription of the two opaque-phase-specific genes OP4 andPEPI (SAPI) immediately turns off. However, when cells shifted to 42°C are shifted back to 25°C prior to the commitment event, transcription of these genes is immediately resumed. After the point of commitment, however, a shift back to 25°C will not turn on transcription of these genes, demonstrating that a fundamental change occurs in the regulation of phase-specific genes at the commitment point. In support of this conclusion, transcription of the white-phase-specific gene WH11 turns on at the commitment event. All three null mutants underwent these “yin-yang” changes in gene expression at the commitment point when shifted from 25 to 42°C. Therefore, EFG1 null mutant cells in the opaque phase undergo temperature-induced phenotypic commitment, switching from an opaque-phase to a white-phase pattern of gene expression, and they undergo changes in wall morphology that include the loss of opaque-phase pimples. Since EFG1 is homologous to known transcription factors in both S. cerevisiae (11) and A. nidulans (21), it seems reasonable to suggest that it plays a similar role in C. albicans and directs the expression of a subset of white-phase-specific genes necessary for the genesis of a round, white-phase budding cell phenotype. EFG1 expression, therefore, is not integral to the basic switch event but, rather, plays a role downstream in the genesis of part of the white-phase phenotype.

The role of the opaque-phase gene transcript.

We have identified a phenotypic defect in white-phase cells of EFG1null mutants that suggests that the white-phase EFG1transcript plays a role in the genesis of the normal, round, shape of white-phase cells. We have found no similar phenotypic defect in opaque-phase cells of the three EFG1 null mutants that would suggest a role for the less-abundant opaque phase transcript. However, before concluding that this latter transcript plays no role in the opaque phase, two points must be considered. First, observing no morphological difference between wild-type and mutant opaque-phase cells does not prove phenotypic equality. Opaque-phase cells differ from white-phase cells in a number of physiological and virulence characteristics (1, 15, 17, 24, 31) that may be expressed independently of the unique opaque-phase cell morphology. Second, we have not demonstrated that the EFG1 protein is in fact expressed in opaque-phase cells. Experiments to resolve these two latter issues are in progress.

ACKNOWLEDGMENTS

We are indebted to William Fonzi of Georgetown University, P. T. Magee and Bebe Magee of the University of Minnesota, and Stewart Scherer of Acacia Biosciences for providing us with specific plasmids. We are also indebted to Lee Enger, Chris Kvaal, Sanjay Gill, and Randy Nessler of the University of Iowa for technical support.
This research was supported, in part, by Public Service grants AI2392 and DE1058 from the National Institutes of Health.

REFERENCES

1.
Anderson J., Cundiff L., Schnars B., Gao M. X., Mackenzie I., and Soll D. R. Hypha formation in the white-opaque transition of Candida albicans.Infect. Immun.571989458-467
2.
Anderson J., Mihalik R., and Soll D. R. Ultrastructure and antigenicity of the unique cell wall pimple of the Candida opaque phenotype.J. Bacteriol.1721990224-235
3.
Anderson J. M. and Soll D. R. Unique phenotype of opaque cells in the white-opaque transition of Candida albicans.J. Bacteriol.16919875579-5588
4.
Aramayo R., Peleg Y., Addison R., and Mutzenberg R. Asm-1+, a Neurospora crassa gene related to transcriptional regulators of fungal development.Genetics1441996991-1003
5.
Balan I., Alarco A.-M., and Raymond M. The Candida albicans CDR3 gene codes for an opaque phase ABC transporter.J. Bacteriol.17919977210-7218
6.
Bedell G. W. and Soll D. R. Effects of low concentrations of zinc on the growth and dimorphism of Candida albicans: evidence for zinc-resistant and -sensitive pathways for mycelium formation.Infect. Immun.261979348-354
7.
Cavallini B., Huet J., Plassat J. L., Sentenac A., Egly J. M., and Chambon P. A yeast activity can substitute for the HeLa cell TATA box factor.Nature334198877-80
8.
Church G. M. and Gilbert W. Genomic sequencing.Proc. Natl. Acad. Sci. USA8119841991-1995
9.
Fonzi W. A. and Irwin M. Y. Isogenic strain construction and gene mapping in Candida albicans.Genetics1341993717-728
10.
Frohman M. A., Dush M. K., and Martin G. R. Rapid production of full-length cDNAs from rare transcripts: amplification using a single gene-specific oligonucleotide primer.Proc. Natl. Acad. Sci. USA8519888998-9002
11.
Gimeno C. J. and Fink G. R. Induction of pseudohyphal growth by overexpression of PHD1, a Saccharomyces cerevisiae gene related to transcriptional regulators of fungal development.Mol. Cell. Biol.1419942100-2112
12.
Hellstein J., Vawter-Hugart H., Fotos P., Schmid J., and Soll D. R. Genetic similarity and phenotypic diversity of commensal and pathogenic strains of Candida albicans isolated from the oral cavity.J. Clin. Microbiol.3119933190-3199
13.
Hube B., Monod M., Schofield D. A., Brown A. J., and Gow N. A. Expression of seven members of the gene family encoding secretory aspartyl proteinases in Candida albicans.Mol. Microbiol.14199487-99
14.
Jones S., White G., and Hunter P. R. Increased phenotypic switching in strains of Candida albicans associated with invasive infections.J. Clin. Microbiol.3219942869-2870
15.
Kennedy M. J., Rogers A. L., Hanselmen L. R., Soll D. R., and Yancey R. J. Jr. Variation in adhesion and cell surface hydrophobicity in Candida albicans white and opaque phenotypes.Mycopathologia1021988149-156
16.
Kurtz M. B., Cortelyou M. W., and Kirsch D. R. Integrative transformation of Candida albicans, using a cloned Candida ADE2 gene.Mol. Cell. Biol.61986142-149
17.
Kvaal C. A., Srikantha T., and Soll D. R. Misexpression of the white-phase-specific gene WH11 in the opaque phase of Candida albicans affects switching and virulence.Infect. Immun.6519974468-4475
18.
Lachke S. A., Srikantha T., Tsai L. K., Daniels K., and Soll D. R. Phenotypic switching in Candida glabrata involves phase-specific regulation of the metallothionein gene MT-II and the newly discovered hemolysin gene HLP.Infect. Immun.682000884-895
19.
Lo H. J., Kohler J. R., DiDomenico B., Loebenberg D., Cacciapuoti A., and Fink G. R. Nonfilamentous C. albicans mutants are avirulent.Cell901997939-949
20.
Lockhart S. R., Nguyen M., Srikantha T., and Soll D. R. A MADS box protein consensus binding site is necessary and sufficient for activation of the opaque-phase-specific gene OP4 of Candida albicans.J. Bacteriol.18019986607-6616
21.
Miller K. Y., Wu J., and Miller B. L. StuA is required for cell pattern formation in Aspergillus.Genes Dev.619921770-1782
22.
Morrow B., Ramsey H., and Soll D. R. Regulation of phase-specific genes in the more general switching system of Candida albicans strain 3153A.J. Med. Vet. Mycol.321994287-294
23.
Morrow B., Srikantha T., Anderson J., and Soll D. R. Coordinate regulation of two opaque-phase-specific genes during white-opaque switching in Candida albicans.Infect. Immun.6119931823-1828
24.
Morrow B., Srikantha T., and Soll D. R. Transcription of the gene for a pepsinogen, PEP1, is regulated by white-opaque switching in Candida albicans.Mol. Cell. Biol.1219922997-3005
25.
Odds F. C. Switch of phenotype as an escape mechanism of the intruder.Mycoses219979-12
26.
Rikkerink E. H., Magee B. B., and Magee P. T. Opaque-white phenotype transition: a programmed morphological transition in Candida albicans.J. Bacteriol.1701988895-899
27.
Schiestl R. H. and Gietz R. D. High efficiency transformation of intact yeast cells using single stranded nucleic acids as a carrier.Curr. Genet.161989339-346
28.
Shore D. and Nasmyth K. Purification and cloning of a DNA binding protein from yeast that binds to both silencer and activator elements.Cell511987721-732
29.
Slutsky B., Buffo J., and Soll D. R. High-frequency switching of colony morphologyin Candida albicans.Science421985666-669
30.
Slutsky B., Staebell M., Anderson J., Risen L., Pfaller M., and Soll D. R. “White-opaque transition”: a second high-frequency switching system in Candida albicans.J. Bacteriol.1691987189-197
31.
Soll D. R. High-frequency switching in Candida albicans.Clin. Microbiol. Rev.51992183-203
32.
Soll D. R. The emerging molecular biology of switching in Candida albicans.ASM News621996415-420
33.
Soll D. R. Gene regulation during high frequency switching in Candida albicans.Microbiology1431997279-288
34.
Soll D. R., Langtimm C. J., McDowell J., Hicks J., and Galask R. High frequency switching in Candida strains isolated from vaginitis patients.J. Clin. Microbiol.2519871611-1622
35.
Soll D. R., Staebell M., Langtimm C. J., Pfaller M., Hicks J., and Rao T. V. G. Multiple Candida strains in the course of a single systemic infection.J. Clin. Microbiol.2619881448-1459
36.
Sonneborn A., Tebarth B., and Ernst J. F. Control of white-opaque phenotypic switching in Candida albicans by the Efg1p morphogenetic regulator.Infect. Immun.6719994655-4660
37.
Srikantha T., Chandrasekhar A., and Soll D. R. Functional analysis of the promoter of the phase-specific WH11 gene of Candida albicans.Mol. Cell. Biol.1519951797-1805
38.
Srikantha T., Klapach A., Lorenz W. W., Tsai L. K., Laughlin L. A., Gorman J. A., and Soll D. R. The sea pansy Renilla reniforms luciferase serves as a sensitive bioluminescent reporter for differential gene expression in Candida albicans.J. Bacteriol.1781996121-129
39.
Srikantha T., Morrow B., Schroppel K., and Soll D. R. The frequency of integrative transformation at phase-specific genes of Candida albicans correlates with their transcriptional state.Mol. Gen. Genet.2461995342-352
40.
Srikantha T. and Soll D. R. A white-specific gene in the white-opaque switching system of Candida albicans.Gene131199353-60
41.
Srikantha T., Tsai L., Daniels K., Enger L., Highley K., and Soll D. R. The two-component hybrid kinase regulator CaNIK1 of Candida albicans.Microbiology14419982715-2729
42.
Srikantha T., Tsai L. K., and Soll D. R. The WH11 gene of Candida albicans is regulated in two distinct developmental programs through the same transcription activation sequences.J. Bacteriol.17919973837-3844
43.
Stoldt V. R., Sonneborn A., Leuker C. E., and Ernst J. F. Efg1p, an essential regulator of morphogenesis of the human pathogen Candida albicans, is a member of a conserved class of bHLH proteins regulating morphogenetic processes in fungi.EMBO J.1619971982-1991
44.
Sundstrom P., Smith D., and Sypherd P. S. Sequence analysis and expression of the two genes for elongation factor 1α from the dimorphic yeast Candida albicans.J. Bacteriol.17219902036-2045
45.
Tuite M. F., Bossier P., and Fitch I. T. A highly conserved sequence in yeast heat shock gene promoters.Nucleic Acids Res.16198811845
46.
Vargas K., Wertz P. W., Drake D., Morrow B., and Soll D. R. Differences in adhesion of Candida albicans 3153A cells exhibiting switch phenotypes to buccal epithelium and stratum corneum.Infect. Immun.6219941328-1335
47.
White T. C., Miyasaki S. H., and Agabian N. Three distinct secreted aspartyl proteinases in Candida albicans.J. Bacteriol.17519936126-6133
48.
Ward M., Gimeno C., Fink G., and Garrett S. SOK2 may regulate cyclic AMP-dependent protein kinase-stimulated growth and pseudohyphal development by repressing transcription.Mol. Cell. Biol.1519956854-6863
49.
Wu J. and Miller B. L. Aspergillus asexual reproduction and sexual reproduction are differentially affected by the transcriptional and translational mechanisms regulating stunted gene expression.Mol. Cell. Biol.1719976191-6201

Information & Contributors

Information

Published In

cover image Journal of Bacteriology
Journal of Bacteriology
Volume 182Number 615 March 2000
Pages: 1580 - 1591
PubMed: 10692363

History

Received: 25 October 1999
Accepted: 23 December 1999
Published online: 15 March 2000

Permissions

Request permissions for this article.

Contributors

Authors

Thyagarajan Srikantha
Department of Biological Sciences, University of Iowa, Iowa City, Iowa 52242
Luong K. Tsai
Department of Biological Sciences, University of Iowa, Iowa City, Iowa 52242
Karla Daniels
Department of Biological Sciences, University of Iowa, Iowa City, Iowa 52242
David R. Soll
Department of Biological Sciences, University of Iowa, Iowa City, Iowa 52242

Metrics & Citations

Metrics

Note:

  • For recently published articles, the TOTAL download count will appear as zero until a new month starts.
  • There is a 3- to 4-day delay in article usage, so article usage will not appear immediately after publication.
  • Citation counts come from the Crossref Cited by service.

Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. For an editable text file, please select Medlars format which will download as a .txt file. Simply select your manager software from the list below and click Download.

View Options

Figures and Media

Figures

Media

Tables

Share

Share

Share the article link

Share with email

Email a colleague

Share on social media

American Society for Microbiology ("ASM") is committed to maintaining your confidence and trust with respect to the information we collect from you on websites owned and operated by ASM ("ASM Web Sites") and other sources. This Privacy Policy sets forth the information we collect about you, how we use this information and the choices you have about how we use such information.
FIND OUT MORE about the privacy policy