• Aucun résultat trouvé

Developmental regulation of DNA replication initiation in Drosophila

N/A
N/A
Protected

Academic year: 2021

Partager "Developmental regulation of DNA replication initiation in Drosophila"

Copied!
156
0
0

Texte intégral

(1)

Developmental Regulation of DNA Replication Initiation

in Drosophila

by

Fang Xie

B.S. in Biology (2001) Beijing University, Beijing, China Submitted to the Department of Biology

in Partial Fulfillment of the Requirement for the Degree of Doctor of Philosophy in Biology

at the

Massachusetts Institute of Technology August, 2007

© 2007 Fang Xie. All rights reserved.

The author hereby grants to MIT permission to reproduce and to distribute publicly paper and electronic copies of this thesis document in whole or in part

in any medium now known or hereafter created.

Signature of Author……… Department of Biology August 17, 2007 Certified by………...………. Terry L. Orr-Weaver Professor of Biology Thesis Supervisor Accepted by………..………. Stephen P. Bell Chair, Committee of Graduate Students Department of Biology

(2)

Developmental Regulation of DNA Replication Initiation

in Drosophila

by Fang Xie

Submitted to the Department of Biology

on August 17, 2007 in Partial Fulfillment of the Requirement for the Degree of Doctor of Philosophy in Biology

ABSTRACT

Developmental gene amplification in the ovarian follicle cells of Drosophila provides a powerful system for the study of metazoan DNA replication. Amplification produces 100kb gradients of amplified DNA through repeated rounds of origin firing and bidirectional movement of replication forks from these origins. The Drosophila Follicle Cell Amplicon at the cytological location 62D, DAFC-62D, is uniquely regulated, with two separate rounds of amplification in developmental stages 10 and 13 of egg chamber development. We investigated mechanisms that control the unusual timing of DAFC-62D origin activation. We first defined origin sequences in DAFC-62D by analyzing the amount of nascent replicative DNA across this amplicon. Surprisingly, the origin coincides with the coding region of a gene named yellow-g2. ORC2 localizes to the origin, as well as two other sites that do not confer origin activity. Both ORC2 and MCM2-7 display differential association with these sequences, corresponding to the two rounds of amplification. All three elements, dispersed in a 7kb central amplified region, are required for either round of DAFC-62D amplification, because deleting any one completely abolished amplification in transposon experiments. Preceded by transcription

yellow-g2 in stage 12, the late round of origin firing was ablated by the RNAPII inhibitor

α-amanitin. This effect was absent from other amplicons and insulated transposons, and was stage-13 specific for amplification at either the endogenous DAFC-62D or

heterologous transposons that did not have functional insulators. Therefore amplification at DAFC-62D in late follicle cell differentiation depends on transcription in cis.

Molecularly, blocking RNAPII transcription compromises MCM2-7 recruitment. Additional transposon and histone modification analyses confirmed the involvement of RNAPII in amplification control, which may be facilitated by favorable chromatin structure. This work provides insights in developmental regulation of origin firing, revealing one mechanism for initiation of metazoan DNA replication: recruitment of MCM2-7 by RNA polymerase II transcription.

Thesis Suporvisor: Terry L. Orr-Weaver Title: Professor of Biology

(3)

Dedicated to Lin Li  李 林 Xingui Liang 梁新桂 and Shangfa Xie 谢尚发

(4)

Acknowledgements

This thesis work would have been impossible without the full support of my advisor Terry Orr-Weaver. She accepted me (and my ideas) with an open mind, her guidance throughout the years has made graduate school less obscure, and she always inspires me to achieve more. I thank her for being the best advisor I could ever have asked for.

I am thankful to all past and current Orr-Weaver lab members with whom I shared numerous memorable moments, scientifically and non-scientifically. Julie Claycomb made the initial observations and lent tremendous help in establishing the DAFC-62D project. Discussions with Eugenia Park and Jane Kim, the replication subgroup people, have been inspiring. Everyone else, especially Tama Resnick, Jillian Pesin, David Doroquez, Yingdee Unhavaithaya, Lena Kashevsky and Raissa Formina, have made the TOW zone such a fun workplace.

My current thesis committee members, Steve Bell, Jianzhu Chen, and Troy Littleton, have been fantastic mentors and incredibly supportive. I am grateful for Professor Nick Dyson’s tremendous help as the outside member of my defense

committee. I also thank Ilaria Rebay and Paul Garrity for their advice in past committee meetings. The regular Drosophila replication meetings with David MacAlpine, Cary Lai and Steve Bell have been a constant driving force of my research.

I would like to express my gratitude to my husband, Lin Li, and my parents, Xingui Liang and Shangfa Xie, for their unconditional love. I never said “thank you” enough, and could never thank you enough. You are my rock. You will stay my rock. In the pursuit of my dreams.

(5)

TABLE OF CONTENTS Chapter One

Introduction: Activation Control of Replication and Amplification Origins 6

Gene amplification as a model for DNA replication 7

Origins of DNA replication 11

Origins of developmental gene amplification 15

Amplification control elements 23

The involvement of transcription factors in gene amplification 27

Chromatin context and amplification activity 34

A general link between replication and transcription 36

Summary of thesis 40

Chapter Two

Identification of a Drosophila Replication Origin

Developmentally Controlled by Transcription 50

Summary 51

Introduction 52

Results 55

Identification of the replication origin and ORC binding sites in DAFC-62D 55

Differential pre-RC binding in DAFC-62D 59

ORC-binding sequences are required for amplification 62 The two rounds of origin firing are interspersed by transcription 66 α-amanitin specifically inhibits DAFC-62D stage 13 amplification 70 Inhibition of transcription affects MCM2-7 localization 75

Discussion 78

Experimental procedures 86

Chapter Three

Conclusions and Future Directions 101

Differential localization of pre-RC 103

Transcriptional regulation of replication initiation 106

Distinct mechanisms of replication regulation 109

Insulators and their insensitivity to α-amanitin 112

Transcription factories 114

Appendix One: Analyses of the ACE3-ori62 Transposon 120

Appendix Two: Histone Acetylation and Amplification Activity 136

(6)

Chapter One

Introduction:

(7)

Gene amplification as a model for DNA replication

Developmental gene amplification is a process that increases the number of DNA molecules as template for transcription at specific developmental points. It has been reported in a variety of organisms as an alternate strategy to produce large quantities of transcripts over relatively short periods of time. Among the first observed examples is the amplification (about a thousand fold) of the genes that code for ribosomal RNA (rRNA) during the development of Xenopus oocytes, in order to stockpile the egg with the translational machinery necessary for rapid embryonic development (Brown and Dawid, 1968; Gall, 1968). Electron microscopy studies suggest that the Amphibian rDNA is amplified via a rolling-circle mechanism (Hourcade et al., 1973). Another example of extrachromosomal gene amplification is the rDNA in the transcriptionally active macronucleus of the protist Tetrahymena (Gall, 1974; Yao et al., 1979). The 21kb minichromosome in the form of a palindrome comprises two copies of the rDNA and is amplified up to 10,000 copies. This differs from the Amphibian oocyte rDNA however, because these palindromic minichromosomes are not produced by a rolling-circle

mechanism, but rather by bidirectional movement of the replication forks initiated from a defined origin (Figure 1A) (Kapler et al., 1996; Prescott, 1994).

Dipteran flies, including Rhynchosciara americana (Glover et al., 1982),

Bradysia hygida (Laicine et al., 1984), and Sciara coprophila (Wu et al., 1993) all utilize

gene amplification at multiple loci throughout the genome in the larval salivary glands, presumably for the production of large quantities of the structural proteins for the

(8)

construction of the cocoon. Note that unlike Tetrahymena rDNA amplification, in these organisms the gene clusters are replicated above the copy number of surrounding

sequences without forming extrachromosomal molecules. The same strategy is employed by another Dipteran fly, Drosophila melanogaster, to amplify at least four groups of genes in the ovarian follicle cells (Claycomb et al., 2004; Spradling, 1981; Spradling et al., 1980). Two of these gene clusters contain genes that encode the major structural proteins of the chorion (eggshell) (Spradling et al., 1980). It is possible that in Dipteran flies the intrachromosomal structures generated by the amplification process may be tolerated, as both the larval salivary gland and ovarian follicle cells are terminally differentiated tissues and are lost during further development. Because these cell types are nondividing, genomic aberrations accumulated during developmental gene

amplification would not be passed on to daughter cells.

Developmental gene amplification in both Tetrahymena and Dipteran flies has been consistently shown to utilize the normal replication machinery to repeatedly initiate DNA replication from dominant origins, resulting in an “onionskin” structure of nested replication bubbles/forks (Figure 1B) (Claycomb and Orr-Weaver, 2005; Tower, 2004). This provides an advantageous model for studying metazoan DNA replication for several reasons. First, amplified regions (amplicons) are relatively well defined especially given the recent employment of Comparative Genomic Hybridization (CGH) arrays (Claycomb et al., 2004). Second, the repeated firing generates a gradient of DNA copy number with the central origin(s) being the most abundant (Claycomb et al., 2002). This allows

(9)

Figure 1. Models of bidirectional replication versus amplification.

(A) Replication origin fires once and only once per cell cycle, followed by bidirectional movement (elongation) of replication forks. (B) During developmental amplification, the origin is activated multiple times consecutively. The resulting multiple replication forks form an “onionskin” structure, with the highest DNA copy number at the amplification origin.

(10)
(11)

and provides insights into the mechanisms by which the usual once-per-cell-cycle control of DNA replication can be thwarted. Third, developmental gene amplification occurs at strategic developmental times, providing the opportunity to study how DNA replication responds to developmental cues. Finally, a range of molecular and genetic tools are available in these model organisms.

Origins of DNA replication

The best-studied eukaryotic origins are those in the yeast Saccharomyces

cerevisiae. Specific, well-defined origins of DNA replication have been revealed

primarily through genetic analyses (Newlon and Theis, 1993). Consisting of an 11bp A-T-rich autonomously replicating sequence (ARS) consensus sequence (ACS) and other elements (B1 and B2) (Figure 2A), the yeast replication origins first recruit a variety of factors known as the pre-replication complex (pre-RC) (Figure 2A) to initiate DNA replication. As a component of the pre-RC, the six-subunit ORC specifically recognizes the ACS and the B1 element. Following loading of ORC, the other pre-RC factors and additional replication factors are recruited and origins are subsequently activated (Bell and Dutta, 2002). Although the protein factors appear to be highly conserved from yeasts to higher eukaryotes, the DNA sequences that define origin activity in different

organisms are not (Cvetic and Walter, 2005). Furthermore, in vitro studies in higher eukaryotes suggest that the metazoan ORC does not rely on sequence specificity to bind DNA (Remus et al., 2004; Vashee et al., 2003).

(12)

Figure 2. Classes of eukaryotic replication origins and composition of the pre-RC. (A) S. cerevisiae ARS1 is a well-defined replication origin. The 11bp ACS (and the B1 element) is specifically recognized by ORC, which together with Cdc6, Cdt1 and MCM2-7 constitute the pre-RC. The Origin of Bidirectional Replication (OBR) has been mapped immediately adjacent to the ORC-binding site. (B) The CHO DHFR locus contains a broad initiation zone, spanning the entire 55kb intergenic region between the DHFR and

2BE212 genes. Three sites show higher initiation activity. (C) The human lamin B2

replicon is markedly less complicated. The exact Transition Point (TP) from

discontinuous to continuous DNA synthesis has been determined by RIP mapping at the nucleotide resolution. See text for references. Blue pointed bars represent the coding frame of genes.

(13)
(14)

Physical mapping techniques have been developed to locate origins of DNA replication. Two-Dimensional (2D) gels separate replicating from nonreplicating molecules and allows the analysis of replication intermediates. This method is of relatively low resolution, revealing origins of replication as small as 2kb (DePamphilis, 1999). More sensitive approaches such as nascent strand analysis that employ PCR to determine the abundance of nascent strands improved the resolution of initiation sites to a few hundred basepairs (Giacca et al., 1994; Kobayashi et al., 1998a). Recently developed Replication Initiation Point (RIP) mapping has achieved nucleotide resolution by accurately defining the Transition Point (TP) from discontinuous to continuous DNA synthesis (Bielinsky and Gerbi, 1998; Gerbi and Bielinsky, 1997), and positioned the origin of bidirectional replication (OBR) immediately adjacent to the ORC-binding site in yeast (Figure 2A) (Bielinsky and Gerbi, 1998; Bielinsky and Gerbi, 1999).

In contrast to our knowledge of yeast origins, what constitutes a replication origin in higher eukaryotes is poorly understood (Bielinsky et al., 2001; DePamphilis et al., 2006; Gilbert, 2004). A handful of metazoan model replicons have been studied in detail in tissue culture cells (Cvetic and Walter, 2005; Gerbi, 2005). The fact that these systems lack convenient genetic assays has restricted metazoan origin studies to physical

biochemical mapping methods. It has been shown by 2D gels that replication of the Chinese hamster DHFR (Burhans et al., 1990; Vaughn et al., 1990) and human rDNA loci (Little et al., 1993; Yoon et al., 1995) initiates in broad regions. These data suggest the existence of large initiation zones (Gilbert, 2001), although in the DHFR locus two to three specific sites are preferred over other initiation sites spread throughout a 55kb

(15)

region (Figure 2B) (Dijkwel et al., 2002; Kobayashi et al., 1998a). On the other hand, studies of human lamin B2 (Figure 2C) and β-globin origins have identified much more defined sites of replication initiation, consistent with the classic replicon/replicator model (Gilbert, 2004; Jacob and Brenner, 1963). Thus, there seem to be two classes of

mammalian origins depending on the locus.

Origins of developmental gene amplification

As previously discussed, developmental gene amplification provides a powerful system for the study of metazoan DNA replication in vivo. In the development of the

Tetrahymena macronucleus, the 10.3kb rDNA locus is specifically excised from the

genome, converted to a ~21kb head-to-head palindrome, and telomeres are added to generate stable linear minichromosomes (Figure 3A). Then over the course of twelve hours, the rDNA minichromosomes are preferentially amplified up to 10,000-fold (Kapler et al., 1996; Prescott, 1994). Amplification initiates from two 430bp sites in

Tetrahymena thermophila (Figure 3A) (MacAlpine et al., 1997), or a single 900bp region

in T. pyriformis (Yue et al., 1998), all located within the 5’ Nontranscribed Spacer region (5’NTS) that is in the center of the palindrome. It appears that some of the amplified molecules separate from each other, as FISH studies demonstrate the presence of several hundred rDNA loci in nucleoli throughout the macronucleus (Ward et al., 1997). Given the small size (21 kb) of these extrachromosomal molecules, it is conceivable that at least some minichromosomes are fully replicated and some portion of the onionskin structures resolve. In contrast, amplification in Dipteran flies only represents a small portion of the

(16)

Figure 3. Origins of developmental gene amplification.

(A) Tetrahymena thermophila rDNA minichromosome. T represents the telomere at either end. Red ovals are nucleosome-free regions within the 5’ NTS (Nontranscribed Spacer) of the rRNA genes that show initiation activity. (B) Sciara coprophila salivary gland DNA puff II/9A. A 1kb origin (ORI) has been mapped upstream of II/9A genes by 2D and 3D gel analyses. ORI contains an ORC-binding site, immediately adjacent to the Transition Point (TP) from discontinuous to continuous DNA synthesis. (C) Drosophila

melanogaster DAFC-66D. The majority of origin activity resides in the intergenic oriβ. ACE3 is an amplification control element necessary for amplification, and it is

functionally separable from the Cp18 promoter. See text for references. Blue pointed bars represent the coding frame of genes.

(17)
(18)

giant polytene chromosomes in which the amplicons (each about 100kb in size) reside. Furthermore, suggested by FISH (Calvi et al., 1998) the onionskin structures remain held together without subsequent rearrangements.

In the Sciara salivary puff II/9A, 2D and 3D gel analyses indicate that initiation occurs over a 5.5kb region, within which a preferred 1kb portion (ORI) accounts for the majority of the origin firing (Figure 3B) (Liang and Gerbi, 1994; Liang et al., 1993). The precise nucleotide (Transition Point, TP) within the 1kb region at which DNA synthesis initiates has been determined by RIP mapping, and both recombinant ORC2 protein from Drosophila and endogenous Sciara ORC2 have been shown to bind to an 80bp segment adjacent to this initiation site (Bielinsky et al., 2001). In the related Sciarid fly,

Rhynchosciara, 2D gel analyses demonstrate that replication initiates in the salivary puff

C3 from at most 3 sites in a zone of about 6kb, and that this zone resides approximately 2kb upstream of the amplified gene C3-22 (Bielinsky et al., 2001).

In Drosophila, gene amplification of four genomic loci in the somatic follicle cells of the ovary occurs at specific stages of egg chamber development (stages 10 to 13, Figure 4A). During stages 9 and 10, the follicle cells surrounding the developing oocyte cease genomic DNA replication and begin to amplify four clusters of genes, which can be visualized as four foci by immunofluorescence following BrdU (bromodeoxyuridine, a thymidine analog) incorporation (Calvi et al., 1998). These Drosophila Amplicons in Follicle Cells (DAFCs) are named according to their cytological locations. Two of the amplicons are the X chromosome (at 7F, DAFC-7F) and the 3rd

chromosome (at 66D,

(19)

demand for the rapid synthesis of chorion proteins (Orr-Weaver, 1991). The other two loci of amplification, DAFC-30B and DAFC-62D, were recently identified by CGH array studies (Claycomb et al., 2004). These amplicons contain genes encoding a variety of proteins, including transporters, proteases, chitin-binding proteins and two putative enzymes Yellow-g and Yellow-g2, thought to be necessary for crosslinking proteins of the vitelline membrane or eggshell (Claycomb et al., 2004).

The third chromosome chorion amplicon DAFC-66D is the best studied (Figure 3C). 2D gel analysis has identified three potential replication origins within the peak amplified region, with one of these, oriβ, the 884bp sequence downstream of the Cp18 chorion protein gene being the preferred site of initiation that contains 70-80% of the origin activity (Delidakis and Kafatos, 1989; Heck and Spradling, 1990). Oriβ has ten out of eleven base pair matches to the Saccharomyces cerevisiae ARS consensus sequence that serves as an essential part of the origin of replication in yeast (Levine and Spradling, 1985). However, the significance of this sequence similarity is not clear, and notably, S. cerevisiae ARS1 origin sequences can not substitute for oriβ, thereby

confirming the sequence specificity of oriβ (Zhang and Tower, 2004). Deletion mapping of oriβ identified a 140 bp 5' element and a 226 bp A/T-rich 3' element called the β region that are necessary and sufficient to induce amplification of transposons (Zhang and Tower, 2004). The high A/T content of the β region might be important, because ORC has been shown to preferentially bind to A/T rich sequences in many species (for review see Bell, 2002). The replication initiation protein ORC2 binds

(20)

directly to oriβ during gene amplification (Austin et al., 1999). However, despite its tight association with oriβ in vivo, ORC does not preferentially bind DAFC amplification origin sequences in in vitro assays (Remus et al., 2004). This is similar to observations in other metazoan systems (Vashee et al., 2003), although contrasting Saccharomyces

cerevisiae in which specific sequences (the ARS Consensus Sequence or ACS) within the

origins are recognized (Newlon and Theis, 1993).

The developmental timing of amplification initiation appears to be highly

regulated and specific to particular amplicons. Real-time PCR suggested that DAFC-62D behaves distinctly from the other three amplicons in the timing of origin firing (Figure 4B) (Claycomb et al., 2004). In DAFC-66D, -7F and -30B, origin firing occurs only in stages 10 and 11. Afterwards there is no more increase in copy number at the central initiation sites. However, for DAFC-62D, an additional round of origin firing is observed in stage 13 within a defined 4kb region. Therefore, DAFC-62D provides a distinct model for studying not only the mechanisms of origin selection and activation, but also its developmental regulation. By identifying cis-regulatory elements in DAFCs that direct amplification as origin(s), and those that regulate amplification as enhancers by sensing differential developmental signals in different stages, we have gained important insights in understanding metazoan DNA replication in vivo.

(21)

Figure 4. Developmental timing of DAFC amplification.

(A) DAPI staining of egg chambers in stages 10 to 13. Follicle cells surround the developing oocytes. Adapted from A. C. Spradling, 1993. (B) Schematic drawing of the developmental timing of DAFC-66D and DAFC-62D amplification. About 30 fold of amplification in the center of DAFC-66D indicates 5 rounds of origin firing, all taking place in stages 10-11. By contrast, amplification of DAFC-62D is activated in two distinct stages, 10 and 13, separated by an elongation-only phase.

(22)
(23)

Amplification control elements

The relative ease of genetically manipulating Drosophila has greatly facilitated the mapping of cis-regulatory elements in DAFCs. In the P-element mediated

transformation systems, transposons that contain proper cis elements are able to amplify at ectopic genomic loci, although levels of amplification are dramatically affected by chromosomal position effects (de Cicco and Spradling, 1984; Orr-Weaver and Spradling, 1986). The introduction of insulators (Suppressor of Hairy-Wing Binding Sites

(SHWBS), Figure 5A) helps to remove position effects (Lu and Tower, 1997). The SHWBS insulators recruit proteins including Su(Hw) (Suppressor of Hairy Wing) that has been suggested to function as barriers between heterochromatin and open chromatin (Figure 5B) (Capelson and Corces, 2004; Gerasimova and Corces, 2001).

A number of studies have delineated in DAFC-66D the 320bp Amplification Control Element on the third chromosome, ACE3, required for high levels of

amplification (Figure 3C) (de Cicco and Spradling, 1984; Delidakis and Kafatos, 1989; Orr-Weaver et al., 1989; Orr-Weaver and Spradling, 1986). ACE3 is evolutionarily well conserved, located approximately 1.5kb upstream of oriβ and at the 5’ end of the s18 chorion gene. Demonstrated by 2D gel analyses (Delidakis and Kafatos, 1989; Heck and Spradling, 1990), ACE3 itself does not function as an origin of DNA replication, as it is not sufficient to support amplification in transposons protected by SHWBS insulators (Lu et al., 2001). Similarly, DAFC-7F contains an ACE element (ACE1) that is important for the amplification of this gene cluster (Spradling et al., 1987).

(24)

Figure 5. The SHWBS insulators remove position effects in P-element mediated transposon systems.

(A) Structure of transposon within the P-element sequences (black boxes). DNA sequence of interest (light-blue box) is flanked by SHWBS insulators. Arrows indicate orientation of SHWBS. mini-white (stippled box) is a reporter gene for the selection of transformation lines. (B) SHWBS binds Su(Hw) (Supressor of Hairy Wing) and additional proteins to form a barrier against surrounding heterochromatin (dark-blue boxes), so that the open chromatin structure (light-blue box) of the transposon is not affected by chromosomal position effects. See text for references.

(25)
(26)

It has been proposed that ACE serves as a developmental control element by stimulating replication initiation at nearby origins (Carminati et al., 1992; Delidakis and Kafatos, 1989; Heck and Spradling, 1990). A 142bp highly evolutionarily conserved “core” region of ACE3 has been determined responsible for the majority of ACE3’s replication stimulatory activity by deletion studies (Zhang and Tower, 2004).

Furthermore, ACE3 is necessary in cooperation with oriβ to achieve high levels of gene amplification, as a SHWBS insulator placed between ACE3 and oriβ in transposons nearly eliminates amplification. Removal of this insulator element by FLP/FRT-mediated recombination then restores amplification (Lu et al., 2001). Additionally, elimination of either ACE3 or oriβ from transposon constructs dramatically reduces amplification levels, indicating that together, ACE3 and oriβ are necessary and sufficient to drive developmental amplification (Lu et al., 2001).

Recent molecular studies provide some clues of how ACE3 might function as an amplification enhancer. ORC binds directly not only to oriβ but also to ACE3 and ACE1 in a site-specific manner, by either in vivo Chromatin Immunoprecipitation (ChIP) or in

vitro binding assays (Figure 6A) (Austin et al., 1999; Royzman et al., 1999). It has hence

been suggested that ACE3 serves as a nucleating site for ORC to spread along the

chromatin, thus influencing the ability of the region to replicate (Austin et al., 1999; Lu et al., 2001). By immunofluorescence, transposons of ACE3 multimers are capable of recruiting ORC2 in vivo (Austin et al., 1999), and support amplification presumably

(27)

initiated from proximal origins (Carminati et al., 1992). The amplification of a minimal transposon buffered by SHWBS containing only ACE3, Cp18, and oriβ is dependent on the orc2 gene product by mutant analysis, though without detection of ORC2 in

immunofluorescence (Lu et al., 2001). Therefore it appears that a certain threshold amount of ORC2 must be recruited. It may not always be detectable by staining methods, but in more sensitive assays such as ChIP ORC2 clearly associates with amplification origins and enhancers (Austin et al., 1999). Cumulatively, these data indicate that ACE3 and oriβ are functionally separable, but act cooperatively to drive gene amplification. The current working model is that ACE3 may nucleate ORC that then spreads along the chromatin to initiate replication at oriβ.

The involvement of transcription factors in gene amplification

Genetic, cytological, and biochemical approaches have also contributed to the understanding of the trans factors involved in developmental gene amplification. It has been clearly demonstrated that the proteins involved in DNA replication during a normal cell cycle are also involved in replication during gene amplification (Claycomb and Orr-Weaver, 2005; Tower, 2004). Hypomorphic mutations in Drosophila genes encoding essential components of the replication machinery lead to female sterility, disrupted eggshells, and severely decreased DAFC amplification, as measured by incorporation of BrdU or Quantitative Southern blotting (Henderson et al., 2000; Landis et al., 1997b; Landis and Tower, 1999; Tower, 2004; Underwood et al., 1990; Whittaker et al., 2000).

(28)

The archetypal DNA replication machinery includes first the formation of the pre-RC at the origins (Bell and Dutta, 2002). The ORC and DUP/Cdt1 proteins are sequentially recruited, which in turn load the putative replication fork helicase complex, MCM2-7 (Aparicio et al., 1997; Bell and Dutta, 2002; Labib et al., 2001).

In addition to conserved replication proteins, the components known to play a role in Drosophila chorion gene amplification include transcription factors. While chromatin immunoprecipitation (ChIP) experiments have shown that ORC binds directly to ACE3 and to oriβ (Austin et al., 1999; Bosco et al., 2001), additional ChIP and binding studies have localized transcription factors E2F1/DP/Rb to ACE3 during amplification stages in a complex containing ORC (Figure 6A) (Bosco et al., 2001). In the normal cell cycle, the E2F1/Rb complex acts as a transcriptional repressor until at the G1 to S phase transition phosphorylation of Rb converts E2F1 into a transcriptional activator for the expression of multiple genes required for S phase entry (Dyson, 1998; Zhu et al., 2005). E2F1 is required for chorion gene amplification because E2f1 mutants in which the DNA-binding domain is disrupted display decreased amplification and no ORC localization; a

hypomorphic Rb mutation or a mutation in E2f1 that removes the Rb binding site results in overamplification and inappropriate genomic replication (Royzman et al., 1999). These data support a model in which E2F1/Rb binds at and/or around ACE3 and represses replication until amplification stages, during which E2F1 positively regulates amplification initiation, with hyperphosphorylated Rb (pRb).

There are two E2f genes in Drosophila, E2f1 and E2f2 (Frolov et al., 2001). E2F1 is a potent activator of transcription, whereas E2F2 has been shown to repress

(29)

transcription (Dyson, 1998). In null E2f2 mutants BrdU incorporation occurs throughout the nucleus during DAFC amplification stages, failing to confine DNA synthesis to

DAFC sites (Cayirlioglu et al., 2001). In parallel, the distribution of pre-RC components

changes from being restricted to DAFC foci into being nuclear in these mutants, suggesting a repressive role of E2F2 in genomic DNA replication (Cayirlioglu et al., 2001). Although in E2f2 mutants there is a mild increase in pre-RC transcript level (Cayirlioglu et al., 2001; Cayirlioglu et al., 2003), it does not exclude the possibility that E2F2 functions directly at genomic origins to repress replication (see below).

Another transcription factor that associates with ACE3 is the Myb oncoprotein. A complex containing Myb, Mip120 (Myb Interacting Protein 120, formerly p120),

Mip130, Mip40, and Caf1 p55 interacts with ORC (Figure 6A) (Beall et al., 2002; Korenjak et al., 2004). Both the Myb and Mip120 subunits exhibit specific binding. Within ACE3, binding sites for Myb and Mip120 have been identified, and deletion of these sites from transposons nearly abolishes amplification compared to the non-deleted control (Beall et al., 2002). These results indicate that the Myb and at least one of the Mip120 binding sites are necessary for amplification. Myb is essential for viability, as

Myb mutants are lethal. Myb mutant follicle cell clones are defective in BrdU

incorporation at DAFCs, showing that Myb is necessary for gene amplification (Beall et al., 2002). The fact that by immunofluorescence ORC2 and DUP/Cdt1 are properly localized to DAFCs in Myb mutant clones indicates that Myb is required for initiation at a later step (Beall et al., 2002). Mip130 mutant females are sterile and have BrdU

(30)

Figure 6. Transcription factors involved in DAFC-66D amplification and origin specification.

(A) E2F1/DP/Rb and a complex containing Myb specifically associate with ACE3. The Rb and Myb proteins may be activated through phosphorylation. ORC is site-specifically recruited and spreads along the chromatin to initiate replication from oriβ. The

Ultraspiracle/ Ecdysone Receptor (USP/EcR) transcription factor also may regulate amplification via an interaction with ACE3. (B) During amplification stages, genomic replication is inhibited by the E2F2-containing dREAM complexes, which excludes ORC from inactive non-DAFC origins.

(31)
(32)

From these data, Mip130 appears to interact with the other Mips in a complex to repress genome-wide replication. At specific chromosomal loci Myb becomes activated in some way, perhaps by phosphorylation (Beall et al., 2004), to initiate replication or

amplification. Such a switch from repressive to active state might be important for Myb to specifically allow the initiation of amplification at the appropriate developmental time at amplification origins.

Strikingly, the Myb and Mip130 mutant phenotypes are very similar to those of

E2f1 and E2f2, respectively. In addition to genetic evidence, molecular and biochemical

studies strongly suggest the E2F and Myb proteins co-regulate replication. E2F1 and the Myb-containing complex, both localized to ACE3, may act coordinately to activate

DAFC-66D amplification (Figure 6A). Although there is no report of a physical

interaction between E2F1 and Myb, a complex containing E2F2, Myb and Mips has been purified from Drosophila embryo extracts (Korenjak et al., 2004; Lewis et al., 2004). These dREAM complexes (Drosophila Rb, E2F and Myb-interacting proteins) bind to repressed chromatin (Korenjak et al., 2004). Based on the mutant phenotypes of Myb and

E2f, it has been proposed that dREAM inhibits genome-wide replication (Korenjak et al.,

2004; Lewis et al., 2004), possibly by excluding pre-RC from genomic origins (Figure 6B); at DAFC-66D this repressive effect is reversed by E2F1 and activated Myb to achieve site-specific amplification (Figure 6A).

E2F and Myb appear to directly regulate amplification without affecting transcription of DAFC-66D genes. The transcription factor Ultraspiracle (USP) on the other hand, has been shown to bind to the promoter of the Cp18 chorion gene of

(33)

DAFC-66D (Shea et al., 1990). USP is a zinc finger protein differentially enriched in the follicle

cells, and it is a developmentally important member of the family of nuclear steroid hormone receptors (Oro et al., 1992; Shea et al., 1990). It heterodimerizes with another member of the nuclear receptor superfamily, ecdysteroid receptor protein (EcR), to function as a receptor for the steroid hormone ecdysone (Christianson et al., 1992; Yao et al., 1992). Ecdysone governs egg chamber development, and maternal EcR is required for normal oogenesis (Buszczak et al., 1999; Carney and Bender, 2000). EcR displays increased activity in follicle cells during amplification stages (Hackney et al., 2007). Dominant negative mutants of EcR (DNEcR) can dimerize with USP and bind DNA, but they do not activate target gene expression (Cherbas et al., 2003; Hackney et al., 2007). Interestingly, introduction of DNEcR into follicle cells not only reduces chorion gene expression, but also results in significantly decreased amplification, and the eggs laid by mutant mother display thin eggshells and shortened dorsal appendages (Hackney et al., 2007). Taken together, these results indicate that the USP/EcR heterodimer mediates ecdysone regulation of chorion gene amplification and transcription (Figure 6A). These two events may be separable from each other because ACE3, discrete from sequences controlling transcription (Orr-Weaver et al., 1989), harbors a good match to Ecdysone Response Element (Hackney et al., 2007).

In Sciara coprophila the amplification of salivary gland DNA puff II/9A maybe similarly influenced by ecdysone, the master regulator of insect development (Crouse, 1968; Foulk et al., 2006). Ecdysone induces transcription of the II/9A genes (Bienz-Tadmor et al., 1991; DiBartolomeis and Gerbi, 1989; Wu et al., 1993). In vitro

(34)

incubation of pre-amplification stage salivary glands with ecdysone induces premature amplification, and injection of ecdysone into pre-amplification stage larvae results in amplification in vivo (Foulk et al., 2006). A putative EcRE is found directly adjacent to the ORC-binding site in the II/9A origin (Bielinsky et al., 2001) and is efficiently bound by the Sciara EcR in in vitro binding assays (Foulk et al., 2006). The Sciara and

Drosophila results indicate that the ecdysone and EcR transcription factor control of developmental gene amplification may be conserved in insects.

Recently proteins TIF1-4 (Type I interacting Factor) that are necessary for rDNA amplification in Tetrahymena have been purified as complexes with differential DNA binding activities within the initiation zone (Mohammad et al., 2000). Notably TIF1 possesses limited homology to a transcription factor in plants, p24 (Mohammad et al., 2000). Together with data from Drosophila and Sciara, it is clear that developmental gene amplification is under complex control that involves a number of transcription factors, acting to modulate the use of origins for amplification. These factors may play repressive or active roles, depending on the developmental stage, the genomic locus and the chromatin context.

Chromatin context and amplification activity

It is not surprising that the replication and amplification machinery requires a favorable chromatin context to access DNA (Groth et al., 2007). Eukaryotic DNA is packaged into an organized, higher-order chromatin structure by histone proteins (Loden and van Steensel, 2005). Post-translational modifications of histones such as acetylation

(35)

of histone N-terminal lysine residues induces chromosomal changes, resulting in the loss of chromosomal repression to allow successful transcription of the underlying genes, as well as replication of DNA molecules (Fukuda et al., 2006). Recent studies in yeast have provided evidence that posttranslational chromatin modification can control the

efficiency and/or timing of chromosomal origin activity (Aparicio et al., 2004; Vogelauer et al., 2002).

Using the DAFC systems, independent groups have demonstrated that histones H3 and H4 at and around ACE3 are hyperacetylated during gene amplification (Aggarwal and Calvi, 2004; Hartl et al., 2007), and the lysine residues that are acetylated are

associated with replication and not transcription (Hartl et al., 2007). Furthermore, the acetylation of H3 and H4 is not the result of histone deposition after replication, as the hyperacetylation is confined to the origins of DAFC-66D and not associated with the replication forks (Hartl et al., 2007). Hyperacetylation of histone H4 leads to

redistribution of ORC2 from amplification foci to a genome-wide staining pattern; tethering histone acetyl transferase (HAT) increases amplification levels of a transposon with ACE3 and oriβ (Aggarwal and Calvi, 2004). Conversely, tethering of a histone deacetylase (HDAC) or a chromatin repressor to ACE3 reduces amplification (Aggarwal and Calvi, 2004). These observations suggest chromatin structure may have a definite role in amplification origin activity and that origin firing may be facilitated by a modification of the chromatin state.

Such modifications may be conducted through recruitment of histone-modifying enzymes and/or chromatin-remodeling proteins by transcription factors (Kohzaki and

(36)

Murakami, 2005). In X. laevis eggs, injected plasmid DNA undergoes site-specific initiation of replication in the presence of a transcription factor that is known to recruit a chromatin-remodeling complex (Danis et al., 2004). This effect does not require active transcription, but rather correlates with the acetylation level of histone H3 at the initiation sites (Danis et al., 2004). The E2F/Rb and Myb complexes are good candidates that may function at DAFC amplification origins to recruit HATs or HDACs to modulate the accessibility of the chromatin at the origin (Beall et al., 2002; Bosco et al., 2001). For example, Rb has been shown in a number of organisms to repress transcription by

remodeling chromatin structure through interaction with proteins involved in nucleosome remodeling, histone acetylation/deacetylation and methylation (Giacinti and Giordano, 2006). Finally, chromatin state and nucleosomal positioning may also play a role in gene amplification in Sciara and Tetrahymena (Clever and Ellgaard, 1970; Giacinti and

Giordano, 2006; MacAlpine et al., 1997; Mok et al., 2001).

A general link between replication and transcription

Transcription factors appear to function by several means at the amplification origins to modulate their activity. Both E2F1 and Myb have been shown to interact with ORC (Beall et al., 2002; Bosco et al., 2001). Possibly with some degree of redundancy, they recruit ORC through direct interaction. Transcription factors may also directly recruit proteins to modify chromatin structure to facilitate the assembly of the replication machinery. It is probably not a pure coincidence, however, that some common chromatin

(37)

features are shared by active transcription and replication, considering the ultimate goal of gene amplification is to augment transcript level.

There is mounting evidence for a general link between transcription and replication. Replication origins are frequently found upstream of transcription units (Mechali, 2001), and there are several examples in which they coincide with promoter sequences (Kohzaki and Murakami, 2005). In the human β-globin and c-myc replicons, transcription regulatory elements have been shown to be essential for replication

initiation (Aladjem et al., 1995; Liu et al., 2003). In addition to studies of specific replication sites, genome-wide mapping of replication origins in eukaryotes has been facilitated by recent advances in DNA microarray technology, and has begun to establish the spatial and temporal program of replication initiation (MacAlpine and Bell, 2005). Microarray analyses of genomic replication in Drosophila and human cells show a correlation between regions undergoing active transcription and early replication (Jeon et al., 2005; MacAlpine et al., 2004; Schubeler et al., 2002; Woodfine et al., 2004). A more extensive study of Drosophila chromosome 2L in Kc cells uncovered a frequent

colocalization of ORC and RNA polymerase II (RNAPII) binding sites, implying a connection between transcription and ORC localization (MacAlpine et al., 2004).

A number of studies report positive effects of transcription factors on DNA replication (Kohzaki and Murakami, 2005). The recruitment of transcription factors alters origin activity on episomal plasmids in both S. cerevisiae and X. laevis eggs (Cheng et al., 1992; Danis et al., 2004). Similarly, expression of a CREB-GAL4 fusion protein restores replication origin activity of the mutant c-myc locus where a GAL4p

(38)

binding cassette replaces all regulatory sequences of the c-myc gene (Ghosh et al., 2004). These results suggest that transcription factor binding can enhance replication origin activity. In Chinese hamster ovary (CHO) cells the dihydrofolate reductase (DHFR) gene locus contains a 55-kb zone of potential initiation sites of replication upstream of the gene (Burhans et al., 1990; Vaughn et al., 1990). In mutants with parts or all of the DHFR promoter deleted such that transcription is undetectable, initiation in the intergenic space is markedly suppressed (but not eliminated); restoration of transcription with either the wild-type Chinese hamster promoter or a Drosophila-based construct restores origin activity to the wild-type pattern (Saha et al., 2004).

However, 2D gel analysis of the promoterless DHFR mutants reveals that initiation occurrs at a low level not only in the intergenic region, but also in the body of the DHFR gene, which had never been observed in the wild-type locus (Saha et al., 2004). Thus transcription seems to suppress replication initiation in the body of the gene, and help define the boundaries of the downstream origin (Saha et al., 2004). In a mutant human c-myc locus with the c-myc promoter replaced by inducible GFP-encoding genes, replication initiation is repressed upon induction of transcription (Ghosh et al., 2004). When basal or induced transcription complexes is slowed by the presence of α-amanitin, origin activity depends on the orientation of the transcription unit (Ghosh et al., 2004). These data suggest that high levels of transcription or the persistence of transcription complexes can repress replication initiation. Taken together, the seemingly dual role of transcription on origin firing may be important to ensure high activity of intergenic

(39)

origins, while suppressing initiation within the gene body to avoid disruption of pre-RCs by the passage of the transcriptional machinery.

Another theme of transcriptional control of origin firing is the involvement of RNAPII. Transcription factors mediate the enhancer --> activator--> mediator -->

RNAPII --> promoter pathway to initiate mRNA transcription via RNAPII in virtually all eukaryotes (Kornberg, 2005). In Chinese hamster ovary cells it has been reported that inhibition of RNAPII transcription by α-amanitin results in deregulation of replication initiation at the DHFR locus (Kornberg, 2005). In Sciara salivary puff II/9A, although transcription of the II-9-1 gene does not begin until amplification is complete, the promoter of II-9-1 is occupied by RNAPII during amplification stages, and it is this presence that is thought to limit the right-hand boundary of the initiation zone during amplification (Sasaki et al., 2006). Furthermore, a direct physical interaction has been reported between RNAPII and MCM2-7 in yeast (Gauthier et al., 2002; Holland et al., 2002), raising the possibility that the transcriptional machinery serves to load the MCM complex to origins in some developmental contexts. Recently the hyperphosphorylated form of RNAPII implicated in transcriptional elongation has been shown to

co-immunoprecipitate with DNA polymerase ε (Rytkonen et al., 2006).

In addition to direct interactions with replication proteins, RNAPII has been shown to be required for histone displacement within the transcriptionally activated gene’s coding region preceding RNAPII (Brown and Kingston, 1997; Lee et al., 2004; Schwabish and Struhl, 2004; Zhao et al., 2005). In the human hsp70 gene, transcription activation leads to nucleosomal disassembly in the first 400 bp coding sequence in front

(40)

of RNAPII (Brown and Kingston, 1997). More recently, it has been demonstrated that histone density throughout the entire Saccharomyces cerevisiae GAL10 coding region is inversely correlated with RNAPII association and transcriptional activity, suggesting efficient eviction of core histones from the DNA by transcription (Schwabish and Struhl, 2004). Additionally, MCM2-7 associated DNA is more susceptible to nuclease digestion, indicating that these chromatin domains may be less tightly compacted, although the causal and consequence relation is not clear (Forsburg, 2004; Holthoff et al., 1998; Richter et al., 1998). Finally, the elongating form of RNAPII is in association with chromatin remodeling and histone-modifying factors (Sims et al., 2004). All together, these physical interactions between promoters, the transcriptional machinery, factors regulating chromatin structures, replication proteins and finally replication and

amplification origins suggest a complex picture of transcription and replication regulation in the chromatin context.

Summary of thesis

This thesis work investigated mechanisms that control the unique timing of

DAFC-62D origin activation using cytological, molecular and genetic methods. We first

defined the origin sequences in DAFC-62D by analyzing the amount of nascent

replicative DNA across this amplicon. Surprisingly the origin coincided with the coding region of a 62D gene. ORC2 localized to the origin, as well as two other sites that did not confer origin activity. Both ORC2 and MCM2-7 displayed differential association with these sequences, corresponding to the two rounds of amplification in two separate

(41)

developmental stages (10 and 13). All three elements, dispersed in a 7kb central amplified region, were required for either round of DAFC-62D amplification, because deleting any one completely abolished amplification in transposon experiments. Preceded by transcription of the 62D gene in stage 12, the late round of origin firing was ablated by the RNAPII inhibitor α-amanitin. This effect was absent from other amplicons and insulated transposons, and specific to the stage 13 amplification at DAFC-62D and transposons that did not have functional insulators. Finally, blocking RNAPII transcription compromised MCM2-7 recruitment as suggested by ChIP analysis.

Our studies of the regulation of DAFC-62D yield several unexpected findings. We find that the positioning of ORC and MCM2-7 can be affected by differentiation stage. Transcription via RNAPII in cis controls localization of replication factors and origin activation. The comparative analyses of DAFC-62D and -66D demonstrate that there are distinct mechanisms for differential regulation of amplification origins during Drosophila follicle cell development. Transposon experiments suggest their distinct behavior than the endogenous amplicon may be accounted for by the insulators’ unique properties.

Together our findings provide critical insights into how metazoan DNA replication is controlled in response to developmental cues.

(42)

REFERENCES

Aggarwal, B. D., and Calvi, B. R. (2004). Chromatin regulates origin activity in Drosophila follicle cells. Nature 430, 372-376.

Aladjem, M. I., Groudine, M., Brody, L. L., Dieken, E. S., Fournier, R. E., Wahl, G. M., and Epner, E. M. (1995). Participation of the human beta-globin locus control region in initiation of DNA replication. Science 270, 815-819.

Aparicio, J. G., Viggiani, C. J., Gibson, D. G., and Aparicio, O. M. (2004). The Rpd3-Sin3 histone deacetylase regulates replication timing and enables intra-S origin control in Saccharomyces cerevisiae. Mol Cell Biol 24, 4769-4780.

Aparicio, O. M., Weinstein, D. M., and Bell, S. P. (1997). Components and dynamics of DNA replication complexes in S. cerevisiae: redistribution of MCM proteins and Cdc45p during S phase. Cell 91, 59-69.

Austin, R. J., Orr-Weaver, T. L., and Bell, S. P. (1999). Drosophila ORC specifically binds to ACE3, an origin of DNA replication control element. Genes Dev 13, 2639-2649. Beall, E. L., Bell, M., Georlette, D., and Botchan, M. R. (2004). Dm-myb mutant lethality in Drosophila is dependent upon mip130: positive and negative regulation of DNA

replication. Genes Dev 18, 1667-1680.

Beall, E. L., Manak, J. R., Zhou, S., Bell, M., Lipsick, J. S., and Botchan, M. R. (2002). Role for a Drosophila Myb-containing protein complex in site-specific DNA replication. Nature 420, 833-837.

Bell, S. P. (2002). The origin recognition complex: from simple origins to complex functions. Genes Dev 16, 659-672.

Bell, S. P., and Dutta, A. (2002). DNA replication in eukaryotic cells. Annu Rev Biochem 71, 333-374.

Bielinsky, A. K., Blitzblau, H., Beall, E. L., Ezrokhi, M., Smith, H. S., Botchan, M. R., and Gerbi, S. A. (2001). Origin recognition complex binding to a metazoan replication origin. Curr Biol 11, 1427-1431.

Bielinsky, A. K., and Gerbi, S. A. (1998). Discrete start sites for DNA synthesis in the yeast ARS1 origin. Science 279, 95-98.

Bielinsky, A. K., and Gerbi, S. A. (1999). Chromosomal ARS1 has a single leading strand start site. Mol Cell 3, 477-486.

Bienz-Tadmor, B., Smith, H. S., and Gerbi, S. A. (1991). The promoter of DNA puff gene II/9-1 of Sciara coprophila is inducible by ecdysone in late prepupal salivary glands of Drosophila melanogaster. Cell Regul 2, 875-888.

Bosco, G., Du, W., and Orr-Weaver, T. L. (2001). DNA replication control through interaction of E2F-RB and the origin recognition complex. Nat Cell Biol 3, 289-295. Brown, D. D., and Dawid, I. B. (1968). Specific gene amplification in oocytes. Science

160, 272-280.

Brown, S. A., and Kingston, R. E. (1997). Disruption of downstream chromatin directed by a transcriptional activator. Genes Dev 11, 3116-3121.

(43)

Burhans, W. C., Vassilev, L. T., Caddle, M. S., Heintz, N. H., and DePamphilis, M. L. (1990). Identification of an origin of bidirectional DNA replication in mammalian chromosomes. Cell 62, 955-965.

Buszczak, M., Freeman, M. R., Carlson, J. R., Bender, M., Cooley, L., and Segraves, W. A. (1999). Ecdysone response genes govern egg chamber development during mid-oogenesis in Drosophila. Development 126, 4581-4589.

Calvi, B. R., Lilly, M. A., and Spradling, A. C. (1998). Cell cycle control of chorion gene amplification. Genes Dev 12, 734-744.

Capelson M., and Corces, V. G. (2004). Boundary elements and nuclear organization. Biol Cell 96, 617-629.

Carminati, J. L., Johnston, C. G., and Orr-Weaver, T. L. (1992). The Drosophila ACE3 chorion element autonomously induces amplification. Mol Cell Biol 12, 2444-2453. Carney, G. E., and Bender, M. (2000). The Drosophila ecdysone receptor (EcR) gene is required maternally for normal oogenesis. Genetics 154, 1203-1211.

Cayirlioglu, P., Bonnette, P. C., Dickson, M. R., and Duronio, R. J. (2001). Drosophila E2f2 promotes the conversion from genomic DNA replication to gene amplification in ovarian follicle cells. Development 128, 5085-5098.

Cayirlioglu, P., Ward, W. O., Silver Key, S. C., and Duronio, R. J. (2003).

Transcriptional repressor functions of Drosophila E2F1 and E2F2 cooperate to inhibit genomic DNA synthesis in ovarian follicle cells. Mol Cell Biol 23, 2123-2134. Cheng, L. Z., Workman, J. L., Kingston, R. E., and Kelly, T. J. (1992). Regulation of DNA replication in vitro by the transcriptional activation domain of GAL4-VP16. Proc Natl Acad Sci U S A 89, 589-593.

Cherbas, L., Hu, X., Zhimulev, I., Belyaeva, E., and Cherbas, P. (2003). EcR isoforms in Drosophila: testing tissue-specific requirements by targeted blockade and rescue.

Development 130, 271-284.

Christianson, A. M., King, D. L., Hatzivassiliou, E., Casas, J. E., Hallenbeck, P. L., Nikodem, V. M., Mitsialis, S. A., and Kafatos, F. C. (1992). DNA binding and heteromerization of the Drosophila transcription factor chorion factor 1/ultraspiracle. Proc Natl Acad Sci U S A 89, 11503-11507.

Claycomb, J. M., Benasutti, M., Bosco, G., Fenger, D. D., and Orr-Weaver, T. L. (2004). Gene amplification as a developmental strategy: isolation of two developmental

amplicons in Drosophila. Dev Cell 6, 145-155.

Claycomb, J. M., MacAlpine, D. M., Evans, J. G., Bell, S. P., and Orr-Weaver, T. L. (2002). Visualization of replication initiation and elongation in Drosophila. J Cell Biol

159, 225-236.

Claycomb, J. M., and Orr-Weaver, T. L. (2005). Developmental gene amplification: insights into DNA replication and gene expression. Trends Genet 21, 149-162. Clever, U., and Ellgaard, E. G. (1970). Puffing and histone acetylation in polytene chromosomes. Science 169, 373-374.

Crouse, H. V. (1968). The role of ecdysone in DNA-puff formation and DNA synthesis in the polytene chromosomes of Sciara coprophila. Proc Natl Acad Sci U S A 61, 971-978.

(44)

Cvetic, C., and Walter, J. C. (2005). Eukaryotic origins of DNA replication: could you please be more specific? Semin Cell Dev Biol 16, 343-353.

Danis, E., Brodolin, K., Menut, S., Maiorano, D., Girard-Reydet, C., and Mechali, M. (2004). Specification of a DNA replication origin by a transcription complex. Nat Cell Biol 6, 721-730.

de Cicco, D. V., and Spradling, A. C. (1984). Localization of a cis-acting element

responsible for the developmentally regulated amplification of Drosophila chorion genes. Cell 38, 45-54.

Delidakis, C., and Kafatos, F. C. (1989). Amplification enhancers and replication origins in the autosomal chorion gene cluster of Drosophila. Embo J 8, 891-901.

DePamphilis, M. L. (1999). Replication origins in metazoan chromosomes: fact or fiction? Bioessays 21, 5-16.

DePamphilis, M. L., Blow, J. J., Ghosh, S., Saha, T., Noguchi, K., and Vassilev, A. (2006). Regulating the licensing of DNA replication origins in metazoa. Curr Opin Cell Biol 18, 231-239.

DiBartolomeis, S. M., and Gerbi, S. A. (1989). Molecular characterization of DNA puff II/9A genes in Sciara coprophila. J Mol Biol 210, 531-540.

Dijkwel, P. A., Wang, S., and Hamlin, J. L. (2002). Initiation sites are distributed at frequent intervals in the Chinese hamster dihydrofolate reductase origin of replication but are used with very different efficiencies. Mol Cell Biol 22, 3053-3065.

Dyson, N. (1998). The regulation of E2F by pRB-family proteins. Genes Dev 12, 2245-2262.

Forsburg, S. L. (2004). Eukaryotic MCM proteins: beyond replication initiation. Microbiol Mol Biol Rev 68, 109-131.

Foulk, M. S., Liang, C., Wu, N., Blitzblau, H. G., Smith, H., Alam, D., Batra, M., and Gerbi, S. A. (2006). Ecdysone induces transcription and amplification in Sciara coprophila DNA puff II/9A. Dev Biol 299, 151-163.

Frolov, M. V., Huen, D. S., Stevaux, O., Dimova, D., Balczarek-Strang, K., Elsdon, M., and Dyson, N. J. (2001). Functional antagonism between E2F family members. Genes Dev 15, 2146-2160.

Fukuda, H., Sano, N., Muto, S., and Horikoshi, M. (2006). Simple histone acetylation plays a complex role in the regulation of gene expression. Brief Funct Genomic Proteomic 5, 190-208.

Gall, J. G. (1968). Differential synthesis of the genes for ribosomal RNA during amphibian oogenesis. Proc Natl Acad Sci USA 60, 553-560.

Gall, J. G. (1974). Free ribosomal RNA genes in the macronucleus of Tetrahymena. Proc Natl Acad Sci U S A 71, 3078-3081.

Gauthier, L., Dziak, R., Kramer, D. J., Leishman, D., Song, X., Ho, J., Radovic, M., Bentley, D., and Yankulov, K. (2002). The role of the carboxyterminal domain of RNA polymerase II in regulating origins of DNA replication in Saccharomyces cerevisiae. Genetics 162, 1117-1129.

Gerasimova, T. I., and Corces, V. G. (2001). Chromatin insulators and boundaries: effects on transcription and nuclear organization. Annu Rev Genet 35, 193-208.

(45)

Gerbi, S. A. (2005). Mapping origins of DNA replication in eukaryotes. Methods Mol Biol 296, 167-180.

Gerbi, S. A., and Bielinsky, A. K. (1997). Replication initiation point mapping. Methods

13, 271-280.

Ghosh, M., Liu, G., Randall, G., Bevington, J., and Leffak, M. (2004). Transcription factor binding and induced transcription alter chromosomal c-myc replicator activity. Mol Cell Biol 24, 10193-10207.

Giacca, M., Zentilin, L., Norio, P., Diviacco, S., Dimitrova, D., Contreas, G., Biamonti, G., Perini, G., Weighardt, F., Riva, S., and et al. (1994). Fine mapping of a replication origin of human DNA. Proc Natl Acad Sci U S A 91, 7119-7123.

Giacinti, C., and Giordano, A. (2006). RB and cell cycle progression. Oncogene 25, 5220-5227.

Gilbert, D. M. (2001). Making sense of eukaryotic DNA replication origins. Science 294, 96-100.

Gilbert, D. M. (2004). In search of the holy replicator. Nat Rev Mol Cell Biol 5, 848-855. Glover, D. M., Zaha, A., Stocker, A. J., Santelli, R. V., Pueyo, M. T., De Toledo, S. M., and Lara, F. J. S. (1982). Gene amplification in Rhynchosciara salivary gland

chromosomes. Proc Natl Acad Sci USA 79, 2947-2951.

Groth, A., Rocha, W., Verreault, A., and Almouzni, G. (2007). Chromatin challenges during DNA replication and repair. Cell 128, 721-733.

Hackney, J. F., Pucci, C., Naes, E., and Dobens, L. (2007). Ras signaling modulates activity of the ecdysone receptor EcR during cell migration in the Drosophila ovary. Dev Dyn 236, 1213-1226.

Hartl, T., Boswell, C., Orr-Weaver, T. L., and Bosco, G. (2007). Developmentally regulated histone modifications in Drosophila follicle cells: initiation of gene amplification is associated with histone H3 and H4 hyperacetylation and H1 phosphorylation. Chromosoma.

Heck, M. M., and Spradling, A. C. (1990). Multiple replication origins are used during Drosophila chorion gene amplification. J Cell Biol 110, 903-914.

Henderson, D. S., Wiegand, U. K., Norman, D. G., and Glover, D. M. (2000). Mutual correction of faulty PCNA subunits in temperature-sensitive lethal mus209 mutants of

Drosophila melanogaster. Genetics 154, 1721-1733.

Holland, L., Gauthier, L., Bell-Rogers, P., and Yankulov, K. (2002). Distinct parts of minichromosome maintenance protein 2 associate with histone H3/H4 and RNA polymerase II holoenzyme. Eur J Biochem 269, 5192-5202.

Holthoff, H. P., Baack, M., Richter, A., Ritzi, M., and Knippers, R. (1998). Human protein MCM6 on HeLa cell chromatin. J Biol Chem 273, 7320-7325.

Hourcade, D., Dressler, D., and Wolfson, J. (1973). The amplification of ribosomal RNA genes involves a rolling circle intermediate. Proc Natl Acad Sci U S A 70, 2926-2930. Jacob, F., and Brenner, S. (1963). [On the regulation of DNA synthesis in bacteria: the hypothesis of the replicon.]. C R Hebd Seances Acad Sci 256, 298-300.

(46)

Jeon, Y., Bekiranov, S., Karnani, N., Kapranov, P., Ghosh, S., MacAlpine, D., Lee, C., Hwang, D. S., Gingeras, T. R., and Dutta, A. (2005). Temporal profile of replication of human chromosomes. Proc Natl Acad Sci U S A 102, 6419-6424.

Kapler, G. M., Dobbs, D. L., and Blackburn, E. (1996). DNA replication in Tetrahymena. In DNA replication in eukaryotic cells, M. L. DePamphilis, ed. (Cold Spring Harbor, N.Y., Cold Spring Harbor Laboratory Press), pp. 915-932.

Kobayashi, T., Rein, T., and DePamphilis, M. L. (1998). Identification of primary initiation sites for DNA replication in the hamster dihydrofolate reductase gene initiation zone. Mol Cell Biol 18, 3266-3277.

Kohzaki, H., and Murakami, Y. (2005). Transcription factors and DNA replication origin selection. Bioessays 27, 1107-1116.

Korenjak, M., Taylor-Harding, B., Binne, U. K., Satterlee, J. S., Stevaux, O., Aasland, R., White-Cooper, H., Dyson, N., and Brehm, A. (2004). Native E2F/RBF complexes

contain Myb-interacting proteins and repress transcription of developmentally controlled E2F target genes. Cell 119, 181-193.

Kornberg, R. D. (2005). Mediator and the mechanism of transcriptional activation. Trends Biochem Sci 30, 235-239.

Labib, K., Kearsey, S. E., and Diffley, J. F. (2001). MCM2-7 proteins are essential components of prereplicative complexes that accumulate cooperatively in the nucleus during G1-phase and are required to establish, but not maintain, the S-phase checkpoint. Mol Biol Cell 12, 3658-3667.

Laicine, E. M., Alves, M. A., de Almeida, J. C., Rizzo, E., Albernaz, W. C., and Sauaia, H. (1984). Development of DNA puffs and patterns of polypeptide synthesis in the salivary glands of Bradysia hygida. Chromosoma 89, 280-284.

Landis, G., Kelley, R., Spradling, A. C., and Tower, J. (1997). The k43 gene, required for chorion gene amplification and diploid cell chromosome replication, encodes the

Drosophila homolog of yeast origin recognition complex subunit 2. Proc Natl Acad Sci

USA 94, 3888-3892.

Landis, G., and Tower, J. (1999). The drosophila chiffon gene is required for chorion gene amplification, and is related to the yeast dbf4 regulator of DNA replication and cell cycle. Development 126, 4281-4293.

Lee, C. K., Shibata, Y., Rao, B., Strahl, B. D., and Lieb, J. D. (2004). Evidence for nucleosome depletion at active regulatory regions genome-wide. Nat Genet 36, 900-905. Levine, J., and Spradling, A. (1985). DNA sequence of a 3.8 kilobase pair region

controlling Drosophila chorion gene amplification. Chromosoma 92, 136-142.

Lewis, P. W., Beall, E. L., Fleischer, T. C., Georlette, D., Link, A. J., and Botchan, M. R. (2004). Identification of a Drosophila Myb-E2F2/RBF transcriptional repressor complex. Genes Dev 18, 2929-2940.

Liang, C., and Gerbi, S. A. (1994). Analysis of an origin of DNA amplification in Sciara coprophila by a novel three-dimensional gel method. Mol Cell Biol 14, 1520-1529. Liang, C., Spitzer, J. D., Smith, H. S., and Gerbi, S. A. (1993). Replication initiates at a confined region during DNA amplification in Sciara DNA puff II/9A. Genes and Dev 7, 1072-1084.

(47)

Little, R. D., Platt, T. H., and Schildkraut, C. L. (1993). Initiation and termination of DNA replication in human rRNA genes. Mol Cell Biol 13, 6600-6613.

Liu, G., Malott, M., and Leffak, M. (2003). Multiple functional elements comprise a Mammalian chromosomal replicator. Mol Cell Biol 23, 1832-1842.

Loden, M., and van Steensel, B. (2005). Whole-genome views of chromatin structure. Chromosome Res 13, 289-298.

Lu, L., and Tower, J. (1997). A transcriptional insulator element, the su(Hw) binding site, protects a chromosomal DNA replication origin from position effects. Mol Cell Biol 17, 2202-2206.

Lu, L., Zhang, H., and Tower, J. (2001). Functionally distinct, sequence-specific replicator and origin elements are required for Drosophila chorion gene amplification. Genes Dev 15, 134-146.

MacAlpine, D. M., Rodriguez, H. K., and Bell, S. P. (2004). Coordination of replication and transcription along a Drosophila chromosome. Genes Dev 18, 3094-3105.

MacAlpine, D. M., Zhang, Z., and Kapler, G. M. (1997). Type I elements mediate replication fork pausing at conserved upstream sites in the Tetrahymena thermophila ribosomal DNA minichromosome. Mol Cell Biol 17, 4517-4525.

Mechali, M. (2001). DNA replication origins: from sequence specificity to epigenetics. Nat Rev Genet 2, 640-645.

Mohammad, M., Saha, S., and Kapler, G. M. (2000). Three different proteins recognize a multifunctional determinant that controls replication initiation, fork arrest and

transcription in Tetrahymena. Nucleic Acids Res 28, 843-851.

Mok, E. H., Smith, H. S., DiBartolomeis, S. M., Kerrebrock, A. W., Rothschild, L. J., Lange, T. S., and Gerbi, S. A. (2001). Maintenance of the DNA puff expanded state is independent of active replication and transcription. Chromosoma 110, 186-196.

Newlon, C. S., and Theis, J. F. (1993). The structure and function of yeast ARS elements. Curr Opin Genet Dev 3, 752-758.

Oro, A. E., McKeown, M., and Evans, R. M. (1992). The Drosophila retinoid X receptor homolog ultraspiracle functions in both female reproduction and eye morphogenesis. Development 115, 449-462.

Orr-Weaver, T. L. (1991). Drosophila chorion genes: cracking the eggshell's secrets. Bioessays 13, 97-105.

Orr-Weaver, T. L., Johnston, C. G., and Spradling, A. C. (1989). The role of ACE3 in Drosophila chorion gene amplification. Embo J 8, 4153-4162.

Orr-Weaver, T. L., and Spradling, A. C. (1986). Drosophila chorion gene amplification requires an upstream region regulating s18 transcription. Mol Cell Biol 6, 4624-4633. Prescott, D. M. (1994). The DNA of ciliated protozoa. Microbiol Rev 58, 233-267. Remus, D., Beall, E. L., and Botchan, M. R. (2004). DNA topology, not DNA sequence, is a critical determinant for Drosophila ORC-DNA binding. Embo J 23, 897-907.

Richter, A., Baack, M., Holthoff, H. P., Ritzi, M., and Knippers, R. (1998). Mobilization of chromatin-bound Mcm proteins by micrococcal nuclease. Biol Chem 379, 1181-1187.

(48)

Royzman, I., Austin, R. J., Bosco, G., Bell, S. P., and Orr-Weaver, T. L. (1999). ORC localization in Drosophila follicle cells and the effects of mutations in dE2F and dDP. Genes Dev 13, 827-840.

Rytkonen, A. K., Hillukkala, T., Vaara, M., Sokka, M., Jokela, M., Sormunen, R., Nasheuer, H. P., Nethanel, T., Kaufmann, G., Pospiech, H., and Syvaoja, J. E. (2006). DNA polymerase epsilon associates with the elongating form of RNA polymerase II and nascent transcripts. Febs J 273, 5535-5549.

Saha, S., Shan, Y., Mesner, L. D., and Hamlin, J. L. (2004). The promoter of the Chinese hamster ovary dihydrofolate reductase gene regulates the activity of the local origin and helps define its boundaries. Genes Dev 18, 397-410.

Sasaki, T., Ramanathan, S., Okuno, Y., Kumagai, C., Shaikh, S. S., and Gilbert, D. M. (2006). The Chinese hamster dihydrofolate reductase replication origin decision point follows activation of transcription and suppresses initiation of replication within transcription units. Mol Cell Biol 26, 1051-1062.

Schubeler, D., Scalzo, D., Kooperberg, C., van Steensel, B., Delrow, J., and Groudine, M. (2002). Genome-wide DNA replication profile for Drosophila melanogaster: a link

between transcription and replication timing. Nat Genet 32, 438-442.

Schwabish, M. A., and Struhl, K. (2004). Evidence for eviction and rapid deposition of histones upon transcriptional elongation by RNA polymerase II. Mol Cell Biol 24, 10111-10117.

Shea, M. J., King, D. L., Conboy, M. J., Mariani, B. D., and Kafatos, F. C. (1990). Proteins that bind to Drosophila chorion cis-regulatory elements: a new C2H2 zinc finger protein and a C2C2 steroid receptor-like component. Genes Dev 4, 1128-1140.

Sims, R. J., 3rd, Belotserkovskaya, R., and Reinberg, D. (2004). Elongation by RNA polymerase II: the short and long of it. Genes Dev 18, 2437-2468.

Spradling, A. C. (1981). The organization and amplification of two clusters of Drosophila chorion genes. Cell 27, 193-201.

Spradling, A. C., de Cicco, D. V., Wakimoto, B. T., Levine, J. F., Kalfayan, L. J., and Cooley, L. (1987). Amplification of the X-linked Drosophila chorion gene cluster requires a region upstream from the s38 chorion gene. Embo J 6, 1045-1053.

Spradling, A. C., Digan, M. E., Mahowald, A. P., Scott, M., and Craig, E. A. (1980). Two clusters of genes for major chorion proteins of Drosophila melanogaster. Cell 19, 905-914.

Tower, J. (2004). Developmental gene amplification and origin regulation. Annu Rev Genet 38, 273-304.

Underwood, E. M., Briot, A. S., Doll, K. Z., Ludwiczak, R. L., Otteson, D. C., Tower, J., Vessey, K. B., and Yu, K. (1990). Genetics of 51D-52A, a region containing several maternal-effect genes and two maternal-specific transcripts in Drosophila. Genetics 126, 639-650.

Vashee, S., Cvetic, C., Lu, W., Simancek, P., Kelly, T. J., and Walter, J. C. (2003). Sequence-independent DNA binding and replication initiation by the human origin recognition complex. Genes Dev 17, 1894-1908.

Références

Documents relatifs

With a growing number of measured values the increase of the probability that the next data point is in the correct class is compared to the results from the special case program

(E) Left panel: western blot analysis of the indicated proteins in total protein extracts obtained from tumors derived from SC35-5 cells and treated (Dox +) or not (Dox-) with 2

Although the mechanical properties of single cells have been associated with cell phenotype and malignancy, it has been difficult to directly correlate each cell’s

Consigne: « Par groupes de 4 ou 5 vous allez devoir schématiser le trajet du sang dans le corps en tenant compte des connaissances acquises lors des dernières

The employed strategy to transport the beam from the exit of the cyclotron downstream to the experimental hall is dependent on the end-station of the line in use and specific

Finally, we analyze different spreading policies of the Tetrys redundancy traffic between available paths and observe that the longer propagation delay path should be preferably used

L’accès à ce site Web et l’utilisation de son contenu sont assujettis aux conditions présentées dans le site LISEZ CES CONDITIONS ATTENTIVEMENT AVANT D’UTILISER CE SITE WEB.

F.3 – Pr´ ecision et nombre de points appari´ es pour des chan- gements d’´ echelle sur les images Leuven.... F.4 – Pr´ ecision et nombre de points appari´ es pour des