Abstract
Free full text
Hydrogen sulfide coordinates glucose metabolism switch through destabilizing tetrameric pyruvate kinase M2
Abstract
Most cancer cells reprogram their glucose metabolic pathway from oxidative phosphorylation to aerobic glycolysis for energy production. By reducing enzyme activity of pyruvate kinase M2 (PKM2), cancer cells attain a greater fraction of glycolytic metabolites for macromolecule synthesis needed for rapid proliferation. Here we demonstrate that hydrogen sulfide (H2S) destabilizes the PKM2 tetramer into monomer/dimer through sulfhydration at cysteines, notably at C326, leading to reduced PKM2 enzyme activity and increased PKM2-mediated transcriptional activation. Blocking PKM2 sulfhydration at C326 through amino acid mutation stabilizes the PKM2 tetramer and crystal structure further revealing the tetramer organization of PKM2-C326S. The PKM2-C326S mutant in cancer cells rewires glucose metabolism to mitochondrial respiration, significantly inhibiting tumor growth. In this work, we demonstrate that PKM2 sulfhydration by H2S inactivates PKM2 activity to promote tumorigenesis and inhibiting this process could be a potential therapeutic approach for targeting cancer metabolism.
Introduction
Cancer metabolism refers to the reprogramming of cellular metabolic pathways in cancer cells to acquire the necessary nutrients to build new biomass. These metabolic rewiring activities in cancer cells are recognized as a general hallmark of cancer1. Among the altered metabolic pathways, the Warburg effect2, a preferential dependence on aerobic glycolysis, is the classical example of cancer-specific metabolism. Although glycolysis is less efficient than oxidative phosphorylation for adenosine triphosphate (ATP) production, glycolysis provides metabolic intermediates required for the biosynthesis of nucleotides and amino acids, which cancer cells depend on for rapid cell proliferation under insufficient nutrient supply3. The observation of this differential metabolic preference in cancer cells unveils a therapeutic potential of targeting this specific metabolic adaptation utilized by cancer cells. Indeed, drugs intended to block aerobic glycolysis in cancer cells are currently under evaluation, including inhibitors against lactate dehydrogenase4 and monocarboxylate transporters5. Further studies of drugs targeting these tumor-specific metabolic pathways are highly warranted.
There are three key rate-limiting enzymes that regulate the flow of glycolysis: hexokinases (HKs), phosphofructokinase-1 (PFK-1), and pyruvate kinases (PKs)6. Among these, PK regulates the final rate-limiting step of glycolysis by catalyzing the phosphoryl transfer from phosphoenolpyruvate (PEP) to adenosine diphosphate (ADP) to produce pyruvate and ATP7. In mammals, two genes encode four mammalian PK isoforms (PKM1, PKM2, PKR, PKL). PKM2 is highly expressed in proliferating cells, including all types of cancer cells and tumors tested thus far8. Despite their sequence similarity, PKM1 and PKM2 have different catalytic properties. PKM1 is constitutively organized as a tetramer with high catalytic activity. In contrast, PKM2 maintains low activity in cancer cells through binding to various ligands or posttranslational modifications (PTMs)9, including phosphorylation10, acetylation11, oxidation12, SUMOylation13, O-GlcNAcylation14, and methylation15. These modifications of PKM2 downregulate its enzymatic activity, thus redirecting the use of glucose from energy production to rapid biomass synthesis, which is needed for fast-proliferating tumor cells. In addition, monomeric or dimeric PKM2 also has non-canonical functions at different cellular locations, such as in the nucleus, mitochondria, and exosomes, to support tumor growth and distant metastasis16.
Hydrogen sulfide (H2S), a colorless, flammable, water-soluble gas, is an endogenously produced gasotransmitter that acts as a critical mediator in multiple physiological processes17, including vasodilation18–20, aging21, and inflammation22. Aberrant upregulation of the H2S-synthesizing enzymes, cystathionine β-synthase (CBS), cystathionine γ-lyase (CTH), and 3-mercaptopyruvate sulfurtransferase (3-MST), is frequently observed in multiple cancer types23. Endogenous H2S promotes tumor growth through antiapoptotic effects, DNA repair, tumor growth, cancer metabolism, metastasis, and angiogenesis23, mainly through the formation of a persulfide (-SSH) bond on cysteine, which is called protein sulfhydration24. Conversely, other studies have indicated that higher concentrations of H2S exert anticancer effects25,26, suggesting a bell-shaped relationship with H2S. Lower concentrations of H2S often promote cancer cell proliferation, while higher concentrations inhibit cell proliferation. Currently, the influences of H2S-mediated protein sulfhydration in cancer remain largely unexplored, warranting further investigations to elucidate how H2S impacts cancer progression.
PKM2 activity can be inhibited by different amino acids9, particularly L-cysteine27. As H2S is mainly produced from L-cysteine28, we investigate whether PKM2 activity could be modulated by H2S. Here, we show that breast cancer cells utilize H2S to destabilize PKM2 tetramers into monomers or dimers with low PK activity. We further determine the sulfhydration sites on PKM2 by mass spectrometry. A single amino acid mutation at cysteine 326 stabilizes the PKM2 tetramer with high PK activity, resulting in the rewiring of glucose metabolism towards mitochondrial respiration, leading to increased cytokinesis failure, as well as reduced cell proliferation and tumor growth in a breast cancer xenograft mouse model. Our data support the hypothesis that sulfhydration is a key posttranslational modification that modulates PKM2 activity. Cysteine 326 on PKM2 may serve as a therapeutic target for future anticancer drug development.
Results
The sulfhydration of PKM2 through H2S promotes the dissociation of the active PKM2 tetramer to the less active PKM2
To investigate the role of H2S in PKM2 activity through protein sulfhydration, we first confirmed that treatment with NaHS, an H2S donor, induced the sulfhydration of PKM2 by using a modified biotin switch assay to pull down sulfhydrated proteins in breast cancer MDA-MB-231 cell lysates (Fig. 1a) and prostate cancer PC3 cell lysates (Fig. S1a). This modification was reversed after treatment with dithiothreitol (DTT), which can cleave disulfide bonds (Fig. 1a, S1a). Moreover, the activities of PKM2 were reduced in the presence of NaHS in both recombinant proteins (Fig. 1b) and cancer cell lysates (Fig. S1b), and this inhibition was reversed by DTT treatment (Fig. 1b), suggesting that disrupting sulfhydration by DTT can restore the PK activity of PKM2. Next, to determine whether H2S-mediated sulfhydration of PKM2 is involved in the oligomerization of PKM2, we performed glycerol gradient ultracentrifugation on purified recombinant PKM2 to differentiate the tetrameric and monomeric/dimeric forms of PKM2 (Fig. 1c). The majority of the PKM2 in solution was a mixture of both monomer/dimer and tetramer forms, while the exposure of PKM2 to its endogenous activator, fructose 1,6-bisphosphate (FBP), resulted in a shift of monomers/dimers to tetramers (Fig. 1c). Interestingly, cotreatment of NaHS resulted in significant dissociation of the FBP-induced tetramer (Fig. 1c), suggesting that allosteric regulation of PKM2 may be modulated by H2S-mediated sulfhydration. We further confirmed that FBP remained the same after NaHS treatment (Fig. S1c) to exclude the possibility that FBP may be dissociated into fructose by NaHS. To ascertain the oligomeric state of sulfhydrated PKM2, we utilized mass photometry, a light scattering-based method that detects protein oligomerization in solution. The results indicate that the majority of NaHS-treated recombinant PKM2 exists as a monomer (Fig. S1d). In the cellular contest, PKM2 can translocate into the nucleus in either monomeric or dimeric state and act as a transcriptional coactivator to stimulate the transcription of genes that are required for tumor growth29,30. We next evaluated PKM2 nuclear translocation and the expression of PKM2 response genes, including cyclin D1 (CCND1), hypoxia-inducible factor 1-α (HIF-1α), lactate dehydrogenase A (LDHA), glucose transporter 1 (GLUT1), GLUT12, hexokinase 2 (HK2), and glutaminase-1 (GLS1). We found that the nuclear translocation of PKM2 was significantly enhanced by NaHS treatment (Fig. 1d), and PKM2-responsive genes were induced in the presence of NaHS (Fig. 1e). Moreover, depletion of H2S using the H2S inhibitor aminooxyacetic acid (AOAA) (Fig. S1e) or knockdown the H2S-producing enzymes CBS and CTH by siRNA (Fig. S1f), both resulted in decreased PKM2 sulfhydration (Fig. S1g). Consequently, H2S depletion by AOAA or siRNA targeting CBS and CTH hindered PKM2 nuclear translocation and led to the inhibition of several PKM2-responsive genes (Fig. S1h-k). Notably, knockdown of CBS and CTH suppressed the proliferation of MDA-MB-231 cells (Fig. 1f). On the other hand, AOAA treatment completely inhibited the proliferation of MDA-MB-231 cells and co-treatment with NaHS only partially mitigated the AOAA-induced reduction in cell proliferation (Figure S1l). This observation suggests that the complete growth inhibition caused by AOAA may not be solely attributed to the depletion of H2S production. Instead, it may involve additional effects mediated by vitamin B6, as AOAA is a broad-spectrum inhibitor of enzymes that require vitamin B6 as a coenzyme31. Overall, our data strongly indicate that H2S-mediated sulfhydration modulates the dissociation of PKM2 tetramers into monomers/dimers, resulting in decreased PK activity and heightened transcriptional activity of PKM2 responsive genes.
Studies have shown that PKM2 and two H2S-synthesizing enzymes, CBS and CTH, are upregulated in breast tumors23,32. To further understand the clinical correlation between PKM2 and H2S-synthesizing enzymes in breast cancer, we analyzed the expression profiles of PKM2, CBS, CTH, and 3-MST from the TCGA database. Upregulation of both PKM and CBS mRNA was observed in breast tumors (Fig. S2a), and CBS expression was associated with poor prognosis in breast cancer patients (Fig. S2b). Interestingly, correlation analysis revealed that the expression of CBS was positively correlated with that of PKM, as well as the expression levels of most PKM2 response genes, including HIF1A, LDHA, GLUT1, GLUT12, and GLS1 (Fig. S2c), suggesting that elevated H2S levels in breast tumors potentially enhance PKM2 tetramer dissociation and facilitate its translocation to the nucleus for transcriptional regulation. We then compared the expression and sulfhydration levels between mammary epithelial MCF-10A cells and breast cancer MDA-MB-231, MCF-7, and HCC-1395 cells. With higher CBS or CTH expression (Fig. 1g), significantly greater sulfhydration was observed in all breast cancer cell lines we examined than in the MCF-10A cells (Fig. 1h). To ascertain whether the stimulation of PKM2 nuclear translocation by H2S and the subsequent regulation of gene expression is a general phenomenon applicable to other cells, we utilized normal breast epithelial MCF-10A cells and prostate cancer PC3 cells for further validation. Our data demonstrated that H2S stimulates PKM2 nuclear translocation (Fig. S2d, e) and induces the expression of PKM2-responsive genes in both cell lines (Fig. S2f, g). A previous study showed that excessive reactive oxygen species (ROS) induced by hypoxia can suppress the PK activity of PKM2 during cancer progression33. As cysteine sulfenylation (S-OH) by ROS is an important prior step for H2S to modify cysteine via protein sulfhydration34, we assessed whether hypoxia could induce PKM2 sulfhydration in MDA-MB-231 and PC3 cells. The level of PKM2 sulfhydration increased under hypoxia, and this increase was reduced by AOAA (Fig. 1i, S2h). Overall, our data suggest that to meet the increased biosynthetic demands in the hypoxic tumor microenvironment, cancer cells utilize H2S in part to stimulate PKM2 sulfhydration. We conducted experiments mostly on cell lines from two different cancer types, breast cancer and prostate cancer, to demonstrate that the destabilization of PKM2 tetramers by H2S is a general phenomenon occurring across various cell types in the body.
Identification of the sulfhydration sites on PKM2 by mass spectrometry and mutation of Cys326 enhances the tetramerization of PKM2
There are ten cysteine residues located on each monomeric PKM2 protein (Fig. 2a). To further characterize the sulfhydration site on PKM2, we carried out liquid chromatography with tandem mass spectrometry (LC-MS/MS) on trypsin-digested recombinant PKM2 protein previously treated with NaHS (Fig. 2b) or endogenous PKM2 isolated from MDA-MB-231 cells (Fig. 2c). The recombinant PKM2 protein was sulfhydrated at cysteines 49 and 326 in the presence of NaHS (Fig. 2b). For endogenous PKM2 sulfhydration, to prevent excessive ROS-induced cysteine oxidation which may lead to artificial sulfhydration, we first added MMTS, which alkylates free thiol groups, to the culture medium before cell lysis. The clarified cell lysate was then subjected to immunoprecipitation to enrich PKM2 for sulfhydration analysis by LC-MS/MS. Only cysteine 326 on PKM2 was found to be sulfhydrated by this approach (Fig. 2c). We also checked whether other cysteine PTMs, including S-nitrosylation, glutathionylation, or oxidation, could be detected. None of these PTMs were detected in our mass spectrometry results (Table S1), although there may be other cysteine PTMs that cannot be detected by this proteomic method, as most cysteine modifications are transient and reversible35. Our proteomic analysis revealed that cysteine 326 endogenously sulfhydrated in MDA-MB-231 cells (Fig. 2c). To further study the effects of sulfhydration on cysteine 326, we substituted this residue with serine (PKM2C326S). Sulfhydration was significantly reduced, but not eliminated, in the HA-tagged PKM2C326S overexpressing MDA-MB-231 cells (Fig. S3a), suggesting that cysteine 326 acts as one of the major sulfhydration sites modulated by H2S. Consistent with the results we observed in cell lysates (Fig. S3a), there was less sulfhydrated recombinant PKM2 protein with the C326S mutation than with wild-type PKM2 (Fig. S3b). More importantly, by performing gel filtration analysis, we observed a striking transition from the monomeric/dimeric form to the tetrameric form in the recombinant PKM2C326S proteins, while the wild-type PKM2 proteins predominantly retained their monomeric/dimeric form (Fig. 2d), suggesting that PKM2 C326 may serve as a crucial allosteric site influencing PKM2 oligomerization. An elevated tetrameric configuration of PKM2 was also observed in cell lysates derived from MDA-MB-231 cells expressing the PKM2C326S (Fig. 2e), despite the complexity of cell lysates, which could encompass some PKM2 interacting proteins that may impact the distribution of PKM2.
Tetrameric PKM2C326S adopts a unique conformation
PKM2 is allosterically regulated by bound ligands. It is believed that tetrameric PKM2 adopts an inactive T-state conformation in the apo state or in the inhibitor-bound state and shifts to an active R-state conformation upon binding to FBP and allosteric activators36,37. Each protomer consists of three functional domains (A, B, and C), with the active site between the A and B domains, and the FBP binding pocket in the C-terminal C domain. Two critical residues, W482 and W515, stabilize the T-state structure at the interface between two neighboring C domains (referred to as the C-C interface). FBP binding leads to global rotation of each protomer, disrupting the intermolecular interaction of W482 and W515, while new intermolecular contacts between K422 on one protomer and Y444 and P403 on the other are merged, giving rise to the fully active R-state.
To delineate which tetrameric state the C326S mutant adopts, we determined the crystal structure of PKM2C326S at 3.1Å resolution in the apo form (Fig. 3a, Table S2). In one asymmetric unit, there are four protomers (A-D) belonging to two tetramers (AD/AD, BC/BC). These protomers have a folding similar to that of the wild type (Fig. S4a–d). To provide the structural basis for C326S mutation-mediated tetramerization, we compared C326S to the T and R state conformers36,38. Structural superimposition revealed that PKM2C326S is arranged in a different tetramer conformation (Fig. 3b, c). Located in the A domain, C326 makes contact with Q310 through hydrogen bonding in the R state, while this interaction is absent in the T state. In the C326S mutant, S326 also establishes a hydrogen bond with Q310 (Fig. 3d). This mutation does not change the overall conformation of the A domain. Instead, the C domains in PKM2C326S display structural differences from a r.m.s.d. of 2.27Å, while the C domains in the T or R state structure are similar to a r.m.s.d. of 0.96Å and 0.64Å, respectively (Fig. S4a–d). The major differences are in the α14-α15 region and the effector loop, whose conformation is one of the key determinants of the T and R states (Fig. 3e–h, S4e–h). The AD/AD tetramer has better structural quality and was used for further structural analyses. Compared to the T-state conformation, the side chain of W515 from protomer D swung away from the C-C interface, breaking the interaction between W515 (protomer D) and W482 (protomer A) (Fig. 3e, f). Additional intermolecular contacts (i.e., cation-π interactions and hydrogen bonding) are established between Y444 and P403 on one protomer and K422 on the other in the α14-α15 region (Fig. 3g, h), which were observed in the R-state conformation. The abovementioned structural differences may result in the elevated tetramerization of the C326S mutant.
We also conducted a molecular dynamics simulation on sulfhydrated PKM2 tetramer. The most stable conformer observed in the simulation was then compared with the T-state, R-state, and PKM2C326S conformers. We found that the side chain conformation of Q310 was different, and no hydrogen bond was observed with sulfhydrated C326 (Fig. S5a). Additionally, the structure of sulfhydrated tetramer had a looser conformation with a smaller interface area between the C-C interface (Fig. S5b, Table S3), and key interactions that stabilize the tetramer were disrupted. The simulation results provide us with structural insights into sulfhydration at position C326.
Disruption of PKM2 sulfhydration at Cys326 promotes mitochondrial OXPHOS by increasing PK activity
To investigate the function of H2S-mediated PKM2 sulfhydration, we established MDA-MB-231 breast cancer cells and PC3 prostate cancer cells stably expressing PKM2wt or PKM2C326S mutant (Fig. 4a, S6a). Consistent with the results we observed in cells with transient transfection (Fig. S3a), there was less sulfhydrated PKM2 protein with the C326S mutation than wild-type PKM2 in MDA-MB-231 cells stably expressing PKM2wt or the PKM2C326S mutant (Fig. 4a). Pyruvate kinase activity was significantly increased in MDA-MB-231 cells expressing PKM2C326S (Fig. 4b). We next examined whether blocking PKM2 sulfhydration at C326 can rewire glucose metabolic fluxes. Remarkably, the expression of PKM2C326S significantly increased the oxygen consumption rates (OCRs) in both MDA-MB-231 cells (Fig. 4c, d) and PC3 cells (Fig. S6b, c), confirming our hypothesis that the disruption of PKM2 sulfhydration redirected cancer cell glucose metabolism to the mitochondrial oxidative phosphorylation (OXPHOS) system. The extracellular acidification rates (ECARs) were only slightly increased in MDA-MB-231 cells and PC3 cells expressing PKM2C326S (Fig. 4e, f, S6d, e) suggesting that the route to glycolysis was not greatly affected by the PKM2C326S mutation. Restoring PK activity strongly facilitated OXPHOS but did not suppress the ECAR in either cell line. To further characterize how PKM2 sulfhydration modulates glucose metabolism, we first evaluated the expression levels of PKM2-response genes, including CCND1, HIF-1α, LDHA, GLUT1, GLUT12, HK2, and GLS1. The expression levels of most genes, except CCND1 and LDHA, were significantly reduced in MDA-MB-231 cells expressing PKM2C326S (Fig. 4g). Metabolite analysis of glycolytic intermediates was then performed using UPLC/Q-TOF-MS/MS. We found that the levels of glucose, glucose-6-phosphate, fructose-1, 6-bisphosphate, ribose-5-phosphate, glyceraldehyde-3-phosphate, and phosphoenolpyruvate were significantly reduced in cells expressing the PKM2C326S mutant (Fig. 4h), revealing that blockade of PKM2 sulfhydration at C326 leads to increased PK activity to facilitate OXPHOS and results in a shortage of glycolytic intermediates and ribose-5-phosphate, which is essential for DNA synthesis. Moreover, lactate level was similar in cells expressing PKM2C326S compared to that in wild-type PKM2-expressing cells, while pyruvate was significantly upregulated in the C326S group (Fig. 4h). The changes in these metabolites were consistent with the results of the Seahorse analysis, in which the OCR was significantly increased in the PKM2C326S mutant while the ECAR was only slightly increased (Fig. 4c–f). We next checked whether PKM2 sulfhydration at C326 is required for its nuclear translocation. As expected, the percentage of PKM2 in the nucleus was much greater in the MDA-MB-231 cells expressing wild-type PKM2 than in the cells expressing PKM2C326S (Fig. 4i). Overall, our data suggest that blocking PKM2 sulfhydration at C326 will facilitate mitochondrial OXPHOS by increasing PK activity and suppressing its nuclear transcriptional activities (as illustrated in Fig. S6f).
PKM2 sulfhydration at Cys326 is required to facilitate cytokinesis during cell division
During cancer cell division, monomeric/dimeric PKM2, which has low PK activity, facilitates biosynthesis from metabolic intermediates and functions in the nucleus to induce genes involved in the cell cycle39, resulting in the promotion of G1/S cell cycle transition40 and chromosome segregation during mitosis41. To determine the effect of PKM2 sulfhydration on cell division, we first analyzed the numbers of MDA-MB-231 cells expressing wild-type PKM2 or PKM2C326S at different phases of the cell cycle. The percentage of polyploid cells (>4N) was increased in cells expressing PKM2C326S (Fig. 5a, b), suggesting that blocking PKM2 sulfhydration at C326 may inhibit cytokinesis during mitosis. Interestingly, approximately 10% of the population of MDA-MB-231 cells with PKM2C326S were giant in size and possessed multiple nuclei (Fig. 5c, d), which are typical characteristics of cell senescence42. β-gal staining further confirmed that these giant cells were undergoing senescence (Fig. S7a). As dimeric PKM2 is known to play a pivotal role during metaphase and cytokinesis41,43, we performed time-lapse microscopy at 10-min intervals continuously for 24h to determine whether disruption of PKM2 sulfhydration at C326 causes mitotic defects (Fig. 5e). Through observation following mitotic progression in individual cells, we found that 22% of MDA-MB-231 cells expressing wild-type PKM2 showed evidence of cytokinesis failure (Fig. 5f), as judged by the failure of the separation of two daughter cells (Fig. 5e). Notably, cells expressing PKM2C326S exhibited an approximately 1.5-fold increase in cytokinesis failure (Fig. 5f). Consistent with this observation, PKM2C326S failed to interact with the spindle checkpoint protein BUB3 during metaphase (Fig. S7b), indicating that PKM2 sulfhydration is required to facilitate chromosome segregation, cytokinesis, and cell cycle progression. To confirm that the increased cytokinesis failure in cells expressing PKM2C326S was mainly due to the inhibition of sulfhydration but not caused by other cysteine PTMs, we depleted H2S production by AOAA or by siRNA knockdown of CBS and CTH in MDA-MB-231 cells. Consistent with the observation in PKM2C326S expressing cells, treatment with AOAA or CBS/CTH knockdown resulted in an increased percentage of polyploid cells (Fig. S7c–g). A similar effect was also observed in PC3 cells after AOAA treatment (Fig. S7h). Overall, our findings indicate that blocking PKM2 sulfhydration at C326 or depleting of H2S by inhibitors or siRNA knockdown leads to impaired cytokinesis and subsequent inhibition of cell division.
Blockade of PKM2 sulfhydration reduces cell proliferation and inhibits tumor growth in a mouse xenograft model
Studies have shown that PKM2 tetramerization suppresses tumor growth by either overexpressing PKM1 or treating tumors with PKM2 activators12. Consistent with this observation, the cell proliferation rate was dramatically reduced in MDA-MB-231 cells expressing the PKM2C326S mutant (Fig. 6a). More significantly, tumor growth was completely suppressed in the C326S group in the mouse xenograft model implanted with MDA-MB-231 cells expressing wild-type PKM2 or PKM2C326S (Fig. 6b–e, S8a, b). Overall, our data demonstrated that disruption of PKM2 sulfhydration at C326 results in significant suppression of tumor growth in a mouse xenograft model, and PKM2 sulfhydration may be an excellent therapeutic target for developing anticancer drugs.
Discussion
Metabolic alteration is one of the major hallmarks of cancer. To grow rapidly, tumor cells display a notable increase in glucose uptake and greater demand for metabolic intermediates for biosynthesis. In this study, we show that H2S mediates PKM2 activity through protein sulfhydration, resulting in the dissociation of the high-activity tetramer to the low-activity monomer/dimer. This PKM2 modification rewires glucose metabolism to meet the demands of high biosynthesis and low energy production for cancer cells to proliferate rapidly. In contrast, blocking PKM2 sulfhydration at C326 by mutation or depletion of H2S via inhibitor or siRNA knockdown prevents tetramer dissociation and nuclear translocation of PKM2, facilitating oxidative phosphorylation for high energy production and low biosynthesis, and leading to significant inhibition of tumor growth (Fig. 6f).
Cysteine is the major target for redox PTMs in response to environmental stresses35. In this study, we investigated the significance of cysteine sulfhydration on PKM2 oligomerization, particularly focusing on cysteine 326. Using a proteomic approach, we demonstrated that cysteine 326 was endogenously sulfhydrated (Fig. 2c). Moreover, none of the other PTMs at cysteine 326 were detected in our LC-MS/MS analysis (Table S1). However, it is challenging to rule out the possibility of other PTMs occurring at this site, as most cysteine modifications are transient and reversible35. In fact, these modifications can sometimes occur sequentially; for instance, H2S can react with oxidized, S-nitrosated cysteines, or cysteine disulfides, to form sulfhydrated cysteines23,34,44. To assess the importance of cysteine 326 sulfhydration on the dynamics of PKM2 functions, we substituted cysteine with serine at position 326. However, this substitution not only blocked sulfhydration but also prevented other potential modifications. Therefore, to unravel whether the significance we observed in PKM2C326S is primarily due to sulfhydration, we examined PKM2 functionality in cells with H2S depletion achieved by AOAA treatment or siRNA knockdown of H2S-producing enzymes, CBS and CTH. Consistent with the results obtained from PKM2C326S (Figs. 4–6), treatment with AOAA or knockdown of CBS and CTH led to the inhibition of PKM2 nuclear localization (Fig. S1h–k), cytokinesis inhibition (Fig. S7c–g), and cell proliferation (Fig. 1f). Our data obtained from AOAA treatment and CBS/CTH knockdown support the hypothesis that the effects observed upon blocking PKM2 at cysteine 326 are largely mediated through sulfhydration.
Structurally, replacing Cys with Ser at residue 326 of PKM2 remotely changes the dynamics of the C domain and promotes tetramerization (Figs. 3, S4). Considering the importance of PKM2 in cancer metabolism, PKM2 has emerged as a promising therapeutic target for the treatment of cancer. Small-molecule inhibitors that selectively activate PKM2, such as TEPP-46, have been developed45,46. These synthetic PKM2 activators promote the association of PKM2 subunits into constitutive tetramers, resulting in increased enzymatic activity and reduced tumor size in mouse xenografts12. Here our study uncovered a regulatory mechanism of PKM2 activity in that PKM2C326S mutation blocks C326 sulfhydration, resulting in an increase in the tetrameric state in the absence of FBP or other activators (Fig. 2d, e). Interestingly, the tetramer organization of PKM2C326S is different from those of the currently known T and R states. Compared to all known PKM2 activators that activate PK activity constitutively, activators targeting PKM2C326 may activate PK activity in a much more moderate manner, which may reduce the incidence of undesired side effects on normal cells pertaining to their growth and function.
Cell cycle checkpoints are surveillance mechanisms that ensure the proper progression of cell division47. Among them, the spindle assembly checkpoint (SAC) at the metaphase-anaphase transition is essential for the checkpoint to delay anaphase until the last chromosome is aligned on the metaphase plate48. PKM2 reportedly interacts with and phosphorylates Bub3 in metaphase, which is important for the recruitment of the Bub3-Bub1 complex to the Blinkin complex and microtubule spindle attachment to kinetochores41. Our data indicate that the PKM2C326S mutant failed to bind Bub3 in metaphase (Fig. S7b), which may lead to chromosomal misalignment, and subsequently, cause abnormal chromosome segregation and lagging chromatids in anaphase41. This condition is called aneuploidy, which results in high-level micronuclei and may further cause cytokinesis failure49, as we observed in cells expressing the PKM2C326S mutant (Fig. 5), as well as in cells with H2S depletion by AOAA (Fig. S7c–f) or siRNA knockdown (Fig. S7g).
Our data showed that cells expressing PKM2C326S under cell culture conditions exhibited only a slight decrease in the cell proliferation rate compared to that of cells expressing PKM2WT (Fig. 6a). However, in the mouse xenograft model, tumor growth was greatly inhibited (Fig. 6b–e). It is important to note that tumor metabolism depends strongly on the metabolites supplied in the surrounding environment50. While the cell culture medium is designed to provide cancer cells with abundant nutrients essential for their continuous proliferation51, it is possible that in cell experiments, cells may rely less on aerobic glycolysis due to the excess metabolites present in the medium compared to the actual tumor microenvironment in vivo. Previous studies have indicated that excess pyruvate and cystine in the medium can provide cancer cells with more energy sources and flexibility to switch between different metabolic pathways, potentially leading to reducing their dependence on glucose metabolism52,53. Additionally, excess pyruvate in the medium may even trigger the activation of HIF1α under nonhypoxic conditions54. These observations help explain why tumor growth was significantly inhibited in the C326S group compared to the proliferation rate observed in cell culture. Future studies should carefully consider how the components of the medium may impact on different metabolic pathways to support cancer cell growth. Interestingly, ectopic expression of wild-type PKM2 in MDA-MB-231 cells did not further enhance metabolism, cell cycle, or tumor growth either in vitro or in vivo compared to that in cells expressing the vector alone (Figs. 4–6). We speculate that these findings may stem from the high expression of endogenous PKM2 in cancer cells, which is adequate for its conversion into the monomeric/dimeric form of PKM2 and subsequent translocation into the nucleus. This translocation exerts significant regulatory effects, including promoting cancer cell metabolism, cell cycle progression, and proliferation. This hypothesis is supported by our data indicating that the average nuclear translocation of PKM2 was 14.52% in control cells (Fig. 1d), while cells overexpressing wild-type PKM2 exhibited a similar average nuclear translocation ratio (15.92%; Fig. 4i). Consequently, it is likely that the level of endogenous PKM2 translocated into the nucleus is sufficient to modulate HIF-1α and other transcription factors, to promote cancer cell metabolism, cell cycle progression, and proliferation. Hence, the efficacy of ectopic expression of wild-type PKM2 was indistinguishable from that of the vector control group.
Most cancer cells undergo metabolic reprogramming in favor of aerobic glycolysis to meet their bioenergetics and biosynthetic demands during tumorigenesis. In breast cancer cells, two H2S-producing enzymes, CBS and CTH, were upregulated compared to those in normal cells (Fig. 1g). Elevated levels of CBS or CTH were positively associated with increased sulfhydrated PKM2 levels in all breast cancer cell lines we examined (Fig. 1h), along with upregulated expression of PKM2-response genes in clinical samples (Fig. S2c), suggesting that cancer cells rewire glucose metabolism by modifying PKM2 activity via H2S-mediated sulfhydration. Moreover, PKM2 sulfhydration was elevated in breast and prostate cancer cells under hypoxia (Fig. 1i, S2h), suggesting that cancer cells may alter their metabolic pathways through PKM2 sulfhydration in response to the hypoxic tumor microenvironment. By blocking H2S-mediated sulfhydration at C326, we provided a promising therapeutic opportunity for the activation of PKM2, resulting in a glucose metabolic switch (Fig. 4) and drastic tumor growth inhibition in a mouse xenograft model (Fig. 6). A previous study indicated that the expression levels of glucose transporters, such as GLUT1, and glucose metabolic enzymes are high in breast cancer, especially in triple-negative breast cancer (TNBC)55, suggesting that glucose metabolism may act as the main source of energy and building materials in breast cancer. Moreover, glucose metabolism is associated with tumor growth, metastasis, and drug resistance in breast cancer56–58. In this study, we showed that breast cancer cells respond significantly to glucose metabolism alterations through the blockade of H2S-mediated PKM2 C326 sulfhydration. The development of inhibitors that block PKM2 sulfhydration could be a metabolic therapy for breast cancer patients.
Studies have indicated that monomeric or dimeric PKM2 exhibits intriguing non-glycolytic functions upon entering the nucleus. Nuclear PKM2 assists in lactate production through the transactivation of HIF-1α, thereby promoting LDHA expression9. However, our data showed that the nuclear translocation of PKM2 and the expression of HIF-1A were notably suppressed in cells expressing PKM2C326S, while LDHA expression remained unaffected (Fig. 4g, i). Consequently, lactate production and ECAR was not inhibited in the PKM2-C326S group (Fig. 4e, f, h). As LDHA expression is modulated by not only HIF-1α but also other transcription factors, including c-MYC, FOXM1, ErbB2, estrogen, EGFR1, and SIRT259, we speculate that LDHA transcription may be influenced by alternative mechanisms even though HIF-1α expression is suppressed in cells expressing PKM2C326S (Fig. 4g). Future investigations could be conducted to examine whether LDHA expression is influenced by the abovementioned transcription factors in PKM2C326S expressing cells. Additionally, studies have demonstrated that cytosolic PKM2 stabilizes EGFR by maintaining its binding with HSP90 to promote cell proliferation60, while another study also showed that mitochondrial PKM2 stabilizes Bcl2 to enhance apoptosis resistance61, suggesting that PKM2 plays non-glycolytic roles in not only the nucleus but also the cytosol and mitochondria. However, it remains unclear whether PKM2 sulfhydration is involved in modulating PKM2 functions in the cytosol and mitochondria. It is possible that through these mechanisms, cells expressing PKM2C326S somehow compensate for the expression of LDHA, leading to the contradictory results shown in Fig. 4.
In addition to the modulation of cancer glucose metabolism, endogenously produced H2S is also essential and sufficient for dietary restriction and an extended lifespan21. A subsequent study further indicated that H2S-induced angiogenesis under amino acid restriction occurs through the inhibition of mitochondrial OXPHOS, increased glycolytic ATP production, and increased glucose uptake in endothelial cells62. Here our data further support their discovery and provide an excellent explanation of the molecular basis by which H2S may facilitate the OXPHOS/glycolysis switch through destabilizing PKM2 tetramerization. We should also consider PKM2 sulfhydration and other known H2S mediated functions, such as vasodilation, immunomodulation, or prosurvival signals, to decipher the underlying mechanisms of PKM2 sulfhydration under both normal physiological and dysregulated disease conditions.
Methods
Ethical statement
The animal study complies with the Institutional Animal Care and Use Committee (IACUC) guidelines at National Tsing Hua University (NTHU-IACUC-10656).
Antibodies
PKM2 (D78A4) XP® Rabbit mAb (#4053) (1:1000 for immunoblotting; 1:100 for immunostaining), HA-Tag (C29F4) Rabbit mAb (#3724) (1:1000), and V5-Tag (D3H8Q) Rabbit mAb (#13202) (1:1000 for immunoblotting; 1:100 for immunostaining) were from Cell Signaling Technology (CST). CTH Antibody (F-1) (#sc-374249) (1:500), GAPDH Antibody (6C5) (#sc-32233) (1:1000), Tubulin Antibody (B-5-1-2) (#sc-23948) (1:100), mouse anti-rabbit IgG-HRP (#sc-2357) (1:5000), and m-IgGκ BP-HRP (anti-mouse) (#sc-516102) (1:5000) were from Santa Cruz Biotechnology. Goat anti-Rabbit IgG (H+L) Cross-Adsorbed Secondary Antibody, Alexa Fluor™ 594 (#A-11012) (1:300) and Goat anti-Mouse IgG (H+L) Cross-Adsorbed Secondary Antibody, Alexa Fluor™ 488 (#A-11001) (1:300) were from Invitrogen. Recombinant Anti-CBS antibody [EPR8579] (#ab140600) (1:1000) and Recombinant Anti-Bub3 antibody [EPR5319(2)] (#ab133699) (1:10000) were from abcam. MPST (3-MST) antibody [C2C3], C-term (#GTX108274) (1:1000) was from GeneTex. Rabbit IgG, Purified (#PP64B) (0.5ug) was from EMD Millipore.
Chemicals, DNA, and siRNAs
S-Methyl methanethiosulfonate (MMTS) (#64306), Iodoacetamide (IAM) (#I6125), and O-(Carboxymethyl)hydroxylamine hemihydrochloride (AOAA) (#C13408) were from Sigma-Aldrich. The PKM2 C326S mutagenesis was generated by site-directed mutagenesis using PfuUltra II Fusion HS DNA polymerase (Agilent Technologies) in PCDNA-HA-PKM2 plasmid63 and pLX304-V5-PKM2. The pLX304-V5-PKM2 was obtained from CCSB-Broad Lentiviral Expression Library (Dharmacon). The Allstars negative control siRNA was from Qiagen. The CBS-specific and CTH-specific siRNAs were purchased from MDBio, Inc (Qingdao, China). All siRNA sequences used in this study are listed in Table S4.
Cell culture and transfection
MDA-MB-231 (ATCC Cat# HTB-26), MCF-7 (ATCC Cat# HTB-22), HCC-1395 (ATCC Cat# CRL-2324), PC3 (ATCC Cat# CRL-1435) were purchased from Bioresource Collection and Research Center (BCRC), Taiwan. Non-tumourigenic mammary epithelial line MCF-10A (ATCC Cat# CRL-10317) was purchased from ATCC directly. MDA-MB-231, MCF-7, and PC3 cell lines were cultured in Dulbecco’s Modified Eagle’s Medium (DMEM, Gibco) while HCC-1395 cells were maintained in Roswell Park Memorial Institute (RPMI, Invitrogen) supplemented with 10% fetal bovine serum (FBS, HyClone) and 1% Penicillin/Streptomycin (P/S, Gibco). MCF-10A cell line was cultured in DMEM/F-12 (Gibco) supplemented with 5% horse serum (HyClone), 20ng/mL epidermal growth factor (EGF, PEPROTECH, #AF-100-15), 0.5μg/mL hydrocortisone (Sigma-Aldrich, #H0888), 10μg/mL insulin (Sigma-Aldrich, #I9278), and 1% Penicillin/Streptomycin (P/S, Gibco). All cells were cultured at 37°C and 5% CO2 with humidity. MDA-MB-231 and PC3 cells stably expressing GFP-Luciferase, together with pLX304, pLX304-V5-PKM2, or pLX304-V5-PKM2-C326S, were generated by lentivirus infection and established as mass culture, selected and maintained in 1μg/mL puromycin (Cyrusbioscience, #101-58-58-2) and 1μg/mL Blasticidine S hydrochloride (BSD, Cyrusbioscience, #101-3513-03-9), respectively. Plasmids were transfected using Opti-MEM (Gibco, #31985070) and TransIT-X2 (Mirus, #MIR 6000). siRNA was transfected using Opti-MEM (Gibco, #31985070) and Lipofectamine RNAiMax (Invitrogen, #13778075).
Immunocytochemistry (ICC)
Cells (0.8–3×104 cells) were seeded on poly-L-lysine (0.01%, Sigma-Aldrich, #P8920) and fibronectin (10μg/mL, Sigma-Aldrich, #FC010) coated glass coverslips overnight. For treatment with NaHS, the cell culture medium was changed to a serum-free medium and incubated in a standard incubator. After 24-h incubation, cells were treated with 1μM NaHS for 1h. For treatment with AOAA, cells were treated with 0.25mM AOAA or H2O control for 24–72h. Then, cells were fixed in 4% formaldehyde for 30min, washed three times with PBS, and permeabilized with 0.1% Triton X- 100 in PBS for 10min. Cells were incubated in blocking buffer (1% BSA in PBS with 5% FBS, filtered before use) for 1h at room temperature before incubation with Rhodamine Phalloidin (Invitrogen, #R415), primary [PKM2 (D78A4) XP® Rabbit mAb, or Anti-α Tubulin Antibody (B-5-1-2)] and secondary [Goat anti-Rabbit IgG (H+L) Cross- Adsorbed Secondary Antibody, Alexa Fluor 594 or Goat anti-Mouse IgG (H+L) Cross-Adsorbed Secondary Antibody, Alexa Fluor™ 488] antibodies. Incubations with both primary and secondary antibodies (diluted in PBST) were done in the presence of 1% FBS. Mounting medium (EverBrite, #23008-T) with 1μg/mL DAPI (Sigma-Aldrich, #D8417) was used to adhere a coverslip to a slide glass. Fluorescence was visualized using a confocal laser scanning microscope (ZEISS LSM 800) or fluorescence microscope (Nikon ECLIPSE Ni and Nikon MODEL ECLIPSE TI2-E). Images were analyzed by ZEN 2012 (Carl Zeiss), NIS-Elements (Nikon), and ImageJ64 software. Images taken from Nikon fluorescence microscope were deconvolved using Richardson-Lucy method with 30 iterations from NIS-Elements software.
Time-lapse live cell microscopy
Cells were seeded in 35mm dishes (Alpha Plus; Scientific Corp., Taoyuan, Taiwan) and incubated overnight. Imaging was performed using a Leica DMI6000 microscope (Leica Microsystems, Wetzlar, Germany) with an HCX PL FL × 20/0.4 numerical aperture objective and an Andor Luca R EMCCD camera (Andor Technology, Belfast, United Kingdom). Differential interference contrast images were captured from multiple positions at a 10min interval for 24h. The percentage of cells that undergo cytokinesis failure was counted manually. For each experiment, at least 40 cells undergoing mitosis were counted. The experiments were repeated three times independently. All images were analyzed using MetaMorph software (Molecular Devices, San Jose, CA, USA).
Quantitative real-time PCR
Cells were treated with serum-free medium containing 1µM NaHS or growth medium containing 0.25mM AOAA for 24h before RNA extraction. RNA was extracted from cells using TRIzol (Invitrogen, #15596018) and Chloroform Replacement (Cyrusbioscience, #CRSR) following protocols supplied by the manufacturer. First-strand cDNA was generated by ReverTra Ace (PURIGO, #PU-TRT-100) using oligo-dT. Real-time PCR was performed on a qTOWER³ real-time PCR detection system (Analytik Jena) and analyzed with qPCRsoft version 3.4 software (Analytik Jena). The KAPA SYBR® FAST qPCR Master Mix (2X) Kit (Roche, KK4600) was used. The mRNA levels were normalized to that of actin or GAPDH. All primer sequences used in this study are listed in Table S4.
Modified biotin switch (S-sulfhydration) assay
The assay was performed with modifications based on a published paper65. In brief, after various treatments, cells were lysed in a HEN buffer [100mM HEPES-NaOH (pH 8.0, Sigma-Aldrich, #H4034)], 1mM EDTA, 0.1mM neocuproine, 100μM deferoxamine, and 1X protease inhibitor (Roche)], and centrifuged at 13,000×g for 30min at 4°C. Recombinant proteins (10–50μg) or cell lysates (0.4–1mg) were treated with 100µM NaHS at 37°C for 30min. [Optional: For certain condition, 1mM DTT was added for additional 30min at 37°C only to verify that the pull-down proteins contain SSH-bond (Fig. 1a, S1A)]. The lysates were then mixed to a HEN buffer with 2.5% SDS and 20mM MMTS at 50°C for 20min with constant vortexing. Proteins were precipitated by acetone at −80°C for at least 20min to overnight. After removing the acetone, the proteins were resuspended in a HEN buffer with additional 1% SDS and 4mM biotin-HPDP (EZ-link, ThermoFisher). Following 3-h incubation at RT, biotinylated proteins were pulled down by streptavidin-agarose beads at 4°C overnight and then washed with buffer containing 25mM HEPES (pH 7.5), 1mM EDTA, 600mM NaCl, and 0.5% Triton X-100. The biotinylated proteins were subjected to Western blot for further analysis.
Western blotting
Cells were lysed in a 1X Lysis buffer [50mM Tris buffer, pH 7.4, 150mM NaCl, 1% Triton X-100, 1mM EDTA, 0.1% SDS, and 1X protease inhibitor (Roche)]. The cell lysates were resolved in 10 or 12.5% SDS-polyacrylamide gel, transferred onto the PVDF membrane, and probed with antibodies. ECL reagent (Perkin Elmer, #NEL105001EA or Thermo Scientific, #34095) was used to capture luminescence by ImageQuant™ LAS 4000 (GE). Protein expression levels were quantified with Image J (NIH, USA). Data are means±SEM from at least 3 independent experiments and presented as fold changes relative to the control. All immunoblotting experiments were repeated at least 3 times with similar results.
Real-time measurements of OCR and ECAR
Seahorse Xfe24 Analyzer (Agilent) was used for cell mito stress test and glycolysis stress test, i.e., measured oxygen consumption rate (OCR) and extracellular acidification rate (ECAR). Seahorse Xfe Wave software (Agilent) was used for data analysis and report generation. Cells were seeded into the wells of the microplate at different densities, which would be a monolayer on the next day, in 100μL of culture medium. After 1–5h, cells were attached and 150μL of culture medium was added. On the day of assay, culture media were removed gently, assay media (DMEM, without buffer reagent, 2% FBS) were added, and cells were incubated at 37°C in the absence of CO2.
Cell mito stress test (Agilent, #103015-100)
The test was performed according to the methods described in the manufacturer manual. Basal OCR was measured three times before adding 1μM oligomycin (inhibitor of ATP synthase). The second injection was carbonyl cyanide-4 (trifluoromethoxy) phenylhydrazone (FCCP), which is an uncoupling agent that collapses the proton gradient and disrupts the mitochondrial membrane potential. The third injection is a mixture of rotenone (inhibitor of complex I) and antimycin A (inhibitor of complex III).
Glycolysis stress test (Agilent, #103020-100)
The test was performed according to the methods described in the manufacturer manual. ECAR of cells was measured three times before and after sequential injections of three different compounds: 10mM glucose, 1μM oligomycin (inhibitor of ATP synthase), and 100mM 2- Deoxy-D-glucose (2-DG, a glucose analog that inhibits glycolysis).
Purification of recombinant proteins
His-tagged WT and mutant PKM2 were expressed in Escherichia coli incubated at 37°C and inducted with 0.1mM isopropylthiogalactopyranoside (IPTG) at 25°C for 5h. Recombinant proteins were then purified by Ni-NTA-agarose (Qiagen) affinity chromatography. The elution buffer contained 20mM Tris (pH 8.0, APOLO Biochemical, #APL-0082), 150mM potassium chloride (KCl, Sigma-Aldrich, #P5405), 1mM β-mercaptoethanol (Cyrusbioscience, #101-60-24-2), 1mM PMSF, and 100-400mM imidazole (Cyrusbioscience, #101-288-32-4). The eluate proteins were further purified by gel filtration chromatography (Superdex 200 Increase 10/300 GL, Cytiva) in PBS buffer (pH 8.0) and 10% glycerol (J.T. Baker, #2136-08). Protein concentration was determined at 280nm using a Colibri microvolume Spectrophotometer (Titertek-Berthold) and an extinction coefficient of 29,910M−1cm−1. All proteins were flash-frozen by liquid nitrogen and stored as aliquots at –80°C.
Pyruvate kinase assay
The pyruvate kinase reaction rate was determined by LDH-coupled enzyme assays. In a 200μL PBS (pH 8.0) buffer containing 5mM MgCl2 (SHOWA chemical, #1301-7260), 1mM ADP (Sigma-Aldrich, #A2754), 0.4mM NADH (Sigma-Aldrich, #N8129), 2 U LDH (Sigma-Aldrich, #L1254), PKM2 proteins were incubated at room temperature for 5min. The addition of 0.5mM PEP (Sigma-Aldrich, #P0564) initiated the PKM-LDH coupled reaction. The absorbance at 340nm was measured every 10 or 12s for 15min by a Synergy HTX plate reader (BioTek). An NADH standard curve was used to determine the reaction rate. All reported rates were normalized by protein concentration. NaHS was incubated with protein on ice for 30min and then assayed immediately. FBP was from Sigma-Aldrich (#F6803). For cell lysates, pyruvate kinase activity was measured using the Pyruvate Assay Kit (Biovision) as per the manufacturer’s instructions. The amount of pyruvate produced was calculated from a standard curve from the provided pyruvate standard.
Mass spectrometry analysis to detect PTM in recombinant proteins
His-tagged recombinant PKM2 proteins were treated with 100μM NaHS for 30min at 37°C and then performed in solution trypsin digestion (n=1). Peptide digests are further purified by C18 Zip-tip and evaporated in a speed vacuum before injection. Samples were then dissolved in 8μL of 2% acetonitrile with 0.1% trifluoroacetic acid. Samples were injected into the Waters NanoAcquity UPLC-Synapt G2 HDMS at National Health Research Institutes mass spectrometry facilities. The sulfhydrated modification on PKM2 was further analyzed by MASCOT.
Mass spectrometry analysis to detect endogenous PTM in cell lysates
On-bead digestion
MDA-MB-231 cells were pre-incubated with 100nM NaHS for 30min followed by 10mM MMTS for 20min before lysed in non-denaturing lysis buffer [20mM Tris-HCl pH 8, 137mM NaCl, 1% NP-40, 2mM EDTA, and 1X protease inhibitor (Roche)] (n=1). The cell lysates were then centrifuged, and the supernatants were collected and diluted to the same concentration. The cell lysates were then mixed with anti-PKM2 antibody and Protein A Mag Sepharose Xtra (Cytiva). The mixture was gently rocked overnight at 4 °C on the rocker. The next day, the beads were washed 3 times with the non-denaturing lysis buffer and then 3 times with PBS. Then, the beads were suspended in 50mM HEPES (pH 8) with 25mM IAM at RT for 30min. Afterward, samples were sequentially digested by Lys-C protease (FUJIFILM Wako Pure Chemical Corporation, #125-05061) at 37°C for 3h, followed by trypsin (Promega, #V511A) at 37°C overnight. After digestion, the peptide mixtures were extracted and desalted using Pierce C18 Spin Tips (Thermo Scientific, #84850) according to the manufacturer’s protocol. Elutes were dried under vacuum and re-suspended in 0.1% formic acid (J.T. Backer, #9834-03).
Shotgun proteomic identifications
NanoLC-nanoESI-MS/MS analysis was conducted using a nanoAcquity system (Waters, Milford, MA), coupled to the Orbitrap Elite hybrid mass spectrometer (Thermo Electron, Bremen, Germany) with a PicoView nanospray interface (New Objective, Woburn, MA). The sample of peptide mixtures was applied onto a 75μm ID, 25cm length C18 BEH column (Waters, Milford, MA) with 1.7μm particles with a pore size of 130Å. Peptide mixtures were separated through a segmented gradient over 30min, starting from 5% to 25% (0 to 27.5min), 25% to 35% (28.5–30min) solvent B containing acetonitrile (J.T. Backer, #9829-03) with 0.1% formic acid as a flow rate of 300 nL/min and a column temperature of 35°C. Solvent A was 0.1% formic acid in ddH2O. The mass spectrometer operated in data-dependent mode, acquiring survey of full scan MS spectra in the orbitrap (m/z 350–1600) with a resolution to 60K at m/z 400 and an automatic gain control (AGC) targeting at 106. The fifteen most intense ions were isolated sequentially for HCD MS/MS fragmentation and detected in the orbitrap, with dynamic exclusion of previously selected ions for 60s. MS/MS analysis was conducted with a resolution of 15,000, an isolation window of 2m/z, and a target value of 50,000ions, with a maximum accumulation time of 200ms. Fragmentation employed a normalized collision energy of 30% and the activation time of 0.1ms. Exclusion criteria: ions with single and unrecognized charge states were excluded. For targeted LC-MS/MS, mass tolerances were set to ±10 ppm as inclusion list, with monoisotopic precursor ion selection enabled.
Database searching
Raw data files were processed using Proteome Discoverer v3.0.1.27 (Thermo Fisher Scientific) and the tandem MS data were then searched using SEQUEST algorithms against the protein sequence of PKM2, Protein A, Ig gamma chain C region, and a human UniProt (Swiss-Prot only) database (released February 2023) with common contaminant proteins. The search parameters included trypsin as the protease with a maximum of 2 missed cleavages allowed; +15.9949Da (oxidation of methionine and cysteine), +0.984Da (deamidation of asparagine and glutamine), +31.990Da (dioxidation of cysteine), +47.985Da (trioxidation of cysteine), +57.021Da (carbamidomethylation of cysteine), and +88.994Da (C2H3NOS, Cys-S-S- CAM, for IAM-tagged persulfides) were set as a dynamic modification. Precursor mass tolerance was set to 10ppm and fragment mass tolerance was set to 0.02Da. The false discovery rates of proteins and peptides were set to 0.01. These data have been deposited in the Proteomics IDEntifications (PRIDE) database (Project accession: PXD044925).
Mass spectrometry analysis for quantitation of metabolites
Cell pellets were re-suspended in 1mL 100% methanol (precooled to –80°C) and quenched in liquid nitrogen. The frozen-quenched cells were thawed, vortexed for 30–60s, and pelleted by centrifugation. The supernatant was transferred to a fresh tube, the cell pellets were re-suspended in 1mL 100% methanol (precooled to –80°C), and the following procedures were repeated. The supernatant was evaporated in a centrifugal evaporator to generate a dry metabolite pellet. Xevo TQ-S triple quadrupole (tandem quadrupole) mass spectrometer (Waters) was used for the quantitation of metabolites involved in the pathways of glycolysis and the TCA cycle.
Cell proliferation assay
The cell proliferation assay was measured with CellTiter 96 Aqueous One Solution (MTS, Promega, #G3580). The assay was performed according to the methods described in the manufacturer’s manual. Briefly, cells (1×103 cells) were seeded in 96-well plates and incubated in a standard incubator. At defined time points, 20μL of CellTiter 96 Aqueous One Solution Reagent was added and incubated for 2h at 37°C. The quantity of formazan product, which is directly proportional to the number of living cells in the culture, was measured by absorbance at 490nm with a 96-well plate reader. In the siRNA knockdown experiment, transfected cells (5×103 cells; post-transfection for 5h) were seeded in 96-well plates and incubated in a standard incubator. At defined time points, live cells were stained with 1μg/mL Hoechst (diluted in DMEM, Sigma-Aldrich, #33342) for cell counting.
Glycerol gradient ultracentrifugation
To perform glycerol gradient ultracentrifugation, 150μg cell lysates or 10μg recombinant PKM2 proteins were loaded on top of 15–35% glycerol gradient. Following centrifugation at 304,000×g for 16h at 4°C in SW 55 Ti Rotor (Beckman Coulter), fractions (100μL each) were taken from top to bottom and the oligomerization of PKM2 was analyzed by Western blot.
PI staining
In total, 0.5–2×106 Cells were fixed in 1mL PBS and 9mL 70% ethanol and stored at –30°C overnight. Cells were collected through centrifugation, washed, and resuspended in 500μL PI/Triton X-100 staining solution [0.1% (v/v) Triton X-100 (Sigma-Aldrich) in PBS add 0.1mg DNAse-free RNAse A (Sigma-Aldrich) and 10μg PI (Sigma-Aldrich)]. The stained samples were incubated for 30min in the dark at room temperature. After PI staining, the stained samples were analyzed using CytoFLEX (Beckman Coulter). For the flow cytometry module, FSC-A and SSC-A was used to gate for the bulk population of cells. Data were collected in 10,000 or 50,000 single cells and used to form a histogram of fluorescence area signals and analyzed by FlowJo or Floreada.io software.
Senescence associated β-galactosidase (SA-β-gal) staining
MDA-MB-231 stable cell lines (5×103 cells) were seeded in a 24-well plate and incubated 4 days before staining. Cells were fixed with 4% formaldehyde and stained with SA-β-gal staining solution [0.1% X-gal, 5mM potassium ferrocyanide, 5mM potassium ferricyanide, 150mM NaCl, and 2mM MgCl2 (SHOWA chemical, #1301-7260) in 40mM citric acid/sodium phosphate solution, pH 6.0] for 3 days. Bright-field images were captured with ×100 magnification.
Cell cycle synchronization
MDA-MB-231 cells were transfected with vector control, PKM2wt, or PKM2C326S mutant 24h before cell cycle synchronization. Cells were synchronized to G2/M transition with double thymidine block (2mM) and RO-3306 (2.5μg/mL), and then released in culture medium for 6h to arrest in metaphase.
Immunoprecipitation (IP)
Cells were lysed in non-denaturing lysis buffer [20mM Tris-HCl pH 8, 137mM NaCl, 1% NP-40, 2mM EDTA, and 1X protease inhibitor (Roche)]. The cell lysates were then centrifuged, and the supernatants were collected and diluted to the same concentration. The cell lysates (500μL) were mixed with 0.5μg anti-Bub3 antibody (Abcam) or IgG control (Millipore), and 50μL protein A agarose bead slurry (25μL packed beads). The mixture was gently rocked overnight at 4°C on the rocker. After the co-IP, the beads were removed from each sample and mixed with 2X sample buffer for western blot analysis.
In vivo xenograft mouse model
The female SCID mice (C.B17/Icr-Prkdcscid/CrlNarl, NARLabs) were randomly assigned to three groups after arriving and injected with 2×106 luciferase stable expressed breast cancer cells (MDA-MB-231 with GFP-luc-2A plus pLX304, pLX304-PKM2, or pLX304-PKM2-C326S) in 100µl PBS containing Matrigel (Corning, #356237) per mouse at the 4th mammary fat pad at age of 8–10 weeks. Mice were housed in a 12h light/12h dark cycle at 22-26 °C with 45–65% humidity, with access to sufficient diet and water. The cages were cleaned and sanitized on regular basis. The cells’ luciferase activity (Promega, #E1483 or PerkinElmer, #6016716) had been confirmed before injection. Mice were measured for body weight and tumor size once per week. Once the largest diameter of the largest tumor reached almost 20mm, mice were intraperitoneally (IP) injected with luciferin (Promega, #P1043) for in vivo imaging system (IVIS, Perkin Elmer). After that, mice were sacrificed, and tumors were harvested for IVIS. All procedures used in our mouse study were performed according to the approved protocol by the Institutional Animal Care and Use Committee of National Tsing Hua University (NTHU-IACUC-10656). The maximal tumor size permitted by our ethics committee was strictly set to be less than 20mm in diameter, and we confirm that this limit was not exceeded at any point during our study.
Mass photometry and gel filtration analyses
Mass photometry was performed using 1.6nM sulfhydrated PKM2 on a TwoMP Mass Photometer (Refeyn Ltd.) in assay buffer containing 20mM Tris (pH. 8.0) and 150mM KCl. For gel filtration, 200μL of purified protein samples at 0.3mg/mL were injected into a Superdex 200 10/300 GL column (Cytiva) equilibrated in 20mM Tris (pH. 8.0) and 50mM KCl at 4°C. Protein profiles were monitored by the absorbance at 280nm.
Crystallization and data collection
For crystallization of the PKM2C326S, purified protein was concentrated to 8mg/mL in 25mM Tris (pH 8.0, APOLO Biochemical, #APL-0082), 100mM KCl (Sigma-Aldrich, #P5405), 1mM DL-Dithiothreitol (DTT, UniRegion Bio-Tech, #UR-DTT), 10% glycerol (J.T.Baker, #2136-08). Protein crystals were grown in 0.4M ammonium phosphate monobasic, 23% polyethylene glycerol 3350, Tris pH 7.0 at 22 °C by hanging drop vapor diffusion method and further improved by microseeding. The diffraction data were collected at a wavelength of 0.97626Å and a temperature of 100K using beamline BL07A at the National Synchrotron Radiation Research Center (Taiwan), merged from several datasets from the same crystal, and processed by HKL200066.
Structural determination and refinement
The crystal structure of PKM2C326S was solved by molecular replacement in Phenix67–69 using wild-type PKM2 (PDB ID: 4QG6) without the B domain as the search model. Structural refinement and model building were carried out by Phenix67–69 and Coot70 iteratively. Structural analysis and presentation were performed by PyMOL71.
Molecular dynamics simulation
The PKM2 model was prepared with BIOVIA Discovery Studio software (Dassault Systèmes BIOVIA, Discovery Studio Modeling Environment, Release 2022, San Diego: Dassault Systèmes, 2016). Chain D of the T state (PDB: 4FXJ) was used as the template to repair the missing residues in each chain by using homology modeling tools. The PKM2 tetramer was solvated in water-filled orthorhombic periodic boxes. The minimum distance from the periodic boundary was set to 7Å. To neutralize the system, sodium chloride was added. Total of 45,787 water molecules, 121 sodium ions, 136 chloride ions were added to the solvated model. Molecular Dynamics (MD) simulations of the PKM2 tetramer were performed using the standard dynamic cascade with CHARMM forcefield. The 100-ns MD run in production step at 300K temperature under 1 reference pressure was performed using the NPT mode. The most stable conformer was analyzed. Interface areas were calculated by PISA72.
Clinical analysis from the cancer genome atlas
For gene expression analysis, data containing 1033 breast cancer (BRCA) cases and 59 normal breast tissues were collected from The Cancer Genome Atlas (TCGA) database (https://portal.gdc.cancer.gov/projects/TCGA-BRCA). Gene expressions were represented as transcripts per million (TPM). After data mining, the student’s t test was performed in R (The R Foundation, v 4.2.1). For Pearson correlation analysis, data collected from TCGA-BRCA were applied, and gene expressions were converted to log2(TPM). After data mining, Pearson correlation analyses were performed in R (v 4.2.1). For survival, Kaplan-Meier Plotter (http://kmplot.com/analysis/) was used to assess the prognostic value of CBS from TCGA-BRCA. The hazard ratios (HRs) with 95% confidence intervals (Cis) and log-rank p-values were also computed.
H2S measurement
H2S production capacity was measured in live cells treated with 0.25mM AOAA in medium supplemented with 10 mM L-Cys and 10μM PLP. A lead acetate H2S detection paper (Sigma) was placed above the liquid phase in a covered 96 well plate and incubated 48h. The amount of lead sulfide (PbS) was quantified with Image J (NIH, USA). Data are means±SD from 3 duplicates of 3 independent experiments and presented as fold changes relative to the control.
Quantification and statistical analysis
The statistical details of the experiments are described in figure legends. Most statistical analyses were performed using GraphPad Prism software (GraphPad Prism, CA, USA) with student’s t test to compare two means. ANOVA followed by Tukey’s post hoc test was used for the statistical analysis when more than two means were compared (*p<0.05; **p<0.01; ***p<0.001; ****p<0.0001).
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Supplementary information
Acknowledgements
Mass spectrometry analyses were performed by the Research Facility for Sharing for Proteomics and Chemistry located at NHRI. LTQ-Orbitrap data were acquired at the Academia Sinica Common Mass Spectrometry Facilities for Proteomics and Protein Modification Analysis located at the Institute of Biological Chemistry, Academia Sinica, supported by Academia Sinica Core Facility and Innovative Instrument Project (AS-CFII-111-209). We thank the experimental facilities and the technical services provided by the “Synchrotron Radiation Protein Crystallography Facility of the National Core Facility Program for Biotechnology, NSTC”, the “National Synchrotron Radiation Research Center”, a national user facility supported by the NSTC, and the “Macromolecular Crystallography Center” at NTHU. We thank the technical support at the confocal imaging core at National Tsing Hua University and the technical services provided by the “Next-Generation Nucleic Acid Drug (NGNAD) Platform of the National Core Facility for Biopharmaceuticals, National Science and Technology Council, Taiwan”. Lastly, we thank Kuan-Wen Chen at the Molecular Sciences and Digital Innovation Center (MSDIC), GGA Corp, for the assistance in molecular dynamics simulation. National Science and Technology Council 108-2314-B-007-003-MY3: KTL National Science and Technology Council 111-2320-B-007-005-MY3: KTL National Science and Technology Council 108-2311-B-007-002-MY3: HCC National Science and Technology Council 111-2311-B-007-009: HCC National Tsing Hua University 111Q2713E1, 112Q2511E1, and 112Q2521E1, 113Q2524E1: KTL National Science and Technology Council 110-2320-B-007-004-MY3: LHW National Health Research Institutes NHRI-EX113-11124BI: LHW National Tsing Hua University 112QI033E1: LHW Ministry of Education CMRC-CENTER-0: LHW National Science and Technology Council 112-2320-B-039-003: LHW National Science and Technology Council 113-2634-F-039-001: LHW National Science and Technology Council 2326-B-038: HJK
Author contributions
Conceptualization: L-HW, H-CC, K-TL Data curation: R-HW, Y-TC, P-RC, Y-CC, Y-HC, C-CC, P-CC, S-YL, Z-LW, M-CT, S-YC, G-SC, W-LC, Y-HW, S-YL, H-CC, K-TL Resources: LW, W-CW, H-JK Investigation: H-CC, K-TL Funding acquisition: H-CC, K-TL Project administration: H-CC, K-TL Supervision: H-CC, K-TL Writing – original draft: R-HW, Y-CC, Y-HC, H-CC, K-TL Writing – review & editing: LW, W-CW, H-JK, L-HW, H-CC, K-TL
Peer review
Peer review information
Nature Communications thanks Todd Holyoak, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Data availability
The coordinates and structure factors of the crystal structure of PKM2 C326S in this study have been deposited in the Protein Data Bank under accession code 8HGF. The proteomic data generated in this study have been deposited in the Proteomics IDEntifications (PRIDE) database under accession code PXD044925. Source data are provided with this paper. The remaining data are available within the Article, Supplementary Information or Source Data file. Source data are provided with this paper.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Contributor Information
Lu-Hai Wang, Email: wt.ude.umc.liam@gnawiahul.
Hui-Chun Cheng, Email: wt.ude.uhtn.efil@gnehcch.
Kai-Ti Lin, Email: wt.ude.uhtn.efil@niltk.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-024-51875-9.
References
Articles from Nature Communications are provided here courtesy of Nature Publishing Group
Citations & impact
This article has not been cited yet.
Impact metrics
Alternative metrics
Discover the attention surrounding your research
https://www.altmetric.com/details/166780150
Data
Data behind the article
This data has been text mined from the article, or deposited into data resources.
BioStudies: supplemental material and supporting data
Protein structures in PDBe (3)
-
(2 citations)
PDBe - 4FXJView structure
-
(1 citation)
PDBe - 4QG6View structure
-
(1 citation)
PDBe - 4B2DView structure
ProteomeXchange
- (1 citation) ProteomeXchange - PXD044925
Similar Articles
To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.
Apoptin Inhibits Glycolysis and Regulates Autophagy by Targeting Pyruvate Kinase M2 (PKM2) in Lung Cancer A549 Cells.
Curr Cancer Drug Targets, 24(4):411-424, 01 Jan 2024
Cited by: 2 articles | PMID: 36284386 | PMCID: PMC10964080
The role of PKM2 in cancer progression and its structural and biological basis.
J Physiol Biochem, 80(2):261-275, 08 Feb 2024
Cited by: 0 articles | PMID: 38329688
Review
PKM2 promotes glucose metabolism and cell growth in gliomas through a mechanism involving a let-7a/c-Myc/hnRNPA1 feedback loop.
Oncotarget, 6(15):13006-13018, 01 May 2015
Cited by: 83 articles | PMID: 25948776 | PMCID: PMC4536995
PKM2 functions as a histidine kinase to phosphorylate PGAM1 and increase glycolysis shunts in cancer.
EMBO J, 43(12):2368-2396, 15 May 2024
Cited by: 0 articles | PMID: 38750259 | PMCID: PMC11183095