Next Article in Journal
Microbial Adhesion and Biofilm Formation on Bioactive Surfaces of Ti-35Nb-7Zr-5Ta Alloy Created by Anodization
Next Article in Special Issue
The Effect of Lithium on the Budding Yeast Saccharomyces cerevisiae upon Stress Adaptation
Previous Article in Journal
Investigations on the Degradation of the Bile Salt Cholate via the 9,10-Seco-Pathway Reveals the Formation of a Novel Recalcitrant Steroid Compound by a Side Reaction in Sphingobium sp. Strain Chol11
Previous Article in Special Issue
RTG Signaling Sustains Mitochondrial Respiratory Capacity in HOG1-Dependent Osmoadaptation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cdk8 Kinase Module: A Mediator of Life and Death Decisions in Times of Stress

by
Brittany Friedson
and
Katrina F. Cooper
*
Department of Molecular Biology, Graduate School of Biomedical Sciences, Rowan University, Stratford, NJ 08084, USA
*
Author to whom correspondence should be addressed.
Microorganisms 2021, 9(10), 2152; https://doi.org/10.3390/microorganisms9102152
Submission received: 4 September 2021 / Revised: 6 October 2021 / Accepted: 8 October 2021 / Published: 15 October 2021

Abstract

:
The Cdk8 kinase module (CKM) of the multi-subunit mediator complex plays an essential role in cell fate decisions in response to different environmental cues. In the budding yeast S. cerevisiae, the CKM consists of four conserved subunits (cyclin C and its cognate cyclin-dependent kinase Cdk8, Med13, and Med12) and predominantly negatively regulates a subset of stress responsive genes (SRG’s). Derepression of these SRG’s is accomplished by disassociating the CKM from the mediator, thus allowing RNA polymerase II-directed transcription. In response to cell death stimuli, cyclin C translocates to the mitochondria where it induces mitochondrial hyper-fission and promotes regulated cell death (RCD). The nuclear release of cyclin C requires Med13 destruction by the ubiquitin-proteasome system (UPS). In contrast, to protect the cell from RCD following SRG induction induced by nutrient deprivation, cyclin C is rapidly destroyed by the UPS before it reaches the cytoplasm. This enables a survival response by two mechanisms: increased ATP production by retaining reticular mitochondrial morphology and relieving CKM-mediated repression on autophagy genes. Intriguingly, nitrogen starvation also stimulates Med13 destruction but through a different mechanism. Rather than destruction via the UPS, Med13 proteolysis occurs in the vacuole (yeast lysosome) via a newly identified Snx4-assisted autophagy pathway. Taken together, these findings reveal that the CKM regulates cell fate decisions by both transcriptional and non-transcriptional mechanisms, placing it at a convergence point between cell death and cell survival pathways.

1. Introduction

Several age-related maladies, including cancers and neurodegenerative diseases, are associated with deteriorating cellular functions. This can result in the loss of cellular homeostasis and decreased ability to elicit correct cell fate responses to unfavorable environmental cues [1]. For instance, high reactive oxygen species (ROS) levels induce protein and organelle damage, stimulating regulated cell death (RCD) pathways in both yeast and mammals [2,3,4,5]. RCD induction involves transcriptional changes and disruption of mitochondrial integrity. In contrast, pro-survival responses are initiated by different cellular cues including nutrient depletion [6]. Under these conditions, the cell induces a different transcriptional response to help recycle cellular components to refill substrate pools and maintain mitochondrial integrity to maximize energy resources. Thus, successfully interpreting a specific stimulus requires the seamless integration of transcriptional remodeling with organelle function.
The molecular details of how cells execute cell death and cell survival pathways are reasonably well understood. Significantly less clear is how cells decide which path to follow in response to different stressors. Our work and those of others in budding yeast and mammalian cells alike have provided evidence that the Cdk8 kinase module (CKM) of the Mediator complex plays a critical role in these decisions [7,8,9,10,11,12,13,14,15,16,17]. The CKM is a dissociable part of the Mediator complex and contains four subunits- Med12, Med13, Cdk8, and cyclin C. In yeast, CKM association with the Mediator predominantly represses transcription [11,18,19], although positive roles have been described [20,21,22,23]. At the mechanistic level, how the CKM executes these different roles is unclear, but likely dependent on specific promoter contexts [24,25]. Deciphering the contribution of the CKM to diverse biological outputs is important as the dysfunction of any of its members is linked to a multitude of human disease, including cancer [26]. Moreover, in mammalian cells, the kinase module can also encompass the paralogs CDK19, MED12L, and MED13L [9]. Genetic variations in these paralogues are also linked to an overlapping spectrum of disorders [27]. Interestingly, no paralogue of cyclin C has been identified. Consistent with this, cyclin C forms a distinct Mediator kinase module with CDK19, which regulates a different transcriptional program to cyclin C-CDK8 [28,29]. In this review, we discuss these transcriptional roles of the CKM. Thereafter, we discuss the significance of cyclin C’s secondary cytoplasmic role as an envoy of nuclear decisions following cell death or survival cues.

2. The CKM Regulates Cell Fate Decisions

In addition to its transcriptional role, the CKM regulates cell fate decisions by modulating the sub-cellular address of cyclin C (Figure 1). Several types of cellular damage induce cyclin C nuclear release and association with the mitochondria [30,31,32,33,34,35,36,37,38,39,40,41,42,43]. This association triggers extensive mitochondrial fission (hyper-fission) and promotes cell death initiation. Underlying the significance of both roles, these “day and night jobs” of cyclin C are conserved from yeast to mammalian cells, although there are some differences that are discussed below and summarized in Table 1 [13,41]. These studies are also consistent with the emerging theme that proteins can have two different functions, coined “day and night jobs”, which are summoned by different external or intrinsic stimuli [44,45]. Although the transcriptional response to nutritional deprivation overlaps with exposure to cytotoxic compounds, the fate of cyclin C is very different. Although still destroyed to allow SRG derepression following nitrogen starvation, cyclin C does not make it to the cytoplasm, as observed with oxidative stress. Rather, it is destroyed via the UPS prior to its nuclear release [37]. Why is this important? Under starvation conditions, the cell wants maximum ATP production, which requires reticular mitochondria. Inducing fragmentation would be counterproductive. Second, mitochondrial localization of cyclin C pushes the cell toward RCD [37]. This is best illustrated by ectopically targeting cyclin C to the mitochondria [41]. Although mitochondrial fission is induced, cell viability is normal. However, with cyclin C at the mitochondria, nitrogen starvation now triggers the cell death pathway, not a survival response [37]. These findings indicate that cellular outputs to a given signal can be switched simply by altering the location of a single protein. Below, we summarize the molecular details gleaned to date on how the CKM controls cell fate decisions, discussing the impact of cyclin C’s transcriptional and post-transcriptional roles (day and night jobs) on this process.

3. The Transcriptional Role of the CKM in Response to Stress

3.1. Association of the CKM with the Core Mediator

The core Mediator complex is a large multi-subunit coactivator complex organized into three functional modules coined Head, Middle, and Tail. Its modular architecture and subunit composition are conserved [46], which allows it to communicate signals originating from a plethora of transcription factors (TFs) to facilitate RNA pol II-directed transcription. This is aided by large intrinsically disordered domains (IDRs) found within its subunits, providing the Mediator with a high degree of structural flexibility [47,48,49]. A combination of in vitro and in vivo approaches (reviewed in [50]) has uncovered mechanistic details into how the CKM interacts with the Mediator (Figure 2A). One model for transcriptional repression proposes CKM binding to the Mediator precludes the interaction of the RNA pol II at gene promotors [9,51,52]. Recent structural studies have suggested that CKM mediated phosphorylation of other core Mediator subunits releases this steric inhibition [53]. CKM release from the Mediator allows pre-initiation complex (PIC) assembly and is a prerequisite for transcription initiation [7,52,54,55,56,57]. In mammalian systems, the CKM is recruited to promoters that are being actively transcribed [14,58]. Several mechanisms have been proposed for its positive role in transcription, including modifying other regulators [59,60]. These studies provide mechanistic insights into how CKM activity stimulates transcription of genes in stress and signal response networks [13,14,61]. Understanding CKM regulation is important, as its dysregulation is implicated in over 100 different human cancers [62,63,64].
In yeast, the CKM predominately negatively regulates a subset of SRG’s, including genes encoding antioxidants, chaperones and proteins needed for autophagy [8,18,30,33,65]. Consistent with this, CKM members were initially identified as suppressors of aberrantly expressed SRG’s in separate genetic screens [8,17,66,67]. Cyclin C and Cdk8 were also isolated in a screen looking for suppressors of truncation of the yeast C terminal tail of RNA pol II [68]. Unfortunately, this has led to significant nomenclature confusion, with CKM members having multiple gene names. To circumnavigate this, in 2004, the Mediator nomenclature unification effort [69] reclassified CKM members to the following names: cyclin C (CNC1- old names: SSN8, UME3, SRB11, RYE2, and GIG3), Cdk8 (CDK8- old names: SSN3, UME5, SRB10, RYE5, and GIG2), Med12 (MED12- old names: SSN5, GIG1, and NUT6) and Med13 (MED13- old names: SSN2, UME2, SRB9, RYE3, and NUT8).

3.2. Cdk8 Is a Non-Canonical Cyclin-Dependent Kinase

Cyclin-dependent kinases (CDKs) are serine/threonine protein kinases that fall broadly into two functional groups as regulators of cell cycle progression or transcription [70]. CDK activation requires cyclin binding with the interaction stabilized by CDK activating kinase (CAK) phosphorylation in the CDK T-loop [71]. However, CDK8 is distinct among the CDK family as it lacks a canonical phosphorylation residue within its T-loop [72]. Instead, biochemical experiments suggest that human MED12 is essential for stabilizing the CDK8–cyclin C association within the module [73]. Importantly, known oncogenic MED12 mutations [74,75] disrupt the ability of MED12 to activate, but not bind, cyclin C-CDK8. These results are consistent with the model that MED12 binding is necessary, but not sufficient, for efficient CDK8 activation. Recent structural studies of the yeast CKM have provided significant insights into this proposed model, revealing that Med12 stabilizes Cdk8 in an activated conformation (reproduced with permission in Figure 2B). This triggers a structural rearrangement that contributes to the stabilization of the Cdk8 T-loop, thereby enabling kinase activity in the absence of canonical T-loop phosphorylation [76]. In the absence of Med12, the Cdk8 T-loop is flexible, and unfavorably poised for substrate binding and phosphorylation. Consistent with this, both Med13 and Med12 are essential for the transcription of Wingless target genes in Drosophila [77].

3.3. Dissociation of the CKM from the Core Mediator

The mammalian CKM both negatively and positively regulates SRG’s in approximately equal numbers [13,78]. Our recent studies of promoter occupancy in mammalian cells revealed that CKM dissociation from promoter regions coincides with derepression of negatively regulated mRNAs [14]. These studies, as well as those by others, showed that chemical inhibition of CDK8 and CDK19 kinases removes their ability to repress the Mediator complex at enhancers [79]. Conversely, genes that require the CKM for transcription induction exhibited increased promoter occupancy. The molecular mechanisms underlying the reversible CKM-Mediator association are incompletely understood but center on Med13, the largest subunit in Mediator, which anchors the CKM to the hook domain formed by Middle module subunits [7,55,80]. Studies in both humans and yeast indicate that Med13 phosphorylation, ubiquitylation, and turnover all control CKM-Mediator association [36,40,81]. Very recent cryo-EM and mass spectrometry (XL-MS) studies [76] have provided a 3D structural representation of the yeast CKM, detailing the interaction of multiple lobes (Figure 2B). The kinase-lobe (composed of cyclin C and Cdk8) and the H-lobe (composed of Med12) protrude from a central-lobe containing Med13 and Med12. Consistent with previous studies, cyclin C joins Cdk8 in the kinase lobe and interacts with the scaffold Med12 [7,76,82]. These findings significantly improve our understanding of how the CKM regulates transcription, providing insights into why the CKM is implicated in a growing number of cancers and developmental diseases [29,64,79,83,84]. Unexpectedly, they also reveal that yeast Med13 has an Argonaute-like bi-lobal architecture, although the functional significance of this finding remains unknown [76].

3.4. Disassembly of Cyclin C from Med13

As the translocation of cyclin C into the cytoplasm following oxidative stress sets cells on a path to RCD, it is vital to understand how cyclin C is released from its CKM anchor, Med13. Genetic manipulations of cyclin C localization in yeast revealed that cyclin C contains a conserved holoenzyme association domain (HAD) located at the N terminus of cyclin C. This domain consists of a short alpha-helical region containing a conserved “KERQK” sequence required for Med13 binding [85]. HAD domain deletion causes cyclin C dissociation from the CKM and reduction of transcriptional repressor ability. Moreover, similar to the med13∆ phenotype, cyclin CHAD mutants colocalize with mitochondria causing fragmentation in unstressed cultures [40]. Taken together, this supports the model that cyclin C disassembly from the CKM plays a role in its localization to the mitochondria [34]. To demonstrate that the cyclin C HAD-MED13 interaction is conserved, Jezek et al. [38] designed a cell-penetrating peptide mimetic (S-HAD) that effectively disrupts Med13-cyclin C interaction. When S-HAD was added to unstressed mouse embryonic fibroblasts (MEFs), it rapidly induced cyclin C nuclear release and mitochondrial fragmentation. The S-HAD-induced mitochondrial fragmentation is dependent on cyclin C, as this response was not observed in Ccnc null (Ccnc−/−) MEFs [38]. These findings indicate that the interaction between Med13 and cyclin C is conserved, and that cyclin C nuclear release is sufficient to induce mitochondrial fragmentation in the absence of additional stress signals.

3.5. Genes Regulated by the CKM in Response to Cell Death Cues

The transcriptome is significantly remodeled to adapt to unfavorable environmental cues [86,87,88,89]. This requires the activation of transcription factors, many of which are conserved, including CKM, FOXO, HSF1, and NRF2 [88,90,91,92,93,94]. For example, oxidative stress stimulates transcription factor activity upregulating many genes encoding antioxidants (catalases) and pro-survival chaperones [95] required to reinstate cellular homeostasis. Failure to neutralize the toxic effects of reactive oxygen triggers cells to switch to upregulating genes required for cell death pathways [2,96]. Mechanistic details on how cells make this molecular switch remain unclear, but the CKM plays a key role. Relief of repression by the CKM is mediated by CKM removal from the core Mediator. Furthermore, quantitative PCR analysis revealed that cyclin C nuclear release in yeast following oxidative stress is required for full mRNA accumulation of DDR2 and CTT1 [32], genes that encode a multi-stress response protein and catalase, respectively. Moreover, a cis-acting cyclin C mutant (A110V) remained nuclear following stress while preserving its normal transcriptional role [32,33]. Lastly, a significant reduction in viability was observed in H2O2-treated cells harboring the A110V mutant as compared with wild type, implicating a role for the cyclin C promoter removal to properly regulate RCD [31].

3.6. Genes Regulated by the CKM in Response to Cell Survival Cues

In response to nitrogen starvation, autophagy (ATG) genes are upregulated to promote survival [97]. This is mediated by various transcriptional regulators including the CKM [37,98,99]. Epistasis analysis revealed that CKM regulation of ATG8, which is required for phagosome formation, occurs within the Rpd3–Sin3–Ume6 histone deacetylase (HDAC) axis [37,100]. The molecular mechanisms controlling this derepression remain unclear. However, the events downstream of CKM dissociation are different from those occurring in oxidative stress. Significantly, following nitrogen starvation, the UPS destroys cyclin C before its detection in the cytoplasm [37,98,101]. The destruction of cyclin C thus contributes to the expression of genes involved in yeast autophagy. While CKM deletion in the absence of stress is not sufficient to induce autophagy, yeast cells lacking cyclin C-Cdk8 demonstrate a growth advantage due to increased ATG8 induction. Likewise, the presence of cyclin C-Cdk8 restricts the growth of cells partially inhibited for TORC1 [37]. These data suggest that loss of cyclin C-Cdk8 following nitrogen starvation stress enhances cell survival by relieving repression of SRG, though more studies are needed to flush out molecular details. The degradation of cyclin C-Cdk8 can thus adjust a cell’s pro-death or pro-survival transcriptional program depending on the environmental trigger. Similarly, the CKM controls cell fate decisions during development of B and T cells as well as during hematopoiesis [102,103].

4. Upstream Signaling Pathways Link Degradation Machines to Stress

4.1. Oxidative Stress Activates the Cell Wall Integrity Pathway

The Cell Wall Integrity (CWI) pathway is the main signaling pathway involved in the regulation of cell wall stress responses [104,105]. A family of cell-surface sensors (Wsc1, Mid2, and Mtl1) communicate the environmental stress signal to a small G protein Rho1, which further activates protein kinase C (Pkc1) [89,106]. Activated Pkc1 triggers the linear mitogen-activated protein kinase (MAPK) cascade (Figure 3). This is comprised of the MAPKKK kinase (Bck1), a pair of redundant MAPKK’s (Mkk1/2), a MAPK (Slt2/Mpk1), and a pseudo-kinase (Ksp1/Mlp1) [89]. Slt2 is a functional homolog of human extracellular signal-regulated kinase 5 (ERK5) that is also activated in response to physical and chemical stresses [107]. Phosphorylation of Slt2 directly activates two transcription factors: Rlm1, which induces transcription of a wide array of cell wall metabolism genes [108], and SBF, a dimeric transcriptional regulator comprised of Swi4 and Swi6, required for upregulation of G1-specific genes [109].

4.2. The CWI MAPK Slt2 Is Required for Cyclin C Nuclear Release in ROS

Slt2 directly phosphorylates both cyclin C and Med13 in response to oxidative stress (Figure 3) [35,36]. The phosphorylation of cyclin C on serine 266 is required for its nuclear release [35], whereas the phosphorylation of Med13 on threonine 835 and 837 triggers its destruction by the multi-subunit E3 ligase SCFGrr1. Genetic analysis also revealed that Med13 destruction requires a priming event mediated by Cdk8 in unstressed cells [36]. Thus, typical of SCF E3 ligases, recognition of the Med13 degron uses two phosphorylation marks, one to prime the degron, and the second for its recognition by ubiquitin ligases [110,111]. Additionally, Med13 destruction requires the direct phosphorylation by Snf1 on a residue adjacent to the Slt2 sites in Med13’s IDR domain [112]. Snf1 is a highly conserved adenosine monophosphate-activated protein kinase (AMPK) that is activated in response to a variety of stresses, including oxidative stress [112,113]. As multiple signaling pathways converge on the CKM, this reveals that cells place several checks on this system. This ensures that releasing cyclin C to the cytoplasm is the correct response to the environmental input.

4.3. Other Factors Required for ROS-Dependent Cyclin C Nuclear Release

The nuclear release of cyclin C is also known to require the Slt2 paralogue Kdx1, the transcription factor Ask10, and the conserved signaling enzyme phospholipase C (Plc1) [33,35,114]. Mechanistically, it remains unknown how these factors contribute to cyclin C nuclear release. This further emphasizes that cyclin C nuclear release is highly regulated.

4.4. Cyclin C Nuclear Release Requires Additional MAPK Pathways in Response to More Stringent ROS

In response to mild oxidative stress (induced by 0.4 mM H2O2), reprograming a single MAPK pathway is sufficient to induce the yeast cyclin C nuclear translocation and eventual destruction. However, increasing the severity of this environmental challenge (1.2 mM H2O2) triggers a more extensive cellular response mediated by cross-pathway signaling. This is accomplished by recruiting Ste11, a MAPKKK traditionally used by the high osmolarity glycerol (HOG) and mating-type signaling pathways [115]. However, under these severe conditions, Ste11 activates Bck1, the MAPKKK of the CWI pathway [116,117]. The HOG and CWI pathways traditionally were thought to be separate [118,119], but now, similar to other signal transduction pathways, exhibit considerable crosstalk in response to different signals [120]. This ability to discriminate stress intensity by remodeling signaling pathway architecture provides the cell a mechanism to speed up the timing of the response as well as providing backup strategies to ensure that the stress signal is effectively transmitted to the nucleus.

4.5. Nitrogen Starvation Inhibits TORC1

The conserved target of rapamycin kinase complex 1 (TORC1) signaling pathway couples environmental and nutritional cues to downstream effectors. TORC1 phosphorylates a wide range of targets that drive protein, lipid, and nucleotide synthesis [121] and is often dysregulated in human diseases, including cancer, type 2 diabetes, and neurodegeneration [122,123]. Its core, ring-like structure is highly conserved, consisting of a heterotrimeric complex that harbors a TOR serine/threonine protein kinase (TOR1 or TOR2 in yeast) and two regulatory proteins called Kog1 and Lst8 [124]. A fourth protein called Tco89 associates with this core complex to adapt its function to species-specific requirements [125].
Nitrogen and amino acid signals are transmitted to TORC1 by the heterodimeric Rag guanosine triphosphatases (GTPases). In yeast, these are encoded by Gtr1 and Gtr2 [126]. In the presence of abundant nitrogen/amino acids, Gtr1 and Gtr2 are in their GTP, and GDP bound forms, respectively, and bind tightly to Kog1 and Tco89 [127]. In contrast, when nutrient or energy levels fall, TORC1 is inhibited, causing cells to switch from anabolic to catabolic metabolism, eventually entering quiescence [128]. The binding of another complex called SEACIT to Gtr1 induces its GTPase activity [129]. The resulting GDP-bound Gtr1 consequently causes a conformational change that weakens the interaction between Gtr1/2 and TORC1 and rapidly inhibits TORC1 signaling [130].
In yeast, Rag GTPase is tethered to the vacuole membrane by the EGO complex (EGOC), resulting in constitutive localization of TORC1 to this organelle [127,131]. Additionally, a newly identified second pool of TORC1 that localizes to the pre-vacuolar endosome [132] has led to an updated model by which these spatially separated pools of TORC1 execute distinct functions [133]. Once activated, the vacuolar pool of TORC1 primarily upregulates protein translation by phosphorylating the AGC-kinase family protein Sch9 [134]. In contrast, the endosomal pool predominantly inhibits autophagy by phosphorylating Atg13 and Vps27 respectively [132,135].

4.6. Nitrogen Starvation Activated Slt2 Does Not Directly Phosphorylate Cyclin C

It is unclear how the stress signals are transmitted to cyclin C following TORC1 inhibition. Similar to ROS stress, Slt2 is activated under these conditions [136], but how TORC1 communicates with the CWI MAPK pathway is unclear. Other roles of activated Slt2 under nitrogen depletion include directing G1 arrest and entrance into quiescence. Here, in a series of elegant experiments, Moreno-Torres et al. [137,138] discovered that Slt2 mediates the phosphorylation of the cell cycle-dependent kinase inhibitor (CDKI) Sic1. This targets Thr173, a different phosphorylation signature from those that trigger Sic1 destruction, converting this CDKI into a potent inhibitor of Clb5-CDK-Cks1 complexes [137,138]. Intriguingly, the only MAPK site in cyclin C (Ser266) is not required for its proteolysis following TORC1 inhibition [37]. These findings suggest that Slt2 functions indirectly to target cyclin C for destruction. Alternatively, as Slt2 can also target tyrosine residues [139,140], this mechanism could be used to transmit the signal to cyclin. Nevertheless, these results are consistent with the model that the differential regulation of cyclin C contributes to its final subcellular address, which consequently contributes to cell fate decisions.

4.7. Cyclin C Destruction in Nitrogen Starvation Is Mediated by the UPS

A recent study in our laboratory uncovered that yeast cyclin C is degraded following nitrogen starvation or rapamycin treatment by the UPS before its detection in the cytoplasm [37]. Similar to ROS stress, cyclin C degradation requires both the 20S catalytic core particle and the 19S regulatory particle of the 26S proteasome. Interestingly, cyclin C is ubiquitinated using the same E2 ubiquitin conjugating enzymes, Ubc4 and Ubc5, following either stress [32,37]. In contrast, the E3 ubiquitin-ligating enzyme Not4, which mediates cyclin C degradation following oxidative stress, is not required for its degradation in nitrogen starvation. Moreover, similar to other substrates [141,142,143,144,145], multiple Ub ligases mediate cyclin C proteolysis during nitrogen starvation, as no single E3 ligase fulfils this role [37]. The necessity for the same E2 enzymes, but a different E3 ligase following each stress, emphasizes the precision in which ubiquitin can tag cyclin C to be recognized by the proteasome at a specific time and cellular location. Uncovering the E3’s that mark cyclin C for degradation following nitrogen starvation will provide insight into the mechanisms used to differentially regulate cyclin C and how this contributes to cell survival.

4.8. TORC1 Inhibition Results in Med13 Degradation by Snx4-Assisted Autophagy

As cyclin C nuclear release requires Med13 destruction by the SCF following cell death cues (particularly ROS), we were surprised to find that cells use a different mechanism to destroy Med13 after nitrogen depletion. In these conditions, Med13 degradation occurs by vacuolar proteolysis requiring the autophagy machinery [42]. In this pathway, outlined in Figure 4, Med13 shuttles through the nuclear pore complex (NPC), associating first with the cytoplasmic nucleoporin Gle1, a member of the RNA remodeling complex [146,147,148,149,150]. Next, Med13 is transported to Atg17-initiated phagophores located on the vacuole [151,152], aided by the sorting nexin heterodimer Snx4-Atg20. Lastly, upon fusion of the autophagosome with the vacuole, Med13 proteolysis occurs with Snx4-Atg20, being recycled back to the cytosol. In addition, the degradation of Rim15 and Msn2, transcriptional activators of ATG genes, also occurs by this pathway [42]. As turnover of both positive and negative regulators of ATG transcription is controlled by Snx4-assisted autophagy (SAA), this pathway permits cells to fine-tune the autophagic response. Moreover, the pathway is distinct from known nucleophagy mechanisms [153,154], which remove unspecified nuclear material through blebbing of the nucleus [155].

4.9. The Role of Snx4 in Autophagy

The evolutionarily conserved sorting nexin Snx4 [156,157] is associated with the etiology of neurodegenerative diseases and cancer [158,159]. Despite this, its precise roles in autophagy and stress-induced cell survival remains elusive. Degrading transcription factors had not been a function previously assigned to Snx4-Atg20. This sub-group of sorting nexins all contain a characteristic BAR (Bin/|Amphiphysin/Rvs) domain that binds curved membranes upon dimerization [160,161]. The BAR domain also binds and transports a variety of cargos, including Med13 [42,162]. Snx4-Atg20 is also required for other selective autophagy pathways, including proteophagy, ribophagy, and a recently identified pathway used to degrade specific translating mRNA’s [163,164]. Underlining its importance, Snx4-Atg20 transports cargos from the vacuole to the Golgi, an essential function for autophagy [165,166]. Despite these varied roles, how Snx4-Atg20 recognizes its autophagic cargos is a topic for further investigations.

5. The Relationship between Mitochondria, the CKM, and Cell Fate Decisions

5.1. Control of Mitochondrial Dynamics

Mitochondria are not only sites of ATP generation by oxidative phosphorylation (OXPHOS), but also have additional roles, including calcium homeostasis and RCD regulation. In yeast and mammals, mitochondria form dynamic tubular networks constantly undergoing fission and fusion events. The equilibrium between fission and fusion is controlled by the activity of conserved molecular machines driven by dynamin-like GTPases [167]. Dysfunctional mitochondrial dynamics is associated with a broad range of human diseases, from cancer to neurodegenerative disorders [168]. Mitochondrial fission in mammalian cells is controlled by the highly conserved dynamin-like GTPase called DRP1 [169]. DRP1 is recruited to the outer mitochondrial membrane (OMM) by several adapter proteins, including mitochondrial fission factor (MFF), mitochondrial dynamics protein 49 and 51 (MiD49, MiD51), and mitochondrial fission 1 protein (hFis1) [170]. DRP1 self-assembles into filaments that form rings around the mitochondria. These rings constrict following GTP hydrolysis with the final scission being executed by dynamin 2 (DNM2) [171,172,173]. This concept is conserved in yeast, where a single OMM protein, Fis1, interacts with one of two adaptor molecules (Mdv1 or Caf4) recruiting the GTPase Dnm1 to mitochondria [174,175,176,177,178,179]. Mitochondrial fusion in yeast is initiated through the Fzo1 (outer membrane, OMM) and Mgm1 (inner membrane, IMM) GTPases [180,181].

5.2. Mitochondrial Dynamics in Unstressed Cells

The balance between fragmented and fused mitochondria in unstressed cells is both asynchronous and adapted according to need [182,183,184]. In unstressed cells, mitochondria predominantly maintain reticular morphology, allowing maximum ATP production as well as membrane and mtDNA damage repair [185,186]. Mitochondria fission in unstressed cells (Figure 5) permits equitable inheritance of mitochondria during cell division [187,188]. In addition, fission is associated with the quality control process of mitophagy that degrades damaged mitochondria [189,190]. In mammalian cells, mitochondrial fission is a prerequisite for mitophagy, whereas in yeast, it enhances the process [189,191,192]. As dysfunctional mitochondria can leak significant levels of ROS from the electron transport chain, mitophagy plays a pivotal role in maintaining cellular homeostasis both during normal and stress conditions [193]. Failure to execute these functions is linked to the pathogenesis of many diseases, including Parkinson’s and Alzheimer’s [194].

5.3. Stress-Induced Mitochondrial Fission Is Linked to Cell Death Pathways

Extensive fission and loss of mitochondrial integrity is associated with RCD induction [195]. In mammals, the pro-death molecules BAX or BAK, which disrupts the OMM, are found at fission sites containing DRP1 and the mitofusin MFN2 [196,197,198,199]. The resulting BAX oligomer formation and MOM permeabilization (MOMP) [200] facilitate the release of pro-apoptotic factors, including cytochrome c and apoptosis-inducing factor (AIF), which initiate caspase-dependent or caspase-independent cell death [201,202,203]. Consistent with this model, DRP1 knockdown cells are resistant to RCD [204]. However, others report that fission itself is not necessary for the timely release of some pro-apoptotic proteins [205,206]. Therefore, the connection between the fission machinery and RCD remains to be clarified. As DRP1 is required for fission during both mitosis and RCD induction, the cell must distinguish these activities to correctly respond to either cell growth or cell death stimuli [184].
In yeast, many regulated cell death hallmarks are observed following stress treatment (e.g., H2O2 or acetic acid) [207,208,209]. A critical difference from mammalian RCD pathways is that yeast lack orthologues of BAX and BAK. However, many executioners of RCD are conserved, including cytochrome c, the nucleases Aif1 and Nuc1 (EndoG), and the metacaspase Mca1/Yca1 [210], although its exact role is still unclear. Stress-induced mitochondrial hyper-fission is also associated with the release of these sequestered apoptotic factors and DNM1 deletion protects yeast cells from RCD [204,211]. As the release of pro-apoptotic factors from the mitochondria is a hallmark of RCD, it is generally accepted that yeast can execute RCD pathways following stress [209]. It also suggests an evolutionary linkage of these regulatory pathways [208].

5.4. Night Job of Cyclin C in Cell Death: A Mitochondrial Response

In addition to its transcriptional role, yeast cyclin C performs its stress-activated night job in the cytoplasm. Following H2O2 stress, cyclin C exits the nucleus and binds to the mitochondrial fission machinery, independent of Cdk8 (see Figure 1) [34]. Here it interacts physically and genetically with Mdv1 [34], stimulating Dnm1 GTPase activity [34]. Stress-induced cyclin C nuclear release is dependent upon Med13 destruction by the the SCFGrr1 E3 ligase complex and the UPS [36,40]. Interestingly, only cyclin C is released into the cytoplasm, leaving Cdk8 in the nucleolus [32]. After cyclin C has stimulated hyper-fission, the E3 ligase Not4 targets it for degradation by the UPS [30,32]. Genetic manipulations of cyclin C localization in yeast have supported this model. Here, constitutively placing cyclin C at the mitochondria by either deleting MED13 [40] or by fusing cyclin C to the OMM binding region of Fis1 [37] results in hyper-fission in the absence of stress. Mutants that retain cyclin C in the nucleus (A110V, S266A) also result in significant reduction in mitochondrial fission following ROS stress as compared with wild-type cells [34]. These studies demonstrate that cyclin C localization to the mitochondria is necessary and sufficient for stress-induced mitochondrial hyper-fission.

5.5. Conservation of the Night Job of Cyclin C

The mitochondrial role of cyclin C is highly conserved. In H2O2 stressed mammalian cells, a portion of nuclear cyclin C translocates to the mitochondria where it directly binds to DRP1, increasing its affinity to GTP and stimulating GTPase activity in vitro [41,43]. Consistent with the yeast results, adding recombinant cyclin C to permeabilized Ccnc−/− MEF cells induces complete fission [41]. Cyclin C also plays a role in the efficient mitochondrial localization and activation of BAX [38]. However, even though mitochondrial-associated cyclin C can recruit BAX to the OMM in the absence of stress, these cells do not undergo cell death as BAX fails to oligomerize. These data support a model in which cyclin C association defines an initial step in BAX-OMM recruitment and provides a physical connection between fission and apoptotic factors. The presence of cyclin C at the mitochondria allows the cell to discriminate stress-induced fission from other types of mitochondrial divisions. Consistent with this activity, cyclin C has been identified as a tumor suppressor in both Acute Lymphoblastic Leukemia [78] and thyroid tumors [64]. Moreover, the CCNC locus (6q21) [212] exhibits loss of heterozygosity in several cancers, including osteosarcoma [213]. CDK8 also has been implicated as an oncogene in an array of cancers including colon, breast, and prostate [214,215,216]. The roles of CDK8 defined in earlier studies [29,217,218] have provided the underlying framework for a partial understanding of its role in tumorigenesis.

5.6. Mitochondrial Dynamics in Nutrient Starved Cells

Much less is known about mitochondrial dynamics in response to amino acid or nitrogen starvation. In mammalian cells, nutrient starvation induces a hyper-fused mitochondrial network, capable of increased ATP production that promotes cell survival [192,219,220,221]. Mitochondrial fusion is achieved in part due to PKA-dependent phosphorylation of DRP1, which blocks OMM recruitment [219,220,222]. Elongated mitochondria are also observed in other conditions associated with increased ATP production [223]. Similarly, yeast grown with a nonfermentable carbon source require increased OXPHOS activity that is accompanied by elongated mitochondrial networks [224]. This is also partially mediated through PKA activity [225] and supports the model that high OXPHOS activity correlates with extensive mitochondrial networks [226].

5.7. A Different Night Shift for Cyclin C in Cell Survival: Preserving Mitochondrial Integrity

TORC1 inhibition also triggers cyclin C destruction, but does so prior to its detection in the cytoplasm [37]. This destruction enhances cell survival by allowing full derepression of SRG’s required for autophagy and other processes. In addition, this nuclear destruction precludes cyclin C relocalization to the cytoplasm, thus preventing mitochondrial hyper-fission. The importance of this switch in destruction strategy was tested by fusing cyclin C to an OMM protein in yeast. This fusion protein localizes to mitochondria and can induce mitochondrial fission without additional stress signals. Interestingly, subjecting these cells to nitrogen starvation initiated the cell death pathway, not the autophagic survival program [37]. This strongly suggests that cyclin C’s subcellular address is intricately linked to cell fates. Thus, these data imply that the nuclear degradation of cyclin C following TORC1 inhibition promotes survival by preventing cyclin C mitochondrial association. These findings argue that the systems controlling cell death and cell survival pathways are intimately involved. Interestingly, cell survival and cell death signals also induce different degradative fates for Med13 (Figure 1), being degraded by the UPS and SAA respectively [42]. These findings suggest that two CKM members have secondary cytoplasmic roles to play following different environmental cues. Moreover, our studies have shown that the correct execution of these roles fortifies the necessary molecular response to the incoming stress [31,34,37].

6. Conclusions: The CKM Is at the Crossroads of Different Cell Fates

Analysis of the CKM following stress has revealed that its role in mediating cell fate decisions extends beyond solely regulating SRG transcription. Relocalization of cyclin C to the mitochondria only in response to cell death cues suggests that the CKM is acting as a signaling system, transmiting decisions made in the nucleus to the organelles. This role is amplified during development, where reduced Cdk8 activity is linked to lower levels of cyclin C and could dictate early mammalian embryo decisions [79]. Thus, in stress and developmental scenarios, cyclin C is the CKM’s envoy, linking nuclear decisions to cellular outcomes in response to cell death cues.
This model has led us to hypothesize that a similar signaling role may be assigned to Med13 in response to cell survival cues (Figure 6). The idea is supported by the observation that Med13 contains an exceedingly large (~700 kDa), intrinsically disordered region (IDR). IDR’s serve as flexible platforms for protein–protein interactions, which allow them to have many binding partners [227]. As such, IDR containing proteins often function as hubs for protein interaction networks, playing crucial roles in many biological processes including signaling pathways, transcriptional regulation, translation, and cell cycle control [228,229]. It comes as no surprise then that studies in Drosophila, Zebrafish, Arabidopsis, and mice have revealed that Med13 plays critical roles in embryonic and tumor development [230,231,232,233,234,235]. Moreover, germline variants of MED12 and MED12L also contain large IDR domains, and are found in several genetic disorders associated with X-linked intellectual disability, neuronal and developmental disorders, failure of hematopoietic-specific transcriptional programs, and cancer [84,102,236,237]. Intriguingly, many mutations in MED13L are found in its IDR [238] and result in a haplo-insufficiency syndrome characterized by a spectrum of symptoms including congenital heart and intellectual disorders [27,238,239]. Additionally, in yeast, the IDR associates with many proteins involved with Snx4-assisted autophagy [42]. Taken together, these findings support a model that this region of Med13 acts as an intricate communication hub for signaling proteins, thus allowing the proper cell fate response to environmental stress.
As highlighted in this review, S. cerevisiae is an excellent experimental model for understanding both cell death and cell survival responses with respect to CKM function [95,207,240]. Given the highly conserved nature of many players (see Table 1), its rapid growth rate, and facile gene manipulations, yeast will continue to provide critical insights into the molecular events that dictate life and death decisions. Importantly, this model system has been key to understanding the molecular events associated with selective and non-selective autophagy mechanisms, with 44 ATG genes known to date, most of which are highly conserved with mammalian autophagy pathways [77]. An immerging area for understanding cell fate decisions is the crosstalk between cell death and cell survival pathways. Rather than separate entities, these pathways share common regulators [4] whose relationships may be better dissected in a more tractable system such as yeast.
Table 1. Summary of the known roles of CKM following ROS stress and TORC1 inhibition in yeast and mammals.
Table 1. Summary of the known roles of CKM following ROS stress and TORC1 inhibition in yeast and mammals.
FunctionS. cerevisiaeMammals
Predominantly negatively regulates SRG’s. Positively regulates a few genes.[18,20,23,68,241,242]No
Equally regulates positive and negative SRGs.No[13,14,28,243]
Cyclin C -Cdk8 phosphorylate the C-terminal domain of the largest subunit of RNA pol II in vitro, to inhibit transcriptional initiation. [68,85,242][244]
Cyclin C -Cdk8 directly phosphorylates other Mediator subunit to negatively regulate transcription.[245][246]
Cyclin C -Cdk8 phosphorylates transcription factors and other targets.[247,248][241,246,249,250]
MAPK mediated phosphorylation of Med13 mediates its nuclear release following cell death cues.[35]unknown
Nuclear release of cyclin C requires Med13 degradation by UPS following cell death cues.[36,40]unknown
Med13 is regulated by SCFGrr1/Fbw7.[36][81]
Cytoplasmic cyclin C mediates mitochondrial hyper-fission by binding to mitochondrial fission complex after ROS stress.[34][41]
Cytoplasmic cyclin C promotes oligomerization of the GTPase DNM1/Drp1.unknown[43]
Cyclin C nuclear release is required for RCD (intrinsic pathway). [31,34][41]
Cyclin C binds to BAX at mitochondria to induce MOMP following ROS stress.No BAX[38]
Cyclin C null mutants are deficient in stress-induced mitochondrial fission.[34][41]
Cyclin C nuclear release after ROS stress promotes the release of pro-apoptotic factors.Assumed [34][38]
Cyclin C is destroyed by the UPS in the cytoplasm in ROS stress after mediating mitochondrial hyper-fission.[30,33]No
ROS induced night job of cyclin C is independent of Cdk8.[32][41]
CKM negatively regulates a subset of ATG genes.[37,42][13,14]
Cyclin C is destroyed by the UPS following TORC1 inhibition [37]unknown
Destruction of cyclin C following TORC1 inhibition promotes cell survival.[37]unknown
Destruction of cyclin C following TORC1 inhibition promotes mitochondrial fusion.[37]unknown
Med13 is destroyed by Snx4-assisted autophagy following TORC1 inhibition.[42]unknown

Author Contributions

B.F. wrote the initial draft of this manuscript. K.F.C. edited and revised the manuscript and drew the figures. All authors have read and agreed to the published version of the manuscript.

Funding

National Institutes of Health grant (GM113196) and an Osteopathic Heritage Foundation NJ, award to K.F.C.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Nicoletta Guaragnella, Tiago Outeiro and Anita Krisko for their invitation to write this review. We also thank Kuang-Lei Tsai (University of Texas Health Science Center at Houston) and Thomas G. Boyer (University of Texas Health Science Center at San Antonio) for permission to reproduce the crystal structure of the CKM (Figure 2B). We especially express gratitude to Randy Strich and members of the Cooper and Strich laboratories for critically reading this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ÅAngstrom
AGC-kinasegroup of kinases related to PKA PKG and PKC
AMPKadenosine monophosphate-activated protein kinase
ATPadenosine triphosphate
CAKCDK-activating kinase
CCcyclin C
CDKCyclin-dependent kinase
CDKIcyclin-dependent protein inhibitor
Cryo-EMCryogenic electron microscopy
EGOCGtr1, Gtr2, Ego1, and Ego3 complex that can activate TORC1
ERK5extracellular signal-regulated kinase 5
GTPguanosine triphosphate
HDACRpd3–Sin3–Ume6 histone deacetylase
HOGhigh osmolarity glycerol
IDRintrinsic disordered region
IMMinner mitochondrial membrane
MAPKmitogen-activated protein kinase
MAPKmitogen-activated protein kinase kinase
MAPKKKmitogen-activated protein kinase kinase kinase
MEFsmouse embryonic fibroblasts
MFFmitochondrial fission factor
MOMPmitochondrial outer membrane permeabilization
mtDNAmitochondrial DNA
NPCnuclear pore complex
OMMouter mitochondrial membrane
OXPHOSoxidative phosphorylation
PASpre-autophagosomal structure
PICpre-initiation complex
Pkc1protein kinase C
PKAprotein kinase A
RNA Pol IIRNA Polymerase II
ROSreactive oxygen species
SAASnx4-assisted autophagy
SCFSkp1-Cullin-F-box E3 ligase complex
SRG’sstress response genes
SEACIT Sea1–Npr2–Npr3 complex that can inhibit TORC1
Serserine
TORC1target of rapamycin kinase complex 1
Thrthreonine
TFTranscription factor
UASupstream activating sequences
UPSubiquitin proteasome system
YFPyellow fluorescence protein

References

  1. Towers, C.G.; Wodetzki, D.; Thorburn, A. Autophagy and cancer: Modulation of cell death pathways and cancer cell adaptations. J. Cell Biol. 2020, 219, e201909033. [Google Scholar] [CrossRef] [PubMed]
  2. Madeo, F.; Frohlich, E.; Ligr, M.; Grey, M.; Sigrist, S.J.; Wolf, D.H.; Frohlich, K.U. Oxygen stress: A regulator of apoptosis in yeast. J. Cell Biol. 1999, 145, 757–767. [Google Scholar] [CrossRef]
  3. Galluzzi, L.; Baehrecke, E.H.; Ballabio, A.; Boya, P.; Bravo-San Pedro, J.M.; Cecconi, F.; Choi, A.M.; Chu, C.T.; Codogno, P.; Colombo, M.I.; et al. Molecular definitions of autophagy and related processes. EMBO J. 2017, 36, 1811–1836. [Google Scholar] [CrossRef]
  4. Cooper, K.F. Till Death Do Us Part: The Marriage of Autophagy and Apoptosis. Oxid. Med. Cell. Longev. 2018, 2018, 4701275. [Google Scholar] [CrossRef] [Green Version]
  5. Carmona-Gutierrez, D.; Bauer, M.A.; Zimmermann, A.; Aguilera, A.; Austriaco, N.; Ayscough, K.; Balzan, R.; Bar-Nun, S.; Barrientos, A.; Belenky, P.; et al. Guidelines and recommendations on yeast cell death nomenclature. Microb. Cell 2018, 5, 4–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Levine, B.; Kroemer, G. Autophagy in the pathogenesis of disease. Cell 2008, 132, 27–42. [Google Scholar] [CrossRef] [Green Version]
  7. Tsai, K.L.; Sato, S.; Tomomori-Sato, C.; Conaway, R.C.; Conaway, J.W.; Asturias, F.J. A conserved Mediator-CDK8 kinase module association regulates Mediator-RNA polymerase II interaction. Nat. Struct. Mol. Biol. 2013, 20, 611–619. [Google Scholar] [CrossRef] [Green Version]
  8. Kuchin, S.; Yeghiayan, P.; Carlson, M. Cyclin-dependent protein kinase and cyclin homologs SSN3 and SSN8 contribute to transcriptional control in yeast. Proc. Natl. Acad. Sci. USA 1995, 92, 4006–4010. [Google Scholar] [CrossRef] [Green Version]
  9. Jeronimo, C.; Robert, F. The Mediator Complex: At the Nexus of RNA Polymerase II Transcription. Trends Cell Biol. 2017, 27, 765–783. [Google Scholar] [CrossRef] [PubMed]
  10. Robinson, P.J.; Trnka, M.J.; Pellarin, R.; Greenberg, C.H.; Bushnell, D.A.; Davis, R.; Burlingame, A.L.; Sali, A.; Kornberg, R.D. Molecular architecture of the yeast Mediator complex. Elife 2015, 4, e08719. [Google Scholar] [CrossRef]
  11. van de Peppel, J.; Kettelarij, N.; van Bakel, H.; Kockelkorn, T.T.; van Leenen, D.; Holstege, F.C. Mediator expression profiling epistasis reveals a signal transduction pathway with antagonistic submodules and highly specific downstream targets. Mol. Cell 2005, 19, 511–522. [Google Scholar] [CrossRef]
  12. Law, M.J.; Ciccaglione, K. Fine-Tuning of Histone H3 Lys4 Methylation During Pseudohyphal Differentiation by the CDK Submodule of RNA Polymerase II. Genetics 2015, 199, 435–453. [Google Scholar] [CrossRef] [Green Version]
  13. Stieg, D.C.; Chang, K.T.; Cooper, K.F.; Strich, R. Cyclin C Regulated Oxidative Stress Responsive Transcriptome in Mus musculus Embryonic Fibroblasts. G3 Genes Genomes Genet. 2019, 9, 1901–1908. [Google Scholar] [CrossRef] [Green Version]
  14. Stieg, D.C.; Cooper, K.F.; Strich, R. The extent of cyclin C promoter occupancy directs changes in stress-dependent transcription. J. Biol. Chem. 2020, 295, 16280–16291. [Google Scholar] [CrossRef]
  15. Allen, B.L.; Taatjes, D.J. The Mediator complex: A central integrator of transcription. Nat. Rev. Mol. Cell Biol. 2015, 16, 155–166. [Google Scholar] [CrossRef] [PubMed]
  16. Yarrington, R.M.; Yu, Y.; Yan, C.; Bai, L.; Stillman, D.J. A Role for Mediator Core in Limiting Coactivator Recruitment in Saccharomyces cerevisiae. Genetics 2020, 215, 407–420. [Google Scholar] [CrossRef] [Green Version]
  17. Strich, R.; Slater, M.R.; Esposito, R.E. Identification of negative regulatory genes that govern the expression of early meiotic genes in yeast. Proc. Natl. Acad. Sci. USA 1989, 86, 10018–10022. [Google Scholar] [CrossRef] [Green Version]
  18. Holstege, F.C.; Jennings, E.G.; Wyrick, J.J.; Lee, T.I.; Hengartner, C.J.; Green, M.R.; Golub, T.R.; Lander, E.S.; Young, R.A. Dissecting the regulatory circuitry of a eukaryotic genome. Cell 1998, 95, 717–728. [Google Scholar] [CrossRef] [Green Version]
  19. Ansari, S.A.; Ganapathi, M.; Benschop, J.J.; Holstege, F.C.; Wade, J.T.; Morse, R.H. Distinct role of Mediator tail module in regulation of SAGA-dependent, TATA-containing genes in yeast. EMBO J. 2012, 31, 44–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Hirst, K.; Fisher, F.; McAndrew, P.C.; Godding, C.R. The transcription factor, the Cdk, its cyclin and their regulator: Directing the transcriptional response to a nutritional signal. EMBO 1994, 13, 5410–5420. [Google Scholar] [CrossRef]
  21. Hirst, M.; Kobor, M.S.; Kuriakose, N.; Greenblatt, J.; Sadowski, I. GAL4 is regulated by the RNA polymerase II holoenzyme-associated cyclin-dependent protein kinase SRB10/CDK8. Mol. Cell 1999, 3, 673–678. [Google Scholar] [CrossRef]
  22. Vincent, O.; Kuchin, S.; Hong, S.P.; Townley, R.; Vyas, V.K.; Carlson, M. Interaction of the Srb10 kinase with Sip4, a transcriptional activator of gluconeogenic genes in Saccharomyces cerevisiae. Mol. Cell. Biol. 2001, 21, 5790–5796. [Google Scholar] [CrossRef] [Green Version]
  23. Larschan, E.; Winston, F. The Saccharomyces cerevisiae Srb8-Srb11 complex functions with the SAGA complex during Gal4-activated transcription. Mol. Cell. Biol. 2005, 25, 114–123. [Google Scholar] [CrossRef] [Green Version]
  24. Xu, W.; Ji, J.Y. Dysregulation of CDK8 and Cyclin C in tumorigenesis. J Genet Genom. 2011, 38, 439–452. [Google Scholar] [CrossRef]
  25. Nemet, J.; Jelicic, B.; Rubelj, I.; Sopta, M. The two faces of Cdk8, a positive/negative regulator of transcription. Biochimie 2014, 97, 22–27. [Google Scholar] [CrossRef]
  26. Berk, A.J. Yin and yang of mediator function revealed by human mutants. Proc. Natl. Acad. Sci. USA 2012, 109, 19519–19520. [Google Scholar] [CrossRef] [Green Version]
  27. Adegbola, A.; Musante, L.; Callewaert, B.; Maciel, P.; Hu, H.; Isidor, B.; Picker-Minh, S.; Le Caignec, C.; Delle Chiaie, B.; Vanakker, O.; et al. Redefining the MED13L syndrome. Eur. J. Hum. Genet. 2015, 23, 1308–1317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Galbraith, M.D.; Allen, M.A.; Bensard, C.L.; Wang, X.; Schwinn, M.K.; Qin, B.; Long, H.W.; Daniels, D.L.; Hahn, W.C.; Dowell, R.D.; et al. HIF1A employs CDK8-mediator to stimulate RNAPII elongation in response to hypoxia. Cell 2013, 153, 1327–1339. [Google Scholar] [CrossRef] [Green Version]
  29. Dannappel, M.V.; Sooraj, D.; Loh, J.J.; Firestein, R. Molecular and in vivo Functions of the CDK8 and CDK19 Kinase Modules. Front. Cell Dev. Biol. 2019, 6, 171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Cooper, K.F.; Mallory, M.J.; Smith, J.B.; Strich, R. Stress and developmental regulation of the yeast C-type cyclin Ume3p (Srb11p/Ssn8p). EMBO J. 1997, 16, 4665–4675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Krasley, E.; Cooper, K.F.; Mallory, M.J.; Dunbrack, R.; Strich, R. Regulation of the oxidative stress response through Slt2p-dependent destruction of cyclin C in Saccharomyces cerevisiae. Genetics 2006, 172, 1477–1486. [Google Scholar] [CrossRef] [Green Version]
  32. Cooper, K.F.; Scarnati, M.S.; Krasley, E.; Mallory, M.J.; Jin, C.; Law, M.J.; Strich, R. Oxidative-stress-induced nuclear to cytoplasmic relocalization is required for Not4-dependent cyclin C destruction. J. Cell Sci. 2012, 125, 1015–1026. [Google Scholar] [CrossRef] [Green Version]
  33. Cooper, K.F.; Mallory, M.J.; Strich, R. Oxidative stress-induced destruction of the yeast C-type cyclin Ume3p requires phosphatidylinositol-specific phospholipase C and the 26S proteasome. Mol. Cell. Biol. 1999, 19, 3338–3348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Cooper, K.F.; Khakhina, S.; Kim, S.K.; Strich, R. Stress-induced nuclear-to-cytoplasmic translocation of cyclin C promotes mitochondrial fission in yeast. Dev. Cell 2014, 28, 161–173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Jin, C.; Strich, R.; Cooper, K.F. Slt2p phosphorylation induces cyclin C nuclear-to-cytoplasmic translocation in response to oxidative stress. Mol. Biol. Cell 2014, 25, 1396–1407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Stieg, D.C.; Willis, S.D.; Ganesan, V.; Ong, K.L.; Scuorzo, J.; Song, M.; Grose, J.; Strich, R.; Cooper, K.F. A complex molecular switch directs stress-induced cyclin C nuclear release through SCF(Grr1)-mediated degradation of Med13. Mol. Biol. Cell 2018, 29, 363–375. [Google Scholar] [CrossRef]
  37. Willis, S.D.; Hanley, S.E.; Beishke, T.; Tati, P.D.; Cooper, K.F. Ubiquitin-proteasome-mediated cyclin C degradation promotes cell survival following nitrogen starvation. Mol. Biol. Cell 2020, 31, 1015–1031. [Google Scholar] [CrossRef]
  38. Jezek, J.; Chang, K.T.; Joshi, A.M.; Strich, R. Mitochondrial translocation of cyclin C stimulates intrinsic apoptosis through Bax recruitment. EMBO Rep. 2019, 20, e47425. [Google Scholar] [CrossRef]
  39. Jezek, J.; Cooper, K.F.; Strich, R. Reactive Oxygen Species and Mitochondrial Dynamics: The Yin and Yang of Mitochondrial Dysfunction and Cancer Progression. Antioxidants 2018, 7, 13. [Google Scholar] [CrossRef]
  40. Khakhina, S.; Cooper, K.F.; Strich, R. Med13p prevents mitochondrial fission and programmed cell death in yeast through nuclear retention of cyclin C. Mol. Biol. Cell 2014, 25, 2807–2816. [Google Scholar] [CrossRef]
  41. Wang, K.; Yan, R.; Cooper, K.F.; Strich, R. Cyclin C mediates stress-induced mitochondrial fission and apoptosis. Mol. Biol. Cell 2015, 26, 1030–1043. [Google Scholar] [CrossRef] [PubMed]
  42. Hanley, S.E.; Willis, S.D.; Cooper, K.F. Snx4-assisted vacuolar targeting of transcription factors defines a new autophagy pathway for controlling ATG expression. Autophagy 2021. [Google Scholar] [CrossRef] [PubMed]
  43. Ganesan, V.; Willis, S.D.; Chang, K.T.; Beluch, S.; Cooper, K.F.; Strich, R. Cyclin C directly stimulates Drp1 GTP affinity to mediate stress-induced mitochondrial hyperfission. Mol. Biol. Cell 2019, 30, 302–311. [Google Scholar] [CrossRef] [PubMed]
  44. Gross, A.; Katz, S.G. Non-apoptotic functions of BCL-2 family proteins. Cell Death Differ. 2017, 24, 1348–1358. [Google Scholar] [CrossRef]
  45. Shamas-Din, A.; Kale, J.; Leber, B.; Andrews, D.W. Mechanisms of action of Bcl-2 family proteins. Cold Spring Harb. Perspect. Biol. 2013, 5, a008714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Bourbon, H.M. Comparative genomics supports a deep evolutionary origin for the large, four-module transcriptional mediator complex. Nucleic Acids Res. 2008, 36, 3993–4008. [Google Scholar] [CrossRef] [Green Version]
  47. Nagulapalli, M.; Maji, S.; Dwivedi, N.; Dahiya, P.; Thakur, J.K. Evolution of disorder in Mediator complex and its functional relevance. Nucleic Acids Res. 2016, 44, 1591–1612. [Google Scholar] [CrossRef] [Green Version]
  48. Malik, S.; Guermah, M.; Yuan, C.X.; Wu, W.; Yamamura, S.; Roeder, R.G. Structural and functional organization of TRAP220, the TRAP/mediator subunit that is targeted by nuclear receptors. Mol. Cell. Biol. 2004, 24, 8244–8254. [Google Scholar] [CrossRef] [Green Version]
  49. Taatjes, D.J.; Tjian, R. Structure and function of CRSP/Med2; a promoter-selective transcriptional coactivator complex. Mol. Cell 2004, 14, 675–683. [Google Scholar] [CrossRef]
  50. Soutourina, J. Transcription regulation by the Mediator complex. Nat. Rev. Mol. Cell Biol. 2018, 19, 262–274. [Google Scholar] [CrossRef]
  51. Jeronimo, C.; Langelier, M.F.; Bataille, A.R.; Pascal, J.M.; Pugh, B.F.; Robert, F. Tail and Kinase Modules Differently Regulate Core Mediator Recruitment and Function In Vivo. Mol. Cell 2016, 64, 455–466. [Google Scholar] [CrossRef] [Green Version]
  52. Petrenko, N.; Jin, Y.; Wong, K.H.; Struhl, K. Mediator Undergoes a Compositional Change during Transcriptional Activation. Mol. Cell 2016, 64, 443–454. [Google Scholar] [CrossRef] [Green Version]
  53. Osman, S.; Mohammad, E.; Lidschreiber, M.; Stuetzer, A.; Bazso, F.L.; Maier, K.C.; Urlaub, H.; Cramer, P. The Cdk8 kinase module regulates interaction of the mediator complex with RNA polymerase II. J. Biol. Chem. 2021, 296, 100734. [Google Scholar] [CrossRef]
  54. Fan, X.; Chou, D.M.; Struhl, K. Activator-specific recruitment of Mediator in vivo. Nat. Struct. Mol. Biol. 2006, 13, 117–120. [Google Scholar] [CrossRef] [PubMed]
  55. Knuesel, M.T.; Meyer, K.D.; Donner, A.J.; Espinosa, J.M.; Taatjes, D.J. The human CDK8 subcomplex is a histone kinase that requires Med12 for activity and can function independently of mediator. Mol. Cell. Biol. 2009, 29, 650–661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Elmlund, H.; Baraznenok, V.; Lindahl, M.; Samuelsen, C.O.; Koeck, P.J.; Holmberg, S.; Hebert, H.; Gustafsson, C.M. The cyclin-dependent kinase 8 module sterically blocks Mediator interactions with RNA polymerase II. Proc. Natl. Acad. Sci. USA 2006, 103, 15788–15793. [Google Scholar] [CrossRef] [Green Version]
  57. Miller, C.; Matic, I.; Maier, K.C.; Schwalb, B.; Roether, S.; Strasser, K.; Tresch, A.; Mann, M.; Cramer, P. Mediator phosphorylation prevents stress response transcription during non-stress conditions. J. Biol. Chem. 2012, 287, 44017–44026. [Google Scholar] [CrossRef] [Green Version]
  58. Donner, A.J.; Szostek, S.; Hoover, J.M.; Espinosa, J.M. CDK8 is a stimulus-specific positive coregulator of p53 target genes. Mol. Cell 2007, 27, 121–133. [Google Scholar] [CrossRef] [Green Version]
  59. Bancerek, J.; Poss, Z.C.; Steinparzer, I.; Sedlyarov, V.; Pfaffenwimmer, T.; Mikulic, I.; Dolken, L.; Strobl, B.; Muller, M.; Taatjes, D.J.; et al. CDK8 kinase phosphorylates transcription factor STAT1 to selectively regulate the interferon response. Immunity 2013, 38, 250–262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Firestein, R.; Bass, A.J.; Kim, S.Y.; Dunn, I.F.; Silver, S.J.; Guney, I.; Freed, E.; Ligon, A.H.; Vena, N.; Ogino, S.; et al. CDK8 is a colorectal cancer oncogene that regulates beta-catenin activity. Nature 2008, 455, 547–551. [Google Scholar] [CrossRef]
  61. Chen, M.; Liang, J.; Ji, H.; Yang, Z.; Altilia, S.; Hu, B.; Schronce, A.; McDermott, M.S.J.; Schools, G.P.; Lim, C.U.; et al. CDK8/19 Mediator kinases potentiate induction of transcription by NFkappaB. Proc. Natl. Acad. Sci. USA 2017, 114, 10208–10213. [Google Scholar] [CrossRef] [Green Version]
  62. Philip, S.; Kumarasiri, M.; Teo, T.; Yu, M.; Wang, S. Cyclin-Dependent Kinase 8: A New Hope in Targeted Cancer Therapy? J. Med. Chem. 2018, 61, 5073–5092. [Google Scholar] [CrossRef]
  63. Clark, A.D.; Oldenbroek, M.; Boyer, T.G. Mediator kinase module and human tumorigenesis. Crit. Rev. Biochem. Mol. Biol. 2015, 50, 393–426. [Google Scholar] [CrossRef]
  64. Jezek, J.; Wang, K.; Yan, R.; Di Cristofano, A.; Cooper, K.F.; Strich, R. Synergistic repression of thyroid hyperplasia by cyclin C and Pten. J. Cell Sci. 2019, 132, jcs230029. [Google Scholar] [CrossRef] [Green Version]
  65. Wahi, M.; Johnson, A.D. Identification of genes required for alpha 2 repression in Saccharomyces cerevisiae. Genetics 1995, 140, 79–90. [Google Scholar] [CrossRef] [PubMed]
  66. Balciunas, D.; Ronne, H. Three subunits of the RNA polymerase II mediator complex are involved in glucose repression. Nucleic Acids Res. 1995, 23, 4421–4425. [Google Scholar] [CrossRef]
  67. Tabtiang, R.K.; Herskowitz, I. Nuclear proteins Nut1p and Nut2p cooperate to negatively regulate a Swi4p-dependent lacZ reporter gene in Saccharomyces cerevisiae. Mol. Cell. Biol. 1998, 18, 4707–4718. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Liao, S.M.; Zhang, J.; Jeffery, D.A.; Koleske, A.J.; Thompson, C.M.; Chao, D.M.; Viljoen, M.; van Vuuren, H.J.; Young, R.A. A kinase-cyclin pair in the RNA polymerase II holoenzyme. Nature 1995, 374, 193–196. [Google Scholar] [CrossRef] [PubMed]
  69. Bourbon, H.M.; Aguilera, A.; Ansari, A.Z.; Asturias, F.J.; Berk, A.J.; Bjorklund, S.; Blackwell, T.K.; Borggrefe, T.; Carey, M.; Carlson, M.; et al. A Unified Nomenclature for Protein Subunits of Mediator Complexes Linking Transcriptional Regulators to RNA Polymerase II. Mol. Cell 2004, 14, 553–557. [Google Scholar] [CrossRef] [PubMed]
  70. Malumbres, M. Cyclin-dependent kinases. Genome Biol. 2014, 15, 122. [Google Scholar] [CrossRef] [Green Version]
  71. Echalier, A.; Endicott, J.A.; Noble, M.E. Recent developments in cyclin-dependent kinase biochemical and structural studies. Biochim. Biophys. Acta 2010, 1804, 511–519. [Google Scholar] [CrossRef] [PubMed]
  72. Schneider, E.V.; Bottcher, J.; Blaesse, M.; Neumann, L.; Huber, R.; Maskos, K. The structure of CDK8/CycC implicates specificity in the CDK/cyclin family and reveals interaction with a deep pocket binder. J. Mol. Biol. 2011, 412, 251–266. [Google Scholar] [CrossRef]
  73. Knuesel, M.T.; Meyer, K.D.; Bernecky, C.; Taatjes, D.J. The human CDK8 subcomplex is a molecular switch that controls Mediator coactivator function. Genes Dev. 2009, 23, 439–451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Turunen, M.; Spaeth, J.M.; Keskitalo, S.; Park, M.J.; Kivioja, T.; Clark, A.D.; Makinen, N.; Gao, F.; Palin, K.; Nurkkala, H.; et al. Uterine leiomyoma-linked MED12 mutations disrupt mediator-associated CDK activity. Cell Rep. 2014, 7, 654–660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Park, M.J.; Shen, H.; Spaeth, J.M.; Tolvanen, J.H.; Failor, C.; Knudtson, J.F.; McLaughlin, J.; Halder, S.K.; Yang, Q.; Bulun, S.E.; et al. Oncogenic exon 2 mutations in Mediator subunit MED12 disrupt allosteric activation of cyclin C-CDK8/19. J. Biol. Chem. 2018, 293, 4870–4882. [Google Scholar] [CrossRef] [Green Version]
  76. Li, Y.C.; Chao, T.C.; Kim, H.J.; Cholko, T.; Chen, S.F.; Li, G.; Snyder, L.; Nakanishi, K.; Chang, C.E.; Murakami, K.; et al. Structure and noncanonical Cdk8 activation mechanism within an Argonaute-containing Mediator kinase module. Sci. Adv. 2021, 7. [Google Scholar] [CrossRef]
  77. Klionsky, D.J.; Abdel-Aziz, A.K.; Abdelfatah, S.; Abdellatif, M.; Abdoli, A.; Abel, S.; Abeliovich, H.; Abildgaard, M.H.; Abudu, Y.P.; Acevedo-Arozena, A.; et al. Guidelines for the use and interpretation of assays for monitoring autophagy (4th edition)(1). Autophagy 2021, 17, 1–382. [Google Scholar] [CrossRef]
  78. Li, N.; Fassl, A.; Chick, J.; Inuzuka, H.; Li, X.; Mansour, M.R.; Liu, L.; Wang, H.; King, B.; Shaik, S.; et al. Cyclin C is a haploinsufficient tumour suppressor. Nat. Cell Biol. 2014, 16, 1080–1091. [Google Scholar] [CrossRef] [Green Version]
  79. Lynch, C.J.; Bernad, R.; Martinez-Val, A.; Shahbazi, M.N.; Nobrega-Pereira, S.; Calvo, I.; Blanco-Aparicio, C.; Tarantino, C.; Garreta, E.; Richart-Gines, L.; et al. Global hyperactivation of enhancers stabilizes human and mouse naive pluripotency through inhibition of CDK8/19 Mediator kinases. Nat. Cell Biol. 2020, 22, 1223–1238. [Google Scholar] [CrossRef] [PubMed]
  80. Harper, T.M.; Taatjes, D.J. The complex structure and function of Mediator. J. Biol. Chem. 2018, 293, 13778–13785. [Google Scholar] [CrossRef] [Green Version]
  81. Davis, M.A.; Larimore, E.A.; Fissel, B.M.; Swanger, J.; Taatjes, D.J.; Clurman, B.E. The SCF-Fbw7 ubiquitin ligase degrades MED13 and MED13L and regulates CDK8 module association with Mediator. Genes Dev. 2013, 27, 151–156. [Google Scholar] [CrossRef] [Green Version]
  82. Klatt, F.; Leitner, A.; Kim, I.V.; Ho-Xuan, H.; Schneider, E.V.; Langhammer, F.; Weinmann, R.; Muller, M.R.; Huber, R.; Meister, G.; et al. A precisely positioned MED12 activation helix stimulates CDK8 kinase activity. Proc. Natl. Acad. Sci. USA 2020, 117, 2894–2905. [Google Scholar] [CrossRef] [PubMed]
  83. Jezek, J.; Smethurst, D.G.J.; Stieg, D.C.; Kiss, Z.A.C.; Hanley, S.E.; Ganesan, V.; Chang, K.T.; Cooper, K.F.; Strich, R. Cyclin C: The Story of a Non-Cycling Cyclin. Biology 2019, 8, 3. [Google Scholar] [CrossRef] [Green Version]
  84. Nizon, M.; Laugel, V.; Flanigan, K.M.; Pastore, M.; Waldrop, M.A.; Rosenfeld, J.A.; Marom, R.; Xiao, R.; Gerard, A.; Pichon, O.; et al. Variants in MED12L, encoding a subunit of the mediator kinase module, are responsible for intellectual disability associated with transcriptional defect. Genet. Med. 2019, 21, 2713–2722. [Google Scholar] [CrossRef]
  85. Cooper, K.F.; Strich, R. Functional analysis of the Ume3p/ Srb11p-RNA polymerase II holoenzyme interaction. Gene Expr. 1999, 8, 43–57. [Google Scholar]
  86. Moye-Rowley, W.S. Transcription factors regulating the response to oxidative stress in yeast. Antioxid. Redox Signal. 2002, 4, 123–140. [Google Scholar] [CrossRef]
  87. Taymaz-Nikerel, H.; Cankorur-Cetinkaya, A.; Kirdar, B. Genome-Wide Transcriptional Response of Saccharomyces cerevisiae to Stress-Induced Perturbations. Front. Bioeng. Biotechnol. 2016, 4, 17. [Google Scholar] [CrossRef] [Green Version]
  88. Gomez-Pastor, R.; Burchfiel, E.T.; Thiele, D.J. Regulation of heat shock transcription factors and their roles in physiology and disease. Nat. Rev. Mol. Cell Biol. 2018, 19, 4–19. [Google Scholar] [CrossRef]
  89. Levin, D.E. Regulation of cell wall biogenesis in Saccharomyces cerevisiae: The cell wall integrity signaling pathway. Genetics 2011, 189, 1145–1175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Vihervaara, A.; Duarte, F.M.; Lis, J.T. Molecular mechanisms driving transcriptional stress responses. Nat. Rev. Genet. 2018, 19, 385–397. [Google Scholar] [CrossRef] [PubMed]
  91. Nataro, J.P.; Corcoran, L.; Zirin, S.; Swink, S.; Taichman, N.; Goin, J.; Harris, M.C. Prospective analysis of coagulase-negative staphylococcal infection in hospitalized infants. J. Pediatr. 1994, 125, 798–804. [Google Scholar] [CrossRef]
  92. Vomund, S.; Schafer, A.; Parnham, M.J.; Brune, B.; von Knethen, A. Nrf2, the Master Regulator of Anti-Oxidative Responses. Int. J. Mol. Sci. 2017, 18, 2772. [Google Scholar] [CrossRef] [Green Version]
  93. Postnikoff, S.D.; Malo, M.E.; Wong, B.; Harkness, T.A. The yeast forkhead transcription factors fkh1 and fkh2 regulate lifespan and stress response together with the anaphase-promoting complex. PLoS Genet. 2012, 8, e1002583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Fuse, Y.; Kobayashi, M. Conservation of the Keap1-Nrf2 System: An Evolutionary Journey through Stressful Space and Time. Molecules 2017, 22, 436. [Google Scholar] [CrossRef]
  95. Perrone, G.G.; Tan, S.X.; Dawes, I.W. Reactive oxygen species and yeast apoptosis. Biochim. Biophys. Acta 2008, 1783, 1354–1368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Guaragnella, N.; Zdralevic, M.; Antonacci, L.; Passarella, S.; Marra, E.; Giannattasio, S. The role of mitochondria in yeast programmed cell death. Front. Oncol. 2012, 2, 70. [Google Scholar] [CrossRef] [Green Version]
  97. Delorme-Axford, E.; Klionsky, D.J. Transcriptional and post-transcriptional regulation of autophagy in the yeast Saccharomyces cerevisiae. J. Biol. Chem. 2018, 293, 5396–5403. [Google Scholar] [CrossRef] [Green Version]
  98. Bartholomew, C.R.; Suzuki, T.; Du, Z.; Backues, S.K.; Jin, M.; Lynch-Day, M.A.; Umekawa, M.; Kamath, A.; Zhao, M.; Xie, Z.; et al. Ume6 transcription factor is part of a signaling cascade that regulates autophagy. Proc. Natl. Acad. Sci. USA 2012, 109, 11206–11210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Jin, M.; He, D.; Backues, S.K.; Freeberg, M.A.; Liu, X.; Kim, J.K.; Klionsky, D.J. Transcriptional regulation by Pho23 modulates the frequency of autophagosome formation. Curr. Biol. 2014, 24, 1314–1322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Kadosh, D.; Struhl, K. Targeted recruitment of the Sin3-Rpd3 histone deacetylase complex generates a highly localized domain of repressed chromatin in vivo. Mol. Cell. Biol. 1998, 18, 5121–5127. [Google Scholar] [CrossRef] [Green Version]
  101. Vlahakis, A.; Lopez Muniozguren, N.; Powers, T. Stress-response transcription factors Msn2 and Msn4 couple TORC2-Ypk1 signaling and mitochondrial respiration to ATG8 gene expression and autophagy. Autophagy 2017, 13, 1804–1812. [Google Scholar] [CrossRef] [PubMed]
  102. Aranda-Orgilles, B.; Saldana-Meyer, R.; Wang, E.; Trompouki, E.; Fassl, A.; Lau, S.; Mullenders, J.; Rocha, P.P.; Raviram, R.; Guillamot, M.; et al. MED12 Regulates HSC-Specific Enhancers Independently of Mediator Kinase Activity to Control Hematopoiesis. Cell Stem Cell 2016, 19, 784–799. [Google Scholar] [CrossRef] [Green Version]
  103. El Khattabi, L.; Zhao, H.; Kalchschmidt, J.; Young, N.; Jung, S.; Van Blerkom, P.; Kieffer-Kwon, P.; Kieffer-Kwon, K.R.; Park, S.; Wang, X.; et al. A Pliable Mediator Acts as a Functional Rather Than an Architectural Bridge between Promoters and Enhancers. Cell 2019, 178, 1145–1158 e1120. [Google Scholar] [CrossRef] [PubMed]
  104. Kamada, Y.; Jung, U.S.; Piotrowski, J.; Levin, D.E. The protein kinase C-activated MAP kinase pathway of Saccharomyces cerevisiae mediates a novel aspect of the heat shock response. Genes Dev. 1995, 9, 1559–1571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Vilella, F.; Herrero, E.; Torres, J.; de la Torre-Ruiz, M.A. Pkc1 and the upstream elements of the cell integrity pathway in Saccharomyces cerevisiae, Rom2 and Mtl1, are required for cellular responses to oxidative stress. J. Biol. Chem. 2005, 280, 9149–9159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Jin, C.; Parshin, A.V.; Daly, I.; Strich, R.; Cooper, K.F. The cell wall sensors Mtl1, Wsc1, and Mid2 are required for stress-induced nuclear to cytoplasmic translocation of cyclin C and programmed cell death in yeast. Oxid. Med. Cell. Longev. 2013, 2013, 320823. [Google Scholar] [CrossRef] [Green Version]
  107. Truman, A.W.; Millson, S.H.; Nuttall, J.M.; King, V.; Mollapour, M.; Prodromou, C.; Pearl, L.H.; Piper, P.W. Expressed in the yeast Saccharomyces cerevisiae, human ERK5 is a client of the Hsp90 chaperone that complements loss of the Slt2p (Mpk1p) cell integrity stress-activated protein kinase. Eukaryot. Cell 2006, 5, 1914–1924. [Google Scholar] [CrossRef] [Green Version]
  108. Garcia, R.; Rodriguez-Pena, J.M.; Bermejo, C.; Nombela, C.; Arroyo, J. The high osmotic response and cell wall integrity pathways cooperate to regulate transcriptional responses to zymolyase-induced cell wall stress in Saccharomyces cerevisiae. J. Biol. Chem. 2009, 284, 10901–10911. [Google Scholar] [CrossRef] [Green Version]
  109. Andrews, B.J.; Moore, L.A. Interaction of the yeast Swi4 and Swi6 cell cycle regulatory proteins in vitro. Proc. Natl. Acad. Sci. USA 1992, 89, 11852–11856. [Google Scholar] [CrossRef] [Green Version]
  110. Nash, P.; Tang, X.; Orlicky, S.; Chen, Q.; Gertler, F.B.; Mendenhall, M.D.; Sicheri, F.; Pawson, T.; Tyers, M. Multisite phosphorylation of a CDK inhibitor sets a threshold for the onset of DNA replication. Nature 2001, 414, 514–521. [Google Scholar] [CrossRef]
  111. Ang, X.L.; Wade Harper, J. SCF-mediated protein degradation and cell cycle control. Oncogene 2005, 24, 2860–2870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Willis, S.D.; Stieg, D.C.; Ong, K.L.; Shah, R.; Strich, A.K.; Grose, J.H.; Cooper, K.F. Snf1 cooperates with the CWI MAPK pathway to mediate the degradation of Med13 following oxidative stress. Microb. Cell 2018, 5, 357–370. [Google Scholar] [CrossRef] [Green Version]
  113. Hong, S.P.; Carlson, M. Regulation of snf1 protein kinase in response to environmental stress. J. Biol. Chem. 2007, 282, 16838–16845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Cohen, T.J.; Lee, K.; Rutkowski, L.H.; Strich, R. Ask10p mediates the oxidative stress-induced destruction of the Saccharomyces cerevisiae C-type cyclin Ume3p/Srb11p. Eukaryot. Cell 2003, 2, 962–970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Garcia, R.; Pulido, V.; Orellana-Munoz, S.; Nombela, C.; Vazquez de Aldana, C.R.; Rodriguez-Pena, J.M.; Arroyo, J. Signalling through the yeast MAPK Cell Wall Integrity pathway controls P-body assembly upon cell wall stress. Sci. Rep. 2019, 9, 3186. [Google Scholar] [CrossRef] [PubMed]
  116. Jin, C.; Kim, S.K.; Willis, S.D.; Cooper, K.F. The MAPKKKs Ste11 and Bck1 jointly transduce the high oxidative stress signal through the cell wall integrity MAP kinase pathway. Microb. Cell 2015, 2, 329–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Moye-Rowley, W.S. Feelin’ it: Differential oxidative stress sensing mediated by Cyclin C. Microb. Cell 2015, 2, 305–307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Junttila, M.R.; Li, S.P.; Westermarck, J. Phosphatase-mediated crosstalk between MAPK signaling pathways in the regulation of cell survival. FASEB J. 2008, 22, 954–965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Saito, H. Regulation of cross-talk in yeast MAPK signaling pathways. Curr. Opin. Microbiol. 2010, 13, 677–683. [Google Scholar] [CrossRef]
  120. Van Drogen, F.; Dard, N.; Pelet, S.; Lee, S.S.; Mishra, R.; Srejic, N.; Peter, M. Crosstalk and spatiotemporal regulation between stress-induced MAP kinase pathways and pheromone signaling in budding yeast. Cell Cycle 2020, 19, 1707–1715. [Google Scholar] [CrossRef]
  121. Gonzalez, S.; Rallis, C. The TOR Signaling Pathway in Spatial and Temporal Control of Cell Size and Growth. Front. Cell Dev. Biol. 2017, 5, 61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Albert, V.; Hall, M.N. mTOR signaling in cellular and organismal energetics. Curr. Opin. Cell Biol. 2015, 33, 55–66. [Google Scholar] [CrossRef]
  123. Sullivan, A.; Wallace, R.L.; Wellington, R.; Luo, X.; Capaldi, A.P. Multilayered regulation of TORC1-body formation in budding yeast. Mol. Biol. Cell 2019, 30, 400–410. [Google Scholar] [CrossRef] [PubMed]
  124. Wullschleger, S.; Loewith, R.; Hall, M.N. TOR signaling in growth and metabolism. Cell 2006, 124, 471–484. [Google Scholar] [CrossRef] [Green Version]
  125. Reinke, A.; Anderson, S.; McCaffery, J.M.; Yates, J., 3rd; Aronova, S.; Chu, S.; Fairclough, S.; Iverson, C.; Wedaman, K.P.; Powers, T. TOR complex 1 includes a novel component, Tco89p (YPL180w), and cooperates with Ssd1p to maintain cellular integrity in Saccharomyces cerevisiae. J. Biol. Chem. 2004, 279, 14752–14762. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Gao, M.; Kaiser, C.A. A conserved GTPase-containing complex is required for intracellular sorting of the general amino-acid permease in yeast. Nat. Cell Biol. 2006, 8, 657–667. [Google Scholar] [CrossRef]
  127. Binda, M.; Peli-Gulli, M.P.; Bonfils, G.; Panchaud, N.; Urban, J.; Sturgill, T.W.; Loewith, R.; De Virgilio, C. The Vam6 GEF controls TORC1 by activating the EGO complex. Mol. Cell 2009, 35, 563–573. [Google Scholar] [CrossRef] [Green Version]
  128. Saxton, R.A.; Sabatini, D.M. mTOR Signaling in Growth, Metabolism, and Disease. Cell 2017, 168, 960–976. [Google Scholar] [CrossRef] [Green Version]
  129. Kim, J.S.; Ro, S.H.; Kim, M.; Park, H.W.; Semple, I.A.; Park, H.; Cho, U.S.; Wang, W.; Guan, K.L.; Karin, M.; et al. Sestrin2 inhibits mTORC1 through modulation of GATOR complexes. Sci. Rep. 2015, 5, 9502. [Google Scholar] [CrossRef] [PubMed]
  130. Hughes Hallett, J.E.; Luo, X.; Capaldi, A.P. Snf1/AMPK promotes the formation of Kog1/Raptor-bodies to increase the activation threshold of TORC1 in budding yeast. Elife 2015, 4, e09181. [Google Scholar] [CrossRef]
  131. Sancak, Y.; Peterson, T.R.; Shaul, Y.D.; Lindquist, R.A.; Thoreen, C.C.; Bar-Peled, L.; Sabatini, D.M. The Rag GTPases bind raptor and mediate amino acid signaling to mTORC1. Science 2008, 320, 1496–1501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Hatakeyama, R.; Peli-Gulli, M.P.; Hu, Z.; Jaquenoud, M.; Garcia Osuna, G.M.; Sardu, A.; Dengjel, J.; De Virgilio, C. Spatially Distinct Pools of TORC1 Balance Protein Homeostasis. Mol. Cell 2019, 73, 325–338 e328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Lahiri, V.; Klionsky, D.J. Spatially distinct pools of TORC1 balance protein homeostasis. Autophagy 2019, 15, 561–564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Urban, J.; Soulard, A.; Huber, A.; Lippman, S.; Mukhopadhyay, D.; Deloche, O.; Wanke, V.; Anrather, D.; Ammerer, G.; Riezman, H.; et al. Sch9 is a major target of TORC1 in Saccharomyces cerevisiae. Mol. Cell 2007, 26, 663–674. [Google Scholar] [CrossRef]
  135. Kamada, Y.; Yoshino, K.; Kondo, C.; Kawamata, T.; Oshiro, N.; Yonezawa, K.; Ohsumi, Y. Tor directly controls the Atg1 kinase complex to regulate autophagy. Mol. Cell. Biol. 2010, 30, 1049–1058. [Google Scholar] [CrossRef] [Green Version]
  136. Torres, J.; Di Como, C.J.; Herrero, E.; De La Torre-Ruiz, M.A. Regulation of the cell integrity pathway by rapamycin-sensitive TOR function in budding yeast. J. Biol. Chem. 2002, 277, 43495–43504. [Google Scholar] [CrossRef] [Green Version]
  137. Moreno-Torres, M.; Jaquenoud, M.; De Virgilio, C. TORC1 controls G1-S cell cycle transition in yeast via Mpk1 and the greatwall kinase pathway. Nat. Commun. 2015, 6, 8256. [Google Scholar] [CrossRef] [Green Version]
  138. Moreno-Torres, M.; Jaquenoud, M.; Peli-Gulli, M.P.; Nicastro, R.; De Virgilio, C. TORC1 coordinates the conversion of Sic1 from a target to an inhibitor of cyclin-CDK-Cks1. Cell Discov. 2017, 3, 17012. [Google Scholar] [CrossRef]
  139. Gonzalez-Rubio, G.; Sellers-Moya, A.; Martin, H.; Molina, M. Differential Role of Threonine and Tyrosine Phosphorylation in the Activation and Activity of the Yeast MAPK Slt2. Int. J. Mol. Sci. 2021, 22, 1110. [Google Scholar] [CrossRef]
  140. Yurko, N.; Liu, X.; Yamazaki, T.; Hoque, M.; Tian, B.; Manley, J.L. MPK1/SLT2 Links Multiple Stress Responses with Gene Expression in Budding Yeast by Phosphorylating Tyr1 of the RNAP II CTD. Mol. Cell 2017, 68, 913–925 e913. [Google Scholar] [CrossRef] [Green Version]
  141. Zhu, J.; Deng, S.; Lu, P.; Bu, W.; Li, T.; Yu, L.; Xie, Z. The Ccl1-Kin28 kinase complex regulates autophagy under nitrogen starvation. J. Cell Sci. 2016, 129, 135–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Finley, D.; Ulrich, H.D.; Sommer, T.; Kaiser, P. The ubiquitin-proteasome system of Saccharomyces cerevisiae. Genetics 2012, 192, 319–360. [Google Scholar] [CrossRef] [Green Version]
  143. Shi, Y.; Chen, X.; Elsasser, S.; Stocks, B.B.; Tian, G.; Lee, B.H.; Shi, Y.; Zhang, N.; de Poot, S.A.; Tuebing, F.; et al. Rpn1 provides adjacent receptor sites for substrate binding and deubiquitination by the proteasome. Science 2016, 351. [Google Scholar] [CrossRef] [Green Version]
  144. Zhang, D.; Chen, T.; Ziv, I.; Rosenzweig, R.; Matiuhin, Y.; Bronner, V.; Glickman, M.H.; Fushman, D. Together, Rpn10 and Dsk2 can serve as a polyubiquitin chain-length sensor. Mol. Cell 2009, 36, 1018–1033. [Google Scholar] [CrossRef] [Green Version]
  145. Saeki, Y. Ubiquitin recognition by the proteasome. J. Biochem. 2017, 161, 113–124. [Google Scholar] [CrossRef] [PubMed]
  146. Noble, K.N.; Tran, E.J.; Alcazar-Roman, A.R.; Hodge, C.A.; Cole, C.N.; Wente, S.R. The Dbp5 cycle at the nuclear pore complex during mRNA export II: Nucleotide cycling and mRNP remodeling by Dbp5 are controlled by Nup159 and Gle1. Genes Dev. 2011, 25, 1065–1077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Tran, E.J.; Zhou, Y.; Corbett, A.H.; Wente, S.R. The DEAD-box protein Dbp5 controls mRNA export by triggering specific RNA:protein remodeling events. Mol. Cell 2007, 28, 850–859. [Google Scholar] [CrossRef]
  148. Aryanpur, P.P.; Regan, C.A.; Collins, J.M.; Mittelmeier, T.M.; Renner, D.M.; Vergara, A.M.; Brown, N.P.; Bolger, T.A. Gle1 Regulates RNA Binding of the DEAD-Box Helicase Ded1 in Its Complex Role in Translation Initiation. Mol. Cell. Biol. 2017, 37. [Google Scholar] [CrossRef] [Green Version]
  149. Alcazar-Roman, A.R.; Tran, E.J.; Guo, S.; Wente, S.R. Inositol hexakisphosphate and Gle1 activate the DEAD-box protein Dbp5 for nuclear mRNA export. Nat. Cell Biol. 2006, 8, 711–716. [Google Scholar] [CrossRef]
  150. Weirich, C.S.; Erzberger, J.P.; Flick, J.S.; Berger, J.M.; Thorner, J.; Weis, K. Activation of the DExD/H-box protein Dbp5 by the nuclear-pore protein Gle1 and its coactivator InsP6 is required for mRNA export. Nat. Cell Biol. 2006, 8, 668–676. [Google Scholar] [CrossRef]
  151. Mao, K.; Chew, L.H.; Inoue-Aono, Y.; Cheong, H.; Nair, U.; Popelka, H.; Yip, C.K.; Klionsky, D.J. Atg29 phosphorylation regulates coordination of the Atg17-Atg31-Atg29 complex with the Atg11 scaffold during autophagy initiation. Proc. Natl. Acad. Sci. USA 2013, 110, E2875–E2884. [Google Scholar] [CrossRef] [Green Version]
  152. Hollenstein, D.M.; Gomez-Sanchez, R.; Ciftci, A.; Kriegenburg, F.; Mari, M.; Torggler, R.; Licheva, M.; Reggiori, F.; Kraft, C. Vac8 spatially confines autophagosome formation at the vacuole in S. cerevisiae. J. Cell Sci. 2019, 132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Millen, J.I.; Krick, R.; Prick, T.; Thumm, M.; Goldfarb, D.S. Measuring piecemeal microautophagy of the nucleus in Saccharomyces cerevisiae. Autophagy 2009, 5, 75–81. [Google Scholar] [CrossRef] [Green Version]
  154. Mochida, K.; Oikawa, Y.; Kimura, Y.; Kirisako, H.; Hirano, H.; Ohsumi, Y.; Nakatogawa, H. Receptor-mediated selective autophagy degrades the endoplasmic reticulum and the nucleus. Nature 2015, 522, 359–362. [Google Scholar] [CrossRef]
  155. Papandreou, M.E.; Tavernarakis, N. Nucleophagy: From homeostasis to disease. Cell Death Differ. 2019, 26, 630–639. [Google Scholar] [CrossRef] [Green Version]
  156. Hanley, S.E.; Cooper, K.F. Sorting Nexins in Protein Homeostasis. Cells 2020, 10, 17. [Google Scholar] [CrossRef]
  157. Haft, C.R.; de la Luz Sierra, M.; Barr, V.A.; Haft, D.H.; Taylor, S.I. Identification of a family of sorting nexin molecules and characterization of their association with receptors. Mol. Cell. Biol. 1998, 18, 7278–7287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Zhang, S.; Hu, B.; You, Y.; Yang, Z.; Liu, L.; Tang, H.; Bao, W.; Guan, Y.; Shen, X. Sorting nexin 10 acts as a tumor suppressor in tumorigenesis and progression of colorectal cancer through regulating chaperone mediated autophagy degradation of p21(Cip1/WAF1). Cancer Lett. 2018, 419, 116–127. [Google Scholar] [CrossRef]
  159. Hu, Y.B.; Dammer, E.B.; Ren, R.J.; Wang, G. The endosomal-lysosomal system: From acidification and cargo sorting to neurodegeneration. Transl. Neurodegener. 2015, 4, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Stanishneva-Konovalova, T.B.; Derkacheva, N.I.; Polevova, S.V.; Sokolova, O.S. The Role of BAR Domain Proteins in the Regulation of Membrane Dynamics. Acta Nat. 2016, 8, 60–69. [Google Scholar] [CrossRef]
  161. Popelka, H.; Damasio, A.; Hinshaw, J.E.; Klionsky, D.J.; Ragusa, M.J. Structure and function of yeast Atg20, a sorting nexin that facilitates autophagy induction. Proc. Natl. Acad. Sci. USA 2017, 114, E10112–E10121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Bean, B.D.; Davey, M.; Conibear, E. Cargo selectivity of yeast sorting nexins. Traffic 2017, 18, 110–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Nemec, A.A.; Howell, L.A.; Peterson, A.K.; Murray, M.A.; Tomko, R.J., Jr. Autophagic clearance of proteasomes in yeast requires the conserved sorting nexin Snx4. J. Biol. Chem. 2017, 292, 21466–21480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Makino, S.; Kawamata, T.; Iwasaki, S.; Ohsumi, Y. Selectivity of mRNA degradation by autophagy in yeast. Nat. Commun. 2021, 12, 2316. [Google Scholar] [CrossRef]
  165. Ma, M.; Burd, C.G.; Chi, R.J. Distinct complexes of yeast Snx4 family SNX-BARs mediate retrograde trafficking of Snc1 and Atg27. Traffic 2017, 18, 134–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Ma, M.; Burd, C.G. Retrograde trafficking and plasma membrane recycling pathways of the budding yeast Saccharomyces cerevisiae. Traffic 2020, 21, 45–59. [Google Scholar] [CrossRef]
  167. Westermann, B. Mitochondrial fusion and fission in cell life and death. Nat. Rev. Mol. Cell Biol. 2010, 11, 872–884. [Google Scholar] [CrossRef]
  168. Chan, D.C. Mitochondrial Dynamics and Its Involvement in Disease. Annu. Rev. Pathol. 2020, 15, 235–259. [Google Scholar] [CrossRef] [Green Version]
  169. Youle, R.J.; van der Bliek, A.M. Mitochondrial fission, fusion, and stress. Science 2012, 337, 1062–1065. [Google Scholar] [CrossRef] [Green Version]
  170. Atkins, K.; Dasgupta, A.; Chen, K.H.; Mewburn, J.; Archer, S.L. The role of Drp1 adaptor proteins MiD49 and MiD51 in mitochondrial fission: Implications for human disease. Clin. Sci. 2016, 130, 1861–1874. [Google Scholar] [CrossRef] [Green Version]
  171. Frohlich, C.; Grabiger, S.; Schwefel, D.; Faelber, K.; Rosenbaum, E.; Mears, J.; Rocks, O.; Daumke, O. Structural insights into oligomerization and mitochondrial remodelling of dynamin 1-like protein. EMBO J. 2013, 32, 1280–1292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Otsuga, D.; Keegan, B.R.; Brisch, E.; Thatcher, J.W.; Hermann, G.J.; Bleazard, W.; Shaw, J.M. The dynamin-related GTPase, Dnm1p, controls mitochondrial morphology in yeast. J. Cell Biol. 1998, 143, 333–349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Bleazard, W.; McCaffery, J.M.; King, E.J.; Bale, S.; Mozdy, A.; Tieu, Q.; Nunnari, J.; Shaw, J.M. The dynamin-related GTPase Dnm1 regulates mitochondrial fission in yeast. Nat. Cell Biol. 1999, 1, 298–304. [Google Scholar] [CrossRef] [PubMed]
  174. Mozdy, A.D.; McCaffery, J.M.; Shaw, J.M. Dnm1p GTPase-mediated mitochondrial fission is a multi-step process requiring the novel integral membrane component Fis1p. J. Cell Biol. 2000, 151, 367–380. [Google Scholar] [CrossRef]
  175. Tieu, Q.; Nunnari, J. Mdv1p is a WD repeat protein that interacts with the dynamin-related GTPase, Dnm1p, to trigger mitochondrial division. J. Cell Biol. 2000, 151, 353–366. [Google Scholar] [CrossRef]
  176. Cerveny, K.L.; Jensen, R.E. The WD-repeats of Net2p interact with Dnm1p and Fis1p to regulate division of mitochondria. Mol. Biol. Cell 2003, 14, 4126–4139. [Google Scholar] [CrossRef] [PubMed]
  177. Motley, A.M.; Ward, G.P.; Hettema, E.H. Dnm1p-dependent peroxisome fission requires Caf4p, Mdv1p and Fis1p. J. Cell Sci. 2008, 121, 1633–1640. [Google Scholar] [CrossRef] [Green Version]
  178. Naylor, K.; Ingerman, E.; Okreglak, V.; Marino, M.; Hinshaw, J.E.; Nunnari, J. Mdv1 interacts with assembled dnm1 to promote mitochondrial division. J. Biol. Chem. 2006, 281, 2177–2183. [Google Scholar] [CrossRef] [Green Version]
  179. Bhar, D.; Karren, M.A.; Babst, M.; Shaw, J.M. Dimeric Dnm1-G385D interacts with Mdv1 on mitochondria and can be stimulated to assemble into fission complexes containing Mdv1 and Fis1. J. Biol. Chem. 2006, 281, 17312–17320. [Google Scholar] [CrossRef] [Green Version]
  180. Anton, F.; Fres, J.M.; Schauss, A.; Pinson, B.; Praefcke, G.J.; Langer, T.; Escobar-Henriques, M. Ugo1 and Mdm30 act sequentially during Fzo1-mediated mitochondrial outer membrane fusion. J. Cell Sci. 2011, 124, 1126–1135. [Google Scholar] [CrossRef] [Green Version]
  181. Sesaki, H.; Jensen, R.E. UGO1 encodes an outer membrane protein required for mitochondrial fusion. J. Cell Biol. 2001, 152, 1123–1134. [Google Scholar] [CrossRef]
  182. Hoppins, S.; Lackner, L.; Nunnari, J. The machines that divide and fuse mitochondria. Annu. Rev. Biochem. 2007, 76, 751–780. [Google Scholar] [CrossRef]
  183. Chan, D.C. Fusion and fission: Interlinked processes critical for mitochondrial health. Annu. Rev. Genet. 2012, 46, 265–287. [Google Scholar] [CrossRef] [Green Version]
  184. Kraus, F.; Roy, K.; Pucadyil, T.J.; Ryan, M.T. Function and regulation of the divisome for mitochondrial fission. Nature 2021, 590, 57–66. [Google Scholar] [CrossRef]
  185. Liesa, M.; Palacin, M.; Zorzano, A. Mitochondrial dynamics in mammalian health and disease. Physiol. Rev. 2009, 89, 799–845. [Google Scholar] [CrossRef] [Green Version]
  186. Liesa, M.; Shirihai, O.S. Mitochondrial dynamics in the regulation of nutrient utilization and energy expenditure. Cell Metab. 2013, 17, 491–506. [Google Scholar] [CrossRef] [Green Version]
  187. Mishra, P.; Chan, D.C. Mitochondrial dynamics and inheritance during cell division, development and disease. Nat. Rev. Mol. Cell Biol. 2014, 15, 634–646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Mishra, P.; Chan, D.C. Metabolic regulation of mitochondrial dynamics. J. Cell Biol. 2016, 212, 379–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Kanki, T.; Wang, K.; Baba, M.; Bartholomew, C.R.; Lynch-Day, M.A.; Du, Z.; Geng, J.; Mao, K.; Yang, Z.; Yen, W.L.; et al. A genomic screen for yeast mutants defective in selective mitochondria autophagy. Mol. Biol. Cell 2009, 20, 4730–4738. [Google Scholar] [CrossRef] [Green Version]
  190. Ding, W.X.; Yin, X.M. Mitophagy: Mechanisms, pathophysiological roles, and analysis. Biol. Chem. 2012, 393, 547–564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Westermann, B. Mitochondrial dynamics in model organisms: What yeasts, worms and flies have taught us about fusion and fission of mitochondria. Semin. Cell Dev. Biol. 2010, 21, 542–549. [Google Scholar] [CrossRef] [PubMed]
  192. Gomes, L.C.; Scorrano, L. Mitochondrial morphology in mitophagy and macroautophagy. Biochim. Biophys. Acta 2013, 1833, 205–212. [Google Scholar] [CrossRef] [PubMed]
  193. Chen, G.; Kroemer, G.; Kepp, O. Mitophagy: An Emerging Role in Aging and Age-Associated Diseases. Front. Cell Dev. Biol. 2020, 8, 200. [Google Scholar] [CrossRef] [Green Version]
  194. Chakravorty, A.; Jetto, C.T.; Manjithaya, R. Dysfunctional Mitochondria and Mitophagy as Drivers of Alzheimer’s Disease Pathogenesis. Front. Aging Neurosci. 2019, 11, 311. [Google Scholar] [CrossRef]
  195. Jezek, J.; Cooper, K.F.; Strich, R. The Impact of Mitochondrial Fission-Stimulated ROS Production on Pro-Apoptotic Chemotherapy. Biology 2021, 10, 33. [Google Scholar] [CrossRef] [PubMed]
  196. Wang, Y.; Subramanian, M.; Yurdagul, A., Jr.; Barbosa-Lorenzi, V.C.; Cai, B.; de Juan-Sanz, J.; Ryan, T.A.; Nomura, M.; Maxfield, F.R.; Tabas, I. Mitochondrial Fission Promotes the Continued Clearance of Apoptotic Cells by Macrophages. Cell 2017, 171, 331–345.e322. [Google Scholar] [CrossRef] [Green Version]
  197. Pena-Blanco, A.; Garcia-Saez, A.J. Bax, Bak and beyond-mitochondrial performance in apoptosis. FEBS J. 2018, 285, 416–431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Karbowski, M.; Lee, Y.J.; Gaume, B.; Jeong, S.Y.; Frank, S.; Nechushtan, A.; Santel, A.; Fuller, M.; Smith, C.L.; Youle, R.J. Spatial and temporal association of Bax with mitochondrial fission sites, Drp1, and Mfn2 during apoptosis. J. Cell Biol. 2002, 159, 931–938. [Google Scholar] [CrossRef] [Green Version]
  199. Tsai, C.J.; Liu, S.; Hung, C.L.; Jhong, S.R.; Sung, T.C.; Chiang, Y.W. BAX-induced apoptosis can be initiated through a conformational selection mechanism. Structure 2015, 23, 139–148. [Google Scholar] [CrossRef] [Green Version]
  200. Antonsson, B.; Montessuit, S.; Sanchez, B.; Martinou, J.C. Bax is present as a high molecular weight oligomer/complex in the mitochondrial membrane of apoptotic cells. J. Biol. Chem. 2001, 276, 11615–11623. [Google Scholar] [CrossRef] [Green Version]
  201. Saito, M.; Korsmeyer, S.J.; Schlesinger, P.H. BAX-dependent transport of cytochrome c reconstituted in pure liposomes. Nat. Cell Biol. 2000, 2, 553–555. [Google Scholar] [CrossRef] [PubMed]
  202. Garcia-Saez, A.J. The BCL-2 family saga. Nat. Rev. Mol. Cell Biol. 2020, 21, 564–565. [Google Scholar] [CrossRef] [PubMed]
  203. Rathore, R.; McCallum, J.E.; Varghese, E.; Florea, A.M.; Busselberg, D. Overcoming chemotherapy drug resistance by targeting inhibitors of apoptosis proteins (IAPs). Apoptosis 2017, 22, 898–919. [Google Scholar] [CrossRef]
  204. Breckenridge, D.G.; Stojanovic, M.; Marcellus, R.C.; Shore, G.C. Caspase cleavage product of BAP31 induces mitochondrial fission through endoplasmic reticulum calcium signals, enhancing cytochrome c release to the cytosol. J. Cell Biol. 2003, 160, 1115–1127. [Google Scholar] [CrossRef] [PubMed]
  205. Estaquier, J.; Arnoult, D. Inhibiting Drp1-mediated mitochondrial fission selectively prevents the release of cytochrome c during apoptosis. Cell Death Differ. 2007, 14, 1086–1094. [Google Scholar] [CrossRef] [Green Version]
  206. Parone, P.A.; James, D.I.; Da Cruz, S.; Mattenberger, Y.; Donze, O.; Barja, F.; Martinou, J.C. Inhibiting the mitochondrial fission machinery does not prevent Bax/Bak-dependent apoptosis. Mol. Cell. Biol. 2006, 26, 7397–7408. [Google Scholar] [CrossRef] [Green Version]
  207. Strich, R. Programmed Cell Death Initiation and Execution in Budding Yeast. Genetics 2015, 200, 1003–1014. [Google Scholar] [CrossRef] [Green Version]
  208. Carmona-Gutierrez, D.; Eisenberg, T.; Buttner, S.; Meisinger, C.; Kroemer, G.; Madeo, F. Apoptosis in yeast: Triggers, pathways, subroutines. Cell Death Differ. 2010, 17, 763–773. [Google Scholar] [CrossRef] [Green Version]
  209. Chaves, S.R.; Rego, A.; Martins, V.M.; Santos-Pereira, C.; Sousa, M.J.; Corte-Real, M. Regulation of Cell Death Induced by Acetic Acid in Yeasts. Front. Cell Dev. Biol. 2021, 9, 642375. [Google Scholar] [CrossRef]
  210. Falcone, C.; Mazzoni, C. External and internal triggers of cell death in yeast. Cell. Mol. Life Sci. 2016, 73, 2237–2250. [Google Scholar] [CrossRef]
  211. Frank, S.; Gaume, B.; Bergmann-Leitner, E.S.; Leitner, W.W.; Robert, E.G.; Catez, F.; Smith, C.L.; Youle, R.J. The role of dynamin-related protein 1, a mediator of mitochondrial fission, in apoptosis. Dev. Cell 2001, 1, 515–525. [Google Scholar] [CrossRef] [Green Version]
  212. Demetrick, D.J.; Matsumoto, S.; Hannon, G.J.; Okamoto, K.; Xiong, Y.; Zhang, H.; Beach, D.H. Chromosomal mapping of the genes for the human cell cycle proteins cyclin C (CCNC), cyclin E (CCNE), p21 (CDKN1) and KAP (CDKN3). Cytogenet. Cell Genet. 1995, 69, 190–192. [Google Scholar] [CrossRef] [PubMed]
  213. Ohata, N.; Ito, S.; Yoshida, A.; Kunisada, T.; Numoto, K.; Jitsumori, Y.; Kanzaki, H.; Ozaki, T.; Shimizu, K.; Ouchida, M. Highly frequent allelic loss of chromosome 6q16-23 in osteosarcoma: Involvement of cyclin C in osteosarcoma. Int. J. Mol. Med. 2006, 18, 1153–1158. [Google Scholar] [CrossRef] [Green Version]
  214. Liang, J.; Chen, M.; Hughes, D.; Chumanevich, A.A.; Altilia, S.; Kaza, V.; Lim, C.U.; Kiaris, H.; Mythreye, K.; Pena, M.M.; et al. CDK8 Selectively Promotes the Growth of Colon Cancer Metastases in the Liver by Regulating Gene Expression of TIMP3 and Matrix Metalloproteinases. Cancer Res. 2018, 78, 6594–6606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Bragelmann, J.; Klumper, N.; Offermann, A.; von Massenhausen, A.; Bohm, D.; Deng, M.; Queisser, A.; Sanders, C.; Syring, I.; Merseburger, A.S.; et al. Pan-Cancer Analysis of the Mediator Complex Transcriptome Identifies CDK19 and CDK8 as Therapeutic Targets in Advanced Prostate Cancer. Clin. Cancer. Res. 2017, 23, 1829–1840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. McDermott, M.S.; Chumanevich, A.A.; Lim, C.U.; Liang, J.; Chen, M.; Altilia, S.; Oliver, D.; Rae, J.M.; Shtutman, M.; Kiaris, H.; et al. Inhibition of CDK8 mediator kinase suppresses estrogen dependent transcription and the growth of estrogen receptor positive breast cancer. Oncotarget 2017, 8, 12558–12575. [Google Scholar] [CrossRef] [Green Version]
  217. Roninson, I.B.; Gyorffy, B.; Mack, Z.T.; Shtil, A.A.; Shtutman, M.S.; Chen, M.; Broude, E.V. Identifying Cancers Impacted by CDK8/19. Cells 2019, 8, 821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Fant, C.B.; Taatjes, D.J. Regulatory functions of the Mediator kinases CDK8 and CDK19. Transcription 2019, 10, 76–90. [Google Scholar] [CrossRef]
  219. Rambold, A.S.; Kostelecky, B.; Elia, N.; Lippincott-Schwartz, J. Tubular network formation protects mitochondria from autophagosomal degradation during nutrient starvation. Proc. Natl. Acad. Sci. USA 2011, 108, 10190–10195. [Google Scholar] [CrossRef] [Green Version]
  220. Rambold, A.S.; Kostelecky, B.; Lippincott-Schwartz, J. Together we are stronger: Fusion protects mitochondria from autophagosomal degradation. Autophagy 2011, 7, 1568–1569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Morita, M.; Prudent, J.; Basu, K.; Goyon, V.; Katsumura, S.; Hulea, L.; Pearl, D.; Siddiqui, N.; Strack, S.; McGuirk, S.; et al. mTOR Controls Mitochondrial Dynamics and Cell Survival via MTFP1. Mol. Cell 2017, 67, 922–935 e925. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Gomes, L.C.; Di Benedetto, G.; Scorrano, L. During autophagy mitochondria elongate, are spared from degradation and sustain cell viability. Nat. Cell Biol. 2011, 13, 589–598. [Google Scholar] [CrossRef] [Green Version]
  223. Tondera, D.; Grandemange, S.; Jourdain, A.; Karbowski, M.; Mattenberger, Y.; Herzig, S.; Da Cruz, S.; Clerc, P.; Raschke, I.; Merkwirth, C.; et al. SLP-2 is required for stress-induced mitochondrial hyperfusion. EMBO J. 2009, 28, 1589–1600. [Google Scholar] [CrossRef] [Green Version]
  224. Egner, A.; Jakobs, S.; Hell, S.W. Fast 100-nm resolution three-dimensional microscope reveals structural plasticity of mitochondria in live yeast. Proc. Natl. Acad. Sci. USA 2002, 99, 3370–3375. [Google Scholar] [CrossRef] [Green Version]
  225. Walden, E.A.; Fong, R.Y.; Pham, T.T.; Knill, H.; Laframboise, S.J.; Huard, S.; Harper, M.E.; Baetz, K. Phenomic screen identifies a role for the yeast lysine acetyltransferase NuA4 in the control of Bcy1 subcellular localization, glycogen biosynthesis, and mitochondrial morphology. PLoS Genet. 2020, 16, e1009220. [Google Scholar] [CrossRef]
  226. Zemirli, N.; Morel, E.; Molino, D. Mitochondrial Dynamics in Basal and Stressful Conditions. Int. J. Mol. Sci. 2018, 19, 564. [Google Scholar] [CrossRef] [Green Version]
  227. Wright, P.E.; Dyson, H.J. Intrinsically disordered proteins in cellular signalling and regulation. Nat. Rev. Mol. Cell Biol. 2015, 16, 18–29. [Google Scholar] [CrossRef]
  228. Anbo, H.; Sato, M.; Okoshi, A.; Fukuchi, S. Functional Segments on Intrinsically Disordered Regions in Disease-Related Proteins. Biomolecules 2019, 9, 88. [Google Scholar] [CrossRef] [Green Version]
  229. Brodsky, S.; Jana, T.; Mittelman, K.; Chapal, M.; Kumar, D.K.; Carmi, M.; Barkai, N. Intrinsically Disordered Regions Direct Transcription Factor In Vivo Binding Specificity. Mol. Cell 2020, 79, 459–471 e454. [Google Scholar] [CrossRef] [PubMed]
  230. Lee, J.H.; Bassel-Duby, R.; Olson, E.N. Heart- and muscle-derived signaling system dependent on MED13 and Wingless controls obesity in Drosophila. Proc. Natl. Acad. Sci. USA 2014, 111, 9491–9496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Tutter, A.V.; Kowalski, M.P.; Baltus, G.A.; Iourgenko, V.; Labow, M.; Li, E.; Kadam, S. Role for Med12 in regulation of Nanog and Nanog target genes. J. Biol. Chem. 2009, 284, 3709–3718. [Google Scholar] [CrossRef] [Green Version]
  232. Miao, Y.L.; Gambini, A.; Zhang, Y.; Padilla-Banks, E.; Jefferson, W.N.; Bernhardt, M.L.; Huang, W.; Li, L.; Williams, C.J. Mediator complex component MED13 regulates zygotic genome activation and is required for postimplantation development in the mouse. Biol. Reprod. 2018, 98, 449–464. [Google Scholar] [CrossRef] [Green Version]
  233. Loncle, N.; Boube, M.; Joulia, L.; Boschiero, C.; Werner, M.; Cribbs, D.L.; Bourbon, H.M. Distinct roles for Mediator Cdk8 module subunits in Drosophila development. EMBO J. 2007, 26, 1045–1054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Gillmor, C.S.; Silva-Ortega, C.O.; Willmann, M.R.; Buendia-Monreal, M.; Poethig, R.S. The Arabidopsis Mediator CDK8 module genes CCT (MED12) and GCT (MED13) are global regulators of developmental phase transitions. Development 2014, 141, 4580–4589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Keightley, M.C.; Layton, J.E.; Hayman, J.W.; Heath, J.K.; Lieschke, G.J. Mediator subunit 12 is required for neutrophil development in zebrafish. PLoS ONE 2011, 6, e23845. [Google Scholar] [CrossRef] [Green Version]
  236. Barbieri, C.E.; Baca, S.C.; Lawrence, M.S.; Demichelis, F.; Blattner, M.; Theurillat, J.P.; White, T.A.; Stojanov, P.; Van Allen, E.; Stransky, N.; et al. Exome sequencing identifies recurrent SPOP, FOXA1 and MED12 mutations in prostate cancer. Nat. Genet. 2012, 44, 685–689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Makinen, N.; Mehine, M.; Tolvanen, J.; Kaasinen, E.; Li, Y.; Lehtonen, H.J.; Gentile, M.; Yan, J.; Enge, M.; Taipale, M.; et al. MED12, the mediator complex subunit 12 gene, is mutated at high frequency in uterine leiomyomas. Science 2011, 334, 252–255. [Google Scholar] [CrossRef] [PubMed]
  238. Asadollahi, R.; Zweier, M.; Gogoll, L.; Schiffmann, R.; Sticht, H.; Steindl, K.; Rauch, A. Genotype-phenotype evaluation of MED13L defects in the light of a novel truncating and a recurrent missense mutation. Eur. J. Med. Genet. 2017, 60, 451–464. [Google Scholar] [CrossRef]
  239. van Haelst, M.M.; Monroe, G.R.; Duran, K.; van Binsbergen, E.; Breur, J.M.; Giltay, J.C.; van Haaften, G. Further confirmation of the MED13L haploinsufficiency syndrome. Eur. J. Hum. Genet. 2015, 23, 135–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Tyler, J.K.; Johnson, J.E. The role of autophagy in the regulation of yeast life span. Ann. N. Y. Acad. Sci. 2018, 1418, 31–43. [Google Scholar] [CrossRef]
  241. Alarcon, C.; Zaromytidou, A.I.; Xi, Q.; Gao, S.; Yu, J.; Fujisawa, S.; Barlas, A.; Miller, A.N.; Manova-Todorova, K.; Macias, M.J.; et al. Nuclear CDKs drive Smad transcriptional activation and turnover in BMP and TGF-beta pathways. Cell 2009, 139, 757–769. [Google Scholar] [CrossRef] [Green Version]
  242. Hengartner, C.J.; Thompson, C.M.; Zhang, J.; Chao, D.M.; Liao, S.M.; Koleske, A.J.; Okamura, S.; Young, R.A. Association of an activator with an RNA polymerase II holoenzyme. Genes Dev. 1995, 9, 897–910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Donner, A.J.; Hoover, J.M.; Szostek, S.A.; Espinosa, J.M. Stimulus-specific transcriptional regulation within the p53 network. Cell Cycle 2007, 6, 2594–2598. [Google Scholar] [CrossRef] [Green Version]
  244. Guo, Y.E.; Manteiga, J.C.; Henninger, J.E.; Sabari, B.R.; Dall’Agnese, A.; Hannett, N.M.; Spille, J.H.; Afeyan, L.K.; Zamudio, A.V.; Shrinivas, K.; et al. Pol II phosphorylation regulates a switch between transcriptional and splicing condensates. Nature 2019, 572, 543–548. [Google Scholar] [CrossRef] [Green Version]
  245. Gonzalez, D.; Hamidi, N.; Del Sol, R.; Benschop, J.J.; Nancy, T.; Li, C.; Francis, L.; Tzouros, M.; Krijgsveld, J.; Holstege, F.C.; et al. Suppression of Mediator is regulated by Cdk8-dependent Grr1 turnover of the Med3 coactivator. Proc. Natl. Acad. Sci. USA 2014, 111, 2500–2505. [Google Scholar] [CrossRef] [Green Version]
  246. Poss, Z.C.; Ebmeier, C.C.; Odell, A.T.; Tangpeerachaikul, A.; Lee, T.; Pelish, H.E.; Shair, M.D.; Dowell, R.D.; Old, W.M.; Taatjes, D.J. Identification of Mediator Kinase Substrates in Human Cells using Cortistatin A and Quantitative Phosphoproteomics. Cell Rep. 2016, 15, 436–450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Chi, Y.; Huddleston, M.J.; Zhang, X.; Young, R.A.; Annan, R.S.; Carr, S.A.; Deshaies, R.J. Negative regulation of Gcn4 and Msn2 transcription factors by Srb10 cyclin-dependent kinase. Genes Dev. 2001, 15, 1078–1092. [Google Scholar] [CrossRef] [Green Version]
  248. Raithatha, S.; Su, T.C.; Lourenco, P.; Goto, S.; Sadowski, I. Cdk8 regulates stability of the transcription factor Phd1 to control pseudohyphal differentiation of Saccharomyces cerevisiae. Mol. Cell. Biol. 2012, 32, 664–674. [Google Scholar] [CrossRef] [Green Version]
  249. Zhao, J.; Ramos, R.; Demma, M. CDK8 regulates E2F1 transcriptional activity through S375 phosphorylation. Oncogene 2013, 32, 3520–3530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Fryer, C.J.; White, J.B.; Jones, K.A. Mastermind recruits CycC:CDK8 to phosphorylate the Notch ICD and coordinate activation with turnover. Mol. Cell 2004, 16, 509–520. [Google Scholar] [CrossRef]
Figure 1. Top panel: Model outlining the different mechanisms used by the CKM to relieve repression on SRG’s in response to either cell death (oxidative stress) or cell survival (nitrogen starvation) cues in yeast. Bottom panel: Live fluorescence images of cyclin C-YFP and DS-Red tagged mitochondria demonstrating the different phenotypes observed after 2 h treatment with H2O2 (left) or following nitrogen starvation (SD-N—right). CKM-Cdk8 kinase module, SRG’s—stress response genes, YFP—yellow fluorescence protein.
Figure 1. Top panel: Model outlining the different mechanisms used by the CKM to relieve repression on SRG’s in response to either cell death (oxidative stress) or cell survival (nitrogen starvation) cues in yeast. Bottom panel: Live fluorescence images of cyclin C-YFP and DS-Red tagged mitochondria demonstrating the different phenotypes observed after 2 h treatment with H2O2 (left) or following nitrogen starvation (SD-N—right). CKM-Cdk8 kinase module, SRG’s—stress response genes, YFP—yellow fluorescence protein.
Microorganisms 09 02152 g001
Figure 2. Model of the interaction of the CKM with the core mediator complex and RNA Pol II. (A). In unstressed cells, the CKM associates with the Mediator at UAS sites found in promotors, by the interaction of Med13 with the mediator hook. This inhibits the Mediator-RNA pol II interaction, preventing PIC formation and transcription of mRNA. Following stress, the CKM is released from the Mediator, permitting PIC assembly and transcription of mRNA. Additional repressors and/or activators also mediate transcription by binding to UAS motifs (Upstream Activating Sequence). Adapted from Cherji et al (2017). NAC. PMID:28575439. (B) Structure of the yeast CKM determined by cryo-EM and mass spectrometry. Reproduced with permission from Li et al (2021) Scientific Advances. PMID: 33390853. CKM—Cdk8 kinase module, RNA Pol II—RNA polymerase II, PIC—preinitiation complex, Å—Angstrom.
Figure 2. Model of the interaction of the CKM with the core mediator complex and RNA Pol II. (A). In unstressed cells, the CKM associates with the Mediator at UAS sites found in promotors, by the interaction of Med13 with the mediator hook. This inhibits the Mediator-RNA pol II interaction, preventing PIC formation and transcription of mRNA. Following stress, the CKM is released from the Mediator, permitting PIC assembly and transcription of mRNA. Additional repressors and/or activators also mediate transcription by binding to UAS motifs (Upstream Activating Sequence). Adapted from Cherji et al (2017). NAC. PMID:28575439. (B) Structure of the yeast CKM determined by cryo-EM and mass spectrometry. Reproduced with permission from Li et al (2021) Scientific Advances. PMID: 33390853. CKM—Cdk8 kinase module, RNA Pol II—RNA polymerase II, PIC—preinitiation complex, Å—Angstrom.
Microorganisms 09 02152 g002
Figure 3. Outline of the Cell Wall Integrity (CWI) signal transduction pathway that mediates cyclin C translocation to the cytoplasm following low oxidative stress (0.4 mM H2O2). Both the MAPK of this pathway (Slt2) and its pseudo-kinase partner, Kdx1 are required for the efficient export and degradation of cyclin C. Cyclin C release is dependent upon Med13 degradation by the SCFGrr1 E3 ligase complex, triggered by Cdk8, Slt2 and the AMPK Snf1-mediated phosphorylation. Following high levels of oxidative stress (1.2 mM H2O2) the activity of Ste11, the MAPKKK of the HOG signal transduction pathway is also required. For clarity, Ask10 and Plc1, which also play roles in cyclin C’s fate are omitted from the figure. MAPK—mitogen activated protein kinase, SCF—Skp1-Cullin-F-box-protein, HOG—high osmolarity glycerol, AMPK—adenosine monophosphate-activated protein kinase, P—phosphorylation event.
Figure 3. Outline of the Cell Wall Integrity (CWI) signal transduction pathway that mediates cyclin C translocation to the cytoplasm following low oxidative stress (0.4 mM H2O2). Both the MAPK of this pathway (Slt2) and its pseudo-kinase partner, Kdx1 are required for the efficient export and degradation of cyclin C. Cyclin C release is dependent upon Med13 degradation by the SCFGrr1 E3 ligase complex, triggered by Cdk8, Slt2 and the AMPK Snf1-mediated phosphorylation. Following high levels of oxidative stress (1.2 mM H2O2) the activity of Ste11, the MAPKKK of the HOG signal transduction pathway is also required. For clarity, Ask10 and Plc1, which also play roles in cyclin C’s fate are omitted from the figure. MAPK—mitogen activated protein kinase, SCF—Skp1-Cullin-F-box-protein, HOG—high osmolarity glycerol, AMPK—adenosine monophosphate-activated protein kinase, P—phosphorylation event.
Microorganisms 09 02152 g003
Figure 4. Upper panel: Outline of the Snx4-assisted autophagy pathway of transcription factors. This pathway transports Med13 from the nucleus to the vacuole for proteolysis following TORC1 inhibition. NPC—nuclear pore complex, PAS—Pre-autophagosomal structure.
Figure 4. Upper panel: Outline of the Snx4-assisted autophagy pathway of transcription factors. This pathway transports Med13 from the nucleus to the vacuole for proteolysis following TORC1 inhibition. NPC—nuclear pore complex, PAS—Pre-autophagosomal structure.
Microorganisms 09 02152 g004
Figure 5. Outline of the different mitochondrial dynamics observed in unstressed and ROS (Reactive oxygen species)-treated yeast cells. See text for details. The fluorescence images are of budding yeast harboring a Mt-DSRed plasmid that localizes to the OMM (outer mitochondrial membrane). MOMP—mitochondria outer membrane permeabilization, RCD—regulated cell death.
Figure 5. Outline of the different mitochondrial dynamics observed in unstressed and ROS (Reactive oxygen species)-treated yeast cells. See text for details. The fluorescence images are of budding yeast harboring a Mt-DSRed plasmid that localizes to the OMM (outer mitochondrial membrane). MOMP—mitochondria outer membrane permeabilization, RCD—regulated cell death.
Microorganisms 09 02152 g005
Figure 6. Similarities and differences between cyclin C and Med13’s state following stress. See text for details. SRG—stress response genes, UPS—ubiquitin proteasome system.
Figure 6. Similarities and differences between cyclin C and Med13’s state following stress. See text for details. SRG—stress response genes, UPS—ubiquitin proteasome system.
Microorganisms 09 02152 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Friedson, B.; Cooper, K.F. Cdk8 Kinase Module: A Mediator of Life and Death Decisions in Times of Stress. Microorganisms 2021, 9, 2152. https://doi.org/10.3390/microorganisms9102152

AMA Style

Friedson B, Cooper KF. Cdk8 Kinase Module: A Mediator of Life and Death Decisions in Times of Stress. Microorganisms. 2021; 9(10):2152. https://doi.org/10.3390/microorganisms9102152

Chicago/Turabian Style

Friedson, Brittany, and Katrina F. Cooper. 2021. "Cdk8 Kinase Module: A Mediator of Life and Death Decisions in Times of Stress" Microorganisms 9, no. 10: 2152. https://doi.org/10.3390/microorganisms9102152

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop