SUMMARY
The arginine dependency of cancer cells creates metabolic vulnerability. In this study, we examine the impact of arginine availability on DNA replication and genotoxicity resistance. Using DNA combing assays, we find that limiting extracellular arginine results in the arrest of cancer cells at S phase and a slowing or stalling of DNA replication. The translation of new histone H4 is arginine dependent and influences DNA replication. Increased proliferating cell nuclear antigen (PCNA) occupancy and helicase-like transcription factor (HLTF)-catalyzed PCNA K63-linked polyubiquitination protect arginine-starved cells from DNA damage. Arginine-deprived cancer cells display tolerance to genotoxicity in a PCNA K63-linked polyubiquitination-dependent manner. Our findings highlight the crucial role of extracellular arginine in nutrient-regulated DNA replication and provide potential avenues for the development of cancer treatments.
Graphical abstract
In brief
Wang et al. find that arginine shortage can temporarily place DNA replication on hold via inhibiting histone H4 translation and stalling fork movement. They show that this is associated with HLTF-mediated K63-linked polyubiquitination of PCNA, which increases PCNA’s presence on nascent DNA strands, and helps cells tolerate DNA damage.
INTRODUCTION
A major hallmark of tumor progression is unchecked DNA replication and adaptation to the high metabolic demands of rapid cell division in a nutritionally deprived microenvironment.1,2 Arginine, a semi-essential amino acid, is a key nutrient for arginine auxotrophic tumors.3 These tumors become reliant on external sources of arginine due to widespread urea cycle dysregulation.4,5 External arginine shortages in the tumor core can result from poor vascularization and dysfunctional blood flow in the tumor microenvironment (TME).6–8 Furthermore, limiting dietary arginine or use of arginine-deprivation agents leads to the inhibition of arginine auxotroph tumor growth.5,9–12 Moreover, the spatiotemporal regulation of DNA replication during S phase of the cell cycle is synchronized with nutrient availability to cells.13–15 However, whether extracellular arginine shortage represents a widespread obstacle to DNA replication and how DNA replication contends with insufficient extracellular arginine availability remain largely unclear.
During DNA replication, nucleosomes are disrupted as the replication forks pass, followed by reassembly of recycled parental histones and incorporation of newly synthesized histones to restore chromatin state.16 As such, the synthesis of replication-dependent histones, essential building blocks for chromatin reassembly behind the replication fork, is tightly coupled to ongoing DNA synthesis.17 Indeed, histone biosynthesis is S-phase specific to govern S-phase timing and cell-cycle progression.18–20 Given that arginine is a proteinogenic amino acid, it is clearly important to understand the role of arginine depletion on replication-dependent histone biosynthesis.
DNA replication stress comprises a multitude of cellular conditions, including physical blockage of DNA replication fork progression along the template, deregulation of the replication initiation or elongation complexes, or deoxyribonucleotide depletion.21 Proliferating cell nuclear antigen (PCNA) plays a vital role in coordinating DNA replication activity and maintaining genome integrity.22,23 PCNA serves as a hub for proteins involved in DNA synthesis, cell-cycle control, and DNA damage response and repair.24–27 PCNA is also subjected to multiple post-translational modifications, including ubiquitination, contributing to the coordination of DNA replication and genotoxic damage tolerance.28,29 Whereas PCNA monoubiquitination in response to DNA damage promotes the bypass of replication-blocking lesions,30 PCNA polyubiquitination has been associated with fork reversal.31 In particular, the mammalian RAD5 ortholog helicase-like transcription factor (HLTF), a ubiquitin ligase for PCNA, has also been shown to be required for replication fork reversal.32,33 However, mechanisms by which arginine levels affect PCNA levels on replicating chromatin and the factors involved in this process remain elusive. In addition, it is currently unclear how arginine availability impacts DNA replication stress response and adaptation to genotoxic insult.
Here, we report that extracellular arginine shortage suppresses DNA replication elongation by reducing newly synthesized histone H4. Transient transfection of recombinant histone H4 protein can partially rescue DNA replication in arginine-starved cells. We present evidence that both HLTF and PCNA play key roles in fork protection in the absence of arginine. Arginine shortage increases HLTF-mediated PCNA K63-linked polyubiquitination and PCNA occupancy on the stalled nascent DNA strands. Next, we find that under replication stress induced by hydroxyurea (HU), arginine-starved cells exhibit vigorous nascent strand degradation, when HLTF is depleted, in a manner dependent on DNA2. Finally, HLTF depletion or treatment with PCNA ligand AOH996 results in a higher DNA damage signaling and sensitivity to genotoxicity under arginine shortage. Our results highlight a previously undescribed replication stress that arginine shortage induces defects on de novo chromatin assembly to interfere with DNA replication and S-phase progression. On the basis of these results, we conclude that extracellular arginine shortage triggers replication stress responses and confers tumor escape from a variety of genotoxicities.
RESULTS
Extracellular arginine regulates DNA replication fork progression
To determine the fate of cells that experience extracellular arginine shortage during cell-cycle progression, asynchronously cycling arginine auxotrophic MDA-MB-231 breast cancer cells were pulse labeled with bromodeoxyuridine (BrdU; t = 0 h), an analog of thymidine, and then chased in medium ± arginine for an additional 10 h. Using flow cytometry, we found that in the presence of arginine, most BrdU-labeled MDA-MB-231 cells were not at S phase (Figure 1A, bottom middle panel), while cells in arginine-free media were retained at S phase (Figure 1A, bottom right panel). Multiple cell lines, irrespective of arginine auxotrophy, displayed a similar impediment to S-phase exit in response to extracellular arginine shortage (Figure 1A, right panel). The overall cell-cycle distribution of G1, S, and G2/M were analyzed, and increased S-phase and decreased M-phase populations were found in arginine-deprived cells (Figure S1A). To determine if this stalled S-phase phenomenon occurs within the S phase, we synchronized MDA-MB-231 cells in early S phase using a double thymidine block. BrdU-labeled synchronized cells did not exit S phase when deprived of arginine (Figure 1B). Further, hallmarks of cell-cycle progression such as the upregulation of cyclin A, cyclin B1, phosphorylated H3S10, and H4 were absent in cells released in arginine-free medium, further confirming the inhibitory effect of arginine shortage on S-phase progression (Figure S1B). This observed S-phase blockade induced by arginine removal was reversible when 2.6–5.3 mg/L arginine was added to the medium (Figure S1C). In parallel, the overall incorporation of a 5-ethynyl-2’-deoxyuridine (EdU) analog in breast cancer-derived BT-549 spheroids decreased in the absence of arginine in a time-dependent manner (Figure S1D). We conclude that arginine is required for the cell-cycle progression of S-phase cells.
Figure 1. Essential role of arginine in DNA replication.
(A) Flow cytometry analysis of cell-cycle progression after BrdU labeling in the presence and absence of arginine. Top left: experimental design. Bottom left: analysis of BrdU-positive MDA-MB-231 cells in G1 phase (purple boxes). Right: percentage of BrdU-positive G1 cells in various cancer cell lines, quantified from three independent experiments with 50,000 events analyzed per experiment.
(B) Flow cytometry analysis of S-phase progression in double thymidine-synchronized MDA-MB-231 cells pulse-labeled with BrdU under ±arginine. Top: experimental design. Bottom: percentage of BrdU-positive S-phase cells at different time periods after thymidine block release, quantified from three independent experiments with 50,000 events analyzed per experiment.
(C and D) DNA fiber analysis using CldU (red) or IdU (green) antibodies in MDA-MB-231 (C) and BT-549 (D) cells under ±arginine. Top left: experimental design. Bottom left: representative images of DNA fibers. Middle: histograms shows percentage frequency of IdU track length. Right: quantitation of IdU track length, elongation ratio (IdU/CldU) and relative new replication origin firing. Each dot represents one fiber or one field with at least 100 tracts, or 40 fields were analyzed per sample. Scale bars, 10 μm.
(E) DNA combing analysis in MDA-MB-231 cells deprived of arginine for varying time frames and then resupplemented with arginine (84 mg/L, the same amount of arginine in full medium). Left: experimental design. Right: quantitation of CldU and IdU track lengths, with at least 100 tracts analyzed per sample.
Mean ± SEM is shown; ns: not significant; ***p < 0.005 (two-tailed unpaired Student’s t test). R, arginine.
We hypothesized that extracellular arginine is required for DNA replication fork progression. To address this possibility, we studied replication fork dynamics in cells at single-molecule resolution under ±arginine conditions. First, MDA-MB-231 and BT-549 cells were pulse labeled with a thymidine analog, 5-chloro-2’-deoxyuridine (CldU), for 40 min to establish the baseline tract lengths. After washing, cells were concomitantly labeled with the second thymidine analog iododeoxyuridine (IdU) under ±arginine conditions for 40 min (Figures 1C and 1D, top left diagrams). The track length of ongoing replication forks (with adjacent CldU [first labeling, red] and IdU [second labeling, green] signals) was examined and quantified (Figures 1C and 1D, bottom left panels). Quantitative analyses confirmed that the frequency of IdU-labeled tracts with reduced lengths increased in both cell lines upon arginine shortage, but this was not due to the new initiation of DNA replication (Figures 1C and 1D, right four panels).
Next, we determined whether the duration of arginine shortage affects DNA replication fork resumption after arginine is resupplied. Cells that were deprived of arginine for up to 120 min were first pulse labeled with CldU in arginine-free medium for an additional 40 min (i.e., up to 160 min) and subsequently pulse labeled with IdU in arginine-containing medium for 40 min (Figure 1E, left diagram). Notably, we found a time-dependent shortening of CldU-labeled tracts up to 100 min of arginine shortage (Figure 1E, middle panel). However, arginine replenishment restored DNA fork progression, visualized with IdU-labeled tracks and track length, despite varying arginine-free exposure times (Figure 1E, right panel). This indicates that transient extracellular arginine shortage is not a catastrophic DNA replication stress event, and replication forks are primed to resume upon arginine readdition. Together, we propose that arginine shortage holds DNA replication forks in a ‘‘standby’’ mode until the arginine shortage is resolved, and concomitantly, arginine shortage decelerates DNA elongation in an immediate, yet reversible, manner.
Extracellular arginine controls DNA replication by promoting the synthesis of histone H4 protein
Since arginine is a proteogenic amino acid and the biosynthesis of core histone proteins are tightly coupled to the cell cycle and peaks specifically during S phase, concomitant with DNA replication,18,34 we surveyed whether arginine promotes DNA replication by regulating translation. We adopted the AHA (L-azidohomoalaine; a methionine analog) incorporation assay to assess the direct impact of arginine shortage on newly synthesized proteins in S-phase cells (Figure S2A). While those Coomassie blue-stained signals were only subtly affected when cells were released in arginine-free medium (Figure S2B, left two panels), AHA-labeled proteins were markedly reduced upon arginine shortage (Figure S2B, right two panels). To determine the effect of arginine availability on newly synthesized histone H2B, H3, and H4, AHA-labeled proteins were isolated using a biotin pull-down strategy followed by western analyses. As shown in Figure 2A, histone H4, unlike H2B and H3, was not synthesized upon arginine shortage. In stark contrast, AHA-labeled ATF4 was induced, serving as an arginine shortage-induced endoplasmic reticulum (ER) stress control.5 To rule out the possibility that arginine shortage also reduced histone H4 mRNA abundance, qRT-PCR analysis was performed. We found that H4 mRNA abundance began to rise at 30 min after release from the double thymidine block, irrespective of arginine availability, validating that cells are in S phase (Figure S2C). However, H4 mRNA level subtly leveled off between 4 and 8 h post-release in full medium, rendering it significantly lower than H4 mRNA in arginine-deprived cells (Figure S2C). These results suggest that arginine shortage leads to a preferential depletion of newly synthesized histone H4 via a translational control.
Figure 2. The transfected recombinant histone H4 protein rescues DNA elongation during arginine shortage.
(A) Analysis of nascent AHA-labeled histones in S-phase-enriched MDA-MB-231 cells. Top: experimental design. Middle: representative western blot. Bottom: densitometric quantification of relative histone and ATF4 levels. The relative protein levels were calculated by dividing the intensity of the target band with that in cells released to +arginine medium for 60 min (set as 1). n = 3 independent experiments.
(B) DNA combing assay using MDA-MB-231 cells transfected with His-tagged histone H4 under ±arginine. Top: experimental design. Middle: non-linear regression curve analysis for IdU track length. Bottom: quantification of IdU track lengths and IdU/CldU ratio. At least 200 tracts were analyzed per sample.
Mean ± SEM is shown; ns: not significant; ***p < 0.005 (two-tailed unpaired Student’s t test).
Next, we tested whether increased newly synthesized H4 rescues DNA replication in arginine-deprived cells. To this end, we used a protein transfection reagent to introduce recombinant human His-tagged histone H4 proteins into cells, modeling an increase in newly synthesized H4, prior to arginine removal (Figure S2D, left panel). After confirming the incorporation of recombinant histone H4 protein into chromatin (Figure S2D, right panel), DNA combing assays showed that transfected recombinant H4 protein, at least in part, rescued both IdU-labeled track lengths and IdU/CldU ratio in arginine-depleted cells (Figure 2B). Together, our data suggest that while histone H4 is not the only protein involved in DNA replication whose synthesis is selectively suppressed by arginine depletion, the transfected recombinant histone H4 is able to partially phenocopy extracellular arginine to promote DNA replication. This also explains why extracellular arginine shortage prevents S-phase exit (Figures 1A and 1B).
Newly synthesized histone H4 marks are surrogate markers of arginine availability within the TME
To search for a marker for arginine availability and/or DNA replication speed, we exposed cells to various arginine concentrations for different durations and assessed histone H4 marks via western blot analysis. Both H4K5ac and H4K12ac marks, which are enriched at newly assembled chromatin,19 decreased in MDA-MB-231 (Figure 3A) and BT-549 cells (Figure S3A) when exposed to suboptimal extracellular levels of arginine. Multiple cell lines displayed similar results upon 2 h of extracellular arginine shortage, regardless of their de novo arginine biosynthesis capacity via ASS1 (Figure S3B).35 Because newly translated histone H4 is acetylated on lysines 5 and 12 in the cytoplasm prior to nuclear translocation, accumulation, and nucleosome deposition,36–38 we assessed the steady-state levels of H4K5ac and H4K12ac in the cytoplasmic and nuclear subfractions of cells released into ±arginine medium for 4 h post-double thymidine block. The lack of nuclear lamin A/C and cell surface CD44 signals in the cytoplasmic and nuclear fractions, respectively, confirmed the relative purity of each fraction. Following arginine shortage, the overall levels of lamin A/C, CD44, histone H3, and actin remained unchanged, and the nuclear accumulation of ATF4 served as an arginine shortage control (Figures 3B and S3C).5 Consistently, a universal decrease of newly synthesized histone H4 marks in total cell lysates and cytoplasmic and nuclear fractions was observed with arginine shortage (Figures 3B and S3C). This arginine removal-mediated decline of histone marks was more notable in H4K5ac and H4K12ac (~50%) but not in H4K16ac, H3K9me3, and H3K4me3 levels (Figure S3D). To support that the arginine shortage contributes to the reduction of newly synthesized histone H4 marks, supplementation of arginine rapidly restored H4K5ac and H4K12ac signals (Figure S3E), correlating well with DNA replication resumption (Figure 1E).
Figure 3. The newly synthesized histone H4 marks show a strong correlation with arginine availability.
(A) Analysis of H4K5ac and H4K12ac in MDA-MB-231 cells grown under varying arginine concentrations. Left: representative western blot. Right: densitometric quantification of relative levels of H4K5ac and H4K12ac normalized to control (84 mg/L) from three independent experiments.
(B) Subcellular fraction analysis of H4K5ac and H4K12ac levels in synchronized S-phase MDA-MB-231 cells grown under ±arginine. Left: representative western blot. Right: densitometric quantification of relative levels of H4K5ac and H4K12ac normalized to control (+arginine) from three independent experiments.
(C) H4K12ac and H4K5ac expression in MDA-MB-231 xenograft tumors. Top: representative IHC image of periphery and non-necrotic core. The corresponding enlarged view of inset area (black square) shows staining for H4K12ac and H4K5ac (brown). Dashed line indicate the boundary between periphery and non-necrotic core. Bottom: quantification of the intensities of H4K12ac-positive and H4K5ac-positive MDA-MB-231 cells from representative IHC images (n = 12) and the quantification of the respective percentage of cells with indicated H4K5ac or H4K12ac intensity. Insets in images show +++ strong, ++ medium, and + weak positive staining. Scale bars, 200 μm (50 μm in zoom images).
(D) Representative IHC images of CD31 (purple) and H4K12ac (brown) expression in MDA-MB-231 and BT-549 xenograft tumors. An overview of full tumor section is shown (top). A corresponding enlarged view of boxed area is shown below for the periphery and the non-necrotic core. Scale bars, 2.5 mm, 50 μm (20 μm in zoom images).
(E) Representative immunofluorescence image of a BT-549 tumor spheroid cryosection stained for H4K12ac (green), EdU (red), and Hoechst 33342 (blue). Scale bars, 100 μm (50 μm, 20 μm and 10 μm in zoom images).
Mean ± SEM is shown; ns, not significant; *p < 0.05; **p < 0.01; ***p < 0.005; two-tailed unpaired Student’s t test. le, long exposure; R, arginine.
A typical feature of solid tumors is their heterogeneous distribution of blood vessels and flow.39 Consequently, suboptimal nutrient availability develops within the avascularized TME. Previous studies have shown that arginine is one of the least available amino acids in the tumor core regions.6 To determine how DNA replication is perturbed by the lack of arginine supply in vivo, we sought to determine whether the distribution of H4K5ac and H4K12ac marks are located together with cell proliferation marker Ki67 in breast cancer MDA-MB-231 and BT-549 xenograft tumors using immunohistochemical (IHC) staining. Quantitative analyses confirmed, in general, the gradual decrease of H4K5ac and H4K12ac signals within the non-necrotic tumor core compared with the peripheral regions of MDA-MB-231 (Figure 3C) and BT-549 (Figure S4A) xenograft tumors. In addition, CD31-positive cells, an endothelial cell marker, and H4K12ac-positive cells were detected in close proximity at the periphery and non-necrotic core regions (Figure 3D), and H4K12ac-negative cells were largely enriched in the CD31-negative regions (Figure S4B). In addition, S-phase marker Ki67 was also enriched at the peripheral region, correlating well with H4K12ac (Figure S4C) in MDA-MB-231 and BT-549 xenografts. Consistent with these observations, most EdU-labeled cells were localized to the periphery of BT-549 spheroids (Figure 3E). Likewise, most H4K12ac-positive cells were found at the periphery of Hs578T and MCF-7 tumor spheroids (Figure S4D). To test the uniqueness of arginine shortage, we deprived other nutrients, such as glutamine, which is limited in tumor core regions, and serum, as well as leucine, which are indispensable for cell proliferation. Results showed that neither removal of serum, glutamine, or leucine affected H4K5ac and H4K12ac signals (Figures S4E–S4G).6,40 Collectively, these data suggest that extracellular arginine availability preferentially affects H4K5ac and H4K12ac abundance in vitro and in vivo.
Extracellular arginine shortage creates genotoxic tolerance
The maintenance of genome stability relies on successful DNA replication and proper DNA damage response in S phase. Next, we hypothesize that arginine availability might impact the DNA replication stress response and/or cell recovery from genotoxicity. When replication fork stress or DNA damage occurs, single-stranded DNA (ssDNA) is created and elicits the DNA damage response.21,41 RPA coating of ssDNA results in the activation of the ATR/Chk1 pathway to promote replication fork stability and genome stability.21,41 To test our hypothesis, we monitored the replication stress response via the phosphorylation of RPA (p-RPA(S33)) and Chk1 (p-Chk1(S345)) in arginine-starved cells. We used HU and camptothecin (CPT) as positive controls; both are known agents that cause DNA replication stress and activate the ATR/Chk1 pathway (Figure 4A).21,41 Arginine shortage (2 h) failed to induce p-Chk1(S345) and phosphorylation of RPA in multiple cell lines (Figures 4A and S5A). Consistent with this, we failed to observe an increase in the chromatin-bound RPA levels, a reflection of the degree of ssDNA exposure, in pre-extracted S-phase cells via flow cytometry post-arginine shortage (4 h) (Figure 4B). This is in contrast to the observations following HU and CPT treatment.42 The inability of extracellular arginine shortage to activate Chk1 phosphorylation and RPA chromatin binding suggests that arginine shortage did not lead to significant ssDNA accumulation while deaccelerating DNA replication.
Figure 4. Arginine shortage promotes resistance to genotoxic stress.
(A) One representative western blot for RPA (S33) and Chk1 (S345) phosphorylation in HU-treated MDA-MB-231 cells under ±arginine from three independent experiments.
(B and C) Quantification of chromatin-bound RPA-positive (B) and γH2AX-positive (C) EdU-labeled MDA-MB-231 cells grown under stated conditions for 4 h, as analyzed by flow cytometry. n = 3 independent experiments. At least 10,000 events were collected and analyzed.
(D) γH2AX expression analysis in BT-549 spheroids treated under stated conditions for 4 h. Top: representative fluorescent images of BT-549 spheroid cryosections stained with DAPI (blue) and γH2AX (red). Bottom: quantification of relative γH2AX intensity normalized to tumor area. n = 4 independent experiments. Scale bars, 100 μm.
(E) Analysis of relative cell recovery of MDA-MB-231 cells under stated conditions for 4 h, followed by recovery in regular full medium. Cell viability was analyzed at indicated time points.
Mean ± SEM is shown; two-tailed unpaired Student’s t test; ns, not significant; *p < 0.05; **p < 0.01; ***p < 0.005. CPT, camptothecin; HU, hydroxyurea; R, arginine.
We next determined whether the DNA damage response to double-strand DNA breaks (DSBs) was impacted under ±arginine conditions by assessing the histone variant H2AX phosphorylation on S139 (γH2AX) in MDA-MB-231 cells.43 To induce DSBs, we used (1) CPT alone and (2) HU, a recognized source of DNA damage by misincorporation, in combination with either Chk1 inhibitor (Chk1i; rabusertib) or an ATR inhibitor (ATRi; BAY1895344).44–47 Chk1i (1 μM) alone, but not ATRi (100 nM) alone, induced p-Chk1(S345) without a marked γH2AX accumulation (Figure S5B, lane #7 vs. #5), and co-treatment of ATRi (100 nM) blocked HU-induced p-Chk1(S345) but induced γH2AX (Figure S5B, lane #9 vs. #3) when serving as a positive control.48 As expected, arginine depletion reduced γH2AX chromatin binding in S-phase cells treated with genotoxic agents as analyzed by flow cytometry (Figure 4C). Notably, arginine shortage attenuated ATRi- or Chk1i-sensitized γH2AX accumulation in HU-treated MDA-MB-231 cells (Figure S5B, lane #9 vs. #10 and lane #11 vs. #12) and γH2AX signals in tumor spheroids treated with CPT (Figure 4D). Taken together, cells exposed to arginine shortage appeared to exhibit reduced DNA replication stress responses than cells maintained in full medium upon treatment with the same genotoxic agents.
To determine the impact of arginine on cell recovery from genotoxic treatments, MDA-MB-231 cells were subjected to CPT (2 μM), HU (2 mM)+ATRi (100 μM), or HU (2 mM)+Chk1i (1 μM) under ±arginine conditions for 4 h and then cultured in drug-free full medium for the indicated time periods, with or without S-phase synchronization (Figure 4E). Similar results were found in CPT-treated BT-549 and MCF-7 cells (Figure S5C). There was a significant gain in resistance to genotoxic agents tested in arginine-starved cells at day 3 post-treatment. A similar result was observed in tumor spheroids (Figure S5D). Collectively, unlike other DNA replication stalling chemicals that induce DNA damage responses, arginine shortage slows down DNA replication without causing DNA damage responses while enabling tolerance to genotoxic insults.
Distinct protein dynamics at arginine shortage-stressed replication forks
We next sought to identify arginine shortage-induced and ATR-independent signaling events that mediate DNA replication slowing in MDA-MB-231 cells using the accelerated isolation of proteins on nascent DNA (aniPOND) method.49 Due to the generation of shorter EdU-incorporated nascent DNA strands under arginine removal (Figures 1C and 1D), we expected less H4K5ac, H4K12ac, H2B, H3, and H4 to be pulled down from EdU-labeled cells exposed to arginine shortage compared with cells grown in full medium (Figure 5A). Furthermore, arginine replenishment following shortage restored these histone levels to similar levels as those observed in cells grown in full medium (Figure 5A). Decreased levels of H4K5ac and H4K12ac marks showed that there was less newly synthesized H4 occupying nascent DNA strands during DNA replication under arginine shortage (Figure 5A). In stark contrast to reduced histone levels, we found that more PCNA was bound to EdU-labeled DNA after arginine removal (Figure 5B), and this did not occur with arginine resupply following arginine shortage (Figure 5B). As previously reported,43,50 CPT and HU treatment did not increase nascent DNA-bound PCNA levels (Figure 5B) when serving as controls. Furthermore, chromatin fractionation experiments revealed that a higher level of chromatin-bound PCNA was exclusively detected in arginine-deprived cells, and both CPT and HU treatments failed to affect occupancy of chromatin-bound PCNA under ±arginine conditions (Figure S6A). These results suggested that there was increased occupancy of PCNA on nascent DNA strands at stalled replication forks upon arginine withdrawal.
Figure 5. Arginine shortage-dependent regulation of PCNA and HLTF.
(A and B) Representative western blots of fork-associated proteins isolated using aniPOND method from EdU-positive MDA-MB-231 cells grown under stated conditions. EdU (−), negative control. n = 3 independent experiments.
(C–F) Western blot analysis of immunoprecipitated PCNA-associated proteins from Myc-ubiquitin-transfected FLAG-PCNA-overexpressed HEK293T cells subjected to indicated treatments. Left: representative western blots are shown for the indicated antibodies. Right: densitometric quantification of relative levels of indicated proteins normalized to control (+arginine) are shown. n = 3 independent experiments.
Mean ± SEM is shown; two-tailed unpaired Student’s t test; ns, not significant; *p < 0.05; **p < 0.01; ***p < 0.005. R, arginine; le, long exposure; −R + R, arginine depletion followed by resupply.
During DNA replication, PCNA is repeatedly loaded and unloaded by the replicating clamp loader, replication factor C (RFC) complex, and an alternative PCNA ring opener, the ATAD5 (ELG1 in yeast)-RFC-like complex (ATAD5-RLC). ATAD5-RLC unloads replication-coupled PCNA.51 Maintaining the balance of PCNA loaded onto nascent DNA strands is crucial for DNA replication.52 To further test the effect of arginine shortage on chromatin-bound PCNA levels, we performed chromatin fractionation to determine the chromatin-associated fraction of subunits of the PCNA unloader, the RLC (ATAD5/RFC2–5) complex.51 Consistent with the increased PCNA levels, chromatin-associated ATAD5, RFC2, and RFC5, subunits of the RLC complex, were significantly reduced in arginine-deprived MDA-MB-231 cells. In addition, the total level of these proteins as shown in whole-cell extracts was also reduced upon arginine removal (Figure S6B). A similar observation was made in arginine-deprived BT-549 cells, except that the ATAD5 level was almost undetectable (Figure S6B). While consistent with the reduced H4K5Ac and H4K12Ac deposition, a marked reduction in chromatin-bound BRD4 (Figure S6B), an acetyl-histone-binding chromatin reader, seems to conflict with increased PCNA loading. A previous study showed that ATAD5, the major PCNA unloader, had dominant effects on PCNA unloading, and overexpression of BRD4 significantly induced PCNA accumulation on chromatin. However, depletion of BRD4 can only slightly decrease the levels of PCNA on nascent DNA.53 Therefore, even though the downregulation of BRD4 should promote PCNA unloading during arginine shortage, the decreased levels of ATAD5, RFC2, and RFC5 could eventually lead to insufficient unloading of PCNA. Together, these findings suggest that the reduced chromatin binding of the PCNA-unloader ATAD5-RLC causes the increased retention of PCNA in the arginine-deprived cells. It remains to be elucidated how arginine removal affects the cross-talk among BRD4, ATAD5, and H4K5ac/H4K12ac to fine-tune PCNA retention on nascent DNA strands.
Given that altered ubiquitination of replication machinery is a hallmark of stressed replication forks,54 we investigated if elevated levels of PCNA might correspond to altered PCNA ubiquitination. PCNA can be either monoubiquitinated or polyubiquitinated to activate translesion synthesis or fork reversal.30,31 To test whether PCNA is ubiquitinated upon arginine removal in a manner similar to that observed under canonical replication stress, HEK-293T cells stably expressing FLAG-tagged PCNA were transiently transfected with Myc-tagged ubiquitin in our biochemical analyses. Following arginine removal (4 h) and immunoprecipitation with an anti-FLAG antibody, increased K63-linked polyubiquitination and K164 monoubiquitination of PCNA were detected (Figure 5C). HLTF is one of the E3 ligases that catalyzes PCNA K63-linked polyubiquitination in a RAD6-RAD18-dependent manner.32,33 Immunoprecipitation followed by western analyses further revealed that arginine removal enhanced the recruitment of RAD6, RAD18, and HLTF to PCNA (Figure 5D). The replenishment of arginine reversed arginine shortage-increased K63-linked polyubiquitination and recruitment of RAD6, RAD18, and HLTF to PCNA without affecting PCNA levels (Figure 5E). Lastly, knockdown of HLTF abolished arginine shortage-induced K63-linked polyubiquitination of PCNA (Figure 5F). Next, we performed a chromatin fractionation and found the levels of chromatin-bound PCNA, HLTF, RAD18, and RAD6 were all significantly higher in arginine-deprived cells, indirectly supporting that increased associations among RAD6, RAD18, HLTF, and PCNA on chromatin resulted in more PCNA K63-linked polyubiquitination at forks (Figure S6A). Based on these observations, we propose that HLTF-mediated K63-linked polyubiquitination of PCNA and/or its associated proteins increases in response to extracellular arginine shortage in a reversible manner. To examine the possibility whether other K63-linked E3 ligases, such as SHPRH,55 are involved in arginine shortage-induced PCNA K63-linked polyubiquitination, SHPRH was knocked down in arginine-deprived cells (Figure S6C). However, depletion of SHPRH failed to reverse arginine shortage-induced K63-linked polyubiquitination. These findings demonstrate that shortage of arginine induces HLTF- , but not SHPRH-, mediated K63-linked polyubiquitination of PCNA. To examine whether there is a link between histone H4 synthesis and PCNA K63-linked polyubiquitination, we impaired histone H4 biosynthesis by knocking down FLASH or SLBP, both of which are essential for histone biosynthesis (Figure S6D).20 Indeed, knockdown of FLASH or SLBP decreased newly synthesized H4 marks, as previously reported, and increased K63-linked polyubiquitination of PCNA (Figure S6D). This observation links arginine shortage-suppressed de novo histone H4 synthesis to HLTF-mediated PCNA K63-ubiquitination.
Arginine shortage promotes HLTF-dependent genotoxin adaption
To mechanistically explore the role of HLTF-mediated K63-linked polyubiquitination of PCNA and/or its associated proteins in response to arginine shortage, we first sought to analyze how HLTF loss altered cellular responses to arginine removal. Previously, HLTF loss reportedly led to an unrestrained fork progression.32,33 To confirm and extend this finding in arginine-starved cells, we used two different small interfering RNAs (siRNAs) to knock down HLTF and monitored fork progression under ±arginine in MDA-MB-231 cells as described in Figure 1C. Compared with siRNA control (siCtrl)-treated cells, knockdown of HLTF alone did not markedly affect the level of H4K5ac and H4K12ac marks (Figure S7A) and the ratio of IdU/CldU track lengths (Figure 6A) in unstarved cells, respectively. However, knockdown of HLTF partially rescued the arginine shortage-suppressed IdU/CldU ratio (Figure 6A). To provide additional evidence supporting the effect of PCNA ubiquitination on fork slowing in arginine-starved cells, isogenic paired HEK-293T wild-type (WT) and HEK-293T-K164R cells were employed.56 Figure 6B shows that the PCNA K164R mutation, at least in part, rescued the IdU/CldU ratio in arginine-starved cells, and inactivating HLTF in the K164R cells had no further effect. Consistent with the idea that HLTF has an activity of slowing fork progression in cells experiencing replication stress,33 our observation confirmed that HLTF restrains replication fork progression in arginine-starved cells.
Figure 6. HLTF and PCNA protects genome stability in arginine-starved cells.
(A) Analysis following MDA-MB-231 cells treated with indicated siRNAs targeting HLTF and ±arginine. Top left: experimental outline. Bottom left: representative western blot analysis of HLTF. Right: quantitation of IdU/CldU ratio. n = 3 independent experiments. Each dot represents one fiber. At least 100 tracts were analyzed per sample.
(B) Analysis of HEK-239T cells overexpressing wild-type (WT) or mutated (K164R) PCNA treated with siRNA targeting HLTF ±arginine. Top left: experimental outline. Bottom left: representative western blot analysis of K164 ubiquitinated PCNA (ub-PCNA) and HLTF. Right: quantitation of IdU/CldU ratio. n = 3 independent experiments. Each dot represents one fiber. At least 100 tracts were analyzed per sample.
(C) Analysis of the effect of DNA2 inhibitor C5 or MRE11 inhibitor mirin on replication fork stability under ±arginine and exposure to HU in HLTF-knockdown MDA-MB-231 cells. Top: experimental outline. Bottom: quantification of IdU/CldU ratio. n = 3 independent experiments. Each dot represents one fiber. At least 100 tracts were analyzed per sample.
(D) Analysis of HLTF knockdown on chromatin-bound RPA and γH2AX in MDA-MB-231 cells grown under ±arginine and subjected to genotoxic insult. Top: experimental outline. Bottom: quantification of chromatin-bound RPA-positive or γH2AX-positive cells following flow analysis. At least 10,000 events were collected and analyzed. n = 3 independent experiments.
(E) Relative cell recovery of control or HLTF-knockdown MDA-MB-231 cells after treatment with indicated genotoxic agents and ±arginine. n = 3 independent experiments.
(F) Relative cell recovery of WT or mutated (K164R) PCNA-expressing HEK-293T cells after treatment with indicated genotoxic agents and ±arginine. n = 3 independent experiments.
Mean ± SEM is shown; ns, not significant; *p < 0.05; **p < 0.01; ***p < 0.005; two-tailed unpaired Student’s t test. R, arginine.
Next, we sought to address the contribution of HLTF to the integrity of arginine shortage-stressed replication forks. In order to do this, we treated HLTF-depleted MDA-MB-231 cells with a high concentration of HU (4 mM), which has been shown to promote fork reversal,57 and measured the IdU/CldU-labeled fiber ratio under ±arginine conditions (Figure 6C). siRNA depletion of HLTF alone had no effect on fork degradation in cells cultured with HU (4 mM) and arginine. In contrast, the removal of arginine and exposure to HU (4 mM) notably accelerated fork degradation in HLTF-depleted cells (Figure 6C, lane #4 vs. #3). Collectively, these findings suggest that HLTF is critical to protect nascent DNA from degradation in arginine-deprived cells when exposed to HU (4 mM). Mirin, an MRE11 inhibitor, did not suppress nascent strand degradation in arginine-starved, HLTF-depleted cells (Figure 6C). In contrast, treatment with C5, an inhibitor of the nuclease DNA2, restored the nascent track length (Figure 6C), indicating the possible involvement of DNA2-mediated nascent DNA nucleolysis in arginine-starved, HLTF-depleted cells. To determine whether the protective function of HLTF relies on its Hiran domain to bind specifically to the 3′ end of ssDNA as previously reported, we tested the fork integrity in cells overexpressing siRNA-resistant WT HLTF and Hiran-mutated (R71E) HLTF.58 Figure S7B shows that knockdown of endogenous HLTF in MDA-MB-231 cells induced the degradation of replication fork; however, both WT and Hiran-mutated HLTF can almost fully restore nascent DNA strands (Figure S7B), indicating that HLTF exerts its protective role in arginine-depleted cells independent of its ssDNA binding ability. To further investigate whether other known translocases share the protective function of HLTF, the fork integrity was tested in SMARCAL1- or ZRANB3-knockdown cells (Figure S7C).59,60 Results showed that knockdown of SMARCAL1 or ZRANB3 did not induce degradation of replication fork in arginine-starved cells (Figure S7C). These observations indicated that HLTF-dependent fork reversal is not involved in resolving arginine shortage-induced fork stalling.
To determine the impact of HLTF on DNA damage signaling upon arginine shortage, we then used a flow cytometry-based method to quantitatively monitor RPA and phospho-H2AX or γH2AX chromatin binding in S-phase cells (Figure 6D, top diagram).42 As expected, brief exposure to a high concentration of CPT (2 μM, 4 h) increased both RPA and γH2AX chromatin binding in MDA-MB-231 S-phase cells cultured in full medium, and arginine shortage attenuated both signals (Figure 6D, bottom panels). Notably, HLTF loss increased both RPA and γH2AX chromatin binding in EdU-positive cells under arginine shortage (Figure 6D, bottom panels, lanes #10 and #12 vs. #8). A similar trend was noted (Figure 6D, bottom panels) in cells treated with HU+ATRi (lanes #16 and #18 vs. #14) or HU+CHK1i (lanes #22 and #24 vs. #20) under arginine shortage. Lastly, we determined the role of HLTF in protecting MDA-MB-231 cells from genotoxins under ±arginine conditions. Compared with Figure 4E, HLTF knockdown mitigated the advantage conferred by arginine removal on cell recovery from indicated genotoxic treatments (Figure 6E). Similar observations, albeit less pronounced, were made in PCNA-mutated (K164R) HEK-293T cells (Figure 6F). While PCNA K63-linked polyubiquitination is likely associated with poor recovery of HLTF-knockdown cells from CPT, HU+ATRi, or HU+Chk1i, the exact contribution to the better recovery remains unclear. It is possible that HLTF-catalyzed PCNA K164 ubiquitination (Figure 5F) is indispensable for arginine shortage to enable cells to better cope with DNA damage induced by a variety of genotoxic agents.
AOH1996 is the lead compound targeting the L126-Y133 region of PCNA to interfere with its interaction with partners.61 To address PCNA function in further detail, we determined the effect of AOH1996 on the fork integrity in arginine-deprived cells. ani-POND assays show that AOH1996 suppressed PCNA accumulation on replication forks (Figure S8A) without affecting the levels of newly synthesized H4 marks (Figure S8B). Like HLTF knockdown, a fork protection assay showed that AOH1996 accelerated fork degradation in arginine-deprived cells (Figure S8C, lane #4 vs. #3). Next, to address whether AOH1996 could reverse arginine shortage-induced K63-linked polyubiquitination of PCNA and genotoxin resistance, immunoprecipitation, a flow cytometry-based assay, and a cell recovery assay were performed.42 We showed that AOH1996 reversed arginine shortage-induced K63-linked polyubiquitination of PCNA (Figure S8D, lane #5 vs. #3). AOH1996 also increased γH2AX accumulation on chromatin in arginine-starved, EdU-positive cells treated with genotoxins (Figure S8E). Additionally, AOH1996 abolished the protection provided by arginine shortage on cells recovered from genotoxic insult (Figure S8F). Taken together, we suggest that PCNA loading and its ubiquitination are important for maintaining fork integrity and recovery from genotoxicity in arginine-deprived cells. Collectively, we propose a model, based on our findings, that arginine shortage inhibits H4 translation to stall DNA replication and enables HLTF-mediated K63-linked polyubiquitination of PCNA and/or its associated proteins and hyper-PCNA accumulation on nascent DNA to protect arginine-depleted cells against genotoxins (Figure S9).
DISCUSSION
We have discovered that arginine, a nutrient that is often lacking in TME,6 plays a crucial role in modulating DNA replication. Our findings indicate that a shortage of arginine significantly impairs the progression of replication forks by inhibiting the translation of histone H4 and leading to an excessive buildup and HLTF-mediated ubiquitination of PCNA on newly synthesized DNA strands. This establishes a clear connection between arginine shortage and the development of genotoxic adaptation.
Replication-dependent histone biosynthesis has long been considered as a crucial factor in regulating DNA replication and cell-cycle progression.18–20 Our studies find that increased levels of histone H4 alone can promote DNA replication elongation under conditions of arginine shortage. Despite being only a moderately abundant amino acid, arginine plays an important role in histone H4 synthesis. This is due to the fact that histone H4 contains a higher percentage of arginine codons (CGC and CGU) compared with other histones, making it more sensitive to changes in arginine availability (Tables 1, 2, and 3). The observed stalling of DNA replication in response to arginine shortage is thought to be regulated by the selective loss of arginine tRNA charging and ribosome pausing at specific arginine codons. Darnell et al. reported that the selective loss of arginine tRNA charging during arginine shortage regulates translation via ribosome pausing at specific arginine codons.62 Altogether, our observations raise the possibility that arginine residues in histone H4 may carry out a nutrient-sensing function by changing its synthesis in response to extracellular arginine availability. This is an important area of investigation, as it may shed light on the mechanisms by which cells respond to changes in nutrient availability.
Table 1.
The mRNA and amino acids composition of Homo sapiens histone H4 (H4C1 H4)
mRNA | AUG | UCU | GGA | CGU | GGU | AAG | GGC | GGG | AAG | GGU | UUG | GGU | AAG | GGG | GGU | GCC | AAG | CGC | CAC | CGC |
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
Protein | MET | SER | GLY | ARG | GLY | LYS | GLY | GLY | LYS | GLY | LEU | GLY | LYS | GLY | GLY | ALA | LYS | ARG | HIS | ARG |
AAG | GUG | UUG | CGU | GAC | AAC | AUC | CAG | GGC | AUC | ACC | AAG | CCG | GCC | AUC | CGG | CGU | CUG | GCC | CGG | |
LYS | VAL | LEU | ARG | ASP | ASN | ILE | GLN | GLY | ILE | THR | LYS | PRO | ALA | ILE | ARG | ARG | LEU | ALA | ARG | |
CGU | GGC | GGU | GUG | AAG | CGG | AUC | UCU | GGU | CUG | AUC | UAC | GAG | GAG | ACU | CGC | GGG | GUG | CUC | AAG | |
ARG | GLY | GLY | VAL | LYS | ARG | ILE | SER | GLY | LEU | ILE | TYR | GLU | GLU | THR | ARG | GLY | VAL | LEU | LYS | |
GUG | UUU | UUG | GAG | AAC | GUG | AUC | CGU | GAC | GCU | GUC | ACC | UAU | ACG | GAG | CAC | GCC | AAG | CGC | AAG | |
VAL | PHE | LEU | GLU | ASN | VAL | ILE | ARG | ASP | ALA | VAL | THR | TYR | THR | GLU | HIS | ALA | LYS | ARG | LYS | |
ACA | GUC | ACU | GCC | AUG | GAC | GUG | GUC | UAC | GCG | CUU | AAG | CGC | CAG | GGA | CGC | ACC | CUU | UAU | GGC | |
THR | VAL | THR | ALA | MET | ASP | VAL | VAL | TYR | ALA | LEU | LYS | ARG | GLN | GLY | ARG | THR | LEU | TYR | GLY | |
UUU | GGC | GGU | UAA | – | – | – | – | – | – | – | – | – | – | – | – | – | – | – | – | |
PHE | GLY | GLY | STOP | – | – | – | – | – | – | – | – | – | – | – | – | – | – | – | – |
Ribosome pause-site arginine codons: CGU, CGC; non-ribosome pause-site arginine codons: CGG, CGA, AGA, AGG. Arginine pausing codon/all codons: 11/104 (10.58%). CGU/all codons: 5/104 (CGU/arginine codon: 5/14; CUG/ribosome pausing arginine codon: 5/11). CGC/all codons: 6/104 (CGC/arginine codon 6/14; CGG/ribosome pausing arginine codon: 6/11).
Table 2.
The mRNA and amino acids composition of Homo sapiens histone H3 (H3C14)
mRNA | AUG | GCC | CGU | ACU | AAG | CAG | ACU | GCU | CGC | AAG | UCG | ACC | GGC | GGC | AAG | GCC | CCG | AGG | AAG | CAG |
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
Protein | MET | ALA | ARG | THR | LYS | GLN | THR | ALA | ARG | LYS | SER | THR | GLY | GLY | LYS | ALA | PRO | ARG | LYS | GLN |
CUG | GCC | ACC | AAG | GCG | GCC | CGC | AAG | AGC | GCG | CCG | GCC | ACG | GGC | GGG | GUG | AAG | AAG | CCG | CAC | |
LEU | ALA | THR | LYS | ALA | ALA | ARG | LYS | SER | ALA | PRO | ALA | THR | GLY | GLY | VAL | LYS | LYS | PRO | HIS | |
CGC | UAC | CGG | CCC | GGC | ACC | GUA | GCC | CUG | CGG | GAG | AUC | CGG | CGC | UAC | CAG | AAG | UCC | ACG | GAG | |
ARG | TYR | ARG | PRO | GLY | THR | VAL | ALA | LEU | ARG | GLU | ILE | ARG | ARG | TYR | GLN | LYS | SER | THR | GLU | |
CUG | CUG | AUC | CGC | AAG | CUG | CCC | UUC | CAG | CGG | CUG | GUA | CGC | GAG | AUC | GCG | CAG | GAC | UUU | AAG | |
LEU | LEU | ILE | ARG | LYS | LEU | PRO | PHE | GLN | ARG | LEU | VAL | ARG | GLU | ILE | ALA | GLN | ASP | PHE | LYS | |
ACG | GAC | CUG | CGC | UUC | CAG | AGC | UCG | GCC | GUG | AUG | GCG | CUG | CAG | GAG | GCC | AGC | GAG | GCC | UAC | |
THR | ASP | LEU | ARG | PHE | GLN | SER | SER | ALA | VAL | MET | ALA | LEU | GLN | GLU | ALA | SER | GLU | ALA | TYR | |
CUG | GUG | GGG | CUG | UUC | GAA | GAC | ACG | AAC | CUG | UGC | GCC | AUC | CAC | GCC | AAG | CGC | GUG | ACC | AUU | |
LEU | VAL | GLY | LEU | PHE | GLU | ASP | THR | ASN | LEU | CYS | ALA | ILE | HIS | ALA | LYS | ARG | VAL | THR | ILE | |
AUG | CCC | AAG | GAC | AUC | CAG | CUG | GCC | CGC | CGC | AUC | CGU | GGA | GAG | CGG | GCU | UAA | – | – | – | |
MET | PRO | LYS | ASP | ILE | GLN | LEU | ALA | ARG | ARG | ILE | ARG | GLY | GLU | ARG | ALA | STOP | – | – | – |
Ribosome pause-site arginine codons: CGU, CGC; non-ribosome pause-site arginine codons: CGG, CGA, AGA, AGG. Arginine pausing codon/all codons: 12/137 (8.76%). CGU/all codons: 2/137 (CGU/arginine codon: 2/18; CUG/ribosome pausing arginine codon: 2/12). CGC/all codons: 10/137 (CGC/arginine codon 10/18; CGG/ribosome pausing arginine codon: 10/12).
Table 3.
The mRNA and amino acids composition of Homo sapiens histone H2 (H2BC21)
mRNA | AUG | CCU | GAA | CCG | GCA | AAA | UCC | GCU | CCG | GCC | CCU | AAA | AAG | GGC | UCC | AAG | AAA | GCC | GUC | ACC |
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
Protein | MET | PRO | GLU | PRO | ALA | LYS | SER | ALA | PRO | ALA | PRO | LYS | LYS | GLY | SER | LYS | LYS | ALA | VAL | THR |
AAA | GCC | CAG | AAG | AAA | GAC | GGC | AAG | AAG | CGC | AAG | CGC | AGC | CGC | AAA | GAG | AGC | UAC | UCC | AUC | |
LYS | ALA | GLN | LYS | LYS | ASP | GLY | LYS | LYS | ARG | LYS | ARG | SER | ARG | LYS | GLU | SER | TYR | SER | ILE | |
UAC | GUG | UAC | AAG | GUG | CUG | AAG | CAG | GUC | CAC | CCC | GAC | ACC | GGC | AUC | UCG | UCC | AAG | GCC | AUG | |
TYR | VAL | TYR | LYS | VAL | LEU | LYS | GLN | VAL | HIS | PRO | ASP | THR | GLY | ILE | SER | SER | LYS | ALA | MET | |
GGC | AUC | AUG | AAC | UCC | UUC | GUC | AAC | GAC | AUC | UUC | GAG | CGC | AUC | GCG | GGA | GAG | GCU | UCC | CGC | |
GLY | ILE | MET | ASN | SER | PHE | VAL | ASN | ASP | ILE | PHE | GLU | ARG | ILE | ALA | GLY | GLU | ALA | SER | ARG | |
CUG | GCG | CAC | UAC | AAC | AAG | CGC | UCC | ACC | AUC | ACA | UCC | CGC | GAG | AUC | CAG | ACG | GCC | GUG | CGC | |
LEU | ALA | HIS | TYR | ASN | LYS | ARG | SER | THR | ILE | THR | SER | ARG | GLU | ILE | GLN | THR | ALA | VAL | ARG | |
CUG | CUG | CUG | CCC | GGC | GAG | CUG | GCC | AAG | CAC | GCC | GUG | UCC | GAG | GGC | ACC | AAG | GCG | GUC | ACC | |
LEU | LEU | LEU | PRO | GLY | GLU | LEU | ALA | LYS | HIS | ALA | VAL | SER | GLU | GLY | THR | LYS | ALA | VAL | THR | |
AAG | UAC | ACC | AGC | UCC | AAG | UGA | – | – | – | – | – | – | – | – | – | – | – | – | – | |
LYS | TYR | THR | SER | SER | LYS | STOP | – | – | – | – | – | – | – | – | – | – | – | – | – |
Ribosome pause-site arginine codons: CGU, CGC; non-ribosome pause-site arginine codons: CGG, CGA, AGA, AGG. Ribosome pause-site encoded arginine: ARG; non-ribosome pause-site encoded arginine: ARG. Arginine pausing codon/all codons: 8/127 (6.3%). CGU/all codons: 0/127 (CGU/arginine codon: 0/8; CUG/ribosome pausing arginine codon: 0/8). CGC/all codons: 8/127 (CGC/arginine codon 8/8; CGG/ribosome pausing arginine codon: 8/8).
In addition to the relationship between extracellular arginine and histone H4, our studies also indicate that arginine is crucial for maintaining DNA replication through its link with PCNA (Figure S9). It is not possible to exclude the involvement of other factors, such as chromatin assembly factor-1 (CAF-1) or other histone chaperones (i.e., FACT and ASF1) in connecting arginine-dependent H4 translation with PCNA’s K63-linked polyubiquitination.63,64 Additionally, changes in arginine availability may impact the interaction between PCNA and other key components involved in DNA replication, such as DNA polymerases, RFC, and RFC-like complexes. These proteins work together to coordinate the recycling of histones and PCNA, ensuring the proper assembly of chromatin on newly replicated DNA.16 However, our results do not suggest the presence of large ssDNA tracts during arginine shortage, although DNA nicks cannot be ruled out, as PCNA interacts with DNA ligase I.52 Increased levels of PCNA K63-linked polyubiquitination in response to arginine deficiency have the potential to serve as a marker for DNA replication stress. Further studies are needed to better understand the role of arginine in regulating PCNA’s residence on nascent DNA strands and maintaining genome stability.
An open question concerns the belief that PCNA is the only protein that significantly accumulates on nascent DNA strands in arginine-deprived cells. However, it is possible that technical issues, such as the efficiency of cross-linking or the conditions used for chromatin purification, may have led to the loss of PCNA-associated factors that are loosely bound. These studies have led to a surprising conclusion: arginine shortage activates Rad6/Rad18- mediated PCNA K164 monoubiquitination and HLTF-catalyzed PCNA K63-linked polyubiquitination. Given that PCNA must be repeatedly recycled during replication to allow for continuous DNA replication,65 it is intriguing to find that arginine plays a role in regulating the timely removal of PCNA from nascent DNA strands by controlling the levels of unloading machinery proteins. The exact role of HLTF-mediated K63-linked polyubiquitination of PCNA and associated proteins in stalling DNA replication remains to be determined. We surmise that the K63-linked E3 ligase activity of HLTF, but not SHPRH, is crucial in maintaining the association of PCNA with the nascent strands and potentially affecting its interactions with other proteins. Another possibility is that K63-linked polyubiquitination of PCNA on newly synthesized DNA strands restricts its sliding, leading to the accumulation of PCNA behind stalled replication forks in arginine-deficient cells. Further research is needed to clarify the regulation of HLTF activity and its relationship with the DNA replication machinery, including the role of K63-linked polyubiquitination in stalling fork progression and enabling cells to withstand genotoxins. Additionally, it will be interesting to examine the coordination between PCNA unloading and histone deposition during DNA replication in response to arginine shortage, as well as the regulation of PCNA activity by arginine to maintain genome stability.
The full effects of arginine shortage on genome stability remain unclear. While the observed S-phase arrest in arginine-deficient cells alone is not enough to account for their increased resistance to genotoxicity, arginine shortage does slow down DNA replication without triggering a damage response. Our study suggests that HLTF plays a role in genome protection in an arginine-dependent manner, potentially through its interactions with other proteins or its E3 ligase activity. As demonstrated by Peng et al., long-term treatment with HU (4 mM) can result in fork degradation in FANCJ-knockout cells, but HLTF depletion significantly reduces this degradation.66 However, we observe that the absence of HLTF exacerbates HU (4 mM)-induced fork degradation in arginine-deficient cells. Furthermore, K63-linked polyubiquitination of PCNA blocks the loading of DNA2 onto replication forks.56 This aligns with our findings that arginine deficiency leads to DNA2-dependent, but MRE11-independent, nucleolytic degradation of nascent DNA at stalled replication forks in HLTF-deficient cells (Figure 6C). Our results are consistent with previous studies, which have shown that HLTF confers resistance to DNA damaging agents and that its loss sensitizes arginine-starved cells.67–70 HLTF is widely recognized for its role in post-replication repair, where it facilitates replication fork reversal to bypass blocks and restart replication in response to stress.71 Additionally, van Toorn et al. recently reported that HLTF is necessary for efficient nucleotide excision repair.72 Our study excluded the role of HLTF and other translocases in fork reversal as a mechanism for managing replication stress caused by arginine shortage (Figures S7B and S7C). Nevertheless, the exact mechanism by which HLTF-mediated K63-linked polyubiquitination of PCNA maintains genome stability independent of fork reversal in the face of arginine shortage remains uncertain and deserves further investigation.
In summary, we have uncovered a hitherto unidentified pathway, the histone H4 translation-arrest and HLTF-mediated PCNA K63-linked polyubiquitination pathway, which provides an immediate and reversible response to slow DNA replication when external arginine availability changes (Figure S9). Our results indicate that the amount of arginine produced endogenously by ASS1 is not enough to support DNA replication and that mechanisms regulating DNA replication, elongation, and preventing replication fork disruption may be more widespread than previously understood. The instant slowing of the replication fork in response to a shortage of extracellular arginine helps maintain genome stability against various agents and prevents harmful consequences in rapidly dividing cells. This arrest of replication also gives any existing lesions time to be repaired, allowing replication to resume once arginine becomes available again. Our findings of H4 translation inhibition and PCNA K63-linked polyubiquitination-mediated arrest preventing genotoxin-induced genomic instability highlight the significance of local nutrient availability in the TME and its impact on tumor cell heterogeneity. Our study emphasizes the protective effect of arginine shortage against genotoxicity, which may be a common phenomenon in the central regions of various types of solid tumors, regardless of their genetic, epigenetic (i.e., as histone and DNA modification), and metabolic characteristics. Additionally, the fact that administering arginine can make tumors more sensitive to radiation therapy supports the importance of considering arginine levels when developing therapeutic approaches and the potential impact on the response to anti-cancer therapy.73
Limitations of the study
Our study sheds light on the role of extracellular arginine in regulating DNA replication through promoting histone H4 synthesis and facilitating PCNA recycling. Our findings are supported by the results of our histone H4 transfection experiment and the involvement of the HLTF-mediated PCNA K63-linked polyubiquitination pathway. However, a limitation of the study is that we were unable to determine the precise mechanism by which arginine shortage leads to the accumulation of PCNA on nascent strands. The results of our immunoprecipitation assay may not provide conclusive evidence for the promotion of K63-linked polyubiquitination of chromatin-bound PCNA in response to arginine shortage. Further research is needed to understand the role of ZRANB3 in this process and to clarify the transition of unmodified PCNA to the K63-linked polyubiquitinated state. Additionally, the impact of newly synthesized histone H3 in DNA replication during arginine shortage remains an area of future investigation.
STAR★METHODS
RESOURCE AVAILABILITY
Lead contact
Further information and requests for resources and reagents should be directed to and will be fulfilled by the lead contact, David Ann (dann@coh.org).
Materials availability
This study did not generate new unique reagents.
Data and code availability
All data reported in this paper will be shared by the lead contact upon request.
This paper does not report original code.
Any additional information required to reanalyze the data reported in this paper is available from the lead contact upon request.
EXPERIMENTAL MODEL AND SUBJECT DETAILS
Animals
Animal experiments were approved by the Institutional Animal Care and Use Committee at City of Hope. MDA-MB-231 or BT-549 cells (3.3 × 3 105) were injected into the mammary fat pads of 6-week-old female NOD.Cg-PrkdcscidIl2rgtm1Wjl/SzJ (NSG) mice. All mice were assigned to the same experimental group. Xenograft tumors were formalin-fixed, paraffin embedded (FFPE) sections were prepared by City of Hope Pathology Core.
Cell lines
Human breast cancer MDA-MB-231, MDA-MB-468, MCF7, BT-549, Hs578T and embryonic kidney HEK-293T cells were purchased from American Type Culture Collection and cultured with Dulbecco’s Modified Eagle’s Medium (DMEM, Cat# 10–013-CV, Corning) containing fetal bovine serum (10%, Cat# 10437028, ThermoFisher) and 1X antibiotic and antimycotic solution (Cat# 30–004-CI, Corning).
METHOD DETAILS
Treatment
Arginine-free media for BT-549, HEK-293T, Hs578T, MCF7, MDA-MB-231 and MDA-MB-468 cells were prepared from DMEM for SILAC (Cat# 88364, ThermoFisher) supplemented with dialyzed FBS (10%, Cat# 26400–044, ThermoFisher), 1X antibiotic and antimycotic solution (Cat# 30–004-CI, Corning), and L-lysine (146.2 μg/ml, Cat# L-9037, Millipore Sigma). For arginine removal, cells were washed with warm arginine-free medium once then incubated in arginine-free medium for indicated times.
Lentivirus preparation
pMD2.G (1 μg; Addgene, Cat# 12259) and psPAX2 (2 μg; Addgene, Cat# 12260) were co-transfected with lentiviral expression construct (3 μg) using Lipofectamine 2000 (14 μL; Life Technologies, 11668–019) into HEK-293T cells to generate lentivirus. Medium was replaced at 24 h post transfection, and viral medium was collected at 72 h post transfection. Viral medium was filtrated with 0.45 μm filter (Millipore Sigma, Cat# SLHAM33SS) and concentrated with Ultra-15 centrifugal filter device (Millipore Sigma, Cat# UFC903024). Concentrated lentivirus-containing medium was recovered and used to transform desired cell lines in the presence of polybrene (10 μg/mL, Millipore Sigma, Cat# TR-1003-G). Clonal cell lines were selected by pre-determined concentration of puromycin (0.5, 1 and 2 μg/ml; Sigma, Cat# P8833) for respective cell line and overexpression of desired protein was confirmed by Western blotting.
Western blotting
Total cell lysates were prepared by lysing cells with sample buffer and analyzed by Western blotting. Following polyacrylamide gel electrophoresis, separated proteins were transferred to polyvinylidene difluoride (PVDF, Millipore) membranes. In general, the PVDF membranes with transferred proteins were blocked with nonfat milk (5%) in PBST (137 mM NaCl, 10 mM phosphate, 2.7 mM KCl, 0.1% Tween 20, pH 7.4) for 40 mins. Proteins of interest were identified with antibodies described in key resources table. Following incubation with primary antibodies (in PBST with non-fat milk (5%)) at 4 °C overnight, PVDF membranes were washed with PBST 3 times, and then incubated with horse radish peroxidase-linked anti-mouse or anti-rabbit secondary immunoglobulin antibodies (1:10000; Cat# 7076S, Cat# 7074S Cell signaling) in PBST for 40 mins, Blots were developed with homemade ECL (2.5 mM luminol, 044 mM p-coumaric acid, 0.1 M Tris-HCl pH 8.5, 0.1% hydrogen peroxide in 100 ml distilled water). ChemiDoc Touch imaging system (Bio-Rad) was used to visualize signals. The captured images were analyzed with Image Lab Software (Bio-Rad, version 5.2.1).
KEY RESOURCES TABLE
REAGENT or RESOURCE | SOURCE | IDENTIFIER |
---|---|---|
Antibodies | ||
| ||
Rabbit polyclonal anti-Ki67 | abcam | Cat#ab15580; RRID:AB_443209 |
Rabbit monoclonal anti-CD31 | Cell signaling | Cat#77699; RRID:AB_2722705 |
Mouse monoclonal anti-CD44 | Cell Signaling | Cat#3570; RRID:AB_2076465 |
Rabbit polyclonal anti-acetyl-histone H4 (Lys5) | Millipore | Cat#07–327; RRID:AB_310523 |
Rabbit polyclonal anti-acetyl-histone H4 (Lys12) | Millipore | Cat#07–595; RRID:AB_310740 |
Rabbit polyclonal anti-acetyl-histone H4 (Lys16) | Millipore | Cat#07–329; RRID:AB_310525 |
Mouse monoclonal anti-Actin | Millipore | Cat#MAB1501; RRID:AB_2223041 |
Rabbit monoclonal anti-ATF4 | Cell Signaling | Cat#11815; RRID:AB_2616025 |
Mouse monoclonal anti-Chk1 | Cell Signaling | Cat#2345; RRID:AB_10693648 |
Rabbit polyclonal anti-cyclin A | Santa Cruz | Cat#sc-751; RRID:AB_631329 |
Mouse monoclonal anti-cyclin B1 | Enzo | Cat#ADI-KAm-CC195; RRID:AB_10615085 |
Rabbit polyclonal anti-eIF2α | Cell Signaling | Cat#9722; RRID:AB_2230924 |
Rabbit polyclonal anti-γH2AX | Bethyl | Cat#A300–081; RRID:AB_203288 |
Mouse monoclonal anti-histone H3 | Active Motif | Cat#39763; RRID:AB_2650522 |
Rabbit monoclonal anti-histone H4 | abcam | Cat#ab177840; RRID:AB_2650469 |
Mouse monoclonal anti-lamin A/C | Cell Signaling | Cat#4777; RRID:AB_10545756 |
Rabbit polyclonal anti-BRCA2 | abcam | Cat#ab27976; RRID:AB_2067760 |
Rabbit polyclonal anti-PCNA | Santa Cruz | Cat#sc-7907; RRID:AB_2160375 |
Rabbit polyclonal anti-phospho-eIF2α | Cell Signaling | Cat#9721; RRID:AB_330951 |
Rabbit monoclonal anti-phospho-Chk1 (Ser296) | abcam | Cat#ab79758; RRID:AB_2244917 |
Rabbit monoclonal anti-phospho-Chk1 (Ser345) | Cell Signaling | Cat#2348; RRID:AB_331212 |
Rabbit polyclonal anti-phospho-histone H3 (Ser10) | Millipore Sigma | Cat#06–570; RRID:AB_310177 |
Mouse monoclonal anti-RPA32/RPA2 | abcam | Cat#ab2175; RRID:AB_302873 |
Rabbit polyclonal anti-phospho-RPA32 (Ser33) | Bethyl | Cat#A300–246A; RRID:AB_2180847 |
Rabbit monoclonal anti-trimethyl-histone H3 (Lys4) | Cell Signaling | Cat#9751; RRID:AB_2616028 |
Mouse monoclonal anti-trimethyl-histone H3 (Lys9) | Active Motif | Cat#61013; RRID:AB_2687870 |
Rabbit monoclonal anti-ubiquityl-PCNA (Lys164) | Cell Signaling | Cat#13439; RRID:AB_2798219 |
Rabbit monoclonal anti-ubiquitin Lys63-specific | Cell Signaling | Cat#5621; RRID:AB_10827985 |
Rabbit polyclonal anti-SMARCA3 (HLTF) | Bethyl | Cat# A300–230A; RRID:AB_2117307 |
Rabbit polyclonal anti-Rad6 | abcam | Cat#ab31917; RRID:AB_777604 |
Rabbit polyclonal anti-Rad18 | abcam | Cat#ab188235; RRID:AB_2935813 |
Rabbit polyclonal anti-RFC5 | Proteintech | Cat#10385; RRID:AB_2178603 |
Rabbit polyclonal anti-RFC2 | Proteintech | Cat#10410; RRID:AB_2285035 |
Rabbit polyclonal anti-BRD4 | Proteintech | Cat#28486; RRID:AB_2918170 |
Rabbit polyclonal anti-ATAD5 | abcam | Cat#ab72111; RRID:AB_1209390 |
Mouse monoclonal anti-flag | Proteintech | Cat#66008–3-Ig; RRID:AB_2749837 |
Rabbit monoclonal anti-FLASH/CASP8AP2 | Cell Signaling | Cat#43339; RRID:AB_2935814 |
Rabbit polyclonal anti-SLBP | Millipore Sigma | Cat#HPA019254; RRID:AB_1856951 |
Mouse monoclonal anti-SHPRH | Santa Cruz | Cat#sc-514395; RRID:AB_2935815 |
Mouse monoclonal anti-SMARCAL1 | Santa Cruz | Cat#sc-376377; RRID:AB_10987841 |
Rabbit polyclonal anti-ZRANB3 | Novus Biologicals | Cat#NBP2–93301; RRID:AB_2935816 |
Rabbit polyclonal anti-phospho-Akt (Ser473) | Cell Signaling | Cat#9271; RRID:AB_329825 |
Rabbit polyclonal anti-Akt | Cell Signaling | Cat#9272; RRID:AB_329827 |
Rat monoclonal anti-BrdU (BU1/75) | abcam | Cat#ab6326; RRID:AB_305426 |
Mouse monoclonal anti-BrdU (B44) | BD Biosciences | Cat#347580; RRID:AB_10015219 |
| ||
Chemicals, peptides, and recombinant proteins | ||
| ||
L-azidohomoalanine | Thermo Fisher | Cat#C10102 |
L-arginine | Millipore Sigma | Cat# A8094 |
Thymidine 5’-monophosphate disodium salt hydrate | Millipore Sigma | Cat# T7004 |
Hydroxyurea | MP Biomedicals | Cat#102023 |
(+)-Camptothecin | ACROS Organics | Cat#AC276721000 |
BAY 1895344 | Selleckchem | Cat#S8666 |
Rabusertib | Selleckchem | Cat#S2626 |
AOH1996 | Gu et al.61 | N/A |
Recombinant human His-tagged histone H4 produced in E. coli | Active Motif | Cat#31493 |
| ||
Critical commercial assays | ||
| ||
FiberPrep | Genomic Vision | Cat#EXT-001 |
Combicoverslips | Genomic Vision | Cat#COV-002-RUO |
| ||
Experimental models: Cell lines | ||
| ||
Human: MDA-MB-231 | ATCC | HTB-26 |
Human: MDA-MB-468 | ATCC | HTB-132 |
Human: MCF7 | ATCC | HTB-22 |
Human: BT-549 | ATCC | HTB-122 |
Human: Hs578T | ATCC | HTB-126 |
Human: HEK-293T | ATCC | CRL-11268 |
Human: HEK-293T-PCNA-WT | Thakar et al.56 | N/A |
Human: HEK-293T-PCNA-K164R | Thakar et al.56 | N/A |
| ||
Experimental models: Organisms/strains | ||
| ||
Mouse:NOD.Cg-PrkdcscidIl2rgtm1Wjl/SzJ(NSG) | The Jackson Laboratory | RRID: IMSR_JAX:005557 |
| ||
Oligonucleotides | ||
| ||
siHLTF | Sigma Aldrich | Cat#SASI_Hs01_00162286 and SASI_Hs01_00029458 |
siFLASH | Sigma Aldrich | Cat#SASI_Hs02_00343647 and SASI_Hs01_00152963 |
siSLBP | Sigma Aldrich | Cat#SASI_Hs01_00147022 and SASI_Hs01_00147023 |
siSHPRH | Sigma Aldrich | Cat#SASI_Hs02_00310973 and SASI_Hs02_00310974 |
siSMARCAL1 | Sigma Aldrich | Cat#SASI_Hs02_00328762 and SASI_Hs02_00328763 |
siZRANB3 | Sigma Aldrich | Cat#SASI_Hs01_00109802 and SASI_Hs01_00109803 |
H4C5 forward primer 5’-GGTGTCAAGCGCATTTCTGGTC-3′ | This paper | N/A |
H4C5 reverse primer 5’-CGTAGACCACATCCATCGCTGT-3′ | This paper | N/A |
| ||
Recombinant DNA | ||
| ||
pMD2.G | Addgene | Cat#12259 |
psPAX2 | Addgene | Cat#12260 |
CMV-FLAG-SV40-eGFP-IRES-puromycin | GeneCopoeia | EX-NEG-Lv203 |
CMV-FLAG-PCNA-SV40-eGFP-IRES-puromycin | GeneCopoeia | EX-B0066-Lv203 |
pHAGE-HLTF-WT | Taglialatela et al.58 | N/A |
pHAGE-HLTF R71E Hiran mutant | Taglialatela et al.58 | N/A |
| ||
Software and algorithms | ||
| ||
ImageJ | NIH | https://imagej.nih.gov/ij/ |
FCS Express | De Novo software | https://denovosoftware.com/ |
GraphPad Prism 7.0 software | GraphPad Prism Software Inc. | https://www.graphpad.com/ |
Image Lab Software 5.2.1 | Bio-rad | https://www.bio-rad.com/en-us/product/image-lab-software?ID=KRE6P5E8Z |
Subcellular fractionation
Equal number (3×106) of cells were harvested in ice-cold nuclear extraction buffer (NEB) (400 μL, 10 mM HEPES pH 7.9, 10 mM KCl, 1.5 mM MgCl2, 0.34 M sucrose, 10% glycerol, 1 mM dithiothreitol, 0.1% Triton X-100) containing protease/phosphatase cocktail inhibitors (Thermo Scientific, Cat# 78446) and incubated on ice for 5 min. 300 μl of whole cell lysates were centrifuged at 1300 × g for 4 min at 4°C. Supernatants were collected as cytoplasmic fractions and nuclei pellets were washed with NEB twice and re-suspended in 300 μl of NEB. Total cell lysates (remaining 100 μl of whole cell lysates), cytoplasmic and nuclear fractions were mixed with 4× Laemli buffer (v/v) and analyzed with desired antibodies by Western blotting.
Cell cycle analyses (flow cytometry)
For double thymidine block, cells were treated with thymidine (2.5 mM; Cat# 194754, MP Biomedicals) for 16 h and washed twice with PBS. Released cells were incubated in complete medium for 8 h and blocked with the 2nd-thymidine treatment. Following 2nd-thymidine block (16 h), cells were washed twice with PBS and released in regular medium for analyses. To enrich mitotic cells, cells were treated with thymidine (2.5 mM, 24 h) and released in regular medium. At 3 h post-releasing in regular medium, cells were treated with nocodazole (100 ng/ml, Cat# M1404, Millipore Sigma) for 12 h and released in regular medium for analysis. For BrdU pulse-labeling, cells were pulse-labeled with BrdU (10 μM, Cat# B9285, Millipore Sigma, 30 min) before or after treatment under indicated conditions. Cells were collected and fixed in ethanol (70%) at −20°C overnight. To analyze BrdU incorporated DNA, cells were washed with PBS to remove ethanol and denatured with HCl (2.5 M, 20 min). Cells were washed twice with PBS then stained with FITC conjugated anti-BrdU antibody (BioLegend, Cat# 364104) (1:100) at 37°C for 1 h. After incubation, cells were washed 3 times with PBS and DNA content were visualized by DAPI (1 μM; Cat# D9542, Millipore Sigma, I h) staining at room temperature. Samples were monitored by Attune NxT flow cytometer (Thermofisher) and results were analyzed and quantified by FCS Express (De Novo software). At least 50000 events were collected and 3 independent experiments were performed and analyzed.
DNA combing assay
Equal number (5×104) of MDA-MB-231 or BT-549 cells were seeded in 60-mm dishes and incubated overnight. Cells were pulse-labeled with CldU (50 μM; Cat# 105478, MP Biomedicals), IdU (250 μM; Cat# 54–42-2, ACROS Organics) and treated as indicated. Cells were washed with PBS and collected. DNA plugs were generated using FiberPrep (Cat# EXT-001, Genomic Vision) and DNA combing was performed using Combicoverslips (Cat# COV-002-RUO, Genomic Vision), according to the manufacturer instructions. DNA combed coverslips were dehydrated with 70%, 90%, and 100% ethanol for 1 min each. DNA combed coverslips were denatured by 0.5 M NaOH plus 1 M NaCl solution for 8 min at room temperature then washed with PBS. The coverslips were dehydrated again with 70%, 90%, and 100% ethanol for 1 min each and air-dried at room temperature for 10 min. Following blocking with 5% BSA in PBS for 5 mins at 37°C in a humidity chamber, coverslips were incubated with primary antibodies (rat anti-BrdU (BU1/75) (1:500; recognizing CldU; Cat# NB500–169, Novus) and mouse anti-BrdU (B44) (1:500; recognizing ldU; Cat# 347580, BD) with 5% BSA in PBS) for 2 h at 37°C in a humidity chamber. Following incubation, coverslips were washed 3 times with PBST (0.05% Tween 20) then dehydrated with 70%, 90%, and 100% ethanol stepwise. Coverslips were incubated in PBS containing both goat anti-rat 647 secondary antibody (1:500; Cat# A-21247, ThermoFisher) and goat anti-mouse 488 secondary antibody (1:500; Cat# A-11001, ThermoFisher) for 1 h at 37°C in a humidity chamber. After incubation, coverslips were washed 3 times with PBST (0.05% Tween 20) then dehydrated with 70%, 90%, and 100% ethanol stepwise. Coverslips were mounted with prolong diamond antifade mountant (Cat# P36970, ThermoFisher), and sealed with nail polish. DNA replication tracts were visualized with a fluorescence microscope (Observer II, Zeiss) and ImageJ was used for measuring the length of individual labeled DNA tract.
Analysis of newly synthesized proteins with AHA labeling
Double thymidine synchronized MDA-MB-231 cells were released in L-Azidohomoalanine (AHA, 200 μM, Cat# C10102, ThermoFisher) containing full or arginine-free medium and cells were harvest in indicated time. Cells were collected and washed with PBS for 3 times. AHA-labeled newly synthesized proteins were analyzed by electrophoresis and detected with anti-biotin antibody (Cat# ab201341, abcam). For visualizing total proteins, whole cell lysates were analyzed by electrophoresis and gels were stained with Coomassie blue and imaged on ChemiDoc Touch imaging system (Bio-Rad). The captured images were analyzed with Image Lab Software (Bio-Rad, version 5.2.1).
To assess levels of specific AHA-labeled proteins, MDA-MB-231 cells were labeled with AHA as described above and lysed in RIPA buffer (1% sodium deoxycholate, 50 mM HEPES, 150 mM NaCl, 1% NP-40, 0.1% SDS, 2.5 mM MgCl2, 10 mM sodium glycerophosphate, 10 mM sodium biphosphate pH 7.2) containing protease and phosphatase inhibitors (Thermo Scientific, Cat# 78446). Approximately 1.5 mg of proteins from each sample was reduced with TCEP (5 mM, room temperature for 10 min, Cat# 580561, Millipore Sigma) then alkylated with chloroacetamide (20 mM, room temperature for 15 min, Cat# C8625G, TCI America). Alkylated proteins were precipitated with methanol/chloroform/H2O and washed 2 times with methanol to remove residual AHA. The precipitated proteins were suspended in suspension buffer (2% SDS, 50 mM HEPES, 150 mM NaCl, pH 7.2, 2.5 mM TCEP) and sonicated briefly with a probe sonicator. Click chemistry reaction was performed with biotin-PEG4-alkyne (100 μM, Cat# 764213, Millipore Sigma), CuSO4 (1 mM, Cat# 422870050, ACROS organics), and L-ascorbate (1 mM, Cat# 352680050, ACROS organics), THPTA ligand (100 μM, Cat# 762342, Millipore Sigma) and mixed on a rotor at room temperature for 2 h. Biotin-alkyne-reacted proteins were precipitated with methanol/chloroform/H2O and washed 2 times with methanol to remove residual biotin-alkyne. The precipitated proteins were suspended in suspension buffer (200 μl, 2% SDS, 50 mM HEPES, 150 mM NaCl, pH 7.2, 2.5 mM TCEP) and sonicated briefly with a probe sonicator and diluted to 600 μl with RIPA buffer. High-capacity streptavidin agarose (20 μl, Cat# 20359, ThermoFisher Scientific) was added to each sample and mixed on a rotor at room temperature overnight. The beads were spun down and washed sequentially twice with RIPA buffer (1 ml), twice with 1M KCl, 0.1 M Na2CO3, 2 M urea in 50 mM HEPES buffer pH 7.5 (1 ml), twice with RIPA buffer (1 ml) and twice with ddH2O (1 ml). After washing, samples were mixed with 2× sample buffer and boiled at 100°C for 15 min to extract the proteins. Elution was analyzed by Western blotting with indicated antibodies.
Protein transfection
Human recombinant proteins were introduced into cells with ProteoJuice protein transfection reagent (Cat# 71281–3, MilliporeSigma) according to manufacturer’s instruction. Briefly, recipient cells were seeded in a 6-well plate, washed with PBS for 3 times and maintained in OPTI-MEM. Approximately 5 μg of recombinant human His-tagged histone H4 produced in E. coli (Cat# 31493, Active Motif) was mixed with 5 ml of ProteoJuice in 100 μl of OPI-MEM and incubated at room temperature for 20 min. After 20 min of incubation, 900 μl of OPI-MEM was added to make a 1000 μl of mixture. After aspirating the OPTI-MEM from the cells, cells were incubated with the transfection mixture at 37°C (5% CO2) for 2–3 h. After incubation, cells were washed 3 times with PBS and incubated in full or arginine-free medium as indicated prior to harvesting or subjected to DNA combing assays.
Immunoprecipitation
Stable expressing FLAG or FLAG-tagged PCNA HEK293T cells were overexpressed with Myc-tagged ubiquitin plasmid (gift from Dr. Hsiu-Ming Shih) for 24 h prior to the indicated treatments. After treatment, cells were harvested by modified RIPA buffer (50mM Tris-HCl pH7.5, 150 mM NaCl, 5 mM EDTA, 0.1% Triton X-100, and 0.1% NP-40) with protease/phosphatase inhibitor (Cat# 78446, ThermoFisher) and incubated in 4°C on rotating mixer for 15 min. Cell debris were removed by 12000 rpm centrifuge for 10 min in 4°C and supernatants were mixed with anti-DYKDDDDK magnetic agarose (Cat# A36797, ThermoFisher) in 4°C overnight on rotating mixer according to manufacturer’s protocol. Magnetic beads were captured with magnetic rack and washed with modified RIPA buffer for 3 times. Samples were eluted by adding 2X SDS-PAGE Sample Buffer and incubated at 100°C for 15 min. Levels of specific proteins were analyzed by Western blotting. To knockdown HLTF, cells were transfected with siRNA targeting HLTF 24 h prior to Myc-ubiquitin overexpression.
Accelerated native iPOND (aniPOND)
A rapid and accelerated method modified from isolation of protein on nascent DNA (iPOND) was performed following previous study.49 In short, MDA-MB-231 cells were pulse labeled by EdU (10 μM) for indicated time along with indicated treatments. Cells were harvested and nuclear extracts were prepared by using the nuclear extraction buffer as described.49 Click reaction was performed to conjugate biotin-azide (Cat# 1265, Click Chemistry Tools) to EdU labeled nascent DNA and sonicated (Bioruptor Pico sonication device). Samples were centrifuged and supernatant was collected and mixed with streptavidin-coated beads (Cat# 20359, ThermoFisher) to pull down biotin-EdU labeled nascent DNA. Beads were centrifuged and washed for 3 times and proteins were eluted by adding 2× sample buffer and boiled at 100˚C for 15 min. Western blotting was performed to analyze the proteins on nascent DNA. Cells without Edu labeling (EdU(−)) served as negative control. To compensate for the reduced EdU-labeled track in perturbed cells, the EdU labeling time of -R, HU, and CPT treated cells were extended to 25 min while +R cells were labeled for 15 min in Figure 5B.
Spheroid formation
Equal number (1.2×103) of cells were seeded in ultralow attachment 96-well plates (Cat# CLS3474, Corning) and incubated for 72 h in a 37°C humidity chamber to form spheroids. Spheroids were washed with medium containing indicated chemical twice then incubated in medium containing same chemical for 2 h at 37°C. After 2 h of incubation, spheroids were washed with complete medium three times and then cultured in complete medium for 8 days. Spheroids were visualized at day 0 and day 8 post-treatment by using Cytation 5 imaging reader (BioTek) and ImageJ was used to process the areas of spheroids.
Cell viability assay
Equal number (5×103) cells were seeded in 96-well plate and cultured under indicated conditions for different time periods. Acid phosphatase (ACP) assay was performed to measure cell viability. Cells were lysed with ACP1 solution (100 μl, 0.1 M sodium acetate, 0.1% Triton X-100, pH 5.7) containing p-Nitrophenyl Phosphate (12 mM) per well and kept in a 37°C incubator. Following 30 min of incubation, the reaction was stopped with NaOH (10 μl, 1 M) and absorbance at 410 nm was recorded using Synergy H1 hybrid multi-mode reader (BioTek). The value of each group measured at day 0 was set as 1.
Immunofluorescence staining
To visualize H4K12ac and EdU incorporation in BT-549 spheroids, equal number (8×103) cells were seeded in ultralow attachment 96-well plates for 72 h. Spheroids were incubated in EdU (10 μM)- and Hoechest33342 (10 μM)-containing medium for 1 h, collected, fixed with paraformaldehyde (4%) for 10 min and washed twice with PBS. To generate blocks for cryosection, spheroids were incubated in sucrose (15%) for 30 min, followed by sucrose (30%) for 30 min, and sucrose (30%) containing optimal cutting temperature compound (50%, O.C.T., Tissue-TeK, Cat# 4583) for 30 min. Spheroids were then embedded in O.C.T. (100%) in cryomold and 8 μm thick sections were cryosectioned using a cryostat (Leica, CM3050S). EdU incorporation was stained according to manufacturer’s protocol (EdU kit for imaging, Cat# C10640, ThermoFisher). Following washing with PBS 3 times and blocking with BSA (4%) in 0.2% Tween-20 in PBS (PBST (0.2%)) for 30 min in a 37°C humidity chamber, slides were stained with an anti-H4K12ac antibody (1:500; Cat# 07–595, Millipore Sigma) in PBST with BSA (2%) slides in a 37°C humidity chamber for 2 h. Slides were washed 3 times with PBST (0.2%) then incubated with a goat anti-rabbit 488 secondary antibody (1:500; Cat# A-11034, ThermoFisher) in PBST with BSA (2%) in a 37°C humidity chamber. Slides were washed 3 times with PBST (0.2%) and mounted with prolong diamond antifade mountant (Cat# P36970, ThermoFisher), and then sealed with nail polish.
To assess H4K12ac signals in BT-549 spheroids, equal number (8×103) cells were seeded in ultralow attachment 96-well plates and incubated with complete medium in a 37°C humidity chamber for 72 h to form spheroids. Cryosectioning and immunofluorescence staining were performed as described above. Briefly, cryosections were probed with an anti-H4K12ac antibody (1:500; Cat# 07–595, Millipore Sigma) and followed with incubation with a goat anti-rabbit 488 secondary antibody (1:500; Cat# A-11034, ThermoFisher). H4K12ac-positive cells were visualized using a fluorescence microscope (Observer II, Zeiss; LSM 700 Confocal microscope, Zeiss). ImageJ was used for processing the images in each cryosection. For visualizing histone H4K12ac signals in Hs578T or MCF7 spheroids, 3×104 cells were seeded in ultralow attachment 96-well plates and incubated in complete medium to form spheroids in a 37°C humidity chamber. After 5 days of incubation, tumor spheroids were collected and fixed with paraformaldehyde (4%) for 10 min and washed twice with PBS. Cryosectioning and immunofluorescence staining were performed as described above.
To assess H4K12ac and γH2AX signals in BT-549 spheres, 8×103 cells were seeded in ultralow attachment 96-well plates and incubated in complete medium for 72 h to form spheres. Cryosection and immunofluorescence staining were performed as described above. Briefly, sections were probed with an anti-H4K12ac antibody (1:500; Cat# 07–595, Millipore Sigma) or an anti-H2A.X phosphor (Ser139) antibody conjugated with Alexa Fluor 647 (1:500; Cat# 613408, BioLegend). For H4K12ac visualization, after primary antibody incubation, sections were incubated with a goat anti-rabbit 488 secondary antibody (1:500; Cat# A-11034, ThermoFisher). Samples were imaged by a fluorescence microscope (Observer II, Zeiss; LSM 700 Confocal microscope, Zeiss), and ImageJ was used for evaluating the overall intensity of indicated signals in each sphere.
Chromatin-bound RPA and γH2AX detection
MDA-MB-231 cells were pulse labeled with EdU (10 μM) for 1 h then treated with full, arginine-free, or medium containing HU (2 mM), CPT (2 μM), ATR inhibitor (100 nM) or Chk1 inhibitor (1 μM) for 4 h. Cells were trypsinized, washed with PBS and extracted with cold 0.2% Triton X-100 in PBS for 7 min on ice. After extraction, cells were washed with PBS and fixed with 4% paraformaldehyde in PBS (Cat# BM-155, Boston BioProducts). Fixed cells were blocked with 0.1% BSA in PBS and RPA signals were stained followed by DAPI staining. Percentage of RPA-positive or γH2AX-positive cells in S-phase (EdU-positive) was analyzed by Attune NxT flow cytometer (Thermofisher) and results were quantified by FCS Express (De Novo software).
Immunohistochemistry
Animal experiments were approved by the Institutional Animal Care and Use Committee at City of Hope. Xenograft tumors were formalin-fixed, paraffin embedded (FFPE) sections were prepared by City of Hope Pathology Core. For detecting the proteins of interest in xenograft tumors, anti-histone 4 lysine 12 acetylation (H4K12ac) antibody (1:500; Cat# ab177793, abcam), anti-histone 4 lysine 5 acetylation (H4K5ac) antibody (1:4000; Cat# ab232507, abcam), anti-Ki67 antibody (1:200, Cat# ab15580, abcam), and anti-CD31 antibody (1:100, Cat# 77699, Cell signaling), respectively, were used for immunohistochemical staining by the pathology core at City of Hope. IHC stain was performed on Ventana Discovery Ultra (Ventana Medical Systems, Roche Diagnostics, Indianapolis, USA) IHC automated stainer. Briefly, tissue samples were sectioned at a thickness of 5 μm and mounted on positively charged glass slides. The sections were deparaffinized and rehydrated. The endogenous peroxidase activity was blocked before antigen retrieval. The antigens were sequentially detected with primary antibody incubation and heat inactivation was performed to prevent antibody cross-reactivity between the same species. Following each primary antibody, DISCOVERY anti-Rabbit HQ and DISCOVERY anti-HQ-HRP were incubated. The positive signals were visualized with DISCOVERY ChromoMap DAB and DISCOVERY Purple Kit, respectively, counterstained with hematoxylin (Ventana) and then coverslipped. Slides were scanned by bright-field microscopy and 4 fields of peripheral and non-necrotic core regions were photograph in each section. H4K5ac and H4K12ac signal intensities were quantified by ImagePro Premier 9.0.
RNA extraction and qRT-PCR
Total RNA was extracted from cells using the Quick-RNA MiniPrep (Zymo Research, Cat# R1055). qRT-PCR was performed by as described previously.5 In brief, the RNA concentration and purity were analyzed by Synergy H1 hybrid multi-mode reader (BioTek). Complementary DNA (cDNA) was synthesized with 1 μg of total RNA by using the iScript cDNA Synthesis Kit (Bio-Rad, Cat# 1708891). cDNA was diluted 20 times and real-time quantitative PCR was performed by using iTaq Universal SYBR Green Supermix (BioRad, Cat# 1725122) with CFX-96 Real-Time PCR Detection System (BioRad). The primers used in this study targeting histone H4C5 are, forward: 5’-GGTGTCAAGCGCATTTCTGGTC-3’, and reverse: 5’-CGTAGACCACATCCATCGCTGT-3’. qRT-PCR data from 3 biological replicates were calculated using 2−ΔΔCt method and normalized to 18S RNA.74 The mean mRNA abundance in the first time point (0 h) was set as 1.
QUANTIFICATION AND STATISTICAL ANALYSIS
Data are presented as mean and standard error of mean (Mean ± SEM). Statistical analyses were performed using GraphPad Prism 7.0 software (GraphPad Prism Software Inc., San Diego, CA). Normal distribution was confirmed using Shapiro-Wilk normality test before performing statistical analyses. For normally distributed data, comparison between two means were assessed by unpaired two-tailed Student’s t test and that between three or more groups were evaluated using one-way analysis of variance followed by Tukey’s post hoc test. A p-value of <0.05 was considered statistically significant.
Supplementary Material
Highlights.
Arginine shortage halts H4 histone synthesis, stalling DNA replication
Fork stalling creates vulnerability of nascent DNA strands to DNA2-mediated degradation
Cells respond to arginine deprivation by HLTF-mediated K63-polyubiquitination of PCNA
HLTF and PCNA polyubiquitylation protect arginine-starved cells from DNA damage agents
ACKNOWLEDGMENTS
We thank Drs. Mark LaBarge, Lei Jiang, Shao-Chun Wang, Anita K. Hopper, Chathurani S. Jayasena, and members of Dr. Ann’s, Stark’s, and Kung’s laboratories for helpful discussion on the manuscript and Dr. Glenn Manthey for editing the manuscript. Immunohistochemistry staining was performed at the Pathology Core of City of Hope by Dr. Aimin Li. This work is supported by funds from the National Institutes of Health (R01CA220693 to D.K.A.; R01CA256989, R01CA197506, and R01CA240392 to J.M.S.; R01CA197774 to A.C.; R01GM134681 to G.-L.M.; T32CA186895 to A.A.K.; T32CA221709–01 for A.K.; and P30CA33572 for City of Hope Core Facilities). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.
Footnotes
SUPPLEMENTAL INFORMATION
Supplemental information can be found online at https://doi.org/10.1016/j.celrep.2023.112296.
DECLARATION OF INTERESTS
The authors declare no competing interests.
REFERENCES
- 1.Boroughs LK, and DeBerardinis RJ (2015). Metabolic pathways promoting cancer cell survival and growth. Nat. Cell Biol 17, 351–359. 10.1038/ncb3124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Lama-Sherpa TD, and Shevde LA (2020). An Emerging regulatory role for the tumor microenvironment in the DNA damage response to double-strand breaks. Mol. Cancer Res 18, 185–193. 10.1158/1541-7786.MCR-19-0665. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Chen CL, Hsu SC, Ann DK, Yen Y, and Kung HJ (2021). Arginine signaling and cancer metabolism. Cancers 13, 3541. 10.3390/cancers13143541. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Lee JS, Adler L, Karathia H, Carmel N, Rabinovich S, Auslander N, Keshet R, Stettner N, Silberman A, Agemy L, et al. (2018). Urea cycle dysregulation generates clinically relevant genomic and biochemical signatures. Cell 174, 1559–1570.e22. 10.1016/j.cell.2018.07.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Cheng CT, Qi Y, Wang YC, Chi KK, Chung Y, Ouyang C, Chen YR, Oh ME, Sheng X, Tang Y, et al. (2018). Arginine starvation kills tumor cells through aspartate exhaustion and mitochondrial dysfunction. Commun. Biol 1, 178. 10.1038/s42003-018-0178-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Pan M, Reid MA, Lowman XH, Kulkarni RP, Tran TQ, Liu X, Yang Y, Hernandez-Davies JE, Rosales KK, Li H, et al. (2016). Regional glutamine deficiency in tumours promotes dedifferentiation through inhibition of histone demethylation. Nat. Cell Biol 18, 1090–1101. 10.1038/ncb3410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Lee SW, Zhang Y, Jung M, Cruz N, Alas B, and Commisso C (2019). EGFR-pak signaling selectively regulates glutamine deprivation-induced macropinocytosis. Dev. Cell 50, 381–392.e5. 10.1016/j.devcel.2019.05.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Wu G (2009). Amino acids: metabolism, functions, and nutrition. Amino Acids 37, 1–17. 10.1007/s00726-009-0269-0. [DOI] [PubMed] [Google Scholar]
- 9.Poillet-Perez L, Xie X, Zhan L, Yang Y, Sharp DW, Hu ZS, Su X, Maganti A, Jiang C, Lu W, et al. (2018). Autophagy maintains tumour growth through circulating arginine. Nature 563, 569–573. 10.1038/s41586-018-0697-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Qiu F, Chen YR, Liu X, Chu CY, Shen LJ, Xu J, Gaur S, Forman HJ, Zhang H, Zheng S, et al. (2014). Arginine starvation impairs mitochondrial respiratory function in ASS1-deficient breast cancer cells. Sci. Signal 7, ra31. 10.1126/scisignal.2004761. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Chen CL, Hsu SC, Chung TY, Chu CY, Wang HJ, Hsiao PW, Yeh SD, Ann DK, Yen Y, and Kung HJ (2021). Arginine is an epigenetic regulator targeting TEAD4 to modulate OXPHOS in prostate cancer cells. Nat. Commun 12, 2398. 10.1038/s41467-021-22652-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Hsu SC, Chen CL, Cheng ML, Chu CY, Changou CA, Yu YL, Yeh SD, Kuo TC, Kuo CC, Chuu CP, et al. (2021). Arginine starvation elicits chromatin leakage and cGAS-STING activation via epigenetic silencing of metabolic and DNA-repair genes. Theranostics 11, 7527–7545. 10.7150/thno.54695. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Bohnsack BL, and Hirschi KK (2004). Nutrient regulation of cell cycle progression. Annu. Rev. Nutr 24, 433–453. 10.1146/annurev.nutr.23.011702.073203. [DOI] [PubMed] [Google Scholar]
- 14.Cuyàs E, Corominas-Faja B, Joven J, and Menendez JA (2014). Cell cycle regulation by the nutrient-sensing mammalian target of rapamycin (mTOR) pathway. Methods Mol. Biol 1170, 113–144. 10.1007/978-1-4939-0888-2_7. [DOI] [PubMed] [Google Scholar]
- 15.Liu GY, and Sabatini DM (2020). mTOR at the nexus of nutrition, growth, ageing and disease. Nat. Rev. Mol. Cell Biol 21, 183–203. 10.1038/s41580-019-0199-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.MacAlpine DM, and Almouzni G (2013). Chromatin and DNA replication. Cold Spring Harbor Perspect. Biol 5, a010207. 10.1101/cshperspect.a010207. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Hammond CM, Strømme CB, Huang H, Patel DJ, and Groth A (2017). Histone chaperone networks shaping chromatin function. Nat. Rev. Mol. Cell Biol 18, 141–158. 10.1038/nrm.2016.159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Günesdogan U, Jäckle H, and Herzig A (2014). Histone supply regulates S phase timing and cell cycle progression. Elife 3, e02443. 10.7554/eLife.02443. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Armstrong C, and Spencer SL (2021). Replication-dependent histone biosynthesis is coupled to cell-cycle commitment. Proc. Natl. Acad. Sci. USA 118, e2100178118. 10.1073/pnas.2100178118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Mejlvang J, Feng Y, Alabert C, Neelsen KJ, Jasencakova Z, Zhao X, Lees M, Sandelin A, Pasero P, Lopes M, and Groth A (2014). New histone supply regulates replication fork speed and PCNA unloading. J. Cell Biol 204, 29–43. 10.1083/jcb.201305017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Zeman MK, and Cimprich KA (2014). Causes and consequences of replication stress. Nat. Cell Biol 16, 2–9. 10.1038/ncb2897. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Moldovan GL, Pfander B, and Jentsch S (2007). PCNA, the maestro of the replication fork. Cell 129, 665–679. 10.1016/j.cell.2007.05.003. [DOI] [PubMed] [Google Scholar]
- 23.Berti M, Cortez D, and Lopes M (2020). The plasticity of DNA replication forks in response to clinically relevant genotoxic stress. Nat. Rev. Mol. Cell Biol 21, 633–651. 10.1038/s41580-020-0257-5. [DOI] [PubMed] [Google Scholar]
- 24.Kelman Z (1997). PCNA: structure, functions and interactions. Oncogene 14, 629–640. 10.1038/sj.onc.1200886. [DOI] [PubMed] [Google Scholar]
- 25.Komatsu K, Wharton W, Hang H, Wu C, Singh S, Lieberman HB, Pledger WJ, and Wang HG (2000). PCNA interacts with hHus1/hRad9 in response to DNA damage and replication inhibition. Oncogene 19, 5291–5297. 10.1038/sj.onc.1203901. [DOI] [PubMed] [Google Scholar]
- 26.Ulrich HD (2009). Regulating post-translational modifications of the eukaryotic replication clamp PCNA. DNA Repair 8, 461–469. 10.1016/j.dnarep.2009.01.006. [DOI] [PubMed] [Google Scholar]
- 27.Kanao R, and Masutani C (2017). Regulation of DNA damage tolerance in mammalian cells by post-translational modifications of PCNA. Mutat. Res 803–805, 82–88. 10.1016/j.mrfmmm.2017.06.004. [DOI] [PubMed] [Google Scholar]
- 28.Ulrich HD, and Takahashi T (2013). Readers of PCNA modifications. Chromosoma 122, 259–274. 10.1007/s00412-013-0410-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Papouli E, Chen S, Davies AA, Huttner D, Krejci L, Sung P, and Ulrich HD (2005). Crosstalk between SUMO and ubiquitin on PCNA is mediated by recruitment of the helicase Srs2p. Mol. Cell 19, 123–133. 10.1016/j.molcel.2005.06.001. [DOI] [PubMed] [Google Scholar]
- 30.Kannouche PL, Wing J, and Lehmann AR (2004). Interaction of human DNA polymerase eta with monoubiquitinated PCNA: a possible mechanism for the polymerase switch in response to DNA damage. Mol. Cell 14, 491–500. 10.1016/s1097-2765(04)00259-x. [DOI] [PubMed] [Google Scholar]
- 31.Hoege C, Pfander B, Moldovan GL, Pyrowolakis G, and Jentsch S (2002). RAD6-dependent DNA repair is linked to modification of PCNA by ubiquitin and SUMO. Nature 419, 135–141. 10.1038/nature00991. [DOI] [PubMed] [Google Scholar]
- 32.Bai G, Kermi C, Stoy H, Schiltz CJ, Bacal J, Zaino AM, Hadden MK, Eichman BF, Lopes M, and Cimprich KA (2020). HLTF promotes fork reversal, limiting replication stress resistance and preventing multiple mechanisms of unrestrained DNA synthesis. Mol. Cell 78, 1237–1251.e7. 10.1016/j.molcel.2020.04.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Kile AC, Chavez DA, Bacal J, Eldirany S, Korzhnev DM, Bezsonova I, Eichman BF, and Cimprich KA (2015). HLTF’s ancient HIRAN domain binds 3’ DNA ends to drive replication fork reversal. Mol. Cell 58, 1090–1100. 10.1016/j.molcel.2015.05.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Mariño-Ramírez L, Jordan IK, and Landsman D (2006). Multiple independent evolutionary solutions to core histone gene regulation. Genome Biol 7, R122. 10.1186/gb-2006-7-12-r122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Keshet R, Szlosarek P, Carracedo A, and Erez A (2018). Rewiring urea cycle metabolism in cancer to support anabolism. Nat. Rev. Cancer 18, 634–645. 10.1038/s41568-018-0054-z. [DOI] [PubMed] [Google Scholar]
- 36.Sobel RE, Cook RG, Perry CA, Annunziato AT, and Allis CD (1995). Conservation of deposition-related acetylation sites in newly synthesized histones H3 and H4. Proc. Natl. Acad. Sci. USA 92, 1237–1241. 10.1073/pnas.92.4.1237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Denu JM (2019). Histone acetyltransferase 1 links metabolism and transcription to cell-cycle progression. Mol. Cell 75, 664–665. 10.1016/j.molcel.2019.08.004. [DOI] [PubMed] [Google Scholar]
- 38.Taddei A, Roche D, Sibarita JB, Turner BM, and Almouzni G (1999). Duplication and maintenance of heterochromatin domains. J. Cell Biol 147, 1153–1166. 10.1083/jcb.147.6.1153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Rausch LK, Netzer NC, Hoegel J, and Pramsohler S (2017). The linkage between breast cancer, hypoxia, and adipose tissue. Front. Oncol 7, 211. 10.3389/fonc.2017.00211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Liu KA, Lashinger LM, Rasmussen AJ, and Hursting SD (2014). Leucine supplementation differentially enhances pancreatic cancer growth in lean and overweight mice. Cancer Metabol 2, 6. 10.1186/2049-3002-2-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Cimprich KA, and Cortez D (2008). ATR: an essential regulator of genome integrity. Nat. Rev. Mol. Cell Biol 9, 616–627. 10.1038/nrm2450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Forment JV, and Jackson SP (2015). A flow cytometry-based method to simplify the analysis and quantification of protein association to chromatin in mammalian cells. Nat. Protoc 10, 1297–1307. 10.1038/nprot.2015.066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Sirbu BM, Couch FB, Feigerle JT, Bhaskara S, Hiebert SW, and Cortez D (2011). Analysis of protein dynamics at active, stalled, and collapsed replication forks. Genes Dev 25, 1320–1327. 10.1101/gad.2053211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Técher H, Koundrioukoff S, Nicolas A, and Debatisse M (2017). The impact of replication stress on replication dynamics and DNA damage in vertebrate cells. Nat. Rev. Genet 18, 535–550. 10.1038/nrg.2017.46. [DOI] [PubMed] [Google Scholar]
- 45.van Harten AM, Buijze M, van der Mast R, Rooimans MA, Martensde Kemp SR, Bachas C, Brink A, Stigter-van Walsum M, Wolthuis RMF, and Brakenhoff RH (2019). Targeting the cell cycle in head and neck cancer by Chk1 inhibition: a novel concept of bimodal cell death. Oncogenesis 8, 38. 10.1038/s41389-019-0147-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Ramaswamy S, Venkatadri N, and Bapna JS (1987). An investigation into the possible development of chronic tolerance to analgesia and dependence on prolactin. Fundam. Clin. Pharmacol 1, 445–449. 10.1111/j.1472-8206.1987.tb00577.x. [DOI] [PubMed] [Google Scholar]
- 47.Kawasumi M, Bradner JE, Tolliday N, Thibodeau R, Sloan H, Brummond KM, and Nghiem P (2014). Identification of ATR-Chk1 pathway inhibitors that selectively target p53-deficient cells without directly suppressing ATR catalytic activity. Cancer Res 74, 7534–7545. 10.1158/0008-5472.CAN-14-2650. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Leung-Pineda V, Ryan CE, and Piwnica-Worms H (2006). Phosphorylation of Chk1 by ATR is antagonized by a Chk1-regulated protein phosphatase 2A circuit. Mol. Cell Biol 26, 7529–7538. 10.1128/MCB.00447-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Leung KHT, Abou El Hassan M, and Bremner R (2013). A rapid and efficient method to purify proteins at replication forks under native conditions. Biotechniques 55, 204–206. 10.2144/000114089. [DOI] [PubMed] [Google Scholar]
- 50.Bétous R, Goullet de Rugy T, Pelegrini AL, Queille S, de Villartay JP, and Hoffmann JS (2018). DNA replication stress triggers rapid DNA replication fork breakage by Artemis and XPF. PLoS Genet 14, e1007541. 10.1371/journal.pgen.1007541. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Arbel M, Choudhary K, Tfilin O, and Kupiec M (2021). PCNA loaders and unloaders-one ring that rules them all. Genes 12, 1812. 10.3390/genes12111812. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Choe KN, and Moldovan GL (2017). Forging ahead through darkness: PCNA, still the principal conductor at the replication fork. Mol. Cell 65, 380–392. 10.1016/j.molcel.2016.12.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Kang MS, Kim J, Ryu E, Ha NY, Hwang S, Kim BG, Ra JS, Kim YJ, Hwang JM, Myung K, and Kang S (2019). PCNA unloading is negatively regulated by BET proteins. Cell Rep 29, 4632–4645.e5. 10.1016/j.celrep.2019.11.114. [DOI] [PubMed] [Google Scholar]
- 54.Nakamura K, Kustatscher G, Alabert C, Hödl M, Forne I, Völker-Albert M, Satpathy S, Beyer TE, Mailand N, Choudhary C, et al. (2021). Proteome dynamics at broken replication forks reveal a distinct ATM-directed repair response suppressing DNA double-strand break ubiquitination. Mol. Cell 81, 1084–1099.e6. 10.1016/j.molcel.2020.12.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Brühl J, Trautwein J, Schäfer A, Linne U, and Bouazoune K (2019). The DNA repair protein SHPRH is a nucleosome-stimulated ATPase and a nucleosome-E3 ubiquitin ligase. Epigenet. Chromatin 12, 52. 10.1186/s13072-019-0294-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Thakar T, Leung W, Nicolae CM, Clements KE, Shen B, Bielinsky AK, and Moldovan GL (2020). Ubiquitinated-PCNA protects replication forks from DNA2-mediated degradation by regulating Okazaki fragment maturation and chromatin assembly. Nat. Commun 11, 2147. 10.1038/s41467-020-16096-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Zellweger R, Dalcher D, Mutreja K, Berti M, Schmid JA, Herrador R, Vindigni A, and Lopes M (2015). Rad51-mediated replication fork reversal is a global response to genotoxic treatments in human cells. J. Cell Biol 208, 563–579. 10.1083/jcb.201406099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Taglialatela A, Alvarez S, Leuzzi G, Sannino V, Ranjha L, Huang JW, Madubata C, Anand R, Levy B, Rabadan R, et al. (2017). Restoration of replication fork stability in BRCA1- and BRCA2-deficient cells by inactivation of SNF2-family fork remodelers. Mol. Cell 68, 414–430.e8. 10.1016/j.molcel.2017.09.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Kolinjivadi AM, Sannino V, De Antoni A, Zadorozhny K, Kilkenny M, Técher H, Baldi G, Shen R, Ciccia A, Pellegrini L, et al. (2017). Smar-cal1-Mediated fork reversal triggers mre11-dependent degradation of nascent DNA in the absence of Brca2 and stable Rad51 nucleofilaments. Mol. Cell 67, 867–881.e7. 10.1016/j.molcel.2017.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Vujanovic M, Krietsch J, Raso MC, Terraneo N, Zellweger R, Schmid JA, Taglialatela A, Huang JW, Holland CL, Zwicky K, et al. (2017). Replication fork slowing and reversal upon DNA damage require PCNA polyubiquitination and ZRANB3 DNA translocase activity. Mol. Cell 67, 882–890.e5. 10.1016/j.molcel.2017.08.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Gu L, Lingeman R, Yakushijin F, Sun E, Cui Q, Chao J, Hu W, Li H, Hickey RJ, Stark JM, et al. (2018). The anticancer activity of a first-in-class small-molecule targeting PCNA. Clin. Cancer Res 24, 6053–6065. 10.1158/1078-0432.CCR-18-0592. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Darnell AM, Subramaniam AR, and O’Shea EK (2018). Translational control through differential ribosome pausing during amino acid limitation in mammalian cells. Mol. Cell 71, 229–243.e11. 10.1016/j.molcel.2018.06.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Smith S, and Stillman B (1989). Purification and characterization of CAF-I, a human cell factor required for chromatin assembly during DNA replication in vitro. Cell 58, 15–25. 10.1016/0092-8674(89)90398-x. [DOI] [PubMed] [Google Scholar]
- 64.Alabert C, Jasencakova Z, and Groth A (2017). Chromatin replication and histone dynamics. Adv. Exp. Med. Biol 1042, 311–333. 10.1007/978-981-10-6955-0_15. [DOI] [PubMed] [Google Scholar]
- 65.Shibahara K, and Stillman B (1999). Replication-dependent marking of DNA by PCNA facilitates CAF-1-coupled inheritance of chromatin. Cell 96, 575–585. 10.1016/s0092-8674(00)80661-3. [DOI] [PubMed] [Google Scholar]
- 66.Peng M, Cong K, Panzarino NJ, Nayak S, Calvo J, Deng B, Zhu LJ, Morocz M, Hegedus L, Haracska L, and Cantor SB (2018). Opposing roles of FANCJ and HLTF protect forks and restrain replication during stress. Cell Rep 24, 3251–3261. 10.1016/j.celrep.2018.08.065. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Unk I, Hajdú I, Fátyol K, Hurwitz J, Yoon JH, Prakash L, Prakash S, and Haracska L (2008). Human HLTF functions as a ubiquitin ligase for proliferating cell nuclear antigen polyubiquitination. Proc. Natl. Acad. Sci. USA 105, 3768–3773. 10.1073/pnas.0800563105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Blastyák A, Hajdú I, Unk I, and Haracska L (2010). Role of double-stranded DNA translocase activity of human HLTF in replication of damaged DNA. Mol. Cell Biol 30, 684–693. 10.1128/MCB.00863-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Cho S, Cinghu S, Yu JR, and Park WY (2011). Helicase-like transcription factor confers radiation resistance in cervical cancer through enhancing the DNA damage repair capacity. J. Cancer Res. Clin. Oncol 137, 629–637. 10.1007/s00432-010-0925-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Ye C, Sun NX, Ma Y, Zhao Q, Zhang Q, Xu C, Wang SB, Sun SH, Wang F, and Li W (2015). MicroRNA-145 contributes to enhancing radiosensitivity of cervical cancer cells. FEBS Lett 589, 702–709. 10.1016/j.febslet.2015.01.037. [DOI] [PubMed] [Google Scholar]
- 71.Quinet A, Tirman S, Cybulla E, Meroni A, and Vindigni A (2021). To skip or not to skip: choosing repriming to tolerate DNA damage. Mol. Cell 81, 649–658. 10.1016/j.molcel.2021.01.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.van Toorn M, Turkyilmaz Y, Han S, Zhou D, Kim HS, Salas-Armenteros I, Kim M, Akita M, Wienholz F, Raams A, et al. (2022). Active DNA damage eviction by HLTF stimulates nucleotide excision repair. Mol. Cell 82, 1343–1358.e8. 10.1016/j.molcel.2022.02.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Marullo R, Castro M, Yomtoubian S, Calvo-Vidal MN, Revuelta MV, Krumsiek J, Cho A, Morgado PC, Yang S, Medina V, et al. (2021). The metabolic adaptation evoked by arginine enhances the effect of radiation in brain metastases. Sci. Adv 7, eabg1964. 10.1126/sciadv.abg1964. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Schmittgen TD, and Livak KJ (2008). Analyzing real-time PCR data by the comparative C(T) method. Nat. Protoc 3, 1101–1108. 10.1038/nprot.2008.73. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All data reported in this paper will be shared by the lead contact upon request.
This paper does not report original code.
Any additional information required to reanalyze the data reported in this paper is available from the lead contact upon request.