Heat shock proteins and aging in Drosophila melanogaster

Heat shock proteins and aging in Drosophila melanogaster

Seminars in Cell & Developmental Biology 14 (2003) 291–299 Heat shock proteins and aging in Drosophila melanogaster Geneviève Morrow, Robert M. Tangu...

236KB Sizes 0 Downloads 89 Views

Seminars in Cell & Developmental Biology 14 (2003) 291–299

Heat shock proteins and aging in Drosophila melanogaster Geneviève Morrow, Robert M. Tanguay∗ Laboratory of Cellular and Developmental Genetics, Department of Medicine and CREFSIP, Université Laval, Pavillon C.E. Marchand, Ste-Foy, Que., Canada G1K 7P4

Abstract Heat shock proteins (Hsps) are conserved molecular chaperones that are upregulated following exposure to environmental stress and during aging. The mechanisms underlying the aging process are only beginning to be understood. The beneficial effects of Hsps on aging revealed in mild stress and overexpression experiments suggest that these proteins are part of an important cell protection system rather than being unspecific molecular chaperones. Among the Hsps families, small Hsps have the greatest influence on aging and the modulation of their expression during aging in Drosophila suggest that they are involved in pathways of longevity determination. © 2003 Elsevier Ltd. All rights reserved. Keywords: Aging; Oxidative stress; Chaperones; Drosophila; sHsp

1. Introduction Aging is a complex process involving hereditary, environmental and life style factors [1,2]. According to the free radical theory of aging, lifespan is determined by the ability of organisms to cope with random somatic damages induced by reactive oxygen species (ROS), the natural by-products of energy metabolism [3]. Superoxide anions and hydroxyl radicals are highly reactive and thought to be responsible for damages to DNA, lipids and proteins [4,5]. As a result, cellular membrane functions are altered [6,7], mitochondrial DNA is preferentially mutated [8,9] telomeric DNA is less stable [10,11] and proteins are subjected to covalent modifications which can alter their functions [12,13]. Therefore, an imbalance between ROS production and removal will lead to the accumulation of cellular damages and subsequent lifespan shortening [3,14]. Model organisms such as Caenorhabditis elegans and Drosophila melanogaster have been used to try to understand the mechanisms underlying the aging process. Recent data from these systems have been shown to apply to mammals implying a conservation of aging mechanisms [15,16]. Mutants displaying increased longevity in both model organisms have been obtained and analyzed for differences in gene expression during the aging process [17–23]. Interestingly in addition to having an extended longevity, many of these mutants displayed increased thermotolerance and resistance ∗ Corresponding

author. Tel.: +1-418-656-3339; fax: +1-418-656-7176. E-mail address: [email protected] (R.M. Tanguay).

1084-9521/$ – see front matter © 2003 Elsevier Ltd. All rights reserved. doi:10.1016/j.semcdb.2003.09.023

to stress [17,18,21]. Heat shock proteins (Hsps) are molecular chaperones that are coordinately expressed following stress and have been shown to be major determinants in the acquisition of thermotolerance and stress resistance [24,25]. Along with the antioxidant defense system and the ubiquitin/proteasome machinery, Hsps are part of the cell-defense mechanisms to prevent accumulation of protein damages [26–29]. As such, Hsps could be important regulators of the aging process. This view takes support from observations on the modulation of Hsp expression during Drosophila lifespan and points to a beneficial effect of some Hsps on aging. 2. Hsps are involved in multiple cellular processes Hsps are divided in many subfamilies on the basis of their molecular weight and the sequence homology of their members. In Drosophila, where the Hsps were originally discovered, there is Hsp83, a homologue of the Hsp90/100 family found in mammals, the Hsp/Hsc70 complex, Hsp60 and the family of small Hsps (sHsps). Many Hsps have been shown to act as molecular chaperones surveying nascent protein from their release from ribosomes to their trafficking across cell membrane to their final destination [27,29]. In vitro, many Hsps also prevent heat-induced aggregation of proteins and help proper refolding of partially denatured proteins to their native state [26,28]. In addition to their rather general and unspecific chaperone function during de novo synthesis of polypeptides or under stress conditions, Hsps are also involved in diverse specific cellular processes. For

292

G. Morrow, R.M. Tanguay / Seminars in Cell & Developmental Biology 14 (2003) 291–299

Table 1 Functions of Drosophila melanogaster Hsps Families

Genes in Drosophila

Functions

References

sHSP

hsp22, hsp23, hsp26, hsp27, hsp67Ba, hsp67Bb, hsp67Bc, l2efl, CG14207, CG4461, CG7409, CG13133

Thermotolerance Protection against apoptosis

[41,103,104]

HSP60

hsp60, hsp60B, hsp64

Involved in spermatid individualization process Folding of nascent proteins

[37,105–109]

HSP70

hsp70, hsp68, hsc-1, hsc-2, hsc-3, hsc-4, hsc-5

Folding of nascent proteins

[38,86,109–112]

Thermotolerance Involved in clathrin-mediated endocytosis HSP90/100

hsp83

example in yeast, Hsp70 and Hsp104 play a role in snRNP assembly under normal conditions and in their maintenance in heat-stressed cells, providing a splicing tolerance essential for the functioning of the mRNA splicing machinery under stress conditions [30]. Hsp90 is required in various processes ranging from signaling of numerous receptors (steroid, tyrosine kinase) [31,32] and proteins (Raf) [33], telomerase [34] and centrosome assembly [35] to spermatogenesis [36]. Many specialized functions of Hsps have been unveiled by studies in Drosophila (Table 1). Indeed, Hsp60B, one of the three members of the Hsp60 family, has been shown to be required for the spermatid individualization process in male flies [37] and the Hsp70 cognate member Hsc-4 is involved in clathrin-mediated endocytosis [38]. In addition to its well-documented function in receptor activity, Hsp90 also acts as a capacitor of morphological evolution buffering genetic variations [39]. The fly genome encodes 12 putative small Hsps, four of which have been more extensively studied as they display distinctive features [40,41]. While Hsp22 is located in the mitochondrial matrix, Hsp23 and Hsp26 are cytoplasmic and Hsp27 is found in the nucleus [40–42]. Following exposure to environmental stresses such as heat shock, these four main sHsps are coordinately upregulated. However, during normal development and throughout lifespan, each sHsp shows a distinct stage- and cell-specific pattern of expression without coordination [40,41].

3. Cell- and stage-specific expression of Hsps during Drosophila development and lifespan As the expression pattern of shsp genes during embryonic development and metamorphosis has been previously reviewed [40,41], we will here summarize certain aspects of this expression during embryogenesis and adulthood that are pertinent to lifespan. Small Hsps display tissue- and stage-specific embryonic expression. Hsp26, Hsp27 and Hsp83 are the first Hsps to accumulate in early embryos as a

Ensure proper centrosome function Maintain of signal transduction pathways and microtubule effectors Capacitor of morphological evolution

[31–33,36,39,113]

result of the translation of their respective maternal mRNAs [43,44]. The zygotic expression of Hsp23 begins at stage 11 and is restricted to the central nervous system (CNS) where this small Hsp is successively expressed in MP2 neuronal cells, in ventral unpaired median (VUM) cells at stage 13 and in dorsal midline glia (MG) during late embryogenesis (Michaud and Tanguay, in preparation). In stage 11 embryos, there is a restricted accumulation of hsp22 mRNA in metameric ectodermal patches [45], but the corresponding protein has yet to be detected. Other sHsp mRNAs such as l(2)efl [46] and hsp67Bc [47] have also been found to be expressed throughout embryogenesis and larval stages until the beginning of pupation. Another evidence for the tight regulation of sHsps expression during development is illustrated by the hsp67Bb gene. Transcription of this gene gives rise to two mRNA products of different lengths, the shortest (560 bp) being expressed during embryogenesis while the longest (780 bp) is restricted to the male germline [48]. During the larval–pupal metamorphosis, the expression of sHsps is regulated by the steroid molting hormone ecdysone and other enhancer elements. The sHsp mRNAs appear in pulses related to those of the molting hormone [49–54]. Hsp23 accumulates in pupae but is almost totally absent from 1-week-old adult flies [55] except in the fat body at Day 1 and in gonads, where it is specifically expressed in cyst and epithelial cells in addition to spermatid bundles [56]. The stress-induction of Hsp23 is also regulated in a cell-specific manner in some tissues such as the eye where its expression is restricted to a single cell type of the ommatidium, the cone cells while Hsp27 is induced in all cells including the cone, pigment and photoreceptor cells [57]. Hsp26 and Hsp27 are also found in gonads and in the CNS where they display distinct levels of expression [56,58]. Table 2 summarizes the present data on the different cell types where each of the small Hsps is expressed during Drosophila development. Like during embryogenesis and morphogenesis, Hsps expression during aging is closely regulated and the heat-shock response following exposure to stress is also modified.

G. Morrow, R.M. Tanguay / Seminars in Cell & Developmental Biology 14 (2003) 291–299

293

Table 2 Small Hsps of Drosophila are expressed in different cell-types throughout development sHsp

mRNA/ protein

Developmental stages

References

Embryonic

Larval

Adult

Hsp22 Hsp23

mRNA Protein

Metameric ectodermal patches CNS (MP2, VUM, MG cells)

Absent Present

Head and thorax Gonads, fat body, CNS

Hsp26

Protein

CNS, male gonads

CNS, gonads

Hsp27

Protein

Ubiquitous

Male gonads, CNS, ventral ganglion CNS, gonads

[45,61] Michaud and Tanguay (in preparation), [40,50,55,58] [56,58,114]

CNS, gonads

[58,115,116]

Fleming et al. [59] were the first to report that the heat-shock response of young and old flies displayed striking differences in terms of Hsps expression. Young and old flies were submitted to heat shock and allowed to recover in presence of filters saturated with [35 S]methionine for 2 h. Flies were then homogenized and loaded on two-dimensional gel electrophoresis. In young flies, the heat shock treatment resulted in the induction of 14 new proteins whereas in old flies, 50 new proteins were found. Proteins common to the two groups were synthesized at a higher rate in old flies. This clear difference between the response of young and old flies suggests an increased sensitivity to environmental stress. This sensitivity is probably due to accumulation of damaged proteins since young flies fed with canavanine, an arginine analogue used to mimic accumulation of damaged proteins, displayed the same set of induced proteins as old flies [59,60]. The sequence requirement for Hsp22 and Hsp70 expression during aging has also been partially defined by using Hsp22:lacZ and Hsp70:lacZ fusion proteins [61,62]. In both cases the presence of three functional heat shock elements (HSE) has been shown to be required for upregulation of expression during aging. Since the hsp22 promoter is preferentially induced during aging comparatively to other hsp genes, it has been suggested that a specific trans-acting factor other than the heat shock factor (HSF) that binds specifically to HSE could be responsible for aging-dependent induction [61,63]. However the existence of such an aging specific trans-acting factor remains to be proven in Drosophila. Recent data obtained in organisms displaying extended longevity phenotypes are in favor of an endocrine regulation of aging featuring ecdysone and insulin mediated signals [16,64]. According to studies in C. elegans, DAF-16/FOXO is the most probable trans-acting factor that acts in concert with HSF to upregulate sHsps expression during aging [65]. Indeed, it was shown that both Daf-16 and Hsf-1 trigger sHsps expression in the insulin/IGF system and after heat shock [65]. Moreover, consensus sequences for Daf-16 were found in most shsp genes of C. elegans [66]. A gene homologous to FOXO has recently been described in Drosophila [67] but its effect on aging remains to be tested in this organism. In addition to displaying a modified stress response, old flies have an altered expression pattern of Hsps under normal

conditions comparatively to young flies. In a genome-wide study aimed at elucidating transcriptional changes accompanying the aging process in Drosophila melanogaster, the expression of Hsp26 and an Hsp60-like protein have been respectively shown to be down and upregulated [19]. An upregulation of shsp genes, ROS detoxifying enzymes and proteins involved in the antimicrobial response has also been observed in a similar study conducted in C. elegans mutants displaying increased longevity [66]. Thus because Hsps are ubiquitous and play a role in a wide variety of cellular processes by interacting with many different proteins, Macario and Macario [68] have suggested that a chaperone failure would have widespread consequences and might be an intrinsic component of the aging process. Further indirect evidence for a role of the stress response in the aging process stems from the findings that, in many species, aging is associated with a reduced heat-shock response [25,69].

4. The stress response in the aging process In attempts to understand the molecular mechanism of aging in D. melanogaster, genome-wide studies have been performed. By characterizing RNA transcript levels of the whole genome of Drosophila, Pletcher et al. [70] found that nearly 23% of the genome varies in transcripts representation with age. Among them, mRNAs for cytochrome P450s, antibacterial and several proteins involved in the stress response were upregulated [70]. In a similar study, aimed at comparing changes in gene transcripts during aging and following oxidative stress, genes of the reproductive system were found to be downregulated while genes involved in detoxification and chaperoning were upregulated in aging flies [19]. Genes associated with immunity response, microtubule organization and muscle function were also shown to be important in lifespan determination [71]. Comparing the data obtained from transcript analysis during aging in different organism reveals common features of aging among eukaryotes. Hence, while stress and inflammatory response genes are upregulated, energy metabolism and growth genes are downregulated [19,66,71–75]. Therefore, aging does not appear to be associated with a widespread disregulation of

294

G. Morrow, R.M. Tanguay / Seminars in Cell & Developmental Biology 14 (2003) 291–299

gene expression but represents a state of precisely regulated changes [70,75–77]. Consistent with the upregulation of stress response genes during Drosophila lifespan, hsp22 and hsp23 mRNAs and Hsp70 protein were demonstrated to be upregulated in aging flies [61,63]. The increase in Hsp70 expression with age is restricted to flight and leg muscles in the absence of heat-shock as assessed by western blotting and in studies of transgenic flies containing a reporter Hsp70-␤-galactosidase construct [63]. The increase in reporter activity is restricted to these specific tissues as it was not detectable by western blot on whole fly extracts (Fig. 1). The expression profile of sHsps throughout lifespan has also been analyzed by northern blots. A preferential induction of hsp22 mRNA was observed during aging [61]. Interestingly, while hsp22 transcripts were upregulated by 60-fold in the head and 16-fold in thorax of 35 days old comparatively to 6 days old flies, the level of hsp26 and hsp27 transcripts remained constant over this period [61]. This preferential induction of hsp22 in aging is delayed at the protein level as demonstrated by western blotting on whole fly extracts (Fig. 1). Indeed, the level of Hsp22 is higher at 90 days of age than at 35. The level of hsp23 mRNA is upregulated by five-fold in flies thorax between Days 6 and 35 [61], but this is not visible on whole fly extracts (Fig. 1). Interestingly, tissues displaying the preferential upregulation of hsp genes either have an increased sensitivity to stress (like neurons [78]) or are submitted to intensive stress (like flying muscles cells [79]). As mentioned before, old flies have a different stress response than young flies [59]. Consistent with an increased sensitivity of old flies to environmental stresses, Minois and Le Bourg [80] have noted that old flies were more sensitive

Fig. 1. Profile of Hsps expression throughout Drosophila lifespan. Wild-type flies of different ages were homogenized and proteins separated on a 12% SDS–polyacrylamide gel. (A) Coomassie blue staining. (B) Western blotting was performed with anti-Hsp70 #799 (1/5000), anti-Hsp60 #37 (1/10,000), anti-Hsp23 #7B12 (1/100) or anti-Hsp22 #36 (1/10,000) and the corresponding secondary antibody coupled with peroxydase.

to starvation, desiccation, heat and cold stress. This points again to the accumulation of damaged proteins during aging since, as discussed above, young flies fed with canavanine display the same set of induced proteins as old flies [59,60].

Fig. 2. The GAL4/UAS system used in the gene search strategy of Seong et al. [89]. Flies having a P-element insertion containing the yeast GAL4 coding sequence are crossed with different fly lines having a P-element insertion containing UAS sequences at both ends. These GAL4-binding sequences are directed to express sequence flanking the P-element. In the progeny kept at 30 ◦ C, the promoter of hsp70 is on and induces the expression of GAL4, which than binds to UAS sequences and activates transcription of gene nearby the insertion (in the example, gene X).

G. Morrow, R.M. Tanguay / Seminars in Cell & Developmental Biology 14 (2003) 291–299

5. sHsps have beneficial effect on the aging process The first evidence that Hsps could have a beneficial effect on aging was suggested from mild stress experiments, which activate the stress response without causing cellular damages [81]. Giving repeated mild heat stresses has been shown to increase Hsps expression and allow subsequent cell survival to otherwise lethal conditions in addition to extending the organism longevity [81–83]. Intriguingly, all these studies failed to demonstrate a correlation between Hsp70 expression and the increase of lifespan by mild stress [83–85], suggesting that the effect is due to the general stress response rather than to the specific expression of Hsp70 [86]. The effect of overexpressing Hsp70 on longevity in Drosophila also failed to demonstrate a protective effect against aging injuries. Indeed, no increase in the mean lifespan (50% survival) of transgenic flies carrying varying

295

numbers of the heat-inducible gene hsp70 and submitted to heat pulse was observed [87,88]. However, flies heat-treated for up to 30 min at 36 ◦ C displayed 10–30% increase in survival in the 2 weeks following the treatment [87]. In order to explain this beneficial effect, Tatar et al. [87] have suggested that Hsp70 could refold damaged proteins and/or interact with other stress-response mechanisms such as with superoxide dismutase enzymes. While the overexpression of Hsp70 failed to substantially increase lifespan, overexpressing the small Hsps seems to be more advantageous. The sole overexpression of Hsp26 in all cells leads to a 15% increase in Drosophila mean lifespan [89]. This phenotype was obtained in a study aimed at identify genes crucial for extension of longevity by mean of a gene search system. This system takes advantage of the GAL4/UAS system, which allows a tight control of targetted gene expression (Fig. 2). In these experiments of Seong et al.

Fig. 3. Hsf and Daf-16 act together to promote lifespan. In C. elegans, under normal growth conditions, Daf-2 signaling results in the phosphorylation of Daf-16 and its subsequent sequestration in the cytoplasm. The involvement of Daf-2 signaling in the inactivation of Hsf have been suggested by Hsu et al. [65] but is not yet clear. In long-lived mutants for daf-2, the phosphorylation of Daf-16 is inhibited resulting in its translocation to the nucleus where it activates the expression of proteins promoting life together with Hsf [65,66]. Daf-16 could regulate the expression of a set of genes on its own (sod-3, mtl-1) while it could act in concert with Hsf to upregulate shsp expression and other genes. Hsf could also be responsible of the transcription of other genes such as hsp70 [65,66].

296

G. Morrow, R.M. Tanguay / Seminars in Cell & Developmental Biology 14 (2003) 291–299

[89] GAL4 expression is driven by an hsp70 promoter and activates UAS sequences of another P-element, which has been inserted randomly. As the UAS sequence is directed to express sequences flanking the P-element, the extension of longevity can be inferred to the overexpression of the gene nearby the P-element. These genes can then be identified by gene rescue. The hsp26 gene was isolated among 25 cDNA obtained from flies displaying increased longevity. Preliminary results in our laboratory also argue for a beneficial effect of sHsps overexpression on aging. While overexpressing Hsp22 ubiquitously using an actin driver resulted in over 30% increase in mean lifespan, overexpressing Hsp23 led to an increase similar (∼15%) to the one reported for Hsp26 (data not shown). Moreover, Hsp22 appears to act on early aging events as the 30% increase in Hsp22 overexpressing flies occurs in their pre-mortality phase (<10% mortality). The 15% increase with Hsp23 was not observed in the pre-mortality phase suggesting that sHsps can affect the aging process by different pathways. As Hsp22, Hsp23 and Hsp26 have distinct intracellular localization (Hsp22 is mitochondrial [42] while Hsp23 and Hsp26 are cytoplasmic [41]), their different efficiency to increase lifespan could reflect the difference in alteration of cell compartment during aging. Interestingly, the upregulation of two mitochondrial Hsps (Hsp22 and a Hsp60-like protein) during aging in D. melanogaster [19,61] is in agreement with the hypothesis that mitochondrial proteins are particularly important in the aging process likely because of their sensitivity to oxidative stress during aging [90,91]. This also supports the idea that mitochondria are key targets of the aging process [9]. As the energy factory of cells, mitochondria play a vital role in many processes such as development, morphogenesis and aging. In addition it is now evident that mitochondria are a key element of the apoptotic process [92–94] that also has influence on aging processes [95,96]. The results obtained with the overexpression of sHsps argue in favor of an increase in protein damages during aging since sHsps are molecular chaperones preventing protein aggregation and maintaining misfolded protein in a refoldable state [28]. A beneficial role of sHsps on the aging process has also been reported in the worm C. elegans, in which the insulin-like signaling pathway known to be a key determinant of lifespan [97,98] has been demonstrated to control the expression of Hsp16 [99]. Walker and Lithgow [99] have shown that introducing extra copies of the gene encoding Hsp16 was sufficient to confer thermotolerance and increase worms longevity. Moreover, they demonstrated that the DAF-16 transcription factor, which is regulated by the insulin-like signaling pathway [97], was required for maximal expression of the sHsp. To assess that Hsp16 was involved in lifespan extension, Hsu et al. [65] performed RNAi experiments targeting the sHsp. As expected, worms submitted to Hsp16 RNAi displayed decreased lifespan as compared to wild-type worms [65]. In a similarly designed experiment, Hsf, which is known to be responsible for the heat-induction of Hsps expression, has been shown to be

involved in the insulin/IGF-1 system [65]. Fig. 3 illustrates a model to link Hsps and the insulin/IGF-1 system to aging. Small Hsps could exert their beneficial effects by preventing aggregation of damaged/oxidized proteins [65].

6. Conclusion Hsps may not be simple unspecific molecular chaperones as their presence is required for the maintenance of specific cellular functions during stress but also through normal lifespan. Modulating the heat-shock response could have many advantages, notably in the treatment or prevention of misfolded diseases. Arguing in that sense, recent work in Drosophila models of Parkinson’s and Poly(Q) diseases have demonstrated the beneficial effect of overexpressing Hsp70 on aggregates toxicity [100–102]. Using sHsp RNAi methods in C. elegans, a beneficial role of sHsps in the onset delay of Poly(Q) diseases has also been demonstrated [65].

Acknowledgements Work in the author’s laboratory was supported by a grant from the Canadian Institutes of Health Research (CIHR) to R.M.T. and a studentship to G.M.

References [1] Sun J, Tower J. FLP recombinase-mediated induction of Cu/Zn-superoxide dismutase transgene expression can extend the lifespan of adult Drosophila melanogaster flies. Mol Cell Biol 1999;19:216–28. [2] Vieira C, Pasyukova EG, Zeng ZB, Hackett JB, Lyman RF, Mackay TF. Genotype–environment interaction for quantitative trait loci affecting lifespan in Drosophila melanogaster. Genetics 2000;154:213–27. [3] Sohal RS. Role of oxidative stress and protein oxidation in the aging process. Free Rad Biol Med 2002;33:37–44. [4] Raha S, Robinson BH. Mitochondria, oxygen free radicals, disease and ageing. Trends Biochem Sci 2000;25:502–8. [5] Raha S, Robinson BH. Mitochondria, oxygen free radicals, and apoptosis. Am J Med Genet 2001;106:62–70. [6] Hrelia S, Fiorentini D, Maraldi T, Angeloni C, Bordoni A, Biagi PL, et al. Doxorubicin induces early lipid peroxidation associated with changes in glucose transport in cultured cardiomyocytes. Biochim Biophys Acta 2002;1567:150–6. [7] Park JE, Yang JH, Yoon SJ, Lee JH, Yang ES, Park JW. Lipid peroxidation-mediated cytotoxicity and DNA damage in U937 cells. Biochimie 2003;84:1198–204. [8] Chomyn A, Attardi G. MtDNA mutations in aging and apoptosis. Biochem Biophys Res Commun 2003;304:519–29. [9] Sastre J, Pallardó FV, Viña J. The role of mitochondrial oxidative stress in aging. Free Rad Biol Med 2003;35:1–8. [10] Djojosubroto MW, Choi YS, Lee HW, Rudolph KL. Telomeres and telomerase in aging, regeneration and cancer. Mol Cell 2003;15:164–75. [11] Liu L, Trimarchi JR, Navarro P, Blasco MA, Keefe DL. Oxidative stress contributes to arsenic-induced telomere attrition, chromosome instability and apoptosis. J Biol Chem 2003;278:31998–2004.

G. Morrow, R.M. Tanguay / Seminars in Cell & Developmental Biology 14 (2003) 291–299 [12] Wright HT. Nonenzymatic deamination of asparaginyl and glutaminyl redidues in proteins. Crit Rev Biochem Mol Biol 1991;26:1– 52. [13] Sun H, Gao J, Ferrington DA, Biesiada H, Williams TD, Squier TC. Repair of oxidized calmodulin by methionine sulfoxide reductase restrores ability to activate the plasma membrane Ca-ATPase. Biochemistry 1999;38:105–12. [14] Butov A, Johnson T, Cypser J, Sannikov I, Volkov M, Sehl M, et al. Hormesis and debilitation effects in stress experiments using the nematode worm Caenorhabditis elegans: the model of balance between cell damage and HSP levels. Exp Gerontol 2001;37:57– 66. [15] Hozenberg M, Dupont J, Ducos B, Leneuve P, Géloën A, Even PC, et al. IGF-1 receptor regulates lifespan and resistance to oxidative stress in mice. Nature 2003;421:182–7. [16] Tatar M, Bartke A, Antebi A. The endocrine regulation of aging by insulin-like signals. Science 2003;299:1346–51. [17] Lin YJ, Seroude L, Benzer S. Extended life-span and stress resistance in the Drosophila mutant methuselah. Science 1998;282: 943–6. [18] Rogina B, Reenan RA, Nilsen SP, Helfand SL. Extended life-span conferred by cotransporter gene mutations in Drosophila. Science 2000;290:2137–40. [19] Zoo S, Meadows S, Sharp L, Jan LY, Nung Jan Y. Genome-wide study of aging and oxidative stress response in Drosophila melanogaster. Proc Natl Acad Sci USA 2000;97:13726–31. [20] Chavous DA, Jackson FR, O’Connor CM. Extension of the Drosophila lifespan by overexpression of a protein repair methyltransferase. Proc Natl Acad Sci USA 2001;98:14814–8. [21] Ekengren S, Tryselius Y, Dushay MS, Liu G, Steiner H, Hultmark D. A humoral stress response in Drosophila. Curr Biol 2001;11:714–8. [22] Seong KH, Matsuo T, Fuyama Y, Aigaki T. Neural-specific overexpression of Drosophila plenty of SH3s DPOSH extends the longevity of adult flies. Biogerontology 2001;2:271–81. [23] Tatar M, Kopelman A, Epstein D, Tu MP, Yin CM, Garofalo RS. A mutant Drosophila insulin receptor homolog that extends life-span and impairs neuroendocrine function. Science 2001;292:107–10. [24] Feder ME, Hofmann GE. Heat-shock proteins, molecular chaperones, and the stess response: evolutionary and ecological physiology. Annu Rev Physiol 1999;61:243–82. [25] Verbeke P, Fonager J, Clark BFC, Rattan SIS. Heat-shock response and ageing: mechanisms and applications. Cell Biol Int 2001;25:845–57. [26] Ehrnsperger M, Gaestel M, Buchner J. Analysis of chaperone properties of small Hsp’s. Methods Mol Biol 2000;99:421–9. [27] Hartl FU, Hayer-Hartl M. Molecular chaperones in the cytosol: from nascent chain to folded protein. Science 2002;295:1852–8. [28] Haslbeck M. sHsps and their role in the chaperone network. Cell Mol Life Sci 2002;59:1649–57. [29] Walter S, Buchner J. Molecular chaperones—cellular machines for protein folding. Angew Chem Int Ed Engl 2002;41:1098–113. [30] Bracken AP, Bond U. Reassembly and protection of small nuclear ribonucleoprotein particles by heat shock proteins in yeast cells. RNA 1999;5:1586–96. [31] Cutforth T, Rubin GM. Mutations in Hsp83 and cdc37 impair signaling by the sevenless receptor tyrosine kinase in Drosophila. Cell 1994;77:1027–36. [32] Nathan DF, Lindquist S. Mutational analysis of Hsp90 function: interactions with a steroid receptor and a protein kinase. Mol Cell Biol 1995;15:3917–25. [33] Van der Straten A, Rommel C, Dickson B, Hafen E. The heat shock protein 83 Hsp83 is required for Raf-mediated signalling in Drosophila. EMBO J 1997;16:1961–9. [34] Holt SE, Aisner DL, Baur J, Tesmer VM, Dy M, Ouellette M, et al. Functional requirement of p23 and Hsp90 in telomerase complexes. Genes Dev 1999;13:817–26.

297

[35] Helmbrecht K, Zeise E, Rensing L. Chaperones in cell cycle regulation and mitogenic signal transduction: a review. Cell Prolif 2000;33:341–65. [36] Yue L, Karr TL, Nathan DF, Swift H, Srinivasan S, Lindquist S. Genetic analysis of viable Hsp90 alleles reveals a critical role in Drosophila spermatogenesis. Genetics 1999;151:1065–79. [37] Timakov B, Zhang P. The hsp60B gene of Drosophila melanogaster is essential for the spermatid individualization process. Cell Stress Chaperones 2001;6:71–7. [38] Chang HC, Newmyer SL, Hull MJ, Ebersold M, Schmid SL, Mellman I. Hsc70 is required for endocytosis and clathrin function in Drosophila. J Cell Biol 2002;159:477–87. [39] Rutherford SL, Lindquist S. Hsp90 as a capacitor for morphological evolution. Nature 1998;396:336–42. [40] Michaud S, Marin R, Tanguay RM. Regulation of heat shock gene induction and expression during Drosophila development. Cell Mol Life Sci 1997;53:104–13. [41] Michaud S, Morrow G, Marchand J, Tanguay RM. Drosophila small heat shock proteins: cell and organelle-specific chaperones? Prog Mol Subcell Biol 2002;28:79–101. [42] Morrow G, Inaguma Y, Kato K, Tanguay RM. The small heat shock protein Hsp22 of Drosophila melanogaster is a mitochondrial protein displaying oligomeric organization. J Biol Chem 2000;275:31204–10. [43] Graziosi G, Micali F, Marzari R, De Cristini F, Savoini A. Variability of response of early Drosophila embryos to heat shock. J Exp Zool 1980;214:141–5. [44] Zimmerman JL, Petri W, Meselson M. Accumulation of a specific subset of D. melanogaster heat shock mRNAs in normal development without heat shock. Cell 1983;32:1161–70. [45] Leemans R, Egger B, Loop T, Kammermeier L, He H, Hartmann B, et al. Quantitative transcript imaging in normal and heat-shocked Drosophila embryos by using high-density oligonucleotide arrays. Proc Natl Acad Sci USA 2000;97:12138–43. [46] Kurzik-Dumke U, Lohman E. Sequence of the new Drosophila melanogaster small heat-shock-related gene, lethal(2) essential for life [l(2)efl], at locus 59F4,5. Gene 1995;154:171–5. [47] Pauli D, Tonka CH. A Drosophila heat shock gene from locus 67B is expressed during embryogenesis and pupation. J Mol Biol 1987;198:235–40. [48] Pauli D, Tonka CH, Ayme-Southgate A. An unusual split Drosophila heat shock gene expressed during embryogenesis. J Mol Biol 1988;200:47–53. [49] Riddiford LM, Cherbas P, Truman JW. Ecdysone receptors and their biological actions. Vitam Horm 2000;60:1–73. [50] Cheney CM, Shearn A. Developmental regulation of Drosophila imaginal disc proteins: synthesis of a heat-shock protein under non-heat-shock conditions. Dev Biol 1983;95:325–30. [51] Arrigo AP, Pauli D. Characterization of Hsp27 and three immunologically related polypeptides during Drosophila development. Exp Cell Res 1988;175:169–83. [52] Pauli D, Tissières A. Developmental expression of the heat shock genes in Drosophila melanogaster. In: Morimoto R, Tissières A, Georgopoulos C, editors. Stress proteins in biology and medicine. New York: Cold Spring Harbour Laboratory Press; 1990. p. 361– 78. [53] Vazquez J. Response to heat shock of gene 1, a Drosophila melanogaster small heat shock gene, is developmentally regulated. Mol Gen Genet 1991;226:393–400. [54] Dubrovsky EB, Dretzen G, Bellard M. The Drosophila broadcomplex regulates developmental changes in transcription and chromatin structure of the 67B heat-shock gene cluster. J Mol Biol 1994; 241:353–62. [55] Arrigo AP. Cellular localization of HSP23 during Drosophila development and following subsequent heat shock. Dev Biol 1987;122:39–48.

298

G. Morrow, R.M. Tanguay / Seminars in Cell & Developmental Biology 14 (2003) 291–299

[56] Marin R, Valet JP, Tanguay RM. hsp23 and hsp26 exhibit distinct spatial and temporal patterns of constitutive expression in Drosophila adults. Dev Genet 1993;14:69–77. [57] Marin R, Demers M, Tanguay R. Cell-specific heat-shock induction of Hsp23 in the eye of Drosophila melanogaster. Cell Stress and Chaperones 1996;1:40–6. [58] Marin R, Landry J, Tanguay RM. Tissue-specific post-translational modification of the small heat shock protein Hsp27 in Drosophila. Exp Cell Res 1996;223:1–8. [59] Fleming JE, Walton JK, Dubitski R, Bensch KG. Aging results in an unusual expression of Drosophila heat shock proteins. Proc Natl Acad Sci USA 1988;85:4099–103. [60] Niedzwiecki A, Kongpachith AM, Fleming JE. Aging affects expression of 70-kDa heat shock proteins in Drosophila. J Biol Chem 1991;266:9332–8. [61] King V, Tower J. Aging-specific expression of Drosophila Hsp22. Dev Biol 1999;207:107–18. [62] Wheeler JC, King V, Tower J. Sequence requirements for upregulated expression of Drosophila hsp70 transgenes during aging. Neurobiol Aging 1999;20:545–53. [63] Wheeler JC, Bieschke ET, Tower J. Muscle-specific expression of Drosophila Hsp70 in response to aging and oxidative stress. Proc Natl Acad Sci USA 1995;92:10408–12. [64] Simon AF, Shih C, Mack A, Benzer S. Steroid control of longevity in Drosophila melanogaster. Science 2003;299:1407–10. [65] Hsu AL, Murphy CT, Kenyon C. Regulation of aging and age-related disease by DAF-16 and heat-shock factor. Science 2003;300:1142–5. [66] Murphy CT, McCarroll SA, Barmann CI, Fraser A, Kamath RS, Ahringer J, et al. Genes that act downstreasm of DAF-16 to influence the lifespan of Caenorhabiditis elegans. Nature 2003;424:277–83. [67] Kramer JM, Dadidge JT, Lockyer JM, Staveley BE. Expression of Drosophila FOXO regulates growth and can phenocopy starvation. BMC Dev Biol 2003;3:5, http://www.biomedcentral.com/1471213X/3/5. [68] Macario AJL, Macario EC. Sick chaperones and ageing: a perspective. Ageing Res Rev 2002;1:295–311. [69] Lee YK, Manalo D, Liu AY. Heat shock response. Biol Signals 1996;5:180–91. [70] Pletcher SD, Macdonald SJ, Marguerie R, Certa U, Stearns SC, Goldstein DB, et al. Genome-wide transcripts profiles in aging and calorically restricted Drosophila melanogaster. Curr Biol 2002;12:712–23. [71] Seroude L, Brummel T, Kapahi P, Benzer S. Spatio-temporal analysis of gene expression during aging in Drosophila melanogaster. Aging Cell 2002;1:47–56. [72] Lee CK, Klopp RG, Weindruch R, Prolla TA. Gene expression profile of aging and its retardation by caloric restriction. Science 1999;285:1390–3. [73] Kayo T, Allison DB, Weindruch R, Prolla TA. Influences of aging and caloric restriction of the transcriptional profile of skeletal muscle from rhesus monkeys. Proc Natl Acad Sci USA 2001;98:5093–8. [74] Lee CK, Weikdruch R, Prolla TA. Gene expression profile of the aging brain in mice. Nat Genet 2000;25:294–7. [75] Seroude L. Differential gene expression and aging. Scientific World J 2002;2:618–31. [76] Rogina B, Helfand SL. Regulation of gene expression is linked to life span in adult Drosophila. Genetics 1995;141:1043–8. [77] Helfand SL, Rogina B. Molecular genetics of aging in the fly: is this the end of the beginning. BioEssays 2003;25:134–41. [78] Boulianne GL. Neuronal regulation of lifespan: clues from flies and worms. Mech Ageing Dev 2001;122:883–94. [79] Sohal RS, Brunk UT. Mitochondrial production of pro-oxidants and cellular senescence. Mutat Res 1992;275:295–304. [80] Minois N, Le Bourg E. Resistance to stress as a function of age in Drosophila melanogaster living in hypergravity. Mech Ageing Dev 1999;109:53–64.

[81] Minois N. Longevity and aging: beneficial effects of exposure to mild stress. Biogerontology 2000;1:15–29. [82] Khazaeli AA, Tatar M, Pletcher SD, Curtsinger JW. Heat-induced longevity extension in Drosophila. I. Heat treatment, mortality, and thermotolerance. J Gerontol A: Biol Sci Med Sci 1997;52:B48– 52. [83] Le Bourg E, Valenti P, Lucchetta P, Payre F. Effects of mild heat shocks at young age on aging and longevity in Drosophila melanogaster. Biogerontology 2001;2:155–64. [84] Le Bourg E, Valentin P, Payre F. Lack of hypergravity-associated longevity extension in Drosophila melanogaster flies overexpressing hsp70. Biogerontology 2002;3:355–64. [85] Hercus MJ, Loeschcke V, Rattan SI. Lifespan extension of Drosophila melanogaster through hormesis by repeated mild heat stress. Biogerontology 2003;4:149–56. [86] Sorensen JG, Loeshcke V. Larval crowding in Drosophila melanogaster induces Hsp70 expression, and leads to increased adult longevity and adult thermal stress resistance. J Insect Physiol 2001;47:1301–7. [87] Tatar M, Khazaeli AA, Curtsinger JW. Chaperoning extended life. Nature 1997;390:30. [88] Minois N, Khazaeli AA, Curtsinger JW. Locomotor activity as a function of age and lifespan in Drosophila melanogaster overexpressing hsp70. Exp Gerontol 2001;36:1137–53. [89] Seong K-H, Ogashiwa T, Matsuo T, Fuyama Y, Aigaki T. Application of the gene search system to screen for longevity genes in Drosophila. Biogerontology 2001;2:209–17. [90] Yan LJ, Sohal RS. Mitochondrial adenine nucleotide translocase is modified oxidatively during aging. Proc Natl Acad Sci USA 1998;95:12896–901. [91] Das N, Levine RL, Orr WC, Sohal RS. Selectivity of protein oxidative damage during aging in Drosophila melanogaster. Biochem J 2001;360:209–16. [92] Granville DJ, Gottlieb RA. Mitochondria: regulators of cell death and survival. Scientific World J 2002;2:1569–78. [93] Kroemer G. Mitochondrial control of apoptosis: an introduction. Biochem Biophys Res Commun 2003;304:433–5. [94] Mayer B, Oberbauer R. Mitochondrial regulation of apoptosis. News Physiol Sci 2003;18:89–94. [95] Sastre J, Pallardo FV, Vina J. Mitochondrial oxidative stress plays a key role in aging and apoptosis. IUBMB Life 2000;49:427–35. [96] Pollack M, Leeuwenburgh C. Apoptosis and aging: role of the mitochondria. J Gerontol A: Biol Sci Med Sci 2001;11:B475–82. [97] Lee SS, Kennedy S, Tolonen AC, Ruvkun G. DAF-16 target genes that control C. elegans life-span and metabolism. Science 2003;300:644–7. [98] Munoz MJ. Longevity and heat stress regulation in Caenorhabditis elegans. Mech Ageing Dev 2003;124:43–8. [99] Walker GA, Lithgow GJ. Lifespan extension in C. elegans by a molecular chaperone dependent upon insulin-like signal. Aging Cell 2003;2:131–9. [100] Auluck PK, Chan HYE, Trojanowski JQ, Lee VM-Y, Bonini NM. Chaperone suppression of ␣-synuclein toxicity in a Drosophila model for Parkinson’s disease. Science 2002;295:865–8. [101] Bonini NM. Chaperoning brain degeneration. Proc Natl Acad Sci USA 2002;99:16407–11. [102] Chan HYE, Warrick JM, Andriola I, Merry D, Bonini NM. Hum Mol Genet 2002;11:2895–904. [103] Rollet E, Lavoie JN, Landry J, Tanguay RM. Expression of Drosophila’s 27 kDa heat shock protein into rodent cells confers thermal resistance. Biochem Biophys Res Comm 1992;185:116– 20. [104] Mehlen P, Kretz-Remy C, Préville X, Arrigo AP. Human Hsp27, Drosophila Hsp27 and human ␣B-crystallin expression-mediated increase in glutathione is essential for the protective activity of these proteins against TNF␣-induced cell death. EMBO J 1996;15:2695– 706.

G. Morrow, R.M. Tanguay / Seminars in Cell & Developmental Biology 14 (2003) 291–299 [105] Lakhotia SC, Singh BN. Synthesis of a ubiquitously present new HSP60 family protein is enhanced by heat shock only in the Malpighian tubules of Drosophila. Experientia 1996;52: 751–6. [106] Adams MD, Celniker SE, Holt RA, Evans CA, Gocayne JD, Amanatides PG, et al. The genome sequence of Drosophila melanogaster. Science 2000;287:2179–85. [107] Rubin GM, Yandell MD, Wortman JR, Gabor Miklos GL, Nelson CR, Hariharan IK. Comparative genomics of the eukaryotes. Science 2000;287:2204–15. [108] Lakhotia SC, Srivastava P, Prasanth KV. Regulation of heat shock proteins, Hsp70 and Hsp64, in heat-shocked Malpighian tubules of Drosophila melanogaster larvae. Cell Stress Chaperones 2002;7:347–56. [109] Bukau B, Horwich AL. The Hsp70 and Hsp60 chaperone machines. Cell 1998;92:351–66. [110] Craig EA, Ingolia TD, Manseau LJ. Expression of Drosophila heat-shock cognate genes during heat shock and development. Dev Biol 1983;99:418–26.

299

[111] Rubin DM, Mehta AD, Zhu J, Shoham S, Chen X, Wells QR, et al. Genomic structure and sequence analysis of Drosophila melanogaster HSC70 gnes. Gene 1993;128:155–63. [112] Krebs RA, Feder ME. Hsp70 and larval thermotolerance in Drosophila melanogaster: how much is enough and when is more too much? J Insect Physiol 1998;44:1091–101. [113] Lange BM, Bachi A, Wilm M, Gonzalez C. Hsp90 is a core centrosomal component and is required at different stages of the centrosome cycle in Drosophila and vertebrates. EMBO J 2000;19:1252– 62. [114] Glaser RL, Wolfner MF, Lis JT. Spatial and temporal pattern of hsp26 expression during normal development. EMBO J 1986;4:747– 54. [115] Pauli D, Tonka CH, Tissieres A, Arrigo AP. Tissue-specific expression of the heat shock protein HSP27 during Drosophila melanogaster development. J Cell Biol 1990;111:817–28. [116] Marin R, Tanguay RM. Stage-specific localization of the small heat shock protein Hsp27 during oogenesis in Drosophila melanogaster. Chromosoma 1996;105:142–9.