Abstract
Protein ubiquitination is essential to govern cells’ ability to cope with harmful environments by regulating many aspects of protein dynamics from synthesis to degradation. As important as the ubiquitination process, the reversal of ubiquitin chains mediated by deubiquitinating enzymes (DUBs) is critical for proper recovery from stress and re-establishment of proteostasis. Although it is known that ribosomes are decorated with K63-linked polyubiquitin chains that control protein synthesis under stress, the mechanisms by which these ubiquitin chains are reversed and regulate proteostasis during stress recovery remain elusive. Here, we showed in budding yeast that the DUB Ubp2 is redox-regulated during oxidative stress in a reversible manner, which determines the levels of K63-linked polyubiquitin chains present on ribosomes. We also demonstrate that Ubp2 can cleave single ubiquitin moieties out of chains and its activity is modulated by a series of repeated domains and the formation of disulfide bonds. By combining cellular, biochemical, and proteomics analyses, we showed that Ubp2 is crucial for restoring translation after stress cessation, indicating an important role in determining the cellular response to oxidative stress. Our work demonstrates a novel role for Ubp2, revealing that a range of signaling pathways can be controlled by redox regulation of DUB activity in eukaryotes, which in turn will define cellular states of health and diseases.
Keywords: deubiquitylation (deubiquitination), redox regulation, translation control, oxidative stress, yeast
Eukaryotic cells are constantly exposed to ever-changing environments, in which they must reprogram several physiological pathways to adapt and thrive. Under stress conditions, cells regulate gene expression at the transcription and translation levels, in addition to reshaping the functional proteome through posttranslation modifications (PTMs), protein interactions, and degradation (1, 2). Several studies have focused on understanding a myriad of regulatory processes that occur upon stress induction (2, 3, 4), but the mechanisms by which cells recover from stress and re-establish proteostasis have remained more elusive. In response to oxidative stress, we observed in the budding yeast Saccharomyces cerevisiae that ribosomes are rapidly and heavily decorated with K63-linked polyubiquitin (K63-ub) chains (5, 6, 7), a less conventional type of polyubiquitin chain that functions independently of the proteasome (8, 9). We showed that the accumulation of these ubiquitin (Ub) chains on ribosomes during oxidative stress is required to pause translation globally at the elongation stage (7, 10, 11). We also determined that ribosome ubiquitination is dependent on the E2 Ub-conjugating enzyme Rad6, which is itself redox-regulated and whose activity is critical for reprogramming protein synthesis upon oxidative stress induction (5, 11). Although we have identified important genes and characterized new processes in the early stage of this pathway of redox control of translation by ubiquitin (RTU), the mechanisms involved in the reversal of this process remain largely unknown.
Ubiquitination is a highly dynamic process, and during cellular recovery from oxidative stress, we have shown that the reversal of K63-ub chains is impaired in the absence of the deubiquitinating enzyme (DUB) Ubp2 (5). Ubp2 is a multifunctional DUB largely known for its role in regulating and antagonizing the E3 Ub ligase Rsp5 during the trafficking of various types of membrane-associated cargoes (12, 13, 14). Ubp2 is also known to participate in mitochondria homeostasis, protein quality control, and DNA damage response (15, 16, 17, 18, 19). Ubp2 is the largest DUB encoded in yeast (146 kDa), and even though it participates in several physiological processes, the structural and functional mechanisms regulating Ubp2’s activity during the stress response are not well understood.
DUBs are proteases that cleave the isopeptide bond between the C-terminus glycine of Ub and the ε-NH2 group of a lysine residue in a substrate or Ub itself when in the form of a chain. DUBs can trim, remove, or even remodel selective Ub chains (20, 21, 22). In addition, DUBs are functionally classified based on their catalytic activity, which are divided into cysteine proteases or metalloproteases (20, 23, 24). Cysteine proteases are far more abundant, and because of their functional and structural diversity, several mechanisms of regulation have been identified, including PTMs, allosteric regulation, protein trafficking, and/or protein–protein interaction (20). In addition, because of their catalytic cysteine, DUBs can have their activity regulated by the oxidation of key cysteine residues (20, 25, 26). A previous study has shown that the activity of DUBs can be either unaffected, enhanced, or activated by the thiol-reducing agent DTT, suggesting that DUB’s protein sequences and structural features can determine their sensitivity to reactive oxygen species (ROS) (27). However, the nature of their regulation and the physiological impact of these redox modifications remain largely unknown. As a cysteine protease from the USP/UBP family, redox regulation of Ubp2 can be critical to regulate proteostasis during cellular stress.
Here, we show that the activity of the cysteine DUB Ubp2 is inhibited by hydrogen peroxide (H2O2) and reversibly regulated during oxidative stress recovery. Furthermore, reactivation of Ubp2 is required to reverse K63-ub chains, and its activity is controlled by selective repeated domains and reactive cysteine residues. Finally, we show that Ubp2 is critical for supporting translation reprogramming after stress cessation. Our work provides new insights on the redox regulation of Ubp2 and proposes new models by which several Ub-mediated pathways could be rapidly induced by common and prominent stressors, with a meaningful impact on cellular physiology and health.
Results
Ubp2 is a linkage-specific DUB regulated by ROS
We previously identified that the deletion of UBP2 impairs the cellular capacity to reverse K63-ub chains that accumulate in response to oxidative stress in yeast (5). However, it remained unknown whether Ubp2 is directly responsible for the cleavage of K63-ub chains as part of the RTU and whether Ubp2’s activity could be regulated by ROS. Expression of WT Ubp2 in ubp2Δ rescues the deubiquitination phenotype, which reverses K63-ub globally and from ribosomes after stress induction (Figs. 1A and S1). By mutating Ubp2’s catalytic cysteine residue to serine (C745S), we confirmed that Ubp2’s activity is required for the reversal of K63-ub from ribosomes that accumulate under H2O2 (Fig. 1A). To further test Ubp2’s role in directly cleaving these K63-ub chains, we purified recombinant Ubp2 and showed that Ubp2 acts preferentially on K63-ub in comparison to K48-ub chains (Figs. 1B and S2). We also showed that Ubp2 is able to generate degradation intermediates of varied lengths when tested against synthetic tetra- and hexa-K63-ub chain substrates (Fig. 1C). Combined, our results suggest that Ubp2 is able to specifically trim K63-ub directly from its targets. Given that Ubp2 is a cysteine protease, we hypothesized that the accumulation of K63-ub under oxidative stress occurs due to oxidative inactivation of Ubp2. To test this possibility, we performed in vitro activity assays, and we observed that the catalytic activity of Ubp2 against the fluorogenic substrate Ub-Rho can be inhibited by H2O2 in a dose-dependent manner (Figs. 1D and S3, A and B). Cleavage of the substrate by DUBs releases the product rhodamine 110, which can be determined fluorimetrically. As expected, the pan DUB inhibitor PR-619 and the cysteine alkylator iodoacetamide (IAA) abrogated Ubp2 enzymatic activity (Figs. 1D and S3C). Importantly, oxidative inhibition of Ubp2 activity can be rescued by the reducing agent DTT, which indicates a reversible regulatory process (Figs. 1D and S3D). These results suggest that H2O2 promotes the oxidation of Ubp2’s catalytic cysteine residue, which can be reversed by thiol reducing agents. This redox inhibition also occurs when Ubp2 is tested against synthetic K63-ub chains (Fig. 1E). While we observed a redox regulation of recombinant Ubp2 in vitro, it was unclear whether this process also occurred in the cellular context. To test this, we immunoprecipitated hemagglutinin (HA)-tagged Ubp2 from yeast cells subjected to oxidative stress and we showed that Ubp2’s activity is partially inhibited by H2O2, which can also be rescued by DTT (Fig. 1F). HA-immunoprecipitation from cells expressing the catalytic dead Ubp2C745S did not display activity against Ub-7-amido-4-methylcoumarin (AMC), demonstrating that this deubiquitinating activity was produced by Ubp2 and not by other contaminant proteases (Fig. 1F). Therefore, our results indicate that Ubp2’s activity is redox-regulated by H2O2, and Ubp2 is responsible for the reversal of the K63-ub chains that accumulate on ribosomes under stress.
Figure 1.
Ubp2 promotes K63-ub cleavage after stress and is reversibly inhibited by H2O2.A, deletion and mutation of UBP2 impairs K63-ub chain removal from ribosomes. Immunoblot anti-K63-ub of isolated ribosomes from cells treated in the presence or absence of 0.6 mM of H2O2 for 30 min (H2O2), followed by 20 min recovery (Rec) in fresh medium. Anti-uL5/Rpl11 was used as loading control for isolated ribosomes and anti-HA for Ubp2 (n = 2). B, Ubp2 preferentially cleaves K63-ub chains. Immunoblot anti-ub of synthetic tetra K63-ub (250 ng) or K48-ub chains (250 ng) incubated in the presence or absence of 3 μg of purified Ubp2 or pan DUB USP2 (n = 2). C, immunoblot anti-ub of synthetic K63-ub chains (250 ng) incubated in the presence or absence of 3 μg of purified Ubp2. Arrows indicate degradation products (n = 2). D, Ubp2’s activity is redox-regulated in vitro. Activity of recombinant Ubp2 (70 ng) was assessed in vitro using 0.75 μM of Ub-Rho fluorophore. Ubp2 was incubated for 5 min with 100 μM or 500 μM of H2O2, 20 mM IAA (cysteine alkylator), or 100 μM PR-619 (general DUB inhibitor). Two millimolars of DTT (reducing agent) was added after 5 min of assay to restore Ubp2 activity. Rho 110 fluorescence was recorded at 535 nm with excitation at 485 nm (n = 2, two technical replicates each). E, Ubp2 deubiquitinating activity is impaired by H2O2. Immunoblot anti-ub of synthetic tetra K63-ub chains (250 ng) incubated in the presence or absence of 3 μg purified Ubp2 with the addition or absence of 7 mM H2O2, 10 mM DTT, or 10 mM IAA (n = 3). F, Ubp2 activity is redox-regulated in yeast cells under stress. Ubp2-HA and Ubp2C745S-HA were immunoprecipitated from yeast cells untreated or treated with 0.6 mM H2O2 for 30 min. Ubp2 activity was assessed in vitro using 1.5 μM of Ub-AMC fluorophore. 20 mM of DTT was added after 60 min as indicated by the arrow to restore Ubp2 activity. Fluorescence was recorded at 445 nm with excitation at 345 nm (n = 2). Reactions shown in (B, C, and E) were incubated for 1 h at 30 °C prior SDS-PAGE and immunoblotting. DUB, deubiquitinating enzyme; H2O2, hydrogen peroxide; HA, hemagglutinin; IAA, iodoacetamide; K63-ub, K63-linked polyubiquitin; MW, molecular weight; Ub, ubiquitin.
Ubp2 has different sensitivity to peroxides
To further understand the redox regulation of Ubp2, we tested whether Ubp2 activity was responsive to other peroxides beyond H2O2. To different extents, we observed that organic peroxides are also able to induce the accumulation of K63-ub chains in yeast (Fig. 2A). As DUBs are a diverse group of enzymes, their sensitivity to ROS can vary substantially. To understand the impact of ROS on DUBs, we determined cellular DUB activity by incubating the whole cell extract or purified proteins with the DUB fluorescent substrate Ub-Rho. We first showed that cellular DUB activity was more affected by the organic peroxides cumene (CHP) and tert-butyl hydroperoxide (t-BHP) than H2O2 at our standard concentration of 0.6 mM, with CHP promoting the strongest effect (Fig. 2B, left). At 2.5 mM, organic and inorganic peroxides had a similar and acute effect on global DUB activity (Fig. 2B, right). These results disagreed with our previous findings that showed higher levels of K63-ub under H2O2 than organic peroxides (Fig. 2A). Therefore, we tested whether Ubp2 itself would be more sensitive to H2O2, while global DUB activity would be more affected by organic peroxides. Using purified Ubp2, we confirmed that H2O2 promoted a stronger reduction of Ubp2 activity (∼60%), while organic peroxide treatments showed only a small reduction at the same concentration (Fig. 2C, ∼5–20%). Upon increasing concentrations, organic peroxides promoted a stronger inhibition of Ubp2 activity (Fig. S4A). Thus, our findings show that Ubp2 can be redox-regulated by organic and inorganic peroxides, and its sensitivity correlates to the levels of K63-ub chains observed in yeast cells (Figs. 2, A and C and S4A). Finally, because several DUBs are cysteine proteases (28), this redox regulation could affect many Ub-mediated pathways in eukaryotic cells at once. Thus, we tested whether Cezanne, a human K11 linkage-specific DUB involved in cell cycle regulation (29), could also be redox-regulated by peroxides. Indeed, we observed an increased sensitivity of the catalytic domain of Cezanne (CezanneCAT) to H2O2, in comparison to t-BHP and CHP (Figs. 2D and S4B). CezanneCAT activity can also be rescued by DTT (Fig. S4C). Therefore, the redox mechanisms that we are exploring for Ubp2 could have further implications for Ub dynamics in a plethora of eukaryotic pathways that are responsive to ROS.
Figure 2.
Ubp2 is sensitive to H2O2.A, accumulation of K63-ub chains is differentially induced by H2O2 and organic peroxides. Immunoblot anti-K63-ub from WT cells exposed to 0.6 mM and 2.5 mM of H2O2, or organic peroxides tert-butyl (t-BHP) and cumene hydroperoxide (CHP) for 30 min. Anti-PGK1 was used as a loading control (n = 3). B, cellular deubiquitinating activity is differentially affected by H2O2 and organic peroxides. WT cells untreated (UT) or treated with either 0.6 mM (left) or 2.5 mM (right) of peroxides (ROOH) for 30 min were lysed, and total DUB activity was determined by incubating the whole-cell extract with the fluorogenic substrate Ub-Rho. Total DUB activity was assessed over time using 0.75 μM of Ub-Rho, whose fluorescence was determined at 535 nm with excitation at 485 nm. Signal was normalized by protein concentration and a standard curve of free rhodamine 110 (n = 2, two technical replicates each). C and D, Ubp2 and CezanneCAT are sensitive to H2O2. C, purified Ubp2 (70 ng) and D, CezanneCAT (3 μg) were incubated with 500 μM H2O2, t-BHP, and CHP for 5 min and DUB activity was assessed as described above. Activity was normalized by Rho 110 fluorescence registered for untreated Ubp2 or CezanneCAT. Bar graphs show mean values ± SD for three biological replicates. CezanneCAT activity was performed in duplicate and error bars represent standard error. Significance was calculated using an unpaired two-tailed Student’s t test, where ∗p < 0.05 and ∗∗p < 0.005. DUB, deubiquitinating enzyme; H2O2, hydrogen peroxide; K63-ub, K63-linked polyubiquitin; MW, molecular weight.
Ubp2 is reactivated during stress recovery
Although we observed that Ubp2 activity can be reversibly regulated by ROS (Fig. 1, D and F), it remained unclear whether this process is required to reverse the K63-ub chains that accumulate under stress. As the intracellular fate of oxidized Ubp2 is unknown, the reversal of K63-ub chains during stress recovery could be mediated either by Ubp2 reactivation or by de novo Ubp2 synthesis. Using the translation inhibitor cycloheximide (CHX), we first showed that the degradation of Ubp2 is not enhanced upon stress induction, and Ubp2 remains present and abundant during the time required for deubiquitination (Fig. 3A). Moreover, yeast cells were able to cleave K63-ub chains even in the presence of CHX, suggesting that de novo synthesis of Ubp2 is not required for K63-ub reversal (Fig. 3A). Accordingly, we showed that cells treated with CHX can regain their DUB activity during stress recovery, further indicating that redox reactivation of DUBs occurs in the cellular context (Fig. 3B). Thus, our data suggest that inactivation and reactivation of Ubp2 are critical to controlling the levels of K63-ub chains following H2O2 stress. Although cysteine oxidation is largely a chemical process (30), intracellular reduction of these residues commonly relies on the glutathione or thioredoxin system as the main thiol-based antioxidant pathways in eukaryotes (31, 32). Therefore, we tested whether alterations of the cellular redox balance could also affect the dynamics of K63-ub chains during stress. Double deletion of thioredoxins (TRX1/TRX2), glutaredoxins (GRX1/GRX2), and single deletion of thioredoxin peroxidases (TSA1 and TSA2) did not show sustained high levels of K63-ub chains during recovery (Figs. 3C and S5). However, deletion of glutathione reductase consistently impaired the cell’s ability to reverse K63-ub chains during stress recovery (Fig. 3C). Our findings show that Ubp2 is reactivated during the recovery phase of stress in a redox-dependent manner and suggest that cell’s reductive capacity, mediated by NADPH-dependent redox systems, participates in this process.
Figure 3.
Ubp2 is reactivated during stress recovery.A, K63-ub chains are reversed independently of translation. Immunoblot anti-K63-ub from cells exposed to 0.6 mM H2O2 in the presence or absence of 150 μg/ml cycloheximide (CHX). Anti-HA shows Ubp2 levels and anti-PGK1 was used as a loading control (n ≥ 3). B, cellular DUB activity recovers from stress independently of translation. DUB activity from WT cells exposed to 0.6 mM H2O2 was assessed over time using 0.75 μM Ub-Rho 110 in the absence (left) or presence (right) of 150 μg/ml CHX. Rho 110 fluorescence was recorded at 535 nm with excitation at 485 nm (n = 2, two technical replicates each). C, reversal of K63-ub chains relies on cellular antioxidant systems. Immunoblot anti-K63-ub from strains incubated with 0.6 mM H2O2. Anti-GAPDH was used as a loading control (n = 3). DUB, deubiquitinating enzyme; H2O2, hydrogen peroxide; HA, hemagglutinin; K63-ub, K63-linked polyubiquitin; MW, molecular weight.
Next, we investigated the role of Ubp2 cysteine residues (Cys) on its redox regulation. Beyond their catalytic cysteine, thiol-dependent enzymes can use additional Cys to aid in their catalytic cycle and redox regulation (30, 31). Ubp2 has 13 cysteine residues, and analysis of its 3D model from AlphaFold2 (33, 34) predicted that the cysteine residues C821 and C944 are in proximity to the catalytic site, which could foster the formation of disulfide bonds upon conformational changes (Fig. 4A). By running nonreducing gels, we observed that Ubp2 forms H2O2-dependent disulfide bonds (Fig. S6, A–C) that are fully reduced by DTT (Fig. S6C). Mutational analysis revealed that the disulfide bond formation was abrogated in the Ubp2C754S and Ubp2C944S mutants following H2O2 treatment, indicating that these residues are involved in disulfide bond formation (Fig. 4B). To determine the effect of these cysteine residues on Ubp2’s activity, we evaluated the cellular growth of strains carrying selective Ubp2 mutations through their cellular sensitivity to the proteotoxic agent L-azetidine-2-carboxylic acid (ADCB). Functional Ubp2 provides resistance to ADCB by deubiquitinating and recycling selective ADCB transporters (13, 35). First, we showed that cells expressing the catalytic dead Ubp2C745S mutant are highly sensitive, while cells expressing WT Ubp2 show tolerance to ADCB (Fig. 4C). Next, we tested whether mutation to other cysteine residues affects cell’s sensitivity to ADCB. While we observed a mild effect for the double mutation C821S/C944S, mutations to cysteine C621 and C645 show no change in ADCB sensitivity (Fig. 4C). Importantly, no growth defects were observed in the absence of ADCB (Fig. S7). Although the expression of the Ubp2 C821S/C944S mutant blocks the formation of Ubp2 disulfide bond and presents sensitivity to ADCB, this strain was still able to remove K63-ub chains in a timely fashion (Fig. 4D). As our experiments are not able to capture all Ubp2 states of cysteine oxidation, which might vary under different phases and magnitudes of stress, the functional role of Cys944 and its oxidation remain to be elucidated. Under the conditions tested, our data show that Ubp2 is reactivated upon stress cessation and is able to form disulfides bonds in the presence of ROS, which correlates to the levels of K63-ub chains that accumulate under stress.
Figure 4.
Ubp2 catalytic cysteine forms disulfide bonds under stress.A, AlphaFold2 structural 3D model of Ubp2 (ID: Q01476). N terminus (gray), repeated domain (RD) 1 (green), RD2 (blue), RD3 (pink), and C terminus (yellow). Inset highlights in orange the catalytic cysteine (C745) and neighboring cysteine residues (C821 and C944) in the catalytic domain with their predicted molecular distances. B, mutation to catalytic cysteine C745 and C944 inhibits the formation of disulfide bond under stress. Immunoblot anti-HA shows Ubp2 levels from cells treated with 0.6 mM H2O2 for 30 min. Lysates were incubated in the presence or absence of 20 mM DTT prior to immunoblotting. Anti-PGK1 was used as a loading control (n ≥ 3). C, cellular sensitivity to the proteotoxic agent ADCB. Cell growth was monitored through absorbance (A600) measurements over time upon addition of 100 μg/ml ADCB. Ubp2FL and Ubp2C745S were used as positive and negative control, respectively. This graph represents at least three independent experiments, each performed with three technical replicates. D, comparative reversal of K63-ub chains during stress recovery. Immunoblot anti-K63-ub from cells expressing Ubp2-HA or its cysteine mutants upon treatment with 0.6 mM H2O2 for the designated times. Anti-GAPDH was used as a loading control (n = 3). ADCB, L-azetidine-2-carboxylic acid; H2O2, hydrogen peroxide; HA, hemagglutinin; K63-ub, K63-linked polyubiquitin; MW, molecular weight.
Ubp2 activity is regulated by a series of repeated domains
In addition to the involvement of key cysteine residues, additional functional elements are necessary to drive Ubp2’s deubiquitinating function. Ubp2 is comprised of a nonconserved N-terminus extension, three repeated domains (RDs), and a conserved C-terminus UBP/USP catalytic domain (Figs. 4A and 5A). While it is known that Ubp2 plays a role in several pathways mediated by K63-ub (5, 36), the protein domains and residues that are critical for its activity have only been broadly characterized in the context of its interaction with the E3 Rsp5 and the cofactor Rup1 in vesicle trafficking (37). As the reversal of K63-ub chains that accumulates under stress works independently of Rup1 (Fig. S8A), we sought to characterize the role of Ubp2’s domains in the RTU. To determine the functional role of Ubp2 domains, we turned again to ADCB, which allowed us to test several experimental conditions in parallel. We observed that expression of the N terminus only (Ubp2Nterm) or the catalytic domain only construct (Ubp2CAT) led to cellular sensitivity toward ADCB (Fig. 5B). Ubp2CAT still retained partial activity against the fluorogenic substrate Ub-AMC (Fig. S8B), suggesting that additional domains are important for Ubp2 cellular function. By producing additional truncated constructs, we showed that cells expressing Ubp2 lacking its N-term but expressing its three RDs and C terminus (Ubp2RD1-3+CAT) are sensitive to ADCB (Fig. 5C). Surprisingly, upon deletion of RD1 (Ubp2RD2-3+CAT), cellular resistance to ADCB was rescued similarly to levels observed in WT cells (Fig. 5C). We then removed Ubp2 RD2 (Ubp2RD3+CAT), which also showed an increased sensitivity to ADCB (Fig. 5C), suggesting that RD2 and the catalytic domain are required to promote resistance against ADCB. To test whether RD2 alone would be sufficient to support Ubp2 function, we created new constructs and showed that cells expressing RD2 with the catalytic domain (Ubp2RD2+CAT) remain sensitive to ADCB (Fig. 5D). Thus, our results indicate that RD2 is required but not sufficient to regulate Ubp2 activity. Addition of RD1 (Ubp2RD1-2+CAT) made cells more sensitive to ADCB again (Fig. 5D). The presence of RD1 consistently reduces Ubp2 activity, suggesting that it might serve as an inhibitory domain that could reduce the exposure of key catalytic residues to the substrate.
Figure 5.
Ubp2 is regulated by a series of repeated domains.A, schematic of Ubp2 domain organization and truncations created in this study. Ubp2 is comprised of a nonconserved N terminus, three repeated domains (RDs), and a conserved C-terminus UBP/USP catalytic domain. Nomenclature for Ubp2 truncations uses the remaining RDs followed by the catalytic domain (CAT). The only truncation that does not contain the CAT domain is the “N term.” Created with BioRender.com. B–D, Ubp2’s RDs modulate cell’s resistance against proteotoxic agent ADCB. ADCB sensitivity growth curves for the Ubp2 truncations as labeled above. Cell growth was monitored through absorbance (A600) over time upon addition of 100 μg/ml ADCB. Ubp2FL and Ubp2C745S were used as positive and negative controls in all growth curves, respectively. This graph represents at least three independent experiments, each performed with three technical replicates. E, Ubp2’s RDs regulate K63-ub chain removal during stress recovery. Immunoblots anti-K63-ub of ubp2Δ cells expressing episomal WT Ubp2, Ubp2RD1-3+CAT, and Ubp2RD2-3+CAT in the presence or absence of 0.6 mM H2O2. Anti-GAPDH was used as loading control (n = 3). ADCB, L-azetidine-2-carboxylic acid; H2O2, hydrogen peroxide; K63-ub, K63-linked polyubiquitin; MW, molecular weight.
Following our initial functional screen of Ubp2’s RDs, we asked whether these regulatory domains also affected the reversal of K63-ub chains that accumulate in response to stress. We have not identified any relationship of these domains with Ubp2’s capacity to bind or associate with ribosomes (Fig. S8, C and D), but they indeed affected the cell’s ability to reverse K63-ub chains. In agreement with the ADCB experiments, cells expressing Ubp2 containing all repeated domains (Ubp2RD1-3+CAT) were inefficient in reversing K63-ub chains, while deletion of RD1 (Ubp2RD2-3) also rescued the phenotype and contributed to reversal of these Ub chains (Fig. 5E). Supporting the notion that RD1 might play an inhibitory effect, we observed that Ubp2RD2-3+CAT forms high levels of disulfides after H2O2 (Figs. 5E and S6A), which are also prevented by the presence of RD1. Here, we showed that Ubp2 activity is regulated by a series of repeated domains that controls cellular resistance to ADCB and the levels of K63-ub chains in the RTU.
Ubp2 supports the re-establishment of translation upon stress cessation
We have previously shown that K63-ub chains that accumulate in response to oxidative stress modifies ribosomes and pauses translation at the elongation stage (5, 7, 10). Following oxidative stress induction, Ubp2 is required for Ub reversal, however, it remains unclear how Ubp2 regulates translation through its deubiquitinating activity. To start to elucidate the role of Ubp2 in translation, we first showed that purified Ubp2 is able to deubiquitinate isolated ribosomes in vitro, and the deubiquitination of ribosomes by Ubp2 can also be impaired by H2O2 (Figs. 6A and S9, A and B). Using polysome profiling, we showed that Ubp2 remains associated to the ribosome fraction in the presence or absence of stress (Fig. 6B). Additionally, we observed that cells lacking Ubp2 retained high levels of K63-ub chains in the monosome and polysome fractions during stress recovery (Fig. 6C), further suggesting that Ubp2 is responsible for removing K63-ub chains from elongating ribosomes. Because of the role of K63-ub chains in ribosome pausing (7, 10), we hypothesized that cells lacking Ubp2 would show impaired resumption of protein synthesis during stress recovery. Therefore, we expressed an episomal inducible GFP-based reporter in the WT and ubp2Δ strains to evaluate whether the absence of UBP2 would lead to a decrease in GFP protein synthesis during recovery. As a stable protein (38), accumulation of GFP is largely driven by synthesis. While both strains have decreased GFP accumulation due to translation shut down after H2O2 treatment, GFP accumulation is rapidly restored after stress in the WT, but remains largely unaltered in ubp2Δ (Fig. 6D). Our findings suggest that Ubp2 plays a role in translation resumption following stress; however, it was unclear whether the results observed for this GFP reporter could be extrapolated to the proteome level.
Figure 6.
Ubp2 deubiquitinates ribosomes and regulates protein synthesis.A, Ubp2 is able to remove K63-ub chains from ribosomes in vitro. Immunoblot anti-K63-ub of ribosomes (40 μg) isolated from ubp2Δ cells. Ribosomes were incubated in the presence or absence of 100 μM PR-619, 10 mM IAA (cysteine alkylator), 10 mM DTT (reducing agent), and purified recombinant Ubp2 (5 μg) for 1 h at 30 °C at 300 rpm. Anti-uL5/Rpl11 was used as a ribosome loading control (n = 3). B, Ubp2 is associated with ribosome fraction in the presence or absence of H2O2. Immunoblot anti-HA to detect Ubp2 levels from polysome profiling fractions from WT cells untreated or treated with H2O2 for 30 min (n = 2). C, Ubp2Δ cells present delayed K63-ub reversal from ribosomes during stress recovery. Immunoblot anti-K63-ub of ribosomes fractions isolated from WT and ubp2Δ upon treatment with 0.6 mM H2O2 for 30 min (top) and 60 min of stress recovery (bottom). Anti-uS3/Rsp3 and anti-uL5/Rpl11 were used as loading controls for 40S and 60S ribosome subunit, respectively (n = 1). D, fluorescence of GFP-reporter in Ubp2-HA and ubp2Δ cells was analyzed as a proxy for translation activity and translation recovery. GFP expression was induced in –Met medium for 100 min, followed by the addition of 0.6 mM H2O2. Increased levels of GFP abundance reflect active protein synthesis. GFP fluorescence was measured every 30 min at 525 nm with excitation at 485 nm, and normalized by cellular absorbance (A). Significance was calculated using an unpaired Student’s t test between WT H2O2/UT and ubp2Δ H2O2/UT (n = 3), ∗p < 0.05. H2O2, hydrogen peroxide; HA, hemagglutinin; IAA, iodoacetamide; K63-ub, K63-linked polyubiquitin; MW, molecular weight; Ub, ubiquitin.
To further investigate Ubp2’s role in restoring protein synthesis, we performed quantitative proteomics to understand how Ubp2 contributes to a global mechanism of translation regulation during different phases of the oxidative stress response. For this purpose, we cultivated WT and ubp2Δ cells and measured protein abundance in cells untreated or incubated with H2O2 for 30 and 120 min using label-free mass spectrometry. We chose 120 min after stress induction because at this time, cells generally start to reverse the K63-ub chains that accumulated under stress and resume translation that had been inhibited following H2O2 treatment (Figs. 5E and 6D). Our processed dataset is comprised of 4422 proteins representing ∼75% of the yeast proteome (39, 40). For a small number of proteins (179 proteins, 4.0% of the dataset) that presented three or fewer missing values across the 18 samples (i.e., 2 strains, 3 time points, 3 replicates), intensity values were imputed using the k-nearest neighbor method (41). To confirm that cells were responding to H2O2, we showed that both strains significantly enhanced the expression of key antioxidant enzymes such as thioredoxin Tsa2 and catalase Ctt1 after stress induction (Fig. S10A). When accounting for basal differences between strains, we observed that after 30 min of stress, 23 proteins were significantly downregulated (<1.5 fold) in ubp2Δ compared to the WT (Fig. S10B). Gene ontology (GO) analysis revealed that proteins involved in mitotic spindle and membrane proteins are significantly downregulated in the ubp2Δ background (Fig. S10D). When focusing on the recovery phase (30–120 min), and accounting for basal differences among strains, we observed that 8 and 27 proteins are significantly downregulated or upregulated, respectively, in the ubp2Δ strain (Fig. S10C). GO analysis confirmed that molecular functions related to exocytosis, vacuolar membrane, and mitosis remain up even 120 min after H2O2 addition to media (Fig. S10E).
Our results using GFP as a reporter protein indicated that Ubp2 is important for the re-establishment of protein synthesis during the recovery phase of stress (Fig. 6D). When comparing 120 min to our initial time point, we observed that 341 proteins are still differentially expressed in ubp2Δ in contrast to the 186 in the WT strain (Fig. S10, F and G). Of that, 90 proteins were shared among strains (Fig. S10G). Here, we show that only a fraction of the proteome is substantially different across strains and that deletion of UBP2 affects specific cellular pathways. However, other meaningful physiological changes might be present below our fold change cut-off. Our findings related to mitosis and mitotic spindle, motivated us to inspect the distribution of functional protein groups that are directly associated with cellular growth/division and stress response. We first observed that proteins from the GCN4 regulon, which are expressed as part of the integrated stress response (42), are significantly lower in ubp2Δ (Fig. 7B). This suggests that translation dysregulation might also affect cells capacity to mount a proper antioxidant defense. We also observed that ribosomal proteins are largely enriched among the upregulated proteins in WT cells (Fig. 7A). However, these proteins are mostly represented among the downregulated proteins and are significantly lower in ubp2Δ (Fig. 7, B and C). As the expression of these proteins is controlled by the target of rapamycin complex pathway (43), their levels are highly correlated with cellular growth and fitness (44), suggesting that different physiological states are present in these strains. Interestingly, we observed that only 16.7% and 9.1% of the proteins that are upregulated or downregulated, respectively, overlap when comparing WT and ubp2Δ strains, which suggests that these cells are mounting different cellular responses following stress induction (Fig. 7D). Taken together, our results support a model where Ubp2 is an integral part of the RTU, in which its redox regulation determines the dynamics of ribosome ubiquitination that is required to achieve proper stress recovery.
Figure 7.
Ubp2 supports translation reprogramming following stress.A and B, volcano plots displaying changes in protein levels for WT and ubp2Δ strains comparing 120 min after H2O2 treatment to untreated (UT) condition. Proteins are color-coded based on their subgroups: antioxidant proteins (pink), ribosomal proteins (orange), GCN4 regulon (blue), and others (gray). The horizontal dashed line indicates significance (p < 0.05), while the vertical dashed lines represent a fold change of ±1.5. C, box plots quantification for proteins belonging to the three functional groups as above. p values derived from unpaired Student’s t test. D, Venn diagrams showing the proteins upregulated (left) and downregulated (right) in WT and ubp2Δ cells between 120 min after H2O2 treatment and untreated (UT) conditions. E, schematic model for Ubp2 role in the RTU during steady state, stress condition, and stress recovery. Created with BioRender.com. H2O2, hydrogen peroxide; RTU, redox control of translation by ubiquitin.
Discussion
Here, we showed that redox regulation of the deubiquitinating enzyme Ubp2 is critical for the modulation of translation under stress. Our results demonstrated that Ubp2 is able to cleave K63-ub chains in vitro and in the cellular context (Fig. 1, A–C), and its activity can be reversibly regulated by peroxides (Figs. 1D and S3D). Although details on the molecular regulation of Ubp2 remain to be further elucidated, our work revealed the importance of several of Ubp2’s domains and amino acid residues to its redox regulation and activity (Figs. 4 and 5). This work opened several directions for research as a large fraction of eukaryotic DUBs are cysteine proteases (20, 23, 24) and have the potential to be redox-regulated under stress. Importantly, while we observe a strong K63-ub chain accumulation upon 0.6 mM H2O2 due to DUB inhibition, high concentrations of H2O2 can also impair the activity of the E2s and E3s (8, 10), altering the global Ub chain dynamics (Fig. 2A). Our findings using different peroxides (Fig. 2B) further demonstrate the nuances of the oxidative stress response, suggesting that different DUBs might be preferentially regulated under specific ROS conditions, triggering selective pathways. In this pathway, the redox balance of the cell and key thiol-specific enzymes (Figs. 3C and S5), likely function in a cascade that donates electrons for Ubp2 reduction and reactivation.
The indication that Ubp2’s repeated domains could have inhibitory or activation functions (Fig. 5, C–E) adds to this complexity, and further research must be conducted to understand the structural and functional conservation of these repeated domains across the evolutionary scale. While Ubp2 RD1 domain shows an inhibitory effect, RD2 seems required but not sufficient to drive Ubp2 function. In this context and based on AlphaFold2 structural predictions, RD3 might possess a structural role in positioning RD2 toward the catalytic residues or aiding in the recognition of selective chains (Fig. 4A). We also showed that the catalytic cysteine C745 forms a disulfide bond with C944 under stress (Fig. 4B), although only C745 mutations significantly impact Ubp2 activity (Fig. 4C). Thus, the role of the disulfide formation in the RTU remains to be elucidated. Additionally, although cells expressing Ubp2C944S/C821S are able to reverse K63-ub chains, it is unclear under different experimental conditions what fraction of Ubp2 remains active, forms disulfides, and is hyperoxidized, which would prevent its redox recycling. These findings support our conclusions that the activity of Ubp2 relies on these series of repeated domains, and future experiments will address the structural details by which Ubp2 recognizes and binds to ribosomes and K63-ub chains.
One of the first steps of the RTU is the accumulation of K63-ub chains on ribosomes during stress (5, 7). We have previously shown that the E2 conjugase Rad6 and the E3 ligase Bre1 are mainly responsible for the burst of the K63-ub chains upon H2O2 exposure (11). Moreover, we have shown that Rad6 itself can be redox-regulated and form an intermolecular disulfide bond with the E1 Uba1 (11). This regulation is part of a feedback loop that controls the total amount of Ub moieties added to ribosomes during the reprogramming of translation (11). In addition to the regulation of Rad6, we showed that Ubp2 plays a central role in the accumulation and reversal of these K63-ub chains (Fig. 1A). Our findings suggest a cycle of ubiquitination and deubiquitination in which Rad6 and Bre1 are constantly counteracted by Ubp2, which determines the amount of K63-ub chains in the system (Fig. 7E). In this scenario, Ubp2 would act as a surveillance system, instead of being recruited to ribosomes upon stress induction (Figs. 6B and S8C). Supporting this notion, we have observed that deletion of UBP2 leads to higher levels of K63-ub chains even in the absence of stress (5). In addition, this early accumulation of K63-ub chains further indicates that Ubp2 is the first protein in the RTU to be inhibited upon H2O2, which leads to a net gain of K63-ub chains. Next, Rad6 becomes inhibited, which defines the total amount of K63-ub added to ribosomes (5). Following stress cessation, Ubp2 is reactivated and is able to cleave these chains from its targets (Figs. 1F and 3A), returning translation to steady state (Figs. 6D and 7, D and E). Considering that several DUBs can be redox-regulated in a similar fashion (27), there is the potential that many pathways beyond the RTU can be controlled by a dynamic cycle of ubiquitination and deubiquitination. This cycle can then be temporarily interrupted depending on the nature of ROS, and the intensity and the duration of the stresses to which cells are exposed.
Although ubiquitination has been traditionally related to protein degradation (45), several new proteasome-independent pathways have been uncovered (46, 47). For example, different steps in gene expression, translation, and protein quality control are known to use Ub in a regulatory manner (48, 49). The regulation of transcription by ubiquitination has been heavily investigated (50, 51), but the rules of translation regulation are far less understood, particularly in response to dynamic environments. As several aspects of translation control can be regulated by Ub (49), DUBs are known to play crucial roles in antagonizing these pathways through the modulation of cycles of ribosomal ubiquitination and deubiquitination. In the RTU, ubiquitination and the activity of Rad6 are required to pause translation and reprogram protein production under stress (10). The ribosomes-associated quality control pathway is also controlled by the ubiquitination of ribosomal proteins and has been suggested to rely on the human DUBs OTUD3 and USP21 (52). Moreover, deubiquitination of the 40S protein eS7 by the yeast DUB Ubp3 regulates translation efficiency, and the DUB Otu2 promotes 40S dissociation from mRNA, participating in the recycling stage (53). In humans, the DUB USP10 has been found to regulate the degradation of stalled 40S ribosomes (54, 55), while USP36 is required for maturation of the 40S ribosomal subunit (56). However, how these additional DUBs are regulated, how the Ub dynamics in the ribosome are affected under stress, and the level of cross-talk among these different pathways remains largely unknown (57, 58). Some of the key questions that remain open in the RTU are how Ub pauses ribosomes under stress and how their deubiquitination allows translation to resume. Because of a preferential pause of ribosomes at the pretranslocation stage of translation elongation with sequence-specific features (7, 10), we have previously proposed that Ub affects the dynamics of binding and recruitment of translation factors to ribosomes (7). However, further research will be necessary to define the molecular and structural characteristics of this pathway. Regardless, the reversible regulation of Ubp2 allows us to propose a mechanism by which Ub temporarily pauses ribosomes, which will resume protein synthesis once they are deubiquitinated (Figs. 1A and 6, A and C).
Throughout evolution, eukaryotic genomes have expanded, and humans currently encode ∼100 DUBs with new functions, specialization, and pathway redundancy (59). Although Ubp2 is not highly conserved in humans, this surveillance mechanism mediated by Ubp2 in association with ribosomes will also support the discovery of new pathways that share similar molecular rules. Other groups have shown that several human DUBs can be activated or enhanced by reducing conditions (21, 27, 60), highlighting that we are just in the beginning of understanding how DUBs control cellular physiology globally in response to stress. As several diseases are caused by mutations and impairment of DUB activity (61), elucidating new mechanisms of DUB regulation can uncover new targets and support the development of new modes of therapy.
Experimental procedures
Yeast strains, plasmids, culture, and protein extraction
All S. cerevisiae strains and plasmids used in this study are described in Table S1. Proteomics processed data, analyses, and metadata are present in Table S2. Unless specified, yeast cells were cultivated into synthetic dextrose minimal medium (SD: 0.67% yeast nitrogen base, 2% dextrose, and required amino acids) at 30 °C at 200 rpm agitation. Cells were harvested at exponential phase A600 0.3 to 0.5. Protein extraction, ribosome isolation, polysome profile, GFP expression, and preparation for immunoblotting assays were performed as described previously (11).
Protein expression and purification
Escherichia coli BL21 and BL21-CodonPlus (DE3)-recombinant inbred line were transformed with pCS (Cold-Shock induced bacterial vector)/UBP2-HA-TEV-His and pGEX/CezanneCAT (catalytic domain)-TEV-glutathione-S-transferase, respectively. Bacterial cells were grown until A600 reached 0.6 to 0.8. Ubp2 expression was induced overnight at 16 °C in the presence of 1 mM IPTG and CezanneCAT was expressed with 0.6 mM IPTG for 4 h at 37 °C. Ubp2’s cell lysis was carried out by incubation with 1 mg/ml lysozyme for 1 h at 4 °C in buffer containing 50 ml Tris–HCl pH 7.5, 500 mM NaCl, 1 mM of DTT, and protease inhibitors (1 mM PMSF and 10 μM leupeptin) during 4× rounds of 2 min of sonication on ice followed by 1 min of rest. E. coli cells expressing CezanneCAT were lysed by sonication in 50 mM Tris–HCl pH 7.5, 150 mM NaCl, 0.5% octylphenoxypolyethoxyethanol, 1 mM DTT, 1 mM PMSF, 1 mM leupeptin, and 1 mg/ml lysozyme. The extract was cleared by centrifugation at 12,000 rpm for 30 min prior to 2 h incubation with nickel affinity beads for Ubp2 purification (GOLDBIO catalog H-355) or glutathione-containing beads (GOLDBIO G-250-5) for 2 h at 4 °C for CezanneCAT. The elution of Ubp2 was carried out with 250 mM imidazole, followed by size-exclusion chromatography column (SEC 200 – Cytiva HiLoad 26/600 Superdex 200 pg, cat. #28989336) in buffer containing 50 mM Tris–HCl pH 7.5, 150 mM and 1 mM DTT. CezanneCAT was eluted by cleavage of the glutathione-S-transferase tag with 15 units/ml tobacco etch virus (TEV) protease (SigmaT4455-1KU) in PBS buffer overnight at 4 °C and desalted using a PD-10 column (GE Healthcare, cat. #17085101) in 50 mM Tris–HCl pH 7.5, 100 mM NaCl buffer. Fractions were combined and concentrated using an Amicon centrifugal filter with 100 kDa or 30 kDa cut-off (Sigma), for Ubp2 and CezanneCAT, respectively.
In vitro ribosome deubiquitinating assay
The in vitro deubiquitinating assay was performed in the presence of 5 μg Ubp2, 10 mM DTT, and 40 μg of isolated ribosomes. All components were pre-incubated in reaction buffer (50 mM Tris pH 7.5, 100 mM NaCl, and 10 mM MgCl2) for 10 min at room temperature, before the addition of ribosomes. When specified, 7 mM H2O2 and/or 10 mM IAA was added to the reaction before the addition of ribosomes. The reaction was incubated for 1 h at 30 °C at 300 rpm, stopped by the addition of 4× Laemmli sample buffer, and subjected to SDS-PAGE gel prior to immunoblotting.
K63-ub chain cleavage assay
The cleavage of K63-ub chains by Ubp2 was assessed through an in vitro reaction containing 3 μg Ubp2, 10 mM DTT, 150 ng of tetra (LifeSensors) or 250 ng hexa K63-ub chains (R&D Systems), and reaction buffer (50 mM Tris pH 7.5 and 100 mM NaCl). The reaction was incubated for 1 h at 30 °C at 300 rpm, stopped by the addition of 4× Laemmli sample buffer, and subjected to SDS-PAGE followed by immunoblotting.
Immunoprecipitation assay
Yeast cells expressing Ubp2-HA and Ubp2C745S-HA were grown until A600 0.3 to 0.4 and challenged with 0.6 mM H2O2. Lysis was carried out in 50 mM Tris–HCl, pH 7.5, 100 mM NaCl, 5 mM MgCl2, 20 mM KCl as described before (11). Twenty five microliters of the normalized lysate was added to pierce protein A/G magnetic beads and incubated for 1 h under mild rotation. Unbound sample was removed and beads were washed twice with lysis buffer. Bound protein was eluted by incubation with 50 ng/μl of HA peptide for 30 min. Protein concentration was normalized via Bradford assay prior to the DUB activity assay.
DUB activity assay
The DUB activity was measured using the fluorogenic substrates Ub-AMC (R&D Systems, ex 345 nm, em 445 nm) or Ub-Rho 110 (LifeSensors, ex 485 nm, em 535 nm) in reaction buffer containing 50 mM Tris–HCl pH 7.5 and 150 mM NaCl at 30 °C. When specified, 5 min preincubation with peroxides was carried out before the start of the reaction. For measuring DUB activity in the total cell extract, cell lysis was performed in the absence of IAA. Enzymatic activity was determined and normalized by standard curves of free AMC and Rho 110.
Immunoblotting
Proteins were separated by standard 10 to 15% SDS-PAGE and transferred to polyvinylidene fluoride membrane (Thermo Fisher Scientific) for immunoblotting. The antibodies used in this work were as follows: anti-K63 Ub (1:8000, EMD Millipore, cat. #05-1308, clone apu3), anti-Ub (1:10000; Cell Signaling Technology, cat. #3936S), anti-uL5/Rpl11 (1:6000; Cell Signaling, cat. #18163), anti-uS3/Rsp3 (1:6000; Cell Signaling, cat. #9538), anti-PGK1 (1:8000; Invitrogen, cat. #22C5D8), anti-GAPDH (1:3000, Abcam, cat. #ab9485), anti-HA (1:3000, Thermo Fisher Scientific, cat. #71-5500), anti-actin (1:5000, Cell Signaling, cat. #4967), and anti-Rabbit IgG (1:400–10000; Cytiva, cat. #NA934). Immunoblots were developed by chemiluminescence using the Amersham ECL Prime (Cytiva, cat. #RPN2232).
ADCB sensitivity assays
Yeast cell cultures were grown into SD-Ura until log phase and diluted back to A600 0.1. Cells were mixed 1:1 with fresh medium or medium containing 200 μg/ml of ADCB (100 μg/ml final concentration). Cells were grown into 96-well plates in triplicate and incubated for up to 96 h at 30 °C under agitation. Absorbance was measured at 600 nm every 15 min in a Tecan Sunrise microplate reader.
3D structural analysis
3D graphical images of Ubp2 structure were generated in Chimera X (33, 34, 62) using the predicted models deposited in Alphafold2 protein database and UniProt (ID: Q01476).
Mass spectrometry analysis
Sample preparation
Three independent biological replicates of WT and ubp2Δ cells were grown in SD complete until A600 0.4. Samples were collected at time 0, 30, and 120 min following incubation with 0.6 mM H2O2. Cell lysis was carried out in buffer containing 50 mM ammonium bicarbonate pH 8, 50 mM NaCl, 1× protease inhibitor (EDTA-free), and 20 mM chloroacetamide. Fifteen micrograms of protein were adjusted to 5% SDS in 50 mM triethylammonium bicarbonate, pH 8.5 (TEAB). Samples were reduced with 10 mM DTT for 10 min at 80 °C using a Thermomixer at 1000 rpm, alkylated with 20 mM IAA for 30 min at room temperature, and then supplemented with a final concentration of 1.2% phosphoric acid and 200 μl of S-Trap (Protifi) binding buffer (90% MeOH/100 mM TEAB). Proteins were captured on the S-Trap micro device and washed 4× with 150 μl binding buffer, digested using 25 μl of 40 ng/μl sequencing-grade modified trypsin (Promega) for 1 h at 47 °C, and eluted using 40 μl 50 mM TEAB, followed by 40 μl of 0.2% formic acid, and 35 μl of 50% acetonitrile/0.2% formic acid. All samples were then lyophilized to dryness and were reconstituted in 30 μl of 1% TFA/2% MeCN. A study pool QC (SPQC) was created by combining equal volumes of each sample.
Quantitative data-independent acquisition LC-MS/MS analysis
Quantitative LC/MS/MS was performed on 1 μl of each sample and replicates of a study pool QC pool, using a Vanquish Neo LC coupled to a Thermo Orbitrap Astral via a Nanospray Flex ionization source. Briefly, the sample was first trapped on a Pepmap Neo Trap Cartridge and separated using a 1.5 μm PepSep 150 μm inner diameter × 8 cm column with a gradient of 5 to 12% MeCN from 0 to 3 min and 12 to 30% MeCN from 3 to 20 min, a flow rate of 500 nl/min with a column temperature of 45 °C. The liquid chromatography (LC) was interfaced to the mass spectrometer (MS) using a PepSep Sprayer and stainless steel (30 μm) emitter. The MS analysis used a 240,000-resolution precursor ion (MS1) scan from 380 to 9080 m/z, automatic gain control target of 500%, and maximum injection time of 50 ms, collected every 0.6 s in centroid mode. Tandem mass spectrometry (MS/MS) was performed using a data-independent acquisition (DIA) method with a default charge state of 3, precursor mass range of 380 to 480, 4 m/z isolation windows, automatic gain control target of 500%, maximum injection time of 6 ms, and a normalized collision energy of 28. A radio frequency lens of 40% was used for MS1 and DIA scans.
Quantitative analysis of DIA data
Raw MS data was demultiplexed and converted to .htrms format using HTRMS converter and processed in Spectronaut 18 (18.4.231017.55695]; Biognosys). A spectral library was built using direct-DIA searches against a S. cerevisiae database, downloaded from UniProt, and appended contaminant sequences using FragPipe. Search settings included N-terminus trypsin/P specificity up to two missed cleavages; peptide length from 7 to 52 amino acids with the following modifications: fixed carbamidomethyl (Cys), oxidation (Met) and acetylation (protein N terminus). For DIA analysis, default extraction, calibration, identification, and protein inference settings were used. Peptide and protein quantification were performed at the MS2 level with q-value sparse settings (precursors that met a q-value <0.01 in at least one run were included for quantification). For analysis in Spectronaut, local normalization was performed and protein abundances were calculated using the MaxLFQ algorithm (63, 64). The intensity values of 179 proteins with missing values (equal to or fewer than three) across the 18 samples were imputed using the impute known function. This function employs the k-nearest neighbor method (k = 5) and is part of the impute R package (41). In total, 4342 proteins were included in the downstream analysis.
Statistical methods and visualization
Statistical tests and visualizations were conducted using R v4.0.2. Pairwise comparisons were assessed using the paired Student’s t test, and unpaired comparisons were performed using the unpaired Student’s t test. p values <0.05 were considered significant. Volcano plots were generated using the ggplot2 package (65). In Figure 7, A–C, proteins were color-coded based on their functions, which included antioxidant proteins, ribosomal proteins, and those associated with the GCN4 regulon (66, 67). Heat maps were created with the R package ComplexHeatmap (68), and clustering was conducted using the k-means method. GO enrichment analysis was done using DAVID functional annotation clustering (69).
Data availability
The LC-MS/MS proteomics data (.RAW files) have been deposited to the Massive repository partnered with ProteomeXchange with the dataset identifier PXD051667. The methodology utilized to generate the proteomics dataset is described under Experimental procedures.
Supporting information
This article contains supporting information (70, 71).
Conflict of interest
The authors declare that they have no conflicts of interest with the contents of this article.
Acknowledgments
We thank Matt Foster and the Duke Proteomics and Metabolomics Shared Resource for support with mass spectrometry data acquisition. We are indebted to Nicholas Brown and Michael Emanuele for their kind donation of plasmids. We also thank Richard Brennan for kindly making SEC available. We also would like to thank the members of the Silva Lab for constructive feedback during the execution of this project and preparation of the manuscript.
Author contributions
C. M. S., B. K. C., D. A. O., C.-Y. C., N. M., N. A. S., V. S., E. J. W., and G. M. S. writing–review and editing; C. M. S. and G. M. S. writing–original draft; C. M. S. methodology; C. M. S., B. K. C., D. A. O., C.-Y. C., N. M., N. A. S., V. S., and G. M. S. investigation; C. M. S., B. K. C., C.-Y. C., N. A. S., and V. S. formal analysis; C. M. S., B. K. C., D. A. O., N. M., N. A. S., V. S., and E. J. W. data curation; D. A. O. and C.-Y. C. visualization; E. J. W. resources; G. M. S. supervision; G. M. S. project administration; G. M. S. funding acquisition; G. M. S. conceptualization.
Funding and additional information
This work was supported with funds from NIGMS R35 Award and the Chan Zuckerberg Initiative (GM137954 and SDL2022-253663, respectively, to G. M. S.). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.
Reviewed by members of the JBC Editorial Board. Edited by Donita C. Brady
Supporting information
References
- 1.Advani V.M., Ivanov P. Translational control under stress: reshaping the translatome. Bioessays. 2019;41 doi: 10.1002/bies.201900009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Vogel C., Silva G.M., Marcotte E.M. Protein expression regulation under oxidative stress. Mol. Cell. Proteomics. 2011;10 doi: 10.1074/mcp.M111.009217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Gasch A.P., Spellman P.T., Kao C.M., Carmel-Harel O., Eisen M.B., Storz G., et al. Genomic expression programs in the response of yeast cells to environmental changes. Mol. Biol. Cell. 2000;11:4241–4257. doi: 10.1091/mbc.11.12.4241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Zhang R., Yuan H., Wang S., Ouyang Q., Chen Y., Hao N., et al. High-throughput single-cell analysis for the proteomic dynamics study of the yeast osmotic stress response. Sci. Rep. 2017;7 doi: 10.1038/srep42200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Silva G.M., Finley D., Vogel C. K63 polyubiquitination is a new modulator of the oxidative stress response. Nat. Struct. Mol. Biol. 2015;22:116–123. doi: 10.1038/nsmb.2955. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Silva G.M., Vogel C. Mass spectrometry analysis of K63-ubiquitinated targets in response to oxidative stress. Data Brief. 2015;4:130–134. doi: 10.1016/j.dib.2015.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Zhou Y., Kastritis P.L., Dougherty S.E., Bouvette J., Hsu A.L., Burbaum L., et al. Structural impact of K63 ubiquitin on yeast translocating ribosomes under oxidative stress. Proc. Natl. Acad. Sci. U. S. A. 2020;117:22157–22166. doi: 10.1073/pnas.2005301117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Komander D. The emerging complexity of protein ubiquitination. Biochem. Soc. Trans. 2009;37:937–953. doi: 10.1042/BST0370937. [DOI] [PubMed] [Google Scholar]
- 9.Back S., Gorman A.W., Vogel C., Silva G.M. Site-specific K63 ubiquitinomics provides insights into translation regulation under stress. J. Proteome Res. 2018;18:309–318. doi: 10.1021/acs.jproteome.8b00623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Meydan S., Barros G.C., Simoes V., Harley L., Cizubu B.K., Guydosh N.R., et al. The ubiquitin conjugase Rad6 mediates ribosome pausing during oxidative stress. Cell Rep. 2023;42 doi: 10.1016/j.celrep.2023.113359. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Simoes V., Cizubu B.K., Harley L., Zhou Y., Pajak J., Snyder N.A., et al. Redox-sensitive E2 Rad6 controls cellular response to oxidative stress via K63-linked ubiquitination of ribosomes. Cell Rep. 2022;39 doi: 10.1016/j.celrep.2022.110860. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Kee Y., Lyon N., Huibregtse J.M. The Rsp5 ubiquitin ligase is coupled to and antagonized by the Ubp2 deubiquitinating enzyme. EMBO J. 2005;24:2414–2424. doi: 10.1038/sj.emboj.7600710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Kee Y., Muñoz W., Lyon N., Huibregtse J.M. The deubiquitinating enzyme Ubp2 modulates Rsp5-dependent Lys63-linked polyubiquitin conjugates in Saccharomyces cerevisiae. J. Biol. Chem. 2006;281:36724–36731. doi: 10.1074/jbc.M608756200. [DOI] [PubMed] [Google Scholar]
- 14.Koepp D.M., Lam M.H.Y., Emili A. Ubp2 regulates Rsp5 ubiquitination activity in vivo and in vitro. PLoS One. 2013;8 doi: 10.1371/journal.pone.0075372. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Jackson S.P., Durocher D. Regulation of DNA damage responses by ubiquitin and SUMO. Mol. Cell. 2013;49:795–807. doi: 10.1016/j.molcel.2013.01.017. [DOI] [PubMed] [Google Scholar]
- 16.Anton F., Dittmar G., Langer T., Escobar-Henriques M. Two deubiquitylases act on mitofusin and regulate mitochondrial fusion along independent pathways. Mol. Cell. 2013;49:487–498. doi: 10.1016/j.molcel.2012.12.003. [DOI] [PubMed] [Google Scholar]
- 17.Cavellini L., Meurisse J., Findinier J., Erpapazoglou Z., Belgareh-Touzé N., Weissman A.M., et al. An ubiquitin-dependent balance between mitofusin turnover and fatty acids desaturation regulates mitochondrial fusion. Nat. Commun. 2017;8 doi: 10.1038/ncomms15832. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Fang N.N., Zhu M., Rose A., Wu K.P., Mayor T. Deubiquitinase activity is required for the proteasomal degradation of misfolded cytosolic proteins upon heat-stress. Nat. Commun. 2016;7 doi: 10.1038/ncomms12907. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Ho H.C., MacGurn J.A., Emr S.D. Deubiquitinating enzymes Ubp2 and Ubp15 regulate endocytosis by limiting ubiquitination and degradation of ARTs. Mol. Biol. Cell. 2017;28:1271–1283. doi: 10.1091/mbc.E17-01-0008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Snyder N.A., Silva G.M. Deubiquitinating enzymes (DUBs): regulation, homeostasis, and oxidative stress response. J. Biol. Chem. 2021;297 doi: 10.1016/j.jbc.2021.101077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Mevissen T.E.T., Komander D. Mechanisms of deubiquitinase specificity and regulation. Annu. Rev. Biochem. 2017;86:159–192. doi: 10.1146/annurev-biochem-061516-044916. [DOI] [PubMed] [Google Scholar]
- 22.Groettrup M., editor. Conjugation and Deconjugation of Ubiquitin Family Modifiers. Vol. 54. Springer; New York, NY: 2010. (Subcellular Biochemistry). [Google Scholar]
- 23.Amerik A.Y., Hochstrasser M. Mechanism and function of deubiquitinating enzymes. Biochim. Biophys. Acta. 2004;1695:189–207. doi: 10.1016/j.bbamcr.2004.10.003. [DOI] [PubMed] [Google Scholar]
- 24.Ambroggio X.I., Rees D.C., Deshaies R.J. JAMM: a metalloprotease-like zinc site in the proteasome and signalosome. PLoS Biol. 2004;2:E2. doi: 10.1371/journal.pbio.0020002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Li Y., Reverter D. Molecular mechanisms of DUBs regulation in signaling and disease. Int. J. Mol. Sci. 2021;22:986. doi: 10.3390/ijms22030986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Ronau J.A., Beckmann J.F., Hochstrasser M. Substrate specificity of the ubiquitin and Ubl proteases. Cell Res. 2016;26:441–456. doi: 10.1038/cr.2016.38. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Lee J.G., Baek K., Soetandyo N., Ye Y. Reversible inactivation of deubiquitinases by reactive oxygen species in vitro and in cells. Nat. Commun. 2013;4:1568. doi: 10.1038/ncomms2532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Finley D., Ulrich H.D., Sommer T., Kaiser P. The ubiquitin-proteasome system of Saccharomyces cerevisiae. Genetics. 2012;192:319–360. doi: 10.1534/genetics.112.140467. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Bonacci T., Suzuki A., Grant G.D., Stanley N., Cook J.G., Brown N.G., et al. Cezanne/OTUD7B is a cell cycle-regulated deubiquitinase that antagonizes the degradation of APC/C substrates. EMBO J. 2018;37 doi: 10.15252/embj.201798701. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Netto L.E.S., de Oliveira M.A., Monteiro G., Demasi A.P.D., Cussiol J.R.R., Discola K.F., et al. Reactive cysteine in proteins: protein folding, antioxidant defense, redox signaling and more. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2007;146:180–193. doi: 10.1016/j.cbpc.2006.07.014. [DOI] [PubMed] [Google Scholar]
- 31.Helmann H.A.J.D. Thiol-based redox switches and gene regulation. Antioxid. Redox Signal. 2011;14:1049–1063. doi: 10.1089/ars.2010.3400. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Ma H., Wang M., Gai Y., Fu H., Zhang B., Ruan R., et al. Thioredoxin and glutaredoxin systems required for oxidative stress resistance, fungicide sensitivity, and virulence of Alternaria alternata. Appl.Environ. Microbiol. 2018;84 doi: 10.1128/AEM.00086-18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Jumper J., Evans R., Pritzel A., Green T., Figurnov M., Ronneberger O., et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596:583–589. doi: 10.1038/s41586-021-03819-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.UniProt C. UniProt: the universal protein knowledgebase in 2023. Nucleic Acids Res. 2023;51:D523–D531. doi: 10.1093/nar/gkac1052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Ren J., Kee Y., Huibregtse J.M., Piper R.C. Hse1, a component of the yeast Hrs-STAM ubiquitin-sorting complex, associates with ubiquitin peptidases and a ligase to control sorting efficiency into multivesicular bodies. Mol. Biol. Cell. 2007;18:324–335. doi: 10.1091/mbc.E06-06-0557. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Zhu L., Zhang Q., Cordeiro C.D., Banjade S., Sardana R., Mao Y., et al. Adaptor linked K63 di-ubiquitin activates Nedd4/Rsp5 E3 ligase. Elife. 2022;11 doi: 10.7554/eLife.77424. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Lam M.H., Emili A. Ubp2 regulates Rsp5 ubiquitination activity in vivo and in vitro. PLoS One. 2013;8 doi: 10.1371/journal.pone.0075372. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Perruca-Foncillas R., Davidsson J., Carlquist M., Gorwa-Grauslund M.F. Assessment of fluorescent protein candidates for multi-color flow cytometry analysis of Saccharomyces cerevisiae. Biotechnol. Rep. (Amst) 2022;34 doi: 10.1016/j.btre.2022.e00735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Hodges P.E., Payne W.E., Garrels J.I. The Yeast Protein Database (YPD): a curated proteome database for Saccharomyces cerevisiae. Nucleic Acids Res. 1998;26:68–72. doi: 10.1093/nar/26.1.68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Cherry J.M., Hong E.L., Amundsen C., Balakrishnan R., Binkley G., Chan E.T., et al. Saccharomyces Genome Database: the genomics resource of budding yeast. Nucleic Acids Res. 2012;40:D700–D705. doi: 10.1093/nar/gkr1029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Hastie T., Tibshirani R., Narasimhan B., Chu G. Impute: Impute: Imputation for Microarray Data R Package Version 1.77.0 Ed. Bioconductor; Seattle, WA: 2023. [Google Scholar]
- 42.Pakos-Zebrucka K., Koryga I., Mnich K., Ljujic M., Samali A., Gorman A.M. The integrated stress response. EMBO Rep. 2016;17:1374–1395. doi: 10.15252/embr.201642195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Lippman S.I., Broach J.R. Protein kinase A and TORC1 activate genes for ribosomal biogenesis by inactivating repressors encoded by Dot6 and its homolog Tod6. Proc. Natl. Acad. Sci. U. S. A. 2009;106:19928–19933. doi: 10.1073/pnas.0907027106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Guerra P., Vuillemenot L.P.E., van Oppen Y.B., Been M., Milias-Argeitis A. TORC1 and PKA activity towards ribosome biogenesis oscillates in synchrony with the budding yeast cell cycle. J. Cell Sci. 2022;135 doi: 10.1242/jcs.260378. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Mallette F.A., Richard S. K48-linked ubiquitination and protein degradation regulate 53BP1 recruitment at DNA damage sites. Cell Res. 2012;22:1221–1223. doi: 10.1038/cr.2012.58. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Jackson M.P., Hewitt E.W. Cellular proteostasis: degradation of misfolded proteins by lysosomes. Essays Biochem. 2016;60:173–180. doi: 10.1042/EBC20160005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Chen R.H., Chen Y.H., Huang T.Y. Ubiquitin-mediated regulation of autophagy. J. Biomed. Sci. 2019;26:80. doi: 10.1186/s12929-019-0569-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Pickart C.M., Eddins M.J. Ubiquitin: structures, functions, mechanisms. Biochim. Biophys. Acta. 2004;1695:55–72. doi: 10.1016/j.bbamcr.2004.09.019. [DOI] [PubMed] [Google Scholar]
- 49.Dougherty S.E., Maduka A.O., Inada T., Silva G.M. Expanding role of ubiquitin in translational control. Int. J. Mol. Sci. 2020;21:1151. doi: 10.3390/ijms21031151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Conaway R.C., Brower C.S., Conaway J.W. Emerging roles of ubiquitin in transcription regulation. Science. 2002;296:1254–1258. doi: 10.1126/science.1067466. [DOI] [PubMed] [Google Scholar]
- 51.Mark K.G., Rape M. Ubiquitin-dependent regulation of transcription in development and disease. EMBO Rep. 2021;22 doi: 10.15252/embr.202051078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Joazeiro C.A.P. Mechanisms and functions of ribosome-associated protein quality control. Nat. Rev. Mol. Cell Biol. 2019;20:368–383. doi: 10.1038/s41580-019-0118-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Takehara Y., Yashiroda H., Matsuo Y., Zhao X., Kamigaki A., Matsuzaki T., et al. The ubiquitination-deubiquitination cycle on the ribosomal protein eS7A is crucial for efficient translation. Iscience. 2021;24 doi: 10.1016/j.isci.2021.102145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Meyer C., Garzia A., Morozov P., Molina H., Tuschl T. The G3BP1-family-USP10 deubiquitinase complex rescues ubiquitinated 40S subunits of ribosomes stalled in translation from lysosomal degradation. Mol. Cell. 2020;77:1193–1205.e5. doi: 10.1016/j.molcel.2019.12.024. [DOI] [PubMed] [Google Scholar]
- 55.Garshott D.M., An H., Sundaramoorthy E., Leonard M., Vicary A., Harper J.W., et al. iRQC, a surveillance pathway for 40S ribosomal quality control during mRNA translation initiation. Cell Rep. 2021;36 doi: 10.1016/j.celrep.2021.109642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.van den Heuvel J., Ashiono C., Gillet L.C., Dörner K., Wyler E., Zemp I., et al. Processing of the ribosomal ubiquitin-like fusion protein FUBI-eS30/FAU is required for 40S maturation and depends on USP36. Elife. 2021;10 doi: 10.7554/eLife.70560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Garshott D.M., Sundaramoorthy E., Leonard M., Bennett E.J. Distinct regulatory ribosomal ubiquitylation events are reversible and hierarchically organized. Elife. 2020;9 doi: 10.7554/eLife.54023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Shannon E., Dougherty G.C.B., Foster M.W., Teo G., Choi H., Silva G.M. Context specific ubiquitin modification of ribosomes regulates translation under oxidative stress. bioRxiv. 2024 doi: 10.1101/2024.05.02.592277. [preprint] [DOI] [Google Scholar]
- 59.Harrigan J.A., Jacq X., Martin N.M., Jackson S.P. Deubiquitylating enzymes and drug discovery: emerging opportunities. Nat. Rev. Drug Discov. 2018;17:57–78. doi: 10.1038/nrd.2017.152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Kulathu Y., Garcia F.J., Mevissen T.E., Busch M., Arnaudo N., Carroll K.S., et al. Regulation of A20 and other OTU deubiquitinases by reversible oxidation. Nat. Commun. 2013;4:1569. doi: 10.1038/ncomms2567. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Bello A.I., Goswami R., Brown S.L., Costanzo K., Shores T., Allan S., et al. Deubiquitinases in neurodegeneration. Cells. 2022;11:556. doi: 10.3390/cells11030556. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Meng E.C., Goddard T.D., Pettersen E.F., Couch G.S., Pearson Z.J., Morris J.H., et al. UCSF ChimeraX: tools for structure building and analysis. Protein Sci. 2023;32 doi: 10.1002/pro.4792. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Callister S.J., Barry R.C., Adkins J.N., Johnson E.T., Qian W.J., Webb-Robertson B.J., et al. Normalization approaches for removing systematic biases associated with mass spectrometry and label-free proteomics. J. Proteome Res. 2006;5:277–286. doi: 10.1021/pr050300l. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Cox J., Hein M.Y., Luber C.A., Paron I., Nagaraj N., Mann M. Accurate proteome-wide label-free quantification by delayed normalization and maximal peptide ratio extraction, termed MaxLFQ. Mol. Cell. Proteomics. 2014;13:2513–2526. doi: 10.1074/mcp.M113.031591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Ginestet C. ggplot2: elegant graphics for data analysis. J. R. Stat. Soc. Ser. A Statistics Soc. 2011;174:245–246. [Google Scholar]
- 66.Fromont-Racine M., Senger B., Saveanu C., Fasiolo F. Ribosome assembly in eukaryotes. Gene. 2003;313:17–42. doi: 10.1016/s0378-1119(03)00629-2. [DOI] [PubMed] [Google Scholar]
- 67.Yan L.L., Zaher H.S. Ribosome quality control antagonizes the activation of the integrated stress response on colliding ribosomes. Mol. Cell. 2021;81:614–628.e4. doi: 10.1016/j.molcel.2020.11.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Gu Z., Eils R., Schlesner M. Complex heatmaps reveal patterns and correlations in multidimensional genomic data. Bioinformatics. 2016;32:2847–2849. doi: 10.1093/bioinformatics/btw313. [DOI] [PubMed] [Google Scholar]
- 69.Sherman B.T., Hao M., Qiu J., Jiao X., Baseler M.W., Lane H.C., et al. DAVID: a web server for functional enrichment analysis and functional annotation of gene lists (2021 update) Nucleic Acids Res. 2022;50:W216–W221. doi: 10.1093/nar/gkac194. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Finley D., Sadis S., Monia B.P., Boucher P., Ecker D.J., Crooke S.T., et al. Inhibition of proteolysis and cell cycle progression in a multiubiquitination-deficient yeast muInhibition of proteolysis and cell cycle progression in a multiubiquitination-deficient yeast mutant. Mol. Cell. Biol. 1994;14:5501–5509. doi: 10.1128/mcb.14.8.5501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Breslow D.K., Cameron D.M., Collins S.R., Schuldiner M., Stewart-Ornstein J., Newman H.W., et al. A comprehensive strategy enabling high-resolution functional analysis of the yeast genome. Nat. Methods. 2008;5:711–718. doi: 10.1038/nmeth.1234. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The LC-MS/MS proteomics data (.RAW files) have been deposited to the Massive repository partnered with ProteomeXchange with the dataset identifier PXD051667. The methodology utilized to generate the proteomics dataset is described under Experimental procedures.