Abstract
Nucleosomes can restrict the access of transcription factors to chromatin. RSC is a SWI/SNF-family chromatin-remodeling complex from yeast that repositions and ejects nucleosomes in vitro. Here, we examined these activities and their importance in vivo. We utilized array-based methods to examine nucleosome occupancy and positioning at more than 200 locations in the genome following the controlled destruction of the catalytic subunit of RSC, Sth1. Loss of RSC function caused pronounced and general reductions in new transcription from Pol I, II, and III genes. At Pol III genes, Sth1 loss conferred a general reduction in RNA Pol III occupancy and a gain in nucleosome density. Notably at the one Pol III gene examined, histone restoration was partly replication-dependent. In contrast, at Pol II promoters we observed primarily single nucleosome changes, including movement. Importantly, alterations near the transcription start site were more common at RSC-occupied promoters than at non-occupied promoters. Thus, RSC action affects both nucleosome density and positioning in vivo, but applies these remodeling modes differently at Pol II and Pol III genes.
Keywords: chromatin, nucleosome, RSC, transcription
Introduction
The packaging of eukaryotic genomes into chromatin provides remarkable compaction and organization within the nucleus. Nucleosomes comprise the basic repeating units of chromatin and consist of an octamer of four histones. Periodically the DNA must be made accessible to multiple types of factors, including DNA-binding factors, polymerases, and repair enzymes. To accomplish this, cells remodel chromatin to allow DNA accessibility through a variety of enzymatic processes (reviewed in Groth et al, 2007; Li et al, 2007). While chromatin remodeling has been characterized in vitro and in vivo at specific loci, a large-scale analysis of remodeling in vivo is needed to understand how these processes are applied and regulated during, for example, transcription.
Typical RNA polymerase II (Pol II) promoters have a chromatin structure that includes a small (∼150 bp) nucleosome-free region (NFR) located ∼200 bp upstream of the coding region (Yuan et al, 2005). The NFR often coincides with the transcriptional start site of transcription, suggesting it is the site of assembly for the Pol II initiation complex. Adjacent to either side of the NFR are well-positioned nucleosomes that often contain inherent nucleosome-positioning sequences (NPS; Ioshikhes et al, 2006; Segal et al, 2006). Furthermore, TATA-containing promoters are more likely to have the TATA box either within, or adjacent to, a strong NPS. Not surprisingly, these promoters are typically regulated by histone acetyltransferases, histone deacetylases, Mediator complex, and chromatin remodelers including SWI/SNF and ISWI, which may collaborate to move proximal nucleosomes. Access to a TATA box and other promoter elements that are occluded by nucleosomes may be accomplished either by nucleosome sliding (Lomvardas and Thanos, 2001; Inai et al, 2007) or by ejection (Adkins et al, 2004; Boeger et al, 2004). Hence, the emergent view of active transcription involves the recruitment of multiple chromatin-modifying complexes to post-translationally modify, move, and/or eject nucleosomes to help establish or expand the NFR and allow the pre-initiation complex to be assembled.
Chromatin remodelers are evolutionarily conserved, multi-subunit complexes that utilize ATP hydrolysis to alter the composition or arrangement of chromatin. Nucleosome remodeling is based on one of two simple strategies: disassembling (ejecting) or sliding the nucleosome (Langst et al, 1999; Lorch et al, 1999; Whitehouse et al, 1999; Boeger et al, 2004). Remodelers are comprised of subunits with specialized roles. The central subunit contains the ATPase engine that provides a DNA translocation force (Saha et al, 2002; Whitehouse et al, 2003). Additional subunits may regulate the ATPase activity (Sif et al, 1998; Yang et al, 2007). Targeting of the remodeler can be accomplished through several means, including DNA binding (Angus-Hill et al, 2001), direct protein–protein interactions with activators (Yudkovsky et al, 1999; Govind et al, 2005; Inai et al, 2007), recruitment by basal transcription machinery (Soutourina et al, 2006), or by recognition of acetylated histones through bromodomains (Hassan et al, 2002; Geng and Laurent, 2004; Kasten et al, 2004).
Remodelers regulate a number of chromatin- and DNA-based processes, including transcriptional activation and repression, transcriptional elongation, replication, and DNA repair (Saha et al, 2006). One well-characterized family of remodelers is the SWI/SNF family, of which multiple homologues exist in vertebrates. The RSC (Remodels Structure of Chromatin) complex is an abundant, essential yeast paralogue of the SWI/SNF complex (Cairns et al, 1996). RSC is comprised of 15 subunits, including several bromodomain-containing subunits. RSC is involved in numerous cell processes throughout the cell cycle, including DNA repair, chromosome segregation, and DNA ploidy maintenance (Hsu et al, 2003; Huang et al, 2004; Chai et al, 2005; Campsteijn et al, 2007; Shim et al, 2007). However, the primary role appears to be transcriptional activation (Cairns et al, 1999; Angus-Hill et al, 2001; Govind et al, 2005). Genome-wide localization of RSC places it at ∼700 targets, nearly all of which are Pol II promoter regions or Pol III genes (Damelin et al, 2002; Ng et al, 2002). Thus, the prevailing hypothesis is that RSC alters the chromatin structure of promoters to assist in transcription.
RSC is recruited to promoters by transcription factors and other transcriptional coactivators to enable chromatin remodeling and Pol II promoter transcription (Govind et al, 2005; Inai et al, 2007). Additionally, recent experiments at Pol III genes show an attenuation of Pol III transcription in RSC mutants (Soutourina et al, 2006). Furthermore, alterations in chromatin structure have been reported at a few individual loci in RSC mutants (Tsuchiya et al, 1998; Moreira and Holmberg, 1999; Soutourina et al, 2006; Inai et al, 2007). However, as RSC can both eject and reposition nucleosomes, a large-scale analysis is required to determine the extent that RSC uses these different modes at different classes of targets. Here, we generated a strong conditional mutant that abolishes RSC function and examined its impact on transcription and promoter chromatin structure. We utilized a custom RSC-specific microarray to explore RSC function at a genomic scale. Surprisingly, loss of RSC had a profound impact on genome-wide transcription. Interestingly, at RSC targets chromatin structure was altered in a class-dependent manner in the absence of RSC function: Pol II promoters primarily displayed single nucleosome events with alterations near the transcription start site (TSS), whereas Pol III genes primarily displayed gross nucleosome gain.
Results
Loss of Sth1 protein results in loss of transcription
To study the impact of RSC on transcription and chromatin structure, we generated an STH1 allele (sth1td) tagged with an N-terminal temperature-sensitive degron to enable rapid and inducible loss of the Sth1 protein (Dohmen et al, 1994). We reasoned that the loss of the catalytic subunit of RSC should inactivate enzymatic activity. For our studies, we employed two isogenic yeast strains carrying the sth1td allele that differed only at UBR1, which encodes the protein that recognizes the degron tag. One strain contained an inducible UBR1 gene (degron strain) while the second was ubr1Δ (control strain; Labib et al, 2000). The degron strain failed to grow under degradation-inducing conditions (galactose media at 37°C), while full growth was observed for the control strain (Figure 1A). Furthermore, western blot analysis of cell lysates showed substantial loss of Sth1 protein selectively in the degron strain (Figure 1B). While the loss of Sth1 protein reduced cell growth, the effect was reversible; cells subjected to loss of Sth1 protein for 2 h and allowed to recover at room temperature showed >95% viability (data not shown). Similar results with another sth1td allele were also recently reported (Campsteijn et al, 2007). Together, these results show the utility of the sth1td allele in analyzing RSC function.
Figure 1.
Sth1 loss results in inviability and general loss of transcription. (A) Control (sth1td ubr1Δ) and degron (sth1td Gal-UBR1) strains were serially diluted and spotted on media containing either raffinose (Raf) or galactose (Gal) and grown at either 24 or 37°C. (B) Whole-cell lysates from wild-type (WT), degron, or control strains were analyzed by western blot and probed with anti-Sth1. Lanes contain extracts from cells representing before (Bef.) or after (Aft.) Sth1 loss. (C) Degron and control strain cells either before or after Sth1 degradation were incubated with 3H-labeled uracil for 25 min before harvesting. RNA samples were separated by size on a denaturing polyacrylamide gel. Half of the gel was stained with ethidium bromide (EtBr) while the remaining half was exposed to film (3H).
To determine the impact of the loss of RSC activity on new transcription, cells were grown in the presence of 3H-labeled uracil for 25 min either before or after Sth1 protein degradation. RNA was isolated and separated by PAGE. Ethidium bromide staining of the gel revealed uniform RNA levels, while autoradiograph exposure revealed the striking loss of new RNA transcription in the degron strain after Sth1 protein loss (Figure 1C). Importantly, transcription by all three RNA polymerases was affected, as tRNAs, rRNAs, and mRNAs were greatly diminished. Thus, the loss of Sth1 protein had a global impact on transcription. Whether these effects were direct or indirect presented a considerable challenge, which we addressed by multiple approaches.
Histone density increases at a Pol III locus with loss of Sth1 protein
RSC may directly impact transcription through alterations of the chromatin structure at RSC targets. If so, then direct targets might share a common chromatin feature. Notably, the largest class of RSC-occupied loci is Pol III genes, primarily tRNA genes (Ng et al, 2002). Furthermore, tRNA genes have the lowest histone occupancy genome-wide (Supplementary Figure 1; Rao et al, 2005). This inverse correlation between histone and RSC occupancy at tRNA genes raised the possibility that RSC helps to keep Pol III genes nucleosome sparse, and that the loss of RSC activity might result in a gain of histone occupancy over time.
To initially investigate RSC function at a Pol III locus, we examined SCR1, the longest Pol III gene in the genome and a known RSC target. SCR1 was previously used as a model Pol III gene to map the distribution of Pol III components over the body of the gene (Roberts et al, 2006). Histone chromatin immunoprecipitation (ChIP) was performed in the degron and control strains both before and after Sth1 protein loss. The precipitated DNA was then analyzed by quantitative PCR using six amplicons distributed across the locus (Figure 2A and B). Histone occupancy was low over the body of the SCR1 gene in both strains before Sth1 loss, consistent with our observations at tRNA genes genome-wide. After 1 h of Sth1 loss, however, we observed a small gain of histone density and a more robust gain after 2 h over the body of SCR1 in the degron strain, but not in the control strain. As an alternative measure, the chromatin structure of SCR1 was probed using micrococcal nuclease (MNase) digestion and analyzed by Southern blot. The body of the SCR1 gene yielded a detectable nucleosomal ladder of hybridization in the degron strain but not the control strain (Figure 2C). Together, these results indicated that Sth1 protein loss results in a progressive gain of nucleosomes at the SCR1 locus.
Figure 2.
Changes in chromatin structure are detected at the Pol III gene SCR1 upon loss of Sth1 protein. (A) The positions of PCR amplicons A–F are mapped onto a diagram of the SCR1 locus drawn to scale. (B) ChIP of histone H4 was performed on chromatin extracts from degron and control strains before and after one or 2 h of Sth1 degradation. Quantitative PCR was performed on the ChIP DNA at six amplicons (labeled A–F) encompassing the SCR1 locus. Enrichment values were normalized to an amplicon from the CDC2 ORF. Error bars indicate standard deviation between three biological replicates. (C) The chromatin structure from degron and control strains following Sth1 degradation was probed with increasing amounts of MNase and analyzed by Southern blot with a probe comprised of amplicons C and D. The blot was stripped and reprobed with amplicon A. (D) ChIP against myc-tagged H4 was performed in STH1+ cells during transcriptional activation and repression of SCR1. Pol III expression is active during nutrient replete conditions (T0), repressed after 25 and 180 min of starvation (T25 and T180), and active during nutrient restoration (T+25; Roberts et al, 2003). (E) Histone density over SCR1 was measured as in panel B in strains that were arrested in G1 using alpha-factor before degrading Sth1.
We next examined whether this progressive return of histone density was due to the loss of transcription, and whether DNA replication was additionally involved. To determine whether this nucleosomal gain was simply due to a lack of transcription, we performed histone ChIP with STH1+ cells during nutrient deprivation, which rapidly (within 10 min) represses Pol III transcription (Clarke et al, 1996; Harismendy et al, 2003; Roberts et al, 2003). Interestingly, histone density was not restored to SCR1 during Pol III repression (Figure 2D). Notably, the Pol III machinery is known to maintain residence at Pol III genes during Pol III repression (Roberts et al, 2006), prompting a test of whether RSC likewise maintains association. Here, we found that during repression RSC remained at the three Pol III loci tested: SCR1 and two tRNA genes (Supplementary Figure 2).
To determine whether replication-dependent chromatin assembly contributes to the restoration of histone density in the absence of RSC, we first arrested Sth1td strains in G1 phase with alpha-factor, degraded Sth1, and then examined histone density over the SCR1 locus after 2 h (Figure 2E). Interestingly, the restoration of histone density was attenuated compared with cycling cells, indicating that replication-dependent chromatin assembly accounts for a large portion of the acquired histone density on Pol III genes lacking RSC. Taken together, our work suggests that RSC helps keep Pol III genes nucleosome-deficient, regardless of their transcriptional status, and that the absence of RSC makes Pol III loci susceptible to replication-dependent assembly into chromatin.
A genome-scale microarray examines chromatin changes at numerous RSC targets
We then determined whether our observations of histone gain at SCR1 were representative of Pol III genes in general, and whether this observation extended to Pol II genes. Toward this end, we generated a custom RSC-specific miniature microarray at relatively high resolution to interrogate a large number of target loci. The array covers 218 specific genomic locations selected by their RSC occupancy levels and genomic content (Figure 3A; Ng et al, 2002). Four general categories of loci were selected: 99 highly-occupied RSC target loci containing 155 Pol II promoters (including 56 divergent pairs), 43 loci containing Pol II promoters not occupied by RSC, 54 loci containing Pol III genes of varying RSC occupancies, and 22 other loci. Each locus, which typically consisted of an intergenic and a portion of the flanking ORFs, was tiled with 60-mer oligonucleotide probes at a mean resolution of 110 bp, which was sufficient for monitoring changes in individual nucleosomes.
Figure 3.
Genome-scale analysis of RSC occupancy using a custom microarray shows dynamic occupancy at Pol II promoters. (A) A pie chart shows the distribution of the 1900 oligo probes from the custom microarray among four general design categories. The categories reflect loci predicted or not predicted to have RSC occupancy based on published results (Ng et al, 2002). (B) ChIP was performed against Rsc3 and Rsc8 in wild-type cells during a mock degron-inducing time course. The mean RSC occupancy values for 259 Pol II promoters (using a 500-bp window from −400 to +100 from the ATG) were determined and the distributions were plotted. The dynamic range of RSC fold enrichment of 0.8–1.2 is fully consistent with published observations (Damelin et al, 2002; Ng et al, 2002). (C) Promoters were assigned to three categories based on RSC occupancy before and after mock Sth1 loss. Non-occupied promoters had enrichment ratios of <0.95; moderate promoters had ratios between 0.95 and 1.05; and RSC-occupied promoters had ratios >1.05. (D) The mean RSC occupancy values for 49 tRNA genes were determined using a 400-bp window encompassing the tRNA and the distributions were plotted.
As an initial test of the microarray, we performed ChIP against RSC. We assessed occupancies of two different RSC subunits, Rsc3 and Rsc8, in wild-type STH1+ cells, since RSC demonstrates low cross-linking efficiency (Ng et al, 2002). ChIP eluates were amplified, labeled, and hybridized to the arrays. We averaged the RSC occupancy of all probes within a 500-bp window at each Pol II promoter on the array, and plotted the distribution of RSC occupancies at all the Pol II promoters (Figure 3B). Under normal growth conditions (mock before), we observed a distinct enrichment at a large number of promoters, consistent with our expectations from our array design. Comparison of our results with published results (Ng et al, 2002) revealed a strong positive correlation (Pearson correlation coefficient=0.87). Since the array represents a small fraction of the genome, the enrichment ratios are not absolute values but rather a relative representation of the RSC occupancy for the promoters on this array. Therefore, we divided promoters into classes representing the least, moderately, and most likely occupied by RSC; for simplicity, these classes are referred to as non-occupied, moderate, and RSC-occupied, respectively (Figure 3C). Based on these criteria, 84% of the promoters predicted to be RSC occupied demonstrated occupancy, and 63% of the promoters predicted not to be occupied demonstrated lack of occupancy; the majority of the remaining promoters fell in the moderate class. These results indicate the validity of using this microarray in analyzing RSC function at its target sites in vivo.
Since analysis of our sth1td allele requires growth in media containing galactose at 37°C, we needed to determine whether the RSC targets we identified under normal conditions remain targets under mock Sth1td protein degradation growth conditions. Therefore, we also performed RSC ChIP under mock degradation conditions (mock after). Interestingly, we observed a redistribution of RSC occupancy (Figure 3B). Presumably, this redistribution of RSC reflects the transcriptional program changes associated with the new growth conditions. Surveying the promoters present on this array, we found fewer but still significant numbers of promoters that we could confidently identify as having RSC occupancy (Figure 3C). Notably, we observed bidirectional shifts, with some promoters gaining occupancy while others losing occupancy. While shifts in RSC occupancy at Pol II promoters occurred, tRNA genes appeared remarkably stable in RSC occupancy during mock after conditions (Figure 3D). Together, these results indicated that RSC occupancy is dynamic at Pol II promoters and stable at Pol III genes under our conditions.
Nucleosomes return to Pol III genes upon loss of Sth1 protein
To examine chromatin changes at Pol III genes, we performed ChIP against tagged H4 in the degron and control strains and hybridized the eluates to the microarray. We first present microarray results from the SCR1 locus (Figure 4A). As a measure of change in histone occupancy between the presence and absence of Sth1 protein, we plotted the ratio of the signal from the degron and control strains, such that a ratio of 1 indicates no change, >1 indicates a gain of nucleosomes in the degron strain, and <1 indicates a loss of nucleosomes. In agreement with our PCR results (Figure 2), we observed a large gain in nucleosome density over the SCR1 gene after loss of Sth1 protein, indicating a gain in nucleosome occupancy.
Figure 4.
Pol III genes generally demonstrate nucleosome gain upon loss of Sth1 protein. (A) ChIP products against tagged histone H4 were applied to the RSC specific microarray. The results from the SCR1 locus are shown. Values representing the ratio between degron and control strains as an indication of histone gain (>1) or loss (<1) were plotted for before (gray) and after (black) Sth1 loss. The ChIP data were plotted against a physical map of the SCR1 locus drawn to scale. The microarray oligo probes are indicated as thick black lines. (B) The mean histone H4 ratio between degron and control strains representing histone gain or loss was determined at tRNA loci both before (gray) and after (black) Sth1 degradation and the distributions were plotted. (C) ChIP against the Pol III subunit Rpc82 was performed before and after Sth1 protein degradation and plotted as in panel B.
We next asked whether the gain of nucleosomes at SCR1 was representative of all Pol III genes. We examined as a class the general chromatin structure of loci containing a tRNA gene by identifying those microarray probes that were located within a 400-bp window centered over the tRNA gene (on average 4–5 probes). We then averaged the degron/control ratios of the probes within this window and plotted the distribution of the resulting values for all the tRNA loci (Figure 4B). We observed a dramatic shift in nucleosome occupancy at tRNA loci following the loss of Sth1 protein (P=2.59E−17), similar to what we observed at SCR1. Examination of two additional Pol III-transcribed genes, RPR1 and SNR6, also revealed general nucleosome gains, including both their coding and 5′ flanking intergenic regions (data not shown). Together, these results suggest that RSC, directly or indirectly, helps to remove nucleosomes from these loci, and that nucleosome density increases in the absence of RSC function. Furthermore, this effect appears to be a general feature at all Pol III loci.
Since an increase in nucleosome density was observed at tRNA loci, we next asked whether this would impact Pol III occupancy. We performed ChIP against the tagged Rpc82 subunit of Pol III in our sth1td strains and analyzed the eluates on our microarray (Figure 4C). Interestingly, we observed a striking, corresponding decrease in Pol III occupancy at tRNA loci in the degron strain upon loss of Sth1 protein (P=3.66E−15). This result suggests a mutually exclusive relationship between histone and Pol III, and may help explain the loss of Pol III transcription upon loss of Sth1 protein.
Loss of Sth1 affects nucleosome positioning at Pol II promoters
We next turned our attention to Pol II promoters. Since the loss of Sth1 had a profound impact on transcription, our ability to attribute transcriptional defects directly to RSC action relied upon observing a more robust change in chromatin structure at RSC-occupied promoters than at non-occupied promoters. To test for this, we focused on those probes located within promoter regions, defined as −400 to +100 bp from the ATG. The mean ratios of the probes within each promoter region were then plotted as a distribution. At non-occupied promoters, we detected a slight histone gain upon loss of Sth1 protein (Figure 5A; P=0.005). The effects at these ‘non-occupied' promoters could either be indirect or due to low or transient RSC occupancy. However, we also observed only modest changes at the vast majority of RSC-occupied promoters (Figure 5B; P=0.014), although there were a small percentage of promoters that did show a notable gain. Therefore, regardless of classification, the majority of Pol II promoters examined did not exhibit the large general gains in histone density observed at Pol III promoters.
Figure 5.
Pol II promoters demonstrate single nucleosome changes upon Sth1 loss. (A, B) The mean histone H4 ratio between degron and control strains representing histone gain or loss was determined at Pol II promoters both before (gray) and after (black) Sth1 degradation. Shown are the distributions for RSC non-occupied (A) and occupied (B) promoters. (C, D) Mononucleosomal DNA was isolated from degron and control strains by MNase digestion and hybridized to the microarray. The MEP represents the single probe within the promoter having the highest gain in nucleosome occupancy in the degron strain relative to the control strain. Shown are the distributions of MEP ratios at RSC non-occupied (C) and occupied (D) promoters. (E, F) The mononucleosome occupancy ratio between degron and control strains was determined for a 50-bp window encompassing the TSS. The distributions were plotted for RSC non-occupied (E) and occupied (F) promoters.
We then considered that RSC-associated chromatin changes at Pol II promoters might be more subtle than the gross nucleosome density changes that are detected by conventional ChIP. To examine chromatin structure at a more detailed level, we performed MNase digestion on chromatin and hybridized the liberated mononucleosomal DNA (∼150 bp) to our microarrays (Sekinger et al, 2005; Yuan et al, 2005). This method can detect either the gain or loss of a single nucleosome centered over an individual probe, and thus monitor single nucleosome events. With traditional ChIP, the fractional gain or loss of a single nucleosome may not be easily detected due to the larger fragment size (∼500 bp) and the associated summation of nucleosome occupancy. Hence, the mononucleosome technique can provide a more sensitive detection of chromatin alterations at the nucleosome level.
We anticipated that three types of chromatin events could be detected by mononucleosome hybridization: movement or sliding of a nucleosome, as inferred by the concordant gain and loss in signal of neighboring probes; single nucleosome gain; and single nucleosome loss. Examples of three candidate loci demonstrating each class of events are presented in Figure 6. In some cases, two separate events were identified within the same promoter or intergenic (Figure 6B and C). In many cases, the MNase procedure revealed a single location where nucleosome density changed, which was either largely (Figure 6A) or partly (Figure 6B) masked in the H4 ChIP data due to the nucleosome averaging effect of ChIP. In other cases, the MNase data correlated well with the H4 ChIP results (Figure 6C); these cases may represent promoters where more than one nucleosome was gained. Together, these examples illustrate the types of nucleosome events occurring in response to loss of RSC function.
Figure 6.
RSC-occupied loci demonstrate single nucleosome movement, gain, and loss. Three loci demonstrating single nucleosome events are illustrated. The upper graphs represent histone H4 ratios between degron and control strains. The middle graph represents the mononucleosome occupancy ratios between degron and control strains. The lower cartoon represents the physical map of the locus as described in Figure 4A. TSSs are represented by bent arrows and ORFs by rectangles. (A) The PMI40/YER004W locus demonstrates nucleosome movement. (B) The SRB2/NCP1 locus demonstrates independent single nucleosome gain and loss. (C) The HHT2/HHF2 locus demonstrates nucleosome movement and general nucleosome gain.
As a simple measure of nucleosome events, we identified the probe in each promoter that gained the most nucleosome occupancy from the mononucleosome hybridization, termed the Most Enriched Probe (MEP), in the degron strain relative to the control strain (Figure 5C and D). The MEP could arise either from a single nucleosome gain or through a nucleosome movement where the gain is associated with the repositioning of the nucleosome completely over a probe location. Examination of the distribution of MEP ratios in non-occupied promoters revealed a slight gain in the magnitude of the MEP (P=0.001), a change similar to what was observed with the ChIP result. However, examination of the RSC-occupied promoters revealed a prominent shift in the distribution (P=2.5E−13), indicating that loss of RSC function resulted in a gain of nucleosome occupancy over at least one probe. These results suggested that single nucleosome events are much more prevalent than changes in gross nucleosome occupancy and are the primary effects due to the loss of RSC function at Pol II promoters.
Single nucleosome events generated by the loss of RSC function may impact transcription, depending upon the location of the event relative to promoter sequence elements. The TSS usually resides within the NFR and may be subject to regulation by nucleosome positioning (Yuan et al, 2005). We identified those probes that overlap a 50-bp window centered over the TSS and examined the distribution of degron/control ratios (Figure 5E and F). We observed no significant shift in the distribution of ratios for non-occupied promoters (P=0.28). However, a clear shift is noted at RSC-occupied promoters (P=1.57E−4), suggesting that one location of RSC action may be the NFR that encompasses the TSS.
Transcriptional shutoff by rpb1-1 results in nucleosome loss in the promoter
The observation of chromatin changes at the TSS prompted us to ask whether the changes we observed in our sth1td mutant were in direct response to the loss of RSC activity or indirectly due to the loss of transcription. To help answer this question we prepared mononucleosomal DNA from an rpb1-1 mutant both at permissive and non-permissive temperature (mimicking our degron time-course protocol). This temperature-sensitive Pol II mutant abrogates new transcription within minutes of shifting to a non-permissive temperature (Nonet et al, 1987). We determined the distribution of MEP ratios (non-permissive over permissive temperature) for both RSC-occupied and non-occupied promoters (Supplementary Figure 3). Surprisingly, we observed a lower mean ratio at RSC-occupied promoters compared with non-occupied promoters (P=1.32E−6), suggesting that RSC promoters generally lose mononucleosomes upon loss of transcription in the rpb1-1 mutant. Furthermore, when the probes corresponding to the TSS were examined, no significant difference in the distribution was observed between the two promoter classes (P=0.59). These results suggested that loss of transcription in rpb1-1 strains is associated with structural changes in promoter chromatin, including a general loss of nucleosome occupancy. The mechanism underlying the loss of nucleosomes in the rpb1-1 strain is unknown, but the result nevertheless highlights the distinction between changes in promoter chromatin structure due to loss of transcription versus loss of RSC activity.
Discussion
Here, we describe the effects on chromatin structure conferred by the loss of RSC activity. Our initial observation with the sth1td mutant was a surprisingly large reduction in new transcription by all three RNA polymerases. Two possibilities are considered for this widespread effect. First, RSC could have a direct role in all transcription. There is some support for this notion, as a direct interaction between the subunit Rsc4 and Rpb5, a conserved subunit shared among all three RNA polymerases, was recently described (Soutourina et al, 2006). Furthermore, RSC appears to be generally recruited by transcriptional machinery at Pol III genes (Ng et al, 2002). This provides the potential for RSC as a direct coactivator at all gene classes in the genome. However, while the association of RSC with RNA polymerases likely occurs at some targets, the vast majority of RSC is not associated with polymerases (Cairns et al, 1996). Importantly, RSC is observed specifically at some promoters and not others, irrespective of transcriptional status (Damelin et al, 2002; Ng et al, 2002). Second, we considered that the profound loss of transcription might be a combination of both direct and indirect effects. Loss of products derived from direct target genes may induce a cascade of cell stresses, which could impact the remainder of the non-target genome. We note that RSC target genes are involved in a wide variety of processes including mitochondrial function, nitrogen and carbohydrate metabolism, and transcriptional regulation (Damelin et al, 2002; Ng et al, 2002). Additionally, RSC mutants affect the expression of cell cycle and ribosomal proteins, Gcn4 target genes, and others (Angus-Hill et al, 2001; Swanson et al, 2003; Soutourina et al, 2006). Our analysis identified key RSC target genes that demonstrated promoter chromatin effects, including SRB2, a mediator subunit; most tRNA genes; mitochondrial genes MAS6, COX19, and OM14; and both loci that encode H3 and H4 (Figures 4 and 6, and data not shown). Simultaneous loss of transcription of key RSC targets would likely have negative global consequences. Hence, we conclude that the global loss of transcription likely reflects both direct and indirect effects.
We observed two classes of chromatin effects due to the loss of RSC function: a change in general nucleosome density and specific single nucleosomal events. These effects are not necessarily mutually exclusive, as for example both nucleosome movement and gain were detected at the HTT2/HHF2 locus. Furthermore, extensive sliding of one nucleosome toward a second neighboring nucleosome may result in the displacement and disassembly of the second nucleosome, a situation that would not be distinguished from simple ejection using our techniques. Gross nucleosome density changes were observed primarily at Pol III genes and at a limited number of Pol II promoters. This effect likely reflects the role of RSC to remove or eject one or more nucleosomes, perhaps to allow room for transcriptional machinery that would normally be occluded by a nucleosome. Indeed, a loss of Pol III occupancy was correlated with an increase in histone density at tRNA loci (Figure 4); this is consistent with the requirement for RSC for full Pol III transcription (Soutourina et al, 2006). In contrast, single nucleosome events, including movement, may represent more subtle chromatin effects at Pol II genes. In this case, a nucleosome that partially or wholly occludes a transcription factor-binding site, such as a TATA box, could be shifted, rather than ejected, for access. In the absence of RSC, nucleosomes may either adopt a more energetically favorable position dictated by an underlying NPS, or be actively deposited and/or repositioned into a more regular, and presumably transcriptionally repressive, chromatin structure by remodelers such as ISWI. The determination of whether RSC will move or eject a nucleosome at a target locus may depend upon specific signals, either in the form of histone modifications on the target nucleosome or from direct interactions with transcriptional machinery. The numerous bromodomains present in RSC are candidates for this instructional mechanism.
The decision to eject nucleosomes (or to perform extensive sliding) may be general at Pol III genes, as we observed occupancy effects at all Pol III genes assayed. This localized gain in histone occupancy could be due to either de novo deposition or possibly a redistribution of nucleosomes from flanking regions. Regardless of the mechanism, the nucleosomes that return to these loci appear to do so in a non-positioned manner, as we were unable to detect gain by mononucleosome hybridization (data not shown); well-positioned nucleosomes should be detected by this technique, whereas randomly distributed nucleosomes may not be detected. This result is not unexpected considering the striking lack of detectable NPS signatures near tRNA genes (Ioshikhes et al, 2006; Segal et al, 2006). In addition, Pol III genes, along with other highly transcribed genes, exhibit high rates of histone turnover (Dion et al, 2007). Together, the observations that tRNA genes normally demonstrate low nucleosome density, high histone turnover, high RSC occupancy, and gain of nucleosome density in the absence of RSC activity clearly suggest that RSC is required to maintain tRNA genes in a nucleosome-sparse environment. Here, the acquisition of nucleosomes is clearly not due to a lack of Pol III transcription, as repressing Pol III transcription by nutrient deprivation was not accompanied by a gain in nucleosomes at SCR1. Furthermore, our occupancy data show that both the Pol III machinery (Roberts et al, 2006) and RSC (Supplementary Figure 2) reside at Pol III genes during Pol III repression (nutrient deprivation for 25 min). This raises the possibility that acutely repressed Pol III genes are bound by a poised Pol III–RSC complex that helps keep these loci nucleosome-deficient, enabling rapid reactivation when environmental conditions improve. A similar role for RSC may also be required upon exit from S phase. Here, replication may displace the Pol III machinery, resulting in a competition for Pol III gene occupancy between histones (facilitated by replication-dependent chromatin assembly) and the Pol III machinery (facilitated by RSC). While RSC may be responsible for the ejection of nucleosomes at tRNA genes, other chromatin remodelers likely help deposit and position new nucleosomes. One candidate is ISW2, which is also localized near tRNA genes (Gelbart et al, 2005), raising the possibility that RSC and ISW2 have antagonistic functions at these loci.
Loss of transcription is often associated with changes in promoter chromatin. For example, repression of the PHO5 promoter involves re-establishment of several promoter nucleosomes (Adkins and Tyler, 2006). In response to Sth1 degradation, we observed alterations in the promoter chromatin at a large number of promoters, which were likely associated with or contributed to the observed transcriptional repression. Interestingly, while a small fraction of promoters demonstrated a change in gross histone occupancy, a larger number exhibited single nucleosome events. In counter distinction to our observation, an interesting recent genome-wide study of CHD remodelers in Schizosaccharomyces pombe found general gains in nucleosome density at the promoters of Pol II genes in the absence of CHD activity (Walfridsson et al, 2007); Pol III genes were not examined. This distinction likely reflects the mode of action: CHD remodelers primarily catalyze the disassembly of nucleosomes at Pol II genes in vivo (Walfridsson et al, 2007), whereas our work suggests that SWI/SNF remodelers (RSC) generally slide nucleosomes at Pol II promoters. We observed two key differences between RSC-occupied and non-occupied Pol II promoters. First, a larger fraction of RSC-occupied promoters had high-value MEP ratios. Second, and more importantly, the change in nucleosome occupancy at the TSS was generally higher at RSC-occupied than non-occupied promoters. Since the TSS is often located within an NFR (Yuan et al, 2005), these observations may suggest that a key role of RSC function at target promoters is to help establish and/or maintain the NFR.
Two lines of evidence suggest that the changes in promoter chromatin structure at RSC target genes were not simply due to the loss of transcription. First, we observed specific chromatin alterations at RSC-occupied promoters relative to non-occupied promoters, even though new transcription at all promoters was reduced. Second, abolishing transcription in the rpb1-1 mutant at the non-permissive temperature resulted in chromatin changes, although distinct from those in the sth1td mutant. The nature of the nucleosome loss in the rpb1-1 strain is not known, although we speculate that the continued action of transcriptional coactivators, including chromatin remodelers such as RSC, results in an apparent decrease in nucleosome occupancy within the promoter. In addition, continued abortive initiation and the lack of productive elongation by Pol II may prevent the restoration of promoter architecture to a non-active state, perhaps by the stalled transcriptional machinery inhibiting nucleosome reassembly.
Together, these results demonstrate a clear role for RSC in the process of transcription in vivo. Our work is consistent with RSC remodeling nucleosomes to allow establishment of RNA polymerase machinery. However, the mode of remodeling is clearly biased; at Pol II promoters single nucleosome events consistent with sliding are more often observed, whereas at Pol III genes the events are consistent with ejection. Future work will explore the nature of this bias.
Materials and methods
Media, genetic methods, and strains
Standard procedures were used for media preparation, transformations, integrations, and genetic analyses. Plasmid and strain constructions will be provided upon request. Strain genotypes (S288C derivatives) are listed in Table I. For degron inductions, cultures were grown in liquid YP containing 2% raffinose and 0.05 mM Cu(II)SO4 at 27°C until OD600 0.4–0.6. Galactose was then added to 2% final concentration and the culture incubated for 60 min to induce UBR1. Afterwards, the temperature was shifted to 37°C for 120 min to induce Sth1 degradation. Samples were cross-linked in 1% formaldehyde for 5 min for mononucleosome analysis, 15 min for histone ChIP, and 30 min for other proteins. For G1 arrest studies, alpha-factor was added to bar1Δ degron and control strains simultaneously with galactose and the time course followed as above. Loss of BAR1 did not affect Sth1 degradation. The nutrient deprivation time-course experiment was performed as described (Roberts et al, 2003).
Table 1.
Yeast strains
Strain | Genotype | Source |
---|---|---|
YBC2191 | MATa ura3-52 trp1Δ63 his3Δ200 leu2∷PET56 ubr1Δ∷pGAL1-UBR1∷HIS3 sth1Δ∷pCUP1-sth1td∷URA3 | This work |
YBC2192 | MATa ura3-52 trp1Δ63 his3Δ200 leu2∷PET56 ubr1Δ∷HIS3 sth1Δ∷pCUP1-sth1td∷URA3 | This work |
YBC2193 | MATa ura3-52 trp1Δ63 his3Δ200 leu2∷PET56 ubr1Δ∷pGAL1-UBR1∷HIS3 sth1Δ∷pCUP1-sth1td∷URA3 RPC82-13myc∷TRP1 | This work |
YBC2194 | MATa ura3-52 trp1Δ63 his3Δ200 leu2∷PET56 ubr1Δ∷HIS3 sth1Δ∷pCUP1-sth1td∷URA3 RPC82-13myc∷TRP1 | This work |
MX1-4C | MATα ura3-52 leu2-3,112 trp1-289 his3Δ1 (hht1-hhf1)Δ (hht2-hhf2)Δ pMS329 [CEN, ARS, URA3, HHT1, HHF1] | Morgan et al (1991) |
YBC2205 | MATα ura3-52 leu2-3,112 trp1-289 his3Δ1 (hht1-hhf1)Δ (hht2-hhf2)Δ pNOY436 [CEN, ARS, TRP1, HHT2, myc-HHF2] | This work |
YBC2449 | MATα ura3-52 leu2-3,112 trp1-289 his3Δ1 (hht1-hhf1)Δ (hht2-hhf2)Δ ubr1Δ∷pGAL1-UBR1∷HIS3 sth1Δ∷pCUP1-sth1td∷URA3 pNOY436 [CEN, ARS, TRP1, HHT2, myc-HHF2] | This work |
YBC2451 | MATα ura3-52 leu2-3,112 trp1-289 his3Δ1 (hht1-hhf1)Δ (hht2-hhf2)Δ ubr1Δ∷HIS3 sth1Δ∷pCUP1-sth1td∷URA3 pNOY436 [CEN, ARS, TRP1, HHT2, myc-HHF2] | This work |
YBC3135 | MATa ura3-52 trp1Δ63 his3Δ200 leu2∷PET56 ubr1Δ∷pGAL1-UBR1∷HIS3 sth1Δ∷pCUP1-sth1td∷URA3 bar1Δ∷KanMX | This work |
YBC3136 | MATa ura3-52 trp1Δ63 his3Δ200 leu2∷PET56 ubr1Δ∷HIS3 sth1Δ∷pCUP1-sth1td∷URA3 bar1Δ∷KanMX | This work |
yHN3 | MATa ura3-52 trp1Δ63 his3Δ200 leu2∷PET56 Rsc3-9myc∷TRP1 | Ng et al (2002) |
yHN4 | MATa ura3-52 trp1Δ63 his3Δ200 leu2∷PET56 Rsc8-9myc∷TRP1 | Ng et al (2002) |
Y260 | MATa ura3-52 rpb1-1 | Nonet et al (1987) |
In vivo RNA labeling
Cells were grown following the degron time course above. Aliquots (2 ml) of culture were incubated with 165 μCi of 3H-labeled uracil for 25 min at each time point. Total RNA was immediately harvested using acid phenol extraction. Messenger RNA was enriched by one round of poly-A selection on oligo-dT beads. Enriched RNA was separated by denaturing polyacrylamide gel electrophoresis, followed by either staining with ethidium bromide or electrophoretic transfer to nylon membrane. The membrane was soaked in Amplify (Amersham) and exposed to film.
Microarray design and analyses
The microarray format was based on Agilent Technologies' 8-pack 22K format. Details on the custom microarray design, ChIP, mononucleosome preparation, hybridization, and data processing are available in Supplementary data. Microarray data are provided as Supplementary data and are present at the NCBI GEO repository under experiment GSE8862.
Genomic and statistical analyses
Promoters were defined as −400 to +100 relative to the translation start site. Regions containing tRNA genes were defined as −150 to +250 relative to the start site. Probes within or overlapping these regions were selected for analysis. TSSs were collected from published sources (Zhang and Dietrich, 2005; Miura et al, 2006). The most commonly used TSS position or, failing that, the position closest to the ATG was selected and used in the analysis. P-values between data sets before and after Sth1 degradation were determined by two-tailed, equal-variance Student's t-test. Distribution histograms for H4 ChIP, MEP, and Pol III ChIP were normalized so that the mean for the Before curve was 1.
Supplementary Material
Supplementary Information
Supplementary Figure 1
Supplementary Figure 2
Supplementary Figure 3
Supplementary Data 1
Supplementary Data 2
Supplementary Data 3
Supplementary Data 4
Supplementary Data 5
Supplementary Data 6
Acknowledgments
We thank Kevin Struhl and Mitch Smith for providing yeast strains, Jürgen Dohmen for the degron tag plasmids, Masayasu Nomura for the H4myc plasmid, Brian Dalley and the HCI microarray facility for microarray hybridizations and analysis, Doug Roberts for technical advice, and members of the Cairns laboratory for critical comments. This work was supported by the National Institutes of Health grant GM60415 (BRC), the Multidisciplinary Cancer Research Training Program grant CA093247 (TJP), and the Howard Hughes Medical Institute. Core facilities are supported by grant CA24014.
References
- Adkins MW, Howar SR, Tyler JK (2004) Chromatin disassembly mediated by the histone chaperone Asf1 is essential for transcriptional activation of the yeast PHO5 and PHO8 genes. Mol Cell 14: 657–666 [DOI] [PubMed] [Google Scholar]
- Adkins MW, Tyler JK (2006) Transcriptional activators are dispensable for transcription in the absence of Spt6-mediated chromatin reassembly of promoter regions. Mol Cell 21: 405–416 [DOI] [PubMed] [Google Scholar]
- Angus-Hill ML, Schlichter A, Roberts D, Erdjument-Bromage H, Tempst P, Cairns BR (2001) A Rsc3/Rsc30 zinc cluster dimer reveals novel roles for the chromatin remodeler RSC in gene expression and cell cycle control. Mol Cell 7: 741–751 [DOI] [PubMed] [Google Scholar]
- Boeger H, Griesenbeck J, Strattan JS, Kornberg RD (2004) Removal of promoter nucleosomes by disassembly rather than sliding in vivo. Mol Cell 14: 667–673 [DOI] [PubMed] [Google Scholar]
- Cairns BR, Lorch Y, Li Y, Zhang M, Lacomis L, Erdjument-Bromage H, Tempst P, Du J, Laurent B, Kornberg RD (1996) RSC, an essential, abundant chromatin-remodeling complex. Cell 87: 1249–1260 [DOI] [PubMed] [Google Scholar]
- Cairns BR, Schlichter A, Erdjument-Bromage H, Tempst P, Kornberg RD, Winston F (1999) Two functionally distinct forms of the RSC nucleosome-remodeling complex, containing essential AT hook, BAH, and bromodomains. Mol Cell 4: 715–723 [DOI] [PubMed] [Google Scholar]
- Campsteijn C, Wijnands-Collin AM, Logie C (2007) Reverse genetic analysis of the yeast RSC chromatin remodeler reveals a role for RSC3 and SNF5 Homolog 1 in ploidy maintenance. PLoS Genet 3: e92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chai B, Huang J, Cairns BR, Laurent BC (2005) Distinct roles for the RSC and Swi/Snf ATP-dependent chromatin remodelers in DNA double-strand break repair. Genes Dev 19: 1656–1661 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clarke EM, Peterson CL, Brainard AV, Riggs DL (1996) Regulation of the RNA polymerase I and III transcription systems in response to growth conditions. J Biol Chem 271: 22189–22195 [DOI] [PubMed] [Google Scholar]
- Damelin M, Simon I, Moy TI, Wilson B, Komili S, Tempst P, Roth FP, Young RA, Cairns BR, Silver PA (2002) The genome-wide localization of Rsc9, a component of the RSC chromatin-remodeling complex, changes in response to stress. Mol Cell 9: 563–573 [DOI] [PubMed] [Google Scholar]
- Dion MF, Kaplan T, Kim M, Buratowski S, Friedman N, Rando OJ (2007) Dynamics of replication-independent histone turnover in budding yeast. Science 315: 1405–1408 [DOI] [PubMed] [Google Scholar]
- Dohmen RJ, Wu P, Varshavsky A (1994) Heat-inducible degron: a method for constructing temperature-sensitive mutants. Science 263: 1273–1276 [DOI] [PubMed] [Google Scholar]
- Gelbart ME, Bachman N, Delrow J, Boeke JD, Tsukiyama T (2005) Genome-wide identification of Isw2 chromatin-remodeling targets by localization of a catalytically inactive mutant. Genes Dev 19: 942–954 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Geng F, Laurent BC (2004) Roles of SWI/SNF and HATs throughout the dynamic transcription of a yeast glucose-repressible gene. EMBO J 23: 127–137 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Govind CK, Yoon S, Qiu H, Govind S, Hinnebusch AG (2005) Simultaneous recruitment of coactivators by Gcn4p stimulates multiple steps of transcription in vivo. Mol Cell Biol 25: 5626–5638 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Groth A, Rocha W, Verreault A, Almouzni G (2007) Chromatin challenges during DNA replication and repair. Cell 128: 721–733 [DOI] [PubMed] [Google Scholar]
- Harismendy O, Gendrel CG, Soularue P, Gidrol X, Sentenac A, Werner M, Lefebvre O (2003) Genome-wide location of yeast RNA polymerase III transcription machinery. EMBO J 22: 4738–4747 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hassan AH, Prochasson P, Neely KE, Galasinski SC, Chandy M, Carrozza MJ, Workman JL (2002) Function and selectivity of bromodomains in anchoring chromatin-modifying complexes to promoter nucleosomes. Cell 111: 369–379 [DOI] [PubMed] [Google Scholar]
- Hsu JM, Huang J, Meluh PB, Laurent BC (2003) The yeast RSC chromatin-remodeling complex is required for kinetochore function in chromosome segregation. Mol Cell Biol 23: 3202–3215 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang J, Hsu JM, Laurent BC (2004) The RSC nucleosome-remodeling complex is required for cohesin's association with chromosome arms. Mol Cell 13: 739–750 [DOI] [PubMed] [Google Scholar]
- Inai T, Yukawa M, Tsuchiya E (2007) Interplay between chromatin and trans-acting factors on the IME2 promoter upon induction of the gene at the onset of meiosis. Mol Cell Biol 27: 1254–1263 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ioshikhes IP, Albert I, Zanton SJ, Pugh BF (2006) Nucleosome positions predicted through comparative genomics. Nat Genet 38: 1210–1215 [DOI] [PubMed] [Google Scholar]
- Kasten M, Szerlong H, Erdjument-Bromage H, Tempst P, Werner M, Cairns BR (2004) Tandem bromodomains in the chromatin remodeler RSC recognize acetylated histone H3 Lys14. EMBO J 23: 1348–1359 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Labib K, Tercero JA, Diffley JF (2000) Uninterrupted MCM2-7 function required for DNA replication fork progression. Science 288: 1643–1647 [DOI] [PubMed] [Google Scholar]
- Langst G, Bonte EJ, Corona DF, Becker PB (1999) Nucleosome movement by CHRAC and ISWI without disruption or trans-displacement of the histone octamer. Cell 97: 843–852 [DOI] [PubMed] [Google Scholar]
- Li B, Carey M, Workman JL (2007) The role of chromatin during transcription. Cell 128: 707–719 [DOI] [PubMed] [Google Scholar]
- Lomvardas S, Thanos D (2001) Nucleosome sliding via TBP DNA binding in vivo. Cell 106: 685–696 [DOI] [PubMed] [Google Scholar]
- Lorch Y, Zhang M, Kornberg RD (1999) Histone octamer transfer by a chromatin-remodeling complex. Cell 96: 389–392 [DOI] [PubMed] [Google Scholar]
- Miura F, Kawaguchi N, Sese J, Toyoda A, Hattori M, Morishita S, Ito T (2006) A large-scale full-length cDNA analysis to explore the budding yeast transcriptome. Proc Natl Acad Sci USA 103: 17846–17851 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moreira JM, Holmberg S (1999) Transcriptional repression of the yeast CHA1 gene requires the chromatin-remodeling complex RSC. EMBO J 18: 2836–2844 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morgan BA, Mittman BA, Smith MM (1991) The highly conserved N-terminal domains of histones H3 and H4 are required for normal cell cycle progression. Mol Cell Biol 11: 4111–4120 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ng HH, Robert F, Young RA, Struhl K (2002) Genome-wide location and regulated recruitment of the RSC nucleosome-remodeling complex. Genes Dev 16: 806–819 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nonet M, Scafe C, Sexton J, Young R (1987) Eucaryotic RNA polymerase conditional mutant that rapidly ceases mRNA synthesis. Mol Cell Biol 7: 1602–1611 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rao B, Shibata Y, Strahl BD, Lieb JD (2005) Dimethylation of histone H3 at lysine 36 demarcates regulatory and nonregulatory chromatin genome-wide. Mol Cell Biol 25: 9447–9459 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roberts DN, Stewart AJ, Huff JT, Cairns BR (2003) The RNA polymerase III transcriptome revealed by genome-wide localization and activity–occupancy relationships. Proc Natl Acad Sci USA 100: 14695–14700 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roberts DN, Wilson B, Huff JT, Stewart AJ, Cairns BR (2006) Dephosphorylation and genome-wide association of Maf1 with Pol III-transcribed genes during repression. Mol Cell 22: 633–644 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saha A, Wittmeyer J, Cairns BR (2002) Chromatin remodeling by RSC involves ATP-dependent DNA translocation. Genes Dev 16: 2120–2134 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saha A, Wittmeyer J, Cairns BR (2006) Chromatin remodelling: the industrial revolution of DNA around histones. Nat Rev Mol Cell Biol 7: 437–447 [DOI] [PubMed] [Google Scholar]
- Segal E, Fondufe-Mittendorf Y, Chen L, Thastrom A, Field Y, Moore IK, Wang JP, Widom J (2006) A genomic code for nucleosome positioning. Nature 442: 772–778 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sekinger EA, Moqtaderi Z, Struhl K (2005) Intrinsic histone–DNA interactions and low nucleosome density are important for preferential accessibility of promoter regions in yeast. Mol Cell 18: 735–748 [DOI] [PubMed] [Google Scholar]
- Shim EY, Hong SJ, Oum JH, Yanez Y, Zhang Y, Lee SE (2007) RSC mobilizes nucleosomes to improve accessibility of repair machinery to the damaged chromatin. Mol Cell Biol 27: 1602–1613 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sif S, Stukenberg PT, Kirschner MW, Kingston RE (1998) Mitotic inactivation of a human SWI/SNF chromatin remodeling complex. Genes Dev 12: 2842–2851 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Soutourina J, Bordas-Le Floch V, Gendrel G, Flores A, Ducrot C, Dumay-Odelot H, Soularue P, Navarro F, Cairns BR, Lefebvre O, Werner M (2006) Rsc4 connects the chromatin remodeler RSC to RNA polymerases. Mol Cell Biol 26: 4920–4933 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Swanson MJ, Qiu H, Sumibcay L, Krueger A, Kim SJ, Natarajan K, Yoon S, Hinnebusch AG (2003) A multiplicity of coactivators is required by Gcn4p at individual promoters in vivo. Mol Cell Biol 23: 2800–2820 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tsuchiya E, Hosotani T, Miyakawa T (1998) A mutation in NPS1/STH1, an essential gene encoding a component of a novel chromatin-remodeling complex RSC, alters the chromatin structure of Saccharomyces cerevisiae centromeres. Nucleic Acids Res 26: 3286–3292 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Walfridsson J, Khorosjutina O, Matikainen P, Gustafsson CM, Ekwall K (2007) A genome-wide role for CHD remodelling factors and Nap1 in nucleosome disassembly. EMBO J 26: 2868–2879 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Whitehouse I, Flaus A, Cairns BR, White MF, Workman JL, Owen-Hughes T (1999) Nucleosome mobilization catalysed by the yeast SWI/SNF complex. Nature 400: 784–787 [DOI] [PubMed] [Google Scholar]
- Whitehouse I, Stockdale C, Flaus A, Szczelkun MD, Owen-Hughes T (2003) Evidence for DNA translocation by the ISWI chromatin-remodeling enzyme. Mol Cell Biol 23: 1935–1945 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang X, Zaurin R, Beato M, Peterson CL (2007) Swi3p controls SWI/SNF assembly and ATP-dependent H2A–H2B displacement. Nat Struct Mol Biol 14: 540–547 [DOI] [PubMed] [Google Scholar]
- Yuan GC, Liu YJ, Dion MF, Slack MD, Wu LF, Altschuler SJ, Rando OJ (2005) Genome-scale identification of nucleosome positions in S. cerevisiae. Science 309: 626–630 [DOI] [PubMed] [Google Scholar]
- Yudkovsky N, Logie C, Hahn S, Peterson CL (1999) Recruitment of the SWI/SNF chromatin remodeling complex by transcriptional activators. Genes Dev 13: 2369–2374 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Z, Dietrich FS (2005) Mapping of transcription start sites in Saccharomyces cerevisiae using 5′ SAGE. Nucleic Acids Res 33: 2838–2851 [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Supplementary Information
Supplementary Figure 1
Supplementary Figure 2
Supplementary Figure 3
Supplementary Data 1
Supplementary Data 2
Supplementary Data 3
Supplementary Data 4
Supplementary Data 5
Supplementary Data 6