Cdc48 (known as VCP in mammals) is a highly conserved ATPase chaperone that plays an essential role in the assembly and disassembly of protein–DNA complexes and in degradation of misfolded proteins. We find that in Saccharomyces cerevisiae budding yeast, Cdc48 accumulates during cellular stress at intranuclear protein quality control sites (INQ). We show that Cdc48 function is required to suppress INQ formation under non-stress conditions and to promote recovery following genotoxic stress. Cdc48 physically associates with the INQ substrate and splicing factor Hsh155, and regulates its assembly with partner proteins. Accordingly, cdc48 mutants have defects in splicing and show spontaneous distribution of Hsh155 to INQ aggregates, where it is stabilized. Overall, this study shows that Cdc48 regulates deposition of proteins at INQ and suggests a previously unknown role for Cdc48 in the regulation or stabilization of splicing subcomplexes.

This article has an associated First Person interview with the first author of the paper.

Protein quality control (PQC) refers to the process of triage for non-native, unassembled or mislocalized proteins through a combination of sequestration, degradation and refolding (Hartl et al., 2011). PQC is a part of normal cellular homeostasis and is particularly important during stress, when non-native proteins assemble in visible aggregate structures at different subcellular locations. Various proteotoxic stresses elicit such a coordinated PQC response, with the best characterized of these being heat stress (Wallace et al., 2015) and expression of amyloid-forming proteins (Kaganovich et al., 2008). PQC responses depend on networks of molecular chaperones, on the action of stress-activated protein aggregase and disaggregase activities (Hartl et al., 2011) and on protein degradation by the ubiquitin–proteasome system and autophagy (Kaganovich et al., 2008; Kumar et al., 2016).

During stress in yeast, specific types of aggregate structures with different components and behaviors form (Saarikangas and Barral, 2016). In the cytoplasm, aggregates at the vacuolar membrane, called the insoluble protein deposit (IPOD), are repositories for irreversibly aggregated proteins destined for degradation, such as amyloids (Kaganovich et al., 2008). Smaller dynamic cytoplasmic aggregates called Q-bodies (quality control bodies) have also been described for protein aggregate reporters such as the temperature-sensitive mutant Ubc9-ts and VHL (Escusa-Toret et al., 2013; Ruan et al., 2017; Zhou et al., 2014). Finally, at the nuclear envelope, the juxtanuclear quality control site (JUNQ) forms peripheral to but outside the nucleus, whereas the intranuclear quality control site (INQ) forms peripheral to and inside the nucleus (Kaganovich et al., 2008; Miller et al., 2015). Most yeast studies of PQC focus on aggregation of the replication fork checkpoint proteins (i.e. Mrc1; Gallina et al., 2015), the activation of DNA repair proteins (i.e. Mus81; Saugar et al., 2017) or rewiring of the spliced transcriptome (i.e. Hsh155; Mathew et al., 2017).

INQ formation is driven by the small heat shock protein Hsp42 and the aggregase protein Btn2 (Ho et al., 2019), whereas INQ dissolution is regulated by Hsp104 and the Hsp70 system (Miller et al., 2015; Saugar et al., 2017). Here, we show that Cdc48 is also a stress-inducible component on INQ structures and that it plays an enzymatic role in INQ turnover. Cdc48 plays numerous normal and stress-responsive roles in cells, including in nuclear PQC (Gallagher et al., 2014; Maric et al., 2017; Pantazopoulou et al., 2016; Verma et al., 2011). Here, we show that Cdc48 is concentrated at INQ following DNA damage, and that its activity is important to suppress the spontaneous aggregation of splicing factor Hsh155 in INQ. Cdc48 physically interacts with Hsh155, even under non-stress conditions, and promotes its assembly with partners in the SF3B complex of the spliceosome. Taken together, these results demonstrate a previously unrecognized role for Cdc48 in protein triage to INQ and suggest unappreciated roles for Cdc48 in splicing.

Cdc48–GFP localizes to intranuclear quality control sites

We previously screened hundreds of genome stability regulatory proteins for relocalization in response to DNA damage, and identified uncharacterized stress-induced relocalization of Cdc48–GFP (Mathew et al., 2017). Cdc48 is widely distributed in unstressed cells, but is enriched at a nuclear periphery/ER localization due to its role in ER-associated degradation (ERAD) (Fig. 1A). Following methyl methanesulfonate (MMS) treatment, Cdc48 relocalized to nuclear and non-nuclear foci (Fig. 1A; Fig. S1A). To test whether these foci were INQ, we expressed Cdc48–GFP together with Hos2–mCherry, a known marker of INQ (Gallina et al., 2015; Tkach et al., 2012). Cdc48–GFP foci colocalized with Hos2–mCherry foci, suggesting that Cdc48 was accumulating at INQ following MMS treatment (Fig. 1B,C). Cdc48 also colocalized with Hos2 at peripheral aggregates, which may be cytoplasmic PQC (CytoQ) sites (Fig. 1B,C). We next ablated INQ formation by deleting the chaperones HSP42 and BTN2, which are essential for INQ protein localization (Gallina et al., 2015; Mathew et al., 2017; Miller et al., 2015). Cdc48–GFP foci were strongly diminished in MMS-treated hsp42Δ or btn2Δ cells, showing that INQ formation was essential for Cdc48 foci (Fig. 1D; Fig. S1B). In contrast, deletion of the Hsp40 homolog Apj1, which has previously been implicated in disaggregation and turnover of INQ-resident proteins, dramatically increased the frequency of Cdc48–GFP foci (Fig. 1D; Fig. S1B) (den Brave et al., 2020). Hos2–mCherry was used as a control, and it showed the expected decrease of INQ formation in cells lacking HSP42 or BTN2 and increased INQ frequency in cells lacking APJ1 (Fig. S1B). Taken together, the results of these experiments suggest that Cdc48 foci mark INQ structures induced by MMS treatment and are influenced by INQ regulatory chaperones.

Fig. 1.

Cdc48 localizes to PQC structures in a chaperone-dependent manner. (A) Cdc48–GFP fusions relocalize into foci upon MMS treatment (white arrowheads). A schematic on left summarizes the movements in both conditions. (B) Cdc48–GFP foci overlap with Hos2–mCherry (mChe) foci after MMS treatment. White arrowheads indicate INQ foci, yellow arrowheads indicate CytoQ foci. (C) Quantification of MMS-induced Cdc48 in perinuclear (INQ) and peripheral (CytoQ) foci. Data are presented as mean±s.e.m., n=3 with >100 cells each. UNT, untreated. ***P<0.001 (Fisher test). (D) Reduction of Cdc48–GFP foci in MMS-treated cells lacking HSP42 (hsp42Δ) or BTN2 (btn2Δ), and increase of Cdc48–GFP foci in apj1Δ cells, compared with those in MMS-treated wild-type cells (wt). Representative images (top) and quantification of percentage of cells with foci (bottom) are shown. Data are presented as mean±s.e.m., n=3 with >50 cells each. **P<0.01 (Fisher test). Asterisks show P-value thresholds in comparison to WT under the same condition. Scale bars: 5 μm.

Fig. 1.

Cdc48 localizes to PQC structures in a chaperone-dependent manner. (A) Cdc48–GFP fusions relocalize into foci upon MMS treatment (white arrowheads). A schematic on left summarizes the movements in both conditions. (B) Cdc48–GFP foci overlap with Hos2–mCherry (mChe) foci after MMS treatment. White arrowheads indicate INQ foci, yellow arrowheads indicate CytoQ foci. (C) Quantification of MMS-induced Cdc48 in perinuclear (INQ) and peripheral (CytoQ) foci. Data are presented as mean±s.e.m., n=3 with >100 cells each. UNT, untreated. ***P<0.001 (Fisher test). (D) Reduction of Cdc48–GFP foci in MMS-treated cells lacking HSP42 (hsp42Δ) or BTN2 (btn2Δ), and increase of Cdc48–GFP foci in apj1Δ cells, compared with those in MMS-treated wild-type cells (wt). Representative images (top) and quantification of percentage of cells with foci (bottom) are shown. Data are presented as mean±s.e.m., n=3 with >50 cells each. **P<0.01 (Fisher test). Asterisks show P-value thresholds in comparison to WT under the same condition. Scale bars: 5 μm.

Cdc48 ATPase regulates the aggregation and INQ localization of Hsh155

Because Cdc48 functions in protein biogenesis and turnover, we suspected that it might have a functional role at INQ. We assembled a set of CDC48 temperature-sensitive alleles, which grew robustly at the permissive temperature of 25°C but exhibited a range of fitness defects (Fig. S1C) (Li et al., 2011). Examining the localization of an INQ substrate in the most severe allele, cdc48-4601 (hereafter called cdc48-4) (Fig. S1E), revealed spontaneous Hsh155–GFP and Hos2–GFP foci at nuclear and cytoplasmic aggregate sites (Fig. 2A; Fig. S1D). MMS treatment further increased the proportion of cells with visible aggregates beyond that seen for MMS-treated wild-type (WT) cells (Fig. 2B). We confirmed that other CDC48 temperature-sensitive alleles also increased spontaneous INQ formation to varying degrees (Fig. 2C). Consistent with a functionally important role, cdc48-4 alleles were significantly impaired in recovery from a 2 h treatment of MMS (Fig. 2D). Although there are many possible mechanisms for the MMS sensitivity of a CDC48 temperature-sensitive allele, our data show that Cdc48 function is important for regulating the INQ localization of Hsh155 in normal and DNA-damaging conditions. Finally, because Cdc48 uses its ATPase activity to disassemble or unfold proteins, we tested the ability of enzymatically active or dead alleles to rescue the cdc48-4 phenotype. Whereas WT CDC48 plasmids suppressed Hsh155 foci, an ATPase-deficient CDC48E315Q allele could not (Fig. 2E,F) (Gallagher et al., 2014). Thus, Cdc48 ATPase activity is essential to suppress aggregation of Hsh155–GFP into INQ structures.

Fig. 2.

Cdc48 ATPase regulates the aggregation and INQ localization of Hsh155. (A) Visualization of Hsh155–GFP foci in cells with temperature-sensitive alleles of CDC48 (cdc48-4) with or without MMS stress at 34°C. White arrowheads show foci in untreated cells. (B) Quantification of foci accumulation in untreated and MMS-treated cdc48-4 cells compared with that in WT cells with or without MMS. (C) Spontaneous Hsh155–GFP foci in three different CDC48 mutant alleles at 34°C. Representative images (top) and quantification (bottom) are shown. (D) Growth curve analysis of cdc48-4 after 2 h of MMS treatment and washout compared with untreated cdc48-4 cells and WT with or without MMS treatment and washout at 34°C. Data are presented as mean±s.e.m., n=3 with triplicates in each. The dotted lines represent the interval wherein all cdc48-4 data points significantly differ from WT after MMS washout (P-value ranges from P<0.05–P<0.0001) and where data points for untreated and treated cdc48-4 cells significantly differ from each other (P<0.05). Two-way ANOVA with Tukey's post hoc test. (E,F) Rescue of Hsh155–GFP foci formation by WT CDC48 but not by the CDC48E315Q ATPase-deficient allele. E is a representative image of data quantified in F. For B,C and F quantifications: data are presented as mean±s.e.m., n=3 with >50 cells each. ***P<0.001; **P<0.01; *P<0.05; ns, not significant (Fisher test). Scale bars: 5 μm.

Fig. 2.

Cdc48 ATPase regulates the aggregation and INQ localization of Hsh155. (A) Visualization of Hsh155–GFP foci in cells with temperature-sensitive alleles of CDC48 (cdc48-4) with or without MMS stress at 34°C. White arrowheads show foci in untreated cells. (B) Quantification of foci accumulation in untreated and MMS-treated cdc48-4 cells compared with that in WT cells with or without MMS. (C) Spontaneous Hsh155–GFP foci in three different CDC48 mutant alleles at 34°C. Representative images (top) and quantification (bottom) are shown. (D) Growth curve analysis of cdc48-4 after 2 h of MMS treatment and washout compared with untreated cdc48-4 cells and WT with or without MMS treatment and washout at 34°C. Data are presented as mean±s.e.m., n=3 with triplicates in each. The dotted lines represent the interval wherein all cdc48-4 data points significantly differ from WT after MMS washout (P-value ranges from P<0.05–P<0.0001) and where data points for untreated and treated cdc48-4 cells significantly differ from each other (P<0.05). Two-way ANOVA with Tukey's post hoc test. (E,F) Rescue of Hsh155–GFP foci formation by WT CDC48 but not by the CDC48E315Q ATPase-deficient allele. E is a representative image of data quantified in F. For B,C and F quantifications: data are presented as mean±s.e.m., n=3 with >50 cells each. ***P<0.001; **P<0.01; *P<0.05; ns, not significant (Fisher test). Scale bars: 5 μm.

SUMOylation regulates INQ formation in cdc48 mutants

Previous studies have linked Cdc48 and its co-factors to proteostasis of nuclear proteins via the San1 E3 ubiquitin ligase and the inner nuclear membrane proteins Asi1–Asi3; however, we saw no significant effect of SAN1 deletion on INQ (Fig. S1F) (Gallagher et al., 2014; Pantazopoulou et al., 2016). In addition, Cdc48 is involved in regulation of protein–DNA complexes, including the removal of RNA polymerase from damaged DNA (Verma et al., 2011), the removal of replisomes during DNA replication termination (Maric et al., 2017), at protein–DNA crosslinks (Amunugama et al., 2018) and during the release of condensin complexes (Thattikota et al., 2018). Thus, Cdc48 is well-positioned to regulate protein sequestration pathways such as INQ formation following DNA damage detection. Other groups have noted the importance of SUMO to INQ formation (Gallina et al., 2015), and we also found that reducing SUMOylation with ubc9-ts (SUMO E2) or SIZ1 (SUMO E3) deletion reduced INQ localization of Hsh155–GFP in MMS-treated WT or cdc48-4 cells (Fig. 3A,B). Indeed, mutation of the SUMO-interacting motif of Cdc48 (CDC48SIM) impairs its ability to rescue cdc48-4 cells, relative to rescue by WT CDC48 (Fig. 3C) (Bergink et al., 2013). Because we used Hsh155–GFP as a marker for INQ, we elected to directly test whether Hsh155 is SUMOylated using a Smt3–His×7 purification scheme. Whereas SUMOylation of the Rfa1–GFP control was easily detectable upon MMS treatment, no detectable pulldown of Hsh155–GFP with an Smt3–His×7 tag was observed with or without MMS (Fig. 3D). Thus, the identity of any SUMO targets important for regulating INQ will require additional study beyond the scope of this work. However, our findings again highlight the importance SUMO in regulating INQ.

Fig. 3.

Testing potential roles of SUMO in Cdc48-induced Hsh155–GFP foci formation. (A) Hsh155–GFP foci induction by MMS in SUMO regulator mutant strains ubc9-ts and siz1Δ. (B) Induction of spontaneous Hsh155–GFP foci in cdc48-4 is reduced by deletion of SUMO E3 ligase SIZ1. (C) Spontaneous Hsh155–GFP foci in cdc48-4 complemented by plasmids bearing WT CDC48 or the CDC48SIM mutant. For A–C, representative images (top) and quantification (bottom) are shown. Data are presented as mean±s.e.m., n=3 with >50 cells each. ****P<0.0001; ***P<0.001; ns, not significant (Fisher test). (D) SMT3–His×7 (His×7SUMO) pulldown assays in untreated or MMS-treated WT and SMT3–His×7-expressing cells. DNA-damage induced SUMOylation of RFA1 in MMS (Psakhye and Jentsch, 2012) is shown as a positive control after Ni-NTA bead pulldown (Ni PD). No bands are detectable when blotting for Hsh155–GFP. Pgk1 is shown as a loading control for input. Input, 4%. Scale bars: 5 μm.

Fig. 3.

Testing potential roles of SUMO in Cdc48-induced Hsh155–GFP foci formation. (A) Hsh155–GFP foci induction by MMS in SUMO regulator mutant strains ubc9-ts and siz1Δ. (B) Induction of spontaneous Hsh155–GFP foci in cdc48-4 is reduced by deletion of SUMO E3 ligase SIZ1. (C) Spontaneous Hsh155–GFP foci in cdc48-4 complemented by plasmids bearing WT CDC48 or the CDC48SIM mutant. For A–C, representative images (top) and quantification (bottom) are shown. Data are presented as mean±s.e.m., n=3 with >50 cells each. ****P<0.0001; ***P<0.001; ns, not significant (Fisher test). (D) SMT3–His×7 (His×7SUMO) pulldown assays in untreated or MMS-treated WT and SMT3–His×7-expressing cells. DNA-damage induced SUMOylation of RFA1 in MMS (Psakhye and Jentsch, 2012) is shown as a positive control after Ni-NTA bead pulldown (Ni PD). No bands are detectable when blotting for Hsh155–GFP. Pgk1 is shown as a loading control for input. Input, 4%. Scale bars: 5 μm.

Cdc48 regulates the stability of Hsh155 and assembly with its partners

If Cdc48 is important for the prevention or dissolution of INQ structures, then it could regulate the stability of INQ substrate proteins under stress. Consistent with this idea, Hsh155–GFP protein was more abundant in the cdc48-4 mutant, and more stable in cycloheximide chase experiments, than it was in WT (Fig. 4A). Because cycloheximide treatment can itself affect protein sequestration (Zhou et al., 2014), we also monitored the impact of Cdc48 on Hsh155 lifetime with an orthogonal approach using in vivo tandem fluorescent timer (tFT) fusions (Khmelinskii et al., 2012). The ratio of GFP to mCherry fluorescence in Hsh155–tFT fusions was significantly red-shifted in cdc48-4 cells, indicating a longer relative protein lifetime (Fig. 4B; Fig. S1G,H). These data are consistent with INQ serving a protective role for substrate proteins, because Hsh155 both becomes more abundant and aggregates in cdc48-4 cells. Cdc48 could affect the stability of Hsh155 through direct protein–protein interactions. Indeed, we observed co-precipitation of Hsh155–GFP with Cdc48–TAP and via reciprocal pulldown (i.e. Cdc48–GFP with Hsh155–TAP) under both untreated and MMS-treated conditions (Fig. 4C,D).

Fig. 4.

Direct effects of Cdc48 on Hsh155 stability and function. (A) Hsh155 protein levels with or without cycloheximide treatment for the indicated times, assayed by anti-GFP western blotting in WT and cdc48-4 cells, relative to Pgk1 levels. CHX, cycloheximide; UNT, untreated. A representative blot from three independent experiments is shown. (B) Tandem fluorescent timer (tFT) analysis of Hsh155–mCherry–GFP dual fusion lifetime in the indicated strains with or without MMS treatment. Ratios of GFP (fast maturing) to mCherry (slow maturing) fluorescence in the nucleus are shown. Quantification of the fluorescent timer indicates the presence of older proteins in the nucleus of cdc48-4 cells compared to those in WT in both conditions. Representative images are shown in Fig. S1G. Box plots show the median and interquartile range, with whiskers indicating the minimum and maximum values. Three replicates, n≥30 per replicate. ****P<0.0001 (Mann–Whitney test). (C,D) Physical interaction of Hsh155 and Cdc48 with or without MMS treatment, as assayed by co-immunoprecipitation (IP) using Cdc48–TAP (C) or Hsh155–TAP (D) followed by western blotting. Control (GFP only) lanes are IgG bead IPs from otherwise isogenic cells expressing Hsh155–GFP or Cdc48–GFP without a TAP fusion to account for any non-specific binding. The IgG bead lane is a no lysate control for the beads. Input, 5%. (E) Co-precipitation of Hsh155 with its partner Cus1 in the SF3B complex in cdc48-4 cells is disrupted in both untreated and MMS conditions, similar to MMS-induced WT cells. Input, 5%. (F) LacZ splicing reporter assay results in WT or cdc48-3 cells treated with MMS. Top: a schematic of the assay (top) shows the intronic versus intronless LacZ constructs that enable quantification of splicing efficiency. Bottom: quantification of relative LacZ splicing in untreated and MMS-treated conditions, normalized to the WT ‘no intron’ control, in both WT and cdc48-3 cells. Data are presented as mean±s.e.m., n=5 with triplicate individual transformants in each. **P<0.01 (Student's t-test). (G) Model of Cdc48 as a chaperone for Hsh155 involved in its quality control and sequestration at INQ sites. Cdc48 binds Hsh155 under normal conditions, but its disruption or DNA damage leads to SF3B disassembly. Cdc48 may also play a role in INQ resolution or turnover, explaining why its loss increases INQ formation.

Fig. 4.

Direct effects of Cdc48 on Hsh155 stability and function. (A) Hsh155 protein levels with or without cycloheximide treatment for the indicated times, assayed by anti-GFP western blotting in WT and cdc48-4 cells, relative to Pgk1 levels. CHX, cycloheximide; UNT, untreated. A representative blot from three independent experiments is shown. (B) Tandem fluorescent timer (tFT) analysis of Hsh155–mCherry–GFP dual fusion lifetime in the indicated strains with or without MMS treatment. Ratios of GFP (fast maturing) to mCherry (slow maturing) fluorescence in the nucleus are shown. Quantification of the fluorescent timer indicates the presence of older proteins in the nucleus of cdc48-4 cells compared to those in WT in both conditions. Representative images are shown in Fig. S1G. Box plots show the median and interquartile range, with whiskers indicating the minimum and maximum values. Three replicates, n≥30 per replicate. ****P<0.0001 (Mann–Whitney test). (C,D) Physical interaction of Hsh155 and Cdc48 with or without MMS treatment, as assayed by co-immunoprecipitation (IP) using Cdc48–TAP (C) or Hsh155–TAP (D) followed by western blotting. Control (GFP only) lanes are IgG bead IPs from otherwise isogenic cells expressing Hsh155–GFP or Cdc48–GFP without a TAP fusion to account for any non-specific binding. The IgG bead lane is a no lysate control for the beads. Input, 5%. (E) Co-precipitation of Hsh155 with its partner Cus1 in the SF3B complex in cdc48-4 cells is disrupted in both untreated and MMS conditions, similar to MMS-induced WT cells. Input, 5%. (F) LacZ splicing reporter assay results in WT or cdc48-3 cells treated with MMS. Top: a schematic of the assay (top) shows the intronic versus intronless LacZ constructs that enable quantification of splicing efficiency. Bottom: quantification of relative LacZ splicing in untreated and MMS-treated conditions, normalized to the WT ‘no intron’ control, in both WT and cdc48-3 cells. Data are presented as mean±s.e.m., n=5 with triplicate individual transformants in each. **P<0.01 (Student's t-test). (G) Model of Cdc48 as a chaperone for Hsh155 involved in its quality control and sequestration at INQ sites. Cdc48 binds Hsh155 under normal conditions, but its disruption or DNA damage leads to SF3B disassembly. Cdc48 may also play a role in INQ resolution or turnover, explaining why its loss increases INQ formation.

Hsh155 is a splicing factor that exists in a complex called SF3B with Hsh49 and Cus1. We previously found that Hsh155 disassembles from the SF3B complex upon MMS treatment (Mathew et al., 2017). To assess the potential impact of Cdc48 on SF3B complex stability, we monitored pulldown of Hsh155–GFP in WT and cdc48-4 cells with or without MMS treatment using co-immunoprecipation. Loss of Cdc48 function greatly diminished the presence of the Hsh155–Cus1 complex, even in untreated cells (Fig. 4E). This suggests that Cdc48 may play a role in the integrity of the SF3B complex even in the absence of DNA-damage signals. Whereas cdc48-4 grew too poorly to easily monitor splicing activity, the cdc48-3 allele that showed an intermediate induction of Hsh155 foci (Fig. 2C) did show reduced splicing flux in a LacZ-based splicing efficiency assay (Fig. 4F) (Galy et al., 2004; Tam et al., 2019).

Context and perspective

Spatial control of PQC is a fundamental feature of stress responses in eukaryotic cells (Sontag et al., 2014). Here, we show that Cdc48, a key regulator of numerous cellular PQC activities, is also associated with INQ, where it interacts with, reduces the accumulation of, or speeds the turnover of the INQ substrate Hsh155 (Fig. 4G). This defines a novel and complementary function for Cdc48 in nuclear protein quality control. Loss of SUMO-conjugating enzyme function or E3 SUMO ligase function reduced INQ formation, consistent with a role for this modification in organized INQ structures. Considerably more work is needed to understand the regulation of INQ substrate protein localization and turnover. Our data specifically implicate Cdc48 in regulating the sequestration, complex formation and stability of the spliceosomal protein Hsh155 within the SF3B complex. We previously found that Hsh155 sequestration may regulate ribosomal protein gene expression through reduced splicing efficiency of ribosomal protein gene transcripts (Mathew et al., 2017). In this regard, it is notable that Cdc48 is also a critical regulator of ribosomal quality control mechanisms and can sense ribosomal stress (Brandman et al., 2012; Defenouillère et al., 2013). It is possible that these responses are coordinated during the DNA-damage response. Regardless, our data define a new function for Cdc48 in regulating the deposition of proteins at the INQ and potentially in the regulation of splicing complexes.

There is currently little evidence linking Cdc48 to splicing. Interestingly, a study of motor neuron transcriptome dynamics in induced pluripotent stem cell differentiation models of amyotrophic lateral sclerosis (ALS), where VCP (the human ortholog of CDC48) was mutated, showed that abnormal intron retention events are increased in VCP-mutated cells (Luisier et al., 2018). Thus, a role for Cdc48/VCP in splicing across species is possible, and we hope that additional research will elucidate how Cdc48 may affect splicing. In this regard, it is notable that the existence of an orthologous human intranuclear quality control response could involve disease genes, including the ALS gene VCP and the Hsh155 ortholog SF3B1, which is frequently mutated in various types of cancer (Yoshida and Ogawa, 2014). Given the proximity of the INQ to the nucleolus, we speculate that human studies highlighting the nucleolus as a phase-separated quality control compartment may reflect a conserved phenomenon homologous to the INQ in yeast (Frottin et al., 2019; Tkach et al., 2012). If so, then studies in model organisms should provide insight to nuclear protein quality control pathways that could modulate human disease.

Yeast strains, MMS treatments and growth analyses

Yeast (Saccharomyces cerevisiae) strains used in this study, including database IDs and genotypes, primers and plasmids, are listed in Table S1. All strains were in the s288c background and were grown under nutrient rich YPD medium (Sherman, 2002) or synthetic medium lacking amino acids (SC; Sherman, 2002) unless otherwise indicated. Schematics of the CDC48 mutant alleles are shown in Fig. S1E. Sequence information came from in-house sequencing or from publications (Simões et al., 2018; Verma et al., 2011). Serial dilution assays and growth curve analyses were performed as described previously (Mathew et al., 2017; Stirling et al., 2011). Briefly, cells with identical optical density (OD) were serially diluted tenfold and spotted on YPD plates using a 48-pin replica pinning manifold then incubated at the indicated temperatures for 72 h. Growth curves were analyzed in a Tecan M200 plate reader monitoring OD600 nm every 30 min for 48 h at 30°C. For all MMS treatment conditions cells were exposed to 0.05% MMS (∼99%; Sigma) in synthetic complete medium for 2 h unless otherwise noted. For growth curves, MMS was then washed out and replaced with fresh medium.

Live-cell imaging, image acquisition, analysis and statistical methods

Imaging and subsequent analysis were performed as described previously (Mathew et al., 2017). Log-phase yeast were mounted on slides pre-treated with concanavalin A, in SC growth medium. Immobilized cells were imaged using an Objective HCX PL APO 1.40 NA oil immersion 100× objective on an inverted Leica DMi8 microscope with a motorized DIC (differential interference contrast) turret (for DIC imaging) and a filter cube set for FITC/TRITC (for GFP and mCherry fluorescence imaging). The images were captured at room temperature using a scientific complementary metal oxide semiconductor (sCMOS) camera (ORCA Flash 4.0 V2; Hamamatsu), collected using MetaMorph Premier acquisition software and post processed [including gamma adjustments, counting of cells with/without foci and foci intensity measurements for tandem fluorescent timer (tFT) experiments] using ImageJ (National Institutes of Health). For all microscopy experiments, the significance of the differences was determined using Prism version 5 or higher (Graphpad). For foci intensity measurements in tFT experiments, samples were compared with t-tests or ANOVA; Prism performs F-tests for variance as part of this analysis. For comparisons of proportions, Fisher tests were used, and P-values were Holm–Bonferroni corrected in the event of multiple comparisons. Sample sizes and specific statistical details for each image analysis are listed in the figure legends.

Co-immunoprecipitation and western blotting

Co-immunoprecipitation was performed using yeast strains containing TAP tagged Cus1, Cdc48 and/or GFP tagged Hsh155, treated with or without MMS. TAP-tagged protein was captured using IgG sepharose fast flow beads (Sigma) and proceeded as described previously (Leung et al., 2016). Immunoblotting was carried out with mouse anti-GFP (1:500; ThermoFisher; Cat#. MA5-15256) and rabbit anti-TAP (1:1000; ThermoFisher; Cat#. CAB1001) antibodies. For western blotting, whole-cell extracts were prepared using tricholoroacetic acid (TCA) extraction and blotted with mouse anti-GFP, rabbit anti-RFA (1:1000; Agrisera; Cat#. AS07 214) or mouse anti-PGK1 (1:500; Santa Cruz; Cat#. Sc130335) antibodies, as described previously (Gallina et al., 2015).

Protein stability timecourse

Overnight cultures of the indicated strains were diluted to below an OD600 nm of 0.2 and allowed to progress into logarithmic phase before collection. The cells were treated with or without cycloheximide (CHX) at a concentration of 200 μg/ml for the indicated times (0–120 min). The final samples of 2×107 cells were collected by centrifugation, and whole-cell extracts were prepared by TCA extraction and used for immunoblotting.

Splicing efficiency assay

The splicing assay protocol was adapted and performed as described previously (Galy et al., 2004). All measurements were taken with three individual transformants per replicate for a total of five replicates. Cells were struck as a patch on SC –leucine plates, then replica plated to glycerol–lactate-containing SC medium (1.6% v/v glycerol and 1.5% w/v lactate) without leucine (GGL −leu). Cells from each patch were inoculated in liquid GGL −leu medium for 2 h at 34°C and induced with galactose (2% final concentration) for 1.5 h before treatment with MMS (0.05% final concentration) for 30 min. Cells were lysed and assayed for β-galactosidase using a Gal-Screen β-galactosidase reporter gene assay system for yeast or mammalian cells (Applied Biosystems), as per the manufacturer's instructions, and read with SpectraMax i3 (Molecular Devices). Relative light units were normalized to cell concentration as estimated by measuring OD600.

Ni-NTA pulldown of SUMOylated proteins

SUMO-conjugated proteins were detected by performing protein pulldowns under denaturing conditions, as described previously (Ulrich and Davies, 2009). In brief, strains of interest were transformed with a plasmid containing His-tagged SUMO (Smt3–His×7; Addgene) under the control of a copper-inducible promoter. Logarithmically growing cells were harvested at an OD600 nm of 0.6–0.8, and a total of 109 cells were collected by centrifugation (4000 rpm, 3724 g; 5 min at 4°C). Pelleted cells were washed and resuspended in 5 ml of pre-chilled water. Cells were then lysed with 800 µl of 1.85 M NaOH containing 7.5% (v/v) 2-mercaptoethanol, followed by 20 min incubation on ice. Protein precipitation was carried out by adding 800 µl 55% TCA on ice for 20 min. Precipitates were collected by centrifugation (8000 g, 20 min, 4°C) and were resuspended in 1 ml Buffer A (6 M guanidine hydrochloride, 100 mM sodium phosphate and 10 mM Tris-HCl, pH 8.0) and incubated on a rotating block at room temperature for 1 h to solubilize the precipitate completely. The resulting solution was then centrifuged (16,000 g, 10 min, 4°C), and supernatant was transferred to tubes containing 30 µl precleared Ni-NTA agarose beads (Qiagen) in the presence of 0.05% Tween-20 and 15 mM imidazole. These were incubated overnight on a rotating block at room temperature. Following binding, the beads were washed twice with Buffer A containing 0.05% Tween-20 and four times with Buffer C (8 M urea, 100 mM sodium phosphate and 10 mM Tris-HCl, pH 6.3) containing 0.05% Tween-20. Beads were centrifuged (200 g, 15 s) and supernatant was completely removed. His–SUMO conjugates on the beads were eluted by adding 30 µl HU sample buffer (8 M urea, 5% SDS, 200 mM Tris-HCl, pH 6.8, 0.1 mM EDTA, 0.5% Bromophenol Blue and 15 mg/ml dithiothreitol) and heating at 70°C for 10 min. Resultant protein extracts were subjected to standard western blotting and probed for SUMO and other proteins of interest using antibodies as described above.

Author contributions

Conceptualization: V.M., P.C.S.; Methodology: V.M., A.K.; Investigation: V.M., A.K., Y.K.J., K.W., A.S.T.; Writing - original draft: V.M., P.C.S.; Writing - review & editing: V.M., P.C.S.; Visualization: A.K., A.S.T., P.C.S.; Supervision: P.C.S.; Project administration: P.C.S.; Funding acquisition: P.C.S.

Funding

P.C.S. is funded by a Natural Sciences and Engineering Research Council of Canada Discovery grant and Discovery Accelerator Supplement (RPGIN 2020-04360), and a Canadian Institutes of Health Research Project grant (MOP136982). P.C.S. is a Canadian Institutes of Health Research New Investigator and Michael Smith Foundation for Health Research Scholar.

Amunugama
,
R.
,
Willcox
,
S.
,
Wu
,
R. A.
,
Abdullah
,
U. B.
,
El-Sagheer
,
A. H.
,
Brown
,
T.
,
McHugh
,
P. J.
,
Griffith
,
J. D.
and
Walter
,
J. C.
(
2018
).
Replication fork reversal during DNA interstrand crosslink repair requires CMG unloading
.
Cell Rep.
23
,
3419
-
3428
.
Bergink
,
S.
,
Ammon
,
T.
,
Kern
,
M.
,
Schermelleh
,
L.
,
Leonhardt
,
H.
and
Jentsch
,
S.
(
2013
).
Role of Cdc48/p97 as a SUMO-targeted segregase curbing Rad51-Rad52 interaction
.
Nat. Cell Biol.
15
,
526
-
532
.
Brandman
,
O.
,
Stewart-Ornstein
,
J.
,
Wong
,
D.
,
Larson
,
A.
,
Williams
,
C. C.
,
Li
,
G.-W.
,
Zhou
,
S.
,
King
,
D.
,
Shen
,
P. S.
,
Weibezahn
,
J.
, et al. 
(
2012
).
A ribosome-bound quality control complex triggers degradation of nascent peptides and signals translation stress
.
Cell
151
,
1042
-
1054
.
Defenouillère
,
Q.
,
Yao
,
Y.
,
Mouaikel
,
J.
,
Namane
,
A.
,
Galopier
,
A.
,
Decourty
,
L.
,
Doyen
,
A.
,
Malabat
,
C.
,
Saveanu
,
C.
,
Jacquier
,
A.
, et al. 
(
2013
).
Cdc48-associated complex bound to 60S particles is required for the clearance of aberrant translation products
.
Proc. Natl. Acad. Sci. USA
110
,
5046
-
5051
.
den Brave
,
F.
,
Cairo
,
L. V.
,
Jagadeesan
,
C.
,
Ruger-Herreros
,
C.
,
Mogk
,
A.
,
Bukau
,
B.
and
Jentsch
,
S.
(
2020
).
Chaperone-mediated protein disaggregation triggers proteolytic clearance of intra-nuclear protein inclusions
.
Cell Rep.
31
,
107680
.
Escusa-Toret
,
S.
,
Vonk
,
W. I. M.
and
Frydman
,
J.
(
2013
).
Spatial sequestration of misfolded proteins by a dynamic chaperone pathway enhances cellular fitness during stress
.
Nat. Cell Biol.
15
,
1231
-
1243
.
Frottin
,
F.
,
Schueder
,
F.
,
Tiwary
,
S.
,
Gupta
,
R.
,
Körner
,
R.
,
Schlichthaerle
,
T.
,
Cox
,
J.
,
Jungmann
,
R.
,
Hartl
,
F. U.
,
Hipp
,
M. S.
, et al. 
(
2019
).
The nucleolus functions as a phase-separated protein quality control compartment
.
Science
365
,
342
-
347
.
Gallagher
,
P. S.
,
Clowes Candadai
,
S. V.
and
Gardner
,
R. G.
(
2014
).
The requirement for Cdc48/p97 in nuclear protein quality control degradation depends on the substrate and correlates with substrate insolubility
.
J. Cell Sci.
127
,
1980
-
1991
.
Gallina
,
I.
,
Colding
,
C.
,
Henriksen
,
P.
,
Beli
,
P.
,
Nakamura
,
K.
,
Offman
,
J.
,
Mathiasen
,
D. P.
,
Silva
,
S.
,
Hoffmann
,
E.
,
Groth
,
A.
, et al. 
(
2015
).
Cmr1/WDR76 defines a nuclear genotoxic stress body linking genome integrity and protein quality control
.
Nat. Commun.
6
,
6533
.
Galy
,
V.
,
Gadal
,
O.
,
Fromont-Racine
,
M.
,
Romano
,
A.
,
Jacquier
,
A.
and
Nehrbass
,
U.
(
2004
).
Nuclear retention of unspliced mRNAs in yeast is mediated by perinuclear Mlp1
.
Cell
116
,
63
-
73
.
Hartl
,
F. U.
,
Bracher
,
A.
and
Hayer-Hartl
,
M.
(
2011
).
Molecular chaperones in protein folding and proteostasis
.
Nature
475
,
324
-
332
.
Ho
,
C.-T.
,
Grousl
,
T.
,
Shatz
,
O.
,
Jawed
,
A.
,
Ruger-Herreros
,
C.
,
Semmelink
,
M.
,
Zahn
,
R.
,
Richter
,
K.
,
Bukau
,
B.
,
Mogk
,
A.
, et al. 
(
2019
).
Cellular sequestrases maintain basal Hsp70 capacity ensuring balanced proteostasis
.
Nat. Commun.
10
,
4851
.
Kaganovich
,
D.
,
Kopito
,
R.
and
Frydman
,
J.
(
2008
).
Misfolded proteins partition between two distinct quality control compartments
.
Nature
454
,
1088
-
1095
.
Khmelinskii
,
A.
,
Keller
,
P. J.
,
Bartosik
,
A.
,
Meurer
,
M.
,
Barry
,
J. D.
,
Mardin
,
B. R.
,
Kaufmann
,
A.
,
Trautmann
,
S.
,
Wachsmuth
,
M.
,
Pereira
,
G.
, et al. 
(
2012
).
Tandem fluorescent protein timers for in vivo analysis of protein dynamics
.
Nat. Biotechnol.
30
,
708
-
714
.
Kumar
,
R.
,
Nawroth
,
P. P.
and
Tyedmers
,
J.
(
2016
).
Prion aggregates are recruited to the insoluble protein deposit (IPOD) via Myosin 2-based vesicular transport
.
PLoS Genet.
12
,
e1006324
.
Leung
,
G. P.
,
Brown
,
J. A. R.
,
Glover
,
J. N. M.
and
Kobor
,
M. S.
(
2016
).
Rtt107 BRCT domains act as a targeting module in the DNA damage response
.
DNA Repair
37
,
22
-
32
.
Li
,
Z.
,
Vizeacoumar
,
F. J.
,
Bahr
,
S.
,
Li
,
J.
,
Warringer
,
J.
,
Vizeacoumar
,
F. S.
,
Min
,
R.
,
Vandersluis
,
B.
,
Bellay
,
J.
,
Devit
,
M.
, et al. 
(
2011
).
Systematic exploration of essential yeast gene function with temperature-sensitive mutants
.
Nat. Biotechnol.
29
,
361
-
367
.
Luisier
,
R.
,
Tyzack
,
G. E.
,
Hall
,
C. E.
,
Mitchell
,
J. S.
,
Devine
,
H.
,
Taha
,
D. M.
,
Malik
,
B.
,
Meyer
,
I.
,
Greensmith
,
L.
,
Newcombe
,
J.
, et al. 
(
2018
).
Intron retention and nuclear loss of SFPQ are molecular hallmarks of ALS
.
Nat. Commun.
9
,
2010
.
Maric
,
M.
,
Mukherjee
,
P.
,
Tatham
,
M. H.
,
Hay
,
R.
and
Labib
,
K.
(
2017
).
Ufd1-Npl4 Recruit Cdc48 for disassembly of ubiquitylated CMG helicase at the end of chromosome replication
.
Cell Rep.
18
,
3033
-
3042
.
Mathew
,
V.
,
Tam
,
A. S.
,
Milbury
,
K. L.
,
Hofmann
,
A. K.
,
Hughes
,
C. S.
,
Morin
,
G. B.
,
Loewen
,
C. J. R.
and
Stirling
,
P. C.
(
2017
).
Selective aggregation of the splicing factor Hsh155 suppresses splicing upon genotoxic stress
.
J. Cell Biol.
216
,
4027
-
4040
.
Miller
,
S. B. M.
,
Ho
,
C.-T.
,
Winkler
,
J.
,
Khokhrina
,
M.
,
Neuner
,
A.
,
Mohamed
,
M. Y. H.
,
Guilbride
,
D. L.
,
Richter
,
K.
,
Lisby
,
M.
,
Schiebel
,
E.
, et al. 
(
2015
).
Compartment-specific aggregases direct distinct nuclear and cytoplasmic aggregate deposition
.
EMBO J.
34
,
778
-
797
.
Pantazopoulou
,
M.
,
Boban
,
M.
,
Foisner
,
R.
and
Ljungdahl
,
P. O.
(
2016
).
Cdc48 and Ubx1 participate in a pathway associated with the inner nuclear membrane that governs Asi1 degradation
.
J. Cell Sci.
129
,
3770
-
3780
.
Psakhye
,
I.
and
Jentsch
,
S.
(
2012
).
Protein group modification and synergy in the SUMO pathway as exemplified in DNA repair
.
Cell
151
,
807
-
820
.
Ruan
,
L.
,
Zhou
,
C.
,
Jin
,
E.
,
Kucharavy
,
A.
,
Zhang
,
Y.
,
Wen
,
Z.
,
Florens
,
L.
and
Li
,
R.
(
2017
).
Cytosolic proteostasis through importing of misfolded proteins into mitochondria
.
Nature
543
,
443
-
446
.
Saarikangas
,
J.
and
Barral
,
Y.
(
2016
).
Protein aggregation as a mechanism of adaptive cellular responses
.
Curr. Genet.
62
,
711
-
724
.
Saugar
,
I.
,
Jiménez-Martín
,
A.
and
Tercero
,
J. A.
(
2017
).
Subnuclear relocalization of structure-specific endonucleases in response to DNA damage
.
Cell Rep.
20
,
1553
-
1562
.
Sherman
,
F.
(
2002
).
Getting started with yeast. Methods Enzymol. 350, 3-41.
Simões
,
T.
,
Schuster
,
R.
,
den Brave
,
F.
and
Escobar-Henriques
,
M.
(
2018
).
Cdc48 regulates a deubiquitylase cascade critical for mitochondrial fusion
.
eLife
7
,
e30015
.
Sontag
,
E. M.
,
Vonk
,
W. I. M.
and
Frydman
,
J.
(
2014
).
Sorting out the trash: the spatial nature of eukaryotic protein quality control
.
Curr. Opin. Cell Biol.
26
,
139
-
146
.
Stirling
,
P. C.
,
Bloom
,
M. S.
,
Solanki-Patil
,
T.
,
Smith
,
S.
,
Sipahimalani
,
P.
,
Li
,
Z.
,
Kofoed
,
M.
,
Ben-Aroya
,
S.
,
Myung
,
K.
,
Hieter
,
P.
, et al. 
(
2011
).
The complete spectrum of yeast chromosome instability genes identifies candidate CIN cancer genes and functional roles for ASTRA complex components
.
PLoS Genet.
7
,
e1002057
.
Tam
,
A. S.
,
Sihota
,
T. S.
,
Milbury
,
K. L.
,
Zhang
,
A.
,
Mathew
,
V.
and
Stirling
,
P. C.
(
2019
).
Selective defects in gene expression control genome instability in yeast splicing mutants
.
Mol. Biol. Cell
30
,
191
-
200
.
Thattikota
,
Y.
,
Tollis
,
S.
,
Palou
,
R.
,
Vinet
,
J.
,
Tyers
,
M.
and
D'Amours
,
D.
(
2018
).
Cdc48/VCP promotes chromosome morphogenesis by releasing condensin from self-entrapment in chromatin
.
Mol. Cell
69
,
664
-
676.e5
.
Tkach
,
J. M.
,
Yimit
,
A.
,
Lee
,
A. Y.
,
Riffle
,
M.
,
Costanzo
,
M.
,
Jaschob
,
D.
,
Hendry
,
J. A.
,
Ou
,
J.
,
Moffat
,
J.
,
Boone
,
C.
, et al. 
(
2012
).
Dissecting DNA damage response pathways by analysing protein localization and abundance changes during DNA replication stress
.
Nat. Cell Biol.
14
,
966
-
976
.
Ulrich
,
H. D.
and
Davies
,
A. A.
(
2009
).
In vivo detection and characterization of sumoylation targets in Saccharomyces cerevisiae
.
Methods Mol. Biol.
497
,
81
-
103
.
Verma
,
R.
,
Oania
,
R.
,
Fang
,
R.
,
Smith
,
G. T.
and
Deshaies
,
R. J.
(
2011
).
Cdc48/p97 mediates UV-dependent turnover of RNA Pol II
.
Mol. Cell
41
,
82
-
92
.
Wallace
,
E. W. J.
,
Kear-Scott
,
J. L.
,
Pilipenko
,
E. V.
,
Schwartz
,
M. H.
,
Laskowski
,
P. R.
,
Rojek
,
A. E.
,
Katanski
,
C. D.
,
Riback
,
J. A.
,
Dion
,
M. F.
,
Franks
,
A. M.
, et al. 
(
2015
).
Reversible, specific, active aggregates of endogenous proteins assemble upon heat stress
.
Cell
162
,
1286
-
1298
.
Yoshida
,
K.
and
Ogawa
,
S.
(
2014
).
Splicing factor mutations and cancer
.
Wiley Interdiscip. Rev. RNA
5
,
445
-
459
.
Zhou
,
C.
,
Slaughter
,
B. D.
,
Unruh
,
J. R.
,
Guo
,
F.
,
Yu
,
Z.
,
Mickey
,
K.
,
Narkar
,
A.
,
Ross
,
R. T.
,
McClain
,
M.
,
Li
,
R.
, et al. 
(
2014
).
Organelle-based aggregation and retention of damaged proteins in asymmetrically dividing cells
.
Cell
159
,
530
-
542
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information