<<

Fungal Genetics and Biology 30, 173–190 (2000) doi:10.1006/fgbi.2000.1221, available online at http://www.idealibrary.com on

REVIEW Genetics and Molecular Physiology of the Kluyveromyces lactis

Raffael Schaffrath1 and Karin D. Breunig

Institut fu¨ r Genetik, Martin-Luther-Universita¨ t Halle-Wittenberg, D-06099 Halle (Saale), Germany

Accepted for publication July 21, 2000

Schaffrath, R., and Breunig, K. D. 2000. Genetics and nable to genetic manipulation almost like Saccharomyces molecular physiology of the yeast Kluyveromyces lac- cerevisiae have been developed, (ii) its evolutionary dis- tis. Fungal Genetics and Biology 30, 173–190. With tance to S. cerevisiae often allows delineation of functional the recent development of powerful molecular genetic domains from sequence comparison of homologous genes, tools, Kluyveromyces lactis has become an excellent (iii) the regulation of carbon and energy re- alternative yeast model for studying the re- flects its adaption to aerobic conditions and contrasts with lationships between genetics and physiology. In par- the well-studied but exceptional physiology of S. cerevi- ticular, comparative yeast research has been providing siae, and (iv) it has specific properties such as the “DNA- insights into the strikingly different physiological strat- killer system” that provide insight into new mechanisms of egies that are reflected by of respiration microbial competition. This review puts particular empha- over fermentation in K. lactis versus Saccharomyces sis on the latter two aspects and points out the important cerevisiae. Other than S. cerevisiae, whose physiology insights that have been gained from studies of K. lactis. is exceptionally affected by the so-called ef- Like the classical S. cerevisiae, K. lactis is an ascomyce- fect, K. lactis is adapted to aerobiosis and its respira- tous budding yeast that belongs to the endoascomycetales. tory system does not underlie glucose repression. As Formerly (mis)classified as S. lactis and K. fragilis var. a consequence, K. lactis has been successfully estab- lactis, K. lactis has been the subject of intensive taxonom- lished in biomass-directed industrial applications and ical studies; using molecular typing techniques such as large-scale expression of biotechnically relevant gene mtDNA-RFLP, rDNA alignments, and PFGE karyotyp- products. In addition, K. lactis maintains species-spe- ing (Herman and Halvorson, 1963; Fuson et al., 1987; cific phenomena such as the “DNA-killer system,” Vaughan-Martini and Martini, 1987; Cai et al., 1996; analyses of which are promising to extend our knowl- Kurtzman and Robnett, 1998), today K. lactis is reconsid- edge about microbial competition and the fundamen- ered to be a unique species. tals of plasmid biology. © 2000 Academic Press Its property of growing on as a sole carbon source distinguishes K. lactis from most other and The yeast Kluyveromyces lactis is attracting increasing can be used to select isolates from dairy industry products attention from molecular biologists working in widely dif- (Valderrama et al., 1999), hence its name dairy or milk ferent fields. Several aspects have contributed to this de- yeast. In fact, it can use a wide range of substrates by velopment: (i) various molecular tools that make it ame- respiration. K. lactis is a so-called “petite-negative” yeast in which maintenance of mitochondrial DNA is essential. It 1 To whom correspondence should be addressed. E-mail: schaffrath@ has a respiratory–fermentative metabolism and most genetik.uni-halle.de. strains can grow on fermentable carbon sources when

1087-1845/00 $35.00 Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. 173 174 Schaffrath and Breunig respiration is blocked. Nevertheless, K. lactis appears to be Despite organization similar to that of other yeast CEN- an obligate aerobe that is not able to grow under strict DNAs, the K. lactis centromeres have been shown to anaerobic conditions. function species specifically (Heus et al., 1990, 1994). Deletion of the K. lactis gene coding for the CDEI- binding protein, Cpf1p, is lethal (Mulder et al., 1994), whereas in S. cerevisiae a cbf1⌬ strain is only partially MENDELIAN GENETICS AND impaired for centromere function and is viable (Mellor et CHROMOSOMAL DNA al., 1990). A CEN-binding factor CBF5 homologue was isolated from K. lactis and implicated in chromosome separation (Winkler et al., 1998). Telomere structure– In contrast to that of S. cerevisiae, mitotic growth of K. function analysis has shown that, other than in S. cerevi- lactis is predominantly restricted to the haplophase, caus- siae, K. lactis maintains a perfect 25-bp telomere repeat ing diploid cells to spontaneously enter meiosis and sporu- present, on average, as 15 tandem copies, suggesting high ␣ lation. Most laboratory strains maintain stable and a fidelity of activity (McEachern and Blackburn, ␣ mating types encoded by the MAT or MATa genes but K. 1994). Telomere addition by telomerase uses a 30-nucle- lactis also carries cryptic mating-type alleles that are si- otide template region of its endogenous 1.2-kb TER1- lenced in a Sir4p-dependent manner and, as in S. cerevi- 1 RNA component that is complementary to 1 ⁄5 telomeric siae, mating-type switching can occur (Astrom and Rine, repeats (Fulton and Blackburn, 1998). Telomere growth is 1998). deregulated in ter1 mutants and dependent on interaction The chromosomal DNA has a GC content of 40%, of the most distal repeat unit with Rap1p, a dsDNA- comparable to that of S. cerevisiae. Karyotyping of K. lactis binding protein of the telomeric cap (Krauskopf and has revealed the presence of only six chromosomes, rang- Blackburn, 1996, 1998). Uncapping or cap complex dis- ing in size from 1 to 3 Mb (Sor and Fukuhara, 1989). The ruption ultimately causes cells to die (Smith and Black- estimated genome size of 12 Mb may not be significantly burn, 1999). The severity of this phenotype can be rescued smaller than that of S. cerevisiae but the larger size and on stabilizing deregulated telomeres by recapping, sug- smaller number of chromosomes are typical of non-Sac- gesting that maintenance of the terminal telomeric cap charomyces yeasts. Codon adaptation indices, studied in complex is essential for telomere function and cellular 47 chromosomal K. lactis genes by Lloyd and Sharp integrity (Smith and Blackburn, 1999). Consistently, lack (1993), follow the S. cerevisiae rules with a notable excep- of functional telomerase due to TER1 deletion results in tion for the unique CYC1 cytochrome c gene, which gradual telomere loss and progressive cellular senescence reflects the importance of respiratory metabolism in K. (Roy et al., 1998; Smith and Blackburn, 1999). Surpris- lactis. Despite the strikingly different karyotypes of K. ingly, cells surviving this catastrophic situation emerge at lactis and S. cerevisiae, the extent to which the local high frequencies. Life without telomerase, however, fully arrangement of gene order along the chromosomes has depends on RAD52, suggesting that recombination acti- been conserved amounts to 75% for 31 adjacent gene vated by telomere uncapping can restore telomere main- neighbors studied (Keogh et al., 1998). Consistent with the tenance and viability (McEachern and Blackburn, recent interpretation of gene arrangement data by K. 1996). Wolfe and co-workers that S. cerevisiae must have had its whole genome duplicated (Wolfe and Shields, 1997; Seoighe and Wolfe, 1998), gene redundancy in K. lactis is generally lower. The genetic map of K. lactis is poor, with NON-MENDELIAN GENETICS AND about only 30 linkage groups established and physically PLASMID BIOLOGY mapped to individual chromosomes (Wesolowski-Louvel et al., 1996). However, large-scale sequencing of random Mitochondrial DNA tags has identified 292 new K. lactis genes corresponding to about 5% of its estimated gene number (Ozier-Kalog- The mitochondrial DNA of K. lactis is a 39-kb circular eropoulos et al., 1998). Also, genetic/physical genome and maintained, on average, as 30 copies per mapping is in progress and recent efforts are being made haploid genome, which roughly equals 10% of the total to initiate a separate K. lactis genome project (M. Bolotin- DNA content (Wesolowski-Louvel et al., 1996). Despite Fukuhara, pers. communication). strain-specific mtDNA polymorphism, partial DNA se-

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. Genetics and Molecular Physiology of K. lactis 175 quence analysis has established a physical map that accom- been repeatedly used for gene cloning and identification. modates the 15S and 21S rRNA genes, 22 tRNA loci, and The Kluyveromyces host range of pKD1 is broader than ORFs encoding apocytochrome B (CYTB), as well as cy- that of the 2-␮m plasmid, allowing maintenance of pKD1 tochrome oxidase (COX1, COX2, and COX3) and mito- vectors in K. marxianus, K. wickerhami, K. thermotoler- chondrial ATPase (ATP6, ATP8, and ATP9) subunits. Con- ans, and K. waltii (Chen et al., 1989). sistent with evidence from mtDNA transcription in other yeast species, the mitochondrial RNA of K. Cytoplasmic Episomal DNA lactis recognizes a conserved nonanucleotide promoter A ( /TTATAAGTA), and all dairy yeast mtORFs are tran- Certain K. lactis strains maintain “killer DNA,” an en- scribed unidirectionally (Osinga et al., 1982). dogenous, cytoplasmic plasmid pair (k1 and k2; Fig. 1) Unlike the high frequency of spontaneously arising S. which encodes an anti-yeast toxin complex known as zy- cerevisiae petite (␳0) mutants, K. lactis is petite-negative. mocin (Gunge et al., 1981; Fig. 1). k1 and k2 have been Usually, loss of mtDNA is lethal for dairy yeast; however, thoroughly studied and represent a model system for an- if mutated at any of three nuclear loci called MGI (mito- alyzing the fundamentals of linear plasmid replication and chondrial genome integrity), K. lactis can be artificially gene expression (for reviews see Stark et al., 1990; Gunge, converted into a ␳0 yeast by continued growth in the 1986, 1995; Fukuhara, 1995; Meinhardt et al., 1997; Schaf- presence of DNA-targeting drugs (Chen and Clark- frath et al., 1999). Genetic analysis has shown that k1 Walker, 1993). Interestingly, disruption of the MGI genes encodes the three zymocin subunits (␣ and ␤: ORF2; ␥: which code for the mitochondrial F1-ATPase subunits ␣, ORF4; see below) and an immunity determinant (ORF3). ␤, and ␥ does not lead to the production of mitochondrial In contrast, the helper plasmid k2 provides vital k1/k2- genome deletion mutants and petite colonies, indicating maintenance functions (Schaffrath and Meacock, 1995), that an assembled F1 complex is needed for the “gain-of- including factors for replication (ORF2, ORF5, and function” of the mgi phenotype (Chen and Clark-Walker, ORF10), mRNA capping (ORF3) (Larsen et al., 1998), 1995, 1996). Inactivation of the F1-ATPase ␣ and ␤ sub- DNA unwinding (ORF4), and a plasmid-specific RNA unit genes in S. cerevisiae converts this petite-positive polymerase (RNP) (ORF6 and ORF7) (Wilson and Mea- yeast into a petite-negative yeast (Chen and Clark-Walker, cock, 1988; Schaffrath et al., 1995a, b, 1997a). k1 and k2 2000). Thus, comparative studies on nuclear genes affect- genes are activated from plasmid-specific promoter motifs ing mtDNA maintenance promise to yield important in- (UCSs) shown to be incompatible with the K. lactis host- sights into nuclear–mitochondrial interactions. encoded RNPs (Romanos and Boyd, 1988). A body of evidence, including terminal proteins attached to k1 and Nuclear Episomal DNA k2 and genes on both k1 (ORF1) and k2 (ORF2) that code for DNA , suggests a viruslike protein-primed The nuclear 2-␮m-type episome, pKD1, originally iso- mode of plasmid replication with replicases that function lated from K. drosophilarum, can be stably maintained in plasmid specifically (Kitada and Gunge, 1988; Salas, 1991; ϩ K. lactis to create cir hosts (Chen et al., 1986; Falcone et Schaffrath et al., 1995b). k2 encodes two other likely al., 1986). In analogy to the 2-␮m plasmid of S. cerevisiae, maintenance factors: TRF1 (ORF10), a plasmid-specific pKD1 is phenotypically cryptic and exists as two isomers replication initiator (McNeel and Tamanoi, 1991), and that can be interconverted by the gene A product, an Orf5p (ORF5) (Schaffrath and Meacock, 1995), which FLP-like site-specific recombinase. Genes B and C prob- binds to single-stranded DNA in vitro and probably pro- ably code for pKD1 partitioning factors, and autonomous tects replicative intermediates from exonucleolytic degra- replication requires a cis-acting ori sequence (Bianchi, dation in vivo (Schaffrath, 1995; R. Schaffrath and P. A. 1992; Bianchi et al., 1991). pKD1 can be used to transform Meacock, submitted). Summing up, in the fundamental S. cerevisiae cir0 strains but plasmid maintenance is very processes of replication and transcription, these plasmids low; vice versa, 2-␮m-based vectors get lost readily in K. function remarkably independent of their host cell. lactis without maintaining constant selective pressure. Mi- Successes in genetically manipulating k1 and k2 by totic stability, however, is increased in both species by allelic replacement have provided powerful tools for introducing the 2-␮m ori into pKD1-based vectors and studying these unusual plasmids at the molecular level ϩ using cir K. lactis or S. cerevisiae hosts carrying resident (Ka¨mper et al., 1989, 1991; Tanguy-Rougeau, 1990; Schaf- pKD1 or 2-␮m plasmids. This type of vector allows gene frath et al., 1992). To do so, the transcriptional barrier and library shuttles between both yeast species and has between host- and plasmid-specific genes was overcome

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. 176 Schaffrath and Breunig

FIG. 1. The K. lactis killer system. (A) Genetic organizations of killer plasmids, k1 and k2. Functionally related genes are identically color–coded according to Schaffrath et al. (1999): zymocin biogenesis/immunity: blue; DNA replication: grey; transcription and mRNA modification: orange; unassigned orphans: red. k1- and k2-specific terminal proteins (TPs) and terminal inverted repeats (TIRs) are illustrated as white/black circles/triangles,

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. Genetics and Molecular Physiology of K. lactis 177

␣ ␤ ␥ by introducing suitable markers fused to UCSs and flanked termed , , and in decreasing order of their Mr (98, 30, by target sequences to direct k1/k2-specific integration. and 28 kDa, respectively), are encoded by the killer plas- This knockout technique has allowed dispensable and es- mid k1 (see above). The amino-terminal residues of the ␣ ␤ ␥ sential genes to be distinguished (Schaffrath et al., 1999). mature , , and polypeptides are A30 (ORF2), G895

In addition, development of a cytoplasmic gene shuffle (ORF2), and A19 (ORF4), respectively. Hence the two system has facilitated the swapping of genes between k1 larger subunits are the processed products of a single and k2 to reintroduce site specifically mutagenized or gene, k1ORF2 (Stark and Boyd, 1986; see above), and modified alleles (Schaffrath et al., 1996, 1999, 2000). Thus, result from proteolysis of an ␣␤ precursor at about two- Orf5p, a protein essential for maintenance of plasmid thirds along the primary ␣␤ translation product; the ␥ integrity (Schaffrath and Meacock, 1995), has been shown subunit, however, is the processed product encoded by to cause severe plasmid instability in vivo when carboxy- k1ORF4 (Stark and Boyd, 1986; see above). As for the terminally mutated and identified by way of epitope tag- latter, its mature NH2 terminus (A19) colocalizes within the ging (Schaffrath and Meacock, 1996; R. Schaffrath and first of three potential signal peptidase recognition sites P. A. Meacock, submitted). Functionally, the tagged vari- 2 2 2 (C16AA A A ). In contrast, despite the existence of such ant is indistinguishable from the wild-type protein and is a recognition site within the ORF2 leader region as abundant as TRF1 (see above), implying a structural 2 ␣ (V19QG ), the amino-terminus of the subunit (A30 of function (Schaffrath and Meacock, 1996). In addition, the ORF2)inthe␣␤ precursor does not appear to result from major transcriptase gene (ORF6) has been swapped and signal peptidase cleavage, but rather from proteolytic pro- ⌬ shown to complement an orf6 knockout created on plas- cession by another endopeptidase, presumably the K. lac- mid k2 when ectopically shuffled onto k1 (Schaffrath et al., tis Kex2p homologue: Kex1p (Stark et al., 1990; see be- 2000). In conclusion, by use of the gene shuffle technique, low). elucidation of k1/k2 gene function, regulation of k1/k2 Zymocin biosynthesis can be prevented in the presence gene expression, and UCS promoter analysis has been of tunicamycin, suggesting that the complex is glycosylated experimentally addressable (Cong et al., 1994; Schaffrath (Sugisaki et al., 1985). As judged from endo-␤-N-acetyl- et al., 1996, 2000; R. Schaffrath and P. A. Meacock, sub- glucosaminidase H (endo H) digestion of purified zymocin mitted). As a consequence, k1 and k2 have acquired con- preparations, the ␣ subunit shows a decrease of about 3.4 siderable interest as an alternative vector system (see be- kDa in its M (Stark and Boyd, 1986). Together with the low; Meinhardt et al., 1997; Schaffrath et al., 1999). r observation that -rich moieties are detected only for the ␣ subunit by use of a concanavalin A conjugated- horseradish peroxidase-based chromogenic assay (Gunge, THE K. lactis ZYMOCIN 1995; Schaffrath et al., 1997b), this indicates a single, N-linked oligosaccharide chain associated with the ␣ Structure and Biogenesis polypeptide. In contrast, none of the other subunits ap- pears to be N- nor O-glycosylated and both are endo H The holo-zymocin secreted by killer strains of K. lactis is resistant (Stark and Boyd, 1986). Structural integrity of the a heterotrimeric glyco–protein complex. Using micro-se- ␣␤␥ glyco–protein complex is maintained by internal cys- quence analysis, it became evident that all three subunits, teine bonds within the ␣ subunit and an intermolecular

respectively. DNP, DNA polymerase; RNP, RNA polymerase; SSB, single-stranded DNA-binding protein; TRF1, terminal recognition factor 1; for explanations see text. (B) Killer phenotype. K. lactis killer (K) and nonkiller (NK) strains were grown on a lawn of sensitive S. cerevisiae cells; growth inhibition is visualized by halo formation (C) Zymocin-mediated cell cycle arrest. Sensitive S. cerevisiae cells were treated with zymocin and analyzed by FACS. Cell fractions with 1n and 2n DNA contents are indicated by arrows. FIG. 2. The LAC/GAL Regulon of K. lactis. (A) Structural genes of the LAC/GAL regulon are shown as arrows (not drawn to scale) with KlGal4p binding sites as green triangles. Regulatory genes are colored. On top, the function of the gene products in the lactose– is shown. The gray bow indicates the plasma membrane with integrated sugar permeases. Galactose can enter the cell via Lac12p and independently of Lac12p. The Kht1p/Rag1p hexose transporter is a candidate. (B) Interactions regulating Gal4p transactivator activity. The Gal80p protein blocks Gal4p function by interacting with the activation domain. Gal1p relieves the inhibition by binding Gal80p in a galactose- and ATP-dependent way. In S. cerevisiae the formation of a Gal4p–Gal80p–Gal3p trimeric complex could be demonstrated (Platt and Reece, 1998).

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. 178 Schaffrath and Breunig disulfide bridge between the ␤ and the ␥ polypeptides. that still contain k2 are phenotypically sensitive toward Hence, treatment of purified zymocin with ϪSH reagents zymocin and indistinguishable from plasmid-free nonkiller causes subunit disintegration and completely abolishes strains (Niwa et al., 1981). In contrast, nonkiller strains biological activity (Stark et al., 1990). carrying the k1-based orf2⌬ knockout plasmids k1-NK2, Zymocin secretion is blocked in a K. lactis kex1 mutant pGKL1S, or pGKL1D retain immunity but fail to produce and can be rescued by single-copy complementation with zymocin (due to being deleted in the ␣␤ subunit structural the KEX2 gene from S. cerevisae. Vice versa, the K. lactis gene, ORF2) (Gunge, 1995). Since strains maintaining the KEX1 gene is able to functionally complement a kex2⌬ k1 orf2-4⌬ deletion plasmids F1 and F2 fail to confer both gene knockout in baker’s yeast (Wesolowski-Louvel et al., autoimmunity and zymocin expression (Kikuchi et al., 1988; Tanguy-Rougeau et al., 1988). Kex2p localizes to the 1985), the k1ORF3 gene was considered to code for im- Golgi and is essentially involved in biogenesis of secreted munity. In fact, on cloning ORF3 into a K. lactis ARS proteins such as prepro-␣-pheromone and the dsRNA- vector and retransformation into a k1-free host strain, encoded K1 killer toxin precursor. Not only do recombi- immunity toward zymocin was partially restored as long as nant k1/k2-carrying S. cerevisiae strains process and se- the cell comaintained k2 (Tokunaga et al., 1987). There- crete authentic zymocin but the kex2⌬ mutant of baker’s fore, k2 was postulated by Stark et al. (1990) to be essen- yeast is also unable to produce secretable zymocin (Toku- tially needed for ORF3 gene expression on the ARS vec- naga et al., 1990; Stark et al., 1990). Thus, zymocin pro- tor; alternatively, it might encode another immunity factor duction appears to be very similar in both yeast species (in addition to k1ORF3). Although there is sequence ho- and processing is obligatorily coupled with secretion dur- mology between the k1ORF3 and the k2ORF1 genes ing its biogenesis and/or assembly into its holo-form. Con- (Stark et al., 1990), a role of the latter in immunity has sistently, in the absence of ␣␤ secretion the ␥ subunit fails been all but ruled out by targeted gene disruption: strains to be secreted but accumulates intracellularly (Tokunaga carrying a k2-based orf1⌬ϻLEU2 null are indis- et al., 1989). A zymocin-resistant S. cerevisiae strain har- tinguishable from wild-type killer strains in expression of boring a conditional galactose-regulatable ORF4 secretion both killer and immunity phenotypes (Schaffrath et al., vector does not allow for ␥ subunit export under inducing 1992). Immunity is also seen in S. cerevisae strains toward conditions; instead, despite its functional secretion signal, expression of intracellular ␥ toxin subunit (see above; which has been shown to efficiently direct the secretion of Tokunaga et al., 1989), evidence which may implicate that heterologous proteins, the ␥ subunit also accumulates in- immunity does not work simply by exclusion of the zymo- tracellularly (Tokunaga et al., 1988, 1989). However, its cin from the cell. Instead, it suggests that immunity func- secretion can be effectively restored by genetically engi- tions by interaction of the Orf3p determinant with the ␥ neering a heterologous pro-sequence from S. cerevisiae subunit itself or with its putative intracellular target(s) in-frame between the ␥ leader peptide and its mature (Stark et al., 1990). coding region (Tokunaga et al., 1989, 1990). If this foreign ␣␤ pro-sequence can circumvent the requirement for the Mode of Action precursor to promote ␥ subunit export, why is ␣␤ needed during zymocin biogenesis? Explanations may include a The K. lactis zymocin is active against the growth of a role in assisting the correct folding of the ␥ polypeptide or variety of sensitive yeast genera, including Saccharomyces, a requirement for the ␣␤ precursor to assemble the ␥ Kluyveromyces, and Candida but not Schizosaccharomy- subunit into a secretable holo-zymocin either by facilitat- ces (Gunge et al., 1981; Gunge and Sakaguchi, 1981; Fig. ing the formation of the intermolecular ␤–␥ disulfide 1). Unlike the killer toxin K1 of S. cerevisiae, the K. lactis bridge (see above) or by providing the unglycosylated ␥ zymocin does not cause sensitive cells to hyperactivate subunit with a glycosylated entity typical for yeast exo- Tok1p and elicit rapid effluxes (Ahmed et al., 1999) but proteins. rather inhibits cell division (Butler et al., 1991a). Zymocin- treated S. cerevisiae strains accumulate as unbudded cells, Immunity implying that a G1 cell cycle-specific event may be af- fected (Butler et al., 1991a; Fig. 1). Further evidence that Not surprisingly, a K. lactis killer strain also expresses the zymocin acts specifically in G1 is provided by the immunity toward its own secreted exo-zymocin. Immunity observation that cells that have been chemically arrested is strictly dependent on the maintenance of plasmid k1: as in S phase by hydroxyurea prior to zymocin treatment are shown by curing experiments, k1-free isolates of K. lactis able to complete one round of cell division and get ar-

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. Genetics and Molecular Physiology of K. lactis 179 rested in the new unbudded G1 cell cycle stage once they losamidin (Butler et al., 1991c) and the hydrophobicity have been released from the chemical S block in the predicted for the ␤ subunit, this suggests that ␣␤ may presence of zymocin (Butler et al., 1991a). Indeed, FACS associate with the plasma membrane. It is therefore analysis demonstrates a prereplicative DNA content (1n) tempting to speculate that ␣␤ may be required for native suggestive for a G1 cell cycle block prior to START (Butler zymocin to act from the cell’s exterior by docking to cell et al., 1991a; L. Fichtner and R. Schaffrath, unpublished wall-associated chitin and mediating ␥ toxin translocation. data; Fig. 1). Consistently, as shown by incorporation of Indeed, depletion of chitin by way of deleting, inactivating, radiolabeled precursors into RNA and protein, treated or mistrafficking major yeast chitin–synthase III activity cells are still metabolically active, allowing for transient renders cells refractory to the inhibitory effects of exo- macromolecular biosynthesis to be permitted (Butler et zymocin. Thus, chitin appears to be the zyomocin receptor al., 1991a; M. J. R. Stark, pers. communication). Also, they required for the initial interaction with sensitive cells (D. increase in volume by some 30–50%, similar to the bona Jablonowski and R. Schaffrath, unpublished data; see be- fide pre-START G1 arrests induced by the mating phero- low). mone cascade or growth of cdc28ts strains at the nonper- Based on their ability to grow in the presence of the missive temperature (Leberer et al., 1997). holo form, zymocin-resistant S. cerevisiae mutants Earlier reports that the zymocin functions on S. cerevi- (termed skt (sensitivity to K. lactis toxin), iki (insensitive to siae by inhibiting the CDC35 gene product, adenylate killer), and kti (K. lactis toxin insensitive), respectively) cyclase, and hence abolishing the roles of cAMP essential have been isolated independently (Kawamoto et al., 1992; for mitotic growth and cell division (Sugisaki et al., 1983) Butler et al., 1994; Kishida et al., 1996). Sensitivity of these have been disproven by knocking out BCY1, which en- mutants toward intracellular, conditional expression of the codes the regulatory subunit of cAMP-dependent protein ␥ toxin from inducible promoters (see above) can distin- ; the bcy1⌬ mutant which no longer requires cAMP guish zymocin binding/uptake (class I) from ␥ toxin target to undergo cellular division does grow in the absence of a site mutants (class II). Consistent with the role that the ␣ functional CDC35 gene product but behaves as sensitive subunit may play in assisting zymocin docking to the cell to and G1 arrestable by the exo-zymocin as its wild-type surface by virtue of its exo-chitinase activity (see above) parent strain (White et al., 1989). Thus, it is highly unlikely are studies on chs (chitin synthesis-deficient) S. cerevisiae that the zymocin exerts its toxicity by inhibiting Cdc35p; mutants (Takita and Castilho-Valavicius, 1993). The ma- yet, studying its mode of action may prove a useful tool for jority of these behave as class I mutants; i.e., they are dissecting stage-specific events in G1. resistant to exo-zymocin but sensitive toward endogenous Interestingly, toxicity of the zymocin complex appears to expression of the ␥ toxin. Moreover, KTI2 is allelic with reside solely within the smallest subunit, the so-called ␥ CHS3, and KTI10 may very well correspond to CHS6 toxin: conditional expression of its structural gene from (M. J. R. Stark, pers. communication); CHS3 codes for the tightly regulated yeast promoters (UASGAL or UASMET) catalytic subunit of chitin synthase III (CSIII), the in vivo mimics exogenous treatment of the holo-zymocin and is activity of which is lacking in chs6 mutants and usually fully lethal to sensitive strains of S. cerevisiae under in- amounts to 90% of a wild-type cell’s chitin synthesis (Cos ducing conditions (Tokunaga et al., 1989; Butler et al., et al., 1998; Ziman et al., 1998). In addition, knocking out 1991b; Schaffrath et al., 1997b; F. Frohloff and R. Schaf- SKT5 (isoallelic with CHS4), which encodes a posttrans- frath, unpublished data). Endogenous expression of ma- lational activator of CSIII, renders cells resistant to exo- ture ␥ toxin (i.e., without its native secretion leader) re- zymocin (Kawamoto et al., 1993; Trilla et al., 1997). Al- sults in biologically active ␥ toxin, whereas, exogenously though the class I gene KTI6 is nonallelic with CHS3, applied, the ␥ polypeptide is not able to inhibit cell growth CHS4, and CHS6, its mutation obviously affects zymocin (Tokunaga et al., 1989). As for the roles played by the ␣ binding or ␥ toxin uptake (M. J. R. Stark, pers. commu- and ␤ subunits within the secreted exo-zymocin, this nication). So it is possible that it corresponds to either strongly implies that the holo form must assist the ␥ toxin CHS5 or CHS7, two genes that have been recently shown subunit to be taken up by the cell. As a precondition for ␥ to function in targeting and trafficking of Chs3p from the subunit entry and action, the holo-zymocin is therefore ER to chitosomes (Ziman et al., 1996; Santos and Snyder, expected to first of all bind to the cell surface to be able to 1997; Trilla et al., 1999). Summing up, we propose that the signal its toxicity (Stark et al., 1990). Together with the primary interaction of the holo-zymocin with its sensitive finding that the ␣ subunit has an essential exo-chitinase cell is facilitated by docking to cell wall chitin. Thus, chitin activity in vitro that can be specifically inhibited by ␤-al- may serve as a receptor molecule and all scenarios that

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. 180 Schaffrath and Breunig lead to a reduction of chitin synthesis due to abolishing, progression into S phase (Di Como and Arndt, 1996; Luke reducing, and/or mistrafficking CSIII activity render S. et al., 1996). Sit4p is required for execution of START and cerevisiae cells resistant toward K. lactis zymocin (D. Jab- sit4ts strains arrest late in G1 prior to START, in part due lonowski and R. Schaffrath, unpublished data). to the role of Sit4p in expression of the G1 cyclin genes The presence of as many as 10 distinct target site class CLN1 and CLN2. Interestingly, sit4⌬ null mutants which II mutants (see above) suggests that a complex pathway are viable only in certain SSD1-␯ backgrounds are fully transduces the zymocin’s inhibitory effect (Butler et al., resistant to partially purified zymocin (Stark, 1996). Thus, 1994). Whereas, some of them may be involved in the though it is tempting to speculate that zymocin might expression of targets inhibited by the ␥ toxin, a number of antagonize G1 cyclin function, we can only conclude that proteins could also participate in the process that is Sit4p is required for cells to respond to it. Also, neither a blocked by it. These might act as a biochemical pathway hyperactive dominant CLN3-1 allele nor overexpression of or, alternatively, form a complex containing several com- CLN2 can significantly reduce zymocin sensitivity (Butler ponents. To identify ␥ toxin target site (tot) mutant(s) we et al., 1994; Schaffrath et al., 1997b). Therefore, the mode are exploiting genetic screens involving two-hybrid inter- of zymocin action still remains unclear. However, we an- action trapping and gene knockout transposon tagging ticipate that molecular genetic analyses of zymocin-resis- (Schaffrath et al., 1997b). As for the latter, a pool of yeast tant mutants (see above) rather than biochemical ap- transformants which carry a Tn3ϻlaczϻLEU2 minitrans- proaches will provide more direct routes to identifying and poson randomly inserted into the genome are screened for characterizing the ␥ toxin target site, TOT. viability and ␥ toxin resistance, the tot phenotype, by conditionally switching on ␥ toxin expression (see above). Candidate clones are then subjected to plasmid rescue in TOOLS FOR GENETIC MANIPULATION and characterized by DNA sequencing. In this way we and others have identified several resistant yeast disruptants. So far seven genes (IKI1/TOT5, IKI3/ The use of K. lactis as a model nonconventional yeast ELP1/TOT1, ELP2/TOT2, ELP3/TOT3, KTI12/TOT4, has been significantly advanced as a result of several tech- SIT4, and SAP155) have been implicated in affecting in- nical developments. Standard prototrophic yeast markers tracellular ␥ toxin action. Iki1p/Tot5p is part of an insol- isolated from S. cerevisiae (URA3, LEU2, TRP1, etc.) were uble fraction in cell extracts and Iki3p/Tot1p is predicted shown to function in the appropriate K. lactis auxotrophic to possess a membrane-spanning region, raising the pos- backgrounds and vice versa; the K. lactis homologues sibility that an insoluble Iki1p/Iki3p-containing compart- could be identified in S. cerevisiae by means of interge- ment is involved in toxin action (Yajima et al., 1997). The neric complementation (Shuster et al., 1987; Stark and recent reidentification of Iki3p as Elp1p, a complex com- Milner, 1989; Zhang et al., 1992; Bai et al., 1999). There ponent of elongating RNA polymerase II holo-, are also dominant markers conferring resistance to gene- which may associate with two other elongator constituents, ticin (G418R), bleomycin (BLER), and aureobasidin A Elp2p/Tot2p and Elp3p/Tot3p, provides substantial evi- (AUR1) or enabling acetamide utilization (amdS) dence for this hypothesis (Otero et al., 1999; Wittschieben (Sreekrishna et al., 1984; Chen and Fukuhara, 1988; et al., 1999; Fellows et al., 2000; F. Frohloff and R. Hashida-Okado et al., 1998; Castrillo et al., 1999). Trans- Schaffrath, unpublished data). Depletion or overproduc- formation techniques include PEG-mediated DNA trans- tion of Kti12p/Tot4p confers ␥ toxin resistance; so Kti12p fer into protoplasts, chemical lithium acetate protocols, may be a potential ␥ toxin target (Butler et al., 1994). If and electroporation (Das and Hollenberg, 1982; Klebe et absent (kti12⌬) or mutated (as in the kti12 background), al., 1983; Meilhoc et al., 1990). The latter yields high the ␥ toxin cannot bind to Kti12p, whereas high KTI12 transformation efficiencies and is generally preferred gene dosage might lead to excess, unbound Kti12p which when using genetic screens based on single and/or multi- competes for a downstream effector molecule, thereby copy genomic libraries. diluting the toxic signal of the ␥ subunit (Schaffrath et al., Currently, there are several plasmid vector systems 1997a). Consistent with this we found that Tot4p interacts available to transform K. lactis cells. These are based on with Elp1-3p/Tot1-3p and Iki1p/Tot5p (L. Fichtner, F. either the multicopy episomal pKD1 plasmid (see above), Frohloff, and R. Schaffrath, unpublished data). Sap155p the K. lactis autonomously replicating sequences (KARSs) associates in a cell cycle-dependent manner with the TOR (Das and Hollenberg, 1982; Thompson and Oliver, 1986), pathway phosphatase Sit4p, which functions late in G1 for or the K. lactis centromeric sequences (K1CENs, see

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. Genetics and Molecular Physiology of K. lactis 181 above) (Heus et al., 1990). Thus, depending on the type of extracellular acid phosphatase, ␣ pheromone, invertase, genetic manipulation, one can choose from a variety of glucoamylase, and serumalbumin (Bui et al., 1996; vectors, including low and multicopy episomal plasmids, Gellissen and Hollenberg, 1997; Ferminan and integrative vectors (Chen, 1996), intergeneric plasmids Dominguez, 1998). and libraries to shuttle DNA between S. cerevisiae, K. High-level gene expression can be approached from lactis, and E. coli (Wesolowski-Louvel et al., 1996), or even either integrative or episomal strategies allowing mitotic shuffle-recombinant DNA constructed in vitro between stability or multicopy dosage effects of the foreign gene. the linear plasmids k1 and k2 in vivo (Schaffrath and Both approaches have been successfully exploited with Meacock, 1996; Schaffrath et al., 1999, 2000). Regulation expression yields on the order of several grams for bovine of gene expression can be studied by standard reporter prochymosin and human serumalbumin per liter fed-batch assays (Chen and Fukuhara, 1988); to utilize the E. coli fermenter culture (van den Berg et al., 1990; Fleer et al., lacZ gene, one should work on a lac4 mutant lacking 1991). Strong transcriptional activation is obtained by both endogenous dairy yeast ␤-galactosidase. constitutive and conditional promoter systems (van den In contrast to the high fidelity of gene replacement in S. Berg et al., 1990; Bui et al., 1996; Ferminan and cerevisiae, knocking out genes by homologous recombina- Dominguez, 1998; Saliola et al., 1999; Mustilli et al., 1999; tion in vivo is less efficient in K. lactis (Wesolowski-Louvel Hsieh and Da Silva, 2000). As far as posttranslational et al., 1996). Nonetheless, targeted gene replacements are modifications are concerned, recombinant ␤-lactoglobulin possible by means of one- or two-step disruption tech- secreted from K. lactis has been shown to be virtually niques and, recently, Bundock et al. (1999) have described indistinguishable from the native protein based on PAGE a modified protocol that avoids illegitimate recombination analysis, reactivity to antibodies, CD, fluorescence spec- and results in a 10-fold increase of homologous targeting troscopy, and N-terminal sequencing (Rocha et al., 1996). studied at the K. lactis TRP1 locus. Also, other than in S. Recently, the zymocin system itself has acquired agri- cerevisiae, knocking out genes essential for viability in a cultural interest for use as anti-yeast additive during model diploid background requires constant selective pressure to silage fermentations. Aerobic spoilage by lactic acid-utiliz- suppress the transitory nature of the diplophase in K. ing yeast genera (Pichia and Candida) was shown to be lactis. significantly reduced in laboratory-scale experiments using lactose and K. lactis killer strains that were defective for growth on lactic acid due to deletion of the K1PCK1 gene BIOENGINEERING WITH K. lactis (coding for PEPCK, a key enzyme for gluconeogenesis; Kitamaoto et al., 1998, 1999). Whether the k1/k2 plasmid pair can be exploited as an alternative vector system awaits K. lactis has received increasing attention from bioin- further analysis. Gene dosage effects (Ն50 per dustries. Attractive properties include its GRAS (generally haploid genome), mitotic stability, and a yeast host spec- regarded as safe) status (Bonekamp and Osterom, 1994), trum potentially broader than that of pKD1 are in favor of molecular genetic accessibility using both integrative and k1 and k2, and pilot expression has been achieved with highly stabilized episomal vector systems (Wesolowski- bacterial and viral genes (Meinhardt et al., 1994; Louvel et al., 1996; Hsieh and Da Silva, 1998), and excel- Schru¨ nder et al., 1996; Schickel et al., 1996). However, lent fermentative characteristics allowing for large-scale gene activation appears to be poor compared to nuclear industrial applications (Fleer, 1992; Swinkels et al., 1993). transcription (Schaffrath et al., 1996, 1999, 2000) and a Moreover, its capability of exporting the zymocin complex comparative study between a k1- and a pKD1-based ex- suggests that K. lactis is suitable for secretion of large pression approach has shown that, despite similar mRNA heterologous proteins. Thus, both the ␣␤ and the ␥ zymo- levels, the latter is superior with regards to total protein cin subunit secretion leader signals have been reproduc- yields (Reliene and Sasnauskas, 1997). Therefore, in ad- ibly used to export heterologous proteins such as human dition to increase in gene activation, another ascet to be interleukin 1␤ and serum albumin, mouse ␣-amylase, hep- optimized is translational initiation. Experimentally, this atitis C virus envelope E2, as well as bacte- should be approachable by gene shuffle-mediated UCS rial/fungal ␤-lactamase and xylanases to name a few (Fleer optimization and introduction of strong ribosome-binding et al., 1991; Tokunaga et al., 1992, 1997; Walsh et al., 1998; sites (Schaffrat et al., 1999, 2000; see above). Saliola et al., 1999; Mustilli et al., 1999; Durand et al., In a number of studies in which identical proteins were 1999). Other efficient secretion signals derive from yeast expressed in S. cerevisiae and K. lactis or other noncon-

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. 182 Schaffrath and Breunig ventional yeasts, the latter have proven superior both in phenotype) has been used to select for mutants affecting Ϫ product yield and in secretion efficiency (Bui et al., 1996; fermentative metabolism (Goffrini et al., 1989). Rag mu- Gellissen and Hollenberg, 1997; Fleer et al., 1991; Van tants depend on oxidative and are unable den Berg et al., 1990). These empirical findings suggest to grow on glucose when respiration is blocked. Most RAG that nonconventional yeasts have particular properties that genes were shown to be involved in controlling the glyco- are different from those of S. cerevisiae, making them lytic flux. They encode glycolytic , a low-affinity attractive hosts for recombinant protein production. glucose transporter, or regulators of glycolysis (reviewed in Breunig et al., 2000). In the absence of respiration inhib- itors many of the rag mutants grow normally. Thus, im- pairment of glycolysis does not have a general negative MOLECULAR PHYSIOLOGY AND influence on cell physiology in K. lactis, in contrast to the PRIMARY CARBON METABOLISM situation in S. cerevisiae. Several metabolic activities have been pointed out that circumvent the necessity for a high The Crabtree Effect glycolytic flux in K. lactis. First, respiratory enzymes are not subject to strong glucose repression (Mulder et al., One of the properties of K. lactis favorable for recom- 1995). Second, the pentose cycle can efficiently binant protein expression is its so-called Crabtree-negative bypass a block in glycolysis (Jacoby et al., 1993). Third, a phenotype (De Deken, 1966). Crabtree-positive yeasts high biosynthetic capacity seems to allow for efficient such as S. cerevisiae form ethanol under aerobic condi- biomass accumulation even at a very low glycolytic flux tions, thus reducing ATP and biomass yield compared to (Breunig et al., 2000). Crabtree-negative yeasts (Verduyn, 1991). Ethanol forma- tion can be avoided by sugar-limited cultivation but this Glucose Uptake results in low specific growth and protein synthesis rates. It was therefore important to understand the molecular Apparently, K. lactis is able to restrict the glycolytic flux basis of the Crabtree effect. by tightly regulating glucose uptake. Only one constitu- Insight was provided from studies of pyruvate metabo- tively expressed glucose transporter encoded by the HGT1 lism in K. lactis. A mutant in which pyruvate dehydroge- gene has been reported (Billard et al., 1996) but this nase was blocked by a mutation in the PDA1 gene (en- high-affinity transporter is insufficient to support fermen- coding its largest subunit) was phenotypically reversed tative growth and thus seems to have a low capacity. The into Crabtree positivity. This result indicated that K. lactis other known transporters are regulated at the transcrip- is able to produce ethanol under aerobiosis. Obviously, in tional level. The first one described encodes a low-affinity the wild type the capacity of Pdh allows channeling of all permease, Rag1p. Deletion of the RAG1 gene leads to a Ϫ glycolytic pyruvate into the TCA cycle by converting it into Rag phenotype and natural isolates have been shown to acetyl-CoA, and only when this step is down-regulated or contain a mutant rag1 allele (Goffrini et al., 1990). RAG1 blocked does the alternative enzyme pyruvate decarboxyl- is induced by several hexoses but induction is rather slow ase convert pyruvate into acetaldehyde and ethanol (Zee- (Chen et al., 1992). Possibly, oxygen limitation in high- man et al., 1998; Kiers et al., 1998). In addition, any density batch cultures contribute to full induction. ethanol eventually produced will be reassimilated effi- The other two known glucose transporters are Kht1p ciently due to the presence of an ethanol-inducible etha- and Kht2p. Kht1p is almost identical to Rag1p except for nol dehydrogenase in K. lactis (Mazzoni et al., 1992). the last two amino acids (Weirich et al., 1997) and is The difference between Crabtree-positive and similarly regulated (C. Milkowski and K. D. Breunig, sub- Crabtree-negative yeasts may thus be due to a difference mitted). In fact, the genes KHT1 and KHT2 replace the in Pdh and/or TCA cycle enzyme activity or to a difference RAG1 gene in some strains. Sequence analysis indicated in regulation of the glycolytic flux or both. that the tandemly arranged KHT genes had undergone a recombination event to give rise to the chimeric RAG1 RAG Genes and Glycolysis gene which is found in most laboratory strains in use (Weirich et al., 1997). Kht2p, like Rag1p/Kht1p, is able to Even in media containing high concentrations of glu- support fermentative growth and is expressed at high cose the glycolytic flux seems to be tightly controlled in K. levels in the absence of glucose. However, in high-glucose lactis. Resistance against antimycin A on glucose (Rag medium the KHT2 gene is repressed, indicating that

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. Genetics and Molecular Physiology of K. lactis 183

Kht1p is the major uptake systems under these conditions permease, respectively, are specific for lactose metabolism (C. Milkowski et al., submitted). Although strains lacking and these genes are sufficient to convert S. cerevisiae into ϩ all known transporters can still grow on glucose, none of a Lac yeast (Sreekrishna and Dickson, 1985). Four the remaining uptake systems that remain to be identified Gal4p-binding sites upstream of LAC4 and LAC12 control is able to support fermentative growth, indicating that the both genes in a coordinate fashion (Go¨decke et al., 1991) regulation of Rag1p (or Kht1p–Kht2p) is of major impor- (Fig. 2). In common with S. cerevisiae is the genetic tance in controlling glycolytic flux. organization and topology of the K. lactis GAL gene clus- ter, indicating evolutionary pressure on this particular ar- Glucose Repression rangement (Webster and Dickson, 1988). Whereas galactose induction does not depend on galac- Several lines of evidence suggest that a restriction of the tose metabolism, lactose induction requires lactose per- glycolytic flux may also contribute to the low sensitivity of mease and ␤-galactosidase activity (Fig. 2). In fact, intra- K. lactis to glucose repression. First, repression is stronger cellular galactose is evidently the activation signal for in strains containing the KHT1–KHT2 gene pair than in Gal4p (Zenke et al., 1996; Cardinali et al., 1997). Partly RAG1 strains. Second, overexpression of KHT1–KHT2 due to comparative studies in S. cerevisiae and K. lactis, further supports repression. Third, mutation of RAG1 or transduction of the galactose signal to the transcription KHT1–KHT2 abolishes glucose repression (Weirich et al., machinery is one of the best-understood examples of nu- 1997; Breunig, 1989). However, the sensitivity to glucose trient signaling in eukaryotes. A crucial finding was the repression is highly strain dependent and seems to be bifunctional nature of the K1GAL1 gene product, which influenced by multiple genetic loci. has galactokinase and an additional regulatory activity Gluconeogenic genes and genes involved in the utiliza- (Meyer et al., 1991). Both functions are separable by tion of alternative carbon sources have been shown to be mutation and only the regulatory function is essential for expressed at a very low level in high-glucose media, at least KlGal4p activation (Vollenbroich et al., 1999; Zenke et al., in the KHT1–KHT2 strain JA6, which has been used for 1996, 1999). In S. cerevisiae, the regulatory function is most studies on glucose repression (Breunig, 1989; Za- mainly conferred by the GAL3 product, which is highly chariae et al., 1993; Kuzhandaivelu et al., 1992; Weirich et similar to ScGal1p and KlGal1p but lacks enzymatic activ- al., 1997; Georis et al., 1999, 2000). However, except for ity (Bhat et al., 1992; Bajwa et al., 1988). the GAL1 gene (Dong et al., 1998), in none of these cases Gene shuffling between S. cerevisiae and K. lactis was does repression seem to be mediated by the K. lactis instrumental in elucidating the regulatory function of Kl- homologue of Mig1p (Georis et al., 1999, 2000), in con- GAL1 and ScGAL3 and provided genetic evidence for trast to S. cerevisiae. Thus, the regulators conferring glu- species-specific interaction between Gal1/Gal3p and cose repression in K. lactis remain to be identified. By Gal80p (Zenke et al., 1996). Gal80p is highly conserved screening for mutants that affect the utilization of glu- between S. cerevisiae and K. lactis (Zenke et al., 1993) and coneogenic carbon sources, two genes, FOG1 and FOG2, negatively regulates Gal4p in the absence of galactose by that affect the derepression of glucose-repressed genes, interacting with its activation domain. KlGal80p–KlGal1p were characterized. FOG1 turned out to encode the K. interaction was confirmed by in vitro studies (Zenke et al., lactis homologue of GAL83, and FOG2 is the K. lactis 1996) and two-hybrid experiments (Vollenbroich et al., SNF1 homologue (Goffrini et al., 1996). 1999). It depends on the regulatory, not the enzymatic activity, of Gal1p and requires galactose and ATP, sug- Lactose and Galactose Metabolism gesting that substrate binding to galactokinase leads to a conformational change that allows interaction with Gal80p The utilization of lactose is a characteristic trait for K. followed by activation of Gal4p (Fig. 2). Galactose phos- lactis and has been studied extensively. Lactose metabo- phorylation is not required for induction; instead, it leads lism is inducible and coregulated with galactose metabo- to inducer consumption, thus providing a direct link be- lism (Sheetz and Dickson, 1980; Wray et al., 1987; Web- tween metabolism and gene regulation (Zachariae, 1994). ster and Dickson, 1988) by a transcriptional activator, Whereas the basics of the “galactose switch” appear to be Lac9p or KlGal4p, homologous to Gal4p of S. cerevisiae essentially the same in S. cerevisiae and K. lactis (Platt et (Salmeron and Johnston, 1986; Ruzzi et al., 1987; Breunig al., 1998; Blank et al., 1997; Yano et al., 1997; Zenke et al., et al., 1987; Zachariae and Breunig, 1993). Two genes, 1996), there may be differences in the way the inhibitory LAC4 and LAC12, encoding ␤-galactosidase and lactose influence of Gal80p on Gal4p is relieved upon Gal80p–

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. 184 Schaffrath and Breunig

Gal1p/Gal3p interaction. ScGal4p is subject to carbon mechanisms of microbial competition, and vector devel- source-dependent phosphorylation (Hirst et al., 1999; My- opment. Supported by the developments of powerful mo- lin et al., 1989; Sadowski et al., 1996, 1991), whereas in K. lecular genetic tools, K. lactis has been successfully estab- lactis KlGal80p is phosphorylated (Zenke et al., 1999). A lished as an alternative system for biomass-directed indus- de- or hypophosphorylated form of the latter accumulates trial applications and large-scale expression of biotechnically upon induction and the KlGal1p regulatory function is relevant gene products. required for this shift, suggesting a functional significance of regulated phosphorylation. Possibly, this difference be- tween S. cerevisiae and K. lactis contributes to the fact that KlGal80p-mediated inhibition of ScGal4p or KlGal4p can- ACKNOWLEDGMENTS not be relieved by ScGal3p (Zenke et al., 1996). Regulation of lactose metabolism has also served as a We thank Lars Fichtner (University of Halle, Germany) who kindly model system for glucose repression. In this case glucose contributed in part to data illustrated in Fig. 1. The work was supported repression is primarily due to an inhibition of induction by grants from the EC Framework IV “Cell Factory” (BIO4-CT96-0003) (Zachariae et al., 1993; Kuzhandaivelu et al., 1992; Zenke to K.D.B. and the Deutsche Forschungsgemeinschaft (DFG) to R.S. et al., 1993; Zachariae, 1994). A low level of LAC12 gene (Scha750) and K.D.B. (Br921). In addition, R.S. acknowledges receipt of a Feodor-Lynen-Fellowship (FLF-DEU/1037031) from the Alexander expression and a Gal4p-inducible KlGAL80 gene is a pre- von Humboldt-Foundation, Germany. requisite for the glucose effect (Zachariae, 1994; Zenke et al., 1993). Repression is relieved by twofold elevated ex- pression of the activator gene LAC9/KlGAL4 (Zachariae et al., 1993; Kuzhandaivelu et al., 1992), indicating that there REFERENCES exists a delicate balance between the transcription activa- tor and its repressor KlGal80p. In contrast, glucose repres- sion of the GAL regulon in S. cerevisiae is retained in a Ahmed, A., Sesti, F., Ilan, N., Shih, T. M., Sturley, S. L., and Goldstein, S. A. 1999. A molecular target for viral killer toxin: TOK1 potassium gal80 mutant due to the activity of Mig1p (Nehlin et al., channels. Cell 99: 283–291. 1991). Obviously, through the improvement of lactose/ Astrom, S. U., and Rine, J. 1998. Theme and variation among silencing galactose metabolism in K. lactis, the Gal80 protein has proteins in and Kluyveromyces lactis. Ge- acquired a more important role in keeping the regulon netics 148: 1021–9102. silent. Bai, X., Larsen, M., and Meinhardt, F. 1999. The URA5 gene encoding orotate-phosphoribosyl of the yeast Kluyveromyces lactis: Cloning, sequencing and use as a selectable marker. Yeast 15: 1393– 1398. CONCLUSIONS Bajwa, W., Torchia, T. E., and Hopper, J. E. 1988. Yeast regulatory gene GAL3: Carbon regulation; UASGal elements in common with GAL1, GAL2, GAL7, GAL10, GAL80, and MEL1; encoded protein strikingly K. lactis has become an excellent model organism for similar to yeast and Escherichia coli galactokinases. Mol. Cell. Biol. 8: studying physiological aspects unique to unconventional 3439–3447. Crabtree-negative yeast species in which regulation of Bhat, P. J., and Hopper, J. E. 1992. Overproduction of the GAL1 or carbon and energy metabolism contrast with the excep- GAL3 protein causes galactose-independent activation of the GAL4 protein: Evidence for a new model of induction for the yeast GAL/ tional physiology of the classical yeast model S. cerevisiae. MEL regulon. Mol. Cell. Biol. 12: 2701–2707. In contrast to S. cerevisiae, which grows on glucose by Bianchi, M. M. 1992. Site-specific recombination of the circular 2 ␮m- alcoholic fermentation with the so-called glucose effect like plasmid pKD1 requires integrity of the recombinase A and of the constituting a major control network for many diverse partitioning genes B and C. J. Bacteriol. 174: 6703–6706. metabolic processes, the petite-negative yeast K. lactis is Bianchi, M. M., Santarelli, R., and Frontali, L. 1991. Plasmid functions adapted to aerobiosis, and its respiratory system does not involved in stable propagation of pKD1 circular plasmid in Kluyvero- underlie glucose repression. Nonetheless, its phylogenetic myces lactis. Curr. Genet. 19: 155–161. relatedness to S. cerevisiae has often allowed the delinea- Billard, P., Me´nart, S., Blaisonneau, J., Bolotin-Fukuhara, M., Fukuhara, H., and We´solowski-Louvel, M. 1996. Glucose uptake in Kluyveromy- tion of functional domains from sequence comparison of ces lactis: Role of the HGT1 gene in glucose transport. J. Bacteriol. homologous genes. Also, K. lactis maintains species-spe- 178: 5860–5866. cific phenomena such as the DNA-killer system, analysis Blank, T. E., Woods, M. P., Lebo, C. M., Xin, P., and Hopper, J. E. 1997. of which promises new insights into linear plasmid biology, Novel Gal3 proteins showing altered Gal80p binding cause constitutive

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. Genetics and Molecular Physiology of K. lactis 185

transcription of Gal4p-activated genes in Saccharomyces cerevisiae. Chen, X-J., and Clark-Walker, G. D. 1995. Specific in alpha- Mol. Cell. Biol. 17: 2566–2575. and gamma-subunits of F1-ATPase affect mitochondrial genome in- Bonekamp, F. J., and Oosterom, J. 1994. On the safety of Kluyveromyces tegrity in the petite-negative yeast Kluyveromyces lactis. EMBO J. 14: lactis: A review. Appl. Microbiol. Biotechnol. 41: 1–3. 3277–3286. Breunig, K. D. 1989. Glucose repression of LAC gene expression in yeast Chen, X-J., and Clark-Walker, G. D. 1996. The mitochondrial genome is mediated by the transcriptional activator LAC9. Mol. Gen. Genet. integrity gene, MGI1, of Kluyveromyces lactis encodes the beta-sub- 216: 422–427. unit of F1-ATPase. Genetics 144: 1445–1454. Breunig, K. D., and Kuger, P. 1987. Functional homology between the Chen, X-J., and Clark-Walker, G. D. 2000. The petite mutation in yeasts: yeast regulatory proteins GAL4 and LAC9: LAC9-mediated transcrip- 50 years on. Int. Rev. Cytol. 194: 197–238. tional activation in Kluyveromyces lactis involves protein binding to a Chen, X-J., and Fukuhara, H. 1988. A gene fusion system using the regulatory sequence homologous to the GAL4 protein-. aminoglycoside 3Ј- gene of the kanamycin-resis- Mol. Cell. Biol. 7: 4400–4406. tance transposon Tn903: Use in the yeast Kluyveromyces lactis and Breunig, K. D., Bolotin-Fukuhara, M., Bianchi, M., Bourgarel, D., Fal- Saccharomyces cerevisiae. Gene 69: 181–192. cone, C., Ferrero, I., Frontali, L., Goffrini, P., Krijger, J. J., Mazzoni, Chen, X-J., Wesolowski-Louvel, M., and Fukuhara, H. 1992. Glucose C., Milkowski, C., Steensma, Y., and Wesolowski-Louvel, M. 2000. transport in the yeast Kluyveromyces lactis. II. Transcriptional regu- Regulation of primary carbon metabolism in Kluyveromyces lactis. lation of the glucose transporter gene RAG1. Mol. Gen. Genet. 233: Enzyme Microbial Technol. 26: 771–780. 97–105. Bui, D. M., Kunze, I., Horstmann, C., Schmidt, T., Breunig, K. D., and Chen, X-J., Bianchi, M. M., Suda, K., and Fukuhara, H. 1989. The host Kunze, G. 1996. Expression of the Arxula Adeninivorans glucoamylase range of the pKD1-derived plasmids in yeast. Curr. Genet. 16: 95–98. gene in Kluyveromyces lactis. Appl. Microbiol. Biotechnol. 45: 102– Chen, X-J., Saliola, M., Falcone, C., Bianchi, M. M., and Fukuhara, H. 106. 1986. Sequence organization of the circular plasmid pKD1 from the Bundock, P., Mroczek, K., Winkler, A. A., Steensma, H. Y., and Hooy- yeast Kluyveromyces drosophilarum. Nucleic Acids Res. 14: 4471– kaas, P. J. 1999. T-DNA from Agrobacterium tumefaciens as an effi- 4481. cient tool for gene targeting in Kluyveromyces lactis. Mol. Gen. Genet. Cong, Y. S., Wesolowski-Louvel, M., and Fukuhara, H. 1994. Creation of 261: 115–121. a functional promoter by rearrangement in a Kluyveromyces lactis Butler, A. R., White, J. H., and Stark, M. J. R. 1991a. Analysis of the linear plasmid. Gene 147: 125–129. response of Saccharomyces cerevisae cells to Kluyveromyces lactis Cos, T., Ford, R., Trilla, J. A., Duran, A., Cabib, E., and Roncero, C. toxin. J. Gen. Microbiol. 137: 1749–1757. 1998. Molecular analysis of Chs3p participation in chitin synthase III Butler, A. R., Porter, M., and Stark, M. J. R. 1991b. Intracellular activity. Eur. J. Biochem. 256: 419–426. expression of Kluyveromyces lactis toxin ␥ subunit mimics treatment Cross, F. R., and Tinkelenberg, A. H. 1991. A potential positive feedback with exogenous toxin and distinguishes two classes of toxin-resistant loop controlling CLN1 and CLN2 gene expression at the start of the mutant. Yeast 7: 617–625. yeast cell cycle. Cell 65: 875–883. Butler, A. R., O’Donnell, R. W., Martin, V. J., Gooday, G. W., and Stark, Das, S., and Hollenberg, C. P. 1982. A high frequency transformation M. J. R. 1991c. Kluyveromyces lactis toxin has an essential chitinase system for the yeast Kluyveromyces lactis. Curr. Genet. 6: 123–128. activity. Eur. J. Biochem. 199: 483–488. De Deken, R. H. 1966. The crabtree effect: A regulatory system in yeast. Butler, A. R., White, J. H., Folawiyo, Y., Edlin, A., Gardiner, D., and J. Gen. Microbiol. 44: 149–156. Stark, M. J. R. 1994. Two Saccharomyces cerevisiae genes which Di Como, C. J., and Arndt, K. T. 1996. Nutrients, via the Tor proteins, control sensitivity to G1 arrest induced by Kluyveromyces lactis toxin. stimulate the association of Tap42 with type 2A phosphatases. Genes Mol. Cell. Biol. 14: 6306–6316. Dev. 10: 1904–1916. Cai, J., Roberts, I. N., and Collins, M. D. 1996. Phylogenetic relation- Dong, J., and Dickson, R. C. 1998. Glucose represses the lactose- ships among members of the ascomycetous yeast genera Brettanomy- galactose regulon in Kluyveromyces lactis through a SNF1 and MIG1- ces, Debaryomyces, Dekkera, and Kluyveromyces deduced by small- dependent pathway that modulates galactokinase (GAL1) gene expres- subunit rRNA gene sequences. Int. J. Syst. Bacteriol 46: 542–549. sion. Nucleic Acids Res. 25: 3657–3664. Cardinali, G., Vollenbroich, V., Jeon, M. S., de Graaf, A. A., and Hol- Durand, R., Rascle, C., and Fevre, M. 1999. Expression of a catalytic lenberg, C. P. 1997. Constitutive expression in gal7 mutants of domain of a Neocallimastix frontalis endoxylanase gene (xyn3)in Kluyveromyces lactis is due to internal production of galactose as an Kluyveromyces lactis and Penicillium roqueforti. Appl. Microbiol. Bio- inducer of the Gal/Lac regulon. Mol. Cell Biol. 17: 1722–1730. technol. 52: 208–214. Castrillo, J. I., Due-as, M., Bassy, O., Castrillo, J. L., and Ugalde, U. Erickson, J. R., and Johnston, M. 1994. Suppressors reveal two classes of 1999. Use of Aspergillus nidulans amdS gene as dominant selectable glucose repression genes in the yeast Saccharomyces cerevisiae. Ge- marker in non-conventional Kluyveromyces yeasts. Curr. Genet. 35: netics 136: 1271–1278. 447. Falcone, C., Saliola, M., Chen, X-J., Frontali, L., and Fukuhara, H. 1986. Chen, X-J. 1996. Low- and high-copy-number shuttle vectors for repli- Analysis of a 1.6-micron circular plasmid from the yeast Kluyveromy- cation in the budding yeast Kluyveromyces lactis. Gene 172: 131–136. ces drosophilarum: Structure and molecular dimorphism. Plasmid 15: Chen, X-J., and Clark-Walker, G. D. 1993. Mutations in MGI genes 248–252. convert Kluyveromyces lactis into a petite-positive yeast. Genetics 133: Fellows, J., Erdjument Bromage, H., Tempst, P., and Svejstrup, J. Q. 517–525. 2000. The elp2 subunit of elongator and elongating RNA polymerase

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. 186 Schaffrath and Breunig

II holoenzyme is a WD40 repeat protein. J. Biol. Chem. 275: 12896– Gunge, N., and Sakaguchi, K. 1981. Intergeneric transfer of deoxyribo- 12899. nucleic acid killer plasmids, pGKL1 and pGKL2, from Kluyveromyces Ferminan, E., and Dominguez, A. 1998. Heterologous protein secretion lactis into Saccharomyces cerevisiae by cell fusion. J. Bacteriol. 147: directed by a repressible acid phosphatase system of Kluyveromyces 155–160. lactis: Characterization of upstream region-activating sequences in the Gunge, N., Tamaru, A., Ozawa, F., and Sakaguchi, K. 1981. Isolation and KIPHO5 gene. Appl. Environ. Microbiol. 64: 2403–2408. characterization of linear deoxyribonucleic acid plasmids from Fleer, R. 1992. Engineering yeast for high level expression. Curr. Opin. Kluyveromyces lactis and the plasmid-associated killer character. J. Biotechnol. 3: 486–496. Bacteriol. 145: 382–390. Fleer, R., Yeh, P., Amellal, N., Maury, I., Fournier, A., Bacchetta, F., Hashida-Okado, T., Ogawa, A., Kato, I., and Takesako, K. 1998. Trans- Baduel, P., Jung, G., L’Hote, H., Becquart, J., Fukuhara, H., and formation system for prototrophic industrial yeasts using the AUR1 Mayaux, J-F. 1991. Stable multicopy vectors for high-level secretion of gene as a dominant selection marker. FEBS Lett. 425: 117–122. recombinant human serum albumin by Kluyveromyces yeasts. Bio- Herman, A. I., and Halvorson, H. O. 1963. Identification of the structural ␤ Technology 9: 968–975. gene for -glucosidase in Saccharomyces lactis. J. Bacteriol. 85: 895– Fukuhara, H. 1995. Linear DNA plasmids of yeasts. FEMS Microbiol. 900. Lett. 131: 1–9. Heus, J. J., Zonneveld, B. J. M., Steensma, H. Y., and van den Berg, J. A. Fulton, T. B., and Blackkburn, E. H. 1998. Identification of Kluyvero- 1990. Centromeric DNA of Kluyveromyces lactis. Curr. Genet. 18: myces lactis telomerase: Discontinuous synthesis along the 30-nucle- 517–522. otide-long templating domain. Mol. Cell. Biol. 18: 4961–4970. Heus, J. J., Zonneveld, B. J. M., Steensma, H. Y., and Van den Berg, J. A. Fuson, G. B., Presley, H. L., and Phaff, H. J. 1987. Deoxyribonucleic acid 1994. Mutational analysis of centromeric DNA elements of Kluyvero- base sequence relatedness among members of the genus Kluyveromy- myces lactis and their role in the species specificity of the highly homologous centromeres from K. lactis and Saccharomyces cerevisiae. ces. Int. J. Syst. Bacteriol. 37: 371–379. Mol. Gen. Genet. 243: 325–333. Gellissen, G., and Hollenberg, C. P. 1997. Application of yeasts in gene Hirst, M., Kobor, M. S., Kuriakose, N., Greenblatt, J., and Sadowski, I. expression studies: A comparison of Saccharomyces cerevisiae, Han- 1999. GAL4 is regulated by the RNA polymerase II holoenzyme- senula polymorpha and Kluyveromyces lactis—A review. Gene 190: associated cyclin-dependent SRB10/CDK8. Mol. Cell 3: 87–97. 673–678. Georis, I., Cassart, J.-P., Breunig, K. D., and Vandenhaute, J. 1999. Hsieh, H-B., and Da Silva, N. A. 1998. An autoselection system in Transcription of the Kluyveromyces lactis invertase gene INV1 is recombinant Kluyveromyces lactis enhances cloned gene stability and glucose-repressed independently from MIG1. Mol. Gen. Genet. 261: provides freedom in medium selection. Appl. Microbiol. Biotechnol. 862–870. 49: 147–152. Georis, I., Krijger, J.-J., Breunig, K. D., and Vandenhaute, J. 2000. Hsieh, H-B., and Da Silva, N. A. 2000. Development of a LAC4 pro- Difference in regulation of yeast gluconeogenesis revealed by Cat8p- moter-based gratuitous induction system in Kluyveromyces lactis. Bio- independent activation of PCK1 and FBP1 genes in Kluyveromyces technol. Bioeng. 67: 408–416. lactis. Mol. Gen. Genet., in press. Jacoby, J., Hollenberg, C. P., and Heinisch, J. J. 1993. Transaldolase Go¨decke, A., Zachariae, W., Arvanitidis, A., and Breunig, K. D. 1991. mutants in the yeast Kluyveromyces lactis provide evidence that glu- Coregulation of the Kluyveromyces lactis lactose permease and ␤-ga- cose can be metabolized through the pentose phosphate pathway. Mol. lactosidase genes is achieved by interaction of multiple LAC9 binding Microbiol. 10: 867–876. sites in a 2.6 kbp divergent promoter. Nucleic Acids Res. 19: 5351– Kawamoto, S., Nomura, M., and Ohno, T. 1992. Cloning and character- 5358. ization of SKT5, a Saccharomyces cerevisiae gene that affects proto- Goffrini, P., Algeri, A. A., Donnini, C., We´solowski-Louvel, M., and plast regeneration and resistance to killer toxin of Kluyveromyces Ferrero, I. 1989. RAG1 and RAG2: Nuclear genes involved in the lactis. J. Ferment. Bioeng. 74: 199–208. dependence/independence on mitochondrial respiratory function for Kawamoto, S., Sasaki, T., Itahashi, S., Hatsuyama, Y., and Ohno, T. 1993. growth on sugars. Yeast 5: 99–106. A mutant allele skt5 affecting protoplast regeneration and killer toxin Goffrini, P., We´solowski-Louvel, M., Ferrero, I., and Fukuhara, H. 1990. resistance has double mutations in its wild-type structural gene in RAG1 gene of the yeast Kluyveromyces lactis codes for a sugar trans- Saccharomyces cerevisiae. Biosci. Biotech. Biochem. 57: 1391–1393. porter. Nucleic Acids Res. 18: 5294. Ka¨mper, J., Esser, K., Gunge, N., and Meinhardt, F. 1991. Heterologous Goffrini, P., Ficarelli, A., Donnini, C., Lodi, T., Puglisi, P. P., and expression on the linear DNA killer plasmid from Kluyveromyces Ferrero, I. 1996. FOG1 and FOG2 genes, required for the transcrip- lactis. Curr. Genet. 19: 109–118. tional activation of glucose-repressible genes of Kluyveromyces lactis, Ka¨mper, J., Meinhardt, F., Gunge, N., and Esser, K. 1989. New recom- are homologous to GAL83 and SNF1 of Saccharomyces cerevisiae. binant linear DNA elements derived from Kluyveromyces lactis killer Curr. Genet. 29: 316–326. plasmids. Nucleic Acids Res. 17: 1781. Gunge, N. 1986. Linear DNA killer plasmids from the yeast Kluyvero- Keogh, R., Seoighe, S., and Wolfe, K. H. 1998. Evolution of gene order myces lactis. Yeast 2: 153–162. and chromosome number in Saccharomyces, Kluyveromyces and re- Gunge, N. 1995. Plasmid DNA and the killer phenomenon in Kluyvero- lated fungi. Yeast 14: 443–457. myces. In The Mycota II Genetics and Biotechnology (Ku¨ ck, Ed.), pp. Kiers, J., Zeeman, A.-M., Luttik, M., Thiele, C., Castrillo, J. I., Steensma, 189–209. Springer-Verlag, Berlin. H. Y., Van Dijken, J. P., and Pronk, J. T. 1998. Regulation of alcoholic

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. Genetics and Molecular Physiology of K. lactis 187

fermentation in batch and chemostat cultures of Kluyveromyces lactis maintains telomeres in Kluyveromyces lactis lacking telomerase. Genes CBS 2359. Yeast 14: 459–469. Dev. 10: 1822–1834. Kikuchi, Y., Hirai, K., Gunge, N., and Hisinuma, F. 1986. Hairpin McNeel, D. G., and Tamanoi, F. 1991. Terminal region recognition plasmid—A novel linear DNA with perfect hairpin structure. EMBO J. factor 1, a DNA binding protein recognising the ITRs of the pGKL 4: 1881–1886. linear plasmids. Proc. Natl. Acad. Sci. USA 88: 11398–11402. Kishida, M., Tokunaga, M., Katayosa, Y., Yajima, H., Kawamura-Watabe, Meilhoc, E., Masson, J. M., and Teissie, J. 1990. High efficiency trans- A., and Hishinuma, F. 1996. Isolation and genetic characterization of formation of intact yeast cells by electric field pulses. BioTechnology 8: pGKL killer-insensitive mutants (iki) from Saccharomyces cerevisiae. 223–227. Biosci. Biotech. Biochem. 60: 798–801. Meinhardt, F., Schaffrath, R., and Larsen, M. 1997. Microbial linear Kitada, K., and Gunge, N. 1988. Palindrome-hairpin linear plasmids plasmids. Appl. Microbiol. Biotechnol. 47: 329–336. possessing only part of the ORF1 gene of the yeast killer plasmid Meinhardt, F., Wodara, C., Larsen, M., and Schickel, J. 1994. A novel pGKL1. Mol. Gen. Genet. 215: 46–52. approach to express a heterologous gene Kluyveromyces lactis linear Kitamoto, H. K., Ohmomo, S., and Iimura, Y. 1998. Isolation and nucle- killer plasmids: Expression of the bacterial aph gene from a cytoplas- otide sequence of the gene encoding phosphoenolpyruvate carboxyki- mic promoter fragment not fused in frame. Plasmid 32: 318–327. nase from Kluyveromyces lactis. Yeast 14: 963–967. Mellor, J., Jiang, W., Funk, M., Rathjen, J., Barnes, C. A., Hinz, T., Kitamoto, H. K., Hasebe, A., Ohmomo, S., Suto, E. G., Muraki, M., and Hegemann, J. H., and Philippsen, P. 1990. CPF1, a yeast protein Iimura, Y. 1999. Prevention of aerobic spoilage of maize silage by a with functions in centromeres and promoters. EMBO J. 9: 4017– genetically modified killer yeast, Kluyveromyces lactis, defective in the 4026. ability to grow on lactic acid. Appl. Environ. Microbiol. 65: 4697– Meyer, J., Walker-Jonah, A., and Hollenberg, C. P. 1991. Galactokinase 4700. encoded by GAL1 is a bifunctional protein required for induction of Klebe, R. J., Harriss, J. V., Sharp, D., and Douglas, M. G. 1983. A general the GAL genes in Kluyveromyces lactis and is able to suppress the gal3 method for polyethyleneglycol-induced transformation of bacteria and phenotype in Saccharomyces cerevisiae. Mol. Cell. Biol. 11: 5454– yeast. Gene 25: 333–341. 5461. Krauskopf, A., and Blackburn, E. H. 1996. Control of telomere growth by Mulder, W., Scholten, I. H. J. M., and Grivell, L. A. 1995. Carbon interactions of RAP1 with the most distal telomeric repeats. Nature catabolite regulation of transcription of nuclear genes coding for mi- 383: 354–357. tochondrial proteins in the yeast Kluyveromyces lactis. Curr. Genet. Krauskopf, A., and Blackburn, E. H. 1998. Rap1 protein regulates telo- 28: 267–273. mere turnover in yeast. Proc. Natl. Acad. Sci. USA 95: 12486–12491. Mulder, W., Winkler, A. A., Scholten, I. H., Zonneveld, B. J., de Winde, Kurtzman, C. P., and Robnett, C. J. 1998. Identification and phylogeny J. H., Steensma, Y. H., and Grivell, L. A. 1994. Centromere promter of ascomycetous yeasts from analysis of nuclear large subunit (26S) factors (CPF1) of the yeasts of Saccharomyces cerevisiae and ribosomal DAN partial sequences. Antonie Van Leeuwenhoek 73: Kluyveromyces lactis are functionally exchangeable despite low overall 331–371. homology. Curr. Genet. 26: 198–207. Kuzhandaivelu, N., Jones, W. K., Martin, A. K., and Dickson, R. C. 1992. Mustilli, A. C., Izzo, E., Houghton, M., and Galeotti, C. L. 1999. The signal for glucose repression of the lactose-galactose regulon is Comparison of secretion of a hepatitis C virus glycoprotein in Saccha- amplified through subtle modulation of transcription of the Kluyvero- romyces cerevisiae and Kluyveromyces lactis. Res. Microbiol. 150: myces lactis K1GAL4 activator gene. Mol. Cell. Biol. 12: 1924–1931. Larsen, M., Gunge, N., and Meinhardt, F. 1998. Kluyveromyces lactis 179–187. killer plasmid pGKL2: Evidence for a viral-like capping enzyme en- Mylin, L. M., Bhat, J. P., and Hopper, J. E. 1989. Regulated phosphor- coded by ORF3. Plasmid 40: 243–246. ylation and dephosphorylation of GAL4, a transcriptional activator. Leberer, E., Thomas, D. Y., and Whiteway, M. 1997. Pheromone sig- Genes Dev. 3: 1157–1165. nalling and polarized morphogenesis in yeast. Curr. Opin. Genet. Dev. Nehlin, J. O., Carlberg, M., and Ronne, H. 1991. Control of yeast GAL 7: 59–66. genes by MIG1 repressor: A transcriptional cascade in the glucose Lloyd, A. T., and Sharp, P. M. 1993. Synonymous codon usage in response. EMBO J. 10: 3373–3377. Kluyveromyces lactis. Yeast 9: 1219–1228. Niwa, O., Sakaguchi, K., and Gunge, N. 1981. Curing of the killer Luke, M. M., Dellaseta, F., Di Como, C. J., Sugimoto, H., Kobayashi, R., dexyribonucleic acid plasmids of Kluyveromyces lactis. J. Bacteriol. and Arndt, K. T. 1996. The SAP, a new family of proteins, associate 148: 988–990. and function positively with the SIT4 phosphatase. Mol. Cell. Biol. 16: Osinga, K. A., De Haan, M., Christianson, T., and Tabak, H. F. 1982. A 2744–2755. nonanucleotide sequence involved in promotion of ribosomal RNA Mazzoni, C., Saliola, M., and Falcone, C. Ethanol-induced and glucose- synthesis and RNA priming of DNA replication in yeast. Nucleic Acids insensitive dehydrogenase in the yeast Kluyveromyces lactis. Res. 10: 7993–8006. 1992. Mol. Microbiol. 6: 2279–2286. Otero, G., Fellows, J., Li, Y., de Bizemont, T., Dirac, A. M., Gustafsson, McEachern, M. J., and Blackburn, E. H. 1994. A C. M., Erdjument-Bromage, H., Tempst, P., and Svejstrup, J. Q. 1999. motif within the exceptionally diverse telomeric sequences of budding Elongator, a multisubunit component of a novel RNA polymerase II yeasts. Proc. Natl. Acad. Sci. USA 91: 3453–3457. holoenzyme for transcriptional elongation. Mol. Cell 3: 109–118. McEachern, M. J., and Blackburn, E. H. 1996. Cap-prevented recom- Ozier-Kalogeropoulos, O., Malpertuy, A., Boyer, J., Tekaia, F., and bination between terminal telomeric repeat arrays (telomere CPR) Dujon, B. 1988. Random exploration of the Kluyveromyces lactis

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. 188 Schaffrath and Breunig

genome and comparison with that of Saccharomyces cerevisiae. Nu- Schaffrath, R., Meinhardt, F., and Meacock, P. A. 1997a. ORF7 of yeast cleic Acids Res. 26: 5511–5524. plasmid pGKL2: Analysis of gene expression in vivo. Curr. Genet. 31: Platt, A., and Reece, R. J. 1998. The yeast galactose genetic switch is 190–192. mediated by the formation of a Gal4p–Gal80p–Gal3p complex. EMBO Schaffrath, R., Meinhardt, F., and Meacock, P. A. 1999. Genetic manip- J. 17: 4086–4091. ulation of Kluyveromyces lactis linear DNA plasmids: Gene tageting Reliene, R., and Sasnauskas, K. 1997. Heterologous gene expression on and plasmid shuffles. FEMS Microbiol. Lett. 172: 201–210. the linear DNA plasmids of Kluyveromyces lactis. Yeast 13: S227. Schaffrath, R., Sasnauskas, K., and Meacock, P. A. 2000. Use of gene Rocha, T. L., Paterson, G., Crimmins, K., Boyd, A., Sawyer, L., and shuffles to study the cytoplasmic transcription system operating on Fothergill-Gilmore, L. A. 1996. Expression and secretion of recombi- Kluyveromyces lactis linear DNA plasmids. Enzyme Microbial. Tech- nant ovine beta-lactoglobulin in Saccharomyces cerevisiae and nol. 26: 664–670. Kluyveromyces lactis. Biochem. J. 313: 927–932. Schaffrath, R., Soond, S. M., and Meacock, P. A. 1995a. The DNA and Romanos, M. A., and Boyd, A. 1988. A transcriptional barrier to expres- RNA polymerase genes of yeast plasmid pGKL2 are essential loci sion of cloned toxin genes of the linear plasmid k1 of Kluyveromyces for plasmid integrity and maintenance. Microbiology 141: 2591– lactis: Evidence that native k1 has novel promoters. Nucleic Acids Res. 2599. 16: 7333–7350. Schaffrath, R., Soond, S. M., and Meacock, P. A. 1995b. Cytoplasmic Roy, J., Fulton, T. B., and Blackburn, E. H. 1998. Specific telomerase gene expression in yeast: A plasmid-encoded transcription system in RNA residues distant from the template are essential for telomerase Kluyveromyces lactis. Biochem. Soc. Trans. 23: 128. function. Genes Dev. 12: 3286–3300. Schaffrath, R., Stark, M. J. R., and Struhl, K. 1997b. Toxin-mediated cell Ruzzi, M., Breunig, K. D., Ficca, A. G., and Hollenberg, C. P. 1987. cycle arrest in yeast: The killer phenomenon of Kluyveromyces lactis. Positive regulation of the ␤-galactosidase gene from Kluyveromyces BIOforum Int. 1: 83–85. lactis is mediated by an upstream activation site that shows homology Schaffrath, R., Stark, M. J. R., Gunge, N., and Meinhardt, F. 1992. to the GAL upstream activation site of Saccharomyces cerevisiae. Mol. Kluyveromyces lactis killer system: ORF1 of pGKL2 has no function in Cell. Biol. 7: 992–997. immunity expression and is dispensable for killer plasmid replication Sadowski, I., Cosat, C., and Dhanawansa, R. 1996. Phosphorylation of and maintenance. Curr. Genet. 21: 357–363. Gal4p at a single C-terminal residue is necessary for galactose-induc- Schickel, J., Helmig, C., and Meinhardt, F. 1996. Kluyveromyces lactis ible transcription. Mol. Cell. Biol. 16: 4879–4887. killer system: Analysis of cytoplasmic promoters of the linear plasmids. Sadowski, I., Niedbala, D., Wood, K., and Ptashne, M. 1991. GAL4 is Nucleic Acids Res. 24: 1879–1886. phosphorylated as a consequence of transcriptional activation. Proc. Schru¨ nder, J., Gunge, N., and Meinhardt, F. 1996. Extranuclear expres- Natl. Acad. Sci. USA 88: 10510–10514. sion of the bacterial xylose- (xylA) and UDP-glucose-dehy- Salas, M. 1991. Protein priming of DNA replication. Annu. Rev. Bio- drogenase (hasB) genes in yeast using Kluyveromyces lactis linear chem. 60: 39–71. killer plasmids as vectors. Curr. Microbiol. 33: 323–330. Saliola, M., Mazzoni, C., Solimando, N., Crisam, A., Falconem, C., Seoighe, C., and Wolfe, K. H. 1998. Extent of genomic rearrangement Jungm, G., and Fleer, R. 1999. Use of the KlADH4 promoter for after genome duplication in yeast. Proc. Natl. Acad. Sci. USA 95: ethanol-dependent production of recombinant human serum albumin 4447–4452. in Kluyveromyces lactis. Appl. Environ. Microbiol. 65: 53–60. Sheetz, R. M., and Dickson, R. C. 1980. Mutations affecting synthesis of Salmeron, J. M., and Johnston, S. A. 1986. Analysis of the Kluyveromyces ␤-Galactosidase activity in the yeast Kluyveromyces lactis. Genetics 95: lactis positive regulatory gene LAC9 reveals functional homology to, 877–890. but sequence divergence from, the Saccharomyces cerevisiae GAL4 Shuster, J. R., Moyer, D., and Irvine, B. 1987. Sequence of the Kluyvero- gene. Nucleic Acids Res. 14: 7767–7781. myces lactis URA3 gene. Nucleic Acids Res. 15: 8573. Santos, B., and Snyder, M. 1997. Targeting of chitin synthase 3 to Smith, C. D., and Blackburn, E. H. 1999. Uncapping and deregulation of polarized growth sites in yeast requires Chs5p and Myo2p. J. Cell Biol. telomeres lead to detrimental cellular consequences in yeast. J. Cell 136: 95–110. Biol. 145: 203–214. Schaffrath, R. 1995. The Killer System of Kluyveromyces lactis: Molecular Sor, F., and Fukuhara, H. 1989. Analysis of chromosome patterns of the Genetic Analysis of Killer Plasmid k2 Gene Function. PhD thesis, genus Kluyveromyces. Yeast 5: 1–10. University of Leicester, UK. Sreekrishna, K., and Dickson, R. C. 1985. Construction of strains of Schaffrath, R., and Meacock, P. A. 1995. Kluyveromyces lactis killer Saccharomyces cerevisiae that grow on lactose. Proc. Natl. Acad. Sci. plasmid pGKL2: Molecular analysis of an essential gene, ORF5. Yeast USA 82: 7909–7913. 11: 615–628. Sreekrishna, K., Webster, T. D., and Dickson, R. C. 1984. Transforma- Schaffrath, R., and Meacock, P. A. 1996. A cytoplasmic gene shuffle tion of Kluyveromyces lactis with the kanamycin (G418) resistance system in Kluyveromyces lactis: Use of epitope-tagging to detect a gene of Tn903. Gene 28: 73–81. killer plasmid encoded gene product. Mol. Microbiol. 19: 545–554. Stark, M. J. R. 1996. Yeast protein /threonine phosphatases: Mul- Schaffrath, R., Meinhardt, F., and Meacock, P. A. 1996. Yeast killer tiple roles and diverse regulation. Yeast 12: 1647–1675. plasmid pGKL2: Molecular analysis of UCS5, a cytoplasmic promoter Stark, M. J. R., and Boyd, A. 1986. The killer toxin of Kluyveromyces element essential for ORF5 gene function. Mol. Gen. Genet. 250: lactis: Characterization of the toxin subunits and identification of the 286–294. genes which encode them. EMBO J. 5: 1995–2002.

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. Genetics and Molecular Physiology of K. lactis 189

Stark, M. J. R., and Milner, J. S. 1989. Cloning and analysis of the K. ianus and Kluyveromyces lactis from dairy products. J. Food Prot. 62: lactis TRP1 gene: A chromosomal locus flanked by genes encoding 189–193. inorganic pyrophosphatase and histone H3. Yeast 5: 35–50. Van den Berg, J. A., Van der Laken, K. J., Van Ooyen, A. J. J., Renniers, Stark, M. J. R., Boyd, A., Mileham, A. J., and Romanos, M. A. 1990. The C. H. M., Rietveld, K., Schaap, A., Brake, A. J., Bishop, R. J., Schultz, plasmid-encoded killer system of Kluyveromyces lactis: A review. Yeast K., Moyer, D., Richman, M., and Shuster, J. R. 1990. Kluyveromyces 6: 1–29. as a host for heterologous gene expression: Expression and secretion of Sugisaki, Y., Gunge, N., Sakaguchi, K., Yamasaki, M., and Tamura, G. prochymosin. BioTechnology 8: 135–139. 1984. Characterization of a novel killer toxin encoded by a double- Vaughan-Martini, A., and Martini, A. 1987. Taxonomic revision of the stranded linear DNA killer plasmid of Kluyveromyces lactis. Eur. yeast genus Kluyveromyces by nuclear deoxyribonucleic acid reasso- J. Biochem. 141: 241–245. ciation. Int. J. Syst. Bacteriol. 37: 380–385. Swinkels, B. W., van Ooyen, A. J. J., and Bonekamp, F. J. 1993. The yeast Verduyn, C. 1991. Physiology of yeasts in relation to biomass yields. Kluyveromyces lactis as an efficient host for heterologous gene expres- Antonie Van Leeuwenhoek 60: 325–353. sion. Antonie Van Leeuwenhoek 64: 187–201. Vollenbroich, V., Meyer, J., Engels, R., Cardinali, G., Menezes, R. A., Takita, M., and Castilho-Valavicius, B. 1993. Absence of cell wall chitin and Hollenberg, C. P. 1999. Galactose induction in yeast involves in Saccharomyces cerevisiae leads to resistance to Kluyveromyces lactis association of Gal80p with Gal1p or Gal3p. Mol. Gen. Genet. 261: killer toxin. Yeast 9: 589–598. 495–507. Tanguy-Rougeau, C., Wesolowski-Louvel, M., and Fukuhara, H. 1988. Walsh, D. J., Gibbs, M. D., and Bergquist, P. L. 1998. Expression and The Kluyveromyces lactis KEX1 gene encodes a subtilisin-type serin secretion of a xylanase from the extreme thermophile, Thermotoga proteinase. FEBS Lett. 234: 464–470. strain FjSS3B.1, in Kluyveromyces lactis. Extremophiles 2: 9–14. Tanguy-Rougeau, C., Chen, X. J., Wesolowski-Louvel, M., and Fuku- Webster, T. D., and Dickson, R. C. 1988. The organization and tran- hara, H. 1990. Expression of a foreign KmR gene in a linear killer DNA scription of the galactose gene cluster of Kluyveromyces lactis. Nucleic plasmid in yeast. Gene 91: 43–50. Acids Res. 16: 8011–8028. Thompson, A., and Oliver, S. G. 1986. Physical separation and functional Weirich, J., Goffrini, P., Kuger, P., Ferrero, I., and Breunig, K. D. 1997. interaction of Kluyveromyces lactis and Saccharomyces cerevisiae ARS Influence of mutations in hexose-transporter genes on glucose repres- elements derived from killer plasmid DNA. Yeast 2: 179–191. sion in Kluyveromyces lactis. Eur. J. Biochem. 249: 248–257. Tokunaga, M., Wada, N., and Hishinuma, F. 1987. A novel yeast secre- Wesolowski-Louvel, M., Breunig, K. D., and Fukuhara, H. 1996. tion vector utilizing secretion signal of killer toxin encoded on the yeast Kluyveromyces lactis In Nonconventional Yeasts in Biotechnology (K. linear DNA plasmid pGKL1. Biochem. Biophys. Res. Commun. 144: Wolf, Ed.), pp. 139–201. Springer-Verlag, Berlin. 613–619. Wesolowski-Louvel, M., Tanguy-Rougeau, C., and Fukuhara, H. 1988. A Tokunaga, M., Wada, N., and Hishinuma, F. 1988. A novel yeast secre- nuclear gene required for the expression of the linear DNA-associated tion signal isolated from 28K killer precursor protein encoded on the killer system in the yeast Kluyveromyces lactis. Yeast 4: 71–81. linear DNA plasmid pGKL1. Nucleic Acids Res. 16: 7499–7511. White, J. H., Butler, A. R., and Stark, M. J. R. 1989. Kluyveromyces lactis Tokunaga, M., Wada, N., and Hishinuma, F. 1989. Expression of pGKL toxin does not inhibit yeast adenylyl cyclase. Nature 341: 666–668. killer 28K subunit in Saccharomyces cerevisiae: Identification of 28K Wilson, D. W., and Meacock, P. A. 1988. Extranuclear gene expression subunit as a killer protein. Nucleic Acids Res. 17: 3435–3446. in yeast: Evidence for a plasmid-encoded RNA polymerase of unique Tokunaga, M., Kawamura, A., Kitada, K., and Hishinuma, F. 1990. structure. Nucleic Acids Res. 16: 8097–8112. Secretion of killer toxin encoded on the linear DNA plasmid Winkler, A. A., Bobok, A., Zonneveld, B. J. M., Steensma, H. Y., and pGKL1 from Saccharomyces cerevisae. J. Biol. Chem. 265: 17274– Hooykans, P. J. J. 1998. The lysine-rich C-terminal repeats of the 17280. centromere-binding factor (Cbf5) of Kluyveromyces lactis are not Tokunaga, M., Kawamura, A., Omori, A., and Hishinuama, F. 1992. essential for function. Yeast 14: 37–48. Characterization of mouse ␣-amylase secreted from Saccharomyces Wittschieben, B. O., Otero, G., de Bizemont, T., Fellows, J., Erdjument- cerevisae by the pGKL 128 kDa killer secretion signal. Biosci. Biotech. Bromage, H., Ohba, R., Li, Y., Allis, C. D., Tempst, P., and Svejstrup, Biochem. 56: 155–156. J. Q. 1999. A novel histone acetyltransferase is an integral subunit of Tokunaga, M., Ishibashi, M., Tatsuda, D., and Tokunaga, H. 1997. elongating RNA polymerase II holoenzyme. Mol. Cell 4: 123–128. Secretion of mouse ␣-amylase from Kluyveromyces lactis. Yeast 13: Wolfe, K. H., and Shields, D. C. 1997. Molecular evidence for an ancient 699–706. duplication of the entire yeast genome. Nature 387: 708–713. Trilla, J. A., Duran, A., and Roncero. 1999. Chs7p, a new protein Wray, L. V. J., Witte, M. W., Dickson, R. C., and Riley, M. I. 1987. involved in the control of protein export from the endoplasmic retic- Characterization of a positive regulatory gene, LAC9, that controls ulum that is specifically engaged in the regulation of chitin synthesis in induction of the lactose-galactose regulon of Kluyveromyces lactis: Saccharomyces cerevisiae. J. Cell Biol. 145: 1153–1163. Structural and functional relationships to GAL4 of Saccharomyces Trilla, J. A., Cos, T., Duran, A., and Roncero. 1997. Characterization of cerevisiae. Mol. Cell. Biol. 7: 1111–1121. CHS4 (CAL2), a gene of Saccharomyces cerevisiae involved in chitin Yajima, H., Tokunaga, M., Nakayama-Murayama, A., and Hishinuma, F. biosynthesis and allelic to SKT5 and CSD4. Yeast 13: 795–807. 1997. Characterization of IKI1 and IKI3 genes conferring pGKL killer Valderrama, M-J., De Siloniz, M. I., Gonzalo, P., and Peinado, J. M. sensitivity on Saccharomyces cerevisiae. Biosci. Biotech. Biochem. 61: 1999. A differential medium for the isolation of Kluyveromyces marx- 704–709.

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. 190 Schaffrath and Breunig

Yano, K.-I., and Fukasawa, T. 1997. Galactose-dependent reversible ylation of the Gal4p inhibitor Gal80p of Kluyveromyces lactis revealed interaction of Gal3p with Gal80p in the induction pathway of Gal4p- by mutational analysis. Biol. Chem. 380: 419–430. activated genes of Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. Zenke, F. T., Zachariae, W., Lunkes, A., and Breunig, K. D. 1993. Gal80 USA 94: 1721–1726. proteins of Kluyveromyces lactis and Saccharomyces cerevisiae are Zachariae, W. 1994. Regulation des Hefetranskriptionsaktivators Lac9. highly conserved but contribute differently to glucose repression of the PhD thesis, Heinrich-Heine-Universita¨t Du¨ sseldorf, Germany. galactose regulon. Mol. Cell. Biol. 13: 7566–7576. Zachariae, W., and Breunig, K. D. 1993. Expression of the transcriptional Zenke, F. T., Engles, R., Vollenbroich, V., Meyer, J., Hollenberg, C. P., activator LAC9 (KlGAL4) in Kluyveromyces lactis is controlled by and Breunig, K. D. 1996. Activation of Gal4p by galactose-dependent autoregulation. Mol. Cell. Biol. 13: 3058–3066. interaction of galactokinase and Gal80p. Science 272: 1662–1665. Zachariae, W., Kuger, P., and Breunig, K. D. 1993. Glucose repression of Zhang, Y-P., Chen, X-J., Li, Y-Y., and Fukuhara, H. 1992. LEU2 gene lactose/galactose metabolism in Kluyveromyces lactis is determined by homolog in Kluyveromyces lactis. Yeast 8: 801–804. the concentration of the transcriptional activator LAC9 (KlGAL4). Ziman, M., Chuang, J. S., and Schekman, R. 1996. Chs1p and Chs3p, two Nucleic Acids Res. 21: 69–77. proteins involved in chitin synthesis, populate a compartment of the Zeeman, A. M., Luttik, M. A., Thiele, C., Van Dijken, J. P., Pronk, J. T., Saccharomyces cerevisiae endocytic pathway. Mol. Biol. Cell 7: 1909– and Steensma, H. Y. 1998. Inactivation of the Kluyveromyces lactis 1919. KlPDA1 gene leads to loss of pyruvate dehydrogenase activity, impairs Ziman, M., Chuang, J. S., Tsung, M., Hamamoto, and Schekman, R. growth on glucose and triggers aerobic alcoholic fermentation. Micro- 1998. Chs6p-dependent anterograde transport of Chs3p from chito- biology. 144: 3437–3446. somes to the plasma membrane in Saccharomyces cerevisiae. Mol. Zenke, F. T., Kapp, L., and Breunig, K. D. 1999. Regulated phosphor- Biol. Cell 9: 1565–1576.

Copyright © 2000 by Academic Press All rights of reproduction in any form reserved.