Molecular Studies of Longevity-Associated Genes in Yeast and Mammalian Cells by Gregory Liszt B.S. Molecular, Cellular, and Developmental Biology Yale University Submitted to the Department of Biology in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY IN BIOLOGY at the Massachusetts Institute of Technology February, 2006 ©2005 by Gregory Liszt. All rights reserved. The author hereby grants to MIT permission to reproduce and distribute publicly paper and electronic copies of this thesis document in whole or in part. I \ Signature of Author 1 ";'-V Certified by - of Biology 1-/~'epartment x Leonard Guarente Novartis Professor of Biology Thesis Supervisor byEm., " Accepted Stephen Bell Chairman, Biology Graduate Committee .A . A. U. IT ,, , ,, ,,, - I MASSACHusEtTTS INS'-"T!iE' OF TECHNOLOGY I Ii NnVnv 4 9flfnsi 1 LIBRARIES ARCHIVES Molecular Studies of Longevity-Associated Genes in Yeast and Mammalian Cells by Gregory Liszt Submitted to the Department of Biology on October 25, 2005 in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in Biology ABSTRACT Aging is a complex process affecting diverse organisms from bacteria to humans. Despite strong evolutionary arguments against the conservation of a single mechanism of aging, a variety of conserved single gene mutations have been found to extend life span and stave off aging in several different species. The study of these mutations yields important insights into the biology of aging. In Saccharomyces cerevisiae, aging can be studied by mutations that extend the replicative potential of mother cells. With successive cell divisions, instability at the rDNA locus and extrachromosomal rDNA circle accumulation exponentially increase the likelihood of senescence and mortality. Aging can be forestalled by caloric restriction, a regimen that increases the activity of Sir2p, an NAD-dependent protein deacetylase and important regulator of aging in yeast and some metazoans. Caloric restriction activates respiration, reducing cellular NADH levels and relieving the competitive inhibition of Sir2p by this metabolite. SSD1 promotes longevity by a Sir2p-independent mechanism that affects neither ERC formation nor rDNA silencing. Ssdlp directly represses the translation of the mitochondrial and cell wall glycoprotein Uthl. This repression, which requires a physical interaction between Ssdlp and the 5'-UTR of the UTH1 mRNA, is necessary and sufficient to account for diverse effects of Ssdlp on cell integrity, stress resistance, and life span. Future studies should determine whether Ssdl plays a role in the maintenance of longevity in higher organisms. Mammalian genomes contain seven Sir2 homologs (SIRT1-7) involved in diverse processes including fat and muscle cell differentiation, p53- and FOXO-dependent apoptosis, stress resistance, and DNA-damage repair. Mouse SIRT6, a nuclear protein, is 3 broadly expressed throughout the body and displays a robust auto-ADPribosyltransferase activity unique among Sir2 family members. SIRT7, a nucleolar homolog of Sir2, physically interacts with RNA polymerase I (Pol I), colocalizing with the Pol I complex at the transcribed regions of the rDNA. SIRT7 activates rDNA transcription by an enzymatic mechanism, suggesting a novel model coordinating cellular energy status with ribosome biogenesis via changes in SIRT7 activity. Thesis Supervisor: Leonard Guarente Title: Novartis Professor of Biology, MIT 4 DEDICATION This thesis is dedicated to my parents, Miki and Harvey Liszt. 5 ACKNOWLEDGEMENTS First, I thank my thesis advisor Leonard Guarente for his guidance, wisdom, patience and understanding, all of which were indispensable to my success in graduate school. I am very lucky to have had you as a teacher. I would also like to thank my thesis committee: Rick Young, Peter Sorger, Paul Garrity, and David Sinclair. I thank Matt Kaeberlein for scientific insight, project ideas, collaboration, and general inspiration. Profound thanks go to the members of Crooked Still (Rushad Eggleston, Corey DiMario, and Aoife O'Donovan) for inspiring me musically and making these years in grad school so unbelievable and adventurous. Special thanks go to Ethan Ford for constantly sharing his insights, teaching me so many lab techniques, collaborating with me scientifically, and keeping me company for so many years. We had a good run of it. Special thanks also go to Nick Bishop, whose passion for the science of aging and beatboxing skills have always kept me inspired. Sorry to have abandoned you so close to the end. And thanks for the help with the Xmas Rap. Thanks to Andy Tolonen, my roommate, classmate, running partner, and good friend. Thanks to Kayvan Zainabadi for scientific insights and personal advice. Thanks to Martin Kurtev for always bringing life to the lab when no one else was around. I would also like to thank the entire Guarente lab for scientific and social camaraderie over the last six years. I have been lucky to share the company of so many brilliant and helpful people. Especially my collaborators, Matt Kaeberlein, Su-Ju Lin, Ethan Ford, and Martin Kurtev, without whom I would never have finished this thesis. Thanks to Sarah Buckley for being a great UROP. Thanks to all the musicians who have enriched my life here in Boston, especially Jake Armerding, Casey Driessen, the Wayfaring Strangers, and everyone from the Cantab. And most importantly, thanks to my family: Miki and Harvey Liszt, for their profound love and support; Jeff, for all that plus being a best friend; Danielle, for being such an inspiring future sister-in-law; Nonni for being such an amazing person and loving grandmother; and Aunt Mimi and Uncle Bev for always being there for me. 6 TABLE OF CONTENTS Title 1 Abstract 3 Dedication 5 Ackowledgements 6 Table of Contents 7 Chapter 1: 11 Introduction - Genetic Pathways of Aging from Yeast to Mammals Why study aging in model organisms? 12 Aging in Yeast 14 Early Studies Extrachromosomal rDNA Circles: A Cause of Aging Sir2p, a Central Regulator of Aging Caloric Restriction and the Regulation of Sir2p Polymorphisms Affecting Aging Chronological Aging Conservation of Aging Pathways in Metazoans Insulin/IGF-1 Pathway 14 17 19 23 27 32 33 33 Sir2 and Calorie Restriction 35 Other Sirtuins Summary and Conclusions 39 41 References 43 Figures 54 Chapter 2: 61 Ssdlp Directly Represses Translation of Uthlp to Increase Longevity, Stress Resistance, and Cell Wall Stability in Saccharomyces cerevisiae Summary Introduction Materials and Methods 62 63 66 Results 72 Discussion References 77 85 Tables and Figures 91 7 Chapter 3: 107 Mouse Sir2 Homolog SIRT6 is a Nuclear ADP-Ribosyltransferase Summary Introduction 108 109 Materials and Methods 112 Results Discussion References Figures 117 123 126 130 Chapter 4: 143 Conclusions Ssdlp and Uthlp 144 Sir2 and CR SIRT6 and SIRT7 References 147 148 150 Appendix A: 153 Saccharomyces cerevisiae SSDI-V Promotes Longevity by a Sir2pIndependent Mechanism Summary Introduction Materials and Methods Results Discussion References Tables and Figures 154 155 157 159 165 169 174 8 Appendix B: 193 Calorie Restriction Extends Life Span by Lowering the Level of NADH Summary Introduction 194 195 Results and Discussion Materials and Methods References Tables and Figures 197 201 204 207 Appendix C: 212 The Mammalian Sir2 Protein SIRT7 is an Activator of RNA Polymerase I Transcription Summary 213 Results and Discussion 214 Materials and Methods References 222 227 Figures 230 9 10 Chapter 1 IntroductionGenetic Pathways of Aging from Yeast to Mammals 11 WHY STUDY AGING IN MODEL ORGANISMS? Although we all possess an intuitive understanding of what it means to age, the scientific study of aging has proven notoriously complex. Aging, after all, is characterized by the modification or breakdown of nearly every system within an organism, and distinguishing the important changes from the unimportant ones poses a daunting problem. Indeed, aging differs drastically from most other basic biological processes in that aging is primarily a nonadaptive trait, conferring no benefit to the individual. As such traits arise precisely because of the lack of natural selection, there is no reason to believe that natural selection would conserve a particular mechanism of aging throughout evolution. For this reason, the study of aging in model organisms, which began in earnest in the 1990's, was met with skepticism (and even some ridicule). In the wild, very few organisms even survive to reach old age (reviewed in (Kirkwood, 2005), instead dying of causes such as predation, starvation, cold, accident, etc. Therefore, throughout evolution very little pressure has selected for alleles that confer benefits late in life. Evolutionary theory thus holds that aging should consist of the incidental failure of several systems rather than the onset of a specific program controlled by particular genes. The discovery of numerous conserved single gene mutations slowing aging and extending life span thus opened a rift between evolutionary theory and experimental reality. As a result, the control of aging has been rethought in evolutionary terms, and is now commonly regarded as possessing significant adaptive benefit. Which is not to say that aging results from the execution of a specific genetic program, merely that the generalized decay of aging can be prevented by means that have evolved over time. 12 Consider the case of dietary restriction, discussed in detail later in this chapter: This intervention, which entails reducing excess food consumption, has been shown to extend the life span of a variety of organisms (Guarente and Picard, 2005; Merry, 2002), likely involving by activating a conserved genetic mechanism. One hypothesis explaining this finding is that organisms with the ability to stave off aging and postpone reproduction during times of nutrient scarcity are at a tremendous selective advantage over those that maintain a constant rate of aging under such conditions. Consistent with this hypothesis, calorie-restricted rodents show signs of increased physiologic maintenance function and reduced reproduction (Weindruch et al., 2002), both of which would prime them for Darwinian success should food conditions improve and their physiology return to normal. This chapter details several of the main advances in model systems aging research, also addressing the extent to which genetic control of aging is conserved between species. Genomic stability, metabolism, and stress resistance appear as common themes in aging from yeast to mammals. In metazoans, control of these processes is superimposed against hormonal signaling pathways, with endocrine regulation emerging as a key means of regulating the rate of aging. In the chapters that follow, the goal of the research is two-fold: first, to arrive at a more complete understanding of aging in the unicellular model organism Saccharomyces cerevisiae, and second, to begin to apply the knowledge of yeast aging to the more complex mammalian systems. 13 AGING IN YEAST Early studies Background: Over 45 years ago, Mortimer and Johnston made the observation that was to become the basis for the application of genetic analysis to the study of aging. Observing the repeated divisions of individual yeast mother cells, these scientists noted that in a genetically identical population of yeast, each cell ceased to divide after giving birth to a finite number of daughter cells (Mortimer and Johnston, 1959). For a given strain of yeast, the likelihood of mortality increased exponentially with increasing numbers of cell divisions. When Mortimer and Johnston plotted their data, they made a striking observation: that the mortality curve for a population of yeast closely resembled the mortality curves for many other organisms, including humans (Figure 1). The study of yeast aging was born. In the decades that followed, radical advances characterized the study of yeast biology. On the forefront of genetic research, yeast provided an invaluable model system for the investigation of such basic biological processes as the cell division cycle, the secretory pathway, and the DNA damage response, among many others. However, it was not until the 1990's that the powerful tools of molecular biology, combined with the resources of the fully sequenced yeast genome, were applied to the study of yeast aging (Sinclair et al., 1998). Budding yeast reproduce by asymmetric cell division, with a pre-existing mother cell giving birth to a smaller daughter cell consisting predominantly of newly-synthesized materials (Guthrie and Fink, 1991). Practically, this asymmetry of cell division enables 14 the microscopic separation of mother and daughter cells upon completion of cytokinesis, forming the experimental basis of the yeast life span assay. Mean and maximum lifespan are relatively constant for a given strain, although different strains of yeast vary drastically over a mean life span range of less than 15 generations to greater than 30 (Sinclair et al, 1998). This initial finding demonstrated a strong genetic influence on longevity, and encouraged the search for single gene mutations affecting the rate of aging. Age-relatedchanges: Aging in yeast is accompanied by a battery of phenotypic changes, ranging from obvious changes in cell size (Mortimer and Johnston, 1959) to more subtle differences in transcription and protein localization (Kennedy et al., 1997; Mortimer and Johnston, 1959; Smeal et al., 1996). Mortimer and Johnston hypothesized, a priori, that the onset of senescence is caused when cell volume reaches some upper limit (Mortimer and Johnston, 1959). Three lines of evidence now refute that hypothesis. First, inducing an increase in size by temporary G1 arrest fails to shorten life span (Kennedy et al, 1994). Second, populations of old senescent cells span a wide range of sizes, defying a critical prediction of Mortimer and Johnston's hypothesis (G.L., unpublished observation). Finally, of the multitude of single gene changes known to extend life span, very few affect cell size (Bitterman et al., 2003). Age-dependent changes in cell surface characteristics also accompany the journey into yeast old age. Notably, each round of cell division leads to the deposition of a chitin bud scar on the surface of the mother (but not the daughter) cell. These bud scars can be stained with calcofluor white, permitting visualization under a microscope and enabling 15 the determination of cell age. Early hypotheses attributing bud scar accumulation as the cause of aging have now been generally refuted (Sinclair et al., 1998), mostly on the grounds that induced chitin accumulation on the cell surface does not curtail life span (Egilmez and Jazwinski, 1989). Also, theoretical models of the yeast cell surface estimate that the cell wall can accommodate more than three times the number of bud scars than usually accumulate over the course of normal aging (Sinclair et al., 1998). The first molecular phenotype associated with aging in S. cerevisiae was a loss of transcriptional silencing (Smeal et al., 1996), and it was the characterization of this phenotype that ultimately led to the discovery of SIR2 as a central regulator of yeast aging (Kaeberlein et al., 1999). Haploid yeast cells typically exist as one of two distinct mating types, known as MATa and MATa, determined by the sequence of the gene expressed at the MAT locus. Although mating type information is only expressed from the single allele present at the MAT locus, haploid yeast harbor silenced copies of both mating type alleles, present at HM loci (Guthrie and Fink, 1991). Under normal conditions, silencing of these loci is accomplished by the SIR (Silent Information Regulator) protein complex, consisting of Sir2p, Sir3p, and Sir4p (Gottschling et al., 1990; Ivy et al., 1986; Rine and Herskowitz, 1987). The SIR complex also directs silencing at telomeres (Gottschling et al., 1990), where gene expression is abolished over telomeric repeats and nearby subtelomeric regions. Sir2p, the catalytic component of this complex, is an NAD+ dependent protein deacetylase (Imai et al., 2000; Landry et al., 2000b; Smith et al., 2000), and is discussed at length later on in this chapter. Progressive sterility has been shown to affect aging yeast mother cells (Muller, 1985), and this phenotype is now known to be attributable to changes in the localization 16 of the SIR complex (Smeal et al., 1996). This molecular phenotype of aging is most likely an effect of the aging process rather than a cause, as deletion of the HM loci cures age-related sterility without extending cellular life span (Kaeberlein et al., 1999; Smeal et al., 1996). Interestingly, telomeric silencing is lost with age (Kim et al., 1996), and this loss of silencing correlates with the absence of the SIR complex. The current explanation for these findings is that the SIR complex relocalizes to the nucleolus to counteract the nucleolar instability that eventually limits mother cell division (Kennedy et al., 1997; Sinclair et al., 1998; Sinclair and Guarente, 1997). Supporting this model, the long-lived SIR4-42 mutant constitutively relocalizes the entire SIR complex to the nucleolus, even in young cells, thereby forestalling nucleolar fragmentation and prolonging life span (Kennedy et al., 1997). Extrachromosomal rDNA Circles: a Cause of Aging The first molecular cause of aging to be determined for yeast synthesized these findings and provided a rationale for the importance of the nucleolus in cellular senescence (Sinclair and Guarente, 1997). The yeast nucleolus houses the rDNA, a tandem array of 150-200 repeated copies of the genes encoding ribosomal RNA (rRNA). Each 9.1 kb repeat encodes a 35S rRNA and a 5S rRS and 5S rRNAs are transcribed in opposing directions, after which the 35S RNA is processed into mature 25S, 18S, and 5.8S rRNA molecules (reviewed in (Nomura, 2001). According to current evidence, the stalling of the replication fork within the rDNA triggers the induction of homologous recombination, as DNA repair enzymes recognize this feature and process it into a double stranded break (Kim and Wang, 1989; Park et al., 1999). Homologous recombination 17 between neighboring rDNA repeats has been shown to result in excision of circular molecules containing one or more complete copies of the rDNA unit (Park et al., 1999), each one of which directs its own replication through the presence of three ARS consensus sequences. Of these three potential origins of replication, it is estimated that an average of one fires each S phase (Miller and Kowalski, 1993). Importantly, the same extrachromosomal rDNA circles (ERCs) that multiply exponentially with each successive round of cell division predominantly fail to segregate to daughter cells, instead accumulating in mother cells at levels approaching 1000 copies per cell (Sinclair and Guarente, 1997). This number of ERCs contains a quantity of DNA roughly equivalent to all the rest of the yeast genome. Several lines of experimental analysis have supported an ERC model of aging, indicating that ERC accumulation is not just a hallmark of yeast aging but actually a cause. First, the ectopic release of an ERC in young cells accelerates the onset of senescence, demonstrating that formation of an ERC is limiting for life span (Defossez et al., 1999; Sinclair and Guarente, 1997). Second, mutation of the rDNA unidirectional replication fork block component Foblp eliminates ERC formation and extends life span (Defossez et al., 1999). Indeed, most mutations extending the yeast life span in a strainindependent manner correlate with reduced ERC levels and can be rationalized by their effects on ERC formation and accumulation (Kaeberlein et al., 2005). 18 Sir2p, a central regulator of aging Geneticanalysis: Multiple studies from the last six years have established an important role for Sir2p as a central regulator of aging (Anderson et al., 2003a; Kaeberlein et al., 1999; Lin et al., 2000; Lin et al., 2002). While the other members of the SIR complex indirectly affect aging through modulation of silencing at the HM loci and telomeres (Kaeberlein et al., 1999), Sir2 plays an independent role in establishing rDNA heterochromatin (Bryk et al., 1997) and repressing the unequal crossing-over responsible for ERC formation (Kobayashi et al., 2004) and, hence, normal aging. At the nucleolus, Sir2p acts within the RENT (regulator of nucleolar silencing and telophase exit) complex, whose other members Netlp, Cdc14p, and Nanl (Ghidelli et al., 2001; Shou et al., 2001) have not been implicated in the aging process. The initial observation that SIR2/sir2 heterozygous diploids displayed an intermediate life span between those of the two homozygotes suggested that Sir2p dosage might be limiting for life span (Kaeberlein et al., 1999). Indeed, addition of a single extra copy of SIR2 extends life span of multiple strains by up to 40%, while the deletion of SIR2 shortens life span by approximately 50% in several strain backgrounds (Kaeberlein et al., 2005; Kaeberlein et al., 1999). Consistent with the ERC model of aging, SIR2 levels correlate not only with life span but also with ERC levels and rate of marker gene loss by rDNA recombination. Interestingly, a recent study indicates that Sir2p influences not the overall rate of rDNA recombination, but rather the relative frequency of unequal sister chromatid crossing over between different rDNA repeats, hence creating the appearance of increased total recombination in rDNA marker loss and ERC accumulation assays (Kobayashi et al., 2004). 19 Providing further support for the role of Sir2p in ERC-mediated aging, genetic analysis has shown that fobl deletion and SIR2 overexpression fail to additively increase life span, and are therefore likely to affect aging through the same pathway (Kaeberlein et al., 1999). By a similar token, fobl deletion restores wild type life span to a sir2 mutant, most likely by rescuing the increased ERC formation caused by sir2 mutation. Biochemistry: Due to its ability to promote silencing, Sir2p was long suspected of possessing histone deacetylase activity. Despite the fact that Sir2p levels were known to correlate with histone deacetlylation at targeted areas of the yeast genome (Braunstein et al., 1993; Grunstein, 1997; Jenuwein and Allis, 2001) attempts to demonstrate an in vitro deacetylation activity for Sir2p met with years of failure. In 1999, however, major strides were made in the enzymology of Sir2p and its metazoan homologs. Initially, Sir2p and several homologs were found to possess an ADP-ribosyltransferase activity, catalyzing the breakdown of a nicotinamide adenine dinucleotide (NAD) substrate and the addition of an ADP-ribose moiety to target proteins (Frye, 1999; Tanny et al., 1999). However, this reaction was soon overshadowed by the discovery of a more robust NAD-dependent histone deacetylase activity present in yeast Sir2p as well as its closest mouse homolog, SIRT1 (Imai et al., 2000; Landry et al., 2000b; Smith et al., 2000). In yeast, this activity is essential for the maintenance of silencing and life span, as mutations abolishing deacetylation generally eliminate silencing in vivo and shorten life span (Imai et al., 2000). Interestingly, some mutations in conserved residues of.Sir2p affect in vitro histone deacetylase activity and in vivo silencing differently (Armstrong et al., 2002), 20 suggesting that Sir2p might influence silencing by deacetylating non-histone substrates within the cell. Despite this inconsistency, Sir2p has been shown to display a substrate preference for lysine 16 of the histone H4 (Imai et al., 2000; Landry et al., 2000b; Smith et al., 2000), a tail residue critical for silencing. Several unique features characterize the Sir2p enzymatic reaction. Unlike other histone deacetylases, the Sir2, or Class III, deacetylases are not inhibited by trichostatin A (Imai et al., 2000). Histone deacetylation by Sir2p also cannot proceed without NAD, nor can other nicotinamide adenine dinucleotides such as NADH, NADP, or NADPH substitute for NAD (Imai et al., 2000). In fact, NADH and nicotinamide both function as effective inhibitors of Sir2p catalytic activity, providing an effective mechanism for regulating Sir2p function in vivo (Anderson et al., 2003a; Bitterman et al., 2002; Lin et al., 2004). The enzymology of NADH inhibition is discussed in detail in Appendix 2, along with the in vivo consequences of this inhibition in calorie-restricted cells. The requirement of NAD for the Sir2 enzymatic reaction is also unexpected for thermodynamic reasons. The deacetylation reaction catalyzed by trichostatin A-sensitive HDACs, a simple hydrolysis reaction, is energetically favorable, begging the question of why Sir2 would couple such a reaction to the breakdown of NAD, a metabolically important compound (Bitterman et al., 2003). Curiously, NAD is not a cofactor in the Sir2-catalyzed reaction but is actually consumed by it, with hydrolysis of a glycosidic bond occurring between the nicotinamide and ADP-ribose moieties. The complete reaction catalyzed by Sir2 encompasses two hydrolysis steps that are likely coupled (Landry et al., 2000a; Min et al., 2001; Tanner et al., 2000). In the first of these, the aforementioned glycosidic bond within NAD is hydrolyzed, releasing a predicted 8.2kcal 21 of free energy per mol (Moazed, 2001). Subsequently, the C-N bond between the acetyl group and lysine is hydrolyzed in a deacetylation reaction that occurs with 1:1 stoichiometry. The final peculiarity of this reaction is that Sir2 catalyzes the transfer of the acetyl group to the ADP-ribose cleaved from NAD, resulting in the formation of Oacetyl ADP-ribose (AAR) (Jackson and Denu, 2002; Liou et al., 2005; Tanner et al., 2000). Thus, the Sir2 reaction produces two major products in addition to deacetylated lysine: nicotinamide and AAR. While Sir2p is generally believed to influence silencing through the deacetylation of histone H4, it is formally possible that any of these three products of the Sir2 reaction could be required for creation and maintenance of the silenced state. Nicotinamide, a precursor of nicotinic acid in the cell, has been shown to strongly inhibit Sir2 enzymatic activity (Bitterman et al., 2002; Jackson et al., 2003) even doing so in several in vivo contexts, discussed below. The function of O-acetyl ADP ribose, however, has been somewhat elusive. This molecule was originally proposed to trigger some sort of signal transduction cascade, as metabolic instability of the type it displays is a hallmark of many signal transduction initiatiors (Jackson and Denu, 2002). Experimentally, injection of AAR has been shown to block cell cycle progression of Xenopus oocytes (Borra et al., 2002), consistent with this model. More recently, however, AAR has been shown to play an integral role in the assembly of the SIR complex in yeast, promoting the association of multiple copies of Sir3p with Sir2p/Sir4p and inducing important structural changes in this complex as a result (Liou et al., 2005). 22 Caloric Restriction and the Regulation of Sir2p Requirement of SIR2 and NAD+ in CR The question of how Sir2 is regulated emerged at the forefront of aging research with the observation that this protein is required for life span extension induced by caloric restriction (Lin et al., 2000). For almost 70 years, it has been known that decreasing the amount of food consumed by laboratory rodents significantly lengthens their life span and promotes youthful vigor (McCay et al., 1989). Studies have since shown that variation of calorie content was the critical factor retarding the aging process (Masoro, 1984a; Masoro, 1984b; Weindruch et al., 1988), and calorie restriction (CR), as it has come to be known, is now known to extend life span of diverse species ranging from yeast (Lin et al., 2000) to possibly even primates (Ingram et al., 2004). In yeast, CR can be accomplished by two different methods: reducing the glucose content of the growth media or inducing genetic mutations that decrease cellular glucose uptake or utilization. Both methods have succeeded in extending replicative life span, likely by the same mechanism (Kaeberlein et al., 2004b; Lin et al., 2000), as both methods combined do not synergize to give an even longer life span. Of the dozens of genes ever reported to extend the yeast life span, those that mimic CR have shown most consistent effect on life span across multiple strain backgrounds (Kaeberlein et al., 2005). Yeast utilize glucose as a preferred source of carbon, tightly regulating the expression of a variety of cell surface glucose transporters of different substrate affinity depending on the concentration of glucose present in the growth media (Ozcan and Johnston, 1999; Rolland et al., 2001). Once inside the cell, glucose is phosphorylated by hexokinase to yield glucose-6-phosphate, a critical substrate 23 in glycolysis. Deletion of HXK2, one of the three genes encoding hexokinase enzymes, extends life span by a mechanism that does not synergize with caloric restriction (Kaeberlein et al., 2004b; Lin et al., 2000). This result has been interpreted to mean that hxk2A mimics low glucose, a very plausible conclusion given that other genetic mutations predicted to reduce carbon flow through glycolysis also fail to synergize with hxk2A (Lin et al., 2000). The cyclic AMP (cAMP)-dependent protein kinase plays a critical role in the cellular response to glucose, increasing in activity when glucose is prevalent and decreasing when glucose in the extracellular medium is depleted (reviewed in (Thevelein and de Winde, 1999). Several mutations reducing PKA signaling activity significantly increase life span, and at least one mutation causing constitutive activation of PKA shortens life span by 40% (Lin et al., 2000). Mutation of the adenylate cyclase (Cdc35p) and the GTP-GDP exchange factor Cdc25p both fall into the former category, and the temperature sensitive cdc25-10 allele has been used as a genetic model for CR in several studies (Lin et al., 2000; Lin et al., 2002). In order to address the question of whether CR extends life span by a regulated mechanism, Lin and coworkers investigated a possible role for Sir2p in this process. Several lines of evidence now point to a key role for Sir2p in coordinating CR with an extension of life span. First, CR fails to extend the life span of both sir2A and sir2A foblA mutants, indicating that Sir2p activity is necessary in order for CR to extend life span (Lin et al., 2000). Second, Nptlp, a component of the predominant cellular pathway of NAD biosynthesis, is also required for CR, presumably to make NAD available as a co-substrate for Sir2p under food-restricted conditions (Lin et al., 2000). Third, CR 24 increases Sir2 activity at the rDNA, increasing silencing, decreasing recombination, and reducing ERC accumulation, all three of which are tightly associated with life span extension by Sir2p (Lin et al., 2000; Lin et al., 2002). Two models of Sir2p activation Very interestingly, CR causes an increase in respiration necessary and sufficient to extend life span in a Sir2p-dependent fashion (Lin et al., 2002). In the presence of sufficient glucose concentrations, yeast cells rely on anaerobic fermentation for energy, even though this process is far less energy-efficient than respiration. Under conditions of CR, however, cells undergo a shift from fermentation to respiration, with the Hap4p transcription factor directing the expression of key genes required for this shift (Forsburg and Guarente, 1989). Overexpression of Hap4p in cells grown on glucose mimics CR by the criteria established for other genetic CR models (Lin et al., 2002). Like these other models, Hap4p overexpression requires Sir2p to extend life span. The exact mechanism by which increased respiration is translated into an increase in Sir2p activity is still controversial, although electron transport per se is definitely required upstream of Sir2p in the pathway. Yeast cells lacking cytochrome 1 (Cytlp), an essential component of the mitochondrial electron transport chain, fail to exhibit an extension of life span or an increase in Sir2p activity when subjected to conditions of CR (Lin et al., 2002). The metabolic shift to respiration induced by CR has been shown to cause an increase in the NAD/NADH ratio, although by decreasing NADH levels (Lin et al., 2004), not increasing NAD levels as originally predicted (Anderson et al., 2003b). As presented in detail in Appendix B, NADH, a key electron donor in respiration, 25 competitively inhibits Sir2p activity, such that a decrement in cellular NADH accounts for a significant increase in Sir2p activity during CR. In agreement with this model, overexpressing NADH dehydrogenase is sufficient to increase Sir2p activity and extend life span (Lin et al., 2004). The major competing model explaining Sir2p activation by CR relies on a similar principle of relief-of-inhibition, although in this case the proposed inhibitor reduced by CR is nicotinamide, not NADH (Anderson et al., 2003a; Anderson et al., 2003b). Mechanistically, this model invokes changes in a key NAD salvage pathway enzyme to account for a reduction in nicotinamide when extracellular glucose is reduced (Anderson et al., 2002). This regimen results in upregulation of the nicotinamidase Pnclp, which converts nicotinamide into nicotinic acid, a compound that does not inhibit Sir2p (Bitterman et al., 2002). Consistent with this model, CR largely fails to extend life span when PNC1 is deleted. Also consistent with this model, overexpression of PNC1 dramatically extends yeast life span in a Sir2p-dependent manner, and PNC1overexpressors do not live any longer when calorie restricted (Anderson et al., 2003a). Interestingly, Pnclp is induced by several other life-extending conditions, including high osmolarity, mild heat stress, and low amino acid concentrations (Anderson et al., 2003a), although of these, only high osmolarity has been experimentally proven to mimic calorie restriction (Kaeberlein et al., 2002). Sir2-independent effects of CR: A recent report indicates that one particularly long-lived yeast strain, BY4742, demonstrates a Sir2-independent life span extension under CR (Kaeberlein et al., 2004b). 26 The authors found that CR elicited a greatly extended life span from sir2A foblA mutants, and that CR synergized with fobl deletion or SIR2-overexpression to create the longest-lived yeast strains ever reported. This interesting finding suggests that CR may activate more than one pathway affecting life span, raising the question of whether this new pathway bears any mechanistic similarity to the one already described. Notably, a study by another group investigating the Sir2-independent effects of CR has implicated Hst2p, the closest yeast Sir2 homolog, as a key player in the process (Lamming et al., 2005). Overexpression of HST2 suppresses rDNA recombination, promotes rDNA silencing, and extends life span. Deletion of this gene largely abrogates the effect of CR in a sir2A foblA background, thus likely accounting for the Sir2p-independent effects of CR observed in the original study. Interestingly, Hstlp, another Sir2p homolog, plays a residual role in CR in the absence of Sir2p and Hst2p (Lamming et al., 2005), suggesting that Sir2 family members display a certain degree of flexibility in their roles within the cell. These results have raised the possibility that in other organisms multiple members of the Sir2 family might cooperate to promote longevity and survival in response to moderate food shortages. The Sir2p response to yeast CR is diagrammed in Figure 2. Natural polymorphisms affecting yeast aging As strain-dependent genetic differences account for much of the controversy surrounding the Sir2-independent effects of CR, it now becomes relevant to discuss the two examples of naturally polymorphic loci known to affect the aging process. These two loci, MPT5 (also known in some studies as UTH4 or PUF5) and SSDI, bear several striking similarities to one another. Both MPT5 and SSDI are large genes characterized 27 by numerous polymorphisms in common laboratory strains (Kennedy et al., 1997; Sutton et al., 1991; Uesono et al., 1997). Both encode RNA-binding proteins that act in parallel to affect cell wall integrity, high temperature growth, and life span, as well as other global processes (Kaeberlein and Guarente, 2002). Both interact genetically with a wide array of mutations in diverse genes, and both function as post-transcriptional regulators (Gerber et al., 2004; Kaeberlein et al., 2004a; Tadauchi et al., 2001). Mpt5p,a polymorphicrepressorof translation The first genetic screen for long-lived mutants in yeast identified mutations in four complementation groups that rescued the stress sensitivity and short life span of the starting strain (Kennedy et al., 1995). These complementation groups, termed UTH1-4, have formed the basis for many subsequent aging studies. In the course of analysis of the UTH mutants, cloning of UTH4 revealed it to be allelic to MPT5, and also led to the observation that the unmutagenized strain used in the screen actually contained a Cterminal truncation of MPT5, resulting in a hypomorphic Mpt5 protein product (Kennedy et al., 1997). Deletion of MPT5 shortens life span by 50%, and overexpression extends life span by up to 40% (Kennedy et al., 1995; Kennedy et al., 1997), underscoring the importance of this post-transcriptional regulator in the aging process. Cells lacking Puf5p/Mpt5p/Uth4p display increased telomeric silencing, decreased rDNA silencing, and an aberrant distribution of Sir3p in the absence of Sir4p (Gotta et al., 1997). In the absence of SSDI-V, mpt5 mutant cells also suffer from several deficiencies associated with a weakened cell wall, including sensitivity to calcofluor white, sensitivity to sodium 28 dodecyl sulfate (SDS), and sorbitol-remedial temperature sensitivity (Kaeberlein and Guarente, 2002). Furthermore, in strains lacking SSD1-V, mpt5A is synthetically lethal in combination with deletion of either SBF or CCR4 (Kaeberlein and Guarente, 2002). SBF, a transcriptional activator consisting of the cell-cycle regulated proteins Swi4p and Swi6p, plays an important role in cell cycle-regulated transcription of the G1 cyclins CLN and CLN2 (Nasmyth and Dirick, 1991) and also acts downstream of protein kinase C (Pkclp) to promote cell wall biosynthesis (Igual et al., 1996). Ccr4p, a component of the cytoplasmic deadenylase complex (Thore et al., 2003; Tucker et al., 2002) as well as a different transcriptional complex, is also required for PKC 1-dependent transcriptional regulation of multiple genes required for proper cell wall structure (Chang et al., 1999). Genetic analysis suggests that Mpt5p, Ssdlp, and Pkclp define three parallel pathways to ensure cell wall integrity (Kaeberlein and Guarente, 2002). Mpt5p contains an RNA-binding domain homologous to that of the Drosophila PUMILIO protein, a translational repressor important for positional patterning in the developing embryo (Parisi and Lin, 2000). Of the five PUMILIO family members in yeast, Mpt5p was the first to be characterized as a post-transcriptional regulator (Tadauchi et al., 2001). Mpt5p has been shown to bind to sequence elements in the 3' UTR of the HO endonuclease (Tadauchi et al., 2001), a protein catalyzing the critical DNA strand cleavage step of haploid mating type switching. The binding of Mpt5p results in repression of HO translation, and Mpt5p is believed to exert a similar effect on other mRNAs to which it binds (Gerber et al., 2004; Tadauchi et al., 2001). Indeed, a recent study systematically identified RNA targets of the yeast PUMILIO family of 29 proteins (Pufl-5p), uncovering a novel mechanism whereby specific RNA-binding proteins physically associate with functionally related groups of target RNA molecules to regulate their post-transcriptional properties (Gerber et al., 2004). This same study identified consensus binding sites for three of the five PUF proteins in yeast, finding that binding motifs typically resided within the 3'-UTRs of target genes, although a significant quantity were found within ORFs or, rarely, the 5'UTR. SSDI, anotherRNA bindingprotein Naturally occurring polymorphisms of the SSD1 gene fall into two phenotypic classes: active SSDI-V alleles and hypomorphic ssdl-d alleles, classified on the basis of their ability to suppress mutations in the Sit4p phosphatase (Sutton et al., 1991). SSDI-V, which suppresses sit4A, has also been isolated as a suppressor of diverse genetic mutations affecting the cell wall, PKA activity, TOR-signaling, and RNA metabolism (summarized in (Kaeberlein et al., 2004a)). Addition of SSD1-V dramatically extends lifespan of ssdl-d strains by a novel, as yet uncharacterized mechanism. This mechanism is Sir2p-independent, affecting neither rDNA silencing nor ERC formation, two processes linked to yeast aging (Kaeberlein et al., 2004a). Significantly, addition of SSD1-V is the only intervention ever shown to extend the life span of a sir2AfoblA double mutant. Although SSD1 is highly conserved from yeast to humans, little is known of its biochemical function. SSD1 is discussed in detail in Chapter 2 as well as Appendix A. 30 Homologsof MPT5 and SSD1 in multicellularorganisms Both Mpt5p and Ssdlp are very highly conserved proteins, although the metazoan homologs of Mpt5p are far better characterized than those of Ssdlp. Interestingly, there are several conceptual similarities linking Mpt5p to its Drosophila and C. elegans homologs PUMILIO and FBF. These three homologs are all translational repressors that bind conserved sequence elements found in the 3'-UTR. And all three proteins possess the eight consecutively repeated PUMILIO homology domains conferring RNA binding activity and defining the PUF family. Early in embryogenesis, the Drosophila PUMILIO protein binds maternal hunchback mRNA to repress hunchback translation at the posterior pole by a mechanism also requiring the NANOS protein (Parisi and Lin, 2000). In addition to this developmental role, PUMILIO is required for the maintenance of germ line stem cells in Drosophila (Forbes and Lehmann, 1998; Lin and Spradling, 1997). C. elegans FBF, which regulates germ cell fate through a repressive interaction with GLD-1 mRNA, is also required for germ line stem cell maintenance in this organism (Crittenden et al., 2002). Thus, both PUMILIO and FBF promote continued cell divisions and delay the onset of an alternate state of differentiation and developmental potency. Intriguingly, Mpt5p also conforms to this common theme, using specific targeted translational repression to prolong cell divisions and maintain potential to produce young cells. The metazoan homologs of Ssdlp still await characterization and may provide likely candidates for regulators of development and cell fate analogous to those of the PUF family. 31 Chronological Aging in Yeast An alternative means of defining and measuring yeast life span is to assess the survival of cells maintained in a non-dividing state. This parameter is known as chronological aging, as it correlates survival with time as measured in days rather than cell division cycles. The most common assay for chronological aging is a post-diauxic survival test in which cells grown to the post-diauxic phase are kept in culture, and viability is measured at successive time points by the plating of cells and measurement of colony formation. Aging under these conditions appears to be largely attributable to damage from reactive oxygen species (ROS), as mutations in catalase and superoxide dismutase accelerate the loss of viability (Longo et al., 1996; Longo et al., 1999) and mutations increasing paraquat resistance prolong survival under post-diauxic conditions (Fabrizio et al., 2003; Fabrizio et al., 2001). Only two genes have so far been reported to increase both post-diauxic aging and replicative aging, and these genes are SCH9 and CYR] (Fabrizio et al., 2004; Fabrizio et al., 2001). SCH9 encodes a serine/threonine kinase that acts in parallel to components of the PKA pathway to affect stress resistance and response to glucose. CYR] encodes adenylate cyclase, which produces cAMP in response to glucose uptake by a mechanism described above. This process is essential for PKA activity, a reduction of which is known to increase stress resistance by downstream activation of the Msn2p and Msn4p transcription factors (Rolland et al., 2001). By this token, the dual effects of cyrl mutation on replicative and post-diauxic aging can be rationalized by downstream activation of stress response genes (predicted to prolong survival in non-dividing cells) and respiration-dependent activation of Sir2p 32 (predicted to prolong replicative life span). It is not clear whether sch9A requires Sir2p in order to extend replicative life span, but it has been shown that sch9 mutation extends chronological life span by a mechanism requiring Sod2p (Fabrizio et al., 2003). Interestingly, the effects of sch9A on replicative life span are very well conserved between several different yeast strains (Kaeberlein et al., 2005), and this gene, like FOB], displays a defect in HOTl-dependent recombination when mutated (Defossez et al., 1999). It is therefore possible that sch9A might promote longevity by correcting a defect similar to that corrected by deletion of FOB]. Future studies should aim to integrate SCH9 into existing genetic pathways of aging. CONSERVATION OF YEAST AGING PATHWAYS IN METAZOANS A critical question facing the model systems approach to aging is to what extent findings from such model systems are conserved across phyla. Intriguingly, the last several years have provided numerous reports confirming that aging and longevity genes can, in fact, affect similar aspects of the aging process in very distantly related species. Insulin/IGF-1 pathway Sch9, like Sir2, has been shown to affect aging in at least two species whose last common ancestor existed a whopping one billion years ago (Guarente and Picard, 2005). The kinase domain of Sch9p is highly conserved, showing 47-49% identity to its closest worm homologs AKT-1 and AKT-2. AKT-1 and AKT-2 act downstream of the DAF-2 insulin receptor homolog, regulating longevity, stress resistance and the diapause state in 33 response to a conserved PI-3 kinase signaling cascade (Guarente and Kenyon, 2000; Paradis and Ruvkun, 1998). Depending on the severity of the mutation, defects in genes of the DAF-2 pathway cause either dauer formation or extended adult life span, with more mild mutations generally favoring the latter result (Johnson, 1990; Kenyon et al., 1993). The longevity of daf-2 mutants is known to require several genes, including DAF16 (a forkhead/winged helix transcription factor), heat shock protein HSF-1, and the AMP-dependent kinase AAK-2 (Garigan et al., 2002; Henderson and Johnson, 2001; Hsu et al., 2003; Lee et al., 2001; Lin et al., 2001; Morley and Morimoto, 2004). The insulin/IGF-1 pathway, like the PKA and Sch9 pathways in yeast, normally downregulates nutrient storage and stress responses (Kenyon, 2005). It is therefore especially interesting that mutating such a pathway would extend life span in both species, as it implies not only conservation of specific aging genes but conservation of an overall approach to regulating the aging process (Figure 3). Other mutations in the insulin/IGF-1 pathway have been shown to extend life span in worms, flies, and even mice. In Drosophila, mutations of the insulin/IGF-l receptor increase life span by variable amounts of up to 80% (Tatar et al., 2001). Similarly, mutations in the downstream chico insulin receptor substrate (IRS-1) extend life span by up to 40% (Clancy et al., 2001; Tu et al., 2002a; Tu et al., 2002b). Although no experiments have yet linked these phenotypes to the DAF-16 homolog FOXO, it is strongly predicted that extension of life span in IRS and insulin/IGF-1 receptor mutants requires FOXO activation. FOXO overexpression has been shown to extend fly life span (Giannakou et al., 2004), and the FOXO transcription factor is known to lie downstream 34 of insulin/IGF-1 signaling for at least one other phenotype, namely reduced cell division (Junger et al., 2003). In mammals, the analogous genetic pathways are far more complex, but similar results have nonetheless been reported. In mice, which have distinct receptors for insulin and IGF-1, female mutants heterozygous for the IGF-1 receptor live about 30% longer than wild type controls (Holzenberger et al., 2003). Also, mice lacking the insulin receptor in adipose tissue display an 18% life span increase compared to controls (Bluher et al., 2003). Taken together, these results suggest that through evolution both the insulin and IGF- 1 pathways have retained an influence on life span regulation. Similar to the case of Drosophila, a downstream connection to FOXO is also strongly suspected but has yet to be demonstrated for long-lived mouse insulin and IGF-1 mutants. FOXO transcription factors act downstream of insulin and IGF-1 to effect changes in metabolism (Burgering and Kops, 2002; Kops et al., 2002), so it is conceivable that they play a similar role in the pathways affecting aging. A recent study has shown that p66shc mutation, which increases stress resistance as well as organismal longevity (Migliaccio et al., 1999), requires FOXO proteins, at least in order to confer cellular resistance to stress (Nemoto and Finkel, 2002). SIR2 and Calorie Restriction In C. elegans, Sir2 ortholog Sir2.1 extends life span, albeit by a mechanism differing from the one at work in yeast cells. Worms likely do not age as a result of ERC accumulation, as no such accumulation has ever been observed in this predominantly post-mitotic organism (Guarente lab, unpublished data). Life span extension by Sir2.1 35 instead requires the forkhead transcription factor DAF-16, as DAF-16 mutants live the same length regardless of Sir2.1 levels (Tissenbaum and Guarente, 2001). DAF-16, a key downstream element in the insulin-like signaling pathway, is therefore likely to act downstream of Sir2.1 as well. Unpublished results from our lab indicate that this is indeed the case (Ala Berdichevsky, personal communication). These findings in worms are especially interesting in light of recent studies uncovering interactions between mammalian SIRT1 and FOXO proteins homologous to the C. elegans DAF-16 (Brunet et al., 2004; Motta et al., 2004; van der Horst et al., 2004). These results are discussed in detail later on in this chapter. Studies from Drosophila melanogaster support a conserved role for Sir2 in the response to CR, a surprising finding considering the vast evolutionary distance between yeast and fruit flies. In flies, food restriction is accomplished by limiting the yeast content of the flies' diet, and this intervention succeeds in extending life span significantly (Clancy et al., 2002). Several lines of evidence support the claim that Sir2 upregulation plays an indispensable role in the fly response to CR. Caloric restriction in flies increases Sir2 mRNA levels (Rogina et al., 2002), and the overexpression of Sir2 using a UAS/Gal4 driver is sufficient to extend life span (Rogina and Helfand, 2004). This extension of life span does not synergizes with CR, suggesting that the two treatments constitute only one pathway (Rogina and Helfand, 2004). Further support for this model comes from experiments with the Sir2-activator resveratrol, which, like CR, extends life span in wild type but not Sir2 mutant flies (Rogina and Helfand, 2004; Wood et al., 2004). Like overexpression of Sir2, resveratrol treatment fails to further extend the long life span of CR flies. 36 These findings hint at a conserved role for Sir2 proteins in life span regulation, especially in response to CR. A significant effort now exists to establish a role for Sir2 in the mammalian response to calorie restriction, and to determine to what extent, if any, findings from lower organisms apply to the more sophisticated mammalian systems. To that end, numerous studies have already defined a role for mammalian Sir2 in the promotion of cell survival and division in the face of potentially apopotic stresses (reviewed in (Blander and Guarente, 2004; Guarente and Picard, 2005). Mammalian genomes contain seven Sir2 homologs , termed sirtuins (SIRTs). Of these, SIRT1, orthologous to the yeast Sir2, is the best characterized and has the broadest substrate specificity (Blander and Guarente, 2003). A nuclear protein, SIRT1 deacetylates several non-histone substrates in vivo, including MyoD, p53, and FOXO transcription factors, thereby affecting cell differentiation and survival under stress (Brunet et al., 2004; Motta et al., 2004; North and Verdin, 2004). In mammals, FOXO transcription factors mediate the metabolic output of IGF-1 and insulin pathways and also play a key role in the apoptotic response to stress. SIRT1 has been shown by several groups to deacetylate FOXO 1, FOXO3, and FOXO4 (Brunet et al., 2004; Daitoku et al., 2004; Motta et al., 2004; van der Horst et al., 2004), thereby repressing FOXO-mediated apoptosis. Strangely, there are also reports of FOXO target genes activated by SIRT1 (Brunet et al., 2004; Daitoku et al., 2004; van der Horst et al., 2004), although no mechanism has yet been established to explain these findings. These FOXO studies are not the only reports linking SIRT1 to regulation of the insulin pathway in mice. Very recent work has also defined a role for SIRT in pancreatic 13-cells,where it enhances insulin secretion in response to glucose, improving 37 glucose tolerance (Moynihan et al., 2005). It is not yet known whether these conditional mutants overexpressing SIRT1 in B-cells show an extension in life span. Recently, mouse SIRT1 was reported to promote the mobilization of fatty acids in white adipose tissue by repressing PPARy, linking Sir2 proteins to the physiology of caloric restriction in mammals (Picard et al., 2004). Several lines of evidence suggest that fat storage in the white adipose tissue (WAT) may play an important role in mammalian aging. Both the FIRKO mice (Bluher et al., 2003) and C/EBPB knock-in mice (Chiu et al., 2004), both genetically engineered to be lean, live longer than controls. These observations can be rationalized in the context of CR where WAT, a known endocrine tissue secreting such important hormones as leptin and adiponectin, shrinks in size as cells lose fat in response to lowered food intake. WAT therefore provides a likely candidate for a diet sensor that could then secrete hormones adjusting the rate of aging in the organism as a whole. Although appealing, this model is not without its inconsistencies. Notably, when leptin-deficient ob/ob mice are subjected to CR, these mice live as long as wild type CR animals, even though the ob/ob mice are fatter than not only wild type CR animals, but wild type animals fed an ad libitum diet (Harrison et al., 1984). Also, one report comparing animals within each CR or ad libitum group describes a slight positive correlation between body fat and longevity within each group (Bertrand et al., 1980). Both of these findings fail to support a key prediction of the WAT model of aging, which is that reduction of WAT should correlate with extended life span. Nevertheless, SIRT1 has been shown to negatively regulate the fat cell differentiation of 3T3-L1 pre-adipocytes (Picard et al., 2004). 38 This effect is explained by the observation that SIRT1 binds to the PPARy negative cofactors NCoR and SMRT and can be found at the promoters of fat-specific genes containing PPARy binding sites (Picard et al., 2004). In animals, SIRT1 also plays a role in the mobilization of fat following fasting. In SIRT1 heterozygous knockouts, lipolysis of triglycerides and release of free fatty acids into the blood was reduced (Picard et al., 2004), although some mechanistic details of this process are still uncharacterized (Bordone and Guarente, 2005), such as how a decrease in food uptake activates SIRT1 in animals. In cultured cells, acute nutrient withdrawal activates FOXO3- and p53-dependent transcription of SIRT1, whose levels consequently increase (Nemoto et al., 2004). Interestingly, p53 and FOXO3 physically interact in response to nutrient stress, and the induction of SIRT1 depends on two p53 binding sites located in the SIRT1 upstream sequence. In rats, caloric restriction increases levels of SIRT1 (Cohen et al., 2004b), inhibiting stress-induced apoptosis by a mechanism involving Ku70 and Bax. Specifically, deacetylation of Ku70 by SIRT1 allows this DNA repair protein to bind to and inactivate Bax, a proapoptotic factor (Cohen et al., 2004a; Cohen et al., 2004b). Other sirtuins Of the remaining SIRTs (SIRT2-7), an in vivo substrate has only been identified for the cytoplasmic SIRT2 (North et al., 2003). This substrate, -tubulin, is specifically deacetylated by SIRT2 (North et al., 2003), although the biological consequences of this reaction are unclear. While the archetypal yeast Sir2p is generally believed to be a histone deacetylase, most mammalian sirtuins are non-nuclear (Michishita et al., 2005), strongly suggesting that they act on non-histone substrates. 39 Using sequence similarity, eukaryotic Sir2 genes have been divided into four broad phylogenetic classes (Frye, 2000), known as classes I-IV (Figure 1). SIRT1, SIRT2, and SIRT3 are class I, SIRT4 is class II, and SIRT5 falls within the predominantly prokaryotic class III. Finally, mammalian SIRT6 and SIRT7, are class IV sirtuins (Frye, 2000). In vitro studies indicate that human SIRT1, 2, 3, and 5 possess NAD-dependent histone deacetylase activity (North et al., 2003), whereas SIRT4, 6, and 7 fail to deacetylate 3 H-labeled acetylated histone H4 peptide (North et al., 2003). The lack of detectable deacetylase activity in these SIRTs may result from their specificity for targets other than those tested, or it may indicate an enzymatic activity other than deacetylation inherent in these sirtuins. To further our understanding of the Sir2 family of proteins in mammals, we have undertaken research on the molecular characteristics and functions of the two class IV sirtuins SIRT6 and SIRT7. As we present in Chapter 3 and Appendix C, SIRT6 and SIRT7 are both nuclear proteins, with SIRT7 found exclusively in the nucleolar subcompartment. While SIRT6 displays an auto-ADP-ribosyltransferase activity uncharacteristic of Sir2 family members, SIRT7 apparently lacks this activity. SIRT7, however, plays a unique role as an activator of RNA polymerase I transcription, physically interacting with the Pol I complex and promoting transcription through the greater association of this complex with the rDNA. 40 Summary and Conclusions Many organisms appear to possess a genetically conserved mechanism by which to forestall aging in response to environmental cues. In yeast, this mechanism hinges on Sir2p, whose activity at the rDNA is essential for longevity. Interestingly, the characterization of Sir2 in yeast has led to several insights about aging in other organisms, and it now seems likely that Sir2 proteins play a conserved role in the response to calorie restriction in flies and even mammals. And Sir2 is not the only longevity-associated gene implicated in a conserved pathway regulating aging. Notably, the insulin/IGF-l hormonal signaling pathway can be mutated to forestall aging in worms, flies and mice, suggesting that current studies have only scratched the surface of a vast network of aging-regulatory genes that also coordinate energy homeostasis, stress resistance, and reproduction. The mission of the field is now to continue the molecular characterization of longevity genes in model systems while simultaneously applying our current knowledge to mammalian systems. With this goal in mind, we present the research of this thesis. Chapter 2 and Appendix A aim to deepen the understanding of yeast aging, providing a model for the action of SSD1-V, one of only two natural polymorphisms known to affect this process. Appendix B visits the issue of Sir2p regulation by calorie restriction, providing evidence that an increase in the cellular NAD/NADH ratio, occurring via a decrease in NADH, results in activation of Sir2p and extension of life span. Finally, Chapter 3 and Appendix C present the first characterizations of class IV sirtuins, SIRT6 and SIRT7. In light of the recent finding that multiple sirtuins can 41 collaborate to extend life span in yeast, it is especially interesting to know to what extent the role of Sir2 has been conserved and distributed among its various homologs. 42 REFERENCES Anderson, R. M., Bitterman, K. J., Wood, J. G., Medvedik, O., Cohen, H., Lin, S. S., Manchester, J. K., Gordon, J. I., and Sinclair, D. A. (2002). Manipulation of a nuclear NAD+ salvage pathway delays aging without altering steady-state NAD+ levels. J Biol Chem 277, 18881-18890. Anderson, R. M., Bitterman, K. J., Wood, J. G., Medvedik, O., and Sinclair, D. A. (2003a). Nicotinamide and PNC 1 govern lifespan extension by calorie restriction in Saccharomyces cerevisiae. Nature 423, 181-185. Anderson, R. M., Latorre-Esteves, M., Neves, A. R., Lavu, S., Medvedik, O., Taylor, C., Howitz, K. T., Santos, H., and Sinclair, D. A. (2003b). Yeast life-span extension by calorie restriction is independent of NAD fluctuation. Science 302, 2124-2126. Armstrong, C. M., Kaeberlein, M., Imai, S. I., and Guarente, L. (2002). Mutations in Saccharomyces cerevisiae gene SIR2 can have differential effects on in vivo silencing phenotypes and in vitro histone deacetylation activity. Mol Biol Cell 13, 1427-1438. Bertrand, H. A., Lynd, F. T., Masoro, E. J., and Yu, B. P. (1980). Changes in adipose mass and cellularity through the adult life of rats fed ad libitum or a life-prolonging restricted diet. J Gerontol 35, 827-835. Bitterman, K. J., Anderson, R. M., Cohen, H. Y., Latorre-Esteves, M., and Sinclair, D. A. (2002). Inhibition of silencing and accelerated aging by nicotinamide, a putative negative regulator of yeast sir2 and human SIRT1. J Biol Chem 277, 45099-45107. Bitterman, K. J., Medvedik, O., and Sinclair, D. A. (2003). Longevity regulation in Saccharomyces cerevisiae: linking metabolism, genome stability, and heterochromatin. Microbiol Mol Biol Rev 67, 376-399, table of contents. Blander, G., and Guarente, L. (2004). The Sir2 family of protein deacetylases. Annu Rev Biochem 73, 417-435. Bluher, M., Kahn, B. B., and Kahn, C. R. (2003). Extended longevity in mice lacking the insulin receptor in adipose tissue. Science 299, 572-574. Bordone, L., and Guarente, L. (2005). Calorie restriction, SIRT1 and metabolism: understanding longevity. Nat Rev Mol Cell Biol 6, 298-305. Borra, M. T., O'Neill, F. J., Jackson, M. D., Marshall, B., Verdin, E., Foltz, K. R., and Denu, J. M. (2002). Conserved enzymatic production and biological effect of O-acetyl-ADP-ribose by silent information regulator 2-like NAD+-dependent deacetylases. J Biol Chem 277, 1263212641. Braunstein, M., Rose, A. B., Holmes, S. G., Allis, C. D., and Broach, J. R. (1993). Transcriptional silencing in yeast is associated with reduced nucleosome acetylation. Genes Dev 7, 592-604. 43 Brunet, A., Sweeney, L. B., Sturgill, J. F., Chua, K. F., Greer, P. L., Lin, Y., Tran, H., Ross, S. E., Mostoslavsky, R., Cohen, H. Y., et al. (2004). Stress-dependent regulation of FOXO transcription factors by the SIRT deacetylase. Science 303, 2011-2015. Bryk, M., Banerjee, M., Murphy, M., Knudsen, K. E., Garfinkel, D. J., and Curcio, M. J. (1997). Transcriptional silencing of Tyl elements in the RDN1 locus of yeast. Genes Dev 11, 255-269. Burgering, B. M., and Kops, G. J. (2002). Cell cycle and death control: long live Forkheads. Trends Biochem Sci 27, 352-360. Chang, M., French-Cornay, D., Fan, H. Y., Klein, H., Denis, C. L., and Jaehning, J. A. (1999). A complex containing RNA polymerase II, Paflp, Cdc73p, Hprlp, and Ccr4p plays a role in protein kinase C signaling. Mol Cell Biol 19, 1056-1067. Chiu, C. H., Lin, W. D., Huang, S. Y., and Lee, Y. H. (2004). Effect of a C/EBP gene replacement on mitochondrial biogenesis in fat cells. Genes Dev 18, 1970-1975. Clancy, D. J., Gems, D., Hafen, E., Leevers, S. J., and Partridge, L. (2002). Dietary restriction in long-lived dwarf flies. Science 296, 319. Clancy, D. J., Gems, D., Harshman, L. G., Oldham, S., Stocker, H., Hafen, E., Leevers, S. J., and Partridge, L. (2001). Extension of life-span by loss of CHICO, a Drosophila insulin receptor substrate protein. Science 292, 104-106. Cohen, H. Y., Lavu, S., Bitterman, K. J., Hekking, B., Imahiyerobo, T. A., Miller, C., Frye, R., Ploegh, H., Kessler, B. M., and Sinclair, D. A. (2004a). Acetylation of the C terminus of Ku70 by CBP and PCAF controls Bax-mediated apoptosis. Mol Cell 13, 627-638. Cohen, H. Y., Miller, C., Bitterman, K. J., Wall, N. R., Hekking, B., Kessler, B., Howitz, K. T., Gorospe, M., de Cabo, R., and Sinclair, D. A. (2004b). Calorie restriction promotes mammalian cell survival by inducing the SIRT1 deacetylase. Science 305, 390-392. Crittenden, S. L., Bernstein, D. S., Bachorik, J. L., Thompson, B. E., Gallegos, M., Petcherski, A. G., Moulder, G., Barstead, R., Wickens, M., and Kimble, J. (2002). A conserved RNA-binding protein controls germline stem cells in Caenorhabditis elegans. Nature 417, 660-663. Daitoku, H., Hatta, M., Matsuzaki, H., Aratani, S., Ohshima, T., Miyagishi, M., Nakajima, T., and Fukamizu, A. (2004). Silent information regulator 2 potentiates Foxo 1-mediated transcription through its deacetylase activity. Proc Natl Acad Sci U S A 101, 10042-10047. Defossez, P. A., Prusty, R., Kaeberlein, M., Lin, S. J., Ferrigno, P., Silver, P. A., Keil, R. L., and Guarente, L. (1999). Elimination of replication block protein Fob extends the life span of yeast mother cells. Mol Cell 3, 447-455. Egilmez, N. K., and Jazwinski, S. M. (1989). Evidence for the involvement of a cytoplasmic factor in the aging of the yeast Saccharomyces cerevisiae. J Bacteriol 171, 37-42. 44 Fabrizio, P., Liou, L. L., Moy, V. N., Diaspro, A., SelverstoneValentine, J., Gralla, E. B., and Longo, V. D. (2003). SOD2 functions downstream of Sch9 to extend longevity in yeast. Genetics 163,35-46. Fabrizio, P., Pletcher, S. D., Minois, N., Vaupel, J. W., and Longo, V. D. (2004). Chronological aging-independent replicative life span regulation by Msn2/Msn4 and Sod2 in Saccharomyces cerevisiae. FEBS Lett 557, 136-142. Fabrizio, P., Pozza, F., Pletcher, S. D., Gendron, C. M., and Longo, V. D. (2001). Regulation of longevity and stress resistance by Sch9 in yeast. Science 292, 288-290. Forbes, A., and Lehmann, R. (1998). Nanos and Pumilio have critical roles in the development and function of Drosophila germline stem cells. Development 125, 679-690. Forsburg, S. L., and Guarente, L. (1989). Identification and characterization of HAP4: a third component of the CCAAT-bound HAP2/HAP3 heteromer. Genes Dev 3, 1166-1178. Frye, R. A. (1999). Characterization of five human cDNAs with homology to the yeast SIR2 gene: Sir2-like proteins (sirtuins) metabolize NAD and may have protein ADP-ribosyltransferase activity. Biochem Biophys Res Commun 260, 273-279. Frye, R. A. (2000). Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochem Biophys Res Commun 273, 793-798. Garigan, D., Hsu, A. L., Fraser, A. G., Kamath, R. S., Ahringer, J., and Kenyon, C. (2002). Genetic analysis of tissue aging in Caenorhabditis elegans: a role for heat-shock factor and bacterial proliferation. Genetics 161, 1101-1112. Gerber, A. P., Herschlag, D., and Brown, P. 0. (2004). Extensive association of functionally and cytotopically related mRNAs with Puf family RNA-binding proteins in yeast. PLoS Biol 2, E79. Ghidelli, S., Donze, D., Dhillon, N., and Kamakaka, R. T. (2001). Sir2p exists in two nucleosome-binding complexes with distinct deacetylase activities. Embo J 20, 4522-4535. Giannakou, M. E., Goss, M., Junger, M. A., Hafen, E., Leevers, S. J., and Partridge, L. (2004). Long-lived Drosophila with overexpressed dFOXO in adult fat body. Science 305, 361. Gotta, M., Strahl-Bolsinger, S., Renauld, H., Laroche, T., Kennedy, B. K., Grunstein, M., and Gasser, S. M. (1997). Localization of Sir2p: the nucleolus as a compartment for silent information regulators. Embo J 16, 3243-3255. Gottschling, D. E., Aparicio, O. M., Billington, B. L., and Zakian, V. A. (1990). Position effect at S. cerevisiae telomeres: reversible repression of Pol II transcription. Cell 63, 751-762. 45 Grunstein, M. (1997). Molecular model for telomeric heterochromatin in yeast. Curr Opin Cell Biol 9, 383-387. Guarente, L., and Kenyon, C. (2000). Genetic pathways that regulate ageing in model organisms. Nature 408, 255-262. Guarente, L., and Picard, F. (2005). Calorie restriction--the SIR2 connection. Cell 120, 473-482. Guthrie, C., and Fink, G. R. (1991). Guide to yeast genetics and molecular biology (San Diego: Academic Press). Harrison, D. E., Archer, J. R., and Astle, C. M. (1984). Effects of food restriction on aging: separation of food intake and adiposity. Proc Natl Acad Sci U S A 81, 1835-1838. Henderson, S. T., and Johnson, T. E. (2001). daf-16 integrates developmental and environmental inputs to mediate aging in the nematode Caenorhabditis elegans. Curr Biol 11, 1975-1980. Holzenberger, M., Dupont, J., Ducos, B., Leneuve, P., Geloen, A., Even, P. C., Cervera, P., and Le Bouc, Y. (2003). IGF- 1 receptor regulates lifespan and resistance to oxidative stress in mice. Nature 421, 182-187. Hsu, A. L., Murphy, C. T., and Kenyon, C. (2003). Regulation of aging and age-related disease by DAF-16 and heat-shock factor. Science 300, 1142-1145. Igual, J. C., Johnson, A. L., and Johnston, L. H. (1996). Coordinated regulation of gene expression by the cell cycle transcription factor Swi4 and the protein kinase C MAP kinase pathway for yeast cell integrity. Embo J 15, 5001-5013. Imai, S., Armstrong, C. M., Kaeberlein, M., and Guarente, L. (2000). Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature 403, 795-800. Ingram, D. K., Anson, R. M., de Cabo, R., Mamczarz, J., Zhu, M., Mattison, J., Lane, M. A., and Roth, G. S. (2004). Development of calorie restriction mimetics as a prolongevity strategy. Ann N Y Acad Sci 1019,412-423. Ivy, J. M., Klar, A. J., and Hicks, J. B. (1986). Cloning and characterization of four SIR genes of Saccharomyces cerevisiae. Mol Cell Biol 6, 688-702. Jackson, M. D., and Denu, J. M. (2002). Structural identification of 2'- and 3'-O-acetyl-ADPribose as novel metabolites derived from the Sir2 family of beta -NAD+-dependent histone/protein deacetylases. J Biol Chem 277, 18535-18544. Jackson, M. D., Schmidt, M. T., Oppenheimer, N. J., and Denu, J. M. (2003). Mechanism of nicotinamide inhibition and transglycosidation by Sir2 histone/protein deacetylases. J Biol Chem 278, 50985-50998. 46 Jenuwein, T., and Allis, C. D. (2001). Translating the histone code. Science 293, 1074-1080. Johnson, T. E. (1990). Increased life-span of age-I mutants in Caenorhabditis elegans and lower Gompertz rate of aging. Science 249, 908-912. Junger, M. A., Rintelen, F., Stocker, H., Wasserman, J. D., Vegh, M., Radimerski, T., Greenberg, M. E., and Hafen, E. (2003). The Drosophila forkhead transcription factor FOXO mediates the reduction in cell number associated with reduced insulin signaling. J Biol 2, 20. Kaeberlein, M., Andalis, A. A., Fink, G. R., and Guarente, L. (2002). High osmolarity extends life span in Saccharomyces cerevisiae by a mechanism related to calorie restriction. Mol Cell Biol 22, 8056-8066. Kaeberlein, M., Andalis, A. A., Liszt, G. B., Fink, G. R., and Guarente, L. (2004a). Saccharomyces cerevisiae SSD1-V confers longevity by a Sir2p-independent mechanism. Genetics 166, 1661-1672. Kaeberlein, M., and Guarente, L. (2002). Saccharomyces cerevisiae MPT5 and SSD1 function in parallel pathways to promote cell wall integrity. Genetics 160, 83-95. Kaeberlein, M., Kirkland, K. T., Fields, S., and Kennedy, B. K. (2004b). Sir2-independent life span extension by calorie restriction in yeast. PLoS Biol 2, E296. Kaeberlein, M., Kirkland, K. T., Fields, S., and Kennedy, B. K. (2005). Genes determining yeast replicative life span in a long-lived genetic background. Mech Ageing Dev 126, 491-504. Kaeberlein, M., McVey, M., and Guarente, L. (1999). The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev 13, 2570-2580. Kennedy, B. K., Austriaco, N. R., Jr., and Guarente, L. (1994). Daughter cells of Saccharomyces cerevisiae from old mothers display a reduced life span. J Cell Biol 127, 1985-1993. Kennedy, B. K., Austriaco, N. R., Jr., Zhang, J., and Guarente, L. (1995). Mutation in the silencing gene SIR4 can delay aging in S. cerevisiae. Cell 80, 485-496. Kennedy, B. K., Gotta, M., Sinclair, D. A., Mills, K., McNabb, D. S., Murthy, M., Pak, S. M., Laroche, T., Gasser, S. M., and Guarente, L. (1997). Redistribution of silencing proteins from telomeres to the nucleolus is associated with extension of life span in S. cerevisiae. Cell 89, 381391. Kenyon, C. (2005). The plasticity of aging: insights from long-lived mutants. Cell 120, 449-460. Kenyon, C., Chang, J., Gensch, E., Rudner, A., and Tabtiang, R. (1993). A C. elegans mutant that lives twice as long as wild type. Nature 366, 461-464. 47 Kim, R. A., and Wang, J. C. (1989). A subthreshold level of DNA topoisomerases leads to the excision of yeast rDNA as extrachromosomal rings. Cell 57, 975-985. Kim, S., Villeponteau, B., and Jazwinski, S. M. (1996). Effect of replicative age on transcriptional silencing near telomeres in Saccharomyces cerevisiae. Biochem Biophys Res Commun 219, 370-376. Kirkwood, T. B. (2005). Understanding the odd science of aging. Cell 120, 437-447. Kobayashi, T., Horiuchi, T., Tongaonkar, P., Vu, L., and Nomura, M. (2004). SIR2 regulates recombination between different rDNA repeats, but not recombination within individual rRNA genes in yeast. Cell 117, 441-453. Kops, G. J., Dansen, T. B., Polderman, P. E., Saarloos, I., Wirtz, K. W., Coffer, P. J., Huang, T. T., Bos, J. L., Medema, R. H., and Burgering, B. M. (2002). Forkhead transcription factor FOXO3a protects quiescent cells from oxidative stress. Nature 419, 316-321. Lamming, D. W., Latorre-Esteves, M., Medvedik, O., Wong, S. N., Tsang, F. A., Wang, C., Lin, S. J., and Sinclair, D. A. (2005). HST2 mediates SIR2-independent life-span extension by calorie restriction. Science 309, 1861-1864. Landry, J., Slama, J. T., and Sternglanz, R. (2000a). Role of NAD(+) in the deacetylase activity of the SIR2-like proteins. Biochem Biophys Res Commun 278, 685-690. Landry, J., Sutton, A., Tafrov, S. T., Heller, R. C., Stebbins, J., Pillus, L., and Sternglanz, R. (2000b). The silencing protein SIR2 and its homologs are NAD-dependent protein deacetylases. Proc Natl Acad Sci U S A 97, 5807-5811. Lee, R. Y., Hench, J., and Ruvkun, G. (2001). Regulation of C. elegans DAF-16 and its human ortholog FKHRL1 by the daf-2 insulin-like signaling pathway. Curr Biol 11, 1950-1957. Lin, H., and Spradling, A. C. (1997). A novel group of pumilio mutations affects the asymmetric division of germline stem cells in the Drosophila ovary. Development 124, 2463-2476. Lin, K., Hsin, H., Libina, N., and Kenyon, C. (2001). Regulation of the Caenorhabditis elegans longevity protein DAF-16 by insulin/IGF-1 and germline signaling. Nat Genet 28, 139-145. Lin, S. J., Defossez, P. A., and Guarente, L. (2000). Requirement of NAD and SIR2 for life-span extension by calorie restriction in Saccharomyces cerevisiae. Science 289, 2126-2128. Lin, S. J., Ford, E., Haigis, M., Liszt, G., and Guarente, L. (2004). Calorie restriction extends yeast life span by lowering the level of NADH. Genes Dev 18, 12-16. 48 Lin, S. J., Kaeberlein, M., Andalis, A. A., Sturtz, L. A., Defossez, P. A., Culotta, V. C., Fink, G. R., and Guarente, L. (2002). Calorie restriction extends Saccharomyces cerevisiae lifespan by increasing respiration. Nature 418, 344-348. Liou, G. G., Tanny, J. C., Kruger, R. G., Walz, T., and Moazed, D. (2005). Assembly of the SIR complex and its regulation by O-acetyl-ADP-ribose, a product of NAD-dependent histone deacetylation. Cell 121, 515-527. Longo, V. D., Gralla, E. B., and Valentine, J. S. (1996). Superoxide dismutase activity is essential for stationary phase survival in Saccharomyces cerevisiae. Mitochondrial production of toxic oxygen species in vivo. J Biol Chem 271, 12275-12280. Longo, V. D., Liou, L. L., Valentine, J. S., and Gralla, E. B. (1999). Mitochondrial superoxide decreases yeast survival in stationary phase. Arch Biochem Biophys 365, 131-142. Masoro, E. J. (1984a). Food restriction and the aging process. J Am Geriatr Soc 32, 296-300. Masoro, E. J. (1984b). Nutrition as a modulator of the aging process. Physiologist 27, 98-101. McCay, C. M., Crowell, M. F., and Maynard, L. A. (1989). The effect of retarded growth upon the length of life span and upon the ultimate body size. 1935. Nutrition 5, 155-171; discussion 172. Merry, B. J. (2002). Molecular mechanisms linking calorie restriction and longevity. Int J Biochem Cell Biol 34, 1340-1354. Michishita, E., Park, J. Y., Burneskis, J. M., Barrett, J. C., and Horikawa, I. (2005). Evolutionarily Conserved and Nonconserved Cellular Localizations and Functions of Human SIRT Proteins. Mol Biol Cell. Migliaccio, E., Giorgio, M., Mele, S., Pelicci, G., Reboldi, P., Pandolfi, P. P., Lanfrancone, L., and Pelicci, P. G. (1999). The p66shc adaptor protein controls oxidative stress response and life span in mammals. Nature 402, 309-313. Miller, C. A., and Kowalski, D. (1993). cis-acting components in the replication origin from ribosomal DNA of Saccharomyces cerevisiae. Mol Cell Biol 13, 5360-5369. Min, J., Landry, J., Sternglanz, R., and Xu, R. M. (2001). Crystal structure of a SIR2 homologNAD complex. Cell 105, 269-279. Moazed, D. (2001). Common themes in mechanisms of gene silencing. Mol Cell 8, 489-498. Morley, J. F., and Morimoto, R. I. (2004). Regulation of longevity in Caenorhabditis elegans by heat shock factor and molecular chaperones. Mol Biol Cell 15, 657-664. 49 Mortimer, R. K., and Johnston, J. R. (1959). Life span of individual yeast cells. Nature 183, 17511752. Motta, M. C., Divecha, N., Lemieux, M., Kamel, C., Chen, D., Gu, W., Bultsma, Y., McBurney, M., and Guarente, L. (2004). Mammalian SIRT represses forkhead transcription factors. Cell 116, 551-563. Moynihan, K. A., Grimm, A. A., Plueger, M. M., Bernal-Mizrachi, E., Ford, E., Cras-Meneur, C., Permutt, M. A., and Imai, S. (2005). Increased dosage of mammalian Sir2 in pancreatic beta cells enhances glucose-stimulated insulin secretion in mice. Cell Metab 2, 105-117. Muller, I. (1985). Parental age and the life-span of zygotes of Saccharomyces cerevisiae. Antonie Van Leeuwenhoek 51, 1-10. Nasmyth, K., and Dirick, L. (1991). The role of SWI4 and SWI6 in the activity of G1 cyclins in yeast. Cell 66, 995-1013. Nemoto, S., Fergusson, M. M., and Finkel, T. (2004). Nutrient availability regulates SIRT1 through a forkhead-dependent pathway. Science 306, 2105-2108. Nemoto, S., and Finkel, T. (2002). Redox regulation of forkhead proteins through a p66shcdependent signaling pathway. Science 295, 2450-2452. Nomura, M. (2001). Ribosomal RNA genes, RNA polymerases, nucleolar structures, and synthesis of rRNA in the yeast Saccharomyces cerevisiae. Cold Spring Harb Symp Quant Biol 66, 555-565. North, B. J., Marshall, B. L., Borra, M. T., Denu, J. M., and Verdin, E. (2003). The human Sir2 ortholog, SIRT2, is an NAD+-dependent tubulin deacetylase. Mol Cell 11, 437-444. North, B. J., and Verdin, E. (2004). Sirtuins: Sir2-related NAD-dependent protein deacetylases. Genome Biol 5, 224. Ozcan, S., and Johnston, M. (1999). Function and regulation of yeast hexose transporters. Microbiol Mol Biol Rev 63, 554-569. Paradis, S., and Ruvkun, G. (1998). Caenorhabditis elegans Akt/PKB transduces insulin receptorlike signals from AGE-1 PI3 kinase to the DAF-16 transcription factor. Genes Dev 12, 24882498. Parisi, M., and Lin, H. (2000). Translational repression: a duet of Nanos and Pumilio. Curr Biol 10, R81-83. 50 Park, P. U., Defossez, P. A., and Guarente, L. (1999). Effects of mutations in DNA repair genes on formation of ribosomal DNA circles and life span in Saccharomyces cerevisiae. Mol Cell Biol 19, 3848-3856. Picard, F., Kurtev, M., Chung, N., Topark-Ngarm, A., Senawong, T., Machado De Oliveira, R., Leid, M., McBurney, M. W., and Guarente, L. (2004). Sirtl promotes fat mobilization in white adipocytes by repressing PPAR-gamma. Nature 429, 771-776. Rine, J., and Herskowitz, I. (1987). Four genes responsible for a position effect on expression from HML and HMR in Saccharomyces cerevisiae. Genetics 116, 9-22. Rogina, B., and Helfand, S. L. (2004). Sir2 mediates longevity in the fly through a pathway related to calorie restriction. Proc Natl Acad Sci U S A 101, 15998-16003. Rogina, B., Helfand, S. L., and Frankel, S. (2002). Longevity regulation by Drosophila Rpd3 deacetylase and caloric restriction. Science 298, 1745. Rolland, F., Winderickx, J., and Thevelein, J. M. (2001). Glucose-sensing mechanisms in eukaryotic cells. Trends Biochem Sci 26, 310-317. Shou, W., Sakamoto, K. M., Keener, J., Morimoto, K. W., Traverso, E. E., Azzam, R., Hoppe, G. J., Feldman, R. M., DeModena, J., Moazed, D., et al. (2001). Netl stimulates RNA polymerase I transcription and regulates nucleolar structure independently of controlling mitotic exit. Mol Cell 8, 45-55. Sinclair, D., Mills, K., and Guarente, L. (1998). Aging in Saccharomyces cerevisiae. Annu Rev Microbiol 52, 533-560. Sinclair, D. A., and Guarente, L. (1997). Extrachromosomal rDNA circles--a cause of aging in yeast. Cell 91, 1033-1042. Smeal, T., Claus, J., Kennedy, B., Cole, F., and Guarente, L. (1996). Loss of transcriptional silencing causes sterility in old mother cells of S. cerevisiae. Cell 84, 633-642. Smith, J. S., Brachmann, C. B., Celic, I., Kenna, M. A., Muhammad, S., Starai, V. J., Avalos, J. L., Escalante-Semerena, J. C., Grubmeyer, C., Wolberger, C., and Boeke, J. D. (2000). A phylogenetically conserved NAD+-dependent protein deacetylase activity in the Sir2 protein family. Proc Natl Acad Sci U S A 97, 6658-6663. Sutton, A., Immanuel, D., and Arndt, K. T. (1991). The SIT4 protein phosphatase functions in late G1 for progression into S phase. Mol Cell Biol 11, 2133-2148. Tadauchi, T., Matsumoto, K., Herskowitz, I., and Irie, K. (2001). Post-transcriptional regulation through the HO 3'-UTR by Mpt5, a yeast homolog of Pumilio and FBF. Embo J 20, 552-561. 51 Tanner, K. G., Landry, J., Sternglanz, R., and Denu, J. M. (2000). Silent information regulator 2 family of NAD- dependent histone/protein deacetylases generates a unique product, -O-acetylADP-ribose. Proc Natl Acad Sci U S A 97, 14178-14182. Tanny, J. C., Dowd, G. J., Huang, J., Hilz, H., and Moazed, D. (1999). An enzymatic activity in the yeast Sir2 protein that is essential for gene silencing. Cell 99, 735-745. Tatar, M., Kopelman, A., Epstein, D., Tu, M. P., Yin, C. M., and Garofalo, R. S. (2001). A mutant Drosophila insulin receptor homolog that extends life-span and impairs neuroendocrine function. Science 292, 107-110. Thevelein, J. M., and de Winde, J. H. (1999). Novel sensing mechanisms and targets for the cAMP-protein kinase A pathway in the yeast Saccharomyces cerevisiae. Mol Microbiol 33, 904918. Thore, S., Mauxion, F., Seraphin, B., and Suck, D. (2003). X-ray structure and activity of the yeast Pop2 protein: a nuclease subunit of the mRNA deadenylase complex. EMBO Rep 4, 11501155. Tissenbaum, H. A., and Guarente, L. (2001). Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature 410, 227-230. Tu, M. P., Epstein, D., and Tatar, M. (2002a). The demography of slow aging in male and female Drosophila mutant for the insulin-receptor substrate homologue chico. Aging Cell 1, 75-80. Tu, M. P., Yin, C. M., and Tatar, M. (2002b). Impaired ovarian ecdysone synthesis of Drosophila melanogaster insulin receptor mutants. Aging Cell 1, 158-160. Tucker, M., Staples, R. R., Valencia-Sanchez, M. A., Muhlrad, D., and Parker, R. (2002). Ccr4p is the catalytic subunit of a Ccr4p/Pop2p/Notp mRNA deadenylase complex in Saccharomyces cerevisiae. Embo J 21, 1427-1436. Uesono, Y., Toh-e, A., and Kikuchi, Y. (1997). Ssdlp of Saccharomyces cerevisiae associates with RNA. J Biol Chem 272, 16103-16109. van der Horst, A., Tertoolen, L. G., de Vries-Smits, L. M., Frye, R. A., Medema, R. H., and Burgering, B. M. (2004). FOXO4 is acetylated upon peroxide stress and deacetylated by the longevity protein hSir2(SIRT1). J Biol Chem 279, 28873-28879. Weindruch, R., Kayo, T., Lee, C. K., and Prolla, T. A. (2002). Gene expression profiling of aging using DNA microarrays. Mech Ageing Dev 123, 177-193. Weindruch, R., Naylor, P. H., Goldstein, A. L., and Walford, R. L. (1988). Influences of aging and dietary restriction on serum thymosin alpha 1 levels in mice. J Gerontol 43, B40-42. 52 Wood, J. G., Rogina, B., Lavu, S., Howitz, K., Helfand, S. L., Tatar, M., and Sinclair, D. (2004). Sirtuin activators mimic caloric restriction and delay ageing in metazoans. Nature 430, 686-689. 53 FIGURES Figure 1. Survival curves from distantly related species display similar shapes. Typical mortality data from Homo sapiens (Sinclair et al., 1998), Mus musculus, Caenorhabditis elegans (Tissenbaum and Guarente, 2001), and S. cerevisiae (Kaeberlein et al., 1999) are presented. This figure adapted from Matt Kaeberlein, Genetic Analysis of Longevity in Saccharomyces cerevisiae, Graduate Thesis, MIT 2002. Reproduced here by permission. 54 HUMAN MOUSE 100 90 100 90 80 80 70 'f: 70 _ :; 60 ~60 ~50 ~50 >40 a: '~ 40 aJ ::::>30 20 CJ) 10 30 20 10 o o o 1‫סס‬oo -30000 20000 0 300 Days WORM 1200 YEAST REPLICATIVE 100 100 90 80 90 _ 70 _ ~8O ~ 600 900 Days 80 70 ~60 50 ~50 ~::;) 40 "J) 30 .~ 40 :J 0030 20 10 20 10 o o 5 10 15 Days 20 o 25 FIGURE 1 55 o 10 20 Generation 30 40 Figure 2. Caloric restriction (CR) in yeast increases budding life span by activating Sir2p. Two separate mechanisms relieve inhibition of Sir2p (and its homolog Hst2p) to promote rDNA stability and longevity. CR also induces a Sir2p- and Hst2p-independent extension of life span by an unknown mechanism. 56 Yeast Caloric Restriction V ft Respiration f NAD+/NADH Ratio 9 A Sir2p (and Hst2) Activity rDNA Stability LONGEVITY FIGURE 2 57 ft Pnclp IL Nicotinamide Figure 3. Components of the insulin/IGF- 1 signaling pathway promote life span in yeast, worms, flies, and mice. Details of each pathway are described in the text. Question marks indicate plausible regulatory connections within each pathway, also discussed in the text. 58 5W- ~ U: 59/ 0 0 t~~~ 0 . c _ el) C c~~ 75~~~~~~~~u 5 C..- 05 ~ \ \ L \ 8_~~~5 CR 59 60 Chapter2 Ssdlp Directly Represses Translation of Uthlp to Increase Longevity, Stress Resistance, and Cell Wall Stability in Saccharomyces cerevisiae This chapter will be submitted for publication. The authors are Gregory Liszt, Matt Kaeberlein, Sarah Buckley, and Leonard Guarente. I contributed Figures lc, Figure 4, Figure 5, Figure 6, and Figure 7, and all unpublished observations referred to in the text. I also developed the mechanistic model and wrote the manuscript. 61 SUMMARY In yeast, natural polymorphisms at the SSDI locus strongly affect aging, stress resistance, pathogenicity, and cell wall integrity. The active SSD1-V allele promotes longevity by a novel mechanism independent of rDNA recombination and extrachromosomal rDNA circle (ERC) formation, two processes tightly linked to known pathways of yeast aging. Here, we report that Ssdlp, an RNA-binding protein, strongly represses the in vivo translation of Uthlp, one of the original proteins implicated in the yeast aging process. Deletion of UTH1 from ssdl-d cells phenocopies the addition of SSD1-V, causing a long life span and unique stress resistance profile. Furthermore, SSD1-V and uthlA both suppress mutations in several apparently unrelated genes including MPT5, SIT4, and CCR4. Uthlp levels are significantly reduced in SSD1-V cells, although the level of the UTH1 message is not influenced by SSD1 under normal growth conditions. Ssdlp physically interacts with the UTH1 mRNA, and these molecules can be copurified from cells. Interestingly, this interaction depends on sequence elements in the 5'-UTR of UTH1. SSD1-V cells lacking the 5'-UTR of UTH1 fail to show repression of Uthlp and phenotypically resemble ssdl-d cells. These results demonstrate that the direct translational repression of Uthlp by Ssdlp is necessary and sufficient to account for several phenotypic hallmarks of SSD1-V cells. These results also suggest that many of the defects of ssdl-d cells arise from a misregulation of Uthlp. 62 INTRODUCTION In Saccharomyces cerevisiae, dividing mother cells undergo an aging process that limits their replicative potential and induces pronounced changes in cell morphology, metabolism, genomic stability, transcriptional silencing, and the cell wall (Bitterman et al., 2003; Sinclair et al., 1998). Mutations that extend the yeast replicative life span have been used to study aging, and several of these mutations have led to the discovery of conserved mechanisms regulating aging in such diverse organisms as worms, flies, and even mice (Guarente and Picard, 2005; Kenyon, 2005). With successive cell divisions, instability at the rDNA locus and extrachromosomal rDNA circle (ERC) accumulation exponentially increase the likelihood of senescence and mortality in yeast mother cells (Sinclair and Guarente, 1997). Aging can be forestalled by caloric restriction, a regimen that increases the activity of Sir2p (Anderson et al., 2003; Lin et al., 2000; Lin et al., 2004; Lin et al., 2002), an important regulator of aging in yeast and some metazoans (Kaeberlein et al., 1999; Rogina and Helfand, 2004; Tissenbaum and Guarente, 2001; Wood et al., 2004). SSD1 is a polymorphic locus affecting diverse cellular processes including high temperature growth, cell integrity (Kaeberlein and Guarente, 2002), pathogenicity (Wheeler et al., 2003), and mother-cell aging (Kaeberlein et al., 2004a). Naturally occurring polymorphisms of the SSD1 gene fall into two phenotypic classes: active SSD1-V alleles and inactive ssdl-d alleles, classified on the basis of their ability to suppress mutations in the Sit4p phosphatase. SSD1-V, which suppresses sit4A lethality, has also been isolated as a suppressor of diverse genetic mutations affecting the cell wall stability, protein kinase A activity, TOR-signaling, and mRNA transcription and 63 processing (summarized in (Kaeberlein et al., 2004a). Addition of SSDI-V dramatically extends lifespan of ssdl-d strains by a novel, as yet uncharacterized mechanism (Kaeberlein et al., 2004a). This mechanism is Sir2p-independent, affecting neither rDNA silencing nor ERC formation, two processes linked to yeast aging (Kaeberlein et al., 2004a). Although SSDI is well conserved from yeast to humans, little is known of its biochemical function. A previous study has demonstrated that Ssdlp, a cytoplasmic protein, can bind RNA in vitro (Uesono et al., 1997), but it is not known whether this activity is physiologically significant. RNA binding proteins have been shown to regulate a variety of posttranscriptional processes including mRNA localization and stability, translation and splicing (Kuersten and Goodwin, 2003). RNA-binding proteins feature prominently in metazoan development, often as repressors of translation (Dreyfuss et al., 2002; Johnstone and Lasko, 2001; Maniatis and Reed, 2002; Mazumder et al., 2003). For example, early in embryogenesis, the Drosophila PUMILIO protein binds maternal hunchback mRNA via sequence elements in the 3' untranslated region (UTR), repressing hunchback translation at the posterior pole by a mechanism also requiring the NANOS protein (Parisi and Lin, 2000). Similarly, the precise control of translation and localization results in highly regulated expression of other mRNAs including BICOID, OSCAR, and NANOS itself (Johnstone and Lasko, 2001). In C. elegans, the RNAbinding protein GLD-1 influences several germ cell fate decisions by repressing the translation of numerous maternal mRNAs, often in a spatially dependent manner (Jan et al., 1999; Lee and Schedl, 2001; Lee and Schedl, 2004; Marin and Evans, 2003; Mootz et al., 2004; Xu et al., 2001). Interestingly, GLD-1 also represses translation of CEP-l/p53 64 to affect germ cell apoptosis in response to DNA damage (Schumacher et al., 2005). The translation of the GLD-1 message is, in turn, regulated by the RNA-binding protein FBF (fem-3 mRNA Binding Factor; (Crittenden et al., 2002; Zhang et al., 1997). FBF consists of two nearly identical proteins, both of which contain the eight consecutively repeated PUMILIO-homology domains required for RNA-binding activity in the Drosophila protein (Zamore et al., 1997). Notably, a recent study systematically identified RNA targets of the yeast PUMILIO-FBF family of proteins (Pufl-5p), uncovering a novel mechanism whereby specific RNA-binding proteins physically associate with functionally related groups of target RNA molecules to regulate their post-transcriptional properties (Gerber et al., 2004). The best characterized of the five yeast Pumilio family members is Puf5/Mpt5/Uth4, a protein important for pheromone response (Chen and Kurjan, 1997), life span regulation (Kennedy et al., 1995; Kennedy et al., 1997), and cell wall integrity (Kaeberlein and Guarente, 2002). Deletion of MPT5 shortens life span by 50%, and overexpression extends life span by up to 40% (Kennedy et al., 1997), underscoring the importance of this post-transcriptional regulator in the aging process. Cells lacking Puf5/Mpt5/Uth4 display increased telomeric silencing, decreased rDNA silencing, and an aberrant distribution of Sir3p in the absence of Sir4p (Gotta et al., 1997). In the absence of SSDI-V, mpt5A mutant cells also suffer from several deficiencies associated with a weakened cell wall, including sensitivity to calcofluor white (CFW), sensitivity to sodium dodecyl sulfate (SDS), and sorbitol-remedial temperature sensitivity (Kaeberlein 65 and Guarente, 2002). Genetic analysis suggests that Mpt5p, Ssdlp, and Pkclp define three parallel pathways to ensure cell wall integrity. Uthlp is a member of a family of proteins specific to fungi, known as the SUN family of proteins because its members include Simlp, Uthlp, and Nca3p as well as the more recently identified Sun4p (Mouassite et al., 2000b). The UTHI gene was originally identified as a loss-of-function suppressor of the short life span and stress sensitivity caused by a C-terminal truncation of Mpt5p (Kennedy et al., 1995). Subsequent studies have shown that deletion of UTHI increases life span, resistance to peroxides (Bandara et al., 1998), and resistance to starvation- and rapamycin-induced autophagy (Kissova et al., 2004). Uthlp localizes to the outer mitochondrial membrane and the cell wall, and has been shown to be highly glycosylated (Velours et al., 2002). We noticed that deletion of UTHI, like addition of SSDI-V, suppresses the cell wall defects and short life span of an mpt5A ssdl-d mutant. This similarity between SSD1-V and uthlA prompted us to further investigate the relationship between these two genes. Here, we present evidence that Ssdlp directly represses Uthlp, and by that repression effects large changes in stress resistance, life span, and cell wall integrity. MATERIALS AND METHODS Strains and genetic techniques. The strains used in this study are listed in Table 1. All strains were derived from either PSY316 (described in (Park et al., 1999) or W303R (described in (Mills et al., 1999). W303R has previously been shown to carry the ssdl-d2 allele (Sutton et al., 1991). PSY316 contains an ssdl-d nonsense mutation predicted to truncate Ssdlp at codon 132 (Nicholas Bishop, personal communication). Yeast transformations were accomplished using the 66 lithium acetate method (Gietz et al., 1992). Genetic crosses, sporulation, and tetrad analysis were carried out by standard techniques (Sherman and Hicks, 1991). Marker segregation in viable spore clones was used to infer the genotype of inviable spore clones from the same tetrad whenever possible. The ccr4::HIS3, swi6::TRPI, mpt5::LEU2, and mptS::HIS3 disruptions were generated using plasmids described in (Kaeberlein and Guarente, 2002). All other gene deletions were done by transforming cells with PCR-amplified disruption constructs. For each gene disruption, the entire open reading frame (ORF) was removed except for SWI4 disruption, which was done'by replacing the first 2kb of the SWI4 ORF with HIS3. All disruptions were verified phenotypically, by PCR, or both. The SSD1-V integrating plasmids p406SSD1 and p405SSD1, as well as the ARS-CEN variants of those plasmids (p416SSD1 and p415SSD1) were previously described (Kaeberlein and Guarente, 2002). Unless otherwise indicated, all SSDI-V strains contain SSD1-V integrated at the marker locus and still carry the ssdl-d allele at the native locus. Deletion of ssdl-d does not affect any of the phenotypes tested, including life span, growth at 30°C, 37°C, or 40°C, or sensitivity to calcofluor white, paraquat or hydrogen peroxide (Kaeberlein and Guarente, 2002; Kaeberlein et al., 2004b). A yeast strain of the BY4741 background containing a tandem affinity purification (TAP) tag integrated in-frame at the 3' end of the SSD1 gene was obtained from OpenBiosystems (Ghaemmaghami et al., 2003). The following primers were used to PCR amplify a region of SSD1-TAP encompassing the final 1785 bp of the SSD1 ORF as well as the entire affinity tag: TAAGGGATAACAATTTTCTTT (forward primer) and TTTGCGGCCGCGAAAGAGTTACTCAAGAATAA (reverse primer). This PCR product was digested with BstBI and NotI and ligated into plasmid p416SSDl cut with the two same enzymes. The resulting plasmid p416SSD1-TAP was transformed into PSY316AR, and expression of an active fusion protein was confirmed by phenotype and Western blot using a rabbit polyclonal anti-TAP antibody (OpenBiosystems). 67 The UTH]TERMNATOR::ADHTERMNATOR-kanMX6 replacement was constructed by one-step PCR-mediated gene disruption as follows. Plasmid pFA6-3HA-kanMX6 was used as a template to generate the disruption construct, as this plasmid contains the ADHTERMINATOR sequence (Longtine et al., 1998). This sequence was amplified with the following primers incorporating regions flanking the UTH1 terminator sequence (indicated in uppercase): TTCAGTTACTTCTGGTTCTGCTAACTTTGTCTTCTACTAGacttctaaataagcgaattt (forward) TGTTATATAATATGCTATATATGATAATATCTAGTGGTATgaattcgagctcgtttaaac (reverse). This construct was transformed into strain GLY3002 and gene replacement was verified by PCR. Spot Assays. Calcofluor white (Fluorescence Brightener 28, Sigma), hydrogen peroxide, and paraquat (Methyl Viologen, Sigma) were added at the designated concentrations to cooled (<60°C), premixed YPD containing 3% agar (w/v). Plates were dried at 37°C and used for serial dilution spot assays the same day. These assays were performed by first growing cells overnight in YPD at 30°C, then diluting cultures back 50-fold in fresh YPD and allowing to grow for an additional 3-4 hours. For each strain, a series of 10-fold dilutions was prepared in fresh YEP over a range of concentrations from 10-' to 10-5 relative to the initial culture. Spots of 5[tL from each dilution series were then plated to the appropriate media using a multi-channel pipettor. Western Blotting. Total protein extracts from yeast were prepared as described (Kushnirov, 2000) by brief NaOH treatment followed by boiling in optimized SDS sample buffer. From 2.5 O.D. units of cells, 50[tL SDS-solubilized proteins were prepared, a 10[tL quantity of which was used for Western analysis. Samples were resolved by SDS-PAGE on a 4-15% Tris-HCl gradient gel (Bio-Rad), transferred to nitrocellulose membrane and blocked using 4% nonfat dry milk in PBS with 0.1% Tween-20 (PBS-T). Uthlp was detected by incubation with rabbit polyclonal anti-Uthlp antisera diluted 1: 2,500 in PBS-T. This antibody, a generous gift from S. Manon, has been characterized previously (Velours. et al., 2002). Peroxidase-conjugated donkey anti-rabbit IgG secondary antibody (Amersham Biosciences) was used at a dilution of 1: 10,000. 68 Chemiluminescent detection was achieved by incubation with the ECL reagent (Amersham Biosciences). Blots were stripped for 20 minutes in 50 ° C stripping buffer (62.5 mM Tris-HCl pH 7.5, 2% SDS, 100 mM beta-mercaptoethanol) and reprobed with anti-c tubulin as a loading control. Northern Blotting. Total cellular RNA was prepared and purified using the Qiagen Rneasy kit according to the manufacture's instructions. Briefly, cells were grown to early log phase, and 2.5 O.D. units of cells were harvested and resuspended in 2 mL buffer Y1 (M sorbitol, 0.1M EDTA pH 7.4, 0.1% -mercaptoethanol, 100 U/mL zymolyase 100-T). Cells were incubated at 30° C for 25 minutes to create spheroplasts. Subsequently, spheroplasts were lysed, and the lysate was applied to a Qiagen Rneasy column, washed three times and eluted two times in 30 [tL Rnase-free water each time. Final RNA concentrations were measured by spectrophotometer, and samples were concentrated in a speed-vac evaporator and resuspended in Rnase-free water at a final concentration of 3tg/tL. For each sample, 12 [tg total RNA (4[tL) was mixed with three volumes of formaldehyde loading buffer (Ambion) and denatured for 15 minutes at 65° C. Ethidium bromide was added to a final concentration of 10 [tg/mL, and samples were run on a formaldehyde denaturing gel (1 % Agarose, lx MOPS, 0.6 M formaldehyde). The gel was run at 100 V for 2.5 hours in gel running buffer (lx MOPS, 0.2 M formaldehyde), bands were visualized and photographed under UV light, and the gel was transferred overnight to GeneScreen membrane (NEN) by capillary transfer in o10X SSC buffer as described (Ausubel, 1999). Membranes were UV-crosslinked two times in a Stratagene Stratalinker and prehybridized for one hour at 42° C in pre-warmed UltaHybe solution (Ambion). During this period, the PrimeIt Random Primer labeling kit (Stratagene) was used to generate corresponding 32 p radiolabeled probe to a 350 bp region of the UTH1 ORF. Blots were probed overnight and on the next day washed twice under both low stringency (2x SSC, 0.1% SDS; 5 minutes at room temperature) and high stringency (0.2x SSC, 0.1% SDS; 15 minutes at 42 ° C) conditions. 69 Blots were analyzed by autoradiography, stripped for 10 minutes in 5% SDS at 95 ° C, and reprobed for ADHI by the same method. Affinity purifications. TAP purifications were conducted essentially as described (Gerber et al., 2004) with a few modifications. A 1 L YPD culture of cells grown to O.D. 0.7 at 30 ° C was harvested by centrifugation, washed twice with ice cold buffer A (20 mM Tris-HC pH 8.0, 140 mM Kcl, 1.8 mM MgC12, 0.1% NP-40, 0.02 mg/mL heparin) and resuspended in 5 mL buffer B (buffer A plus 0.5mM DTT, 20 U/mL DNAse I, 100 U/mL RnaseOut (Invitrogen), 0.2 mg/mL heparin, and protease inhibitors (Roche)). Cells were subjected to glass bead lysis and whole cell extracts were incubated with 400[tL (50% (v/v)) IgG-agarose beads (Sigma) for 4h at 4 C. Beads were washed four times for 15 min each wash at 4 ° C in buffer C (20 mM Tris-HCl ph 8.0, 140 mM Kcl, 1.8 mM MgC12, 0.5 mM DTT, 0.01% NP-40, 10 U/mL RnaseOut). TAP-tagged Ssdlp was released from the beads by cleavage with 100 U of TEV protease for 1.5 h at 16 ° C with gentle periodic agitation. RNA was isolated from final eluates as well as whole cell extracts using the Rneasy kit (Qiagen). Final RNA concentrations were determined spectrophotometric measurement. RT-PCR. One Step RT-PCR (Qiagen) was performed on RNA purified from whole cell extracts and final TEV eluates. Each reaction was first optimized on 500 ng whole cell RNA to determine the most effective number of PCR cycles for gene detection. The following primers were used: UTHI forward-TCGGCGTCTCCCAGATCAAAG,UTHI reverseCCGTCTGCGTTGACGTACTG, ACT1 forward-ATTCTGAGGTTGCTGCTT, AACTCTCAATTCGTTGTAGA, CBKI forward- ATGTATAATAGCAGCACCA, TGCACCTGCTCCAGCGGACA, ACT] reverse- CBKI reverse- ATCATCTGATTTGAAGCTG, NCA3 forward- NCA3 reverse-AGTAGTAGCCATCCTTACAT. Reactions were carried out in a 50 [tL volume of the following composition: 1X OneStep RTPCR Buffer (Qiagen) containing 400 [tM of each dNTP, 0.6 [tM each primer, 5 units Rnasin, 500 70 by ng template RNA (or equivalent volume from mock IP samples), and 2 [tL OneStep RT-PCR Enzyme Mix (Qiagen). RT-PCR cycles were as follows: 30 min at 50°C (reverse transcription); 15 min at 95°C (initial PCR activation step); either 30 (ACT]) or 35 (CBKI, UTH1, NCA3) cycles of the following three steps: 30 sec at 94°C (denaturation); 1 min at 65°C (annealing); 1 minute at 72°C (extension); 10 minute final extension. 71 RESULTS We have previously found that SSD1-V suppresses the short life span and cell wall defects caused by deletion of MPT5 in an ssdl-d background (Kaeberlein and Guarente, 2002). To determine whether uthlA would have a similar effect on these mpt5A phenotypes, we deleted UTH1 from ssdl-d mpt5A cells. As expected, mpt5A alone severely impaired growth at 37 ° C and caused dose-dependent sensitivity to the cell wall-perturbing agent calcofluor white (CFW) (Figure 1A). Strikingly, when UTHI was also deleted from these cells, they grew well at 37 ° C and nearly as well as wild type cells in the presence of CFW (Figure 1A). Similar to what is observed in UTH1 MPT5 SSDI- V cells (Kaeberlein and Guarente, 2002), uthlA MPTS ssdl-d cells are more resistant to CFW than UTH1 MPTS ssdl-d cells. Thus, deletion of UTH1 behaves phenotypically like addition of SSD1-V, with both modifications strongly promoting cell integrity. Spurred by this finding, we sought to identify other phenotypic similarities between uthlA and SSD1-V cells. Uthlp was originally cloned by complementation of the sensitivity to paraquat displayed by uthl mutant cells (Kennedy et al., 1995). Paraquat generates superoxide radicals, suggesting that cells lacking Uthlp are sensitive to oxidative stress. However, this view is complicated by the fact that uthlA also causes resistance to hydrogen peroxide (Bandara et al., 1998; Kennedy et al., 1995). We were interested in determining what effect, if any, SSD1-V would have on resistance to these oxidative agents. As expected, uthlA in strain PSY316 resulted in resistance to hydrogen peroxide and sensitivity to paraquat (Figure 2). Very similarly, addition of SSDI-V caused resistance to hydrogen peroxide and sensitivity to paraquat, demonstrating that both SSD1-V and Uthlp affect the oxidative stress response of cells. 72 During the course of our analysis, we observed that SSD1-V and uthlA cells both have a greater maximum growth temperature than ssdl-d cells. Strain PSY316, which harbors the ssdl-d allele, is unable to grow at 40° C; however, both SSD1-V and uthlA cells grow very well at this temperature (Figure 2). This phenotype could be the result of a strengthened cell wall, or could reflect some other function shared by these proteins. We next examined the relative and combined effects of SSD1-V and uthlA on life span. Like SSD1-V, uthlA extends wild type ssdl-d life span. SSD1-V UTH1 cells have a much longer life span than ssdl-d uthl cells, suggesting that uthlA has a weaker longevity-promoting effect than addition of SSD1-V (Figure C). Interestingly, SSD1-V uthlA cells have a life span comparable to ssdl-d uthlA cells and shorter than those of the SSD1-V UTH1 genotype. These results indicate that the shorter-lived uthl mutation is epistatic to SSDI-V with respect to life span. We sought to establish a genetic relationship between UTH1 and SSD1. SBF, a transcriptional activator consisting of the cell-cycle regulated proteins Swi4p and Swi6p, acts downstream of protein kinase C (Pkclp) to promote cell wall biosynthesis (Igual et al., 1996). Ccr4p, a component of the cytoplasmic deadenylase complex (Thore et al., 2003; Tucker et al., 2002) also functions in an RNA Polymerase II-containing complex required for PKCl-dependent transcriptional regulation of multiple genes required for proper cell wall structure (Chang et al., 1999). Mutation of either CCR4 or SBF is lethal in combination with mutation of MPT5 in an ssdl-d or ssdlA background (Kaeberlein and Guarente, 2002). In order to determine whether uthlA, like addition of SSD1-V, would suppress this synthetic lethality, we generated the following diploid mutants in a W303R (ssdl-d) background: ccr4A/CCR4 mpt5A/MPT5 uthlA/UTH]; swi4A/SWI4 73 mpt5A/MPT5 uthlA/UTHI; swi6A/SWI6 mpt5A/MPTS uthlA/UTH1. These strains were sporulated, tetrads were dissected, and the genotypes of the surviving spores were determined. While ccr4A mpt5A ssdl-d UTH1, swi4A mpt5A ssdl-d UTH1 and swi6A mpt5A ssdl-d UTH1 spores were never viable, we were able to recover ccr4A mpt5A ssdl-d uthlA, swi4A mpt5A ssdl-d uthlA and swi6A mpt5A ssdl-d uthlAcells (Table 2). Deletion of UTH1 also partially rescued the growth defects and sensitivity to CFW caused by mutation of CCR4 or SBF (data not shown). These results suggest that Uthlp acts as a negative regulator of cell integrity in a pathway parallel to Mpt5p. Moreover, these results underline the striking similarity between mutation of UTH1 and addition of SSD1-V. We therefore wished to determine whether this similarity extended to processes other than cell integrity and life span. SSD1-V was first defined based on its ability to confer viability to a strain carrying a mutation in SIT4. In order to test whether uthlA could also confer viability to a sit4A mutant, a sit4A/SIT4 ssdl-d/ssdl-d uthlA/UTHI doubly heterozygous diploid was sporulated and tetrads were analyzed. The results are compiled in Table 2 and shown in Figure 3. As expected, 0/21 sit4A ssdl-d UTH1 spore clones were able to form viable colonies, confirming that sit4A and ssdl-d are synthetically lethal. In contrast, 14/14 sit4A ssdl-d uthlA spore clones formed colonies (Table 2), indicating that uthlA is an effective suppressor of the sit4A ssdl-d synthetic lethality. Like sit4A SSD1-V UTH1 cells, sit4A ssdl-d uthlA cells are slow growing, but are capable of vegetative growth indefinitely (Figure 3). Interestingly, maximal benefit to the sit4A cells appears to be obtained by either SSD 1-V or mutation of UTH1, as sit4A 74 SSD1-V uthlA cells display no greater growth advantage compared to sit4A ssdl-d uthlA and sit4A SSD1-V UTHI mutants (Figure 3). Taken together, these results suggest a model whereby Ssdlp represses Uthlp to affect cell integrity, life span, and stress resistance. Using Western blot analysis, we tested this possibility by comparing Uthlp levels in SSDI1-V and ssdl-d strains (Figure 4a). SSDI-V and ssdl-d cells grown in rich media were harvested during early-, mid-, and late-log phase, and protein extracts were subjected to Western blotting. As reported (Velours et al., 2002), a rabbit polyclonal antibody specific to Uthlp predominantly recognized a thick band of approximately 60 KDa corresponding to glycosylated Uthlp (Figure 4a). This band was absent from extracts of uthlA cells. Importantly, SSD1-V caused a dramatic reduction in Uthlp levels in all growth phases without significantly affecting the levels of P-tubulin. Similar results were obtained from cells grown at 37° C (data not shown), indicating that SSD -V represses Uthlp under a wide variety of culture conditions. Interestingly, in the absence of SSDI-V, Uthlp was upregulated during latelog phase growth, while in the presence of SSDI-V Uthlp remained barely detectable as nutrients in the culture became exhausted (Figure 4a). To determine whether SSDI-V repressed Uthlp at the transcriptional level, we performed Northern analysis on ssdl-d and SSDI-V cells. During log phase growth, UTHI mRNA levels were unchanged by SSDI-V (Figure 4b), despite the fact that this allele caused a striking reduction in Uthl protein. During late-log phase, as glucose was depleted from the growth medium, UTHI mRNA levels decreased significantly in SSDIV cells but not in ssdl-d cells (Figure 4b). All blots were stripped and reprobed for ADHI expression to demonstrate equal RNA loading. From these data, we conclude that 75 under normal growth conditions, SSDI-V reduces Uthl protein expression without decreasing UTH1 mRNA levels. The repression of Uthlp by SSD1-V under these conditions must therefore be post-transcriptional. Under conditions of high culture density and reduced glucose availability, SSDI-V further reduces Uthlp expression by a mechanism that decreases the prevalence of the UTHI mRNA. As all phenotypic analyses (including life span assays) were conducted under early log phase growth conditions, we sought to explain the regulation of Uthlp by Ssdlp under these circumstances. Two possible models might account for the post- transcriptional repression of the Uthl protein by Ssdlp. By the first model, Ssdlp would reduce Uthlp by lowering that protein's stability, making it more susceptible to degradation. To test this possibility, early-log phase ssdl-d and SSDI-V cells were treated with 100 ug/mL cycloheximide to inhibit protein synthesis. Proteins were extracted by boiling in modified SDS-sample buffer before cycloheximide treatment and every 70-seconds subsequently. Samples were then subjected to Western blotting, and Uthlp was detected using an anti-Uthlp polyclonal antibody. We observed that Uthlp decayed with identical kinetics in cycloheximide-treated cells of both the ssdl-d and SSDI-V genotypes (Figure 5), indicating that the observed decrease in Uthlp in log phase SSDI-V cells was not the result of accelerated protein degradation. In both cases, an initially rapid period of degradation (Figure 5a) was followed by a very long period (>2hr) during which continued protein decay was undetectable (Figure 5b). Over the first 225 seconds, decay in both ssdl-d and SSDI-V strains occurred with a half-life time constant of approximately one minute, very short even for a yeast protein (Beyer et al., 2004; Varshavsky, 1996). Following this period, a fraction of the existing Uthlp in the 76 cell remained stable for hours. Importantly, this unusual protein stability profile applied to cells of both ssdl-d and SSD1-V genotypes, leading us to conclude that SSDI-V does not repress Uth lp at the level of protein stability. A second model postulates that SSDI, an RNA-binding protein, might be acting as a post-transcriptional repressor by inhibiting the translation of the UTHI message. In order to purify ribonucleoparticles associated with Ssdlp, a single-copy plasmid containing the SSDI1-Vgene fused to an in-frame C-terminal tandem affinity purification (TAP) tag sequence was transformed into strain PSY316AR. The same plasmid lacking the TAP tag sequence was used to obtain transformants of strain PSY316AR to provide a control for non-specific enrichment during the purification process. Affinity purifications were performed from these two strains using agarose-IgG beads under conditions previously described to preserve protein-RNA interactions (Gerber et al., 2004). RNA was then isolated from purified TEV protease-eluted protein samples by column chromatography using the Qiagen Rneasy kit. From SSD1-TAP cells, we obtained between 0.2-0.6 [tg RNA per 1 L of starting culture (data not shown), a result comparable to what has been reported for other yeast RNA binding proteins (Gerber et al., 2004). We did not obtain a significant quantity of RNA from the final elution of untagged control samples, suggesting that the RNA fraction obtained from Ssdlp-TAP cells consisted of RNA molecules specifically associated with the tagged protein. To confirm this finding and to test for a specific physical association between Ssdlp-TAP and UTHI mRNA, purified RNA fractions were probed for several candidate genes using RT-PCR. Primers specific to -actin, encoded by ACT], detected that gene's cDNA in input but not affinity-purified RNA preparations from both tagged and untagged 77 strains, indicating that this abundant transcript was not significantly associated with Ssdlp-TAP or control fractions. Similar results were obtained when extracts were probed for the UTHI family member NCA3 and the unrelated CBK1, encoding a kinase whose protein product physically interacts with Ssdlp (Ho et al., 2002; Racki et al., 2000). When used to probe total cellular RNA fractions from Ssdlp-TAP and mock affinity purified eluates, each pair of primers detected a band, confirming the efficacy of these primer pairs in the RT-PCR reaction. However, of the four genes analyzed, only primers specific to UTH1 succeeded in amplifying a product from Ssdlp-TAP purified fractions. This UTH] band was sequenced to confirm that it represented the predicted gene (not shown). This band was absent from the TEV eluate of untagged extracts. These results indicate that the UTH1 mRNA physically associates with Ssdlp-TAP, and that this association is specific, as Ssdlp-TAP fails to stably associate with the abundant ACT1 transcript or the NCA3 or CBKl transcripts. To identify the region of UTH1 mediating this interaction and to ask whether this interaction is required for the regulation of Uthlp by Ssdlp, mutants were generated substituting the upstream region (including the 5'-UTR) or the downstream region (including the 3'-UTR) of the ADH1 gene for that of UTH1 (Figure 7). These mutations were first analyzed in an SSD1-V background to determine whether either substitution affected the regulation of Uthlp by Ssdlp. Surprisingly, substitution of the 3'-UTR of UTH1 did not hinder the repressive effects of Ssdlp on Uthlp (Figure 7b), and SSD1-V UTHl 3 rUTR::ADH 3 UTRcells grew as well as SSD1-V cells at 40°C (Figure 7c). However, substitution of the promoter and 5'-UTR of UTHI with a truncated, less active form of the ADHI promoter and its 5'-UTR did abolish the repression by Ssdlp, also severely 78 limiting the ability of these cells to grow at 40°C (Figure 7). Therefore, we conclude that repression of Uthlp by Ssdlp requires the 5' upstream region of UTH1, and that this repression is required for the SSDI-V phenotype of high temperature growth. Preliminary results from affinity co-purification experiments indicate that the substitution of the UTHI promoter and upstream region abolishes the physical interaction between Ssdlp and the UTH1 mRNA. Validation of these results will be included upon submission of this manuscript. DISCUSSION Here, for the first time, we provide a molecular mechanism to explain the action of Ssdlp, a highly conserved protein with diverse effects on yeast physiology. Several lines of evidence now lead us to conclude that Ssdlp directly represses Uthlp by a posttranscriptional mechanism, and this repression is necessary and sufficient to account for vast effects of Ssdlp on life span, cell integrity, high temperature growth, and stress resistance. First, deletion of UTHI can effectively compensate for the null ssdl]-d allele, restoring longevity, increased cell wall stability, resistance to high temperature, and the unusual combination of peroxide resistance and paraquat sensitivity. Second, deletion of UTHI, like addition of SSDI-V, suppresses several genetic mutations in an ssdl-d background, including the lethalities caused by mpt5 swi4 ssdl-d, mpt5 swi6 ssdl-d, mpt5 ccr4 ssdl-d, and sit4 ssdl-d genotypes. These findings demonstrate that abrogation of Uthlp expression is sufficient to recapitulate widely varying phenotypes associated with SSDI -V. 79 Additional data show that not only is Uthlp reduction sufficient to vary stress resistance, life span, and cell wall stability in a manner reminiscent of SSDI-V, Uthlp is actually strongly reduced SSDI1-Vcells. This downregulation occurs without a significant effect on UTHI mRNA levels under normal culture conditions, pointing to a posttranscriptional mechanism of repression. Indeed, Ssdlp-TAP forms a specific physical association with the UTHI mRNA in vivo, and this association is required for the proper repression of Uthlp translation. SSDI-V cells in which this interaction has been abrogated by replacement of the UTH] 5' UTR fail to resist oxidative stress and high temperature, indicating that downregulation of Uthlp is required for these phenotypic hallmarks of SSD -V cells. A previous study established that Ssdlp possesses an in vitro RNA binding activity, mediated by a central domain within the protein (Uesono et al., 1997). For the first time, our results link this activity to the biological function of Ssdlp, raising the possibility that Ssdlp may regulate other transcripts in addition to that of UTH. Consistent with this possibility, Ssdlp-TAP co-purified with significant amounts of RNA even in mutants in which the interaction with UTHI was apparently abolished (preliminary data, not shown). Future experiments should aim to systematically identify transcripts bound by Ssdlp in vivo, as these represent likely candidates for posttranscriptional co-regulation by Ssdlp. The majority of mRNA-protein interactions characterized to date occur via sequence elements in the 3' UTR of the bound mRNA (Gerber et al., 2004; Kuersten and Goodwin, 2003). However, our data indicate that the interaction between Ssdlp and UTH1 depends on sequence elements within the 5' UTR of UTHI, and possibly 80 additional elements located elsewhere within the UTHI mRNA. So far, we have been unable to identify the precise nucleotide sequence sufficient to direct Ssdlp binding to an mRNA (our unpublished observations), possibly because of a requirement for additional sequence or structural characteristics within the ORF or even the 3' UTR. Interestingly, for the three yeast Pumilio family members for which consensus binding sites have been determined (Puf3-5), a significant fraction of these motifs reside within the open reading frames (ORFs) of target genes. We therefore cannot rule out the possibility that the UTH1 ORF contains additional sequence elements or structural characteristics required to direct Ssdlp binding to the mRNA. Perhaps a comprehensive identification of Ssdlp targets will also yield insight into the potentially complex issue of the sequence specificity of Ssdlp. While Ssdlp likely acts as post-transcriptional repressor of Uthlp during normal growth conditions, presence of Ssdlp results in a dramatic reduction in UTHI mRNA levels at the onset of the diauxic shift. Incidentally, we have also observed a similar reduction in UTHI mRNA in SSDI-V cells treated with rapamycin (G.L. ad L.G., unpublished observations), used in some studies as a mimic of starvation (Albig and Decker, 2001; Reinke et al., 2004). This Ssdlp-induced reduction in UTHI message could result from indirect effects of Ssdlp on transcription under these conditions. Alternatively, Ssdlp might specifically target UTHI mRNA for decay as starvation sets in and the cellular mechanisms favoring RNA decay over synthesis and translation begin to predominate. Both starvation and rapamycin treatment have been shown to activate mRNA turnover via the nonsense mediated decay pathway, destabilizing particular RNAs 81 (Albig and Decker, 2001). Determination of the effect of Ssdlp on the half-life of UTHI mRNA would help distinguish between these two possibilities. Mutation of UTHI results in a shorter life span extension than SSDI-V, and is epistatic to SSDI-V for life span. One possible explanation for this is that mutation of UTHI results in two opposing effects on life span, and a maximum life span is only obtained at an optimal level of the protein. Consistent with this model, SSD1-V cells do maintain a low level of Uthlp that might be responsible for their increased longevity compared to uthl null mutants. Although studies of Uthlp have primarily focused on the benefits conferred by the absence of this protein, there are reports indicating that uthlA mutants have subtle defects in mitochondrial function stemming from slightly reduced synthesis of critical mitochondrial proteins (Camougrand et al., 2000; Mouassite et al., 2000b). It is possible that these defects somehow begin to curtail life span past a certain very old age. Remaining members of the SUN family of proteins can compensate for the loss of any individual SUN protein to some extent. This is evidenced by the fact that double mutations within this family typically show much more severe phenotypes than single mutations (Camougrand et al., 2000; Mouassite et al., 2000b). Consistent with this notion, we have observed by transcriptional profiling that the SUN family member NCA3 is strongly induced by either uthlA or addition of SSDI-V (Kaeberlein et al., 2004a), possibly as a means of compensating for the loss of its homolog in both cases. It is interesting to note that Nca3p, although 63% identical to Uthlp, appears to be regulated very differently, showing induction rather than repression as a result of Ssdlp. This gene 82 therefore provided an excellent negative control in the purification of RBPs associated with Ssdlp-TAP in vivo. Exactly how Uthlp affects so many cellular processes has been the subject of several studies (Bandara et al., 1998; Camougrand et al., 2003; Camougrand et al., 2000; Kissova et al., 2004), but still remains a significant mystery. Effects on cell wall stability and structure could conceivably be mediated directly by Uthlp, which is not only partially localized to the cell wall but also bears some homology to a B-glucosidase of Candida wickerhamii (Mouassite et al., 2000a; Velours et al., 2002). No enzymatic activity has yet been shown for Uthlp. And while effects of uthlA on stress resistance have been hypothesized to account for its extension of replicative life span (Camougrand et al., 2004), we feel this is unlikely to be the case, as genetic modifications made specifically to increase peroxide resistance are insufficient to extend life span (Van Zandycke et al., 2002). Moreover, superoxide sensitivity resulting from mutation of superoxide dismutase generally causes a shortened life span (Fabrizio et al., 2004). A deletion of UTH1 was found to increase resistance to death induced by the ectopic expression of the mammalian apoptosis protein Bax (Camougrand et al., 2003). As yeast have been shown to display a form of cell death resembling apoptosis in some ways (Madeo et al., 2002), it is possible that dysregulation of Uthlp in ssdl-d cells shortens life span by increasing susceptibility to this form of death in very old cells. While Uthlp repression accounts for many phenotypes seen in SSD1-V cells, there is also evidence that Ssdlp has Uthlp-independent roles in the cell. Even in a UTHI mutant with the ADHI 5' UTR substituting for the endogenous 5'UTR, there still is a small effect on high temperature growth and stress resistance caused by SSDl-V, 83 despite the lack of Uthlp repression in this strain. This residual effect likely results from a parallel activity of Ssdlp affecting these two phenotypes. Furthermore, we have observed an SSDI1-Vgenetic interaction that is not recapitulated in uthlA cells. Namely, deletion of components of the Cbklp kinase complex, required for cell morphogenesis and the transcription of daughter-specific genes, is lethal in combination with SSD1-V but not with uthlA (G.L. and L.G., unpublished observations). Ssdlp now becomes the second translational repressor implicated in the coordinated maintenance of yeast longevity, cell integrity, and stress resistance. The first such protein, Mpt5p, has many well-studied homologs in metazoans including Drosophila PUMILIO and C. elegans FBF, both of which are specific repressors of translation. Several studies now indicate a conserved role for these homologs in the maintenance of germ line stem cells in Drosophila (Forbes and Lehmann, 1998; Lin and Spradling, 1997) and C. elegans (Crittenden et al., 2002). Thus, both PUMILIO and FBF promote continued cell divisions and delay the onset of an alternate state of differentiation and developmental potency. Intriguingly, Mpt5p and Ssdlp also conform to this common theme, using specific targeted translational repression to prolong cell divisions and maintain potential to produce young cells. The metazoan homologs of Ssdlp still await characterization and may provide likely candidates for regulators of development and cell fate analogous to those of the PUF family. 84 References: Albig, A. R., and Decker, C. J. (2001). The target of rapamycin signaling pathway regulates mRNA turnover in the yeast Saccharomyces cerevisiae. Mol Biol Cell 12, 3428-3438. Anderson, R. M., Bitterman, K. J., Wood, J. G., Medvedik, O., and Sinclair, D. A. (2003). Nicotinamide and PNC 1 govern lifespan extension by calorie restriction in Saccharomyces cerevisiae. Nature 423, 181-185. Ausubel, F. M. (1999). Short protocols in molecular biology: a compendium of methods from Current protocols in molecular biology, 4th edn (New York: Wiley). Bandara, P. D., Flattery-O'Brien, J. A., Grant, C. M., and Dawes, I. W. (1998). Involvement of the Saccharomyces cerevisiae UTH1 gene in the oxidative-stress response. Curr Genet 34, 259268. Beyer, A., Hollunder, J., Nasheuer, H. P., and Wilhelm, T. (2004). Post-transcriptional expression regulation in the yeast Saccharomyces cerevisiae on a genomic scale. Mol Cell Proteomics 3, 1083-1092. Bitterman, K. J., Medvedik, O., and Sinclair, D. A. (2003). Longevity regulation in Saccharomyces cerevisiae: linking metabolism, genome stability, and heterochromatin. Microbiol Mol Biol Rev 67, 376-399, table of contents. Camougrand, N., Grelaud-Coq, A., Marza, E., Priault, M., Bessoule, J. J., and Manon, S. (2003). The product of the UTH1 gene, required for Bax-induced cell death in yeast, is involved in the response to rapamycin. Mol Microbiol 47, 495-506. Camougrand, N., Kissova, I., Velours, G., and Manon, S. (2004). Uthlp: a yeast mitochondrial protein at the crossroads of stress, degradation and cell death. FEMS Yeast Res 5, 133-140. Camougrand, N. M., Mouassite, M., Velours, G. M., and Guerin, M. G. (2000). The "SUN" family: UTH1, an ageing gene, is also involved in the regulation of mitochondria biogenesis in Saccharomyces cerevisiae. Arch Biochem Biophys 375, 154-160. Chang, M., French-Cornay, D., Fan, H. Y., Klein, H., Denis, C. L., and Jaehning, J. A. (1999). A complex containing RNA polymerase II, Paflp, Cdc73p, Hprlp, and Ccr4p plays a role in protein kinase C signaling. Mol Cell Biol 19, 1056-1067. Chen, T., and Kurjan, J. (1997). Saccharomyces cerevisiae Mpt5p interacts with Sst2p and plays roles in pheromone sensitivity and recovery from pheromone arrest. Mol Cell Biol 17, 34293439. Crittenden, S. L., Bernstein, D. S., Bachorik, J. L., Thompson, B. E., Gallegos, M., Petcherski, A. G., Moulder, G., Barstead, R., Wickens, M., and Kimble, J. (2002). A conserved RNA-binding protein controls germline stem cells in Caenorhabditis elegans. Nature 417, 660-663. 85 Dreyfuss, G., Kim, V. N., and Kataoka, N. (2002). Messenger-RNA-binding messages they carry. Nat Rev Mol Cell Biol 3, 195-205. proteins and the Fabrizio, P., Pletcher, S. D., Minois, N., Vaupel, J. W., and Longo, V. D. (2004). Chronological aging-independent replicative life span regulation by Msn2/Msn4 and Sod2 in Saccharomyces cerevisiae. FEBS Lett 557, 136-142. Forbes, A., and Lehmann, R. (1998). Nanos and Pumilio have critical roles in the development and function of Drosophila germline stem cells. Development 125, 679-690. Gerber, A. P., Herschlag, D., and Brown, P. 0. (2004). Extensive association of functionally and cytotopically related mRNAs with Puf family RNA-binding proteins in yeast. PLoS Biol 2, E79. Ghaemmaghami, S., Huh, W. K., Bower, K., Howson, R. W., Belle, A., Dephoure, N., O'Shea, E. K., and Weissman, J. S. (2003). Global analysis of protein expression in yeast. Nature 425, 737741. Gietz, D., St Jean, A., Woods, R. A., and Schiestl, R. H. (1992). Improved method for high efficiency transformation of intact yeast cells. Nucleic Acids Res 20, 1425. Gotta, M., Strahl-Bolsinger, S., Renauld, H., Laroche, T., Kennedy, B. K., Grunstein, M., and Gasser, S. M. (1997). Localization of Sir2p: the nucleolus as a compartment for silent information regulators. Embo J 16, 3243-3255. Guarente, L., and Picard, F. (2005). Calorie restriction--the SIR2 connection. Cell 120, 473-482. Ho, Y., Gruhler, A., Heilbut, A., Bader, G. D., Moore, L., Adams, S. L., Millar, A., Taylor, P., Bennett, K., Boutilier, K., et al. (2002). Systematic identification of protein complexes in Saccharomyces cerevisiae by mass spectrometry. Nature 415, 180-183. Igual, J. C., Johnson, A. L., and Johnston, L. H. (1996). Coordinated regulation of gene expression by the cell cycle transcription factor Swi4 and the protein kinase C MAP kinase pathway for yeast cell integrity. Embo J 15, 5001-5013. Jan, E., Motzny, C. K., Graves, L. E., and Goodwin, E. B. (1999). The STAR protein, GLD-1, is a translational regulator of sexual identity in Caenorhabditis elegans. Embo J 18, 258-269. Johnstone, O., and Lasko, P. (2001). Translational regulation and RNA localization in Drosophila oocytes and embryos. Annu Rev Genet 35, 365-406. Kaeberlein, M., Andalis, A. A., Liszt, G. B., Fink, G. R., and Guarente, L. (2004a). Saccharomyces cerevisiae SSD1-V confers longevity by a Sir2p-independent mechanism. Genetics 166, 1661-1672. 86 Kaeberlein, M., and Guarente, L. (2002). Saccharomyces cerevisiae MPT5 and SSD1 function in parallel pathways to promote cell wall integrity. Genetics 160, 83-95. Kaeberlein, M., Kirkland, K. T., Fields, S., and Kennedy, B. K. (2004b). Sir2-independent life span extension by calorie restriction in yeast. PLoS Biol 2, E296. Kaeberlein, M., McVey, M., and Guarente, L. (1999). The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev 13, 2570-2580. Kennedy, B. K., Austriaco, N. R., Jr., Zhang, J., and Guarente, L. (1995). Mutation in the silencing gene SIR4 can delay aging in S. cerevisiae. Cell 80, 485-496. Kennedy, B. K., Gotta, M., Sinclair, D. A., Mills, K., McNabb, D. S., Murthy, M., Pak, S. M., Laroche, T., Gasser, S. M., and Guarente, L. (1997). Redistribution of silencing proteins from telomeres to the nucleolus is associated with extension of life span in S. cerevisiae. Cell 89, 381391. Kenyon, C. (2005). The plasticity of aging: insights from long-lived mutants. Cell 120, 449-460. Kissova, I., Deffieu, M., Manon, S., and Camougrand, N. (2004). Uthlp is involved in the autophagic degradation of mitochondria. J Biol Chem 279, 39068-39074. Kuersten, S., and Goodwin, E. B. (2003). The power of the 3' UTR: translational control and development. Nat Rev Genet 4, 626-637. Kushnirov, V. V. (2000). Rapid and reliable protein extraction from yeast. Yeast 16, 857-860. Lee, M. H., and Schedl, T. (2001). Identification of in vivo mRNA targets of GLD-1, a maxi-KH motif containing protein required for C. elegans germ cell development. Genes Dev 15, 24082420. Lee, M. H., and Schedl, T. (2004). Translation repression by GLD- 1 protects its mRNA targets from nonsense-mediated mRNA decay in C. elegans. Genes Dev 18, 1047-1059. Lin, H., and Spradling, A. C. (1997). A novel group of pumilio mutations affects the asymmetric division of germline stem cells in the Drosophila ovary. Development 124, 2463-2476. Lin, S. J., Defossez, P. A., and Guarente, L. (2000). Requirement of NAD and SIR2 for life-span extension by calorie restriction in Saccharomyces cerevisiae. Science 289, 2126-2128. Lin, S. J., Ford, E., Haigis, M., Liszt, G., and Guarente, L. (2004). Calorie restriction extends yeast life span by lowering the level of NADH. Genes Dev 18, 12-16. 87 Lin, S. J., Kaeberlein, M., Andalis, A. A., Sturtz, L. A., Defossez, P. A., Culotta, V. C., Fink, G. R., and Guarente, L. (2002). Calorie restriction extends Saccharomyces cerevisiae lifespan by increasing respiration. Nature 418, 344-348. Longtine, M. S., McKenzie, A., 3rd, Demarini, D. J., Shah, N. G., Wach, A., Brachat, A., Philippsen, P., and Pringle, J. R. (1998). Additional modules for versatile and economical PCRbased gene deletion and modification in Saccharomyces cerevisiae. Yeast 14, 953-961. Madeo, F., Herker, E., Maldener, C., Wissing, S., Lachelt, S., Herlan, M., Fehr, M., Lauber, K., Sigrist, S. J., Wesselborg, S., and Frohlich, K. U. (2002). A caspase-related protease regulates apoptosis in yeast. Mol Cell 9, 911-917. Maniatis, T., and Reed, R. (2002). An extensive network of coupling among gene expression machines. Nature 416, 499-506. Marin, V. A., and Evans, T. C. (2003). Translational repression of a C. elegans Notch mRNA by the STAR/KH domain protein GLD-1. Development 130, 2623-2632. Mazumder, B., Seshadri, V., and Fox, P. L. (2003). Translational control by the 3'-UTR: the ends specify the means. Trends Biochem Sci 28, 91-98. Mills, K. D., Sinclair, D. A., and Guarente, L. (1999). MEC 1-dependent redistribution of the Sir3 silencing protein from telomeres to DNA double-strand breaks. Cell 97, 609-620. Mootz, D., Ho, D. M., and Hunter, C. P. (2004). The STAR/Maxi-KH domain protein GLD-1 mediates a developmental switch in the translational control of C. elegans PAL-1. Development 131, 3263-3272. Mouassite, M., Camougrand, N., Schwob, E., Demaison, G., Laclau, M., and Guerin, M. (2000a). The 'SUN' family: yeast SUN4/SCW3 is involved in cell septation. Yeast 16, 905-919. Mouassite, M., Guerin, M. G., and Camougrand, N. M. (2000b). The SUN family of Saccharomyces cerevisiae: the double knock-out of UTH1 and SIM1 promotes defects in nucleus migration and increased drug sensitivity. FEMS Microbiol Lett 182, 137-141. Parisi, M., and Lin, H. (2000). Translational repression: a duet of Nanos and Pumilio. CuffrrBiol 10, R81-83. Park, P. U., Defossez, P. A., and Guarente, L. (1999). Effects of mutations in DNA repair genes on formation of ribosomal DNA circles and life span in Saccharomyces cerevisiae. Mol Cell Biol 19, 3848-3856. Racki, W. J., Becam, A. M., Nasr, F., and Herbert, C. J. (2000). Cbklp, a protein similar to the human myotonic dystrophy kinase, is essential for normal morphogenesis in Saccharomyces cerevisiae. Embo J 19, 4524-4532. 88 Reinke, A., Anderson, S., McCaffery, J. M., Yates, J., 3rd, Aronova, S., Chu, S., Fairclough, S., Iverson, C., Wedaman, K. P., and Powers, T. (2004). TOR complex 1 includes a novel component, Tco89p (YPL180w), and cooperates with Ssdlp to maintain cellular integrity in Saccharomyces cerevisiae. J Biol Chem 279, 14752-14762. Rogina, B., and Helfand, S. L. (2004). Sir2 mediates longevity in the fly through a pathway related to calorie restriction. Proc Natl Acad Sci U S A 101, 15998-16003. Schumacher, B., Hanazawa, M., Lee, M. H., Nayak, S., Volkmann, K., Hofmann, E. R., Hengartner, M., Schedl, T., and Gartner, A. (2005). Translational repression of C. elegans p53 by GLD-1 regulates DNA damage-induced apoptosis. Cell 120, 357-368. Sinclair, D., Mills, K., and Guarente, L. (1998). Aging in Saccharomyces cerevisiae. Annu Rev Microbiol 52, 533-560. Sinclair, D. A., and Guarente, L. (1997). Extrachromosomal rDNA circles--a cause of aging in yeast. Cell 91, 1033-1042. Sutton, A., Immanuel, D., and Arndt, K. T. (1991). The SIT4 protein phosphatase functions in late G1 for progression into S phase. Mol Cell Biol 11, 2133-2148. Thore, S., Mauxion, F., Seraphin, B., and Suck, D. (2003). X-ray structure and activity of the yeast Pop2 protein: a nuclease subunit of the mRNA deadenylase complex. EMBO Rep 4, 11501155. Tissenbaum, H. A., and Guarente, L. (2001). Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature 410, 227-230. Tucker, M., Staples, R. R., Valencia-Sanchez, M. A., Muhlrad, D., and Parker, R. (2002). Ccr4p is the catalytic subunit of a Ccr4p/Pop2p/Notp mRNA deadenylase complex in Saccharomyces cerevisiae. Embo J 21, 1427-1436. Uesono, Y., Toh-e, A., and Kikuchi, Y. (1997). Ssdlp of Saccharomyces cerevisiae associates with RNA. J Biol Chem 272, 16103-16109. Van Zandycke, S. M., Sohier, P. J., and Smart, K. A. (2002). The impact of catalase expression on the replicative lifespan of Saccharomyces cerevisiae. Mech Ageing Dev 123, 365-373. Varshavsky, A. (1996). The N-end rule: functions, mysteries, uses. Proc Natl Acad Sci U S A 93, 12142-12149. Velours, G., Boucheron, C., Manon, S., and Camougrand, N. (2002). Dual cell wall/mitochondria localization of the 'SUN' family proteins. FEMS Microbiol Lett 207,165-172. 89 Wheeler, R. T., Kupiec, M., Magnelli, P., Abeijon, C., and Fink, G. R. (2003). A Saccharomyces cerevisiae mutant with increased virulence. Proc Natl Acad Sci U S A 100, 2766-2770. Wood, J. G., Rogina, B., Lavu, S., Howitz, K., Helfand, S. L., Tatar, M., and Sinclair, D. (2004). Sirtuin activators mimic caloric restriction and delay ageing in metazoans. Nature 430, 686-689. Xu, L., Paulsen, J., Yoo, Y., Goodwin, E. B., and Strome, S. (2001). Caenorhabditis elegans MES-3 is a target of GLD-1 and functions epigenetically in germline development. Genetics 159, 1007-1017. Zamore, P. D., Williamson, J. R., and Lehmann, R. (1997). The Pumilio protein binds RNA through a conserved domain that defines a new class of RNA-binding proteins. Rna 3, 14211433. Zhang, B., Gallegos, M., Puoti, A., Durkin, E., Fields, S., Kimble, J., and Wickens, M. P. (1997). A conserved RNA-binding protein that regulates sexual fates in the C. elegans hermaphrodite germ line. Nature 390, 477-484. 90 TABLES AND FIGURES Table 1. Yeast strains used in this study. Strain Relevant Genotype W303R MKY1324 MKY641 MKY643 MKD133 PSY316AR MATa ADE2::RDNI ade2 his3 leu2 trpl ura3 ssdl-d2 RAD5 W303R mpt5::LEU2 W303R mpt5::LEU2 uthl::URA3 W303R mpt5::LEU2 uthl::TRPI W303R/W303R MATa/a ccr4::HIS3/CCR4 mpt5::LEU2/MPT5 uthl ::TRP/UTH1 W303R/W303R MATa/a swi4::HIS3/SWI4 mpt5::LEU2/MPT5 uthl ::TRP/UTH1 W303R/W303R MATa/a swi6::HIS3/SWI6mpt5::LEU2/MPT5 uthl ::TRP/UTH1 MATa ADE2::RDN1 ade2-101 his3-A200 leu2-3,112 lys2-801, ura3-52, ssdl-d GLY3002 GLY3003 PSY316AR SSD1-V/LEU2 PSY316AR SSD1-V/URA3 MKD 134 MKD194. GLY3004 PSY316AR uthl ::HIS3 MKY2132 PSY316AR uthl::URA3 MKD212 MKD202 MKD213 PSY316AR/PSY316AR MATa/a sit4::HIS3/SIT4 uthl ::URA3/UTH1 PSY316AR/PSY316ARMATa/a sit4::HIS3/SIT4 SSD1-V/URA3 PSY316AR/PSY316ARMATa/a sit4::HIS3/SIT4 uthl::URA3/UTH1 SSD1- GLY3101 GLY3102 GLY30 17 V/LEU2 PSY316AR p416SSD1 PSY316AR p416SSD1-TAP TERINATOR-kanMX6 PSY3 16AR SSD1- V/LEU2 UTH1 TERMINATOR::ADH1 GLY3018 PSY316AR SSD1-V/LEU2 UTHlPROMOTER::ADHl pRooTER* 91 TABLE2. Spore clone viability. Tetrads were dissected onto YPD and incubated at 30°. Spore clones that failed to form visible colonies after 4 days were examined microscopically to verify inviability. Numbers in parentheses give the total number of representative spores analyzed. - Spore Genotype mpt5 ccr4 ssdl-d UTH1 mpt5 ccr4 SSDI1-VUTHI mpt5 ccr4 ssdl-d uthl Spore Clone Percent Viability - 083(23) 83 (6) 83 (6) mpt5 swi4 ssdl-d UTH1 0 (74) mpt5 swi4 SSD1-V UTH1 84 (79) mpt5 swi4 ssdl-d uthl 30 (20) mpt5 swi6 ssdl-d UTH1 0 (65) mpt5 swi6 SSD1-V UTH1 100 (10) mpt5 swi6 ssdl-d uthl 67 (18) sit4 ssdl-d UTH1 0 (21) sit4 SSD1-V UTH1 100 (15) sit4 ssdl-d uthl 100 (14) sit4 SSD1-V uthl 100 (7) 92 Figure 1. Effect of UTH1 on cell integrity and life span. (A) 5 RLaliquots of 10-fold serial dilutions were plated onto the designated media and cultured at 30°C or 37°C for 2 days, as indicated. (B) Life spans were determined for PSY316(*), PSY316uthl (), PSY316mpt5 (A), and PSY316mpt5 uthl (X). Mean life spans and number of cells analyzed were: PSY316 21.9 (n=40), PSY316 uthl 26.4 (n=40), PSY316 mpt5 8.9 (n=40), and PSY316 mpt5 uthl 19.2 (40). (C) Life spans were determined for PSY316 (*), PSY316 uthl (), PSY316 SSD1- V (A), and PSY316uthl SSD1-V (X). Mean life spans and number of cells analyzed were: PSY316 19.4 (n=40), PSY316 uthl 23.6 (n=40), PSY316 SSD1-V 34.0 (n=40), and PSY316 mpt5 SSD1-V 24.4 (40). 93 A 37°C 0.025 mglmL CFW 0.05 mglmL CFW c B __ PSY376 (79.4) __ PSY376 (27.9) -+-PSY376 uth7 (26.4) -.-PSY376 mptS (8.9) -M-PSY376 mptS uth7 (79.2) 0.8 -+-PSY376uth7 0.8 -.-PSY376 (23.6) SSD7-V (34.0) -M-PSY316 uthl SSD1-V (24.4) CI> ~ 0.6 0.6 :> c: o "B ~ 0.4 0.4 0.2 0.2 o o o 10 30 20 40 so o 20 40 Generation Generation FIGURE 1 94 60 80 Figure 2. SSD1-Vand UTH1 both regulate resistance to oxidative damage and growth at high temperature. 10 pL aliquots of 10-fold serial dilutions were plated onto the designated media and cultured at 30°C or 40°C for 2 days, as indicated. H20 2 indicates YPD prepared with hydrogen peroxide. PQ indicates YPD prepared with paraquat. 95 o 2mM == 3mM N N SmM FIGURE 2 96 Figure 3. Mutation of UTH1 suppresses the lethality caused by deletion of SIT4. (A) Strain MKD213was sporulated and tetrads were dissected onto YPD. Growth after 4 days at 30° is shown. Genotypes are shown for sit4 ssdl-d UTH1 (0), sit4 ssdl-d uthl (A), sit4 SSD1-V UTH (), and sit4 SSD1-V uthl (<>)spore clones. (B) Spore clones were streaked onto YPD and cultured at 30°C. The pictures show cell growth after 3 days. 97 A B FIGURE 3 98 Figure 4. Effects of SSD1-V on Uthlp and UTH1 mRNA levels. Cells of the indicated genotypes were grown in YPD and harvested at early log phase (OD600=0.6), mid-log phase (OD600=2.O),or late-log phase (OD600=4.5). (A) Protein extracts were prepared and analyzed by SDS-PAGEfollowed by Western blotting. Blots were probed with rabbit polyclonal anti-Uthlp, then stripped and reprobed with mouse monoclonal antibody specific for beta- tubulin. (B) Total RNA was extracted from the indicated samples and subjected to Northern blotting. Blots were probed with 3 2 P-labelledcDNA corresponding to UTH1 (top panel) or ADH1 (bottom panel). 99 A. Uthl p Tub2p ,~ early log late log B. UTHl i x > ~ ' ?~~''''c< :'" :, ~ "- 'i1 early-log ~>- •. ¥ ADHl 1' late-log FIGURE 4 100 Figure 5. SSD1 does not affect Uthlp decay kinetics. Early-log phase cells harboring the either the ssdl-d or SSD1-V allele were treated with 100 Rg/mL cycloheximide to inhibit protein synthesis. Samples were harvested at the given time points, and extracts were prepared by boiling cells in modified SDS sample buffer (Kushnirov, 2000). (A) Rate of Uthlp decay was assessed by Western blotting using rabbit antisera specific to Uthlp. Decay of Tub2p, included here as a control, was undetectable over this time period. (B) Over 120 minute time course, a fraction of Uthlp resists degradation in both SSD1-V and ssdl-d cells. 101 Duration cycloheximide treatment: Uthlp ssdl-d SSD1-V ].It;II.'.'m.'.' '. '1..'1 ..'1'.:.1J.,......•............... :~.\ '.I.' ''.'.~.'..'.. .\.' .""""""""""""," /.",l i.e,. x.•.•.•..•..• Duration cycloheximide treatment: Uthlp SSD1-V ssdl-d FIGURE 5 102 Figure 6. Ssdlp specifically associates with UTH1 mRNA in vivo. Yeast cells expressing either Ssdlp-TAP or Ssdlp (mock IP) were grown to OD600=0.6 and harvested. Cell extracts were prepared by glass bead lysis and immunoprecipitated using rabbit IgG agarose beads. RNA purified from the supernatant and pellet fractions was subjected to RT-PCR using primers specific to UTH1, ACT1, CBK1, and NCA3. RT-PCR reactions were performed on two different starting volumes of RNA input to confirm that amplification occurred within a linear range. 103 UTHl input 1 5 IP 1 IJL input 5 Mock IP UTHl ACTl 1 input IP input 5 1 1 5 IP 1 5 5 JlL input .5 IJL input SSD1-TAP IP CBKl NCA3 1 input IP input 5 1 1 5 SSD1-TAP IP FIGURE 6 104 IP 5 1 Figure 7. Repression of Uthlp by Ssdlp is necessary for growth at high temperature and requires the UTH1 5' UTR. (A) Two mutations were constructed: the tUTHI::tADHI replacement substituted the downstream terminator sequence and 3'UTR of ADHI for that of UTH. The pUTHI :.pADHI * substitution replaced the promoter and 5'UTR of UTHI with a truncated, weakened promoter from ADHI. (A) These mutants were assayed for growth at 41°C and (B) the effects of each mutation on Uthlp expression was compared in ssdl-d and SSDI-V cells. 105 B. A. YPD 30' C ssdl-d uthl tUTHl ::tADH1SSD1-V pUTHl ::pADHl *SSD1-V FIGURE 7 106 Chapter3 Mouse Sir2 Homolog SIRT6 is a Nuclear ADP-Ribosyltransferase This chapter was published in the Journal of Biological Chemistry, Volume 280, Issue 22, pages 21313-20. The authors were Gregory Liszt, Ethan Ford, Martin Kurtev, and Leonard Guarente. I contributed all figures except Figure 2a and Figure 4a. I also wrote the manuscript. 107 SUMMARY Members of the Sir2 family of NAD-dependent protein deacetylases regulate diverse cellular processes including aging, gene silencing, and cellular differentiation. Here, we report that the distant mammalian Sir2 homolog SIRT6 is a broadly expressed, predominantly nuclear protein. Northern analysis of embryonic samples and multiple adult tissues reveals mSIRT6 mRNA peaks at day El 1I,persisting into adulthood in all eight tissues examined. At the protein level, mSIRT6 is readily detectable in the same eight tissue types, with the highest levels in muscle, brain and heart. Subcellular localization studies using both C- and N-terminal GFP fusion proteins show mSIRT6GFP to be a predominantly nuclear protein. Indirect immunofluorescence using antibodies to two different mSIRT6 epitopes confirms that endogenous mSIRT6 is also largely nuclear. Consistent with previous findings, we do not observe any NAD+ dependent protein deacetylase activity in preparations of mSIRT6. However, purified recombinant mSIRT6 does catalyze the robust transfer of radiolabel from 32 P-NAD to mSIRT6. Two highly conserved residues within the catalytic core of the protein are required for this reaction. This reaction is most likely mono-ADP ribosylation, as only the modified form of the protein is recognized by an antibody specific mono-ADP-ribose. Surprisingly, we observe that the catalytic mechanism of this reaction is intra-molecular, with individual molecules of mSIRT6 directing their own modification. These results provide the first characterization of a Sir2 protein from phylogenetic class IV. 108 INTRODUCTION Members of the Sir2 family of enzymes, conserved from bacteria to man, utilize oxidized nicotinamide adenine dinucleotide (NAD+) as a cosubstrate in the deacetylation of a wide variety of proteins, thereby regulating diverse processes including aging, genomic silencing, recombination, cell fate, and metabolism (Blander and Guarente, 2004). In yeast, the archetypal Sir2p localizes to chromatin sites at the telomeres, rDNA, and silent mating type loci, facilitating genomic silencing through the deacetylation of lysines within the histone H3 and H4 tails (Gasser and Cockell, 2001). In this context, yeast Sir2p promotes mother cell longevity by repressing rDNA recombination and formation of toxic extra-chromosomal rDNA circles (Bitterman et al., 2003; Hekimi and Guarente, 2003). Overexpression of Sir2p extends yeast lifespan (Kaeberlein et al., 1999), and Sir2p activity increases in response to caloric restriction (Lin et al., 2004), contributing to the resultant extension of life span (Lin et al., 2000). Intriguingly, Sir2 orthologues in worm and fly also promote longevity, though probably by different molecular mechanisms (Rogina and Helfand, 2004; Tissenbaum and Guarente, 2001; Wood et al., 2004). Mammalian genomes contain seven Sir2 homologs, termed sirtuins (SIRTs). Of these, SIRT1, orthologous to the yeast Sir2, is the best characterized and has the broadest substrate specificity (North and Verdin, 2004). A nuclear protein, SIRT1 deacetylates several substrates in vivo, including MyoD, p53, and FOXO transcription factors, thereby affecting cell differentiation and survival under stress (Brunet et al., 2004; Motta et al., 2004; North and Verdin, 2004). Recently, mouse SIRT1 was reported to promote the mobilization of fatty acids in white adipose tissue by repressing PPARy, 109 linking Sir2 proteins to the physiology of caloric restriction in mammals (Picard et al., 2004). Of the remaining SIRTs (SIRT2-7), an in vivo substrate has only been identified for the cytoplasmic SIRT2 (North et al., 2003). This substrate, -tubulin, is specifically deacetylated by SIRT2 (North et al., 2003), although the biological consequences of this reaction are unclear. Yeast Sir2 and its orthologues, including mouse SIRT1 and bacterial CobB, catalyze the tightly-coupled cleavage of NAD+ and protein deacetylation, producing nicotinamide and 2-O-acetyl-ADP-ribose reaction products (Sauve and Schramm, 2004). While Sir2 proteins are generally thought to be NAD-dependent protein deacetylases, most also display a less robust mono-ADP-ribosyltransferase activity in vitro (Frye, 1999; Tanner et al., 2000; Tanny et al., 1999; Tanny and Moazed, 2001). Mutations in phylogenetically conserved residues within the catalytic core of Sir2 and SIRT1 have failed to separate the two enzymatic activities (Armstrong et al., 2002). The initial observation that ADP-ribosylation by Sir2 is stronger on acetylated substrates (Imai et al., 2000), combined with subsequent mechanistic studies (Tanner et al., 2000; Tanny and Moazed, 2001), has led to the conclusion that the deacetylase and phosphoribosyltransferase activites of Sir2 proteins are coupled (Sauve and Schramm, 2004). Using sequence similarity, eukaryotic Sir2 genes have been divided into four broad phylogenetic classes (Frye, 2000), known as classes I-IV (Figure 1). SIRT1, SIRT2, and SIRT3 are class I, SIRT4 is class II, and SIRT5 falls within the predominantly prokaryotic class III. Finally, mammalian SIRT6 and SIRT7, are class IV 110 sirtuins (Frye, 2000). In vitro studies indicate that human SIRT1, 2, 3, and 5 possess NAD-dependent histone deacetylase activity (North et al., 2003), whereas SIRT4, 6, and 7 fail to deacetylate 3 H-labeled acetylated histone H4 peptide (North et al., 2003). The lack of detectable deacetylase activity in these SIRTs may result from their specificity for targets other than those tested, or it may indicate an enzymatic activity other than deacetylation inherent in these sirtuins. Mono-ADP-ribosylation is emerging as a common mechanism of reversible protein modification within mammalian cells. Originally described as the acting mechanism for specific bacterial toxins, ADP-ribosylation is typically performed by separate families of intra- and extracellular enzymes in vertebrates (Corda and Di Girolamo, 2003). Known targets of the intracellular class of enzymes include molecular chaperone GRP78/BiP, translational elongation factor 2, and -subunit of heterotrimeric G-proteins (Corda and Di Girolamo, 2003). Extracellular mammalian ADP- ribosyltransferases, generally found in cells of the immune system, modify substrates important for the immune response, as well as integrin ca7 and the antimicrobial peptide defensin (Corda and Di Girolamo, 2003). In most known cases, ADP-ribosylation of arginine residues results in reversible inactivation of the substrate protein (Riese et al., 2002), although this modification may enhance certain enzymatic functions within the substrate (Weng et al., 1999). In this report, we characterize the tissue-specific expression, subcellular localization, and in vitro enzymatic activity of mouse SIRT6. This investigation, the first detailed study of a Class IV sirtuin, reveals mouse SIRT6 to be a widely expressed, predominantly nuclear protein with a robust auto-ADP-ribosyltransferase activity. 111 MATERIALS AND METHODS Multiple sequence alignments Sequences of mouse proteins SIRT1, SIRT4, SIRT5, and SIRT6, as well as yeast Sir2p were obtained from the Proteome (https://proteome.incyte.com/control/tools/proteome). BioKnowledge® Library Core domains were aligned by ClustalW using software from the DNASTAR package. PlasmidconstructionIMAGE clone 1259892, which contains the mouse SIRT6 cDNA, was obtained from ATCC. The mSIRT6 coding sequence was amplified by PCR and cloned into the EcoRI and BamHI sites of pEGFP-N1 and the BglII and EcoRI sites of pEGFP-C2, creating pmSIRT6-EGFP and pEGFP-mSIRT6. respectively. Similarly, the mSIRT6 coding sequence was amplified by PCR and cloned into the NheI and BamHI sites of pET28a (Novagen) and pET-GST (gift from Robert Marciniak,) creating pET28a-6xHis-SIRT6 and pET-GST-SIRT6. pET28a-6xHis-mSIRT6-S56A and pET28a-6his-SIRT6-H133Y were created using the Strataene QuikChange site directed mutagenesis kit according to the manufacturer's instructions. PET-GST-SIRT6 (274-355) was created by amplifying the portion of the mSIRT6 cDNA corresponding to amino acids 274-355 and ligating into pET-GST cut with NheI and BamHI. Multiple tissue Northern blots Probe containing the SIRT6 open reading frame was created using the PrimeIt Random Primer Labeling kit (Stratagene) according to the manufacturer's instructions. This 32 p_ labeled fragment was used to probe Stratagene Multiple Tissue Northern Blots 112 (Stratagene) according to the manufacturer's instructions. The blots were then stripped and reprobed with the actin probe included in the kit. AntibodyproductionpET28a-6xHis-mSIRT6 and pET-GST-mSIRT6 (274-355) were transformed into BL21(DE3)-Codon Plus-RP cells (Stratagene),and inoculated into 4 or 2 liter cultures of LB (50 mg/ml kanamycin, 35 mg/ml chloramphenicol), respectively. The cells were grown at 37 ° C to an optical density of 0.7. Protein expression was induced by the addition of IPTG to a final concentration of 0.4 mM for 2 h, and the cells were harvested by centrifugation. The cells containing the pET28a-6xHis-mSIRT6 plasmid were resuspended in lysis buffer (0.1 M NaH2 PO4, 10mM Tris, 8 M urea, pH 8.0) and lysed by freezing in liquid N2 . The lysate was centrifuged at 19,000 rpm in a Sorvall SS-34 rotor. The supernatant was incubated with 4 ml of Ni-NTA agarose (Stratagene) for 3 h at 4 ° C, loaded onto a column, and washed extensively with wash buffer (0.1 M NaH2PO4, 10 mM Tris, 8 M urea, pH 6.4). The purified protein was eluted with elution buffer (0.1 M NaH 2PO4 , 10 mM Tris, 8M urea, pH 3.0) in 1.5 ml fractions. The pH was neutralized by the immediate addition of 100 ml 1 M Tris base. The fractions containing protein were pooled and dialyzed against PBS. This protein was sent to Covance Reseach Bioproducts for antibody production. The cells containing the pET-GST-SIRT6 (274-355) plasmid were resuspended in PBSTG (PBS, 0.1% Tween-20, 10% glycerol, 1 mM DTT) and 0.1 g lysozyme, incubated at 37° C for 20 min, and frozen in liquid N2 . The lysis was completed by sonication and the lysate was centrifuged at 19,000 rpm for 30 min. The supernatant was 113 incubated with 4 ml GST-agarose (Sigma) for 3 h at 4° C and loaded onto a column. The column was washed extensively with PBSTG and the protein was eluted with elution buffer (50 mM Tris, 10 mM glutathione, pH 7.5). Fractions containing protein were dialyzed against PBS and sent to Covance Reseach Bioproducts for antibody production. The antibodies were affinity purified as follows. Aliquots of the proteins used to generate the antibodies were attached tol ml NHS-activated sepharose columns (Amersham) according to the manufacturer's instructions. 10 ml of the final bleed of each antibody was combined with 10 ml antibody wash buffer (20 mM Hepes, pH 7.5, 150 mM NaCl, 0.1% Tween-20) and applied to the column. Each column was washed extensively with antibody wash buffer and the antibody was eluted with 100 mM glycine pH 2.5. The pH was neutralized by the addition of 1/10 volume of 1 M Hepes-KOH, pH 7.5. The antibodies were then subjected to a second round of affinity purification, but this time the anti-mSIRT6 (full length) antibody was applied to the SIRT6 (274-355) column and the anti-mSIRT6(274-355) antibody was applied to the SIRT6(full length) column. Western blot analysis Homogenized mouse tissues from four mice of the FVB genetic background were obtained as a generous gift from L. Bordone, MIT. From these samples, SDS-solubilized proteins were prepared (Kain et al., 1994), equal amounts of which were used for Western analysis. For Western detection, samples were resolved on 4-15% Tris-HCl gradient gels (Bio-Rad), transferred to PVDF membrane, and blocked using 4% nonfat dry milk in PBS with 0.1% Tween-20 (PBS-T). Primary anti-mSIRT6 antibody was diluted 1:400 in 114 PBS-T before incubation with membranes. Rabbit anti-ADP-ribose antibody (Kain et al., 1994), a generous gift from H. Hilz, was used at a 1:100 dilution in PBS-T. Anti-actin C4 (Sigma) was used as a loading control (1:5,000) in multiple tissue Western blots. Secondary antibodies were goat anti-rabbit or sheep anti-mouse horseradish peroxidaseconjugated antibodies (1:10,000). Chemiluminescent detection was achieved by incubation with the ECL reagent (Amersham Biosciences). Immunofluorescenceand GFP microscopy NIH 3T3 cells were grown in Dulbecco's Modified Eagle's Medium for 48 hours on gelatin-treated 18mm circular glass coverslips in 6-well plates. Cells were fixed for 10 minutes with 4% paraformaldehyde and washed twice with PBS, then permeablized by 3 minute treatment with 0.2% Triton X-100. Coverslips were gently washed twice with PBS and blocked with 10% BSA for 15 minutes, transferred to a rack, and washed in PBS. Affinity purified primary antibody was added at a 1:100 dilution. Staining proceeded for 45 minutes, and cells were then washed twice for 5 minutes in fresh PBS. FITC-conjugated chicken anti-rabbit IgG was diluted in 1% BSA and incubated with coverslips for 45 minutes. Coverslips were mounted with VectaShield mounting media containing 4',6-diamidino-2-phenyindole, dilactate (Vector, Burlingame, CA). Microscopy was performed on a Nikon Eclipse E500 fluorescence microscope. Digital images were obtained using a SPOT RT CCD camera and software using identical exposure times for comparable samples. NIH 3T3 cells were transfected with 5 g of each GFP construct using FuGene 6 (Roche) according to manufacturer's instructions. Cells were grown for 48 hours on glass coverslips. Samples were fixed, DAPI-stained, and visualized as described above. 115 Purificationof recombinantSIRT6protein BL21(DE3)-Codon Plus-RP E. coli (Stratagene) were transformed with the mSIRT6 expression plasmids. Protein purifications were performed essentially as described for antibody purification. Briefly, fusion protein expression was induced in 0.4 mM IPTG at 37° C for 1 h. The induced 6 x His-tagged proteins were purified with Ni-NTA agarose beads (Qiagen) under native conditions. The control eluate was prepared from a bacterial clone carrying pET28a vector alone. GST-tagged proteins were purified by a similar method, with affinity purification accomplished with glutathione agarose beads (Pierce). Control eluate for GST-tagged proteins was prepared from bacteria carrying pET28-GST. Aliquots of recombinant proteins were stored at -70 ° C. ADP-ribosylationassaysand detectionof ADP-ribosylatedproteins ADP ribosylation assays were performed essentially as described (Imai et al., 2000), with minor modifications. Reactions were performed in 50 [tL total volume containing 5 g recombinant SIRT6 in 50 mM Tris-HCl, pH 8.0 (at 22°C), 150 mM NaCl, 10 mM DTT, 1 uM unlabeled NAD, and 8 Ci of NAD 5'-[- 32 P]triphosphate (800 Ci mmol-1, Amersham Pharmacia). Reactions also contained 2.5 pg core histones as indicated. For reactions containing both His-tagged and GST-tagged recombinant SIRT6 protein, 2.5 ug of each protein were added. Prior to SDS-PAGE analysis, all reactions were purified from unincorporated 32 P-NADusing Micro Bio-Spin chromatography columns (Bio-Rad) with a 6,000 Da exclusion limit. Gels were transferred to PVDF membrane before autoradiography and subsequent immunoblot analysis. Blots were stained with Coomassie, and the Coomassie-stained bands and Western signals were aligned to verify that the radioactive bands were SIRT6. 116 Quantitation of ADP-ribosylationADP-ribosylation modifications. reactions were performed as above, but with the following Total reaction volume was 40 FL, including approximately 0.5 protein, 0.8 L of 6.25 gM 32 P-NAD,and 4 g gL 10 gM NAD. Purified reaction products were separated by SDS-PAGE, and BSA standards of 5 jtg, 2 g, 0.8 g, 0.4 g, and 0.1 jg were included on the same gel. SIRT6 protein concentrations in each lane were determined by comparison to these standards upon Coomassie staining. Gel was analyzed by autoradiography, and radioactive SIRT6 bands were excised. Radioactivity was determined by scintillation counting on a Beckman LS 6500 Liquid Scintillation Counter. A standard curve relating NAD quantity to radioactivity was generated by scintillation counting of known pmol quantities of 32 P-NADdiluted to reflect the ratio of unlabeled to labeled NAD in the original reaction. This curve was used to estimate pmol NAD incorporated in SIRT6 bands. For each band, pmol of NAD was divided by pmol of SIRT6. This final value reflects the molar ratio of total incorporated ADP-ribose to SIRT6. RESULTS mSIRT6 expression in mice and embryos To determine the expression pattern of mSIRT6, we analyzed RNA extracts from eight adult mouse tissues by Northern blotting (Fig. 2a). Using a cDNA probe corresponding to the 5' region of mSIRT6, expression of mSIRT6 mRNA was observed in every tissue type tested (Fig. 2a). Confirming the predictions of sequence analysis 117 (Frye, 2000), mSIRT6 was detected as a single 1.0 Kb transcript, suggesting that this gene is not found in alternate splice forms. When normalized to actin, the highest levels of mSIRT6 mRNA were seen in the brain, heart, and liver, with the lowest expression level observed in skeletal muscle. To evaluate mSIRT6 levels during development, we probed Northern blots of RNA from four mouse embryonic stages from E7 to E17 (Fig. 2b). Mouse SIRT6 transcript was readily detectable in all embryonic samples, reaching a peak at E 1. Having established the prevalence of mSIRT6 message in embryonic and adult tissues, we wished to determine the distribution of mSIRT6 protein. Affinity-purified antiserum specific to mSIRT6 was used to probe Western blots of proteins isolated from eight adult murine tissue types (Fig. 2c). For each tissue type, samples from four individual animals were obtained and analyzed by immunoblot. As illustrated in Figure 2c, mSIRT6 protein was detectable in all tissue types as a single 40 KDa band. As expected from Northern analysis, levels were high in brain, liver, and heart when normalized to actin protein levels. Surprisingly, mSIRT6 protein was most strongly expressed in muscle (Fig. 2c), despite the relative paucity of its transcript in this tissue. This observation might indicate increased mSIRT6 transcript stability in muscle, or an enhanced rate of translation in this tissue type relative to others tested. Visualization of mSIRT6-GFP in nuclei In order to observe the subcellular localization of mSIRT6, we engineered two GFP-fusion expression vectors under the control of the CMV promoter. Both a Cterminal (mSIRT6-GFP) and an N-terminal (GFP-mSIRT6) fusion construct were transfected into NIH 3T3 cells grown on glass coverslips. Samples were cultured for 48 118 hours, at which point they were fixed, washed and DAPI-stained to reveal nucleic acids. Fluorescence microscopy of cells harboring each GFP-fusion construct showed strong localization of mSIRT6-GFP to the nucleus, accompanied by diffuse cytoplasmic staining, as judged by colocalization of DAPI-stained nuclei (Fig. 3a). As expected, transfection of GFP alone caused intense fluorescence throughout the cytoplasm and nucleus. Indirect immunofluorescence of mSIRT6 To visualize endogenous levels of mSIRT6, affinity-purified rabbit antisera to two different mSIRT6 antigens were used to probe NIH 3T3 cells by indirect immunofluorescence. As a control, cells were incubated without primary antibody and subsequently stained with DAPI and Cy-3 conjugated anti-rabbit IgG. Very faint background staining was seen throughout the cytoplasm and nucleus in these control samples (Fig. 3b). Strikingly, mSIRT6 showed strong staining predominantly in the nucleus (Fig. 3b), regardless of which anti-mSIRT6 antibody was used in the primary incubation. In both cases, endogenous mSIRT6 appeared to be excluded from the nucleolus. Importantly, immunofluorescence and GFP-localization studies revealed the same predominantly nuclear localization pattern for mSIRT6. We therefore conclude that the majority of mSIRT6 is found in the nucleus, while a small fraction may be localized to the cytoplasm. mSIRT6 catalyzes the transfer of 3 2 p from a 32 P--NADdonor Of the seven mammalian Sir2 homologs, only SIRT1, SIRT2, SIRT3, and SIRT5 have been shown to have NAD+-dependent protein deacetylase activity in vitro (North et 119 al., 2003). Consistent with published reports, we did not observe any protein deacetylase activity of mouse or human SIRT6 using a variety of experimental conditions and potential substrates (data not shown). Several sirtuins also possess a mono-ADP-ribosyltransferase activity (Tanny et al., 1999). We therefore sought to test whether mSIRT6 could catalyze the transfer of radiolabel from 32 P--NAD*to histones, an assay of ADP-ribosyltransferase activity. Six- histidine-tagged human SIRT1 and mSIRT6 were expressed in E. coli and purified by chromatography. Equal amounts of hSIRT1 and mSIRT6 were incubated with 32 P-NAD+ in the presence and absence of core histones (Fig. 4a). Whereas hSIRT1 transferred label efficiently to histones, mSIRT6 failed to do so (Fig. 4a). However, mSIRT6 catalyzed the robust transfer of label to itself, regardless of whether histones were present in the reaction (Fig. 4a). Control extracts purified from E. coli containing vector alone possessed no detectable activity under these conditions (data not shown). We wished to determine if 32 p transfer stimulated by mSIRT6 required the catalytic activity of the enzyme. Two point mutations in highly-conserved residues within mSIRT6 core domain were introduced by site-directed mutagenesis (Figure 1). The histidine residue at position 133 was changed to tyrosine (mSIRT6-H133Y), a mutation previously demonstrated to abolish enzymatic activity in human and yeast Sir2 proteins (Frye, 1999; Tanny et al., 1999). In a separate construct, the phylogenetically invariant serine residue at position 56 was mutated to alanine (mSIRT6-S56A). Serine 56 is buried deep in the active site of the enzyme, and does not appear to be structurally important (Finnin et al., 2001). Both mutant proteins were expressed in E coli, purified, and incubated in the presence of 32 P-NAD+ as described above. Unincorporated NAD 120 was removed by chromatography column. Proteins were separated by SDS-PAGE, transferred to nitrocellulose membrane, and exposed to film. Whereas wild type mSIRT6 efficiently transferred 32 p radiolabel to itself, no label transfer was detected in mutants mSIRT6-S56A and mSIRT6-H133Y (Fig. 4c). Presence of mutant proteins on the membrane was verified by Western blotting using affinity-purified antibody to mSIRT6 (Fig. 4b). Based on this requirement for two catalytic core residues, we conclude that mSIRT6 catalyzes the transfer of radiolabel from 32 P-NADby an enzymatic mechanism. Labeling of proteins with 32 P-NAD+can also occur nonenzymatically by the covalent binding of NAD. Such labeling, as observed in the case of GAPDH (Zhang and Snyder, 1992), can be misinterpreted as ADP-ribosylation of specific amino acid acceptors (Itoga et al., 1997; Zhang and Snyder, 1992). We have observed low levels of nonspecific radiolabeling of proteins incubated in the presence of 32 P-NAD+. Under our reaction conditions, such modification occurs in core histones and enzymatically-inactive Sir2 point mutants, generally not labeling more than 0.001% of total protein after 1-hour incubation at 30° C (Fig. 5), as determined by scintillation counting of protein bands. In contrast, in the case of mSIRT6 we observed incorporation of a molar quantity of ADPribose equivalent to 15% of total moles of mSIRT6 present in the reaction (Fig 5). For this reason, and because of the difference in labeling efficiency between wild type and point-mutant varieties of mSIRT6, we conclude that the transfer of via a robust enzymatic mechanism. 121 32 P-radiolabeloccurs Auto-ADP-ribosylation of mSIRT6 To test the hypothesis that mSIRT6 catalyzed the mono-ADP-ribosylation of itself, we probed the reaction products from the above label-transfer reactions with an antibody specific to mono-ADP-ribose (Meyer and Hilz, 1986). This antibody recognized a band corresponding to the radiolabeled mSIRT6 following incubation with 32 P--NAD (Fig. 6a). No band was present at the corresponding position in preparations from mutant mSIRT6 (Fig. 6a), demonstrating the specificity of the antibody for the modified form of mSIRT6. This evidence strongly suggests that this modified form of mSIRT6 is mono-ADP-ribosylated. While our results do not rule out the possibility that the modification of wild type mSIRT6 occurs in E. coli before purification, the absence of signal in point mutant SIRT6 indicates that mono-ADP ribosylation depends on the catalytic activity of mSIRT6. Having confirmed mono-ADP-ribosylation as the nature of the modification of mSIRT6, we wished to ascertain whether this modification was catalyzed in an inter- or intra-molecular fashion. mSIRT6 and two different mutants (H133Y and S56A) were used in a molecular cis/trans test for enzymatic activity. GST-mSIRT6 was purified from E. coli and incubated in combination with His-tagged mutant and wild-type mSIRT6 in the presence of 32 P--NAD. Reactions were purified as before and subjected to autoradiography. Labeled GST-mSIRT6 migrated as an approximately 80 KDa band in all reactions (Fig. 6b). Six histidine-tagged mSIRT6 migrated as a strongly labeled 39 KDa band (Fig. 6b). However, this band was absent from the corresponding reactions of both His-tagged mutants, demonstrating that GST-mSIRT6 was unable to ADP-ribosylate those proteins. Coomassie staining of the gel prior to film exposure revealed similar 122 amounts of protein in all three lanes (data not shown). We therefore conclude that the auto-ADP-ribosylation reaction catalyzed by mSIRT6 is intra-molecular. DISCUSSION Here we provide the first characterization of a class IV sirtuin, mouse SIRT6. This sirtuin shows a relatively weak sequence homology to the yeast Sir2 (25%), suggesting a significant divergence in function and enzymatic activity. Consistent with published findings, we failed to detect any NAD-dependent protein deacetylase activity in preparations of mSIRT6 expressed in bacterial and mammalian cells. Surprisingly, mSIRT6 did show a robust ADP-ribosyltransferase activity in vitro, indicating that recombinant bacterial preparations indeed contained active protein. We conclude that mSIRT6 is a mono-ADP-ribosyltransferase for the following reasons. First, the purified protein can catalyze the transfer of radiolabel from 32 P--NAD, a reaction also catalyzed by other Sir2 family members (Frye, 1999; Imai et al., 2000; Tanny et al., 1999) and previously shown to be mono-ADP-ribosylation (Tanny et al., 1999). Second, this activity requires the catalytic function of mSIRT6, as two different point mutations in phylogenetically invariant residues predicted to abolish enzymatic activity eliminated the transfer of label from NAD to mSIRT6. Third, only the enzymatically-modified form of mSIRT6 is recognized by an anti-ADP ribose antibody specific to mono-ADP-ribosylated proteins. Finally, the transfer of label from 32 P--NAD to mSIRT6 is accomplished in by an intra-molecular mechanism, indicating that SIRT6 may use ADP-ribosylation as a way to auto-regulate its activity. 123 Recombinant mSIRT6 expressed in E. coli and incubated with NAD was able to incorporate a quantity of ADP-ribose equivalent to 15% of the moles mSIRT6 present in the reaction. This level of ADP-ribosylation is impressive considering the intra- molecular nature of the reaction, by which a molecule of SIRT6 is only active on a single substrate molecule. We failed to observe a band shift resulting from ADP-riboyslation, suggesting that the number of amino acid residues ADP-ribosylated on mSIRT6 is very small. Assuming a single ADP-ribosylation per molecule of SIRT6, our results indicate a strong 15% activity in our recombinant protein preparations. Mouse SIRT6 is broadly expressed at the RNA and protein levels throughout development and during adulthood in at least eight different tissue types, indicating broad expression throughout the organism. The SIRT6 protein is particularly abundant in brain and liver. Interestingly, mSIRT6 protein expression is highest in muscle, even though mRNA levels in this tissue are low relative to those of actin. This might indicate increased SIRT6 protein stability in muscle compared to other tissue types. Alternatively, the disproportionate amount of SIRT6 protein observed in muscle may result from an increased rate of translation of SIRT6 mRNA in this tissue type. Localization studies of mSIRT6 using N- and C-terminal GFP fusions as well as antibodies to two different epitopes showed that expression of mSIRT6 was largely nuclear, with only a diffuse presence in cytoplasm. Interestingly, endogenous levels of SIRT6 do not permeate certain intra-nuclear regions, probably corresponding to the nucleolus (Figure 3b). The only other mouse class IV sirtuin, mSIRT7, is localized to the nucleolus (E. F. and L. G., unpublished data). We therefore speculate that mSIRT6 and mSIRT7, despite extensive sequence homology (39%), have non-overlapping functions 124 within the cell. Like SIRT6, SIRT7 has not been shown to possess any in vitro protein deacetylase activity (North et al., 2003). Several ADP-ribosyltransferases are capable of reversible auto-modification resulting in altered enzymatic activity. Pseudomonas ExoS, a bifunctional enzyme, is a Rho GTPase-activating protein (GAP) as well as an ADP-ribosyltransferase (Riese et al., 2002). Auto-ADP-ribosylation at arginine-146, observed in vitro and in vivo, reduces GAP activity (Riese et al., 2002), suggesting a mechanism for intramolecular regulation. In mammals, auto-modification of ADP-ribosyltransferase 5 (ART5) converts the protein from NADase to transferase (Weng et al., 1999), again suggesting a mechanism for regulation of enzyme activity. We speculate that SIRT6 might regulate its own activity in vivo by ADP-ribosylation of specific residues required for activity. At this time, we have not yet identified the physiological targets of SIRT6. Unlike class I sirtuins including yeast Sir2 and SIRT1, SIRT6 appears to be highly specific in vitro, even failing to catalyze ADP-ribosylation of SIRT6 molecules in trans. The discovery of auto-ADPribosyltransferase activity in preparations of recombinant mouse SIRT6 is the first report of enzymatic activity for any class IV sirtuin. Future experiments will aim to identify physiological substrates ADP-ribosylated by SIRT6. 125 FOOTNOTES *We thank H. Hilz (Hamburg, Germany) for the ADPR antibody. We also thank Laura Bordone (Guarente lab, MIT) for her kind contribution of mouse tissue protein samples. Finally, we would like to thank Marcia Haigis and Gil Blander for scientific input and critical discussions. 1Abbreviations used in this study: mSIRT6, mouse SIRT6; GFP, Green Fluorescent Protein; DAPI, 4',6-diamidino-2-phenyindole; ADPR, ADP-ribose. REFERENCES Armstrong, C. M., Kaeberlein, M., Imai, S. I., and Guarente, L. (2002). Mutations in Saccharomyces cerevisiae gene SIR2 can have differential effects on in vivo silencing phenotypes and in vitro histone deacetylation activity. Mol Biol Cell 13, 1427-1438. Bitterman, K. J., Medvedik, O., and Sinclair, D. A. (2003). Longevity regulation in Saccharomyces cerevisiae: linking metabolism, genome stability, and heterochromatin. Microbiol Mol Biol Rev 67, 376-399, table of contents. Blander, G., and Guarente, L. (2004). The Sir2 family of protein deacetylases. Annu Rev Biochem 73,417-435. Brunet, A., Sweeney, L. B., Sturgill, J. F., Chua, K. F., Greer, P. L., Lin, Y., Tran, H., Ross, S. E., Mostoslavsky, R., Cohen, H. Y., et al. (2004). Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. Science 303, 2011-2015. Corda, D., and Di Girolamo, M. (2003). Functional aspects of protein mono-ADP-ribosylation. Embo J 22, 1953-1958. 126 Finnin, M. S., Donigian, J. R., and Pavletich, N. P. (2001). Structure of the histone deacetylase SIRT2. Nat Struct Biol 8, 621-625. Frye, R. A. (1999). Characterization of five human cDNAs with homology to the yeast SIR2 gene: Sir2-like proteins (sirtuins) metabolize NAD and may have protein ADP-ribosyltransferase activity. Biochem Biophys Res Commun 260, 273-279. Frye, R. A. (2000). Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochem Biophys Res Commun 273, 793-798. Gasser, S. M., and Cockell, M. M. (2001). The molecular biology of the SIR proteins. Gene 279, 1-16. Hekimi, S., and Guarente, L. (2003). Genetics and the specificity of the aging process. Science 299, 1351-1354. Imai, S., Armstrong, C. M., Kaeberlein, M., and Guarente, L. (2000). Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature 403, 795-800. Itoga, M., Tsuchiya, M., Ishino, H., and Shimoyama, M. (1997). Nitric oxide-induced modification of glyceraldehyde-3-phosphate dehydrogenase with NAD+ is not ADP-ribosylation. J Biochem (Tokyo) 121, 1041-1046. Kaeberlein, M., McVey, M., and Guarente, L. (1999). The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev 13, 2570-2580. Kain, S. R., Mai, K., and Sinai, P. (1994). Human multiple tissue western blots: a new immunological tool for the analysis of tissue-specific protein expression. Biotechniques 17, 982987. Lin, S. J., Defossez, P. A., and Guarente, L. (2000). Requirement of NAD and SIR2 for life-span extension by calorie restriction in Saccharomyces cerevisiae. Science 289, 2126-2128. Lin, S. J., Ford, E., Haigis, M., Liszt, G., and Guarente, L. (2004). Calorie restriction extends yeast life span by lowering the level of NADH. Genes Dev 18, 12-16. Meyer, T., and Hilz, H. (1986). Production of anti-(ADP-ribose) antibodies with the aid of a dinucleotide-pyrophosphatase-resistant hapten and their application for the detection of mono(ADP-ribosyl)atedpolypeptides. Eur J Biochem 155, 157-165. Motta, M. C., Divecha, N., Lemieux, M., Kamel, C., Chen, D., Gu, W., Bultsma, Y., McBurney, M., and Guarente, L. (2004). Mammalian SIRT1 represses forkhead transcription factors. Cell 116, 551-563. 127 North, B. J., Marshall, B. L., Borra, M. T., Denu, J. M., and Verdin, E. (2003). The human Sir2 ortholog, SIRT2, is an NAD+-dependent tubulin deacetylase. Mol Cell 11, 437-444. North, B. J., and Verdin, E. (2004). Sirtuins: Sir2-related NAD-dependent protein deacetylases. Genome Biol 5, 224. Picard, F., Kurtev, M., Chung, N., Topark-Ngarm, A., Senawong, T., Machado De Oliveira, R., Leid, M., McBurney, M. W., and Guarente, L. (2004). Sirtl promotes fat mobilization in white adipocytes by repressing PPAR-gamma. Nature 429, 771-776. Riese, M. J., Goehring, U. M., Ehrmantraut, M. E., Moss, J., Barbieri, J. T., Aktories, K., and Schmidt, G. (2002). Auto-ADP-ribosylation of Pseudomonas aeruginosa ExoS. J Biol Chem 277, 12082-12088. Rogina, B., and Helfand, S. L. (2004). Sir2 mediates longevity in the fly through a pathway related to calorie restriction. Proc Natl Acad Sci U S A 101, 15998-16003. Sauve, A. A., and Schramm, V. L. (2004). SIR2: the biochemical mechanism of NAD(+)dependent protein deacetylation and ADP-ribosyl enzyme intermediates. Curr Med Chem 11, 807-826. Tanner, K. G., Landry, J., Sternglanz, R., and Denu, J. M. (2000). Silent information regulator 2 family of NAD- dependent histone/protein deacetylases generates a unique product, -O-acetylADP-ribose. Proc Natl Acad Sci U S A 97, 14178-14182. Tanny, J. C., Dowd, G. J., Huang, J., Hilz, H., and Moazed, D. (1999). An enzymatic activity in the yeast Sir2 protein that is essential for gene silencing. Cell 99, 735-745. Tanny, J. C., and Moazed, D. (2001). Coupling of histone deacetylation to NAD breakdown by the yeast silencing protein Sir2: Evidence for acetyl transfer from substrate to an NAD breakdown product. Proc Natl Acad Sci U S A 98, 415-420. Tissenbaum, H. A., and Guarente, L. (2001). Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature 410, 227-230. Weng, B., Thompson, W. C., Kim, H. J., Levine, R. L., and Moss, J. (1999). Modification of the ADP-ribosyltransferase and NAD glycohydrolase activities of a mammalian transferase (ADPribosyltransferase 5) by auto-ADP-ribosylation. J Biol Chem 274, 31797-31803. Wood, J. G., Rogina, B., Lavu, S., Howitz, K., Helfand, S. L., Tatar, M., and Sinclair, D. (2004). Sirtuin activators mimic caloric restriction and delay ageing in metazoans. Nature 430, 686-689. Zhang, J., and Snyder, S. H. (1992). Nitric oxide stimulates auto-ADP-ribosylation of glyceraldehyde-3-phosphate dehydrogenase. Proc Natl Acad Sci U S A 89, 9382-9385. 128 129 FIGURES Figure 1. Sequence alignment of mouse sirtuin core domains. Conserved core domains of mouse SIRT1, SIRT4, SIRT5, and SIRT6 were aligned with the yeast Sir2 core using ClustalW. In the sirtuin phylogenetic tree, SIRT1 is class I, SIRT4 is class II, SIRT5 is class III, and SIRT6 is class IV. At each position of the alignment, identical residues are boxed and shaded in gray. Amino acids filled in black correspond to mSIRT6 residues targeted for mutagenesis in this study. 130 1.. ySir2p 255 mSIRTl mSIRT4 mSIRTS mSIRT6 246 L 61 Ll~M ySir2p mSIRTl mSIRT4 mSIRTS mSIRT6 ySir2p mSIRTl mSIRT4 mSIRT5 mSIRT6 R M •'.'.' .•.. •.•.. '.~ .•. ':•. IlL S1 4S S S v a.~ ..•... L A I I V F H - - I -KHLGIl,.'.?".DD DVFNYNI MHDPSVIlIYN LAV-DFPDlpD AMFDIEY KDPRP~FK TDRRPIQHID- - -I.VRSAPV~QRYWARNF - -K - I . , ',: .........•. .•. om I~P.-AUI Q'A 'Q D LAP IANMVLPPEK'IYSPLFAKEIYPGQVGWPQFSSHQ~NPAI:nw Y R REV - -HSFIKM~QMK - FQPSLC-HKFI~LSDKE ~SNWE-Rlin- L.~•... N P G •..B L A R ~:8 S K TiftM M R S K E - - D T T FEN ySir2p A •. '.~ .•.. ' .A •.R N P S - - - Q v W E ~ Y H K fE - - - - - - - - - - - - - - - A T • V: .. T M E ERG L R N Y L R N Y HWLV *.;;e .••i.'.'. 1;;;;:,MiW* S TDK~VQCI~S IWlSFLV ySir2p ATATIVTqHWNLPGERIFNKJRNLELI~L mSIRTl mSIRT4 mSIRTS mSIRT6 - S ri G - T - - - - - - V A E N Y R S - FVEEbpKIKTQYVRDT~VGTM- N M -'2 I Ie. 'RI -GLKATGR II Y K K R R E Y F l.Pjx"l G Y N N K V G ~ A A S Q G S M S I E;RI Y ~A ~- LAG ~ ~ Q ~K G~ ~1Wl.h.' .•.~.' ~'" ..~ ... ~ A lmI E 1;P,JE;1 T ~ ~ ~ ~ ~ L~T - ySir2p P P Y I ~N mSIRTl mSIRT4 mSIRT5 mSIRT6 T L R G D I F N Q V V , .E..•. R .. ~'. ~.f mSIRTl mSIRT4 mSIRT5 mSIRT6 ySir2p NIIiJIJ~.~11 ~. ~ ySir2p mSIRTl mSIRT4 mSIRTS mSIRT6 '(; S F GC M P R D KIAILG e'.'•'.'. ...•.. .. li T F F - - Q R , L Q C - S Q R '8' T iE L - - T K G A T A S .• L I " M .G .•~ .' • •.•...•.. 'K ..• ... Y K V DeE A HRVL.LNDG-EQTARRELQERFQALNisws F K T R U V Q - - - - '13 E R MG N T A E mSIRTl mSIRT4 mSIRTS mSIRT6 R V V V I -- '. A Q C E l!!] ~ R.~..$>'.C." .. R D Q .G....• R .... M. ,00 V A K T R - - - - - - AG - - - - - - - - - I C I KIArimV II.~.""I , K .:.'~'.R ~."'. V A LIP L ,y; V C L N: V S V T l ySirZp mSIRTl mSIRT4 mSIRTS mSIRT6 ySir2p DIAAMVAQKCrmW 526 mSIRTl mSIRT4 mSIRTS mSIRT6 V I 483 323 I N E L C H R LlQI ElLPLIDPRRQH KTLPEALAPHET EVMCR LMKHL@]L - - - - - - A 'NKFHKS N EQFHRA DTVNPDKVDF EN D P A I LEE LDWEDS' DRDLML S Y I,RED I LllE. ,'. C MKYDKDEV VHRRVKiA VDRELALC ADEASRTA Q~~ V~ ~D ~K ~,~ ~~..HIDE ~R L IRe - - Q D A RIP J06 271 131 S E I V N M V 'QiYYSGYRF VNcY~AAMFAPQV Q I R [!ijs S S I LTA I G N L P L A T Figure 2. Analysis of mouse SIRT6 and actin expression in adult tissues and developing embryos. (A) RNA was isolated from the indicated tissues in wild-type adult mice, resolved by electrophoresis, and subjected to Northern blotting. Blots were probed with 32 P-labeledcDNA specific to mSIRT6 (top panel) or actin (lower panel). Muscle actin found in heart and skeletal muscle samples migrated as a distinct band. (B) RNA from 7day, 11-day, 15-day, and 17-day mouse embryos was analyzed by Northern blotting as in A. (C) Protein extracts from the indicated wild-type mouse tissues were resolved by SDSPAGE, and analyzed by Western blotting using antibodies specific to mSIRT6 (top panel) or actin (lower panel). For each tissue type, samples from four different mice were included. 132 ~ 1j Vj =e A. ... .-= ~~= ~ ~ =: eu eu == - 00 t)I) == ~ -... - :s ~;0.. eu ~ ~ ~ 00 B. ~ ~ "'Cl .~ ... ~ ~ .-= ~ eu Vj "'Cl I I' ~ eu "'Cl I 'I""'C 'I""'C ..... Liver 'I""'C Actin Actin Pancreas --_._--:--~ Kidney eu SIRT6 Heart .. '••.•.... ~•.... _ SIRT6 .... Spleen 'I""'C ~ "'Cl I I' SIRT6 c. Brain ~ ~ "'Cl I tn Muscle '. I Actin Testis - ..' I SIRT6 I Actin ~~:~::-:b=--- 1----- - r 133 Figure 3. Subcellular localization of mSIRT6 by GFP and indirect immunofluorescence microscopy. (A) NIH 3T3 cells were transfected with GFP (top row), an N-terminal eGFP-mSIRT6 fusion (middle row), or a C-terminal mSIRT6-eGFP fusion (bottom row). Transfections were performed on glass coverslips in six-well plates. 48 hours posttransfection, cells were washed, fixed with paraformaldehyde, stained with DAPI, and analyzed by fluorescence microscopy. The FITC channel (left column) shows GFP fluorescence, and the DAPI channel (middle column) reveals nucleic acid localization. Images from both channels are merged in the rightmost column. For each transfection, a representative field of cells is pictured. (B) NIH 3T3 cells grown on coverslips were stained with affinity-purified rabbit polyclonal antibodies to two different mSIRT6 epitopes (full-length mSIRT6, middle row; mSIRT6 amino acids 274-355, bottom row). Samples were then stained with DAPI and Cy3-conjugated anti-rabbit IgG and visualized by fluorescence microscopy. Cy3 fluorescence is shown in the left column, DAPIstained nucleic acids are shown in the middle column, and the merged image is shown in the right column. 134 135 A. GFP DAPI FITC DAPI GFP GFP- SIRT6 SIRT6GFP B. SIRT6 (full length) SIRT6 (274- 355) 136 Figure 4. NAD+-dependentmodification of mSIRT6. (A) Recombinant hSIRTl (left) and mSIRT6 (right) were incubated in the presence of 32 P-NAD+with and without core histones. Samples were purified by gel filtration chromatography, subjected to SDSPAGE, and exposed to film for 2 hrs. (B) Wild-type and two mutant forms of mSIRT6 (S56A; H133Y) were incubated with 32 P-NAD+, purified, and resolved by SDS-PAGE. Proteins were transferred to nitrocellulose membrane and developed by autoradiography (right panel). Presence of wild type and mutant SIRT6 was confirmed by probing membrane with affinity-purified anti-mSIRT6 rabbit polyclonal antibody (left panel). 137 A. mSIRTl mSIRT6 mSIRT6 . Histones - Histones: + + 32p, c. B. < \C ~ ~ ~ l/) fIIII4 rJj < \C ~ \C == \C \C an • \C ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ I 00 00 I 00. CI'J CI'J == • \C ~ ~ r.rJ 1- ~ 32p anti-mSIRT6 138 ~ ~ ~ ~ 00 Figure 5. Quantification of auto-ADP ribosylation by mSIRT6. (A) 0.5 [LLrecombinant wild-type SIRT6 and two mutant forms (S56A, H133Y) were incubated with 3 2 p -NAD, purified by gel filtration chromatography, and analyzed by SDS-PAGE. The gel was stained with Coomassie and photographed. SIRT6 quantities were determined by comparison to BSA standards run on the same gel. Wild type SIRT6 was estimated at 0.250 g, or 6.4 pmol. (B) The stained gel from (A) was analyzed by autoradiography to confirm radiolabeling of SIRT6. (C) SIRT6 bands were excised from the gel and 3 2 p radioactivity was quantitated by scintillation counting. Results are normalized to measurements from the corresponding region excised from pET28a vector control lanes. (D) An NAD standard curve was generated by scintillation counting of known quantities of 32 p -NAD diluted to reflect the ratio of unlabeled NAD to 32 p -NAD in the reaction. (E) Ratio of incorporated ADP-ribose to SIRT6 was determined as follows. Radioactivity measured in SIRT6 bands was converted to pmol NAD by comparison to the standard curve. Pmol of NAD was divided by pmol of SIRT6 to attain the fraction of labeled mSIRT6. 139 A. B. SlRT6 Coomassie c. 32p D. 200 200 CJSIRT6 lBS56A _H133Y CO) 0 ~ )( E CO) )( 100 E 0. r2 =0.996 0 ....... 100 0. U U 0 SIRT6 S56A 0 0.000 H133Y 0.275 0.550 0.825 NAD(pmol) E. .20 CJSIRT6 iiiJS56A _H133Y T "0<0 Q)~ -0:: eou) 0.0 L.. ..- 8 Q) .Q :g .10 .5 ~ cod. 0::0 ~ 0 SIRT6 S56A H133Y 140 1.00 Figure 6. Auto-ADP-ribosylation of mSIRT6. (A) Equal amounts of recombinant mSIRT6 and mutant mSIRT6-H133Y protein were incubated with 32 -NAD*and blotted to nitrocellulose as described in Fig. 4. The blot was probed with antibody specific to mono-ADP-ribose, stripped, and reprobed with antibody specific to mSIRT6 as a control. (B) GST-SIRT6 was incubated with 32 P-NAD* and either mSIRT6 (lane 1), mSIRT6- S56A (lane 2), or mSIRT6-H133Y (lane 3). Reactions were analyzed by autoradiography as in Fig. 4b. Coomassie staining of the gel prior to film exposure revealed similar amounts of recombinant protein in all three lanes (not shown). 141 A. B. rI) :: < \C I .... rI) ~ ('t") ('t") ,.... ,.... ('t") :: \C == \C \C ~ I-ol ~ I-ol ~ a-c I 00 I ~== ~ ~ I-ol 00 + 00 ~ 00 ~ 00 ~ ~ CI:J CI:J I \C \C CI:J 00 ~ ~ ~ ~ I-ol lI) 00 I I ~('t") .... rI) .,... \C ~ ~ a-c + I :: ~ ('t") ('t") ,.... == ~ I ~ I-ol 00 + ~ 00 ~ I \C \C 00 00 ~ ~ I-ol ~ ~ -+- SIRT6-GST ~ anti-mSIRT6 ~ anti-ADPR -+- SIRT6-His 32p 142 Chapter4 Conclusions 143 SSD1 and UTH1 Naturally occurring polymorphisms can account for significant effects on longevity and the rate of aging, not only in yeast but in humans, where a locus linked to extreme long life span has been mapped to chromosome 4 (Puca et al., 2001). Different alleles of S. cerevisisae SSDI have been isolated from both laboratory and wild yeast strains, indicating that this is a true polymorphic locus and suggesting that variation of SSD1 positively impacts survival under certain environmental conditions (Sutton et al., 1991; Wheeler et al., 2003). The observation that SSDI-V affects aging adds to an already long list of cellular processes influenced by this gene. A summary of SSDI's reported genetic interactions, included in Appendix A, Table 1, lists more than 50 genes whose phenotypes are modified by SSD1. It is therefore notable that we present the first mechanistic study of the action of Ssdlp in cells, establishing this protein as a direct repressor of translation of at least one key target mRNA. Ssdl, like Mpt5/Puf5/Uth4, is likely to have a conserved regulatory role of some sort, probably in its capacity as an RNA binding protein and repressor of translation. However, the key downstream target of Ssdlp, Uthlp, is a protein specific to fungi. Interestingly, several genes associated with yeast longevity play a conserved role in evolution, despite the fact that yeast aging is predominantly controlled by a mechanism specific to yeast cells (discussed in Chapter 1). It is therefore possible that Ssdl proteins play a role in maintaining life span and stress resistance in other organisms. Exactly how Ssdlp and Uthlp affect life span is unclear, but genetic data clearly indicate that the mechanism is distinct from the main yeast aging pathways discussed in detail in Chapter 1. SSDJ-V extends the life span of sir2A and sir2AfoblA strains, 144 strongly suggesting that the longevity caused by Ssdlp is independent of Sir2p activity. Indeed, Ssdlp does not reduce ERC accumulation or rDNA marker loss, strongly supporting this conclusion. A recent report indicates that the Sir2p homolog Hst2p promotes longevity during caloric restriction by a mechanism very similar to the one already described for Sir2p (Lamming et al., 2005). This effect, while genetically independent of SIR2, still occurs through a decrease in rDNA silencing and a reduction in ERC formation and accumulation. For this reason, we believe that Hst2p is an unlikely candidate to account for the Sir2p-independent extension of life span by Ssdlp. Ssdlp exerts a profound effect on cell wall stability and cell integrity. Although the bud scar hypothesis of aging has been widely discredited, it is nevertheless possible that cell wall stability becomes a limiting factor for extremely old cells. Senescence is typically accompanied by eventual cell lysis, and it is therefore possible that stabilizing the cell wall might prevent this lysis and allow cell division to continue. A small portion of the extension of life span observed in SSDI - V cells is attributable to an increase in Nca3p, a mitochondrial homolog of Uthlp. NCA3 mRNA is upregulated in both SSD1-V and UTHI cells, possibly as a means of compensating for a reduction in Uthlp function in both cases. Overexpression of NCA3 slightly extends life span, and deletion of this gene slightly shortens life span in an SSDI-V background. Despite extensive homology to UTHI, NCA3 does not appear to associate with Ssdlp in cells, and is regulated quite differently in response to SSD1-V. NCA3 has been reported to affect mitochondrial biogenesis (Pelissier et al., 1995), and might exert its modest effects on aging through an influence on metabolism. 145 Ssdlp is the second translational repressor found to strongly affect yeast aging. The first, Mpt5/Uth4/Puf5, also plays a conserved role in germ line stem cell development in C. elegans and Drosophila (Crittenden et al., 2002; Lin and Spradling, 1997). It will be interesting to see if Ssdl and Mpt5 homologs have co-evolved with similar roles in these two organisms. In yeast, the similarities between Ssdlp and Mpt5 are striking. Both are large polymorphic repressors of RNA translation affecting cell wall stability, stress resistance, and life span through the regulation of critical target molecules. The discovery of these two genes as key polymorphic regulators suggests that control of RNA translation and stability is a key mechanism controlling phenotypic variation in the wild. Very little is known of the homologs of SSDI in higher organisms, although SSD1 is a conserved gene. The finding, presented in Chapter 2, that Ssdlp can act as a translational repressor should provide a starting point for studies of Ssdlp-related proteins in other organisms. We believe the next major step in the research of Ssdlp will be to systematically identify the RNA molecules associated with Ssdlp in vivo, as these represent excellent targets for regulation by Ssdlp. This approach has succeeded in identifying targets of several RNA-binding proteins in yeast and worms (Gerber et al., 2004; Lee and Schedl, 2001), and shows promise in other systems as well. It will be interesting to see if the other targets of Ssdlp are connected by a common theme, such as cell wall regulation or metabolism. It will also be interesting to see whether Ssdl homologs regulate similar targets in different species. 146 SIR2 and CR In Appendix B, we discuss the mechanism by which CR increases Sir2p activity to promote rDNA stability and longevity. The results presented in this thesis support the model wherein CR increases respiration, effectively increasing the NAD/NADH ratio through a reduction in NADH levels. By this model, Sir2p activity is increased by relief of competitive inhibition by NADH. An opposing model posits that Pnclp, a nicotinamidase involved in NAD synthesis, is increased by CR, thereby lowering the levels of nicotinamide that normally limit Sir2p function (Anderson et al., 2003). There is also a third possible model that has not been explicitly addressed by any published studies. Although Sir2p is a nuclear protein, there is reason to believe that NAD is freely diffusible between the nucleus and the cytoplasm. It is therefore possible that the major NAD-utilizing glycolytic enzymes of the cytosol compete with Sir2p for NAD binding, and it is the reduction of these enzymes during CR that activates Sir2p by increasing available (but not total cellular) NAD. Consistent with this model, deletion of glyceraldehyde 3-phosphate dehydrogenase (GAPDH), which converts glyceraldehydes3-phosphate to 1,3-diphosphoglycerate while reducing one molecule of NAD to NADH, extends life span in a Sir2p-dependent manner (Lin et al., 2000). Overexpression of GAPDH isoform TDH3 decreases Sir2p activity and rescues toxicity resulting from extremely high SIR2 overexpression (Matecic et al., 2002). Future studies should investigate whether reduction of major NAD-utilizing enzymes of glycolysis is sufficient to activate Sir2p, and whether this activation might result from increased NAD availability to Sir2p. 147 SIRT6 and SIRT7 In Chapter 3 and Appendix C, we present the first characterization of class IV members of the Sir2 phylogenetic tree (Frye, 2000). Interestingly, despite extensive conservation within the catalytic core, both SIRT6 and SIRT7 fail to demonstrate an in vitro NAD-dependent deacetylase activity. Mouse SIRT6 does, however, catalyze a robust auto-ADP-ribosyltransferase reaction, the discovery of which counters the assumption, widely held in the field, that Sir2 proteins are deacetylases first and foremost, only weakly catalyzing ADP-ribosylation. It remains to be seen what is the biological significance of this ADP-ribosylation reaction. In most known cases, ADPribosylation of arginine residues results in reversible inactivation of the substrate protein (Riese et al., 2002), although this modification may enhance certain enzymatic functions within the substrate (Weng et al., 1999). Future studies should identify ADP-ribosylated targets of SIRT6. SIRT7, is a broadly expressed nucleolar protein that strongly activates RNA polymerase I transcription by an enzymatic mechanism. It remains to be seen what the exact molecular target of SIRT7 is, but it is likely a member of the enormous Pol I complex. SIRT7 displays neither deacetylase activity nor ADP-ribosyltransferase in vitro, but this may result from substrate specificity issues rather than an inherent lack of activity. SIRT1, orthologous to the yeast Sir2, has the broadest substrate specificity of the mammalian sirtuins (North and Verdin, 2004) and is also the best characterized (reviewed in Chapter 1). The discovery of SIRT7 as an activator of rDNA transcription provides a potential link between energy metabolism, NAD availability, and ribosome biogenesis, all of which one would expect to be tightly linked. It will therefore be 148 important and interesting to establish a model of SIRT7 regulation analogous to those established for the yeast Sir2p. 149 REFERENCES Anderson, R. M., Bitterman, K. J., Wood, J. G., Medvedik, O., and Sinclair, D. A. (2003). Nicotinamide and PNC1 govern lifespan extension by calorie restriction in Saccharomyces cerevisiae. Nature 423, 181-185. Crittenden, S. L., Bernstein, D. S., Bachorik, J. L., Thompson, B. E., Gallegos, M., Petcherski, A. G., Moulder, G., Barstead, R., Wickens, M., and Kimble, J. (2002). A conserved RNA-binding protein controls germline stem cells in Caenorhabditis elegans. Nature 417, 660-663. Frye, R. A. (2000). Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochem Biophys Res Commun 273, 793-798. Gerber, A. P., Herschlag, D., and Brown, P. 0. (2004). Extensive association of functionally and cytotopically related mRNAs with Puf family RNA-binding proteins in yeast. PLoS Biol 2, E79. Lamming, D. W., Latorre-Esteves, M., Medvedik, O., Wong, S. N., Tsang, F. A., Wang, C., Lin, S. J., and Sinclair, D. A. (2005). HST2 mediates SIR2-independent life-span extension by calorie restriction. Science 309, 1861-1864. Lee, M. H., and Schedl, T. (2001). Identification of in vivo mRNA targets of GLD-1, a maxi-KH motif containing protein required for C. elegans germ cell development. Genes Dev 15, 2408-2420. Lin, H., and Spradling, A. C. (1997). A novel group of pumilio mutations affects the asymmetric division of germline stem cells in the Drosophila ovary. Development 124, 2463-2476. Lin, S. J., Defossez, P. A., and Guarente, L. (2000). Requirement of NAD and SIR2 for life-span extension by calorie restriction in Saccharomyces cerevisiae. Science 289, 2126-2128. Matecic, M., Stuart, S., and Holmes, S. G. (2002). SIR2-induced inviability is suppressed by histone H4 overexpression. Genetics 162, 973-976. North, B. J., and Verdin, E. (2004). Sirtuins: Sir2-related NAD-dependent protein deacetylases. Genome Biol 5, 224. Pelissier, P., Camougrand, N., Velours, G., and Guerin, M. (1995). NCA3, a nuclear gene involved in the mitochondrial expression of subunits 6 and 8 of the Fo-FI ATP synthase of S. cerevisiae. Curr Genet 27, 409-416. Puca, A. A., Daly, M. J., Brewster, S. J., Matise, T. C., Barrett, J., Shea-Drinkwater, M., Kang, S., Joyce, E., Nicoli, J., Benson, E., et al. (2001). A genome-wide scan for linkage to human exceptional longevity identifies a locus on chromosome 4. Proc Natl Acad Sci U S A 98, 10505-10508. Riese, M. J., Goehring, U. M., Ehrmantraut, M. E., Moss, J., Barbieri, J. T., Aktories, K., and Schmidt, G. (2002). Auto-ADP-ribosylation of Pseudomonas aeruginosa ExoS. J Biol Chem 277, 12082-12088. Sutton, A., Immanuel, D., and Arndt, K. T. (1991). The SIT4 protein phosphatase functions in late G1 for progression into S phase. Mol Cell Biol 11, 2133-2148. Weng, B., Thompson, W. C., Kim, H. J., Levine, R. L., and Moss, J. (1999). Modification of the ADP-ribosyltransferase and NAD glycohydrolase activities of a mammalian transferase (ADP-ribosyltransferase Chem 274,31797-31803. 150 5) by auto-ADP-ribosylation. J Biol Wheeler, R. T., Kupiec, M., Magnelli, P., Abeijon, C., and Fink, G. R. (2003). A Saccharomyces cerevisiae mutant with increased virulence. Proc Natl Acad Sci U S A 100, 2766-2770. 151 152 Appendix A Saccharomyces cerevisiae SSDJ-V Promotes Longevity by a Sir2p-Independent Mechanism This chapter was published in Genetics Volume 166 Issue 4 pages 1661-1672. The authors were Matt Kaeberlein, Alex Andalis, Gregory Liszt, Gerald Fink and Leonard Guarente. I contributed Table 3, Figure 3, and Figure 4. 153 SUMMARY TheSSDI gene of Saccharomyces cerevisiae affects diverse cellular processes including cell integrity, cell cycle progression, and growth at high temperature. We show here that the SSDI-V allele is necessary for cells to achieve extremely long life span. Furthermore, we demonstrate that addition of SSDI - V to cells can increase longevity independently of SIR2 or calorie restriction. Microarray analysis differentiates SSDI1-Vfrom four other long-lived cell types and suggests that the presence of SSDI -V results in altered transcript levels for genes involved in many cellular processes including cell wall biosynthesis, DNA repair and stress response, and RNA processing. We propose that SSD1-V defines a previously undescribed pathway affecting cellular longevity and aging. 154 INTRODUCTION Aging in Saccharomyces cerevisiae can be studied by mutations that extend the replicative life span of mother cells, defined as the number of daughters produced by a given mother cell prior to senescence. We report here a new longevity-promoting factor, SSD1-V, that extends mean life span up to 85% and promotes longevity in the absence of Sir2p. One cause of aging in yeast is the accumulation of extrachromosomal ribosomal DNA circles (ERCs) (SINCLAIR and GUARENTE 1997). The yeast ribosomal DNA (rDNA) is organized into a tandem array of 100-200 copies of a 9.1 kb repeat (PETES and BOTSTEIN 1977; PHILIPPSEN et al. 1978; RUSTCHENKO and SHERMAN 1994). Homologous recombination between adjacent repeats can result in excision of an ERC. ERCs are self-replicating and asymmetrically segregated to the mother cell nucleus during S-phase, resulting in an exponential increase in ERC copy number with age ( SINCLAIR and GUARENTE 1997). Once a threshold level of ERCs is achieved, it is thought that a cell senescence pathway is activated that culminates in cell death. A central regulator of life span in yeast is the Sir2 protein (KAEBERLEIN et al. 1999). Sir2p is an NAD-dependent histone deacetylase (IMAI et al. 2000; LANDRY et al. 2000; SMITH et al. 2000) required for transcriptional silencing at telomeres (GOTTSCHLING et al. 1990), silent mating (HMA)loci (IVY et al. 1986; RINE and HERSKOWITZ 1987), and the rDNA (BRYK et al. 1997; SMITH and BOEKE 1997). Mutation of SIR2 results in increased rDNA recombination (GOTTLIEB and ESPOSITO 1989), increased ERC formation (KAEBERLEIN et al. 1999), and decreased life span (KENNEDY et al. 1995), wheras overexpression extends life span by 30-40% (KAEBERLEIN et al. 1999). Sir2p is also required for life span extension by calorie restriction (CR), demonstrating the importance of this protein as a central regulator of longevity (LIN et al. 2000). Intriguingly, overexpression of a Sir2p homolog, Sir-2.1, was found to extend life span in the nematode C. elegans, suggesting that Sir2 proteins also regulate aging in higher eukaryotes (TISSENBAUM and GUARENTE 2001). 155 Genetic interaction between MPT5 and SSDI: SSDI is a polymorphic locus that affects diverse cellular processes. Two allele classes have been identified for SSD1, designated SSDI1-V and ssdl-d. SSDI-V alleles confer viability in the absence of the Sit4 protein phosphatase and code for functional Ssdl protein. In contrast, strains carrying ssdl-d alleles are inviable in the absence of Sit4p (SUTTON et al. 1991), and d-type alleles are likely null for Ssdlp function. Both V and d type alleles have been found in natural isolates as well as laboratory strains of S. cerevisiae. A recent report (WHEELER et al. 2003) suggests that SSDI allele type affects pathogenicity of yeasts, suggesting that the SSD1 locus is susceptible to strong selective pressure under certain einvironmental conditions. A potential role for SSD 1-V as a regulator of cell life span was suggested by the observation that SSD1-V suppresses many phenotypes associated with mutation of the MPT5IUTH4 gene (KAEBERLEIN and GUARENTE 2002). MPT5 is a post- transcriptional regulator (TADAUCHI et al. 2001) involved in regulating pheromone response (CHEN and KURJAN 1997), cell wall stability (KAEBERLEIN and GUARENTE 2002), telomere silencing (COCKELL et al. 1998), and longevity (KENNEDY et al. 1995). Like SIR2, MPT5 is a limiting factor for longevity: overexpression of MPT5 extends life span, while deletion has the opposite effect (KENNEDY et al. 1997). SSDI1-Vsuppresses the temperature sensitive growth defect caused by mutation of Mpt5p as well as the sensitivity to calcofluor white (CFW) and sodium dodecyl sulfate (SDS)(KAEBERLEIN and GUARENTE 2002). Furthermore, in strains lacking SSDI-V, deletion of MPT5 is synthetically lethal in combination with loss of function in either SBF or CCR4 transcriptional complexes (KAEBERLEIN and GUARENTE 2002), both of which function downstream of protein kinase C (Pkclp) to promote cell wall biosynthesis (CHANG et al. 1999; IGUAL et al. 1996; MADDEN et al. 1997). These 156 results were interpreted to suggest that Mpt5p, Ssdlp, and Pkclp define three parallel pathways that function to ensure cell integrity (KAEBERLEIN and GUARENTE 2002). In addition to suppressing the cell integrity defects, SSD1-V suppresses the shortened life span caused by deletion of MPT5 (KAEBERLEIN and GUARENTE 2002). We were therefore interested in examining the possibility that SSDI-V might also promote longevity in wild type cells. Here we show that addition of a single copy of SSDI-V to ssdl-d wild type cells extends life span in at least two different strain backgrounds. Furthermore, life span extension by SSDI-V does not require the Sir2 protein, although the presence of both SSDI-V and SIR2 is necessary for maximal longevity. MATERIALS AND METHODS Strains and genetic techniques: The strains used in this study are listed in Table 2. All strains were derived from W303R (described in MILLS et al. 1999), PSY316 (described in PARK et al. 1999), or BKY5 (described in KENNEDY et al. 1995).Genetic crosses, sporulation, and tetrad analysis were carried out as described (SHERMAN and HICKS 1991). The genotype of inviable spore clones was inferred when possible based on marker segregation in viable spore clones from the same tetrad. Unless otherwise noted, cells were cultured in YPD or synthetic media prepared using conventional methods (GUTHRIE and FINK 1991). Yeast transformation was accomplished by the lithium acetate method (GIETZ et al. 1992). The sir2 fob] strain was constructed as described (LIN et al. 2000). All other gene deletions were generated by transforming cells with PCR amplified disruption cassettes as described (KAEBERLEIN et al. 1999). In each case, the entire open reading frame was removed. All disruptions were verified phenotypically or by PCR. The SSDI1-Vintegrating plasmids p406SSD1 and p405SSDl1 were previously described (KAEBERLEIN and GUARENTE 2002). Unless otherwise indicated, all SSDI- V strains contain SSDI- V integrated at the marker locus and still carry the ssdl-d allele at the SSD1 locus. Deletion of ssdl-d does not affect any of the phenotypes tested, including life span, growth at 30 ° , 37° , or 40 ° , or sensitivity to calcofluor white. 157 Determination of SSD1 allele: In order to determine which SSDI allele was present in PSY316, we deleted one copy of the SIT4 gene in diploid cells. Sporulation of these cells revealed that deletion of SIT4 always resulted in lethality (n>20 sit4A spore clones). This lethality was suppressed by integration of a single copy of SSD1-V at the URA3 locus. Therefore, we conclude that in PSY316, the SSDI allele is a ssdl-d allele. Life span, recombination and ERC analysis: Life spans were performed as described (KAEBERLEIN and GUARENTE 2002). Statistical significance was determined by a Wilcoxon rank sum test. Average life span is stated to be different for p<0.05. Each figure represents data derived from a single experiment, unless otherwise stated. ERC levels were determined as described (KAEBERLEIN et al. 1999; DEFOSSEZ et al. 1999). ERCs were separated on a 0.6% agarose gel without addition of ethidium bromide at 1 Volt/cm for 48 hr. rDNA recombination rate was determined as described (KAEBERLEIN et al. 1999). Microarray analysis: RNA isolation and microarray analysis was performed essentially as described (LIN et al. 2002). Candidate genes with altered transcript levels in SSD1-V cells relative to wild type were defined as such if the ratio of SSD1-V (Cy5) to ssdl-d (Cy3) was greater than 1.5 or less than 0.667 (1.5-fold decrease) in 2/2 independent experiments. All such genes are listed in the supplemental data (Supplemental Table 1). As a control, microarray analysis was performed on ssdlA (Cy5) cells relative to ssdl-d (Cy3) cells. The 124 genes defined as regulated by CR represent the subset of genes found to show significant changes in mRNA expression both in cells lacking HXK2 and in cells grown on 0.5% glucose (LIN et al. 2002). Cluster analysis was performed using Cluster and visualized with TreeView (EISEN et al. 1998). Statistical significance of the overlap between regulated genes in different experiments shown in Figure 4 was calculated using a hypergeometric distribution. P-values were obtained from the online hypergeometric distribution calculator at http://www.alewand.de/stattab/tabdiske.htm. Supplemental Table 1 as well as the entire microarray data sets for all experiments presented in this paper are available on the World Wide Web at 158 http://web.mit.edu/biology/guarente/microarray/kaeberlein. Files may be downloaded as Microsoft Excel spreadsheets. RESULTS SSDI-V extends life span and improves growth at high temperature: Mpt5p is a limiting factor for longevity and functions in a pathway parallel to SSDI -V for cell integrity (KAEBERLEIN and GUARENTE 2002). Based on this genetic interaction, we hypothesized that SSD1-V might also regulate longevity. We first determined that our wild type strain PSY316 carries the ssdl-d allele at the SSDI locus, based on inviability caused by deletion of SIT4 in this background (see Materials and Methods). A single copy of SSD1-V integrated at the URA3 locus results in an approximately 50% increase in mean life span (Figure 1A). Integration of SSD1-V at the SSDI locus has a similar effect on life span (not shown). Deletion of the chromosomal ssdl-d allele of PSY316 has no effect on life span and does not affect life span extension by SSD1-V (Figure 1A). Therefore ssdl-d is a null allele with respect to life span. All subsequent experiments were carried out in the parental ssdl-d background. PSY316 is a moderately long-lived yeast strain, however many yeast aging studies have been carried out in short lived strain backgrounds having mean life spans of 10-15 generations. In order to determine whether the life span extension by SSD1-V was strain specific, we integrated SSD1-V into the short-lived strain BKY5. We verified that BKY5 carries an ssdl-d allele (see Materials and Methods) as well as a previously identified C- terminal truncated allele of MPT5 (KENNEDY et al. 1997). Addition of SSD- V to BKY5 results in an 85% increase in mean life span (Figure B). Thus, SSD1-Vpromotes long life span in at least two different ssdl-d strain backgrounds. We had previously observed that cells from strain PSY316 grow normally at 37 ° , but are incapable of sustained growth at 40 ° . Cells grown at 40 ° generally arrest as large budded cells with a significant fraction undergoing lysis (data not shown), consistent with a loss 159 of cell wall integrity at the restrictive temperature. Addition of SSDI-V fully suppresses these phenotypes and allows growth of PSY316 at 40° (Figure 1C). Addition of SSD1-V to PSY316 also improves growth in the presence of the cell wall perturbing agents CFW and SDS (data not shown), as previously reported for strain W303R (KAEBERLEIN and GUARENTE 2002). SSDI-V extends life span in the absence of SIR2: The Sir2 protein is a central regulator of yeast longevity, necessary for life span extension in response to environmental signals such as reduced nutrient availability (LIN et al. 2000) and osmotic stress (KAEBERLEIN et al. 2002). In order to place SSD1-V into a genetic pathway relative to Sir2p, we integrated the SSDI1-Vallele into a sir2 fob] double mutant. The sir2fob] double mutant was utilized because it has a nearly wild-type life span, rather than the extremely short life span observed for sir2 FOB] cells (KAEBERLEIN et al. 1999). As predicted by our model previously proposed model (LIN et al. 2000; LIN et al. 2002), calorie restriction (CR) by growth on low glucose fails to extend life span in the absence of SIR2 (Figure 2A). In contrast, sir2 fob] SSD1-V cells grown on 2% glucose have a life span that is significantly longer than sir2 fob] ssdl -d cells. However, sir2 fob] SSD1-V cells do not live as long as SIR2 FOB] SSD1-V cells (Figure 2B). Therefore, SSDI-V acts in a novel, SIR2-independent pathway for longevity, although Sir2p is required for maximum longevity in SSD1-V cells. 160 The Sir2-dependent life span extension caused by CR is the result of a metabolic shift from fermentation to respiration (LIN et al. 2002). It is possible that a portion of the longevity conferred by SSD1-V is achieved by causing the cell to undergo a similar metabolic shift. To address this possibility, the effect of SSD1-V on life span was examined in a respiration deficient strain lacking the CYT1 gene encoding cytochrome cl. Mutation of CYT1 prevents life span extension by CR or by overexpression of the Hap4p transcription factor (LIN et al. 2002). In contrast, SSD 1-V extends the life span of cells lacking CYT1 to the same extent as wild type cells (Figure 2C), suggesting that the mechanism of life span extension by SSD 1-V is independent of mitochondrial function and respiration. SSDI-V acts independently of ERC formation or accumulation: Sir2p and calorie restriction promote longevity by decreasing the formation and accumulation of ERCs in mother cells (KAEBERLEIN et al. 1999; LIN et al. 2000). In order to determine whether the long life span of SSD1-V cells is also due to fewer ERCs, we measured the rate of ERC formation and the amount of ERCs present in ssdl-d and SSD1-V cells. ERC formation was estimated by determining the frequency at which an ADE2 marker integrated into the rDNA is lost, as demonstrated by the presence of half-sectored colonies (KAEBERLEIN et al. 1999). No difference was observed between SSD1-V and ssdl-d cells by this assay (Figure 3A). Upon direct quantitation of ERCs from unsorted cells, we observed that SSD1-V cells often had higher levels of ERCs than ssdl-d cells (Figure 3B), indicating that the life span extension caused by SSD 1-V is unlikely to be the result of decreased ERC formation or accumulation. This is consistent with the observation that SSD1-V does not require Sir2p to extend life span. Transcriptional analysis of SSDI-V: In order to further understand the effect of SSD1-V on cell physiology and longevity, we used microarray analysis to examine the transcriptional profile of SSD1-V cells relative to wild type ssdl-d cells. In 2/2 independent experiments, 83 genes were down-regulated and 217 genes were up- regulated at least 1.5-fold by SSDI-V (see Supplemental Table 1). 161 The most common phenotype associated with Ssdlp is the ability to suppress temperature sensitivity caused by mutations that affect the cell wall (Table 1). Recently, SSDI -Vhas been found to increase resistance to the cell wall perturbing agents CFW and SDS (KAEBERLEIN and GUARENTE 2002) and to alter the amount of chitin, 3-1,6-glucan, -1,3-glucan, and mannan present in the cell wall (WHEELER et al. 2003). Microarray analysis suggests that these observations can be at least partially explained by altered expression of cell wall genes in SSDI-V cells. Messenger RNA levels for several cell wall proteins, including the Chs5p chitin synthase and Kre5p, a protein required for 1,6-glucan biosynthesis, are elevated by SSDI-V. Interestingly, we also observe increased transcript levels for several proteins involved in sporulation and spore wall formation. In particular, SWM1 and TEP1, two genes involved in assembly of the spore cell wall are up-regulated 2.2-fold and 2.1-fold respectively, perhaps suggesting a fundamental alteration in the cell wall structure of vegetatively growing SSD1-V cells. We have previously demonstrated a correlation between cell wall stability and longevity (KAEBERLEIN and GUARENTE 2002), suggesting that enhanced cell wall integrity may be necessary for cells to attain extreme longevity. SSD1-V has also recently been reported to promote resistance to the plant antifungal osmotin (IBEAS et al. 2000; IBEAS et al. 2001). It has been suggested that this increased resistance is due to an altered composition of cell wall mannoproteins in SSDIVcells. Indeed, we find that mRNA coding for the cell wall mannoprotien, Cwplp, is decreased 2.4-fold in SSD1-V cells. Cwpl is known to directly bind osmotin (IBEAS et al. 2000), suggesting that reduction of Cwplp might promote the osmotin resistance observed in SSD1-V cells. SSD1-Vhas also been found to suppress the temperature sensitivity of several splicing defective mutations (LUUKKONEN and SERAPHIN 1999). Consistent with these observations, we find that several genes involved in RNA processing are differentially transcribed in SSD1-V cells relative to ssdl-d cells, perhaps accounting for this phenotype. In addition to these changes, SSD1-V cells also show altered levels of mRNA coding for proteins involved in DNA repair, stress response, mating, and mitochondrial 162 function. Reasonable hypotheses could be made as to why each of these functional categories might be important for the effect of SSD1-V on life span. Therefore, further experimental analysis of specific candidate genes will be required in order to determine which are important for longevity. One particularly interesting candidate gene that shows altered mRNA levels in SSDI1-V cells is NCA3. Nca3p is a member of the SUN family of proteins (Siml, Uthl, Nca3, and Sun4) and functions to promote maturation of the mitochondrially encoded ATP8-ATP6 co-transcript (PELISSIER et al. 1995). Up-regulation (3.4-fold) of NCA3 by SSD -V is striking because Nca3p shares extensive homology (60%) with the aging protein Uthlp. Interestingly, we find that NCA3 mRNA is increased 4-fold in long-lived cells lacking Uthlp (data not shown). Overexpression of NCA3 from the ADHI promoter results in a slight, but reproducible increase in life span (Figure 4A), suggesting that Nca3p dosage can affect longevity. However, NCA3 is not required for the majority of the life-span extension seen in SSDI-V cells, as demonstrated by the finding that nca3 SSDI-V cells have a life span comparable to NCA3 SSD1-V cells (Figure 4B). Therefore, we conclude that increased transcription of NCA3 accounts for, at most, a minor fraction of the longevity promoting activity of SSD1-V. Gene expression analysis of long-lived cell types: We have previously shown that calorie restriction by growth in 0.5% glucose, which promotes long life span, causes characteristic changes in gene expression that are reproduced in two genetic models of CR, namely overexpression of HAP4 and deletion of HXK2 (LIN et al. 2002). More recently, we demonstrated that growth in the presence of high external osmolarity (HEO) also extends life span in a SIR2-dependent manner and results in a gene expression profile with significant similarity to calorically restricted cells (KAEBERLEIN et al. 2002). We were therefore interested in determining how closely the changes in gene expression caused by addition of SSD1-V would match those caused by CR and HEO. Two dimensional cluster analysis (EISEN et al. 1998) of the SSD1-V gene expression data and previously published data sets for 0.5% glucose, HAP4-overexpression, HXK2 163 deletion (LIN et al. 2002), and 20% glucose (KAEBERLEIN et al. 2002) was performed over the set of genes previously identified as being regulated by CR (see Materials and Methods; LIN et al. 2002). Two independently derived data sets were included for each condition, except for cells lacking HXK2, for which only one microarray experiment was performed. In each case, the two data sets corresponding to a particular condition cluster together (Figure 5A). As expected, cells grown in 0.5% glucose cluster most closely with cells lacking HXK2 and define the CR group (Figure 5B). Cells overexpressing HAP4 also cluster with this group, although on a separate branch. Cells exposed to HEO cluster with the HAP4/CR group on the next level of the tree, and together these four cell types comprise the SIR2dependent longevity group. Strikingly, SSDI-V cells cluster separately from all the other data sets at the highest level of the tree. This analysis is further supported by a comparison of genes co-regulated by each of the longevity promoting conditions above. Of the 124 genes regulated by CR, 55 are similarly regulated by HAP4-overexpression and 48 are similarly regulated by growth in 20% glucose (Figure 5C). In contrast, only 8/290 genes differentially transcribed by addition of SSD1-V overlap with the CR-regulated genes (Figure 5C). Thus, the gene expression changes observed by microarray analysis strongly support the genetic data and suggest that SSD1-V promotes longevity by a novel mechanism unrelated to CR. MPT5 and SIR2 affect longevity in a pathway parallel to SSDI-V: We have previously demonstrated that MPT5 and SSD1-V act in parallel pathways to promote cell integrity and that SSD1-V suppresses the short life span caused by deletion of MPT5 in the W303R strain background (KAEBERLEIN AND GUARENTE 2002). As expected, addition of SSD1-V also suppresses the short life span of the mpt5 deletion in PSY316 (Figure 6A). It is interesting to note, however, that mpt5 SSD1-V cells have a life span intermediate between cells lacking both MPT5 and SSD1-V and cells with functional copies of both genes. In fact, mpt5 SSD1-V cells have a life span not significantly different than the wild type MPT5 ssdl-d strain, suggesting that MPT5 and SSDJ- V have 164 additive effects on longevity, as would be expected for genes functioning in parallel pathways. Like SSDI-V, overexpression of MPT5 increases mother cell life span (KENNEDY et al. 1997). Since SSDI-V is capable of extending the life span of cells lacking SIR2, we wished to determine whether overexpression of MPT5 would have a similar effect. In contrast to SSD1-V, MPT5-overexpression fails to extend the life span of sir2 fobl cells (Figure 6B), suggesting that MPT5 and SIR2 act in the same pathway to promote longevity. It had been previously observed that cells with altered dosage of MPT5 display changes in telomeric and rDNA silencing (KENNEDY 1996), suggesting a further link between MPT5 and SIR2. Overexpression of MPT5 increases rDNA silencing and decreases telomeric silencing, while deletion has an opposite effect (Table 3). Integration of SSD1V, in contrast, has no detectable effect on silencing at either locus. Consistent with the inability of MPT5 overexpression to extend life span in the absence of SIR2, the enhanced rDNA silencing observed in cells overexpressing MPT5 is fully suppressed by deletion of SIR2 (Table 3). Based on these results, we propose that overexpression of MPT5 increases life span by relocalizing Sir2p from telomeres to the rDNA (see Discussion), thus enhancing ability of Sir2p to inhibit ERC accumulation in aging mother cells. DISCUSSION One cause of aging in yeast is the accumulation of ERCs (SINCLAIR and GUARENTE 1997). A central regulator of ERC formation and longevity is the Sir2p histone deacetylase (KAEBERLEIN et al. 1999). Several genes that regulate yeast life span act by altering Sir2p activity or dosage (KAEBERLEIN et al. 1999; LIN et al. 2000; LIN et al. 2002; KAEBERLEIN et al. 2002). Here we present evidence that SSD1-V defines a novel Sir2p-independent pathway necessary for cells to achieve extreme longevity. 165 Two pathways promoting longevity: We initially began studying SSDI-Vbased on its ability to suppress the temperature sensitivity caused by mutation of the UTH4/MPT5 gene. Like Sir2p, Mpt5p is limiting for life span in wild type cells (KENNEDY et al. 1997). Overexpression of Mpt5p increases life span and rDNA silencing in a Sir2p- dependent manner (Figure 6B, Table 3), suggesting that Mpt5p promotes longevity by increasing Sir2p activity at the rDNA. In contrast to overexpression of MPT5, addition of a single copy of SSD1-V extends life span in both SIR2 wild type and sir2 fob] double mutant cells (Figure 2C). However, the sir2fob] SSD1-V strain has a life span that is shorter than the SIR2 FOB1 SSDI-V strain, demonstrating that Sir2p is required for maximum longevity in SSDI1-Vcells. This is consistent with the observation that MPT5 SSDI -V cells have a longer life span than mpt5 SSDI-V cells (Figure 6A), and suggests a model whereby Mpt5p and Sir2p function in one pathway to increase life span while Ssdlp functions in a parallel pathway (Figure 7). Mechanism of life span extension by SSDI-V: How is SSD -V acting to extend life span? The effect of Sir2p on life span is, at least partially, due to its ability to deacetylate rDNA histones and inhibit ERC formation (KAEBERLEIN et al. 1999). We have observed no evidence to suggest that SSDI-V affects the rate of ERC formation or accumulation. Addition of SSD1-Vhad no effect on rDNA recombination (Figure 3A) or on rDNA silencing (Table 3) in PSY316, and we failed to detect a decrease in ERC levels in SSDI-V cells relative to ssdl-d cells (Figure 3B). While it is still possible that SSDI-V affects ERC replication or segregation specifically in aged cells, we feel that this is unlikely to be the case. An alternative possibility is that SSDI -V makes cells more resistant to ERCs, rather than reducing ERC levels. In support of this hypothesis, we often observed that steady-state ERC levels were increased in unsorted SSDI-V cells relative to wild-type ssdl-d cells (Figure 3B), although this was not always the case. The mechanism by which ERCs induce senescence is currently unknown. One hypothesis is that ERCs bind to and titrate key cellular replication or transcription factors away from 166 their normal targets. Alternatively, the rapid amplification of rDNA sequence could alter rRNA transcription and/or processing, resulting in ribosome dysregulation. Ssdlp has been shown to bind RNA and is predicted to have RNase activity (UESONO et al. 1997). Perhaps, SSDI -V somehow alters rRNA or ribosome biogenesis in a manner that makes cells more resistant to ERCs. One attractive hypothesis is that SSDI-Vpromotes longevity by increasing cell wall stability and cell integrity. SSDI-V suppresses several temperature sensitive mutations that weaken the cell wall (Table 1) and has been found to directly affect cell wall composition (WHEELER et al. 2003). Furthermore, microarray analysis suggests that mRNA transcripts coding for cell wall regulatory and biosynthetic proteins are differentially expressed in SSD1-V cells (Supplemental Table 1). SSD1-V also improves resistance to the cell wall perturbing agents CFW and SDS (KAEBERLEIN and GUARENTE 2002), and increases the maximum temperature at which PSY316 is capable of growth (Figure 1C). Perhaps, the cell wall becomes limiting in very old cells and SSDI-V extends life span by stabilizing it. How might cell wall stability limit replicative life span? The terminal phenotype of yeast cells in the life span assay is cell cycle arrest often accompanied by cell lysis (MCVEY et al. 2002). Enhanced cell wall stability may prevent cell lysis late in life and allow additional cell divisions to occur. 167 Genetic diversity and the study of aging: The data presented here identify a genetic polymorphism that has a profound effect on mother cell life span. Genetic polymorphisms have also been proposed to affect the likelihood of achieving extreme longevity in human populations (PUCA et al. 2001), as well as other model systems. Since both ssdl-d and SSD1-V allele types have been isolated from natural yeast populations, the SSD1 locus represents a true polymorphic locus affecting longevity. In the past, researchers studying aging in yeast have tended to avoid using long-lived wild type backgrounds. We speculate that the majority (if not all) of these shorter lived yeast strains carry ssdl-d alleles. A comprehensive reevaluation of previously identified mutations affecting life span in a long-lived SSD1-V background would be of value to the field. SSD1-V is a polymorphic locus that confers extreme longevity on yeast mother cells by a pathway independent of Sir2p. Sir2p has been found to extend life span in animals and, like SIR2, SSDI homologs are present in yeast, worms, flies, and mammals. Might SSD1 family members also promote longevity outside of yeast? The mechanism by which SSDI1-V cells achieve up to 85% longer life span is still unknown. Further work should be devoted to testing candidate longevity genes regulated by SSD1-V and to defining the molecular function of Ssdlp in cells. 168 REFERENCES Bryk, M., M. Banerjee, M. Murphy, K. E. Knudsen, D. J. Garfinkel et al., 1997 Transcriptional silencing of Tyl elements in the RDN1 locus of yeast. Genes Dev 11: 255-269. Chang, M., D. French-Cornay, H. Y. Fan, H. Klein, C. L. Denis et al., 1999A complex containing RNA polymerase II, Paflp, Cdc73p, Hprlp, and Ccr4p plays a role in protein kinase C signaling. Mol Cell Biol 19: 1056-1067. Chen, T., and J. Kurjan, 1997 Saccharomyces cerevisiae Mpt5p interacts with Sst2p and plays roles in pheromone sensitivity and recovery from pheromone arrest. Mol Cell Biol 17: 34293439. Cockell, M., H. Renauld, P. Watt and S. M. Gasser, 1998 Sif2p interacts with Sir4p amino-terminal domain and antagonizes telomeric silencing in yeast. Curr Biol 8: 787-790. Costigan, C., S. Gehrung and M. Snyder, 1992 A synthetic lethal screen identifies SLK1, a novel protein kinase homolog implicated in yeast cell morphogenesis and cell growth. Mol Cell Biol 12: 1162-1178. Cvrckova, F., and K. Nasmyth, 1993 Yeast G1 cyclins CLN1 and CLN2 and a GAP-like protein have a role in bud formation. Embo J 12: 5277-5286. Defossez, P. A., R. Prusty, M. Kaeberlein, S. J. Lin, P. Ferrigno et al., 1999 Elimination of replication block protein Fobl extends the life span of yeast mother cells. Mol Cell 3: 447-455. Doseff, A. I., and K. T. Arndt, 1995 LAS 1 is an essential nuclear protein involved in cell morphogenesis and cell surface growth. Genetics 141: 857-871. Du, L. L., and P. Novick, 2002 Paglp, a novel protein associated with protein kinase Cbklp, is required for cell morphogenesis and proliferation in Saccharomyces cerevisiae. Mol Biol Cell 13: 503-514. Eisen, M. B., P. T. Spellman, P. 0. Brown and D. Botstein, 1998 Cluster analysis and display of genome-wide expression patterns. Proc Natl Acad Sci U S A 95: 14863-14868. Evans, D. R., and M. J. Stark, 1997 Mutations in the Saccharomyces cerevisiae type 2A protein phosphatase catalytic subunit reveal roles in cell wall integrity, actin cytoskeleton organization and mitosis. Genetics 145: 227-241. Fernandez-Sarabia, M. J., A. Sutton, T. Zhong and K. T. Arndt, 1992 SIT4 protein phosphatase is required for the normal accumulation of SWI4, CLN1, CLN2, and HCS26 RNAs during late G1. Genes Dev 6: 2417-2428. Fortes, P., J. Kufel, M. Fornerod, M. Polycarpou-Schwarz, D. Lafontaine et al., 1999 Genetic and physical interactions involving the yeast nuclear cap- binding complex. Mol Cell Biol 19: 65436553. 169 Gietz, D., A. St Jean, R. A. Woods and R. H. Schiestl, 1992 Improved method for high efficiency transformation of intact yeast cells. Nucleic Acids Res 20: 1425. Gottlieb, S., and R. E. Esposito, 1989 A new role for a yeast transcriptional silencer gene, SIR2, in regulation of recombination in ribosomal DNA. Cell 56: 771-776. Gottschling, D. E., O. M. Aparicio, B. L. Billington and V. A. Zakian, 1990 Position effect at S. cerevisiae telomeres: reversible repression of Pol II transcription. Cell 63: 751-762. Guthrie, C., and G. R. Fink, 1991 Guide to yeast genetics and molecular biology. Methods Enzymol 194. Ibeas, J. I., H. Lee, B. Damsz, D. T. Prasad, J. M. Pardo et al., 2000 Fungal cell wall phosphomannans facilitate the toxic activity of a plant PR-5 protein. Plant J 23: 375-383. Ibeas, J. I., D. J. Yun, B. Damsz, M. L. Narasimhan, Y. Uesono et al., 2001 Resistance to the plant PR-5 protein osmotin in the model fungus Saccharomyces cerevisiae is mediated by the regulatory effects of SSD1 on cell wall composition. Plant J 25: 271-280. Igual, J. C., A. L. Johnson and L. H. Johnston, 1996 Coordinated regulation of gene expression by the cell cycle transcription factor Swi4 and the protein kinase C MAP kinase pathway for yeast cell integrity. Embo J 15: 5001-5013. Imai, S., C. M. Armstrong, M. Kaeberlein and L. Guarente, 2000 Transcriptional silencing and longevity protein Sir2 is an NAD- dependent histone deacetylase. Nature 403: 795-800. Ivy, J. M., A. J. Klar and J. B. Hicks, 1986 Cloning and characterization of four SIR genes of Saccharomyces cerevisiae. Mol Cell Biol 6: 688-702. Jacoby, J. J., S. M. Nilius and J. J. Heinisch, 1998 A screen for upstream components of the yeast protein kinase C signal transduction pathway identifies the product of the SLG1 gene. Mol Gen Genet 258: 148-155. Kaeberlein, M., A. A. Andalis, G. R. Fink and L. Guarente, 2002 High osmolarity extends life span in Saccharomyces cerevisiae by a mechanism related to calorie restriction. Mol Cell Biol 22: 8056-8066. Kaeberlein, M., and L. Guarente, 2002 Saccharomyces cerevisiae MPT5 and SSD1 function in parallel pathways to promote cell wall integrity. Genetics 160: 83-95. Kaeberlein, M., M. McVey and L. Guarente, 1999 The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev 13: 2570-2580. Kennedy, B. K., 1996 Genetic and Molecular Analysis of Aging in Yeast, pp. 233 in Biology. Massachusetts Institute of Technology, Boston. Kennedy, B. K., N. R. Austriaco, Jr., J. Zhang and L. Guarente, 1995 Mutation in the silencing gene SIR4 can delay aging in S. cerevisiae. Cell 80: 485-496. 170 Kennedy, B. K., M. Gotta, D. A. Sinclair, K. Mills, D. S. McNabb et al., 1997 Redistribution of silencing proteins from telomeres to the nucleolus is associated with extension of life span in S. cerevisiae. Cell 89: 381-391. Kikuchi, Y., Y. Oka, M. Kobayashi, Y. Uesono, A. Toh-e et al., 1994 A new yeast gene, HTR1, required for growth at high temperature, is needed for recovery from mating pheromone-induced G1 arrest. Mol Gen Genet 245: 107-116. Kim, Y. J., L. Francisco, G. C. Chen, E. Marcotte and C. S. Chan, 1994 Control of cellular morphogenesis by the Ipl 2/Bem2 GTPase-activating protein: possible role of protein phosphorylation. J Cell Biol 127: 1381-1394. Kosodo, Y., K. Imai, A. Hirata, Y. Noda, A. Takatsuki et al., 2001 Multicopy suppressors of the sly 1 temperature-sensitive mutation in the ER-Golgi vesicular transport in Saccharomyces cerevisiae. Yeast 18: 1003-1014. Landry, J., A. Sutton, S. T. Tafrov, R. C. Heller, J. Stebbins et al., 2000 The silencing protein SIR2 and its homologs are NAD-dependent protein deacetylases. Proc Natl Acad Sci U S A 97: 5807-5811. Li, B., and J. R. Warner, 1998 Genetic interaction between YPT6 and YPT1 in Saccharomyces cerevisiae. Yeast 14: 915-922. Lin, S. J., P. A. Defossez and L. Guarente, 2000 Requirement of NAD and SIR2 for life-span extension by calorie restriction in Saccharomyces cerevisiae. Science 289: 2126-2128. Lin, S. J., M. Kaeberlein, A. A. Andalis, L. A. Sturtz, P. A. Defossez et al., 2002 Calorie restriction extends Saccharomyces cerevisiae lifespan by increasing respiration. Nature 418: 344348. Lorenz, M. C., and J. Heitman, 1998 Regulators of pseudohyphal differentiation in Saccharomyces cerevisiae identified through multicopy suppressor analysis in ammonium permease mutant strains. Genetics 150: 1443-1457. Luukkonen, B. G., and B. Seraphin, 1999 A conditional U5 snRNA mutation affecting premRNA splicing and nuclear pre-mRNA retention identifies SSD1/SRK1 as a general splicing mutant suppressor. Nucleic Acids Res 27: 3455-3465. Madden, K., Y. J. Sheu, K. Baetz, B. Andrews and M. Snyder, 1997 SBF cell cycle regulator as a target of the yeast PKC-MAP kinase pathway. Science 275: 1781-1784. Martin, H., M. C. Castellanos, R. Cenamor, M. Sanchez, M. Molina et al., 1996 Molecular and functional characterization of a mutant allele of the mitogen-activated protein-kinase gene SLT2(MPK1) rescued from yeast autolytic mutants. Curr Genet 29: 516-522. McDonald, H. B., A. H. Helfant, E. M. Mahony, S. K. Khosla and L. Goetsch, 2002 Mutational analysis reveals a role for the C terminus of the proteasome subunit Rpt4p in spindle pole body duplication in Saccharomyces cerevisiae. Genetics 162: 705-720. 171 McVey, M., M. Kaeberlein, H. A. Tissenbaum and L. Guarente, 2001 The short life span of Saccharomyces cerevisiae sgs 1 and srs2 mutants is a composite of normal aging processes and mitotic arrest due to defective recombination. Genetics 157: 1531-1542. Mills, K. D., D. A. Sinclair and L. Guarente, 1999 MEC 1-dependent redistribution of the Sir3 silencing protein from telomeres to DNA double-strand breaks. Cell 97: 609-620. Moriya, H., and K. Isono, 1999 Analysis of genetic interactions between DHH1, SSD1 and ELM indicates their involvement in cellular morphology determination in Saccharomyces cerevisiae. Yeast 15: 481-496. Park, P. U., P. A. Defossez and L. Guarente, 1999 Effects of mutations in DNA repair genes on formation of ribosomal DNA circles and life span in Saccharomyces cerevisiae. Mol Cell Biol 19: 3848-3856. Pelissier, P., N. Camougrand, G. Velours and M. Guerin, 1995 NCA3, a nuclear gene involved in the mitochondrial expression of subunits 6 and 8 of the Fo-F1 ATP synthase of S. cerevisiae. Curr Genet 27: 409-416. Petes, T. D., and D. Botstein, 1977 Simple Mendelian inheritance of the reiterated ribosomal DNA of yeast. Proc Natl Acad Sci U S A 74: 5091-5095. Philippsen, P., M. Thomas, R. A. Kramer and R. W. Davis, 1978 Unique arrangement of coding sequences for 5 S, 5.8 S, 18 S and 25 S ribosomal RNA in Saccharomyces cerevisiae as determined by R-loop and hybridization analysis. J Mol Biol 123: 387-404. Puca, A. A., M. J. Daly, S. J. Brewster, T. C. Matise, J. Barrett et al., 2001A genome-wide scan for linkage to human exceptional longevity identifies a locus on chromosome 4. Proc Natl Acad Sci U S A 98: 10505-10508. Rempola, B., A. Kaniak, A. Migdalski, J. Rytka, P. P. Slonimski et al., 2000 Functional analysis of RRD1 (YIL153w) and RRD2 (YPL152w), which encode two putative activators of the phosphotyrosyl phosphatase activity of PP2A in Saccharomyces cerevisiae. Mol Gen Genet 262: 1081-1092. Rine, J., and I. Herskowitz, 1987 Four genes responsible for a position effect on expression from HML and HMR in Saccharomyces cerevisiae. Genetics 116: 9-22. Rosenwald, A. G., M. A. Rhodes, H. Van Valkenburgh, V. Palanivel, G. Chapman et al., 2002 ARL1 and membrane traffic in Saccharomyces cerevisiae. Yeast 19: 1039-1056. Rustchenko, E. P., and F. Sherman, 1994 Physical constitution of ribosomal genes in common strains of Saccharomyces cerevisiae. Yeast 10: 1157-1171. Sherman, F., and J. Hicks, 1991 Micromanipulation and dissection of asci. Methods Enzymol 194: 21-37. Sinclair, D. A., and L. Guarente, 1997 Extrachromosomal rDNA circles--a cause of aging in yeast. Cell 91: 1033-1042. 172 Smith, A. M., J. E. Archer and F. Solomon, 1998 Regulation of tubulin polypeptides and microtubule function: Luvlp [correction of Rkilp] interacts with the beta-tubulin binding protein Rbl2p. Chromosoma 107: 471-478. Smith, J. S., and J. D. Boeke, 1997 An unusual form of transcriptional silencing in yeast ribosomal DNA. Genes Dev 11: 241-254. Smith, J. S., C. B. Brachmann, I. Celic, M. A. Kenna, S. Muhammad et al., 2000 A phylogenetically conserved NAD+-dependent protein deacetylase activity in the Sir2 protein family. Proc Natl Acad Sci U S A 97: 6658-6663. Stettler, S., N. Chiannilkulchai, S. Hermann-Le Denmat, D. Lalo, F. Lacroute et al., 1993 A general suppressor of RNA polymerase I, II and III mutations in Saccharomyces cerevisiae. Mol Gen Genet 239: 169-176. Strunnikov, A. V., L. Aravind and E. V. Koonin, 2001 Saccharomyces cerevisiae SMT4 encodes an evolutionarily conserved protease with a role in chromosome condensation regulation. Genetics 158: 95-107. Sutton, A., D. Immanuel and K. T. Arndt, 1991 The SIT4 protein phosphatase functions in late G1 for progression into S phase. Mol Cell Biol 11: 2133-2148. Tadauchi, T., K. Matsumoto, I. Herskowitz and K. Irie, 2001 Post-transcriptional regulation through the HO 3'-UTR by Mpt5, a yeast homolog of Pumilio and FBF. Embo J 20: 552-561. Tissenbaum, H. A., and L. Guarente, 2001 Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature 410: 227-230. Uesono, Y., A. Toh-e and Y. Kikuchi, 1997 with RNA. J Biol Chem 272: 16103-16109. Ssdlp of Saccharomyces cerevisiae associates Vannier, D., P. Damay and D. Shore, 2001 A role for Sds3p, a component of the Rpd3p/Sin3p deacetylase complex, in maintaining cellular integrity in Saccharomyces cerevisiae. Mol Genet Genomics 265: 560-568. Watanabe, Y., K. Irie and K. Matsumoto, 1995 Yeast RLM1 encodes a serum response factor-like protein that may function downstream of the Mpkl (Slt2) mitogen-activated protein kinase pathway. Mol Cell Biol 15: 5740-5749. Wheeler, R. T., M. Kupiec, P. Magnelli, C. Abeijon and G. R. Fink, 2003 A Saccharomyces cerevisiae mutant with increased virulence. Proc Natl Acad Sci U S A 100: 2766-2770. Wilson, R. B., A. A. Brenner, T. B. White, M. J. Engler, J. P. Gaughran et al., 1991 The Saccharomyces cerevisiae SRK1 gene, a suppressor of bcyl and ins 1, may be involved in protein phosphatase function. Mol Cell Biol 11: 3369-3373. Yashiroda, H., T. Oguchi, Y. Yasuda, E. A. Toh and Y. Kikuchi, 1996 Bull , a new protein that binds to the Rsp5 ubiquitin ligase in Saccharomyces cerevisiae. Mol Cell Biol 16: 3255-3263. 173 TABLES AND FIGURES TABLE 1. Reported genetic interactions with SSDI-V. Gene Function Genetic Reference Interaction ARL1 Membrane trafficking ts BCK1 Protein kinase involved in cell wall integrity ts, other ROSENWALD et al. 2002 COSTIGAN et al. 1992 BEM2 GTPase activating protein required for polarized cell growth Protein that binds ubiquitin ligase. ts KIM etal. 1994 ts YASHIRODA et al. BULl CBK1 CCR4 CBC2 Protein kinase involved in cell morphogenesis Transcriptional regulator of glucose-repressed and cell wall genes. Component of cap binding complex. CDC28 CYR1 sl(mpt5), ts 1996 DU and NOVICK 2002 KAEBERLEIN and GUARENTE 2002 sl (stol) FORTES et al. 1999 Cyclin-dependent protein kinase Adenylate cyclase ts ts SUTTON et al. 1991 CLN1, CLN2 DHH1 G1 cyclins. G ATP-dependent RNA helicase sl1 (elml) ELM1 Serine/threonine protein kinase that regulates sl (dhhl) pseudohyphal growth Component of DNA end-joining repair pathway ts HDF1 JNM1 LAS 1 LUV 1 Protein required for proper nuclear migration during mitosis. Essential gene required for bud formation and morphogenesis. Protein involved in protein sorting in the late Golgi M. Kaeberlein and L. Guarente, unpublished CVRCKOVA and NASMYTH 1993 MORIYA and ISONO 1999 MORIYA and ISONO 1999 M. Kaeberlein and L. Guarente, unpublished ts WILSON et al. 1991 ts DOSEFF and ARNDT 1995 sl((rbl2) SMITH et al. 1998 MEP1, MEP2 Ammonium permeases involved in regulating pseudohyphal growth G LORENZ and HEITMAN 1998 MPT5 Protein required for high temperature growth and normal life span ts, sl(ccr4, swi4, swi6) KAEBERLEIN and GUARENTE 2002 174 Gene - -- ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~------Function Genetic Reference Interaction PAG1 Protein that functions with Cbklp to regulate cell morphogenesis D DU and NOVICK 2002 PDE2 3',5'-Cyclic-nucleotide phosphodiesterase misc WILSON et al. 1991 PPH21, Catalytic subunits of protein phosphatase 2A. ts, other PRP18 U5 snRNA-associated protein ts RBL2 Putative tubulin cofactor A sl(luvl) EVANS and STARK 1997 LUUKKONEN and SERAPHIN 1999 SMITH etal. 1998 RLMI1 Transcription factor downstream of MPK ts WATANABE et al. RPC31 RPC53 RPD3 RNA polymerase III RNA polymerase III Histone deacetylase component of the Rpd3pSin3p complex ts 1995 STETTLER etal. 1993 sl(swi6) STETTLER et al. 1993 VANNIER et al. 2001 PRT4 Proteosome ATPase ts MCDONALD et al. PPH22 2002 RRD 1 RRD2 SDS3 Phosphotyrosyl phosphatase activator Protein involved in rapamycin sensitivity Component of the Rpd3p-Sin3p histone deacetylase complex sl(rrd2) sl(swi6) REMPOLA et al. 2000 REMPOLA et al. 2000 VANNIER et al. 2001 SIN3 Component of the Rpd3p-Sin3p histone deacetylase complex sl(swi6) VANNIER et al. 2001 SIT4 Phosphatase required for GI -S transition. L SUTTON et al. 1991 SLG 1 Protein required for maintenance of cell wall integrity ts JACOBY et al. 1998 SLT2 ts MARTIN etal. 1996 SLY 1 MAP kinase that functions downstream of BCK1/SLK1 Protein involved in vesicle trafficking ts KOSODO et al. 2001 SMC2 Subunit of condensin protein complex ts sl(rrdl) STRUNNIKOV et al. 2001 SMC4 Subunit of condensin protein complex ts STRUNNIKOV et al. SNP1 U1 snRNA-associated protein ts 2001 LUUKKONEN and SERAPHIN 1999 STO 1 Component of cap binding complex. sl(cbc2) FORTES et al. 1999 SWI4 Cell cycle specific transcription factor. ts, sl(mpt5) KAEBERLEIN and GUARENTE 2002 175 Gene Function Genetic Reference Interaction SWI6 Cell cycle specific transcription factor. ts, sl(mpt5) U5AI U5 snRNA ts YPT1 GTP-binding protein involved in the secretory pathway GTP-binding protein involved in the secretory pathway ts KAEBERLEIN and GUARENTE 2002 LUUKKONEN and SERAPHIN 1999 LI and WARNER 1998 ts LI and WARNER 1998 YPT6 Abbreviations for genetic interactions are as follows: ts = suppresses temperature sensitivity, sl(x) = suppresses synthetic lethality between given gene and gene x, G = improves growth, L = suppresses lethality, D = causes death. 176 TABLE 2. Yeast strains used in this study Strain Relevant Genotype PSY316AR MATa ADE2::rDNI ade2-101 his3-A200 leu2-3,112 lys2-801 MKY2018 MKY2012 MK2017 MKY2076 MKY2077 MKY2085 MKY2086 MKY2147 MKY2153 MKY2020 MKY2021 MKY2152 MKY2149 MKY2152 MKD202 BKY5 PSY316AR pRS406/URA3 PSY316AR SSDI-V/URA3 PSY316AR SSDI-V/LEU2 PSY316AR ssdl-dA::HIS3 PSY316AR ssdl-dA::HIS3 SSD1-V/LEU2 PSY316AR sir2A::TRP1fobl A::LEU2pRS406/URA3 PSY316AR sir2A::TRP1fob A::LEU2 SSD1-V/URA3 PSY316AR cytlA::kanMX PSY316AR cytlA::kanMX SSD-V/URA3 PSY316AR mpt5A::LEU2 PSY316AR mpt5A::LEU2 SSD1-V/URA3 PSY316pADH_NCA3 PSY316AR nca3::HIS3 PSY316AR nca3::HIS3 SSD1-V/URA3 PSY316AR a/a sit4::HIS3/SIT4 SSD1-V/URA3 MATa adel-100 his4-519 leu2-3,112 lys2-801 ura3-52 uth4-14c ura3-52ssdl-d ssdl-d MKY3021 BKY5 SSD1-V/URA3 W303R MKY1324 MATa ADE2::rDNJ his3 leu2 trpl ura3 ssdl-d2 RAD5 W303R mpt5::LEU2 MKY 1332 W303R mpt5::LEU2 SSD1 -V/URA3 MKY183 MKY596 MKY606 MKY580 MKY574 MKY590 W303R sir2::TRPJ W303RpADH_MPT5 W303R sir2::TRPJ pADHMPT5 W303R rpd3::LEU2 W303R sir2::TRPJ rpd3::URA3 W303R mpt5::LEU2 rpd3::URA3 177 TABLE 3. Effects of MPT5, SIR2, and SSDI-V on silencing Genptype Telomere Silencing rDNA Silencing 4 4 4 1 mpt5A sir2A pADHMPT5 pADHMPT5 sir2A 4 4 4 ssdlA SSDI-V > > > Stated effects on silencing are relative to silencing in the wild type ssdl-d background. Telomere and rDNA silencing were measured by color formation using an ADE2 marker integrated into subtelomeric or ribosomal DNA. In order to detect decreased rDNA silencing relative to wild type in the sir2A strain, an rpd3A mutation was introduced into the wild type background. 178 Figure 1 - SSD1-V extends life span and improves cell integrity. (A) Life spans were determined for PSY316 (*), PSY316 SSDI-V (), PSY316 ssdlA::HIS3 (X), and PSY316 ssdlA::HIS3 SSD1-V (A). Mean life spans and number of cells analyzed were: PSY316 22.2 (n=40), PSY316 SSDI-V 36.2 (n=40), PSY316 ssdlA::HIS3 21.9 (n=40), and PSY316 ssdlA::HIS3 SSD1-V 38.0 (40). (B) Life spans were determined for BKY5 (*) and BKY5 SSD1-V (A). Mean life spans and number of cells analyzed were: BKY5 13.0 (n=40) and BKY5 SSDI-V 24.3 (n=40). (C) PSY316 cells are unable to grow at temperatures greater than 39° unless SSD1-V is present. 10-fold serial dilutions of a log phase culture plated onto YPD and incubated at either 30 ° or 40° for 48 hours are shown. 179 B. A. Generation Generation c. ~ 180 Figure 2. - SSDI -V and Sir2p act in different pathways to promote longevity. (A) Life spans were determined for PSY316 (*), PSY316 sir2fob (), PSY316 0.5% glucose (A), and PSY316 sir2fobl 0.5% glucose (X). Mean life spans and number of cells analyzed were: PSY316 22.8 (n=40), PSY316 sir2fobl 21.2 (n=39), PSY316 0.5% glucose 27.0 (n=40), and PSY316 sir2fobl 0.5% glucose 21.4 (40). (B) Life spans were determined for PSY316 (*), PSY316 sir2fob (), PSY316 SSDI-V (A), and PSY316 sir2fobl SSDI-V (X). Mean life spans and number of cells analyzed were: PSY316 22.8 (n=40), PSY316 sir2fobl 21.2 (n=39), PSY316 SSDI-V 37.3 (n=40), and PSY316 sir2 fobl SSDI-V 30.0 (40). (C) Life spans were determined for PSY316 (*), PSY316 SSDI-V (), PSY316 cyt (), and PSY316 SSDI-Vcytl (X). Mean life spans and number of cells analyzed were: PSY316 21.7 (n=40), PSY316 SSDI-V39.3 (n=40), PSY316 cytl 22.2 (n=40), and PSY316 SSDI-Vcytl 181 37.9 (40). B. A. 0. jII0. 25 III - iIs io0. 0. 50 40 30 20 10 0 Generation C" 0.8 3 0.6 E 0.4 0.2 0 10 20 30 40 Generation 50 60 70 80 182 0 10 20 40 30 Generation 50 60 70 Figure 3. - SSD1-V does not extend life span by decreasing ERC levels. (A) SSD1-Vhas no detectable effect on rDNA recombination. rDNA recombination was measured by the frequency at which an ADE2 marker integrated into the rDNA is lost. A total of 33,000 colonies from 3 independent derived isolates was examined for each strain. (B) DNA from unsorted cells was isolated and electrophoresed as described (Kaeberlein et al. 1999). The gel was transferred and probed with sequence homologous to the rDNA. Extrachromosomal rDNA circles (ERCs) are denoted by arrowheads. The dark band present in both lanes corresponds to genomic rDNA. In this isolate, the long-lived SSDIVcells have a greater steady-state amount of ERCs than ssdl-d cells, as quantitated by the ratio of ERC DNA to genomic rDNA. However, in one independently derived SSDIV transformant we were unable to detect a significant difference in ERC levels relative to wild type cells. In no case, did we detect fewer ERCs in SSDI-V cells. 183 A. B. 2.5 --.. 0 0 0 x ......... >. u c: 2 1.5 Q) :J rr Q) ~ u. (f) (f) 0 .....J ~ ~~ ro 0.5 ~ Q) a ssdl -d SSOl -v "'0 ...... I "'0 en en 184 > ...... Cl en en I Figure 4. - Effect of NCA3 cell life span. (A) Life spans were determined for PSY316 (*) and PSY316 pADH_NCA3 (U). Mean life spans and number of cells analyzed were: PSY316 23.4 (n=80) and PSY316pADH_NCA3 27.6 (n=120). This figure represents data pooled from two different experiments. (B) Life spans were determined for PSY316 (*), PSY316 nca3 (U), PSY316 SSDI-V (A) and PSY316 nca3 SSDI-V(X). Mean life spans and number of cells analyzed were: PSY316 23.3 (n=40), PSY316 nca3 22.3 (n=40), PSY316 SSD1-V 37.5 (n=40) and PSY316 nca3 SSDI-V 32.8 (n=40). 185 B. A. 0.8 z 0.6 > i .2 2 0.4 0.2 0 0 10 30 20 50 40 Generation 186 10 20 30 Generation 40 50 60 70 Figure 5. - Microarray Analysis of long-lived cells. (A) Gene expression profiles for long-lived cells were clustered over the 55 most highly expressed genes previously identified as being regulated by CR (Lin et al. 2002). (B) Identical clustering as in A, except over all 124 genes regulated by CR. The CR group clusters most closely together and consists of cells grown on 0.5% glucose or deleted for HXK2. Life span extension by CR, HAP4-overexpression or 20% glucose is SIR2 -dependent, and these data sets cluster together on one of the two primary branches. Cells containing a single copy of SSDI -V cluster separately from the other long-lived cell types. (C) Comparison of the overlap in co-regulated genes between long-lived cell types. 55/124 genes transcriptionally altered by CR are similarly altered by overexpression of HAP4. This overlap is statistically significant (p < 10-4°). 48/124 genes transcriptionally altered by CR are similarly altered by growth in 20% glucose. This overlap is statistically significant (p < 10-50). 8/124 genes transcriptionally altered by CR are similarly altered by introduction of SSD1-V. This overlap is not statistically significant (p = 0.1). 187 B. A. nR412W YBR012W-B SRL3 mH1 E0l23 ERG4 0IS5 Yn089W YOL003C SDS3 nR45411 PHO'O HGG1 YI!L00311 S1iR1 c. DR525W DR2.2C DR290W YEROO2W nR44fiW YUR306C YHL311C DL094C SHU23 HAP4 CR 20% GLU CR SSDI-V CR CTF4 SOF1 J.lET13 RPL318 KAP122 JmPL28 DR314C DR543C YU.062C SIT 1 YUR3.2W nR321C T1511 HXK1 YPLOO5W YUR3I3C JRH1 YPLO.IV DR534C »!R103C YGL133V TFB3 !'US 1 Y.:JL213W LIF1 }IF t 1LPHJl) 2 SAG1 ~1 YER028C DtR095V ..J HXTl 188 Figure 6. - MPT5 and SIR2 determine longevity in a pathway parallel to SSD1-V. (A) Life spans were determined for PSY316 (*), PSY316 SSDI-V (U), PSY316 mpt5 (), and PSY316 mpt5 SSD1-V (X). Mean life spans and number of cells analyzed were: PSY316 24.3 (n=40), PSY316 SSDI-V35.5 (n=40), PSY316 mpt5 15.3 (n=40), and PSY316 mpt5 SSDI-V 22.9 (40). (B) Life spans were determined for W303R (), W303R pADH_MPT5 (), W303R sir2 (A), W303R sir2 pADHMPT5 (X), W303R SIR2/URA3 (*), and W303R SIR21URA3 pADHMPT5 (0). This experiment was performed in the W303R strain background because overexpression of MPT5 from the pADH_MPT5 plasmid causes slow growth and decreased viability in strain PSY316. Mean life spans and number of cells analyzed were: W303R 20.9 (n=41), W303RpADHMPT5 25.8 (n=70), W303R sir2 11.2 (n=41), W303R sir2 pADH_MPT5 10.1 (n=70), W303R SIR2/URA3 27.4 (n=40), and W303R SIR2/URA3 pADH_MPT5 25.8 (n=40). 189 LA .Z. 3~. R. 1 0.8 1 0.8 c) 7 0.6 .Lg - 0.6 S E! 0.4 i 0.4 O.Z 0.Z 0 N 0 10 20 30 40 Generation 50 60 70 10 30 20 Generation 190 40 50 Figure 7. - Genetic model for Ssdlp as a regulator of longevity. Ssdlp acts parallel to Sir2p to extend life span. This could involve increasing the cell's resistance to ERCs or could represent an ERC-independent longevity promoting function. One likely possibility is that SSDI-Vresults in an altered cell wall structure that allows mother cells to achieve extreme old age. 191 Caloric Restriction Mpt5p Sir2p Ssd p Cell wall? ERC 0 Longevity 192 APPENDIX B Caloric Restriction Extends Yeast Life Span By Lowering the Level of NADH This chapter was published in Genes and Development, Volume 18, Issue 1, pages 12-16. The authors were Su-Ju Lin, Ethan Ford, Marcia Haigis, Gregory Liszt, and Leonard Guarente. I contributed to the generation of the data presented in Figure 4. 193 SUMMARY Calorie restriction extends life span in a wide variety of species. Previously, we showed that calorie restriction increases the replicative life span in yeast by activating Sir2, a highly conserved NAD-dependent deacetylase. Here we test whether CR activates Sir2 by increasing the NAD/NADH ratio or by regulating the level of nicotinamide, a known inhibitor of Sir2. We show that CR decreases NADH levels, and that NADH is a competitive inhibitor of Sir2. A genetic intervention that specifically decreases NADH levels increases life span, validating the model that NADH regulates yeast longevity in response to CR. 194 INTRODUCTION Calorie restriction (CR) extends life span in a wide spectrum of organisms and for decades was the only regimen known to promote longevity in mammals (Weindruch and Walford 1998; Roth et al. 2001). CR has also been shown to delay the onset or reduce the incidence of many age-related diseases including cancer and diabetes (Weindruch and Walford 1998; Roth et al. 2001). Although it has been suggested that CR may work by reducing the levels of reactive oxygen species due to a slowing in metabolism (Weindruch and Walford 1998), the mechanism by which CR extends life span is still uncertain. To study the mechanism by which CR extends life span, we established a model of CR in the budding yeast Saccharomyces cerevisiae. In this system, life span can be extended by limiting glucose content in the media from 2% to 0.5% or by reducing the activity of the glucose-sensing cAMP-dependent kinase (PKA) (Lin et al. 2000). The benefit of CR requires NAD (nicotinamide adenine dinucleotide, oxidized form) and Sir2 (Lin et al. 2000; Lin et al. 2002), a key regulator of life span in both yeast and animals (Kaeberlein et al. 1999; Tissenbaum and Guarente 2001). Sir2 is an NAD- dependent histone deacetylase and is required for chromatin silencing and life span extension (Imai et al. 2000; Landry et al. 2000; Smith et al. 2000). The requirement of NAD for Sir2 decaetylase activity suggested CR may activate Sir2 by increasing the available pool of NAD for Sir2 (Guarente 2000). Our previous studies suggested that there was a second mechanism to activate 195 Sir2 and extend the life span in yeast - osmotic stress (Kaeberlein and Guarente 2002). This stress mechanism was genetically distinguishable from CR. Mutations knocking out electron transport prevented CR from extending life span, but had no effect on the longevity conferred by osmotic stress (Lin et al. 2002; Kaeberlein and Guarente 2002). This requirement for electron transport during CR is because of a shunting of carbon metabolism from fermentation to the mitochondrial TCA cycle. The concomitant increase in respiration is necessary and sufficient for the activation of Sir2-mediating silencing and extension in life span (Lin et al. 2002). The fact that respiration produces NAD from NADH (Bakker et al. 2001), reinforces the idea that changes in the NAD/NADH ratio regulates Sir2 during CR. A recent report, however, challenged this model by claiming that both stress and CR activated Sir2 by a different mechanism, namely by decreasing intracellular levels of nicotinamide (Anderson et al., 2003), a non-competitive inhibitor of Sir2 (Bitterman et al. 2002). This change in nicotinamide levels was triggered by activation of the PNC1 gene, encoding a nicotinamidase in the NAD salvage pathway (Ghislain et al. 2002). This gene was strongly activated by osmotic stress, but to a lesser degree by CR. Also consistent with this model, PNCI was shown to be required for the extension in life span by CR (Anderson et al. 2003). The authors conclude that CR reduces nicotinamide levels by upregulating PNC1, and this reduction activates Sir2 and extends the life span. In this report we attempt to determine whether changes in nicotinamide, the NAD/NADH ratio, or both up-regulate Sir2 during CR. We show that NADH is a 196 competitive inhibitor of Sir2, and that CR reduces the level of NADH in cells. These findings provide a simple model for activation of Sir2 and extension of the life span by CR, which we further support by genetic experiments. RESULTS AND DISCUSSION To study whether CR activates Sir2 by increasing the NAD/NADH ratio, we first measured the NAD and NADH levels in cells grown in 2% and 0.5% glucose (CR). Surprisingly, the NAD levels in cells under CR were not changed (Fig. 1A), whereas the NADH levels were decreased to 50% of non-CR cells, (Fig. B). To validate the efficacy and sensitivity of our assay, we measured the NAD levels of the nptlA mutant. It has been shown that cells lacking the NPT1 gene (encoding nicotinic acid phosphoribosyl transferase in the NAD salvage pathway) exhibit a 40% to 60% decrease in NAD levels (Smith et al. 2000; Anderson et al. 2002). Consistent with previous reports (Smith et al. 2000; Anderson et al. 2002), we detected a similar decrease in the NAD levels in nptlA mutants. As a further test, we measured NAD and NADH levels in cells that overexpress the transcription factor Hap4, which activates nuclear genes encoding mitochondrial proteins (Forsburg and Guarente 1989; de Winde and Grivell 1993). This manipulation was shown to extend life span in a Sir2 dependent manner (Lin et al. 2002). Similar to CR, Hap4 over-expression causes a switch of metabolism from fermentation toward respiration (Blom et al. 2000) and, as shown in Fig. B, triggered a 50% decrease in NADH levels without altering NAD concentration (Fig. A). Since respiration is required for life span extension in CR, we tested whether a functional electron transport chain is required for the decrease in NADH levels. As shown in Fig. D, deleting the 197 CYTI gene, which encodes the cytochrome c , abolished the decrease in NADH levels induced by CR. All these studies suggest that CR increases the intracellular NAD/NADH ratio by up-regulating respiration thereby decreasing NADH levels. We also measured NAD and NADH levels in a sir2A mutant grown in 2% and 0.5% glucose. The sir2A mutant exhibited the normal reduction in NADH levels in response to CR (shown in Fig. 1C, D), indicating that it functions downstream of the metabolic changes. Since NADH levels responded to CR but NAD levels did not, it seemed possible that NADH, and not NAD, regulated Sir2, perhaps by inhibiting its activity. To test this hypothesis, we measured activity of Sir2 in the presence of various concentrations of NADH using purified recombinant GST-tagged Sir2 proteins. Sir2 activity was determined by quantitating the NAD-dependent deacetylation of histone H4 peptides labeled with [3 H]-acetyl coenzyme A (Armstrong et al. 2002). If NADH is indeed a competitive inhibitor of Sir2 activity, the presence of NADH should increase the apparent KM(the Michaelis-Menten binding constant) of Sir2 for NAD without affecting the Vmax (maximum velocity) (Stryer 1995). Kinetic analysis with a wide range of substrate (NAD) concentrations is shown as a Lineweaver-Burk double reciprocal plot (Fig. 2A). This analysis estimates the KM(Fig. 2A, X intercept) of Sir2 to be 30 yM, consistent with previous report (Imai et al. 2000; Tanner et al. 2000). As NADH concentrations were stepped up, the apparent KMfor NAD was increased, reaching an increase of 3 fold at 250 FM NADH. The Vmax(Fig. 2A, Y intercept) of Sir2 activity was not significantly changed in the presence of NADH. These data suggest NADH functions as a competitive 198 inhibitor of Sir2. As shown in Fig. 2B, NADH also competitively inhibited the human SIRT1. In the presence of 300 FM NADH, the KMof SIRT1 increased 3 fold. Intracellular concentrations of total NAD plus NADH of about 1mM have been reported (de Koning and van Dam 1992; Richard et al. 1993). We have calculated values for NAD and NADH from the data in Fig 1, assuming a yeast cell size of 70 mm3 (Guthrie and Fink 2002). Table 1 shows our values and those of other recent reports. The data are in reasonable agreement with the previous reports and with each other, and small differences may by partly due to different estimates of cell size. The concentration of free NADH in cells must be less than the values in Table 1, and appears to be in a sensitive range to regulate Sir2 activity. We sought genetic data to support the idea that the increase in the NAD/NADH ratio is what extends life span in CR. During respiratory growth, both cytosolic and mitochondrial NADH are re-oxidized, in part by the NADH dehydrogenases in the respiratory chain (Bakker et al. 2001). To determine whether yeast life span could be extended by simply activating NADH dehydrogenase, we over-expressed two related mitochondrial NADH dehydrogenases, Ndel and Nde2 (Luttik et al. 1998). Similar to CR, over-expressing Ndel and Nde2 significantly decreased the NADH levels without changing the NAD levels (Fig. 3A). Moreover, cells over-expressing Ndel and Nde2 exhibited a longer life span on 2% glucose to a degree similar to cells grown on 0.5% glucose (Fig. 3B). Further, 0.5% glucose did not extend the life span of cells overexpressing the NADH dehydrogenases (Fig. 3B), suggesting that CR and the NADH 199 dehydrogenases function in the same pathway, i.e. to decrease NADH levels and extend the life span. Our findings appear to challenge the claim that the Pncl nicotinamidase triggers the life span extension in CR (Anderson et al. 2003). To address this paradox, we first repeated life span analysis of the pnclA mutants on 2% and 0.5% glucose. Consistent with the previous report, the pnclA mutation largely prevented the life span extension by 0.5% glucose (at best a 10-15% increase) (Fig. 4A) when compared with wild type cells (~30% increase) (Fig. 4B). Since the above data indicated that CR functions by decreasing NADH, we surmised that hyper-accumulation of nicotinamide in the pnclA mutants might mask regulation by NADH. Nntl, a putative nicotinamide methyl transferase, appears to modify nicotinamide in yeast (Anderson et al. 2003). Thus, we over expressed Nntl to reduce nicotinamide levels in the pnclA mutant. Strikingly, overexpressing Nntl restored the ability of CR to extend life span (30%) in pnclA mutants (Fig. 4B). These data show that the reduction in NADH can activate Sir2 and give a full extension of the life span in a pnclA mutant, as long as the excess nicotinamide is depleted. This result shows that PNCI, the NAD salvage pathway, and glucosedependent changes in nicotinamide levels are not required for the extension of life span by CR. Our studies show that a switch to oxidative metabolism during CR increases the NAD/NADH ratio by decreasing NADH levels. NADH is a competitive inhibitor of Sir2, implying that a reduction in this dinucleotide activates Sir2 to extend the life span in 200 CR. Indeed, over-expression of the NADH dehydrogenase specifically lowers NADH levels and extends the life span, providing strong support for this hypothesis. Regulation of the life span by NADH is also consistent with the earlier finding that electron transport is required for longevity during CR (Lin et al., 2002). The NAD/NADH ratio reflects the intracellular redox state and is a read out of metabolic activity. Our findings suggest that this ratio can serve a critical regulatory function, namely the determination of the life span of yeast mother cells. It remains to be seen whether this ratio will serve related regulatory functions in higher organisms. MATERIALS AND METHODS Yeast strain PSY316 MATa ura3-52 leu2-3, 112 his3-A200 ade2-101 lys2-801 RDN1.::ADE2has been previously described (Park et al. 1999). Rich media YPD and synthetic media were made as described (Sherman et al. 1978). The ADHI -driven integrating ppp35 (URA3) vector is a derivative of ppp81 (LEU2) vector (Lin et al. 2002) made by Peter Park. Over-expression constructs of Ndel1, Nde2 and Nntl were made as described previously (Lin et al. 2002) using ppp81 (Ndel) or ppp35 (Nde2 and Nntl). All constructs made for this study were verified by sequencing. The yeast Sir2 expression construct pGEX-SIR2 was a gift from J. Tanny and D. Moazed at Harvard Medical School (Tanny et al. 1999). The human SIRT1 expression construct pGEX-GST-hSIRTl was a gift from F. Ishikawa at Tokyo Institute of Technology, Japan (Takata and Ishikawa 2003). All gene deletions in this study were done by replacing the wild type genes with the Kanr marker as described in (Guldener et al. 1996) and verified by Polymerase Chain Reaction (PCR) using oligonucleotides flanking the genes of interest. 201 Life span analyses were carried out as previously described (Lin et al. 2000). All life span analyses in this study were carried out on YPD plates at least twice independently with more than 45 cells per strain per experiment. Results from a single experiment are shown. Yeast GST-Sir2 and human GST-SIRT1 fusion proteins were made using the T7 expression system as previously described (Ford and Hernandez 1997) with the following modifications. pGEX-SIR2 and pGEX-GST-hSIRTl were transformed into the E. coli strain BL21(DE3) Codon-Plus-RIL (Stratagene) and BL21(DE3) respectively. In addition, the procedure for yeast GST-Sir2 was scaled up to a 10 liter culture and 5 ml of glutathione-agarose resin. The purified protein was concentrated on a Centriprep-30 (Amacon) column. The deacetylation assay was carried out as described previously (Armstrong et al. 2002). In brief, histone deacetylase activity was measured using a peptide corresponding to the N-terminal tail of the histone H4 (Upstate Biotechnology) labeled with tritiated Acetyl Coenzyme A (PerkinElmer). Reactions were carried out in 50 yl buffer containing 50 mM Tris-HCl, pH 8, 4 mM MgCl 2, 0.2 mM DTT and a variable concentration of NAD and NADH at 30°C for 2.5 hours (yeast Sir2) or 37°C for 2 hours (human SIRT1). Measurement of the NAD and NADH nucleotides were performed as described previously (Lin et al. 2001) with a few modifications. In brief, cells were grown to an OD6 00 of 0.5 then 107 cells were harvested in duplicates by centrifugation in 2 x 1.5 ml tubes. Acid extraction was performed in one tube to obtain NAD and alkali extraction was performed in the other to obtain NADH. 2 l of neutralized cell extract (105 cells) was used for enzymatic cycling reaction as previously described (Lin et al. 2001). The concentration of nucleotides was measured fluorometrically with excitation at 365 nm 202 and emission monitored at 460 nm. Standard curves for determining NAD and NADH concentrations were obtained as follows: NAD and NADH were added into the acid and alkali buffer to a final concentration of 0 pM, 2.5 pM, 5 pM, 7.5 yM which were then treated with the same procedure along with other samples. The fluorometer was calibrated each time before use with 0 FM, 5 FM, 10 uM, 20 pM, 30 pM and 40 pM NADH to ensure the detection was within a linear range. Acknowledgment We thank members of the Guarente laboratory for discussions; J. Gordon for advice with the NAD nucleotide measurements; J. Tanny and D. Moazed for the yeast Sir2 expression construct and discussions; F. Ishikawa for the human SIRT1 expression construct. Supported by NIH, the Ellison Medical Foundation, the Seaver Institute, and the Howard and Linda Stern Fund (LG); Individual National Research Service Award (S.-J. L., E. F., M. H.). 203 REFERENCES Anderson, R.M., K.J. Bitterman, J.G. Wood, O. Medvedik, H. Cohen, S.S. Lin, J.K. Manchester, J.I. Gordon, and D.A. Sinclair. 2002. Manipulation of a nuclear NAD+ salvage pathway delays aging without altering steady-state NAD+ levels. J Biol Chem 277: 18881-90. Anderson, R.M., K.J. Bitterman, J.G. Wood, O. Medvedik, and D.A. Sinclair. 2003. Nicotinamide and PNC1 govern lifespan extension by calorie restriction in Saccharomyces cerevisiae. Nature 423: 181-5. Armstrong, C.M., M. Kaeberlein, S.I. Imai, and L. Guarente. 2002. Mutations in Saccharomyces cerevisiae gene SIR2 can have differential effects on in vivo silencing phenotypes and in vitro histone deacetylation activity. Mol Biol Cell 13: 1427-38. Bakker, B.M., K.M. Overkamp, A.J. van Maris, P. Kotter, M.A. Luttik, J.P. van Dijken, and J.T. Pronk. 2001. Stoichiometry and compartmentation of NADH metabolism in Saccharomyces cerevisiae. FEMS Microbiol Rev 25: 15-37. Bitterman, K.J., R.M. Anderson, H.Y. Cohen, M. Latorre-Esteves, and D.A. Sinclair. 2002. Inhibition of silencing and accelerated aging by nicotinamide, a putative negative regulator of yeast Sir2 and human SIRT1. J Biol Chem 23: 23. Blom, J., M.J. De Mattos, and L.A. Grivell. 2000. Redirection of the respiro-fermentative flux distribution in Saccharomyces cerevisiae by overexpression of the transcription factor Hap4p. Appl Environ Microbiol 66: 1970-3. de Koning and K. van Dam. 1992. A method for the determination of changes of glycolytic metabolites in yeast on a subsecond time scale using extraction at neutral pH. Anal Biochem 204:118-23 de Winde, J.H. and L.A. Grivell. 1993. Global regulation of mitochondrial biogenesis in Saccharomyces cerevisiae. Prog Nucleic Acid Res Mol Biol 46: 51-91. Ford, E. and N. Hernandez. 1997. Characterization of a trimeric complex containing Oct-1, SNAPc, and DNA. JBiol Chem 272: 16048-55. Forsburg, S.L. and L. Guarente. 1989. Identification and characterization of HAP4: a third component of the CCAAT-bound HAP2/HAP3 heteromer. Genes Dev 3:1166-78. Fulco, M., R.L. Schiltz, S. ezzi, M.T. King, P. Zhao, Y. Kashiwaya, E. Hoffman, R.L. Veech, and V. Sartorelli. 2003. Sir2 regulates skeletal muscle differentiation as a potential sensor of the redox state. Mol Cell 12: 51-62. 204 Ghislain, M., E. Talla, and J.M. Francois. 2002. Identification and functional analysis of the Saccharomyces cerevisiae nicotinamidase gene, PNC1. Yeast 19: 215-24. Guarente, L. 2000. Sir2 links chromatin silencing, metabolism, and aging. Genes Dev 14: 1021-6. Guldener, U., S. Heck, T. Fielder, J. Beinhauer, and J.H. Hegemann. 1996. A new efficient gene disruption cassette for repeated use in budding yeast. Nucleic Acids Res 24: 2519-24. Guthrie, C. and G.R. Fink. 2002. Guide to Yeast Genetics and Molecular and Cellular Biology. In Methods in Enzymology. Academic Press. Imai, S., C.M. Armstrong, M. Kaeberlein, and L. Guarente. 2000. Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature 403: 795-800. Kaeberlein, M., M. McVey, and L. Guarente. 1999. The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev 13: 2570-80. Kaeberlein, M., A.A. Andalis, G.R. Fink, and L. Guarente. 2002. High osmolarity extends life span in Saccharomyces cerevisiae by a mechanism related to calorie restriction. Mol Cell Biol 22: 8056-66. Landry, J., A. Sutton, S.T. Tafrov, R.C. Heller, J. Stebbins, L. Pillus, and R. Sternglanz. 2000. The silencing protein SIR2 and its homologs are NAD-dependent protein deacetylases. Proc Natl Acad Sci U S A 97: 5807-11. Lin, S.-J., M. Kaeberlein, A.A. Andalis, L.A. Sturtz, P.-A. Defossez, V.C. Culotta, G.R. Fink, and L. Guarente. 2002. Calorie restriction extends life span by shifting carbon toward respiration. Nature 418: 344-8. Lin, S.J., P.A. Defossez, and L. Guarente. 2000. Requirement of NAD and SIR2 for life-span extension by calorie restriction in Saccharomyces cerevisiae. Science 289: 2126-8. Lin, S.J. and L. Guarente. 2003. Nicotinamide adenine dinucleotide, a metabolic regulator of transcription, longevity and disease. Curr Opin Cell Biol 15: 241-6. Lin, S.S., J.K. Manchester, and J.I. Gordon. 2001. Enhanced gluconeogenesis and increased energy storage as hallmarks of aging in Saccharomyces cerevisiae. J Biol Chem 276: 36000-7. Luttik, M.A., K.M. Overkamp, P. Kotter, S. de Vries, J.P. van Dijken, and J.T. Pronk. 1998. The Saccharomyces cerevisiae NDE1 and NDE2 genes encode separate mitochondrial NADH dehydrogenases catalyzing the oxidation of cytosolic NADH. J Biol Chem 273: 2452934. 205 Park, P.U., P.A. Defossez, and L. Guarente. 1999. Effects of mutations in DNA repair genes on formation of ribosomal DNA circles and life span in Saccharomyces cerevisiae. Mol Cell Biol 19: 3848-56. Richard, P., B. Teusink, H.V. Westerhoff, and K. van Dam. 1993. Around the growth phase transition S. cerevisiae's make-up favours sustained oscillations of intracellular metabolites. FEBS Lett 318: 80-2 Roth, G.S., D.K. Ingram, and M.A. Lane. 2001. Caloric restriction in primates and relevance to humans. Ann N YAcad Sci 928: 305-15. Rutter, J., M. Reick, L.C. Wu, and S.L. McKnight. 2001. Regulation of clock and NPAS2 DNA binding by the redox state of NAD cofactors. Science 293: 510-4. Sherman, F., G.R. Fink, and C.W. Lawrence. 1978. Methods in Yeast Genetics. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Smith, J.S., C.B. Brachmann, I. Celic, M.A. Kenna, S. Muhammad, V.J. Starai, J.L. Avalos, J.C. Escalante-Semerena, C. Grubmeyer, C. Wolberger, and J.D. Boeke. 2000. A phylogenetically conserved NAD+-dependent protein deacetylase activity in the Sir2 protein family. Proc Natl Acad Sci U S A 97: 6658-63. Stryer, L. 1995. Biochemistry. W. H. Freeman and Company, New York. Takata, T. and F. Ishikawa. 2003. Human Sir2-related protein SIRT1 associates with the bHLH repressors HES 1 and HEY2 and is involved in HES 1- and HEY2-mediated transcriptional repression. Biochem Biophys Res Commun 301: 250-7. Tanner, K.G., J. Landry, R. Sternglanz, and J.M. Denu. 2000. Silent information regulator 2 family of NAD- dependent histone/protein deacetylases generates a unique product, 1-Oacetyl-ADP-ribose. Proc Natl Acad Sci U S A 97: 14178-82. Tanny, J.C., G.J. Dowd, J. Huang, H. Hilz, and D. Moazed. 1999. An enzymatic activity in the yeast Sir2 protein that is essential for gene silencing. Cell 99: 735-45. Tissenbaum, H.A. and L. Guarente. 2001. Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature 410: 227-30. Weindruch, W. and R.L. Walford. 1998. The retardation of aging and diseases by dietary restriction. Charles C. Thomas, Springfield, Illinois, USA. Zhang, Q., D.W. Piston, and R.H. Goodman. 2002. Regulation of corepressor function by nuclear NADH. Science 295: 1895-7. 206 TABLES AND FIGURES Table 1: Intracellular concentrations of NAD and NADH. Normal condition (mM) NAD NADH Calorie restriction (mM) NAD/NADH ratio NAD NADH NAD/NADH ratio 2.14 a* 1.14 - 2-3 - 2C 0.78 1.26 0.0 6 d 0.85 0.13 a Reported unit was b 1.14~~~~~~~~~~~~~~~~~~~b* ~Ashrafi - 2-3 Smith et al 2000 etal 2000 Lin et al 2001 - 2.56 1.48 References - 1.19 0.08 0.39 0.11 - Anderson et al 2002 3.05 This study 1.5x10-4 pmole/cell. Reported unit was about 80 amole/cell. c Reported units were 23.7 amole/pg protein and 9.3 amole/pg protein for NAD and NADH respectively. *We converted reported units (a to c) into mM (per cell) assuming a yeast cell size of 70 femtoliter (70 jum3) and the total protein of 6 pg/cell (Guthrie and Fink 2002). d Numbers show mean ±+standard deviations derived from data shown in Fig 1. 207 Figure 1. Calorie restriction decreases intracellular NADH levels. Measurements of intracellular levels of NAD (A and C) and NADH (B and D) in various yeast strains grown in 2% or 0.5% glucose. Results show average of 3 independent experiments each conducted in duplicates. Error bars denote standard deviations. NAD and NADH levels are shown as (pM/p,g): levels of NAD or NADH (PM) normalized to the concentrations of proteins from the extracts (pg) present in the reaction. Reactions contain cell extracts from 105 cells. WT, wild type yeast strain PSY316; +Hap4-oe: wild type cells over-expressing Hap4; 2%: cells grown in 2% glucose; 0.5%: cells grown in 0.5% glucose. nptl~, cytl ~ and sir2~hmr 1~ are isogenic derivatives of PSY316. B A 0.75 i 0.5 C) :::L ..-. T ..-. J. 0 '# 1 9 c. ~ co :c D T T T T T T T T T T It) ..-. i 0.5 . ~ CD 9 Xca :c + 1 C) C) :::L "'-' 0.75 T 0.5 ::J: C « Z ,. 0 ~ . 0.75 . 1 0.25 '# C'l + T - :::L "'-' 0.5 CD ~ i ~ :::L "'-' « Z C'l ..-. 0.75 C C 0.25 C C) ::J: "'-' « Z Lin_Figure 1 1 1 0.25 - 0 - C « Z T T T ""'I'" . '# C'l ~ #It) 0 ~ . '#. #C'l It) 0 #C'l . '# 1I?~C'l #- .'# II? <f~o ~ o ... <f ... & <f~-~<f -OJe~& .c: .c:e ~ ~ ~ 208 0.25 0 #- iP- It)~C'l II? . <f~o ~.-. <f _OJe'O)~ .c: e u .c: Figure 2. NADH inhibits Sir2 NAD-dependent histone deacetylase activity. (A) Kinetic analysis of the Sir2 NAD-dependent histone deacetylase activities in the presence of 0 /M (), 50 AM (2), 100 pM () or 250 pM (O) NADH. 150 ng of recombinant GST-tagged yeast Sir2 protein was assayed with various concentrations of NAD and NADH at 30°C for 2.5 hrs. Data are shown as a Lineweaver-Burk double reciprocal plot of 1/V (CPM/hr) versus 1/[NAD] (M). Results show average of three independent experiments each measured in duplicates. (B) Kinetic analysis of the human SIRT1 NAD-dependent histone deacetylase activities in the presence of 0 PM (U), 50 pM (01), 150 FM () or 300 pM (O) NADH. Experiments were carried out as in (A), except that 50 ng of recombinant GST-tagged human SIRT1 protein (Takata and Ishikawa 2003) was used and reaction was carried out at 37°C for 2 hrs. B A LinFigure 2 C O D :I 0 1/[NAD] (M) 1/[NAD] (M) 209 Figure 3. Over-expressing NADH dehydrogenase increases the intracellular NAD/NADH ratio and life span. (A) Measurements of NADH (closed bars, left panel) and NAD (open bars, right panel) levels in cells over-expressing Ndel and Nde2. Results show average of 4 independent experiments each conducted in duplicates. Error bars denote standard deviations. (B) Life span analysis of cells over expressing Ndel and Nde2 grown in 2% and 0.5% glucose. Average life spans on 2% glucose: +vectors, 19.4; +Ndel-oe, +Nde2-oe: 27.2. Average life spans on 0.5% glucose: +vectors, 25.8; +Ndel-oe, +Nde2-oe: 27.1. + vectors: cells carrying control vectors ppp35 and ppp81; +Ndel-oe, +Nde2-oe: cells over-expressing Ndel and Nde2. A Lin_Figure 3 1 1 .-. .-. ~ ~ en 0.7 ~ en 1 1 0.7 :L 0.5 "'-' :t: C c( Z C c( Z 0.2 0.2 0 +vectors +Ndel-oe +Nde2-oe +vectors +Nde1-oe +Nde2-oe B ___ .!l. +vectors, .2% -0- +vectors, 0.5% -e-- +Ndel-oe, Nde2-oe,2% --cr- +Ndel-oe, Nde2-oe, 0.5% 1 Gi () .!0.7 .a .! > ....o c o ;: g 0.2 a- U. .,.. Generations 210 Figure 4. Calorie restriction in pnclA mutants. (A) Calorie restriction slightly extends life span in a pnclA mutant. (B) Over-expressing Nntl restores the full life span extension by calorie restriction in a pnclA mutant. Average life spans on 2% glucose: pnclA + vector, 20.9; WT + vector, 20.16; Apncl+Nntl-oe, 21.6. Average life spans on 0.5% glucose: pnclA + vector 23.9; WT + vector, 27.3; pnclA +Nntl-oe, 28. WT: wild type PSY316 cells; + vector: cells carrying a control vector ppp35; + Nntl -oe: cells over-expressing Nntl. pnclA was derived from PSY316. LinFigure 4 ~~~~~A A0 0 3 ao a- .1 0 a- U. 0 10 20 30 40 50 Generations B 0 ', 2% 0 e,0.5% 0 0 eAs a c U W i) IL 0 10 20 30 40 Generations 211 50 APPENDIX C The Mammalian Sir2 Protein SIRT7 is an Activator of RNA Polymerase I Transcription This chapter has been submitted for publication. The authors are Ethan Ford, Renate Voit, Gregory Liszt and Leonard Guarente. My contributions include Figure 1 and Supplemental Figure S 1, as well as contributing to the writing of the paper. 212 SUMMARY The yeast silent information regulator Sir2 is a highly conserved NAD*dependent protein deacetylase. Here, we show that one mammalian homolog, SIRT7, is preferentially localized in nucleoli, where it is associated with the coding regions of the rRNA genes (rDNA). SIRT7 interacts with RNA polymerase I (Pol I) as well as with histones. Overexpression of SIRT7 upregulates Pol I transcription, whereas knock-down of SIRT7 via siRNAs decreases Pol I transcription. SIRT7 increases the amount of Pol I at the coding region of the rDNA. Our results establish SIRT7 as a novel activating component of the mammalian Pol I transcription apparatus. 213 RESULTS AND DISCUSSION RNA polymerase I (Pol I) is the most efficient of the mammalian RNA polymerases, accounting for up to 65% of total transcription in metabolically active mammalian cells (Grummt, 2003). To attain this high level of ribosomal RNA (rRNA) synthesis, the ribosomal DNA (rDNA) comprises approximately 200 tandemly repeated rRNA genes. Each of the 43 kb gene repeats encodes a single rRNA precursor transcript that is later cleaved and processed into the mature 18S, 28S, and 5.8S rRNAs. Intergenic spacer regions (IGS) separate adjacent transcribed rDNA sequences. In all eukaryotes Pol I transcription is highly coordinated with cellular metabolism and cell proliferation (Grummt, 1999). Nutrient starvation, aging, growth factor deprivation, DNA damage, and other conditions that slow down cellular division cause a dramatic decrease in prerRNA synthesis (Grummt, 2003). Along these lines, the tumor suppressors pRb and p53 downregulate Pol I transcription (Budde and Grummt, 1999; Cavanaugh et al., 1995; Voit et al., 1997; Zhai and Comai, 2000). Likewise, conditions that stimulate cellular division increase pre-rRNA synthesis (Zhao et al., 2003). The Sir2 family of enzymes, termed sirtuins, is conserved from bacteria to humans and regulates a wide range of processes including gene silencing, aging, cellular differentiation, and metabolism (Blander and Guarente, 2004). Sirtuins contain a conserved catalytic core domain flanked by divergent amino- and carboxy-terminal regions (Frye, 2000). The catalytic domain confers NAD*-dependentprotein deacetylase as well as ADP-ribosyltransferase activity. The mammalian genome encodes seven homologs of the yeast SIR2 gene, termed SIRT1-7 (Frye, 2000). 214 While several physiologically significant SIRT1 substrates have been identified including p53 (Luo et al., 2001; Vaziri et al., 2001), MyoD (Fulco et al., 2003), FOXO3 (Brunet et al., 2004; Motta et al., 2004), PPARy (Picard et al., 2004), and NF-kB (Yeung et al., 2004), little is known about the biological functions of the other mammalian sirtuins. SIRT2 is a cytoplasmic protein and catalyzes together with HDAC6 the deacetylation of tubulin (North et al., 2003), whereas so far no target has been identified for SIRT3, which is located in mitochondria (Onyango et al., 2002). To begin to address the function of SIRT7, the expression pattern of SIRT7 was investigated by Northern and Western blot analysis using mRNA and cell extracts derived from eight different adult mouse tissues. Antibodies were raised against the Nterminus of mSIRT7, which is not conserved in SIRT1-6. These antibodies recognized one single polypeptide of 45 kD in nuclear extracts, but not in cytoplasmic fractions of mouse and human cells (Fig. A). SIRT7 mRNA was detected in each sample except skeletal muscle, and showed highest levels in the liver (Fig. B). In addition, SIRT7 mRNAs were readily detectable in samples taken from different mouse embryonic stages between day E7 and day E17. SIRT7 protein was also broadly expressed except in heart and muscle (Fig. 1C). Next, we investigated the subcellular localization of human SIRT7 using GFPtagged proteins and indirect immunofluorescence. SIRT7-GFP was located almost exclusively in nucleoli (Fig. D). Moreover, the SIRT7-specific antibodies detected endogenous SIRT7 within distinct nuclear foci, which were co-stained with antibodies against Pol I and therefore represent nucleoli (Fig. E). Nucleolar localization was observed throughout interphase. During M-phase, when nucleoli dissolve, SIRT7 was 215 not retained at the Nucleolus Organizer Region (NOR), but distributed over the entire condensed chromatin (Fig. F). The nucleolar localization raised the possibility that SIRT7 is associated with rDNA. We investigated this by chromatin immunoprecipitation (ChIP). Crosslinked chromatin from 293T cells was immunoprecipitated with anti-SIRT7 antibodies, and precipitated DNA was analyzed by PCR using four primer sets that amplify specific regions of the 43 kb human rDNA repeat except for primer set H23/H27 which recognizes a repeat in the intergenic spacer (IGS) region. (Fig. 2A). ChIP-PCR revealed that anti-SIRT7 antibodies specifically precipitated rDNA when tested with primer sets that amplify two parts of the transcribed region (HI and H13) and the promoter sequence (H0) (Fig. 2B top). No signal above the IgG control was observed with the H23/H27 primer set that amplifies sequences in the IGS, demonstrating the specificity of the interaction of SIRT7 with the transcribed region of the rDNA. The distribution of SIRT7 on rDNA was almost identical to that of Pol I, viewed by precipitation with anti-RPA 16 antibodies, which recognize the second largest subunit of Pol I (Fig. 2B middle). As expected, acetylated histone H4 was present across the entire rDNA repeat including the IGS (Fig. 2B bottom). Since SIRT7 and Pol I were detected at the same rDNA regions, we next tested whether the two proteins interact. Lysates from 293T cells overexpressing FLAG-tagged SIRT7 or mock-transfected 293T cells were immunoprecipitated with anti-FLAG (M2) antibodies, and co-precipitation of Pol I was analyzed by Western blot with RPA1 16 antibodies. While equal amounts of RPA1 16 were present in both cell lysates, it was only detected in immunoprecipitates from cells expressing FLAG-tagged SIRT7 (Fig. 216 2C). Simmalarly, RNA polymerase I remains associated with SIRT7 after tandem affinity purification (TAP)(Fig. 2D). We next tested whether SIRT7 and Pol I interact in vivo at their endogenous levels. Antibodies against Pol I, but not control antibodies, coimmunoprecipitated endogenous SIRT7 along with Pol I from a partially purified mouse nuclear extract (Fig. 2E). Thus, SIRT7 and Pol I associate in vivo at the rDNA. Interestingly, like the yeast SIR protein complex, SIRT7 fused to GST also interacts with histones isolated from HeLa cells in vitro (Fig. 2F). To test whether SIRT7 modulates Pol I transcription, a human rDNA minigene reporter (Voit et al., 1999) and an expression plasmid for FLAG-tagged SIRT7 were cotransfected into human 293T cells. Transcripts derived from the rDNA reporter gene were analyzed by Northern blot. Overexpression of SIRT7 stimulated transcription of the reporter gene up to six-fold over basal levels in a dose-dependent manner (Fig. 3A). A comparable activation of Pol I transcription was also observed in human U2OS cells and mouse 3T3 fibroblasts that were overexpressing SIRT7 (data not shown). This stimulation was Pol I-specific, since overexpression of SIRT7 did not activate transcription of a Pol II-driven luciferase reporter gene (data not shown). These results demonstrate that SIRT7 specifically increases transcription by Pol I in vivo. To gain further evidence that SIRT7 functions in transcriptional activation, cells were depleted of SIRT7 by RNA interference (RNAi). Western blots from cells transfected with anti-SIRT7 small inhibitory RNAs (siRNAs) showed that SIRT7, but not actin protein levels were efficiently decreased by anti-SIRT7 siRNAs (Fig. 3B left). PrerRNA levels were measured using RT-PCR and pairs of primers that amplify part of the 5' external transcribed spacer of the pre-rRNA. Downregulation of SIRT7 by specific 217 siRNAs resulted in significant reduction of pre-rRNA synthesis, whereas expression of GAPDH was not affected (Fig. 3B right). These results indicate that depletion of SIRT7 reduces cellular rRNA synthesis and substantiate the claim that SIRT7 is a positive regulator of Pol I transcription. To determine whether activation of Pol I transcription by SIRT7 is functionally linked to the enzymatic activities of sirtuins, we generated point mutants, replacing serine 112 with alanine and histidine 188 with tyrosine (fig. S1). These residues within the catalytic domain are invariant among sirtuins. Moreover, the homologous mutations have been shown to be required for enzymatic activity in SIRT1 and yeast Sir2 (Frye, 1999; Liszt et al., 2005; Vaziri et al., 2001). Wild-type and mutant proteins were transiently expressed in human 293T cells, and 45S pre-rRNA levels were analyzed on Northern blots. Notably, a two- to three-fold increase of the level of pre-rRNA was observed in cells expressing wild-type FLAG-SIRT7, but not in cells overexpressing 88 Y (Fig. 3C). Both mutant proteins were correctly localized to the SIRT7Sll2A or SIRT7Hl nucleolus (Fig 3D). These findings indicate that an NAD+-dependent enzymatic activity of SIRT7 is required for transcriptional activation but not for nucleolar localization. In addition, pre-rRNA levels were determined in cells treated with nicotinamide, which is a potent inhibitor of sirtuins. As expected, treatment of NIH3T3 cells with 5 mM nicotinamide strongly inhibited pre-rRNA synthesis (Fig. 3E). In contrast, TSA, an inhibitor of class I and II HDACs, did not affect the level of pre-rRNA synthesis, suggesting that SIRT7, but not the class I and II HDACs, supports Pol I transcription in vivo. Inhibition by nicotinamide is not due to cellular redistribution of SIRT7, since the localizatioin of GFP-SIRT7 is not altered in the presence or absence of nicotinamide (Fig. 218 3F). In contrast, addition of actinomycin D at a concentration that selectively inhibits Pol I transcription released GFP-SIRT7 from the nucleoli resulting in redistribution throughout the nucleus. These results suggest that nucleolar localization of SIRT7 is tightly linked to ongoing Pol I transcription. Based on the observation that SIRT7 is (i) bound to rDNA, (ii) interacts with Pol I, and (iii) stimulates Pol I transcription, we asked whether SIRT7 changes the occupancy of Pol I at the rDNA. To investigate this, we performed ChIP using anti-RPA 16 antibodies and cell lines stably expressing FLAG-SIRT7 wild-type or mutant proteins. Western blot analysis demonstrated equivalent overexpression of SIRT7Wt,SIRT7s t2 A, 88 Y (Fig. 4A). Both input chromatin and chromatin precipitated with antiand SIRT7Hl RPA116 antibodies were used as template for PCR with primers corresponding to the rDNA promoter (HO), coding region (H1 and H13), and IGS (H23/H27). Notably, the amount of rDNA precipitated with Pol I-specific antibodies was significantly higher in SIRT7-overexpressing cells when compared to control cells that had been transfected with the empty vector (Fig. 4B). However, when cell lines overexpressing the SIRT7 point mutants S 112Aand H188Y were analyzed, the levels of precipitated rDNA was similar to that of the empty vector (Fig. 4B). When a primer pair to the IGS was used, a region in which Pol I is not bound, the background ChIP signal was identical in all four cell lines, demonstrating that the differences between the cell lines seen with primer pairs HO, H1, and H13 is not due to variations in the preparation of the chromatin. These data demonstrate that overexpression of SIRT7 increases the amount of Pol I at the rDNA. In addition, enhancement of association of Pol I with the rDNA requires conserved residues within the catalytic domain of SIRT7. 219 In a converse experiment, we analyzed the association of Pol I with rDNA in cells transfected with nonspecific or SIRT7-specific siRNAs. Western blot analysis demonstrated significant reduction of SIRT7 protein levels by SIRT7-specific siRNA, whereas the levels of actin were the unaffected (Fig. 4C). Crosslinked chromatin was precipitated from cells treated with nonspecific and SIRT7-specific siRNAs using antiRPA116 antibodies, and quantified by PCR with the rDNA-specific primer sets H0, H1, H13, and H23/H27. Anti-RPA1 16 antibodies precipitated significantly less rDNA from cells treated with anti-SIRT7 siRNAs compared to cells treated with nonspecific siRNAs (Fig. 4D). This demonstrates that reduction of cellular SIRT7 protein levels by RNAi concomitantly reduced the amount of Pol I associated with rDNA. These results are in agreement with our observation that SIRT7 is critical for activation of Pol I transcription. Here, we establish that SIRT7 is an activating component of the RNA polymerase I transcriptional machinery. SIRT7 strongly stimulates Pol I transcription in vivo by an enzymatic mechanism. Importantly, these data suggest a novel mechanism by which mammalian cells may regulate rRNA transcription. In yeast, a change in the NAD+/NADH ratio translates the metabolic shift of calorie restriction into lifespan extension via Sir2p (Lin et al., 2004). Such a model is also believed to explain the regulation of SIRT1 in liver and muscle cells, where changes in the NAD+/NADHratio can arise from diet or exercise and influence SIRT1 activity (Fulco et al., 2003; Rodgers et al., 2005). This paradigm suggests that diet-induced changes in the NAD+/NADH ratio may regulate SIRT7 to couple changing energy status with levels of rRNA transcription and ribosome production. 220 Sir2 homologs positively regulate longevity in yeast, worms, and flies despite extraordinary phylogenetic distance between these organisms (Hekimi and Guarente, 2003). In yeast, Sir2p promotes longevity predominantly through its silencing role at the rDNA (Kaeberlein et al., 1999; Sinclair and Guarente, 1997). It is interesting that the effect of sirtuins at the rDNA has been inverted in mammals. SIRT7 activates RNA polymerase I transcription, which is typically proportional to the requirement for cell growth. We suspect that a link between SIRT7 and growth may be part of a larger set of physiological changes involving all seven sirtuins that is induced by diet. 221 MATERIALS AND METHODS Plasmid constructs, antibodies and antibody production The human SIRT7 cDNA was purchased from ATCC (IMAGE 222518), amplified by PCR, and cloned into pCMV-Tag4a (Stratagene) cut with EcoRI and XhoI to make the plasmid pCMVFLAG-SIRT7. The plasmid pCMV-FLAG-SIRT7 served as template for site directed mutagenesis using the Stratagene Quick Change mutagenesis kit to produce the plasmids pCMVFLAG-SIRT7sl 12A and pCMV-FLAG-SIRT7Hl 88Y. pEGFP-SIRT7 was constructed by amplifying the SIRT7 coding sequence by PCR and cloning it into the EcoRI and BamHI sites of pEGFP-N1 (Clontech). pET-GST-SIRT7 was created by cloning the SIRT7 coding sequence into the NheI and BamHI sites of the vector pET-GST (kind gift from R. Marciniak, University of Texas, San Antonio). To generate TAP-tagged SIRT7, the stop codon of the SIRT7 cDNA was removed by site directed mutagenesis and substituted by a BglII site, and the coding sequence was cloned into the Bam HI site of the vector pZome-lC (Cellzome). The N-terminal 81 amino acids of mouse SIRT7 were expressed in E. coli as a glutathione S-transferase (GST)-fusion and purified over a glutathione-Sepharose column. The purified protein was injected into NZW rabbits by Covance Bioproducts to create SIRT7 antiserum. Full-length SIRT7 and SIRT7(1-81) proteins were bound to NHS-Sepharose according to manufacturer instructions (Amersham). The crude serum was first passed over the SIRT7(1-81)-NHS-Sepharose resin. The resin was washed extensively. The purified antibody was eluted with 0.1 M glycine, pH 2.5, and the pH was neutralized by the addition of 1/104 volume of 1 M Hepes, pH 7.9. The process was repeated a second time using the full-length SIRT7-NHS-Sepharose resin. Other antibodies used were anti-actin C4 (ICN), anti-Flag M2 (Sigma), anti-acetylated histone H4 (Upstate), and anti-histone H4 (Upstate), a purified rabbit polyclonal antibody against RPA-116, and human autoimmune serum against the largest subunit of Pol I (provided by U. Scheer, University of Wuerzburg). Multiple tissue Northern and Western blots A hybridization probe containing the SIRT7 open reading frame was created using the Prime-It Random Primer Labeling kit (Stratagene) and was used to probe Multiple Tissue Northern Blots (Stratagene) according to the manufacturer's instructions. The blots were then stripped and reprobed with the actin probe included in the kit. 222 Homogenized mouse tissues from two mice of the FVB genetic background were obtained as a generous gift from L. Bordone, MIT. From these samples, SDS-solubilized proteins were prepared, separated by SDS-PAGE, and analyzed by Western blotting. Cell culture, transfections and RNA analysis 293T, U2OS and NIH3T3 cells were cultured in DMEM supplemented with 10% FCS. Where indicated cells were treated with 0.050 yg/ml actinomycin D, 5 mM nicotinamide, or 40 nM TSA. For transient expression, 293T or U2OS cells were transfected in 60 mm dishes with the calcium phosphate precipitation technique using different amounts of pCMV-SIRT7 plasmids and 2 ug rDNA reporter plasmid pHr-P-BH, which contains a fragment from -401 to +378 (+1, transcription start site) of the human rDNA repeat fused to a BamHI-Hinfl fragment encompassing the transcription termination sites T1 and T2. Reporter transcripts were analyzed on Northern blots as described (Voit et al., 1999). To analyze pre-rRNA synthesis, cellular RNA was isolated and subjected to Northern blot analysis using a 32 P-labeled antisense RNA encompassing the 5'-terminal rDNA sequences from -57 to +183. To normalize for gel loading, the blots were reprobed with radiolabeled riboprobes against cytochrome c oxidase 1 (cox 1). For stable expression of TAP-tagged SIRT7, U2OS cells were transfected with pZome-SIRT7 or pZome alone followed by selection with 2 g/ml of puromycin. Similarly, for stable expression of FLAG-tagged SIRT7 cell lines, 293T cells were transfected with pCMVTAG-4a (Stratagene), pCMV-FLAG-SIRT7, pCMV-FLAG-SIRT7sl 2A, or pCMV-FLAG88 Y in combination with pBABE-PURO at a 10:1 ratio and selected with 2 mg/ml SIRT7Hl puromycin. Immunofluorescence U2OS cells were fixed with paraformaldehyde as described before (Zatsepina et al., 1993) and stained with purified anti-SIRT7 antibodies, anti-Pol I antibodies, anti-rabbit Alexa-488 antibodies (MoBiTec), anti-human Cy3 antibodies (Dianova), and Hoechst 33342. Images were visualized using a Zeiss Axiophot microscope. Localization of GFP-tagged proteins was visualized with a Nikon miscroscope in live cells that had been treated with the dye Hoechst 33342. Immunoprecipitations One 10 cm plate of 293T cells stably transfected with pCMV-Tag4a and pCMV-FLAG-SIRT7 was resuspended in 1 ml NP-40 lysis buffer (50 mM Tris-HC1, pH 8.0, 150 mM NaC1, 1% NP- 223 40) and transferred to a 1.5 ml tube. The lysis reaction was allowed to proceed for 30 min on ice and centrifuged for 15 min at 14,000 rpm. 10 ul of FLAG(M2)-protein A-Sepharose beads were added to the lysates and incubated for 1 hr. After extensive washing the beads were boiled in SDS loading buffer and the eluted proteins were analyzed by SDS-PAGE and Western blotting. Chromatin immunoprecipitations Chromatin immunoprecipitation (ChIP) experiments were performed essentially as described (Weinmann et al., 2001) with a few modifications. For each ChIP, one 15 cm plate of 293T cells was grown to 75% confluency. was added directly to the media to a final Formaldehyde concentration of 1% and incubated for 10 min at room temperature. The reaction was stopped by adding glycine to a final concentration of 0.125 M for 5 min. Cells were washed twice with cold PBS, incubated with 2 ml of 2.5% trypsin at 37 ° C for 10 min, harvested with 10 ml of PBS plus Complete Protease Inhibitors (Roche), and washed twice with PBS plus protease inhibitors. The final cell pellet was resuspended in cell lysis buffer (5 mM HEPES pH 8.0, 85 mM KCl, 0.5% NP-40, and protease inhibitors), incubated on ice for 5 min, and centrifuged at 5,000 rpm in a microfuge for 10 min. The pellet was resuspended in nuclei lysis buffer (50 mM Tris-Cl pH 8.1, 10 mM EDTA, 1% SDS, and protease inhibitors) and incubated on ice for 10 min. The chromatin was sonicated to an average length of approximately 500 bp as visualized by ethidium bromide agarose gel electrophoresis. The resulting chromatin was cleared of insoluble material by spinning at 14,000 rpm for 15 min in a microfuge and diluted 5-fold in IP dilution buffer (0.01% SDS, 1.1% Triton X-100, 1.2 mM EDTA, 16.7 mM Tris-Cl pH 8.1, 167 mM NaCl). Chromatin was pre-cleared by adding 80 ,pulof salmon sperm DNA/BSA/protein-A slurry (0.4 mg/ml salmon sperm DNA, 1 mg/ml BSA, 10 mM Tris-Cl pH 8.0, 1 mM EDTA, 50% protein A sepharose beads). In separate 1.5 ml tubes, antibody beads were prepared by incubating 10 1 of specific antibodies with protein-A Sephaorse in 500 ml IP dilution buffer plus 50 mg/ml BSA and 0.4 mg/ml ssDNA. After 1 hour the supernatant was removed from the antibody beads. The pre- cleared chromatin was added and incubated for 3 hours at 4° C. The beads were washed two times with each of the following buffers: low salt wash buffer (0.1% SDS, 1% Triton X-100, 2 mM EDTA, 20 mM Tris-Cl pH 8.1, 150 mM NaCl), high salt wash buffer (0.1% SDS, 1% Triton X-100, 2 mM EDTA, 20 mM Tris-Cl pH 8.1, 500 mM NaCl), LiCl wash buffer (1% deoxycholic acid, 1% NP-40, 1 mM EDTA, 10 mM Tris-Cl pH 8.0, 250 mM LiCl), and TE. After the last wash the chromatin was eluted from the beads by incubating twice for 15 minutes with 150 p IP elution buffer (100 mM NaHCO3, 1% SDS). Formaldehyde crosslinking was reversed by incubating at 65 ° C for 4 hours in the presence of 0.3 M NaCl and 0.1 mg/ml RNase A. DNA was 224 ethanol precipitated and resuspended in 100 pl of H 20. 2 6.5, and 1.5 l 0.5 M EDTA, 4 l 1 M Tris-Cl pH l of 20 mg/ml Proteinase K were added and incubated for 2 hours at 45 ° C. Samples were extracted with phenol/chloroform/isoamyl alcohol and ethanol precipitated. were washed with 70% ethanol, dried and resuspended in 30 0.1 mM EDTA). Pellets l TE/10 (10 mM Tris-Cl pH 7.5, Increasing amounts of precipitated chromatin was used as template in quantitative PCR reactions as previously described (O'Sullivan et al., 2002). Purification of proteins and protein-protein interactions GST and recombinant GST-tagged SIRT7 were expressed in E. coli Codon Plus-RP cells (Stratagene) and affinity purified over glutathione-Sepharose according to standard protocols. Acetylated core histones were purified from butyric acid treated HeLa cells by hydroxylapatite chromatography essentially as described (Ausubel et al., 1997). To assay for interaction of SIRT7 and histones, GST and GST-SIRT7 were immobilized on glutathione-Sepharose and incubated with 2 pg of purified core histones for 2 h at 4° C in buffer AM-400 (400 mM KCl, 20 mM Tris-Cl pH 7.9, 5 mM MgCl 2 , 0.1 mM EDTA, 10% glycerol, 0.5 mM EDTA) supplemented with 0.2% NP-40 and protease inhibitors. After five washes in incubation buffer, bound proteins were eluted with SDS-loading buffer, separated on 15% polyacrylamide gels by SDS-PAGE, and visualized by staining with Coomassie. U2OS cells stably transfected with pZome or pZome-SIRT7 were grown on ten 150 mm dishes to 80% confluency. Cells were harvested in buffer A (10 mM Hepes pH 7.9, 1.5 mM MgCl 2, 10 mM KCl, 0.5 mM DTT, protease inhibitors), incubated for 15 min at 4 ° C, and the nuclei pelleted by centrifugation. The nuclei were resuspended in buffer C (20 mM Hepes pH 7.9, 420 mM NaCl, 1.5 mM MgCl2, 25% glycerol, 0.2 mM EDTA, protease inhibitors), homogenized, and incubated at 4° C for 30 min. The lysate was cleared by centrifugation at 13000 rpm for 30 min at 4 ° C, and the extract was dialysed against buffer AM-100 (100 mM KCl, 20 mM Tris-Cl pH 7.9, 5 mM MgCl 2, 0.1 mM EDTA, 10% glycerol, 0.5 mM EDTA, 0.5 mM PMSF). Batch purification of the IgG binding domain-tagged proteins was performed using 300 l of IgG-Sepharose 6 Fast Low resin (Amersham Biosciences). The protein was eluted from the beads by washing three times in TEV-cleavage buffer (25 mM Tris-Cl pH 8.0, 150 mM NaCl, 0.1% NP-40, 0.5 mM EDTA, 1 mM DTT) and rotating over night at 4° C in TEV-cleavage buffer containing 50 units AcTEV protease (Invitrogen). Supernatants from the TEV-cleavage reaction containing SIRT7 and interacting proteins were collected and substituted with CaCl2 (final concentration 2 mM) and 3 volumes of calmodulin binding buffer (25 mM Tris-Cl pH 8.0, 150 mM NaCl, 1 mM magnesium acetate, 1 mM imidazol, 225 2 mM CaCl 2, 10 mM B- mercaptoethanol, 0.1% NP-40). The proteins were batch-purified with 200 Pl of calmodulin beads (Stratagene) under rotation for 2 h at 4 C. The beads were washed three times in calmodulin binding buffer, transferred into a column and bound proteins eluted with calmodulin elution buffer (25 mM Tris-Cl pH 8.0, 150 mM NaCl, 1 mM magnesium acetate, 1 mM imidazol, 20 mM EGTA, 10 mM B-mercaptoethanol, 0.02% NP-40). RNAi experiments To knock-down cellular SIRT7 by RNAi, dsRNAs were synthesized as a custom "Smart Pool" against SIRT7 and "non-specific control duplexes - XIII" (Dharmacon). To determine the effect of these siRNAs on pre-rRNA levels, the duplexes were transfected into U2OS cells using RNAifect (Qiagen). After incubation for 48 h, cells were harvested and lysed with SDS loading buffer for Western blot analysis or with GITC for isolation of cellular RNA. Cellular RNA was reverse transcribed with random primers. The amount of pre-rRNA determined by 32 semiquantitative PCR in the presence of a- P-dCTP with primer 1 (forward) 5'-GCT GTC CTC TGG C-3' and primer 2 (reverse) 5'-CGG CAG GCG GCT CAA G-3' that amplify a fragment from +9 to +120 of the external transcribed spacer of the human rDNA repeat. Radiolabeled PCR fragments were analyzed on 8% polyacrylamide gels and quantified using a Phosphorimager. GAPDH-cDNA was amplified by PCR using the primers 5'-CCA TCA CCA TCT TCC AGG AG-3' and 5'-CCT GCT TCA CCA CCT TCT TG-3'. 293T cells for ChIP assays were transfected with the above siRNAs using Oligofectamine (Invitrogen) according to the manufacturer's instructions. 48 hours post-transfection the cells were transfected a second time and 48 hours after the second transfection the cells were harvested for ChIP experiments as described above. 226 REFERENCES Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G., Smith, J. A., and Struhl, K. (1997). Current Protocols in Molecular Biology (New York: John Wiley and Sons). Blander, G., and Guarente, L. (2004). The Sir2 family of protein deacetylases. Annu Rev Biochem 73,417-435. Brunet, A., Sweeney, L. B., Sturgill, J. F., Chua, K. F., Greer, P. L., Lin, Y., Tran, H., Ross, S. E., Mostoslavsky, R., Cohen, H. Y., et al. (2004). Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. Science 303, 2011-2015. Budde, A., and Grummt, I. (1999). p53 represses ribosomal gene transcription. Oncogene 18, 1119-1124. Cavanaugh, A. H., Hempel, W. M., Taylor, L. J., Rogalsky, V., Todorov, G., and Rothblum, L. I. (1995). Activity of RNA polymerase I transcription factor UBF blocked by Rb gene product. Nature 374, 177-180. Frye, R. A. (1999). Characterization of five human cDNAs with homology to the yeast SIR2 gene: Sir2-like proteins (sirtuins) metabolize NAD and may have protein ADP-ribosyltransferase activity. Biochem Biophys Res Commun 260, 273-279. Frye, R. A. (2000). Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochem Biophys Res Commun 273, 793-798. Fulco, M., Schiltz, R. L., Iezzi, S., King, M. T., Zhao, P., Kashiwaya, Y., Hoffman, E., Veech, R. L., and Sartorelli, V. (2003). Sir2 regulates skeletal muscle differentiation as a potential sensor of the redox state. Mol Cell 12, 51-62. Grummt, I. (1999). Regulation of mammalian ribosomal gene transcription by RNA polymerase I. Prog Nucleic Acid Res Mol Biol 62, 109-154. Grummt, I. (2003). Life on a planet of its own: regulation of RNA polymerase I transcription in the nucleolus. Genes Dev 17, 1691-1702. Hekimi, S., and Guarente, L. (2003). Genetics and the specificity of the aging process. Science 299, 1351-1354. Kaeberlein, M., McVey, M., and Guarente, L. (1999). The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev 13, 2570-2580. Lin, S. J., Ford, E., Haigis, M., Liszt, G., and Guarente, L. (2004). Calorie restriction extends yeast life span by lowering the level of NADH. Genes Dev 18, 12-16. 227 Liszt, G., Ford, E., Kurtev, M., and Guarente, L. (2005). Mouse Sir2 homolog SIRT6 is a nuclear ADP-ribosyltransferase. J Biol Chem. Luo, J., Nikolaev, A. Y., Imai, S., Chen, D., Su, F., Shiloh, A., Guarente, L., and Gu, W. (2001). Negative control of p53 by Sir2alpha promotes cell survival under stress. Cell 107, 137-148. Motta, M. C., Divecha, N., Lemieux, M., Kamel, C., Chen, D., Gu, W., Bultsma, Y., McBurney, M., and Guarente, L. (2004). Mammalian SIRT1 represses forkhead transcription factors. Cell 116, 551-563. North, B. J., Marshall, B. L., Borra, M. T., Denu, J. M., and Verdin, E. (2003). The human Sir2 ortholog, SIRT2, is an NAD+-dependent tubulin deacetylase. Mol Cell 11,437-444. O'Sullivan, A. C., Sullivan, G. J., and McStay, B. (2002). UBF binding in vivo is not restricted to regulatory sequences within the vertebrate ribosomal DNA repeat. Mol Cell Biol 22, 657-668. Onyango, P., Celic, I., McCaffery, J. M., Boeke, J. D., and Feinberg, A. P. (2002). SIRT3, a human SIR2 homologue, is an NAD-dependent deacetylase localized to mitochondria. Proc Natl Acad Sci U S A 99, 13653-13658. Picard, F., Kurtev, M., Chung, N., Topark-Ngarm, A., Senawong, T., Machado De Oliveira, R., Leid, M., McBurney, M. W., and Guarente, L. (2004). Sirtl promotes fat mobilization in white adipocytes by repressing PPAR-gamma. Nature 429, 771-776. Rodgers, J. T., Lerin, C., Haas, W., Gygi, S. P., Spiegelman, B. M., and Puigserver, P. (2005). Nutrient control of glucose homeostasis through a complex of PGC- lalpha and SIRT . Nature 434, 113-118. Sinclair, D. A., and Guarente, L. (1997). Extrachromosomal rDNA circles--a cause of aging in yeast. Cell 91, 1033-1042. Vaziri, H., Dessain, S. K., Ng Eaton, E., Imai, S. I., Frye, R. A., Pandita, T. K., Guarente, L., and Weinberg, R. A. (2001). hSIR2(SIRT1) functions as an NAD-dependent p53 deacetylase. Cell 107, 149-159. Voit, R., Hoffmann, M., and Grummt, I. (1999). Phosphorylation by GI-specific cdk-cyclin complexes activates the nucleolar transcription factor UBF. Embo J 18, 1891-1899. Voit, R., Schafer, K., and Grummt, I. (1997). Mechanism of repression of RNA polymerase I transcription by the retinoblastoma protein. Mol Cell Biol 17, 4230-4237. Weinmann, A. S., Bartley, S. M., Zhang,T., Zhang, M. Q., and Farnham, P. J. (2001). Use of chromatin immunoprecipitation to clone novel E2F target promoters. Mol Cell Biol 21, 68206832. 228 Yeung, F., Hoberg, J. E., Ramsey, C. S., Keller, M. D., Jones, D. R., Frye, R. A., and Mayo, M. W. (2004). Modulation of NF-kappaB-dependent transcription and cell survival by the SIRTI1 deacetylase. Embo J 23, 2369-2380. Zatsepina, O. V., Voit, R., Grummt, I., Spring, H., Semenov, M. V., and Trendelenburg, M. F. (1993). The RNA polymerase I-specific transcription initiation factor UBF is associated with transcriptionally active and inactive ribosomal genes. Chromosoma 102, 599-611. Zhai, W., and Comai, L. (2000). Repression of RNA polymerase I transcription by the tumor suppressor p53. Mol Cell Biol 20, 5930-5938. Zhao, J., Yuan, X., Frodin, M., and Grummt, I. (2003). ERK-dependent phosphorylation of the transcription initiation factor TIF-IA is required for RNA polymerase I transcription and cell growth. Mol Cell 11, 405-413. Acknowledgements The authors thank Marcia Haigis, and Gil Blander for thoughtful discussion and careful reading of the manuscript. LG and EF are supported by NIH grants. 229 FIGURES Fig. 1. SIRT7 is a nucleolar protein. (A) SIRT7 is a nuclear component of mammalian cells. 20 mg cytoplasmic and nuclear extracts from NIH-3T3 and U2OS cells were subjected to Western blot analysis with anti-SIRT7 antibodies. (B) RNA was isolated from the indicated tissues (left panels) and embryos (right panels) in wild-type adult mice, resolved by electrophoresis, and subjected to Northern blotting. Blots were probed with 32 -labeled cDNA specific to SIRT7 (top panels) or actin (lower panels). Muscle actin found in heart and skeletal muscle samples migrated as a distinct band. (C) Protein extracts from the indicated wild-type mouse tissues were resolved by SDS-PAGE, and analyzed by Western blotting using antibodies specific to SIRT7 (top panel) or actin (lower panel). For each tissue type, samples from two different mice were included. (D) Nucleolar localization of GFP-SIRT7 in live human U2OS cells. (E) Endogenous SIRT7 and Pol I were visualized by immunofluorescence microscopy in U2OS cells using anti- SIRT7 (green) and anti-Pol I antibodies (red). The DNA was counter-stained with Hoechst 33342. (F) During mitosis SIRT7 is associated with condensed chromatin. SIRT7-GFP was visualized in live U2OS cells stained with Hoechst 33342. 230 A Celllysates NIH 3T3 w z () HeLa w () mSIAT? z actin - ... -.- 2 3 4 6 GFP-SIAT? • 8 9 10 11 13 SIAT7 C .... ---- actin 1-1 D 7 2 ---,- 3 4 R 5 6 __ V ~ 7 8 9 .-... _ "!",...I IIJ" _I mS IAT? actin 10 11 13 14 15 16 17 phase F me~~hase anaphase ._ 1111 Figure 1 Fig. 2. SIRT7 is associated with the rDNA and RNA polymerase I. (A) Diagram depicting the location of the primer pairs used: (H0) promoter, (H1) 5' external transcribed spacer, (H13) end of 28S rRNA, and (H23/H27) intergenic region of the human rDNA gene. (B) Results of ChIP analysis showing the localization of SIRT7 and RNA pol I within transcribed region of the rDNA gene repeats. Chromatin from 293T cells was precipitated by the indicated antibodies and analyzed by PCR with the indicated primer pairs. (C) Coimmunoprecipitation of SIRT7 and RNA polymerase I (RPA116). Anti-FLAG immunoprecipitations were performed from whole cell extracts prepared from 293T cells harbouring empty vector (pCMV) or a plasmid expressing FLAG-tagged SIRT7 (pCMV-SIRT7-FLAG). Immunoprecipitates were analyzed by Western blot as indicated. (D) Pol I is a component of the SIRT7-TAP complex. U2OS cell lines harboring the empty vector or SIRT7-TAP were established. TAP-tagged complexes were purified on IgG and Calmodulin resins, and the eluted protein complexes tested for copurifying proteins by Western blot analysis. Lane 1, 50 /g of cell extract; lane 2, eluate from mock transfected cells; lane 3, eluate from a cell line expressing SIRT7-TAP. The Western blot was probed with an antibody specific for RPA116. (E) Endogenous SIRT7 interacts with RNA polymerase I. Partially purified human nuclear extract was immunoprecipitated with control non-specific human antibodies (lane 2) and anti-RNA polymerase I antibodies (lane 3). The precipitated proteins were analyzed by western blotting with anti-SIRT7 and anti-RPA-1 16 antibodies as indicated. (F) SIRT7 interacts with histones. Histones were isolated from butyric acid treated HeLa cells by ion exchange chromatography incubated with GST-SIRT7-Sepharose and GST- Sepharose beads. The proteins were eluted and separated by SDS-PAGE and visualized by staining with Coomassie. 232 A 18S POETS 28S 5.8S • 3'ETS IGS /;-- I H23/H27 H13 HOH1 H23/H27 B HO a-SIRT711 5 II I 5I IgG I I HI IgG I I H13 a-SIRT711 IgG a-SIRT71 5 II I 5 I I I 5 II I 5 ~JW I II ' ~3'1"l~'1[1<s;F' .... ~ H23/H27 Oligo Pair IgG a-SIRT71 I.P. I 5 II I 5 I ~I Input ,~'~'~Y"":~ '!it~~k*~~",,, <._ ~*\m:~~~~ J'L~'.<=i~~N«~ ""il/" > "< HO IgG 5 I I II HI a-Poll I I IgG I 5I I 1 5 II J ..'.~'" ~~~~ ~;. ""'$ f::.._:aYIi;o.;~ +\,~~ SIRTl H13 H23/H27 Oligo Pair a-Poll II IgG a-Pol II I IgG a-Poll I I.P. ' 1 5 I I 1 5 II 1 5 II 1 5 II 1 5 I ~I Input Poll HO IgG 5 I I II HI a-H4 II IgG 1 5I I I 5 II Hl3 H23/H27 Oligo Pair a-H4 II IgG a-H4 I I IgG a-H4 I I.P. 1 5 I I 1 5 II I 5 II I 5 II I 5 I ~I Input Histone H4 2 4 3 C 5% Input I (!)II a. 10 2Q) 0 13 !!; ~0 en c:i. E ~ 0 () :> :E a. a. - .. 1 U _ - 16 17 <5 a.. 6 (!) S? - .- en ~ 15 '5 a. E !!; U 14 E ~ Q) Cti 11 ~ > a. D (/) :> :E 9 8 ~~ LL a: U5 U 7 (!)I ~ :E 6 aFLAG I.P. ~~ U. > 5 - RPA116 _ ... 2 a-RPA-116 a-SIRT7 3 RPA116~,k ..... en -SIRT7 2 3 (!) /H3 -H2A1B 4 -H4 2 3 2 3 Figure 2 Fig. 3. SIRT7 is an activator of Pol I transcription. (A) Overexpression of SIRT7 stimulates Pol I transcription from an rDNA minigene reporter. 293T cells were cotransfected with a Pol I reporter plasmid and increasing amounts of an expression plasmid encoding SIRT7-Flag. RNA was isolated from transfected cells. The amount of reporter transcripts was determined by Northern blot analysis and quantified using a PhosphorImager. Bottom panel: Western blot of the transfected cells probed with antiFlag antibody. (B) Knock-down of SIRT7 by RNAi impairs rRNA synthesis. Left panel: Western blot of extracts from U2OS cells transfected with control-dsRNA (lane 1) or SIRT7-specific dsRNA (lane 2). Right panel: Pol I transcription was analyzed by RTPCR. RNA was isolated from cells treated with the respective dsRNAs. Increasing amounts of cDNA were used for PCR with primers that amplify a region of the 5' external transcribed spacer of the human pre-rRNA. Results from two independent experiments (lanes 1,2,5,6 and 3,4,7,8, respectively) are shown. As a control, RT-PCR was performed to analyze expression of GAPDH. (C) Stimulation of cellular Pol I transcription by SIRT7 requires reisdues H188 and S112. Northern blot analysis of prerRNA transcripts from 293T cells transfected with different amounts of plasmids expressing wild-type SIRT7 or the point mutants H188Y and S112A (top panel). Expression levels of SIRT7-Flag, SIRT7/H188Y-Flag, and SIRT7/Sl12A-Flag were monitored on Western blots using anti-Flag antibodies (bottom panel). (D) The SIRT7 mutants H188Y and S1 12A were expressed as GFP-tagged proteins in U2OS cells and their cellular localization was examined in live cells. (E) Nicotinamide represses rRNA synthesis. NIH3T3 cells were cultured for 6h in medium containing 40 nM TSA or 5 mM nicotinamide, and pre-rRNA levels were measured by Northern blot analysis. The blot was subsequently reprobed for cytochrome c oxidase (coxl) mRNA. (F) Treatment with actinomycin D, but not with nicotinamide releases SIRT7-GFP from nucleoli. U2OS cells expressing SIRT7-GFP were cultured in the presence of either 50 mg/ml of actinomycin D for 2h to inhibit Pol I activity or 5 mM nicotinamide for 6h to inhibit NAD+-dependent deacetylase activity. Localization of SIRT7-GFP was examined in live cells. 234 A B - 0.5 Northern blot: - 1 2 •• reporter transcript 2 5.9 6.5 EtBr: _ ..~.... :..... ' .'..' ..".• " WB a-Flag: _ ., c WB: 28S rRNA a-SIRT7 • a-actin •• Flag-Sirt? 3 4 .1 - 0.5 1 0.5 1 0.5 ...... _ ...... ______ 1 - >'tII "-" ....... "'. GAPDH - ...... - .... -- -- ~ ~~ -2 2 pre-rRNA 28S rRNA '" pre-rRNA I1g pCMV-Flag-Sirt? lilIIIlW EtBr: """ ,"'''' - - .. 3 4 5 6 ? 8 D WT H188Y S112A Northern blot: RT-PCR: fold stimulation 18S rRNA •• 1 2 SIRT7-siRNA control-siRNA I1g pCMV-Flag-Sirt? 18S rRNA ~ •• Q) en cu ... WB a-Flag: 2 3 4 Flag-Sirt? .t: C. 56? E Northern blot: •• 2 pre-rRNA cox1 3 Figure 3 Fig. 4. SIRT7 stimulates the association of RNA Polymerase I with the rDNA (A) Western blot showing overexpression levels of SIRT7. (B) Enrichment of RNA pol I at the transcribed region of the rDNA genes in the presence of ectopic SIRT7. Chromatin was prepared from cells harbouring empty vector (pCMV) or plasmids expressing wildtype SIRT7 (SIRT7-wt) and mutant SIRT7 (SIRT7-S112A and SIRT7 H188Y). The chromatin was precipitated with anti-RPA1 16 antibodies and analyzed by PCR with the primer pairs indicated on the right side of the figure. (C) Following siRNA transfections, whole-cell lysates were analyzed by immunoblot for SIRT7 and B-actin. (D) ChIP analysis of Pol I levels in cells transfected with siRNAs. Chromatin was prepared from cells transfected with non-specific (ns) or anti-SIRT7 siRNA. The chromatin was precipitated with anti-RPA1 16 antibodies and analyzed by PCR with the primer pairs indicated on the right side of the figure 236 A B (\J ><Xl U; I <{ ..... == > Input SIRT7- SIRT7SI~T7- S112A H188Y <Xl ..... pCMV ---- ~ ~ ~ a: a: c: ~ 0 0.. U5 U5 5 1 5 5 1 1 5 anti-Poll IgG SIRT7- SIRT7pCMV SI~T7- ---1 5 5 1 S112A 1 5 H188Y 1 5 SIRT7- SIRT7- SIRT7pCMV ~ S112A H188Y 5 1 5 1 5 5 JlI Input U5 HO -SIRT7 2 3 4 H1 H13 H23/H27 4 2 c D <{ z <{ z 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 anti-Poll Input a: "(jj a: "(jj Jllinput (/) c ~:~ ..w' -~ - HO SIRT7 -actin H1 2 H13 H23/H27 2 3 4 5 6 7 8 9 10 11 12 Figure 4 SUPPORTING FIGURE Fig. S1. Sequence alignment of mouse sirtuin core domains. Conserved core domains of mouse SIRT1, SIRT4, SIRT5, and SIRT7 were aligned with the yeast Sir2 core using ClustalW. In the sirtuin phylogenetic tree, SIRT1 is class I, SIRT4 is class II, SIRT5 is class III, and SIRT7 is class IV. At each position of the alignment, identical residues are boxed and shaded in gray. Amino acids filled in black correspond to SIRT7 residues targeted for mutagenesis in this study. 238 -I.. ySir2p hSIRTI hSIRT4 hSIRT5 hSIRTI ySir2p hSIRTI hSIRT4 hSIRT5 hSIRTI ySir2p hSIRTI hSIRT4 hSIRT5 hSIRT7 ySir2p hSIRTI hSIRT4 hSIRT5 hSIRTI ySir2p hSIRTI hSIRT4 hSIRT5 hSIRTI K V." L - A. G N G V - - - 1 K H L GID R L A V D F P D P R T D R R P 1 Q H G -WRKWQAQDIIAT - - - -WTLLQKGR S ~~I~~~~ ~~~11~~~~~~~~~~~1~~~ I -I'.. Fl'. wi R AS - - - --IIQ.M - - S -FQ 1 Y S Q F S S H Q S KEN - ,- L S EAE P ,I R Q R Y WAR N F V G HN~SRVWEFYHYRREVMG D - - - - - - - SLC L N PAW A G TLT QV - - S G'LYW---LV K L L R N Y -K A T K G QtlJR R V V V 1 HEQKLVQHVV E G R K T K N - -ilL. L R L P R TAl ; L I ....•K Y :..........•.... .• ""'.:.::. V TI ..••.....•..... H W \'. L D. G E .~. T S .G I;'. T S '.. ' - - K N Q - V D C E AI." . R LII..........................• G E R 1 F T_ - R G .. L - - - - -. '.' V V N R E Y Ii- G N Q A V OliN Q K 1 .R. N L E R . Q E N Y K R V D - II S Y - - - - - IIR R G V A EFT RRPKLYIV RE DIK IG ETA LQW L L - - TWSAEAHBL_DVF ALSGKGE R TAL H R H Q TIII- - - - V 1 :I~ - T Q - - - - Y - - - PGT - - - - - - - - - - - ~I:::: DTV- --N ~'--~ - A P V S EIV N M V P RPVAL PSSIP SGYRF LTAWE PAAMFAPQVAA P R L W C M T K P P S H A E -IDIISILIY HLHDVE RSDDLAC_K T N RtliR KDDWAA_K_H 1 S Y C Y K K R R E Y F_ R C I V'L QILI IAIL FATAT FAT A S M D R V L LFKTR NMYI EV Q.. ~ ,..•. .•...N K F H K SIR E OIL C .. EQFHRAMKYDKDV DKVD-FVHKRVK A A I L E - E V D R E L A H C I LNWE-AATEAASRA S HIQ HE KKL AL-K M - -. A L S T Wi RAilA E C MS ---- Q C 1 S T 0 KItIV L R - L 1 V E - KF S ~ ~ ~ ~ ~.~I .. DI Ell....•. N T Kim. K L L R N Y ySir2p hSIRTI hSIRT4 hSIRT5 hSIRTI ySir2p hSIRTI hSIRT4 hSTRT5 hSIRTI K - D - . I E G.....•.•. F L 0 NSR F H F Qlp K I A A M V A Q K C I I N E L C H R L L L P LID P C GTTLPEALACHEN DVM R L L MAE L I W 526 G 495 314 III L 306 328 Figure 81 240 MITLibraries Document Services Room 14-0551 77 Massachusetts Avenue Cambridge, MA 02139 Ph: 617.253.5668 Fax: 617.253.1690 Email: docs@mit.edu http://libraries.mit.edu/docs DISCLAIMER OF QUALITY Due to the condition of the original material, there are unavoidable flaws in this reproduction. We have made every effort possible to provide you with the best copy available. If you are dissatisfied with this product and find it unusable, please contact Document Services as soon as possible. Thank you. Some pages in the original document contain pictures or graphics that will not scan or reproduce well.