In eukaryotes, origin recognition complex (ORC) proteins establish the pre-replicative complex (preRC) at the origins, and this is essential for the initiation of DNA replication. Open chromatin structures regulate the efficiency of preRC formation and replication initiation. However, the molecular mechanisms that control chromatin structure, and how the preRC components establish themselves on the chromatin remain to be understood. In human cells, the ORC is a highly dynamic complex with many separate functions attributed to sub-complexes or individual subunits of the ORC, including heterochromatin organization, telomere and centromere function, centrosome duplication and cytokinesis. We demonstrate that human Orc5, unlike other ORC subunits, when ectopically tethered to a chromatin locus, induces large-scale chromatin decondensation, predominantly during G1 phase of the cell cycle. Orc5 associates with the H3 histone acetyl transferase GCN5 (also known as KAT2A), and this association enhances the chromatin-opening function of Orc5. In the absence of Orc5, histone H3 acetylation is decreased at the origins. We propose that the ability of Orc5 to induce chromatin unfolding during G1 allows the establishment of the preRC at the origins.

In eukaryotes, the initiation of DNA replication requires the coordinated action of a multiprotein pre-replication complex (preRC) at the origins (Bell and Dutta, 2002). The origin recognition complex (ORC), a six-subunit complex, binds to replication origins during the G1 phase of the cell cycle, and this is followed by a sequential assembly of other preRC components (Bell and Stillman, 1992). In addition to their role in replication initiation, ORC subunits contribute to other cellular processes including transcriptional silencing, heterochromatin organization, sister chromatid cohesion, centrosome duplication, telomere maintenance and cytokinesis (Sasaki and Gilbert, 2007).

Open chromatin structures are known to regulate the efficiency of preRC formation, thereby facilitating replication initiation (Papior et al., 2012). However, the molecular mechanisms that affect chromatin structure and how the preRC components establish themselves on the chromatin remain to be understood. The accessibility of the replication factors is influenced by the chromatin structure, and the chromatin architecture dictates the efficiency of origin usage and firing (Brown et al., 1991; Ferguson and Fangman, 1992; Simpson, 1990; Stevenson and Gottschling, 1999). Histone acetylation is known to play a key role in the regulation of origins of DNA replication in yeast and Drosophila, and there is accumulating evidence that the deacetylation of histones negatively affects origin activity (Aggarwal and Calvi, 2004; Groth et al., 2007a; Knott et al., 2009b; Unnikrishnan et al., 2010; Vogelauer et al., 2002). Furthermore, the replication timing of the β-globin gene domain in human cells is also modulated by histone modifications at the origin (Goren et al., 2008). There is accumulating evidence that histone acetyl transferases act as positive regulators of replication origins in yeast and Drosophila as well as human cells (Groth et al., 2007b; Knott et al., 2009a). In yeast, GCN5p (also known as KAT2A in humans), a histone acetyl transferase (HAT), has been found to positively stimulate DNA replication by negating the inhibitory effect of the histone deacetylases (Espinosa et al., 2010; Vogelauer et al., 2002). Further, Hat1p and its partner Hat2p interact with the ORC (Suter et al., 2007). In Drosophila, the HATs Chameau (Chm) and CBP (Nejire) stimulate origin activity (Aggarwal and Calvi, 2004; McConnell et al., 2012). In human cells, HBO1 (also known as KAT7), another HAT, associates with ORC and is required for the loading of the minichromosome maintenance complex (MCM) onto chromatin and for replication fork progression (Iizuka et al., 2006; Iizuka and Stillman, 1999; Miotto and Struhl, 2008, 2010). A recent study has pointed out that the acetylation of some histone lysine residues depends on the binding of ORC to the origin and that the acetylation is at its maximum on the nucleosomes adjacent to one side of the major initiation site (Liu et al., 2012). How the ORC regulates such chromatin modifications and how the chromatin structure at origins is organized remain to be defined.

The ORC comprises six subunits, and in human cells they are highly dynamic. The largest subunit, Orc1 is degraded at the end of G1, and rebinding of the protein to chromatin is an obligatory step for the establishment of the preRC in G1 (Méndez et al., 2002). The smallest subunit of ORC, Orc6, binds to the ORC in a transient manner and also has independent roles in cytokinesis (Bernal and Venkitaraman, 2011; Prasanth et al., 2002). Orc2, Orc3, Orc4 and Orc5 in human cells constitute the core ORC and are associated with each other throughout the cell cycle (Dhar et al., 2001; Vashee et al., 2001). Orc1, Orc4 and Orc5 are members of the AAA+ family of ATPases and contain consensus motifs. Mutations in the ATP-binding sites on Orc4 and Orc5 impair complex assembly, whereas the ATP binding of Orc1 is dispensable (Ranjan and Gossen, 2006; Siddiqui and Stillman, 2007). Orc2 and Orc3 also have a structure that is common to AAA+ proteins, but they do not possess a consensus ATP-binding motif.

Multiple subunits of human ORC – including Orc1, Orc2, Orc3 and Orc5 – and the protein ORC-associated (ORCA) have roles in heterochromatin organization (Giri et al., 2015; Prasanth et al., 2010; Shen et al., 2010). In yeast, Orc5 has distinguishable functions in replication initiation and silencing (Dillin and Rine, 1997). Further, in Drosophila and humans, the loss of multiple ORC subunits leads to chromosome segregation defects (Pflumm and Botchan, 2001; Prasanth et al., 2004b). In this manuscript, we report that Orc5 has a distinct function in chromatin unfolding. Ectopic tethering of Orc5 to a chromatin locus leads to dramatic chromatin decondensation, predominantly during G1 phase of the cell cycle. This chromatin-opening role of Orc5 requires the activity of the HAT GCN5. We propose that the Orc5 subunit of the ORC plays a key role in mediating large-scale chromatin-opening during G1 that, in turn, facilitates the loading of other preRC components onto the origins.

Ectopic tethering of Orc5 induces large-scale chromatin decondensation

To investigate the chromatin changes that occur when preRC proteins, including ORC proteins, bind to origins, we tethered individual subunits of the ORC to a heterochromatic locus using an in vivo human U2OS osteosarcoma cell system (CLTon) (Fig. 1A). This reporter carries a stably integrated 200-copy transgene array with lac operator repeats, and this heterochromatic locus is visualized through the stable expression of Cherry–lac-repressor (Cherry–LacI). Upon transcriptional activation of this reporter locus, through the addition of doxycycline, this locus shows chromatin decondensation (Janicki et al., 2004; Shen et al., 2010). We generated triple-fusion proteins of YFP–LacI–ORCs, and these were tethered to the CLTon locus. Targeting YFP–LacI to this locus showed association of LacI with the heterochromatic CLTon locus (Fig. 1Ba). Surprisingly, tethering of YFP–LacI–Orc5 caused dramatic decondensation at the CLTon locus, whereas none of the other ORC subunits, including Orc1, Orc2, Orc3, Orc4 and Orc6, caused any changes to the chromatin architecture at the locus (Fig. 1Ba). 81% of YFP–LacI–Orc5-tethered cells showed decondensation of the heterochromatic locus (Fig. 1Bb). Furthermore, the extent of decondensation upon tethering Orc5 to the locus was determined by calculating the area of the decondensed chromatin. Measurement of the area of decondensation upon tethering Orc5 revealed a range of chromatin decondensation, ranging 2–35 µm2 (Fig. 1Bc), whereas the control YFP–LacI cells showed condensed loci with sizes in the range 0.2–1.3 µm2 (Fig. 1Bc). The average area of the U2OS nuclei was found to be 360±101 µm2 (n=52 cells). Based on the area of decondensation, we categorized the Orc5-mediated decondensation phenotype into three categories: medium (2–6 µm2), large (6–10 µm2) and very large (10–35 µm2) (Fig. 1C). The tethering of Orc5 to the locus resulted in 37%, 34% and 29% of cells showing medium, large and very large ranges of decondensation, respectively.

Fig. 1.

Orc5 causes chromatin decondensation. (A) Schematic of the heterochromatic locus in U2OS 2-6-3 CLTon cells. The copy numbers for the indicated regions are shown. (Ba) Chromatin decondensation upon tethering YFP–LacI and the indicated YFP–LacI-tagged ORC proteins to the heterochromatic locus of CLTon cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm. (Bb) The percentage of cells with open loci upon tethering YFP–LacI or the indicated YFP–LacI-tagged ORC proteins to the heterochromatic locus of CLTon cells. Error bars represent s.d., n=3 independent experiments (100 loci examined upon tethering each construct). ***P<0.001 (Student's t-test). (Bc) Area of heterochromatic loci upon tethering YFP–LacI and YFP–LacI–Orc5. Error bars represent s.d., n=3 independent experiments (area of 25–50 loci measured upon tethering each construct). ****P<0.0001 (Student's t-test). Lines represent the median, and the boxes represent the 25–75th percentiles. (C) Chromatin decondensation upon tethering YFP–LacI and YFP–LacI–Orc5 to the heterochromatic locus of CLTon cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm. (D) Chromatin decondensation upon tethering YFP–LacI and the indicated YFP–LacI-tagged Orc proteins to the heterochromatic locus of AO3 cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm.

Fig. 1.

Orc5 causes chromatin decondensation. (A) Schematic of the heterochromatic locus in U2OS 2-6-3 CLTon cells. The copy numbers for the indicated regions are shown. (Ba) Chromatin decondensation upon tethering YFP–LacI and the indicated YFP–LacI-tagged ORC proteins to the heterochromatic locus of CLTon cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm. (Bb) The percentage of cells with open loci upon tethering YFP–LacI or the indicated YFP–LacI-tagged ORC proteins to the heterochromatic locus of CLTon cells. Error bars represent s.d., n=3 independent experiments (100 loci examined upon tethering each construct). ***P<0.001 (Student's t-test). (Bc) Area of heterochromatic loci upon tethering YFP–LacI and YFP–LacI–Orc5. Error bars represent s.d., n=3 independent experiments (area of 25–50 loci measured upon tethering each construct). ****P<0.0001 (Student's t-test). Lines represent the median, and the boxes represent the 25–75th percentiles. (C) Chromatin decondensation upon tethering YFP–LacI and YFP–LacI–Orc5 to the heterochromatic locus of CLTon cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm. (D) Chromatin decondensation upon tethering YFP–LacI and the indicated YFP–LacI-tagged Orc proteins to the heterochromatic locus of AO3 cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm.

We investigated the role of Orc5 in chromatin decondensation by utilizing another system, in this case a CHO-derived A03 cell line that contains 90 Mb of a homogenously staining region generated through stable integration and amplification of the LacO-DHFR vector (Li et al., 1998). Tethering Orc5 to the A03 locus also resulted in dramatic decondensation of this locus (Fig. 1D). The decondensation upon tethering of Orc5 was in the range 4.5–27 µm2, whereas tethering of YFP–LacI resulted in decondensation in the range 0.6–1.2 µm2.

We next determined the minimum domain of Orc5 that is required for its ability to mediate chromatin decondensation. Triple-fusion Orc5 truncation mutants (including fusions comprising amino acid residues 1–100, 101–200, 201–300, 301–400 and 301–435 of Orc5) were generated (Fig. 2A), and their ability to cause chromatin unfolding was examined (Fig. 2B). As described earlier, full-length YFP–LacI–Orc5 caused decondensation of 81% of the CLTon locus (Fig. 2C). The Orc5 truncation mutants 1–100, 101–200, 201–300 and 301–400 failed to mediate chromatin unfolding, but the N-terminal truncation 301–435 mutant resulted in chromatin unfolding in ∼40% of cells (Fig. 2B,C). The decondensation upon tethering YFP–LacI–Orc5 (301–400) was in the range 0.5–1.0 µm2, whereas tethering of YFP–LacI–Orc5 (301–435aa) resulted in decondensation in the range 2.5–19.0 µm2 (Fig. 2D). Our results indicate that the last 35 amino acid residues at the C-terminus of Orc5 are crucial for its chromatin decondensation function.

Fig. 2.

The last 35 amino acids of Orc5 are necessary for its function in chromatin decondensation. (A) Schematic representation of various truncation mutants of Orc5 containing a T7-epitope on the N-terminus. The specific domains that can cause decondensation are indicated with ‘+’. AAs, amino acid residues. (B) Chromatin decondensation upon tethering YFP–LacI and various truncation mutants of YFP–LacI–Orc5 to the heterochromatic locus of CLTon cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm. (C) The percentage of cells with open loci upon tethering either YFP–LacI or various truncation mutants of YFP–LacI–Orc5 to the heterochromatic locus of CLTon cells. Error bars represent s.d., n=3 independent experiments (100 loci examined upon tethering each construct). **P<0.01, ***P<0.001 (Student's t-test). The YFP–LacI and YFP–LacI–Orc5 data from Fig. 1Bb has been replotted here for ease of comparison with the YFP–LacI-tagged Orc5 mutants. (D) Area of heterochromatic loci upon tethering YFP–LacI, YFP–LacI–Orc5 and YFP–LacI–GCN5. Error bars represent s.d., n=3 independent experiments (areas of 25–50 loci measured upon tethering each construct). ***P<0.001, ****P<0.0001 (Student's t-test). Lines represent the median, and the boxes represent the 25–75th percentiles. The YFP–LacI and YFP–LacI–Orc5 data from Fig. 1Bc has been replotted here for ease of comparison with the YFP–LacI-tagged Orc5 mutants.

Fig. 2.

The last 35 amino acids of Orc5 are necessary for its function in chromatin decondensation. (A) Schematic representation of various truncation mutants of Orc5 containing a T7-epitope on the N-terminus. The specific domains that can cause decondensation are indicated with ‘+’. AAs, amino acid residues. (B) Chromatin decondensation upon tethering YFP–LacI and various truncation mutants of YFP–LacI–Orc5 to the heterochromatic locus of CLTon cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm. (C) The percentage of cells with open loci upon tethering either YFP–LacI or various truncation mutants of YFP–LacI–Orc5 to the heterochromatic locus of CLTon cells. Error bars represent s.d., n=3 independent experiments (100 loci examined upon tethering each construct). **P<0.01, ***P<0.001 (Student's t-test). The YFP–LacI and YFP–LacI–Orc5 data from Fig. 1Bb has been replotted here for ease of comparison with the YFP–LacI-tagged Orc5 mutants. (D) Area of heterochromatic loci upon tethering YFP–LacI, YFP–LacI–Orc5 and YFP–LacI–GCN5. Error bars represent s.d., n=3 independent experiments (areas of 25–50 loci measured upon tethering each construct). ***P<0.001, ****P<0.0001 (Student's t-test). Lines represent the median, and the boxes represent the 25–75th percentiles. The YFP–LacI and YFP–LacI–Orc5 data from Fig. 1Bc has been replotted here for ease of comparison with the YFP–LacI-tagged Orc5 mutants.

Upon examination of the C-terminal 400–435 residues, we found that it was enriched with acidic residues (Fig. S1A). It has previously been reported that targeting of acidic activators to heterochromatic chromatin domains can cause large-scale chromatin decondensation (Carpenter et al., 2005). We generated a mutant of Orc5 where multiple aspartic acid residues were replaced by alanine residues (Fig. S1A). However, this mutant, when tethered to the locus, showed similar levels of chromatin decondensation, suggesting that the ‘acidic domain’ within Orc5 is not required for decondensation (Fig. S1B). We next determined whether the ATP-binding ability of Orc5 is required for its chromatin-unfolding function. We generated Walker A mutant (K43A) and an arginine finger mutant (R166A) (Fig. S1A), and tethered these to the CLTon locus (Fig. S1C). The extent of chromatin decondensation upon tethering these mutants was comparable to that of the wild-type Orc5, suggesting that the ATP-binding ability of Orc5 is also dispensable for its chromatin decondensation function.

Orc5 associates with the histone acetyl transferase GCN5

Histone acetylation, catalyzed by various HATs, is linked with the open chromatin state and is known to facilitate transcription (Narlikar et al., 2002). Because Orc5 was found to induce chromatin decondensation, we asked whether it achieves this through its association with known HATs. We transfected T7–Orc5 into U2OS cells and performed immunoprecipitation with an antibody against T7 (Fig. 3A). We found that T7–Orc5 interacted strongly with endogenous GCN5 (Fig. 3A). Next, we co-transfected T7–Orc5 and Flag–GCN5, and performed immunoprecipitation with an antibody against Flag. Orc5 and GCN5 were found to interact with one another (Fig. 3B). A reverse immunoprecipitation experiment recapitulated the interaction (Fig. 3E, full-length T7–Orc5 lane). Further, tethering of GCN5 to the CLTon locus also showed strong recruitment of Orc5 to the site, corroborating the immunoblot results (Fig. 3C).

Fig. 3.

Orc5 interacts with GCN5. (A) Semi-endogenous immunoprecipitation (IP) in U2OS cells expressing T7–Orc5 and analysis by immunoblotting for T7 and GCN5. (B) Immunoprecipitation in U2OS cells expressing T7–Orc5 and Flag–GCN5 by using an anti-Flag antibody and analysis by immunoblotting for T7 and Flag. (C) Tethering of YFP–LacI–GCN5 to the CLTon locus shows recruitment of Orc5 to the site. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 μm. (D) Schematic representation of various truncation mutants of Orc5 containing a T7-epitope on the N-terminus. The specific domains that can associate with GCN5 and Orc2 based on immunoprecipitation and immunoblotting analysis (E) are depicted as ‘+’. (E) Immunoprecipitation from U2OS cells expressing various T7-tagged Orc5 mutants and Flag–GCN5 using an antibody against T7 and immunoblotting for T7, Flag and Orc2. (Fa) Chromatin decondensation upon tethering YFP–LacI–Orc5(100–435aa) to the heterochromatic locus of CLTon cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm. (Fb) Area of heterochromatic loci upon tethering YFP–LacI, YFP–LacI–Orc5(100-435aa) and full-length YFP–LacI–Orc5. Error bars represent s.d., n=3 independent experiments (areas of 25–50 loci measured upon tethering each construct). ****P<0.0001 (Student's t-test). Lines represent the median, and the boxes represent the 25–75th percentiles. The YFP–LacI and YFP–LacI–Orc5 data from Fig. 1Bc has been replotted here for ease of comparison with the YFP–LacI–Orc5(100–435aa) mutant.

Fig. 3.

Orc5 interacts with GCN5. (A) Semi-endogenous immunoprecipitation (IP) in U2OS cells expressing T7–Orc5 and analysis by immunoblotting for T7 and GCN5. (B) Immunoprecipitation in U2OS cells expressing T7–Orc5 and Flag–GCN5 by using an anti-Flag antibody and analysis by immunoblotting for T7 and Flag. (C) Tethering of YFP–LacI–GCN5 to the CLTon locus shows recruitment of Orc5 to the site. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 μm. (D) Schematic representation of various truncation mutants of Orc5 containing a T7-epitope on the N-terminus. The specific domains that can associate with GCN5 and Orc2 based on immunoprecipitation and immunoblotting analysis (E) are depicted as ‘+’. (E) Immunoprecipitation from U2OS cells expressing various T7-tagged Orc5 mutants and Flag–GCN5 using an antibody against T7 and immunoblotting for T7, Flag and Orc2. (Fa) Chromatin decondensation upon tethering YFP–LacI–Orc5(100–435aa) to the heterochromatic locus of CLTon cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm. (Fb) Area of heterochromatic loci upon tethering YFP–LacI, YFP–LacI–Orc5(100-435aa) and full-length YFP–LacI–Orc5. Error bars represent s.d., n=3 independent experiments (areas of 25–50 loci measured upon tethering each construct). ****P<0.0001 (Student's t-test). Lines represent the median, and the boxes represent the 25–75th percentiles. The YFP–LacI and YFP–LacI–Orc5 data from Fig. 1Bc has been replotted here for ease of comparison with the YFP–LacI–Orc5(100–435aa) mutant.

We next mapped the region within Orc5 that associates with GCN5. We generated several truncation mutants of Orc5 (Fig. 3D) and co-transfected each of these with Flag–GCN5. Immunoprecipitation with the antibody against T7 revealed that full-length Orc5 and the 100–435 fragment efficiently associated with GCN5 (Fig. 3E), suggesting that the first 100 amino acids within the N-terminus of Orc5 are dispensable for GCN5 binding. Orc2 was found to bind to full-length Orc5 and to fragments 100–435 and 300–435 (Fig. 3D,E). These results suggest that Orc2 and GCN5 could associate with Orc5 simultaneously and that the binding might not be mutually exclusive. Because the T7–Orc5 100–435 mutant could bind to GCN5, we tested the ability of this mutant to cause chromatin decondensation. We tethered it to the CLTon locus (Fig. 3Fa) and found that it could cause chromatin decondensation comparable to that of the full-length protein (Fig. 3Fb). It is interesting to note that the fragment of Orc5 comprising amino acids 300–435 did not interact with Flag–GCN5 (Fig. 3E), but did cause partial chromatin decondensation (Fig. 2B–D).

We next examined the status of various chromatin marks at the Orc5-tethered locus. Immunofluorescence analysis using an antibody against trimethylated histone 3 (H3) at residue K9 (H3K9me3) showed robust accumulation of this mark at the LacI-containing heterochromatic locus (Fig. S3A). However, upon tethering Orc5 to the locus, H3K9me3 was distinctly devoid at these sites (Fig. S3A). Because Orc5 associates with GCN5, we examined whether H3 acetylation accumulated in Orc5-tethered cells. We did not observe robust accumulation of acetylated H3 marks in the highly decondensed Orc5-tethered cells; however, the control cells were clearly devoid of this mark at the condensed CLTon locus (Fig. S3Ba–Bc).

We then evaluated whether GCN5 could itself cause chromatin decondensation in our CLTon assay. YFP–LacI–GCN5 was tethered to the locus, and the status of chromatin architecture at the CLTon locus was evaluated (Fig. 4Aa). GCN5 could also mediate chromatin decondensation; however, this decondensation looked visually different from that observed for Orc5. Tethering YFP–LacI–Orc5 caused considerable large-scale decondensation, whereas tethering YFP–LacI–GCN5 caused a smaller scale ‘puffy’ chromatin appearance, but still significant decondensation when compared to that of YFP–LacI control (Fig. 4Aa). Although the range of chromatin decondensation upon tethering YFP–LacI–Orc5 varied from 2 to 35 µm2, the extent was much smaller for YFP–LacI–GCN5, which showed decondensation in the range 3–9.5 µm2 (Fig. 4Ab).

Fig. 4.

Orc5 causes decondensation in a GCN5-dependent manner. (Aa) Chromatin decondensation upon tethering YFP–LacI, YFP–LacI–Orc5 and YFP–LacI–GCN5 to the heterochromatic locus of CLTon cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm. (Ab) Area of heterochromatic loci upon tethering YFP–LacI, YFP–LacI–Orc5 and YFP–LacI–GCN5. Error bars represent s.d., n=3 independent experiments (areas of 25–50 loci measured upon tethering each construct). ****P<0.0001 (Student's t-test). The YFP–LacI and YFP–LacI–Orc5 data in Fig. 1Bc has been replotted here for ease of comparison with that for YFP–LacI–GCN5. (B) Immunoblot showing efficient siRNA-mediated knockdown of GCN5. (C) The percentage of YFP–LacI–Orc5-tethered cells with open loci in control and GCN5-knockdown (GCN5 si) cells. Error bars represent s.d., n=3 independent experiments (100 loci examined upon tethering each construct). **P<0.01 (Student's t-test). (D) Extent of decondensation upon tethering YFP–LacI–Orc5 in control and GCN5-knockdown cells. Error bars represent s.d., n=3 independent experiments (100 loci examined upon tethering each construct). (E) Area of heterochromatic loci upon tethering YFP–LacI and YFP–LacI–Orc5 (in control and GCN5-knockdown cells). Error bars represent s.d., n=3 independent experiments (area of 25–50 loci measured upon tethering each construct). ****P<0.0001 (Student's t-test). Lines represent the median, and the boxes represent the 25–75th percentiles. The YFP–LacI and YFP–LacI–Orc5 data in Fig. 1Bc has been replotted here for ease of comparison with the decondensation caused by YFP–LacI–Orc5 upon GCN5 knockdown.

Fig. 4.

Orc5 causes decondensation in a GCN5-dependent manner. (Aa) Chromatin decondensation upon tethering YFP–LacI, YFP–LacI–Orc5 and YFP–LacI–GCN5 to the heterochromatic locus of CLTon cells. Insets represent 200% magnifications of the boxed regions. Scale bar: 10 µm. (Ab) Area of heterochromatic loci upon tethering YFP–LacI, YFP–LacI–Orc5 and YFP–LacI–GCN5. Error bars represent s.d., n=3 independent experiments (areas of 25–50 loci measured upon tethering each construct). ****P<0.0001 (Student's t-test). The YFP–LacI and YFP–LacI–Orc5 data in Fig. 1Bc has been replotted here for ease of comparison with that for YFP–LacI–GCN5. (B) Immunoblot showing efficient siRNA-mediated knockdown of GCN5. (C) The percentage of YFP–LacI–Orc5-tethered cells with open loci in control and GCN5-knockdown (GCN5 si) cells. Error bars represent s.d., n=3 independent experiments (100 loci examined upon tethering each construct). **P<0.01 (Student's t-test). (D) Extent of decondensation upon tethering YFP–LacI–Orc5 in control and GCN5-knockdown cells. Error bars represent s.d., n=3 independent experiments (100 loci examined upon tethering each construct). (E) Area of heterochromatic loci upon tethering YFP–LacI and YFP–LacI–Orc5 (in control and GCN5-knockdown cells). Error bars represent s.d., n=3 independent experiments (area of 25–50 loci measured upon tethering each construct). ****P<0.0001 (Student's t-test). Lines represent the median, and the boxes represent the 25–75th percentiles. The YFP–LacI and YFP–LacI–Orc5 data in Fig. 1Bc has been replotted here for ease of comparison with the decondensation caused by YFP–LacI–Orc5 upon GCN5 knockdown.

Orc5-mediated chromatin decondensation is GCN5-dependent

To gain an insight into the functional relevance of the Orc5 and GCN5 interaction, we evaluated whether GCN5 is required for the Orc5-mediated chromatin decondensation. We depleted GCN5 using small interfering (si)RNA in CLTon cells (Fig. 4B) and examined whether tethering of Orc5 could still induce chromatin decondensation. Remarkably, in GCN5-depleted cells, we observed a significant decrease (32% decrease, **P<0.01, Student's t-test) in the extent of Orc5-mediated chromatin decondensation (Fig. 4C–E). In control cells, 84.3% of the cells showed chromatin decondensation upon tethering Orc5. By contrast, this number reduced to 52% upon loss of GCN5 (Fig. 4C). To better understand this reduction in Orc5-mediated chromatin decondensation, we scored CLTon cells based on medium or large decondensation of the heterochromatic locus. Upon loss of GCN5, there was a striking reduction in the percentage of cells showing large decondensation (from 44% in control cells to 16% in GCN5-knockdown cells) (Fig. 4D). This result was corroborated by examining the area of the loci in control and GCN5-depleted cells. The area of the CLTon locus varied from 2 to 35 µm2 in control cells and 1 to 7 µm2 in GCN5-depleted cells (Fig. 4E).

Orc5-mediated chromatin decondensation occurs predominantly during G1 phase

We next determined whether the Orc5-mediated chromatin decondensation occurred throughout the cell cycle or was restricted to a specific stage of the cell cycle. To investigate this, we tethered YFP–LacI–Orc5 to the CLTon locus and performed double fluorescence labeling of MCM (through labeling of MCM3) and PCNA. MCM and PCNA show distinct temporal patterns of chromatin loading (Prasanth et al., 2004a). MCM, the replication helicase, loads onto chromatin in G1, shows spatio-temporal patterns during S phase and is gradually lost by the end of S phase, whereas PCNA, the clamp, is associated with chromatin only during S phase. Based on the patterns of MCM and PCNA loading, it is possible to discern whether the cell is in G1 or S phase of the cell cycle (G1 cells are MCM+ and PCNA−, whereas S-phase cells are PCNA+). We observed that the YFP–LacI–Orc5-expressing cells with the most dramatic chromatin decondensation at the CLTon locus were positive for MCM staining (Fig. 5Aa). At the same time, YFP–LacI–Orc5-tethered cells in S phase (PCNA+ cells) showed a lower degree of chromatin decondensation (Fig. 5Ab, inset 3). The most dramatic decondensation of the CLTon loci happened in MCM+ PCNA− cells (Fig. 5Ab, inset1 and 2, and Fig. 5Ac). These results suggest that the Orc5-mediated decondensation occurs predominantly during G1 phase of the cell cycle.

Fig. 5.

Orc5 causes maximal chromatin decondensation in G1 and early replication upon tethering to heterochromatic locus. (Aa) Immunofluorescent labeling of MCM3 upon tethering CFP–LacI–Orc5 to heterochromatic loci in CLTon cells. Scale bar: 10 µm. (Ab) Immunofluorescent labeling of PCNA and MCM3 upon tethering CFP–LacI–Orc5 to heterochromatic loci in CLTon cells. Scale bar: 10 µm. Insets represent 200% magnifications of the boxed regions. (Ac) Percentage of CLTon loci showing dramatic decondensation upon tethering YFP–LacI–Orc5 in cells that are PCNA+ or MCM3+. Error bars represent s.d., n=3 independent experiments (50 PCNA+ and 50–100 MCM3+ cells used for analysis). (Ba) PCNA immunofluorescence upon tethering YFP–LacI to CLTon loci. Scale bar: 10 µm. Note the accumulation of PCNA at the locus in late S-phase. (Bb) Percentage of YFP–LacI or YFP–LacI–Orc5-tethered CLTon loci replicating in early and late S-phase. Error bars represent s.d., n=3 independent experiments (100 cells used for analysis).

Fig. 5.

Orc5 causes maximal chromatin decondensation in G1 and early replication upon tethering to heterochromatic locus. (Aa) Immunofluorescent labeling of MCM3 upon tethering CFP–LacI–Orc5 to heterochromatic loci in CLTon cells. Scale bar: 10 µm. (Ab) Immunofluorescent labeling of PCNA and MCM3 upon tethering CFP–LacI–Orc5 to heterochromatic loci in CLTon cells. Scale bar: 10 µm. Insets represent 200% magnifications of the boxed regions. (Ac) Percentage of CLTon loci showing dramatic decondensation upon tethering YFP–LacI–Orc5 in cells that are PCNA+ or MCM3+. Error bars represent s.d., n=3 independent experiments (50 PCNA+ and 50–100 MCM3+ cells used for analysis). (Ba) PCNA immunofluorescence upon tethering YFP–LacI to CLTon loci. Scale bar: 10 µm. Note the accumulation of PCNA at the locus in late S-phase. (Bb) Percentage of YFP–LacI or YFP–LacI–Orc5-tethered CLTon loci replicating in early and late S-phase. Error bars represent s.d., n=3 independent experiments (100 cells used for analysis).

Because we observed a decrease in the extent of Orc5-mediated chromatin decondensation in cells that lacked GCN5, we examined if this was due to an effect of cell cycle perturbation. We evaluated the cell cycle profile of control and GCN5-depleted cells with flow cytometry (Fig. S2Aa,Ab) and observed a marginal increase in the G1 population of cells upon loss of GCN5. Because the Orc5-mediated chromatin decondensation is most dramatic in G1, the loss of the decondensation in the absence of GCN5 is not due to perturbation of cell cycle progression. We propose that GCN5 and Orc5 cooperate in mediating this chromatin decondensation.

Chromatin changes are related to the efficiency of DNA replication. We therefore profiled the replication timing of the locus upon tethering YFP–LacI–Orc5. To determine the stage of S phase during which the locus replicates, we performed immunofluorescence analysis of PCNA. PCNA shows distinct localization patterns during S phase, and based on the PCNA patterns, S phase can be categorized into early, mid or late stages (O'Keefe et al., 1992). We tethered YFP–LacI or YFP–LacI–Orc5 to the locus and examined the accumulation of PCNA at the locus (Fig. 5Ba,Bb, Fig. S4). The replication timing of the CLTon locus was discerned by examining the PCNA pattern of the cell nucleus. Upon tethering YFP–LacI, the locus did not show any accumulation of PCNA in early S-phase (Fig. 5Ba). In late S-phase, we could observe robust accumulation of PCNA at the locus, indicating that the YFP–LacI-labeled locus replicated in late S-phase (Fig. 5Ba,Bb). This wasn't surprising owing to the inherent heterochromatic nature of the locus. Next, we tethered YFP–LacI–Orc5 to the locus and examined the replication timing of the locus (Fig. 5Bb; Fig. S4). We could observe robust accumulation of PCNA at the locus preferentially in early S-phase and not in late S-phase, indicating that the replication timing of the locus shifted to early S-phase upon tethering of Orc5. Note that even the few dramatically decondensed Orc5-tethered loci that we observed during S phase replicated early (Fig. S4).

Loss of Orc5 causes a reduction in histone acetylation at origins

Histone acetylation is required for origin activation during S phase (Unnikrishnan et al., 2010), and it has also been shown to regulate the timing of replication origin firing (Vogelauer et al., 2002). We asked whether Orc5-mediated chromatin-opening facilitates histone acetylation in cooperation with GCN5 at the origins of replication. We conducted acetylated H3 chromatin immunoprecipitation (ChIP) in control and Orc5-depleted cells (Fig. 6A), and found a significant reduction in the acetylation of H3 at select origins – including those of lamin B1, MCM5 and lamin B2 – but not at distal sites (either upstream or downstream) of the specific origins (Fig. 6Ba–Bd) or other origins – including those of AKNA, TRPM1, MCM6 and MCM3 (data not shown). We next wanted to determine whether the reduced acetylation of origins causes inefficient origin firing and replication defects in S phase. Therefore, we depleted Orc5 and performed double immunofluorescence staining of MCM and PCNA to examine the cell cycle profile upon loss of Orc5 (Fig. 6C). We observed a 21% reduction in S-phase cells (PCNA+ cells) from 58 to 37%, and a concomitant increase in G1 (MCM+ PCNA− cells) from 28% to 46% upon loss of Orc5, indicating that cells accumulated in G1 in the absence of Orc5.

Fig. 6.

Orc5 regulates histone acetylation in a subset of origins. (A) Immunoblot showing efficient siRNA-mediated knockdown of Orc5 (Orc5 si). (B) Acetylated H3 ChIP analysis at lamin B1, MCM5 and lamin B2 gene origins and non-origins in U2OS cells after transfection with control siRNA or Orc5-knockdown. Error bars represent s.d., n=3. *P<0.1, **P<0.01 (Student's t-test). (C) Percentage of cells that were PCNA+, or MCM3+ and PCNA− in U2OS cells after transfection with control siRNA or Orc5-knockdown. Error bars represent s.d., n=3 independent experiments (1000 cells used for analysis). (D) Cartoon depicting Orc5-mediated chromatin opening. H3Ac, acetylated H3; S, S phase.

Fig. 6.

Orc5 regulates histone acetylation in a subset of origins. (A) Immunoblot showing efficient siRNA-mediated knockdown of Orc5 (Orc5 si). (B) Acetylated H3 ChIP analysis at lamin B1, MCM5 and lamin B2 gene origins and non-origins in U2OS cells after transfection with control siRNA or Orc5-knockdown. Error bars represent s.d., n=3. *P<0.1, **P<0.01 (Student's t-test). (C) Percentage of cells that were PCNA+, or MCM3+ and PCNA− in U2OS cells after transfection with control siRNA or Orc5-knockdown. Error bars represent s.d., n=3 independent experiments (1000 cells used for analysis). (D) Cartoon depicting Orc5-mediated chromatin opening. H3Ac, acetylated H3; S, S phase.

Our results suggest that Orc5 might make the chromatin more accessible for the establishment of preRCs during G1 at specific origins (Fig. 6D), and this in turn is important for efficient origin firing and DNA replication in S phase. Our results provide new insights into the specific role of one of the ORC subunits in facilitating chromatin opening for the establishment of preRCs.

Replication of DNA occurs once and only once per cell division cycle. The licensing of replication origins requires the sequential binding of the ORC, Cdc6 and Cdt1 that is then needed for the loading of the Mcm2-7 complex onto the chromatin (Bell and Dutta, 2002). ORC, comprising six subunits, serves as the landing pad for the assembly of this multiprotein complex at the origins of replication. The contribution of individual subunits in this process remains to be understood. We demonstrate that the Orc5 subunit is unique, and when tethered ectopically to a transgene array, induces large-scale chromatin decondensation. It associates with the HAT GCN5, which acetylates H3. The ability of Orc5 to cause chromatin decondensation requires GCN5 (Fig. 6D), and the loss of Orc5 causes decreased acetylation at specific origins. Our data suggest that there could be multiple mechanisms for the chromatin decondensation that is mediated by Orc5 during G1, a GCN5-dependent mechanism involving residues 100–200 of Orc5 and a GCN5-independent mechanism involving the last 35 amino acids of Orc5. Once the chromatin is decondensed during G1, it facilitates origin licensing, and it is likely that Orc5 and GCN5 cooperate at select origins and that Orc5 cooperates with other ‘unknown factors’ at other origins.

GCN5 is a global regulator of gene expression (Baker and Grant, 2007; Robert et al., 2004). More recently, GCN5 has been found to associate with yeast origins, albeit weakly (Espinosa et al., 2010), and positively regulates DNA replication by counteracting the inhibitory effects of HDACs. In addition, GCN5-mediated acetylation of H3 lysine residues has also been proposed to function in replication-coupled nucleosome assembly (Burgess et al., 2010). We propose that Orc5 at origins helps the recruitment of GCN5 to these sites and that this in turn facilitates the opening of chromatin, thus enabling the loading of other preRC components.

The acetylation of histones H3 and H4 is known to be dynamically regulated around the origins of replication that facilitate origin firing (Unnikrishnan et al., 2010). Furthermore, acetylation of H4 at residue K79 is enriched at origins; acetylation of H4 at residue K16 is enriched at early-firing origins and also limits the spread of heterochromatin at origins. Acetylation of H3 at residues K9 or K14, and of H4 at residues K5, K8 or K12 is enriched at active origins, and these modifications are believed to promote firing (Dorn and Cook, 2011).

Hbo1 is an H4-specific HAT that interacts with the human ORC and MCM proteins (Iizuka and Stillman, 1999). It is required for replication licensing (Iizuka et al., 2006) and is known to associate with replication origins (Miotto and Struhl, 2008). It has recently been demonstrated that Hbo1-mediated hyperacetylation of H4 increases Mcm2-7 loading onto chromatin (Miotto and Struhl, 2010). It is generally believed that histone acetylation loosens up the compacted chromatin, thereby increasing the accessibility for loading of the MCM helicase complex, or acts as a molecular tag to which the helicase is tethered (Chadha and Blow, 2010). Hbo1 association to origins is dependent on Cdt1, and it has been shown that Cdt1 can modulate chromatin accessibility through temporal recruitment of Hbo1 to origins (Miotto and Struhl, 2008). Interestingly, ectopic tethering of Cdt1 is also known to induce large-scale chromatin unfolding at a transgene array (Wong et al., 2010). It is well established that the origins comprise open chromatin during G1 and then become less accessible as cells exit out of G1 phase (Djeliova et al., 2002; Pemov et al., 1998). The replication origins are known to exhibit temporal dynamics in chromatin structure, with a highly open structure during G1 and more closed architecture during S phase. Elegant work, using a quantitative-PCR-based approach on DNase-1-treated chromatin samples, has shown that endogenous replication origins, including those of MCM4 and lamin, display a more open chromatin structure during G1 than in S phase (Wong et al., 2010).

In addition, tethering of replication protein Cdc45 leads to chromatin decondensation in a Cdk2-dependent manner (Alexandrow and Hamlin, 2005). Such decondensation is mediated by phosphorylation of H1. The phosphorylation of H1, in turn, is Cdk2-dependent and leads to chromatin decondensation in S phase. These observations point towards a role for Cdc45 in replication fork progression by regulating large-scale chromatin changes (Alexandrow and Hamlin, 2005). Our observations of the CLTon locus – which reveal that Orc5 can efficiently cause chromatin decondensation, predominantly during the G1 phase of the cell cycle – provide a strong indication that similar events occur at endogenous origins, whereby Orc5 might cause local chromatin changes in collaboration with GCN5. Another piece of evidence to support this idea comes from a recent study that shows preferential association of Orc5 with acetylated H3 at K27 (Ji et al., 2015) as compared to several other markers of activation, such as trimethylated H3 at K4, dimethylated H3 at K79 and trimethylated H3 at K36. We propose that at an endogenous locus at which ORC binds, the amount of Orc5 present would be sufficient for a smaller scale of decondensation, which might not be observable at the resolution of light microscopy. But that decondensation would be sufficient for the cognate cellular function, such as replication initiation.

Orc5 is one of the ORC subunits, and ATP binding to Orc5 is involved in efficient ORC formation (Siddiqui and Stillman, 2007; Takahashi et al., 2004). Orc5 has also been implicated in silencing at the HML and MHR loci in yeast and in heterochromatin organization in human cells (Dillin and Rine, 1997; Prasanth et al., 2010). The gene encoding Orc5 maps to chromosome 7q22 and is frequently deleted in adult acute myeloid leukemia, myelodysplastic syndrome, uterine leiomyomas and malignant myeloid diseases (Fröhling et al., 2001; Quintana et al., 1998). We have observed dose-dependent chromatin decondensation at the CLTon locus, and this was directly correlated with the expression level of Orc5. This finding is supported by the fact that tethering of Orc5 to the CLTon locus can cause dramatic chromatin decondensation. However, tethering of other ORC subunits does not, despite the fact that Orc5 can be recruited to the locus by other ORC subunits. Our data imply that excessive levels of Orc5 in the cell could result in aberrant chromatin decondensation and cause genomic instability.

pEGFP-LacI vector was a kind gift from Dr Miroslav Dundr (Rosalind Franklin University of Medicine and Science, North Chicago, IL) (Kaiser et al., 2008) and used to generate the pEYFP-LacI vector. YFP-LacI-Orc1, -Orc2, -Orc3, -Orc4, -Orc5 and -Orc6 were cloned by amplifying and inserting the ORC proteins into the pEYFP-LacI vector. YFP–LacI–Orc5 D414A,D426A,D433A; YFP–LacI–Orc5 K43A; YFP–LacI–Orc5 R166A were generated using site-directed mutagenesis (Stratagene) with YFP–LacI–Orc5 wild type as template. T7–Orc5 full length and truncations were cloned by amplifying and inserted into pCGT vector. Flag–GCN5 was a kind gift from Brian Freeman (University of Illinois at Urbana–Champaign, Champaign, IL).

Cell culture

U2OS osteosarcoma cells were grown in high glucose Dulbecco's modified Eagle's medium (DMEM) supplemented with 10% fetal bovine serum (FBS; Hyclone). U2OS-2-6-3 CLTon cells were cultured in DMEM supplemented with 10% Tet-system-approved FBS (Clonetech).

Immunofluorescence and fluorescent protein visualization

For visualizing YFP–LacI-tagged proteins, cells were fixed with 2% formaldehyde in PBS (pH 7.4) for 15 min at room temperature, followed by permeabilization with 0.5% Triton X-100 in PBS for 7 min on ice and by blocking. For immunofluorescent labeling of H3K9me3, acetylated H3, MCM3 and PCNA, cells were pre-extracted before fixing with 0.5% Triton X-100 in cytoskeletal buffer (CSK; 100 mM NaCl, 300 mM sucrose, 3 mM MgCl2, 10 mM PIPES at pH 6.8) for 5 min on ice followed by fixing with 1% formaldehyde in PBS (pH 7.4) for 5 min at room temperature and then blocking. This was followed by blocking for 30 min with 1% normal goat serum in PBS. After that, cells were incubated for 1 h with primary antibody in a humidified chamber, followed by incubation with secondary antibody for 25 min. Nuclei were then stained with DAPI and mounted using Vectashield (Vector Laboratories). The following antibodies were used for immunofluorescence: anti-H3K9me3 (1:200, Millipore 07-523), anti-acetylated-H3 (1:500, Millipore 06-599), anti-MCM3 (1:300, 738 rabbit polyclonal antibody) and anti-PCNA mAbPC10 (1:150, PC10 monoclonal antibody).

Microscopy

For observing cells and obtaining statistics (area of decondensation, and PCNA and MCM3 staining), we used a Zeiss Axioimager z1 fluorescence microscope (Carl Zeiss Inc.) equipped with chroma filters (Chroma Technology). The images were acquired using a 60×, 1.4NA oil immersion objective. Digital imaging was performed using an Hamamatsu ORCA cooled CCD camera. Axiovision software (Zeiss) was used for processing the images.

A delta vision optical sectioning deconvolution instrument (Applied precision) on an Olympus microscope was also used for observing cells and for obtaining the images used in the manuscript. The images were acquired using a 60×, 1.4NA oil immersion objective. softWoRx imaging workstation was used to calculate the area of the decondensed locus. The ‘Edit Polygon’ tool was used as it is suitable for calculating irregular-area statistics. The extent of overlap of H3K9me3 with YFP–LacI or YFP–LacI–Orc5 was measured by using the ‘Colocalization’ tool and defining a region of interest (ROI); the Pearson coefficient of colocalization was then determined. Line profiles for H3K9me3 and acetylated H3 immunofluorescence were generated by using the ‘Line Profile’ tool with the band value set to 2.

siRNA-mediated depletion of Orc5

U2OS cells were grown to 30% confluence, followed by two rounds of knockdown with a control siRNA against the luciferase gene (Shen et al., 2010) or with an siRNA against Orc5 (5ʹ-UGACUUUGUUCGCUUGUUUUU-3ʹ) (Prasanth et al., 2010), 24 h apart at a final concentration of 100 nM using Lipofectamine RNAiMax (Invitrogen). This was followed by collection of cells for ChIP analysis 24 h after the second knockdown.

For GCN5 knockdown, CLTon cells were grown to 30% confluence on coverslips followed by two rounds of knockdown with a control siRNA against the luciferase gene (Shen et al., 2010) or with an siRNA against GCN5 (Palhan et al., 2005), 24 h apart at a final concentration of 40 nM. YFP–LacI–Orc5 was transfected into cells while carrying out the second round of knockdown, and the cells were fixed for microscopy analysis 24 h later. The first round of knockdown was performed using Lipofectamine RNAimax (Invitrogen), and the second round of knockdown with transfection of the YFP-tagged constructs was performed using Lipofectamine 2000 (Invitrogen).

Immunoprecipitations and immunoblots

For co-immunoprecipitation experiments, co-transfections were performed in U2OS cells. Flag–HATs and T7–Orc5 were transiently transfected, and cells were lysed, 24 h post-transfection, in immunoprecipitation buffer (50 mM HEPES pH 7.9, 10% Glycerol, 200 mM NaCl, 0.1% Triton X-100, 1 mM CaCl2) supplemented with the protease and phosphatase inhibitors. 4U of MNase (Sigma) was then added per 10-cm plate of sample, followed by rocking at room temperature for 20 min. EDTA (final concentration 5 mM) was added to stop the reaction, and samples were centrifuged at 12,500 rpm (Eppendorf centrifuge 5424), for 5 min at 4°C. The supernatant was pre-cleared. Pre-clearing was performed using Gammabind Sepharose beads for 1 h, and the lysates were incubated with appropriate antibody overnight. For pulldown of the antibody-bound complexes, agarose beads were washed in the same immunoprecipitation buffer and were incubated with lysate containing antibodies for 2 h. This was followed by three washes of the pulled-down complexes, and finally denaturation of the pulled-down proteins by the addition of Laemmli buffer and incubation in a heat block for 10 min. The complexes were then analyzed by western blotting.

For semi-endogenous immunoprecipitation, the cells were initially lysed in IP-100 buffer (50 mM Tris pH 7.5, 100 mM NaCl, 1 mM MgCl2, 0.2% NP-40, 1% glycerol supplemented with protease and phosphatase inhibitors). The lysate was then sonicated for 2 min with Bioruptor Power-up (Diagenode) using 30-s-on and 30-s-off cycles. Benzonase (250 U/107 cells) was then added to the lysate, and the sample was incubated at room temperature for 30 min with rotation. The reaction was stopped by adding EDTA (final concentration 5 mM), and the salt concentration was raised to 200 mM with NaCl. The sample was then pre-cleared, and subsequent steps of immunoprecipitation were performed as explained in the previous paragraph. All subsequent washes were performed with IP-200 buffer (50 mM Tris pH 7.5, 200 mM NaCl, 1 mM MgCl2, 0.2% NP-40, 1% glycerol supplemented with protease and phosphatase inhibitors).

For immunoprecipitations and immunoblotting, the following antibodies were used: anti-Flag M2 (1:500, Sigma), anti-T7 (1:5000, Novagen), anti-ORC2 polyclonal antibody 205-6 (1:1000) and anti-GCN5 (1:200, Cell Signaling).

Chromatin immunoprecipitation

Control and Orc5-knockdown cells were crosslinked for 10 min at room temperature through addition of formaldehyde (Sigma) to culture medium to a final concentration of 1%. The reaction was stopped by addition of glycine at a final concentration of 0.125 M. Fixed cells were then washed twice (quickly) in chilled PBS and harvested with PBS. Chromatin was prepared using two subsequent extraction steps (10 min at 4°C) with Buffer 1 (50 mM HEPES with KOH pH 7.5, 140 mM NaCl, 1 mM EDTA, 10% glycerol, 0.5% NP-40, 0.25% Triton X-100) and Buffer 2 (200 mM NaCl, 1 mM EDTA, 0.5 mM EGTA, 10 mM Tris-HCl pH 8). Nuclei were then pelleted by centrifugation, resuspended in SDS lysis buffer (50 mM Tris-HCl pH 8, 0.1% SDS, 1% NP-40, 0.1% sodium deoxycholate, 10 mM EDTA, 150 mM NaCl) and subjected to sonication with a Bioruptor Power-up instrument (Diagenode) for 45 min (three times of 15 min of sonication, each cycle comprising 30-s on and 30-s off). This yielded genomic DNA fragments that were 300 bp in length. The obtained chromatin was then precleared with Protein A/G ultralink beads (53133, Pierce) for 1 h at 4°C, and immunoprecipitation with the anti-acetylated-H3 antibody and rabbit IgG was performed overnight at 4°C. Immune complexes were pulled down by adding pre-blocked protein A/G ultralink beads, and the mixture was incubated for 2 h at room temperature. Beads with immune complexes bound to them were washed once with low-salt buffer (0.1% SDS, 1% Triton X-100, 2 mM EDTA, 20 mM Tris-HCl pH 8, 150 mM NaCl), once with high-salt buffer (0.1% SDS, 1% Triton X-100, 2 mM EDTA, 20 mM Tris-HCl pH 8, 500 mM NaCl), once with LiCl wash buffer (10 mM Tris-HCl pH 8.0, 1% sodium deoxycholate, 1% NP-40, 250 mM LiCl, 1 mM EDTA) and twice with Tris-EDTA. Beads were then eluted twice (10 min each) in Tris-EDTA+1% SDS+0.1% NaHCO3 at 65°C, and the cross-links were then reversed overnight at 65°C. DNA from the eluted material was then treated with RNase (10 µg/ml) for 1 h at 37°C followed by 2 h of treatment with proteinase K (4 μl 0.5 M EDTA, 8 μl 1 M Tris-HCl pH 6.9, 1 μl proteinase K 20 mg/ml) at 42°C. DNA was isolated by using the QIAquick PCR purification kit (Qiagen) resuspended in elution buffer, and quantitative PCR was performed using PowerSYBR Green PCR Master mix (Applied Biosystems) and analyzed on a 7300 PCR System (Applied Biosystems). ChIP–quantitative-PCR results were analyzed and plotted as percentages (%) of immunoprecipitation/input signal (% input).

The primer sequences of the regions analyzed were as follows: lamin B1 origin +4.2 Kb, forward 5ʹ-CCTCCCAAAGTGCTAGGATTAC-3ʹ, reverse 5ʹ-CCTAAGGCCCAACCAAGAAT-3ʹ; lamin B1 origin forward 5ʹ-AAGATTCTACGCCTGCTTTAGG-3ʹ, reverse 5ʹ-AGCTAACCAGCTAGCAAACC-3ʹ; MCM5 origin −4 Kb forward 5ʹ-TAACCTCTCTGAGGCTCCAT-3ʹ, reverse 5ʹ-GTGTATGGCGTTGGCATTTC-3ʹ; MCM5 origin forward 5ʹ-CTAGGTGATTGTACGACCTGTG-3ʹ, reverse 5ʹ-CCTGCAACCTCATGTTCTCT-3ʹ; lamin B2 origin −2.2 Kb forward 5ʹ-GAGAATCACTTGAACCCAGGAG-3ʹ, reverse 5ʹ-CCACTAGCTGCCAAGCTTAT-3ʹ; lamin B2 origin forward 5ʹ-TGAGGAGTCCCTCAGATCTTTA-3ʹ, reverse 5ʹ-CAGAATCCGATCATGCACCT-3ʹ; MCM4−5 Kb forward 5ʹ-TTCACATCCACCCAGCTTATC-3ʹ, reverse 5ʹ-AGAGCATTCTTCCCCTGATG-3ʹ; MCM4 origin forward 5ʹ-TTGGGTGGCTACTTGGTGTT-3ʹ, reverse 5ʹ-TAGGCCCCTCGCTTGTTT-3ʹ.

We thank members of the Prasanth laboratory for discussions and suggestions. We thank Drs B. Freeman, C. Mizzen (phosphorylated histone H1 antibodies, University of Illinois at Urbana–Champaign, Champaign, IL), M. Dundr, D. Spector (Cold Spring Harbor Laboratory, Cold Spring Harbor, NY) and B. Stillman (Cold Spring Harbor Laboratory, Cold Spring Harbor, NY) for providing reagents and suggestions. We would like to thank Dr D. Rivier, K. Dovalovsky and A. Matur for critical reading of the paper.

Author contributions

S.G. and A.C. designed and performed the experiments. K.M.S. made the initial observation. S.G., K.V.P. and S.G.P. designed the experiments and wrote the manuscript

Funding

This work was supported by NSF-CMMB-IGERT and F31 [grant number CA180616]; a National Institutes of Health (NIH) fellowship to S.G.; American Cancer Society [grant number RSG 11-174-01RMC]; NIH [grant number 1RO1GM088252] award to K.V.P.; and National Science Foundation career [grant number 1243372] and NIH [grant number 1RO1GM099669] awards to S.G.P. Deposited in PMC for release after 12 months.

Aggarwal
,
B. D.
and
Calvi
,
B. R.
(
2004
).
Chromatin regulates origin activity in Drosophila follicle cells
.
Nature
430
,
372
-
376
.
Alexandrow
,
M. G.
and
Hamlin
,
J. L.
(
2005
).
Chromatin decondensation in S-phase involves recruitment of Cdk2 by Cdc45 and histone H1 phosphorylation
.
J. Cell Biol.
168
,
875
-
886
.
Baker
,
S. P.
and
Grant
,
P. A.
(
2007
).
The SAGA continues: expanding the cellular role of a transcriptional co-activator complex
.
Oncogene
26
,
5329
-
5340
.
Bell
,
S. P.
and
Dutta
,
A.
(
2002
).
DNA replication in eukaryotic cells
.
Annu. Rev. Biochem.
71
,
333
-
374
.
Bell
,
S. P.
and
Stillman
,
B.
(
1992
).
ATP-dependent recognition of eukaryotic origins of DNA replication by a multiprotein complex
.
Nature
357
,
128
-
134
.
Bernal
,
J. A.
and
Venkitaraman
,
A. R.
(
2011
).
A vertebrate N-end rule degron reveals that Orc6 is required in mitosis for daughter cell abscission
.
J. Cell Biol.
192
,
969
-
978
.
Brown
,
J. A.
,
Holmes
,
S. G.
and
Smith
,
M. M.
(
1991
).
The chromatin structure of Saccharomyces cerevisiae autonomously replicating sequences changes during the cell division cycle
.
Mol. Cell. Biol.
11
,
5301
-
5311
.
Burgess
,
R. J.
,
Zhou
,
H.
,
Han
,
J.
and
Zhang
,
Z.
(
2010
).
A role for Gcn5 in replication-coupled nucleosome assembly
.
Mol. Cell
37
,
469
-
480
.
Carpenter
,
A. E.
,
Memedula
,
S.
,
Plutz
,
M. J.
and
Belmont
,
A. S.
(
2005
).
Common effects of acidic activators on large-scale chromatin structure and transcription
.
Mol. Cell. Biol.
25
,
958
-
968
.
Chadha
,
G. S.
and
Blow
,
J. J.
(
2010
).
Histone acetylation by HBO1 tightens replication licensing
.
Mol. Cell
37
,
5
-
6
.
Dhar
,
S. K.
,
Delmolino
,
L.
and
Dutta
,
A.
(
2001
).
Architecture of the human origin recognition complex
.
J. Biol. Chem.
276
,
29067
-
29071
.
Dillin
,
A.
and
Rine
,
J.
(
1997
).
Separable functions of ORC5 in replication initiation and silencing in Saccharomyces cerevisiae
.
Genetics
147
,
1053
-
1062
.
Djeliova
,
V.
,
Russev
,
G.
and
Anachkova
,
B.
(
2002
).
DNase I sensitive site in the core region of the human beta-globin origin of replication
.
J. Cell Biochem.
87
,
279
-
283
.
Dorn
,
E. S.
and
Cook
,
J. G.
(
2011
).
Nucleosomes in the neighborhood: new roles for chromatin modifications in replication origin control
.
Epigenetics
6
,
552
-
559
.
Espinosa
,
M. C.
,
Rehman
,
M. A.
,
Chisamore-Robert
,
P.
,
Jeffery
,
D.
and
Yankulov
,
K.
(
2010
).
GCN5 is a positive regulator of origins of DNA replication in Saccharomyces cerevisiae
.
PLoS ONE
5
,
e8964
.
Ferguson
,
B. M.
and
Fangman
,
W. L.
(
1992
).
A position effect on the time of replication origin activation in yeast
.
Cell
68
,
333
-
339
.
Fröhling
,
S.
,
Nakabayashi
,
K.
,
Scherer
,
S. W.
,
Dohner
,
H.
and
Dohner
,
K.
(
2001
).
Mutation analysis of the origin recognition complex subunit 5 (ORC5L) gene in adult patients with myeloid leukemias exhibiting deletions of chromosome band 7q22
.
Hum. Genet.
108
,
304
-
309
.
Giri
,
S.
,
Aggarwal
,
V.
,
Pontis
,
J.
,
Shen
,
Z.
,
Chakraborty
,
A.
,
Khan
,
A.
,
Mizzen
,
C.
,
Prasanth
,
K. V.
,
Ait-Si-Ali
,
S.
,
Ha
,
T.
, et al. 
(
2015
).
The preRC protein ORCA organizes heterochromatin by assembling histone H3 lysine 9 methyltransferases on chromatin
.
eLife
4
,
332
.
Goren
,
A.
,
Tabib
,
A.
,
Hecht
,
M.
and
Cedar
,
H.
(
2008
).
DNA replication timing of the human beta-globin domain is controlled by histone modification at the origin
.
Genes Dev.
22
,
1319
-
1324
.
Groth
,
A.
,
Corpet
,
A.
,
Cook
,
A. J. L.
,
Roche
,
D.
,
Bartek
,
J.
,
Lukas
,
J.
and
Almouzni
,
G.
(
2007a
).
Regulation of replication fork progression through histone supply and demand
.
Science
318
,
1928
-
1931
.
Groth
,
A.
,
Rocha
,
W.
,
Verreault
,
A.
and
Almouzni
,
G.
(
2007b
).
Chromatin challenges during DNA replication and repair
.
Cell
128
,
721
-
733
.
Iizuka
,
M.
and
Stillman
,
B.
(
1999
).
Histone acetyltransferase HBO1 interacts with the ORC1 subunit of the human initiator protein
.
J. Biol. Chem.
274
,
23027
-
23034
.
Iizuka
,
M.
,
Matsui
,
T.
,
Takisawa
,
H.
and
Smith
,
M. M.
(
2006
).
Regulation of replication licensing by acetyltransferase Hbo1
.
Mol. Cell. Biol.
26
,
1098
-
1108
.
Janicki
,
S. M.
,
Tsukamoto
,
T.
,
Salghetti
,
S. E.
,
Tansey
,
W. P.
,
Sachidanandam
,
R.
,
Prasanth
,
K. V.
,
Ried
,
T.
,
Shav-Tal
,
Y.
,
Bertrand
,
E.
,
Singer
,
R. H.
, et al. 
(
2004
).
From silencing to gene expression: real-time analysis in single cells
.
Cell
116
,
683
-
698
.
Ji
,
X.
,
Dadon
,
D. B.
,
Abraham
,
B. J.
,
Lee
,
T. I.
,
Jaenisch
,
R.
,
Bradner
,
J. E.
and
Young
,
R. A.
(
2015
).
Chromatin proteomic profiling reveals novel proteins associated with histone-marked genomic regions
.
Proc. Natl. Acad. Sci. USA
112
,
3841
-
3846
.
Kaiser
,
T. E.
,
Intine
,
R. V.
and
Dundr
,
M.
(
2008
).
De novo formation of a subnuclear body
.
Science
322
,
1713
-
1717
.
Knott
,
S. R. V.
,
Viggiani
,
C. J.
and
Aparicio
,
O. M.
(
2009a
).
To promote and protect: coordinating DNA replication and transcription for genome stability
.
Epigenetics
4
,
362
-
365
.
Knott
,
S. R. V.
,
Viggiani
,
C. J.
,
Tavare
,
S.
and
Aparicio
,
O. M.
(
2009b
).
Genome-wide replication profiles indicate an expansive role for Rpd3L in regulating replication initiation timing or efficiency, and reveal genomic loci of Rpd3 function in Saccharomyces cerevisiae
.
Genes Dev.
23
,
1077
-
1090
.
Li
,
G.
,
Sudlow
,
G.
and
Belmont
,
A. S.
(
1998
).
Interphase cell cycle dynamics of a late-replicating, heterochromatic homogeneously staining region: precise choreography of condensation/decondensation and nuclear positioning
.
J. Cell Biol.
140
,
975
-
989
.
Liu
,
J.
,
McConnell
,
K.
,
Dixon
,
M.
and
Calvi
,
B. R.
(
2012
).
Analysis of model replication origins in Drosophila reveals new aspects of the chromatin landscape and its relationship to origin activity and the prereplicative complex
.
Mol. Biol. Cell
23
,
200
-
212
.
McConnell
,
K. H.
,
Dixon
,
M.
and
Calvi
,
B. R.
(
2012
).
The histone acetyltransferases CBP and Chameau integrate developmental and DNA replication programs in Drosophila ovarian follicle cells
.
Development
139
,
3880
-
3890
.
Méndez
,
J.
,
Zou-Yang
,
X. H.
,
Kim
,
S.-Y.
,
Hidaka
,
M.
,
Tansey
,
W. P.
and
Stillman
,
B.
(
2002
).
Human origin recognition complex large subunit is degraded by ubiquitin-mediated proteolysis after initiation of DNA replication
.
Mol. Cell
9
,
481
-
491
.
Miotto
,
B.
and
Struhl
,
K.
(
2008
).
HBO1 histone acetylase is a coactivator of the replication licensing factor Cdt1
.
Genes Dev.
22
,
2633
-
2638
.
Miotto
,
B.
and
Struhl
,
K.
(
2010
).
HBO1 histone acetylase activity is essential for DNA replication licensing and inhibited by Geminin
.
Mol. Cell
37
,
57
-
66
.
Narlikar
,
G. J.
,
Fan
,
H.-Y.
and
Kingston
,
R. E.
(
2002
).
Cooperation between complexes that regulate chromatin structure and transcription
.
Cell
108
,
475
-
487
.
O'Keefe
,
R. T.
,
Henderson
,
S. C.
and
Spector
,
D. L.
(
1992
).
Dynamic organization of DNA replication in mammalian cell nuclei: spatially and temporally defined replication of chromosome-specific alpha-satellite DNA sequences
.
J. Cell Biol.
116
,
1095
-
1110
.
Palhan
,
V. B.
,
Chen
,
S.
,
Peng
,
G.-H.
,
Tjernberg
,
A.
,
Gamper
,
A. M.
,
Fan
,
Y.
,
Chait
,
B. T.
,
La Spada
,
A. R.
and
Roeder
,
R. G.
(
2005
).
Polyglutamine-expanded ataxin-7 inhibits STAGA histone acetyltransferase activity to produce retinal degeneration
.
Proc. Natl. Acad. Sci. USA
102
,
8472
-
8477
.
Papior
,
P.
,
Arteaga-Salas
,
J. M.
,
Gunther
,
T.
,
Grundhoff
,
A.
and
Schepers
,
A.
(
2012
).
Open chromatin structures regulate the efficiencies of pre-RC formation and replication initiation in Epstein-Barr virus
.
J. Cell Biol.
198
,
509
-
528
.
Pemov
,
A.
,
Bavykin
,
S.
and
Hamlin
,
J. L.
(
1998
).
Attachment to the nuclear matrix mediates specific alterations in chromatin structure
.
Proc. Natl. Acad. Sci. USA
95
,
14757
-
14762
.
Pflumm
,
M. F.
and
Botchan
,
M. R.
(
2001
).
Orc mutants arrest in metaphase with abnormally condensed chromosomes
.
Development
128
,
1697
-
1707
.
Prasanth
,
S. G.
,
Prasanth
,
K. V.
and
Stillman
,
B.
(
2002
).
Orc6 involved in DNA replication, chromosome segregation, and cytokinesis
.
Science
297
,
1026
-
1031
.
Prasanth
,
S. G.
,
Mendez
,
J.
,
Prasanth
,
K. V.
and
Stillman
,
B.
(
2004a
).
Dynamics of pre-replication complex proteins during the cell division cycle
.
Philos. Trans. R. Soc. Lond. B Biol. Sci.
359
,
7
-
16
.
Prasanth
,
S. G.
,
Prasanth
,
K. V.
,
Siddiqui
,
K.
,
Spector
,
D. L.
and
Stillman
,
B.
(
2004b
).
Human Orc2 localizes to centrosomes, centromeres and heterochromatin during chromosome inheritance
.
EMBO J.
23
,
2651
-
2663
.
Prasanth
,
S. G.
,
Shen
,
Z.
,
Prasanth
,
K. V.
and
Stillman
,
B.
(
2010
).
Human origin recognition complex is essential for HP1 binding to chromatin and heterochromatin organization
.
Proc. Natl. Acad. Sci. USA
107
,
15093
-
15098
.
Quintana
,
D. G.
,
Thome
,
K. C.
,
Hou
,
Z.-h.
,
Ligon
,
A. H.
,
Morton
,
C. C.
and
Dutta
,
A.
(
1998
).
ORC5L, a new member of the human origin recognition complex, is deleted in uterine leiomyomas and malignant myeloid diseases
.
J. Biol. Chem.
273
,
27137
-
27145
.
Ranjan
,
A.
and
Gossen
,
M.
(
2006
).
A structural role for ATP in the formation and stability of the human origin recognition complex
.
Proc. Natl. Acad. Sci. USA
103
,
4864
-
4869
.
Robert
,
F.
,
Pokholok
,
D. K.
,
Hannett
,
N. M.
,
Rinaldi
,
N. J.
,
Chandy
,
M.
,
Rolfe
,
A.
,
Workman
,
J. L.
,
Gifford
,
D. K.
and
Young
,
R. A.
(
2004
).
Global position and recruitment of HATs and HDACs in the yeast genome
.
Mol. Cell
16
,
199
-
209
.
Sasaki
,
T.
and
Gilbert
,
D. M.
(
2007
).
The many faces of the origin recognition complex
.
Curr. Opin. Cell Biol.
19
,
337
-
343
.
Shen
,
Z.
,
Sathyan
,
K. M.
,
Geng
,
Y.
,
Zheng
,
R.
,
Chakraborty
,
A.
,
Freeman
,
B.
,
Wang
,
F.
,
Prasanth
,
K. V.
and
Prasanth
,
S. G.
(
2010
).
A WD-repeat protein stabilizes ORC binding to chromatin
.
Mol. Cell
40
,
99
-
111
.
Siddiqui
,
K.
and
Stillman
,
B.
(
2007
).
ATP-dependent assembly of the human origin recognition complex
.
J. Biol. Chem.
282
,
32370
-
32383
.
Simpson
,
R. T.
(
1990
).
Nucleosome positioning can affect the function of a cis-acting DMA element in vivo
.
Nature
343
,
387
-
389
.
Stevenson
,
J. B.
and
Gottschling
,
D. E.
(
1999
).
Telomeric chromatin modulates replication timing near chromosome ends
.
Genes Dev.
13
,
146
-
151
.
Suter
,
B.
,
Pogoutse
,
O.
,
Guo
,
X.
,
Krogan
,
N.
,
Lewis
,
P.
,
Greenblatt
,
J. F.
,
Rine
,
J.
and
Emili
,
A.
(
2007
).
Association with the origin recognition complex suggests a novel role for histone acetyltransferase Hat1p/Hat2p
.
BMC Biol.
5
,
38
.
Takahashi
,
N.
,
Yamaguchi
,
Y.
,
Yamairi
,
F.
,
Makise
,
M.
,
Takenaka
,
H.
,
Tsuchiya
,
T.
and
Mizushima
,
T.
(
2004
).
Analysis on origin recognition complex containing Orc5p with defective Walker A motif
.
J. Biol. Chem.
279
,
8469
-
8477
.
Unnikrishnan
,
A.
,
Gafken
,
P. R.
and
Tsukiyama
,
T.
(
2010
).
Dynamic changes in histone acetylation regulate origins of DNA replication
.
Nat. Struct. Mol. Biol.
17
,
430
-
437
.
Vashee
,
S.
,
Simancek
,
P.
,
Challberg
,
M. D.
and
Kelly
,
T. J.
(
2001
).
Assembly of the human origin recognition complex
.
J. Biol. Chem.
276
,
26666
-
26673
.
Vogelauer
,
M.
,
Rubbi
,
L.
,
Lucas
,
I.
,
Brewer
,
B. J.
and
Grunstein
,
M.
(
2002
).
Histone acetylation regulates the time of replication origin firing
.
Mol. Cell
10
,
1223
-
1233
.
Wong
,
P. G.
,
Glozak
,
M. A.
,
Cao
,
T. V.
,
Vaziri
,
C.
,
Seto
,
E.
and
Alexandrow
,
M.
(
2010
).
Chromatin unfolding by Cdt1 regulates MCM loading via opposing functions of HBO1 and HDAC11-geminin
.
Cell Cycle
9
,
4351
-
4363
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information