Next Article in Journal
The Effects of Mycovirus BmPV36 on the Cell Structure and Transcription of Bipolaris maydis
Next Article in Special Issue
Processes Controlling the Contractile Ring during Cytokinesis in Fission Yeast, Including the Role of ESCRT Proteins
Previous Article in Journal
Galleria mellonella Model of Coccidioidomycosis for Drug Susceptibility Tests and Virulence Factor Identification
Previous Article in Special Issue
The Roles of Septins in Regulating Fission Yeast Cytokinesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Meiotic Cytokinesis in Saccharomyces cerevisiae: Spores That Just Need Closure

by
Matthew Durant
,
Xheni Mucelli
and
Linda S. Huang
*
Department of Biology, University of Massachusetts Boston, Boston, MA 02125, USA
*
Author to whom correspondence should be addressed.
J. Fungi 2024, 10(2), 132; https://doi.org/10.3390/jof10020132
Submission received: 12 January 2024 / Revised: 30 January 2024 / Accepted: 4 February 2024 / Published: 6 February 2024
(This article belongs to the Special Issue Yeast Cytokinesis)

Abstract

:
In the budding yeast Saccharomyces cerevisiae, sporulation occurs during starvation of a diploid cell and results in the formation of four haploid spores forming within the mother cell ascus. Meiosis divides the genetic material that is encapsulated by the prospore membrane that grows to surround the haploid nuclei; this membrane will eventually become the plasma membrane of the haploid spore. Cellularization of the spores occurs when the prospore membrane closes to capture the haploid nucleus along with some cytoplasmic material from the mother cell, and thus, closure of the prospore membrane is the meiotic cytokinetic event. This cytokinetic event involves the removal of the leading-edge protein complex, a complex of proteins that localizes to the leading edge of the growing prospore membrane. The development and closure of the prospore membrane must be coordinated with other meiotic exit events such as spindle disassembly. Timing of the closure of the prospore membrane depends on the meiotic exit pathway, which utilizes Cdc15, a Hippo-like kinase, and Sps1, an STE20 family GCKIII kinase, acting in parallel to the E3 ligase Ama1-APC/C. This review describes the sporulation process and focuses on the development of the prospore membrane and the regulation of prospore membrane closure.

1. Introduction

Meiosis is a specialized form of cell division that results in the formation of haploid gametes used for sexual reproduction. In the budding yeast Saccharomyces cerevisiae, meiosis occurs as cells undergo sporulation, which happens when diploid cells are starved of nitrogen and a fermentable carbon source. Four haploid spores are formed within the diploid mother cell upon completion of sporulation; the plasma membrane and cell wall from the former mother cell become the ascus that encases the four spores.
Sporulation involves both the processes of meiosis and spore morphogenesis (Figure 1). Meiosis in S. cerevisiae occurs within the cytoplasm of the mother cell. The formation of the spores within the mother cell requires the de novo generation of two distinct cellular structures: a new phospholipid membrane for each daughter cell called the prospore membrane, and a new cell wall that contains different components from the vegetative yeast cell wall called the spore wall. Cellularization of the spore occurs when the prospore membrane closes; this is the cytokinetic event of meiosis in budding yeast. The sporulation process was reviewed extensively in [1]. The chromosomal duplication, recombination, and segregation aspects of meiosis, including meiosis I and the transition to meiosis II, have been recently reviewed [2]. Here, we focus on more recently published work related to meiosis II and meiotic cytokinesis in S. cerevisiae.

2. Sporulation in S. cerevisiae Involves Meiosis and Spore Morphogenesis

During sporulation, cells undergo both meiosis and spore morphogenesis. The early phase of sporulation is marked by gene expression shifts from mitotic genes to early meiotic genes, which begin when diploid cells lack nitrogen and lack a fermentable carbon source (reviewed in [3]). During starvation, diploid cells will alter gene expression profiles as they exit the mitotic cell cycle in G1 and enter the premeiotic S phase in preparation for the events of meiosis, including recombination, homology pairing, and two rounds of chromosomal segregation (reviewed in [2]).
Spore morphogenesis begins at meiosis II, with the modification of the cytoplasmic face of the spindle pole body (SPB) with a protein coat called the meiotic outer plaque, which will serve as membrane nucleation sites for the initiation of the prospore membrane [4,5,6]. The SPB is the yeast equivalent of the centrosome, and functions as the sole microtubule-organizing center in S. cerevisiae cells. In meiosis, the SPBs duplicate both at the beginning of meiosis I and again at the beginning of meiosis II, resulting in the four SPBs required for the second nuclear division. During meiosis II, the four SPBs change composition and function both in microtubule nucleation and membrane nucleation [4]. The modified SPBs are expanded and replace the microtubule nucleator Spc72 with a group of proteins that form the meiotic outer plaque. The meiotic outer plaque consists of Spo21 (also called Mpc70), Mpc54, Spo74, and Ady4 [7,8,9]. The deposition of Spo21 and Spo74 depend on Pfs1 (also called Ady1) [10,11]. Spo21 and Mpc54 are both coiled-coil proteins that have their N-termini facing the cytoplasmic face of the meiotic outer plaque [12]. The C-termini of both Spo21 and Mpc54 face the N-terminus of Cnm67, a constitutive SPB member [13].
As cells undergo meiosis, the number of nuclei that are packaged into spores can vary from one (monad) to up to four (tetrad). The reasons for packaging fewer than the expected four spores can vary. Disruption of the meiotic division process can lead to a situation where two diploid spores are packaged, as seen in spo12∆ and spo13∆ mutants [14]. Sometimes, cells package fewer than four spores, even when the meiotic divisions occur. The situations where two haploid spores are packaged (forming dyads) can be categorized into two genetically distinct types: those that package nonsister haploid spores [11,15,16,17] versus those that package spores randomly (such as in ady3∆ mutants [18,19]).
There are at least two distinct mechanisms that cells use to package fewer than four haploid spores. The process by which cells package nonsister dyads (as well as triads of three spores and monads of one spore) has been termed spore number control, and depends on the availability of acetate [15,16]. Cells may choose to package fewer spores if the nutritional environment is unfavorable; this process depends on the Ras/cAMP/Protein Kinase A pathway [11]. Spore number control occurs by modifying fewer SPBs, which leads to the formation of fewer prospore membranes and thus fewer spores. This is in contrast to the random packaging of spores, as described in ady3∆ mutants, which utilizes a different (and not yet fully understood) mechanism: ady3∆ cells can produce four prospore membranes but will often have fewer than four spores reaching maturity [18,19].
Sporulating cells modify their SPBs depending on the nutrients present in the environment. An intermediate metabolite in the glyoxylate pathway is required for modification of all four SPBs [15]. In the absence of this metabolite, cells preferentially modify the two daughter SPBs formed in the transition from meiosis I to meiosis II, leading to the packaging of only two spores and the formation of dyads (two spores packaged within a mother cell ascus) [15].
The production of the components of the meiotic outer plaque (Mpc54, Spo74, and Mpc70) relies on the concentration of acetate in the media of sporulating cells [16]. The levels of meiotic outer plaque components directly correlate to sporulation efficiency: lower levels of meiotic outer plaque protein expression lead to fewer SPB modifications, ultimately resulting in fewer than four spores [16]. A core component of the mitotic exit network (the NDR/LATS kinase complex Mob1-Dbf2 [20]) is repurposed for spore number control in meiosis to modify the SPB components [17,21].
During spore morphogenesis, the prospore membrane grows through the fusion of post-Golgi vesicles at the meiotic outer plaque to surround the new meiotic products (Figure 2) [5,6]. Lipid droplets directly contact the prospore membrane in meiosis II and may also contribute to the expansion of the prospore membrane [22]. The four prospore membranes grow as double lipid bilayers to encapsulate each haploid nucleus, rounding up and closing at the end of meiosis. The outer bilayer lyses during spore wall deposition while the inner bilayer becomes the plasma membrane of the newly formed spore [23]. Closure of the prospore membrane leads to cellularization and is the cytokinetic event of meiosis, as this event separates the nuclei of the daughter cells from the mother cell cytoplasm and also captures some cytoplasmic components from the mother cell. Aging-related factors are kept in the mother cell and segregated away from the spore during sporulation, creating robust gametes (reviewed in [24]).
The prospore membrane is important for spore morphogenesis, as the spore wall is deposited within the lumen of the double lipid bilayer of the prospore membrane. The spore wall differs from the vegetative cell wall and is important for the ability of the spore to survive harsh environmental conditions [28]. Spore wall deposition utilizes the prospore membrane as a template and is an important part of spore maturation and viability [23,29]. The spore wall is made up of four distinct layers: two inner layers made up of mannan and β-glucan, and two outer layers made of chitosan and dityrosine (reviewed in [1,30]). The mannan layer is first deposited into the lumen of the prospore membrane, followed by the glucan layer [31]. Following deposition of the inner spore wall layers, the outer lipid bilayer of the prospore membrane lyses [23]. The deposition of the glucan layer is negatively regulated by the sporulation-specific MAP kinase SMK1, which is needed for cells to transition from the deposition of the inner spore wall to the deposition of the outer chitosan layer [32]. The chitosan layer is needed so that the dityrosine layer can be properly deposited [23]. The formation of the outer spore wall layers requires lipid droplets, which have been proposed to activate chitin synthesis in a manner that requires SRT1 [22,33,34]. During spore maturation, spores form interspore bridges that are continuous with the outer layers of the spore wall; these bridges hold spores together when the ascal wall is disrupted [23]. Spore maturation also involves the rupture of the mother cell vacuole, leading to digestion of incompletely developed spores and compaction of the ascus [35,36].
The spore wall is distinct from the cell wall of vegetatively growing cells, which are composed of an inner layer of chitin and β-glucan and an outer mannan layer; this order is different from the spore wall whose innermost layer is the mannan layer (reviewed in [30,37,38,39]). Furthermore, the two outer layers (chitosan and dityrosine) are not present in the vegetative cell wall and allow the spore to survive harsher conditions (including the Drosophila intestinal tract [40]) compared to a vegetatively growing cell [29,41,42].
Autophagy is also essential for sporulating cells (reviewed in [43]), as cells lacking important components of the autophagic pathways do not form spores [44,45,46,47,48,49]. Autophagy is required for the clearance of amyloid-like Rim4 aggregates, which prevent translation of middle and late meiotic transcripts [50,51,52,53]. In addition to the role of Rim4 in mRNA sequestration and regulation, Rim4 also stimulates autophagy through activation of the Atg1 kinase [54,55]. Autophagy is needed for progression through meiosis through the degradation of cyclins and the polo-like kinase Cdc5 as well as the release of the phosphatase Cdc14 from the nucleolus during anaphase II [50]. Interestingly, Cdc14 also appears to promote autophagy during sporulation through the dephosphorylation of Atg13 [56] and by regulating the localization and autophagic destruction of Rim4 [55]. Thus, autophagy is an important regulator of meiotic progression.

3. Sporulation Is Regulated by Changes in Gene Expression

The early phase of sporulation involves the initiation of transcription and translation of several meiotic genes (reviewed in [3]). The increase in transcription of many early meiotic genes is regulated by the master regulator of meiotic genes, IME1 [57,58]. IME1 is expressed by diploid cells in response to nitrogen starvation [57,59]. Among the many IME1-dependent genes is another master regulator of meiotic transcription, NDT80 [60,61,62]. NDT80 binds to specific sequences in the yeast genome known as middle sporulation elements (MSEs). The translation of Ndt80 and the subsequent upregulation of middle sporulation genes marks the transition from early sporulation to middle sporulation.
Several middle sporulation genes have their timing of expression regulated through multiple levels of transcriptional and translational control (reviewed in [63]). SUM1 encodes a transcriptional repressor that also binds to some of the same MSEs that are bound by Ndt80 [64,65,66]. Sum1 must be removed from these MSEs prior to Ndt80 binding to promote transcription of these Ndt80 targets. Genes with promoter elements bound by Sum1 are termed middle–late sporulation genes, as their expression is delayed beyond the initial expression of Ndt80. Sum1 is removed from MSEs both through phosphorylation by Ime2, Cdk1, and Cdc7, as well as high levels of Ndt80 expression [66,67,68,69,70,71]. Phosphorylation of Sum1 weakens repressor binding to MSEs and allows Ndt80 to outcompete phosphorylated Sum1 for binding to the same DNA sequences, leading to transcriptional upregulation of affected genes. Sum1 also directly represses NDT80 transcription, creating a feed-forward loop that also includes the increase in IME2 transcription by Ndt80 [72].
During sporulation, there is significant post-transcriptional regulation of transcripts [73]. Sequestration of mRNA transcripts by aggregates of Rim4 may contribute to this regulation, as Rim4 aggregates sequester several transcripts in sporulation and prevent their translation until the aggregates are cleared by Ime2 phosphorylation [51,52,53]. In addition to Rim4 sequestration, the Pes4 and Mip6 RNA-binding proteins play a similar role in delaying the translation of mRNA transcripts in sporulation [74].

4. The Development of the Prospore Membrane Involves Initiation and Elongation

Cellularization of the spore occurs as the prospore membrane grows to surround the meiotic products (Figure 2). The fusion of post-Golgi vesicles at the meiotic outer plaque initiates prospore membrane growth and is mediated by SNAREs present on the meiotic outer plaque [6]. ADY4 is required for a stable interaction between the meiotic outer plaque and the prospore membrane. ADY4 was initially identified as a meiotic outer plaque component [9] and has been shown to stabilize the meiotic outer plaque through the recruitment of the lipid kinase Mss4 [75]. Furthermore, SPO21, which encodes another meiotic outer plaque component [7,8], appears to be important for proper prospore membrane number control. Modifications to the N-terminus of Spo21, which change positively charged residues to neutral residues, result in more than four prospore membranes initiating [75].
Many genes have been implicated in various aspects of prospore membrane development. SPO71, SPO73, and VPS13 all appear to have roles in prospore membrane extension, as they are required for prospore membranes to become the appropriate size [27,76,77,78,79,80,81]. SPO73 encodes a dysferlin domain-only protein that is involved in prospore membrane extension [78,79]. SPO71 encodes a pleckstrin homology domain protein expressed specifically in sporulating cells [27,82,83]. Despite forming smaller rounded prospore membranes at the end of meiosis II, cells deficient for SPO71 do not prematurely close their prospore membranes and do elongate prospore membranes in a timely manner [27]. SPO71 is required to translocate Vps13 from the endosome to the prospore membrane [77]. SPO71 and SPO73 act cooperatively in sporulation to regulate prospore membrane morphogenesis and are together required for proper Vps13 localization to the prospore membrane [77,79].
Vps13 belongs to a family of lipid transporters that act at membrane contact sites [84,85]. In vegetatively growing cells, Vps13 localization can change depending on the nutrient conditions of cells, localizing to endosome–mitochondria junction sites on rich (glucose) media, while instead localizing to nuclear–vacuole junctions on poor (acetate) media [80,86]. VPS13 can also facilitate lipid transfer from the Golgi to the vacuole [87,88]. Vps13 mediates lipid transfer through membrane contact sites via a large hydrophobic channel that covers the junction between two organelle membranes [86,89,90,91].
During sporulation, Vps13 localizes to prospore membranes during sporulation [76]. Localization of Vps13 to the prospore membrane is facilitated by an adaptor complex made up of Spo71 and Spo73. The Spo71-Spo73 adaptor complex also mediates the ability of Vps13 to form membrane contact sites between the ER and the prospore membrane [81]. Specifically, VPS13 is required for the contact sites between the ER and the plasma membrane and can link ER–plasma membrane tethers along the prospore membrane. The Vps13 ER–plasma membrane contact sites are mediated by the presence of PI4P and facilitated through the Spo71-Spo73 complex [81].
Septins are also present during prospore membrane development, but do not play the same role they do during mitotic cytokinesis. The septins are present along elongating prospore membranes [92]. However, unlike in mitotically growing cells, the septins do not recruit actomyosin contractile ring components (which are not utilized during prospore membrane closure [93] (see below)). During sporulation, the septin structures are composed of slightly different multimers than in mitosis, with Spr28 and Spr3 replacing the mitotic septins Shs1 and Cdc12 [66,94]. Although Cdc12 is detectable during sporulation, it does not form higher-order structures with the other septins present in sporulation [95]. The septin genes are individually dispensable for sporulation [92,96]. However, loss of both SPR28 and SPR3 leads to a defect in sporulation that cannot be rescued by the overexpression of other septins [97]. SPR28 and SPR3 likewise cannot complement the loss of mitotic septins in mitosis, indicating that there must be some difference between the meiosis specific function of the meiotic septin complex and the mitotic septin complex [97]. It is possible that the probable loss of GTPase activity in both Spr3 and Spr28 plays some role in differentiating septin function between mitosis and meiosis [98].
Although individual septin mutants have a minimal effect on sporulation efficiency in most strain backgrounds, a role for septins during sporulation in prospore membrane elongation around the nuclear envelope was seen in the BY4743 strain background [44,45,99]. [The SK1 yeast strain background [100] is more typically used in sporulation studies, due to its ability to sporulate efficiently and somewhat synchronously]. In the BY4743 background, cells lacking SPR3 or SPR28 mislocalize Ady3, a component of the leading-edge protein complex (LEP; discussed below) found at the growing edge of the prospore membrane [99]. Interestingly, the septin Cdc10 appears to play a role after spore morphogenesis by marking polarization sites on the daughter spores, directing subsequent growth away from sister spores within the ascus [101,102].
Septin localization to the prospore membrane in sporulation depends on the protein phosphatase Glc7 and its sporulation-specific regulator Gip1 [31,103]. Additionally, the Glc7-Gip1 phosphatase complex is required for prospore membrane extension similar to the requirement for the Spo71-Spo73-Vps13 complex, although the two seem to operate in parallel [103]. Gip1 and Glc7 are also required for the generation of endoplasmic reticulum exit sites (ERESs), which are regenerated within the developing spores during the time of membrane extension [104]. SPO71 is required for the localization of the sporulation-specific septin Spr28, indicating a possible link between septins and prospore membrane elongation [27].
Interestingly, the actin cytoskeleton appears to be dispensable for the development and closure of the prospore membrane. The actin cytoskeleton plays only a minor role in facilitating the transport of prospore membrane precursors and appears to be dispensable for prospore membrane initiation, extension, or closure, but is required for later spore wall deposition [93].

5. Removal of the Leading-Edge Protein Complex (LEP) Is Important for Prospore Membrane Closure

A protein complex that consists of at least four protein components (Don1, Ssp1, Ady3, and Irc10) is found at the leading edge of the growing prospore membrane [7,18,19,105]; this complex has been termed the leading-edge protein complex (LEP). The LEP forms a ring at the open end of each growing prospore membrane. The precise function of each protein is unknown, but it has been proposed that the LEP is responsible for exerting force to keep elongating prospore membranes open [106,107]. This protein complex is removed from the prospore membrane before it closes [106]. To date, LEP removal is the last known event to occur before the prospore membranes round up and close. Prospore membrane closure requires fusion of the open ends of the membrane to create the double lipid bilayers that fully surround the newly formed spore. Whether LEP removal is concurrent with closure or whether an additional event that promotes membrane fusion occurs after the necessary removal of the LEP is currently unknown.
Don1, the first member of the LEP to be identified, was found through a screen for meiotically expressed coiled-coil proteins and was subsequently found to form a ring-like shape at the open end of the prospore membranes [7]. This novel localization suggested that there may be a protein structure at the open end of prospore membranes. Loss of DON1 does not result in any obvious phenotypes either in prospore membrane development or spore formation [7,25].
SSP1 and ADY3 were identified as members of the LEP through a Yeast Two-Hybrid screen for proteins associated with the meiotic outer plaque [18]. Though the meiotic outer plaque and the LEP are at opposite ends of the growing prospore membrane, Ssp1 and Ady3 both co-localize with Don1 at the leading edge during prospore membrane elongation, indicating that both are members of the LEP [18,19]. Loss of SSP1 results in a severe prospore membrane formation defect in which prospore membranes collapse on the nuclear envelope, rather than maintaining a tubular structure during elongation. The C-terminal region of Ssp1 appears to be required for the degradation of Ssp1, as truncations of the protein lead to persistent expression during sporulation [106]. ADY3 has not been shown to be integral for prospore membrane development but is required for the proper localization of Don1 to the LEP [18,19,99]. Ady3 is also required for the proper synthesis of the β-glucan and chitosan components of the spore wall [19] and proper mitochondrial segregation during sporulation [108].
The most recently discovered member of the LEP is Irc10. Identified through a visual screen of GFP-tagged ORFs expressed during meiosis, Irc10 also co-localizes with Don1. Loss of IRC10 alone results in no obvious phenotype. However, loss of both IRC10 and ADY3 phenocopies loss of SSP1, indicating that these two members of the LEP are collectively important for the formation of the whole complex [105]. Genetic analysis utilizing loss of function alleles for the LEP members suggests a model with an order of assembly for the LEP in which Ssp1 is present on the meiotic outer plaque at prospore membrane initiation, Ady3 is recruited by Ssp1 to the LEP, then as prospore membranes elongate, both Don1 and Irc10 are recruited to the LEP [18,19,105]. After anaphase II, Ssp1 degradation requires the meiotic specific activator of the APC/C, Ama1 [109]. Ssp1 may require the Sps1 kinase for its removal from the prospore membrane for timely prospore membrane closure [110].
The sporulation MAP Kinase Smk1 also localizes to the leading edge of the prospore membrane late in meiosis II, moving from the elongating prospore membrane to briefly co-localizing with the LEP around anaphase II [25]. Interestingly, the Ssp2 activator of Smk1 [111,112,113,114,115,116] shares a similar localization as Smk1 [25], suggesting that Smk1 may be active at the LEP before prospore membrane closure. The ability of Smk1 to localize to the leading edge requires ADY3 [25]. The localization of Smk1 without ADY3 resembles the localization of Ssp1-YFP (which does not fully complement), which localizes ectopically in discrete areas along the prospore membrane in both wild-type and the ady3irc10∆ double mutant [105]. SMK1 also plays a role in late prospore membrane development, as SMK1 is required for normal rounding of the prospore membrane [25]; rounding up of the prospore membrane correlates with the closure of the membrane and the cellularization of the spore [109,110]. The role SMK1 plays in prospore membrane development may be related to its localization at the leading edge, although the exact function of Smk1 at the leading edge is still unclear.
The LEP is an important structure required for proper prospore membrane development. However, the mechanism underlying LEP localization is still unclear as is the relationship between LEP removal and prospore membrane closure.

6. Exit from Meiosis II: Similarities and Differences with Mitotic Exit

From the perspective of cellular events, meiosis is more complex than mitosis. Al-though both meiotic and mitotic cells must undergo DNA replication, meiosis involves two rounds of chromosome segregation: homolog segregation in meiosis I and sister chromatid segregation in meiosis II. Furthermore, the goal of meiosis is not to make a faithful copy but to create haploid gametes; this process involves shuffling the genomic content through the random segregation of homologous chromosomes plus recombination among homolog pairs. These meiotic-specific events require distinct regulatory processes (reviewed in [2]). Although some of the machinery that controls the mitotic cell cycle, like the use of cyclin-dependent kinase regulation and the spindle apparatus are similar, the details of the regulation are distinct. For example, the mitotic cyclin CLB2 is not expressed in meiosis [117,118]. The only CDK in mitosis, Cdk1 (also known as Cdc28), is required for meiosis but acts in conjunction with the CDK-like kinase Ime2 [72,119,120,121] and utilizes some mitotic cyclins and not others; CDK regulation in meiosis has been recently reviewed [122].
In budding yeast, the mitotic division is an asymmetric event involving the formation of a bud, polarized growth, and the movement of the mitotic spindle into the newly formed bud before exit from mitosis. On the other hand, meiosis is symmetrical, with the four spores forming within the mother cell ascus. During mitosis, an actomyosin ring constricts for cytokinesis while prospore membrane closure is the cytokinetic event in meiosis. The prospore membranes round up and close after anaphase II; this process does not require the actin cytoskeleton [93].
This closure event marks the point when the four daughter cells have cellularized. Meiosis II happens quickly in the SK1 strain background, with anaphase II lasting about 10–20 min [53] and prospore membrane closure occurring soon after the completion of anaphase II [110]. Meiotic cytokinesis is regulated by the meiotic exit pathway [21,110,123] similar to how the end of mitosis is regulated by the mitotic exit network (sometimes known as “MEN”; reviewed in [20,124,125]) (Figure 3).
The mitotic exit network involves the activation of the Tem1 GTPase, which is activated at the spindle pole body and leads to the activation of Cdc15, a hippo-like kinase [126,127,128,129,130,131,132,133,134]. Cdc15 subsequently phosphorylates Nud1, another component of the SPB, which leads to the recruitment of the NDR-LATS kinase complex Dbf2-Mob1 [133,135,136,137]. NDR-LATS activity then leads to sustained release of the phosphatase Cdc14 from the nucleolus, which inhibits cyclin-dependent kinase (CDK) activity and promotes mitotic exit [138,139,140,141].
The Cdc14 phosphatase is an important effector of mitotic exit, reversing CDK-dependent phosphorylations and inhibiting further CDK activity so that cells return to G1 (reviewed in [142]). Cdc14 is regulated through sequestration in the nucleolus by its inhibitor Net1 until anaphase [139,143]. Two pathways control the release of Cdc14 from the nucleolus: the Cdc14 early anaphase release (FEAR) network (reviewed in [144]) and the mitotic exit network (MEN) (reviewed in [20]). The FEAR network leads to a transient release of Cdc14 in early anaphase and is partially regulated by mitotic CDKs [145,146,147]. The transient release of Cdc14 facilitates many anaphase-specific events, including nucleolar segregation and chromosome compaction [148,149,150]. Transiently released Cdc14 also promotes MEN activity by opposing the inactivation of MEN kinases by CDKs, ultimately leading to sustained Cdc14 release in late anaphase II [151,152,153]. In mitosis, Net1 is phosphorylated by the Dbf2-Mob1 kinase complex; this kinase complex is transported from the SPB to the nucleus after being phosphorylated by Cdc15 [125,154,155]. Cdc14 is also important for redistributing the GTPase Cdc42 during mitotic cytokinesis, ultimately resulting in proper cell size determination [156].
Cdc14 also affects mitotic progression by acting on an E3 ligase, the anaphase-promoting complex/cyclosome (APC/C), by dephosphorylating the late mitotic-activating subunit, Cdh1 [157]. Cdh1 activity is inhibited by Cdk1-dependent phosphorylations from the S phase until mitotic exit [158,159]. Cdh1-APC/C promotes mitotic exit through the degradation of securin, leading to the separation of sister chromatids, as well as the degradation of cyclins [160,161,162,163].
The meiotic exit pathway, which controls the timing of prospore membrane closure, requires the Cdc15 Hippo-like kinase but does not require the Tem1 GTPase or the Dbf2-Mob1 NDR/LATS kinase complex [21,164,165] (Figure 3). Instead, the Sps1 STE20 family GCKIII kinase acts downstream of Cdc15 [21]. The Nud1 SPB component, which acts as a scaffold for the mitotic exit components, is also not involved in the exit from meiosis II [165,166]. How Cdc15 is activated and whether there is a scaffold playing a role like Nud1 for the Cdc15-Sps1 pathway in meiosis is unclear. The specific role of the Cdc14 phosphatase in meiotic exit has also not been defined. Inactivating Cdc14 at metaphase II appears to block spore formation but does not seem to affect the disassembly of the meiosis II spindle [167]. The sustained release of Cdc14 from the nucleolus requires the Cdc15-Sps1 pathway [21] as well as the Hrr25 casein kinase 1 [161], although the significance of this release and the identity of which substrates require Cdc14 dephosphorylation during exit from meiosis II is currently unknown. The relationship between the Hrr25 kinase and the Cdc15-Sps1 pathway has also not been defined.
Like in mitosis, the APC/C is involved in regulating meiotic exit [109,110,168], although it utilizes a meiotic-specific activator, Ama1 [169], instead of Cdh1, as is used in mitosis. The switch in APC/C activators between mitosis and meiosis occurs because Ime2 phosphorylates many Cdk1 substrates (including Cdh1), using different phosphosites than Cdk1; these Ime2-specific phosphorylations are resistant to dephosphorylation by Cdc14, leading to the suppression of mitotic Cdk1 targets throughout meiosis [170]. During meiotic prophase, Ama1 is needed for the proteolysis of mitotic M-phase regulators like Ndd1, Cdc5, and Clb4 [171]. Ama1 activity is repressed by Mnd2, another APC/C subunit, until meiosis II [172,173]. Mnd2 is removed from the APC/C in meiosis II, at which point Ama1 targets the APC/C to degrade various proteins in sporulation including another activator of the APC/C, Cdc20 [168]. For exit from meiosis II, Ama1 (and presumably APC/C activity) is required for several processes including removal of the LEP coat, prospore membrane closure, and spindle disassembly [109,110,161,168,174,175,176].

7. Controlling the Timing of Meiotic Cytokinesis

The Cdc15-Sps1 pathway and the Ama1-APC/C work in parallel to regulate the proper timing of meiotic cytokinesis, as the removal of both pathways leads to continued prospore membrane elongation without closure [110]. Cells lacking AMA1 have a partial prospore membrane closure defect, with about 30% of ama1∆ cells closing their prospore membranes [109,110]. Cells lacking SPS1 also have a partial prospore membrane closure defect, with about 70% of sps1∆ cells rounding up and closing their prospore membranes [110]. Both SPS1 and AMA1 affect the timing of prospore membrane closure, as loss of each results in a delay of at least an hour without affecting the timing of prospore membrane initiation or progression through meiosis [21,110]. Strikingly, the sps1ama1∆ double mutants have an additive effect: all cells form hyperelongated prospore membranes that remain open and hyperelongated [110].
While the mechanism by which SPS1 and AMA1 affect prospore membrane closure timing is unclear, both affect the removal and/or degradation of LEP components. Sps1 and the LEP component Ssp1 coimmunoprecipitate, indicating the two proteins complex [109]. AMA1 is required for the normal degradation of Ssp1, as seen by immunoblotting [109]. Similarly, fluorescent microscopy indicates that both AMA1 and SPS1 are required for the timely removal of Don1 from the LEP [110].

8. Meiotic Exit Requires the Coordination of Prospore Membrane Closure and Spindle Disassembly

The closure of the prospore membrane and cellularization of the newly formed spore requires the coordination of other cellular events such as spindle disassembly, organelle segregation, nuclear envelope separation, disassembly of the meiotic outer plaque (to disassociate the nucleus from the new plasma membrane), and recycling of the nuclear pore complex through the gametogenesis uninherited nuclear compartment (known as the “GUNC”) [106,177,178,179,180,181,182]. Little is known about the coordination of these events, although the coordination of prospore membrane closure and spindle disassembly are both regulated by the Cdc15-Sps1 pathway working in conjunction with Ama1-APC/C [21,110,177].
The first indication that both the Cdc15-Sps1 and Ama1-APC/C pathways were involved in coordinating spindle disassembly and prospore membrane closure was that all mutants in these pathways produce hyperelongated prospore membranes and have meiosis II spindle disassembly defects. In addition to the hyperelongated prospore membranes seen in cells lacking CDC15, SPS1, or AMA1, these mutant cells also had persistent tubulin fragments that remained after anaphase II. This observation led to the question of whether the delay in the closure of the prospore membrane seen in these mutants was due to a problem with meiosis II spindle disassembly. In wild-type cells, the spindle disassembles before the prospore membrane closes (Figure 4).
Interestingly, disassembly of the meiosis II spindle in cells with hyperelongated prospore membranes does not cause prospore membrane closure, consistent with the idea that these two events are regulated separately [177]. In wild-type cells, the addition of a benomyl–nocodozole cocktail that depolymerized microtubules after anaphase II resulted in normal prospore membrane closure shortly after the spindles were disassembled. In contrast, although the addition of the benomyl–nocodozole cocktail to cells lacking either SPS1 and AMA1 completely disassembled the spindles, the prospore membranes continued to elongate for up to an hour after the addition of the benomyl–nocodozole cocktail, indicating that the hyperelongated prospore membranes are not due to persistent intact microtubule fragments seen in these mutants.
Although both prospore membrane closure and spindle disassembly are regulated by Cdc15-Sps1 and Ama1-APC/C [21,109,110,164,165,177], these pathways likely act on distinct downstream targets for each process and thus, the activity of these pathways provides the coordination of the two events. Potential targets for Cdc15-Sps1 and Ama1-APC/C for spindle disassembly include the microtubule-binding proteins Bim1, Cin8, and Ase1, whose localization in meiosis II is impacted by the loss of either SPS1 or AMA1 [177]. Ssp1 and other components of the LEP may be important targets for prospore membrane closure [106,109,110]. Cdc14, an effector for mitotic exit, may also play a role in meiotic exit, as the Cdc15-Sps1 pathway and the Hrr25 casein kinase affect Cdc14 release from the nucleolus in anaphase II [21,167]. However, the specific requirement for Cdc14 release in meiosis II has not been determined.

9. Conclusions

Progress has been made in understanding the mechanisms underlying meiotic cytokinesis in S. cerevisiae, but there are still important areas to explore. Sporulation involves a diverse array of cell biological processes, including the creation of a new membrane, microtubule spindle assembly and disassembly, and organellar segregation. Such processes are important in all eukaryotes and operate across a variety of cellular contexts, so insights obtained from the study of sporulation can help inform our understanding of fundamental cell biological processes more broadly.
Furthermore, a better understanding of this cellularization process can inform our understanding of other systems. Although prospore membrane closure occurs using mechanisms distinct from typical animal cell cytokinesis, atypical cellularization events occur in both mitosis and meiosis across species from insects to mammals [183]. The fission yeast Schizosaccharomyces pombe (which is thought to have diverged from the budding yeast S. cerevisiae 350 million years ago [184]) also produces spores though the production of a forespore membrane that grows from a meiotically modified SPB [185,186]. However, unlike the budding yeast, the closure of the forespore membrane in S. pombe requires a meiotic actin ring [187]. The cellularization of the nuclei in early Drosophila development is one of the better-studied examples of atypical cytokinesis and involves extensive membrane remodeling and growth to surround the mitotic nuclei (reviewed in [188]). In mammalian females, the germline undergoes incomplete cytokinesis through intercellular bridges, resulting in a syncytium structure [189]. Better understanding of prospore membrane closure as the meiotic cytokinetic event may inform our understanding of other atypical cytokinetic events. Thus, the further examination of meiotic cytokinesis using the powerful tools available in S. cerevisiae is likely to have impacts well beyond the creation of the spore in budding yeast.

Funding

M.D., X.M. and L.S.H. were funded in part by a grant from the National Institutes of Health, NIGMS GM086805; M.D. was also supported in part by a University of Massachusetts Boston CSM Dean’s Fellowship.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

We thank Kailee Charland, Erin Klee, and Emily Vu for comments on this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Neiman, A.M. Sporulation in the Budding Yeast Saccharomyces cerevisiae. Genetics 2011, 189, 737–765. [Google Scholar] [CrossRef] [PubMed]
  2. Börner, G.V.; Hochwagen, A.; MacQueen, A.J. Meiosis in budding yeast. Genetics 2023, 225, iyad125. [Google Scholar] [CrossRef]
  3. Mitchell, A.P. Control of meiotic gene expression in Saccharomyces cerevisiae. Microbiol. Rev. 1994, 58, 56–70. [Google Scholar] [CrossRef] [PubMed]
  4. Moens, P.B.; Rapport, E. Spindles, spindle plaques, and meiosis in the yeast Saccharomyces cerevisiae (Hansen). J. Cell Biol. 1971, 50, 344–361. [Google Scholar] [CrossRef] [PubMed]
  5. Byers, B. Cytology of the Yeast Life Cycle; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 1981; pp. 59–96. [Google Scholar]
  6. Neiman, A.M. Prospore Membrane Formation Defines a Developmentally Regulated Branch of the Secretory Pathway in Yeast. J. Cell Biol. 1998, 140, 29–37. [Google Scholar] [CrossRef] [PubMed]
  7. Knop, M.; Strasser, K. Role of the spindle pole body of yeast in mediating assembly of the prospore membrane during meiosis. EMBO J. 2000, 19, 3657–3667. [Google Scholar] [CrossRef] [PubMed]
  8. Bajgier, B.K.; Malzone, M.; Nickas, M.; Neiman, A.M. SPO21 Is Required for Meiosis-specific Modification of the Spindle Pole Body in Yeast. Mol. Biol. Cell 2001, 12, 1611–1621. [Google Scholar] [CrossRef]
  9. Nickas, M.E.; Schwartz, C.; Neiman, A.M. Ady4p and Spo74p Are Components of the Meiotic Spindle Pole Body That Promote Growth of the Prospore Membrane in Saccharomyces cerevisiae. Eukaryot. Cell 2003, 2, 431–445. [Google Scholar] [CrossRef]
  10. Deng, C.; Saunders, W.S. ADY1, A Novel Gene Required for Prospore Membrane Formation at Selected Spindle Poles in Saccharomyces cerevisiae. Mol. Biol. Cell 2001, 12, 2646–2659. [Google Scholar] [CrossRef]
  11. Jungbluth, M.; Mösch, H.-U.; Taxis, C. Acetate Regulation of Spore Formation Is under the Control of the Ras/Cyclic AMP/Protein Kinase A Pathway and Carbon Dioxide in Saccharomyces cerevisiae. Eukaryot. Cell 2012, 11, 1021–1032. [Google Scholar] [CrossRef]
  12. Mathieson, E.M.; Suda, Y.; Nickas, M.; Snydsman, B.; Davis, T.N.; Muller, E.G.D.; Neiman, A.M. Vesicle Docking to the Spindle Pole Body Is Necessary to Recruit the Exocyst During Membrane Formation in Saccharomyces cerevisiae. Mol. Biol. Cell 2010, 21, 3693–3707. [Google Scholar] [CrossRef]
  13. Schaerer, F.; Morgan, G.; Winey, M.; Philippsen, P. Cnm67p Is a Spacer Protein of the Saccharomyces cerevisiae Spindle Pole Body Outer Plaque. Mol. Biol. Cell 2001, 12, 2519–2533. [Google Scholar] [CrossRef]
  14. Klapholz, S.; Esposito, R.E. Recombination and Chromosome Segregation During the Single Division Meiosis in spo12–1 and spo13–1 Diploids. Genetics 1980, 96, 589–611. [Google Scholar] [CrossRef]
  15. Nickas, M.E.; Diamond, A.E.; Yang, M.-J.; Neiman, A.M. Regulation of Spindle Pole Function by an Intermediary Metabolite. Mol. Biol. Cell 2004, 15, 2606–2616. [Google Scholar] [CrossRef] [PubMed]
  16. Taxis, C.; Keller, P.; Kavagiou, Z.; Jensen, L.J.; Colombelli, J.; Bork, P.; Stelzer, E.H.K.; Knop, M. Spore number control and breeding in Saccharomyces cerevisiae. J. Cell Biol. 2005, 171, 627–640. [Google Scholar] [CrossRef] [PubMed]
  17. Renicke, C.; Allmann, A.-K.; Lutz, A.P.; Heimerl, T.; Taxis, C. The Mitotic Exit Network Regulates Spindle Pole Body Selection During Sporulation of Saccharomyces cerevisiae. Genetics 2017, 206, 919–937. [Google Scholar] [CrossRef]
  18. Moreno-Borchart, A.C.; Strasser, K.; Finkbeiner, M.G.; Shevchenko, A.; Shevchenko, A.; Knop, M. Prospore membrane formation linked to the leading edge protein (LEP) coat assembly. EMBO J. 2001, 20, 6946–6957. [Google Scholar] [CrossRef]
  19. Nickas, M.E.; Neiman, A.M. Ady3p Links Spindle Pole Body Function to Spore Wall Synthesis in Saccharomyces cerevisiae. Genetics 2002, 160, 1439–1450. [Google Scholar] [CrossRef] [PubMed]
  20. Bardin, A.J.; Amon, A. MEN and SIN: What’s the difference? Nat. Rev. Mol. Cell Biol. 2001, 2, 815–826. [Google Scholar] [CrossRef]
  21. Paulissen, S.M.; Hunt, C.A.; Seitz, B.C.; Slubowski, C.J.; Yu, Y.; Mucelli, X.; Truong, D.; Wallis, Z.; Nguyen, H.T.; Newman-Toledo, S.; et al. A Noncanonical Hippo Pathway Regulates Spindle Disassembly and Cytokinesis During Meiosis in Saccharomyces cerevisiae. Genetics 2020, 216, 447–462. [Google Scholar] [CrossRef]
  22. Hsu, T.-H.; Chen, R.-H.; Cheng, Y.-H.; Wang, C.-W. Lipid droplets are central organelles for meiosis II progression during yeast sporulation. Mol. Biol. Cell 2017, 28, 440–451. [Google Scholar] [CrossRef]
  23. Coluccio, A.; Bogengruber, E.; Conrad, M.N.; Dresser, M.E.; Briza, P.; Neiman, A.M. Morphogenetic Pathway of Spore Wall Assembly in Saccharomyces cerevisiae. Eukaryot. Cell 2004, 3, 1464–1475. [Google Scholar] [CrossRef] [PubMed]
  24. Sing, T.L.; Brar, G.A.; Ünal, E. Gametogenesis: Exploring an Endogenous Rejuvenation Program to Understand Cellular Aging and Quality Control. Annu. Rev. Genet. 2022, 56, 89–112. [Google Scholar] [CrossRef]
  25. Durant, M.; Roesner, J.M.; Mucelli, X.; Slubowski, C.J.; Klee, E.; Seitz, B.C.; Wallis, Z.; Huang, L.S. The Smk1 MAPK and Its Activator, Ssp2, Are Required for Late Prospore Membrane Development in Sporulating Saccharomyces cerevisiae. J. Fungi 2021, 7, 53. [Google Scholar] [CrossRef]
  26. Nakanishi, H.; de los Santos, P.; Neiman, A.M. Positive and Negative Regulation of a SNARE Protein by Control of Intracellular Localization. Mol. Biol. Cell 2004, 15, 1802–1815. [Google Scholar] [CrossRef] [PubMed]
  27. Parodi, E.M.; Baker, C.S.; Tetzlaff, C.; Villahermosa, S.; Huang, L.S. SPO71 Mediates Prospore Membrane Size and Maturation in Saccharomyces cerevisiae. Eukaryot. Cell 2012, 11, 1191–1200. [Google Scholar] [CrossRef]
  28. Lynn, R.R.; Magee, P.T. Development of the spore wall during ascospore formation in Saccharomyces cerevisiae. J. Cell Biol. 1970, 44, 688–692. [Google Scholar] [CrossRef]
  29. Briza, P.; Ellinger, A.; Winkler, G.; Breitenbach, M. Characterization of a DL-dityrosine-containing macromolecule from yeast ascospore walls. J. Biol. Chem. 1990, 265, 15118–15123. [Google Scholar] [CrossRef] [PubMed]
  30. Lesage, G.; Bussey, H. Cell Wall Assembly in Saccharomyces cerevisiae. Microbiol. Mol. Biol. Rev. 2006, 70, 317–343. [Google Scholar] [CrossRef]
  31. Tachikawa, H.; Bloecher, A.; Tatchell, K.; Neiman, A.M. A Gip1p–Glc7p phosphatase complex regulates septin organization and spore wall formation. J. Cell Biol. 2001, 155, 797–808. [Google Scholar] [CrossRef]
  32. Huang, L.S.; Doherty, H.K.; Herskowitz, I. The Smk1p MAP kinase negatively regulates Gsc2p, a 1,3-β-glucan synthase, during spore wall morphogenesis in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 2005, 102, 12431–12436. [Google Scholar] [CrossRef]
  33. Lin, C.P.-C.; Kim, C.; Smith, S.O.; Neiman, A.M. A Highly Redundant Gene Network Controls Assembly of the Outer Spore Wall in S. cerevisiae. PLoS Genet. 2013, 9, e1003700. [Google Scholar] [CrossRef] [PubMed]
  34. Hoffmann, R.; Grabińska, K.; Guan, Z.; Sessa, W.C.; Neiman, A.M. Long-Chain Polyprenols Promote Spore Wall Formation in Saccharomyces cerevisiae. Genetics 2017, 207, 1371–1386. [Google Scholar] [CrossRef] [PubMed]
  35. Eastwood, M.D.; Cheung, S.W.T.; Lee, K.Y.; Moffat, J.; Meneghini, M.D. Developmentally Programmed Nuclear Destruction during Yeast Gametogenesis. Dev. Cell 2012, 23, 35–44. [Google Scholar] [CrossRef] [PubMed]
  36. Eastwood, M.D.; Meneghini, M.D. Developmental Coordination of Gamete Differentiation with Programmed Cell Death in Sporulating Yeast. Eukaryot. Cell 2015, 14, 858–867. [Google Scholar] [CrossRef] [PubMed]
  37. Orlean, P. Architecture and Biosynthesis of the Saccharomyces cerevisiae Cell Wall. Genetics 2012, 192, 775–818. [Google Scholar] [CrossRef] [PubMed]
  38. Rij, N.J.W.K. Ultrastructure of the ascospores of the new yeast genus Sporopachydermia Rodrigues de Miranda. Antonie Leeuwenhoek 1978, 44, 451–456. [Google Scholar] [CrossRef]
  39. Smits, G.J.; van den Ende, H.; Klis, F.M. Differential regulation of cell wall biogenesis during growth and development in yeast. Microbiology 2001, 147, 781–794. [Google Scholar] [CrossRef]
  40. Coluccio, A.E.; Rodriguez, R.K.; Kernan, M.J.; Neiman, A.M. The Yeast Spore Wall Enables Spores to Survive Passage through the Digestive Tract of Drosophila. PLoS ONE 2008, 3, e2873. [Google Scholar] [CrossRef]
  41. Briza, P.; Ellinger, A.; Winkler, G.; Breitenbach, M. Chemical composition of the yeast ascospore wall. The second outer layer consists of chitosan. J. Biol. Chem. 1988, 263, 11569–11574. [Google Scholar] [CrossRef]
  42. Pammer, M.; Briza, P.; Ellinger, A.; Schuster, T.; Stucka, R.; Feldmann, H.; Breitenbach, M. DIT101 (CSD2, CAL1), a cell cycle-regulated yeast gene required for synthesis of chitin in cell walls and chitosan in spore walls. Yeast 1992, 8, 1089–1099. [Google Scholar] [CrossRef]
  43. Barve, G.; Manjithaya, R. Cross-talk between autophagy and sporulation in Saccharomyces cerevisiae. Yeast 2021, 38, 401–413. [Google Scholar] [CrossRef]
  44. Deutschbauer, A.M.; Williams, R.M.; Chu, A.M.; Davis, R.W. Parallel phenotypic analysis of sporulation and postgermination growth in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 2002, 99, 15530–15535. [Google Scholar] [CrossRef] [PubMed]
  45. Enyenihi, A.H.; Saunders, W.S. Large-Scale Functional Genomic Analysis of Sporulation and Meiosis in Saccharomyces cerevisiae. Genetics 2003, 163, 47–54. [Google Scholar] [CrossRef] [PubMed]
  46. Piekarska, I.; Kucharczyk, R.; Mickowska, B.; Rytka, J.; Rempola, B. Mutants of the Saccharomyces cerevisiae VPS genes CCZ1 and YPT7 are blocked in different stages of sporulation. Eur. J. Cell Biol. 2010, 89, 780–787. [Google Scholar] [CrossRef] [PubMed]
  47. Sarkar, S.; Dalgaard, J.Z.; Millar, J.B.A.; Arumugam, P. The Rim15-Endosulfine-PP2ACdc55 Signalling Module Regulates Entry into Gametogenesis and Quiescence via Distinct Mechanisms in Budding Yeast. PLoS Genet. 2014, 10, e1004456. [Google Scholar] [CrossRef] [PubMed]
  48. Straub, M.; Bredschneider, M.; Thumm, M. AUT3, a serine/threonine kinase gene, is essential for autophagocytosis in Saccharomyces cerevisiae. J. Bacteriol. 1997, 179, 3875–3883. [Google Scholar] [CrossRef] [PubMed]
  49. Wen, F.-P.; Guo, Y.-S.; Hu, Y.; Liu, W.-X.; Wang, Q.; Wang, Y.-T.; Yu, H.-Y.; Tang, C.-M.; Yang, J.; Zhou, T.; et al. Distinct temporal requirements for autophagy and the proteasome in yeast meiosis. Autophagy 2016, 12, 671–688. [Google Scholar] [CrossRef] [PubMed]
  50. Wang, F.; Zhang, R.; Feng, W.; Tsuchiya, D.; Ballew, O.; Li, J.; Denic, V.; Lacefield, S. Autophagy of an Amyloid-like Translational Repressor Regulates Meiotic Exit. Dev. Cell 2020, 52, 141–151.e5. [Google Scholar] [CrossRef] [PubMed]
  51. Berchowitz, L.E.; Gajadhar, A.S.; van Werven, F.J.; De Rosa, A.A.; Samoylova, M.L.; Brar, G.A.; Xu, Y.; Xiao, C.; Futcher, B.; Weissman, J.S.; et al. A developmentally regulated translational control pathway establishes the meiotic chromosome segregation pattern. Gene Dev. 2013, 27, 2147–2163. [Google Scholar] [CrossRef]
  52. Berchowitz, L.E.; Kabachinski, G.; Walker, M.R.; Carlile, T.M.; Gilbert, W.V.; Schwartz, T.U.; Amon, A. Regulated Formation of an Amyloid-like Translational Repressor Governs Gametogenesis. Cell 2015, 163, 406–418. [Google Scholar] [CrossRef]
  53. Carpenter, K.; Bell, R.B.; Yunus, J.; Amon, A.; Berchowitz, L.E. Phosphorylation-Mediated Clearance of Amyloid-like Assemblies in Meiosis. Dev. Cell 2018, 45, 392–405.e6. [Google Scholar] [CrossRef]
  54. Zhang, R.; Feng, W.; Qian, S.; Li, S.; Wang, F. Regulation of Rim4 distribution, function, and stability during meiosis by PKA, Cdc14, and 14-3-3 proteins. Cell Rep. 2023, 42, 113052. [Google Scholar] [CrossRef]
  55. Zhang, R.; Feng, W.; Qian, S.; Wang, F. Autophagy-mediated surveillance of Rim4-mRNA interaction safeguards programmed meiotic translation. Cell Rep. 2023, 42, 113051. [Google Scholar] [CrossRef]
  56. Feng, W.; Argüello-Miranda, O.; Qian, S.; Wang, F. Cdc14 spatiotemporally dephosphorylates Atg13 to activate autophagy during meiotic divisions. J. Cell Biol. 2022, 221, e202107151. [Google Scholar] [CrossRef]
  57. Kassir, Y.; Granot, D.; Simchen, G. IME1, a positive regulator gene of meiosis in S. cerevisiae. Cell 1988, 52, 853–862. [Google Scholar] [CrossRef]
  58. Mitchell, A.P.; Driscoll, S.E.; Smith, H.E. Positive Control of Sporulation-Specific Genes by the IME1 and IME2 Products in Saccharomyces cerevisiae. Mol. Cell. Biol. 1990, 10, 2104–2110. [Google Scholar] [CrossRef]
  59. Smith, H.E.; Su, S.S.Y.; Neigeborn, L.; Driscoll, S.E.; Mitchell, A.P. Role of IME1 Expression in Regulation of Meiosis in Saccharomyces cerevisiae. Mol. Cell. Biol. 1990, 10, 6103–6113. [Google Scholar] [CrossRef] [PubMed]
  60. Pak, J.; Segall, J. Regulation of the Premiddle and Middle Phases of Expression of the NDT80 Gene during Sporulation of Saccharomyces cerevisiae. Mol. Cell. Biol. 2002, 22, 6417–6429. [Google Scholar] [CrossRef] [PubMed]
  61. Chu, S.; Herskowitz, I. Gametogenesis in Yeast Is Regulated by a Transcriptional Cascade Dependent on Ndt80. Mol. Cell 1998, 1, 685–696. [Google Scholar] [CrossRef] [PubMed]
  62. Xu, L.; Ajimura, M.; Padmore, R.; Klein, C.; Kleckner, N. NDT80, a Meiosis-Specific Gene Required for Exit from Pachytene in Saccharomyces cerevisiae. Mol. Cell. Biol. 1995, 15, 6572–6581. [Google Scholar] [CrossRef]
  63. Winter, E. The Sum1/Ndt80 Transcriptional Switch and Commitment to Meiosis in Saccharomyces cerevisiae. Microbiol. Mol. Biol. Rev. 2012, 76, 1–15. [Google Scholar] [CrossRef]
  64. Xie, J.; Pierce, M.; Gailus-Durner, V.; Wagner, M.; Winter, E.; Vershon, A.K. Sum1 and Hst1 repress middle sporulation-specific gene expression during mitosis in Saccharomyces cerevisiae. EMBO J. 1999, 18, 6448–6454. [Google Scholar] [CrossRef]
  65. Lindgren, A.; Bungard, D.; Pierce, M.; Xie, J.; Vershon, A.; Winter, E. The pachytene checkpoint in Saccharomyces cerevisiae requires the Sum1 transcriptional repressor. EMBO J. 2000, 19, 6489–6497. [Google Scholar] [CrossRef]
  66. Pierce, M.; Benjamin, K.R.; Montano, S.P.; Georgiadis, M.M.; Winter, E.; Vershon, A.K. Sum1 and Ndt80 Proteins Compete for Binding to Middle Sporulation Element Sequences That Control Meiotic Gene Expression. Mol. Cell. Biol. 2003, 23, 4814–4825. [Google Scholar] [CrossRef]
  67. Sopko, R.; Raithatha, S.; Stuart, D. Phosphorylation and Maximal Activity of Saccharomyces cerevisiae Meiosis-Specific Transcription Factor Ndt80 Is Dependent on Ime2. Mol. Cell. Biol. 2002, 22, 7024–7040. [Google Scholar] [CrossRef]
  68. Lo, H.-C.; Wan, L.; Rosebrock, A.; Futcher, B.; Hollingsworth, N.M. Cdc7-Dbf4 Regulates NDT80 Transcription as Well as Reductional Segregation during Budding Yeast Meiosis. Mol. Biol. Cell 2008, 19, 4956–4967. [Google Scholar] [CrossRef]
  69. Moore, M.; Shin, M.E.; Bruning, A.; Schindler, K.; Vershon, A.; Winter, E. Arg-Pro-X-Ser/Thr Is a Consensus Phosphoacceptor Sequence for the Meiosis-Specific Ime2 Protein Kinase in Saccharomyces cerevisiae. Biochemistry 2007, 46, 271–278. [Google Scholar] [CrossRef] [PubMed]
  70. Ahmed, N.T.; Bungard, D.; Shin, M.E.; Moore, M.; Winter, E. The Ime2 Protein Kinase Enhances the Disassociation of the Sum1 Repressor from Middle Meiotic Promoters. Mol. Cell. Biol. 2009, 29, 4352–4362. [Google Scholar] [CrossRef] [PubMed]
  71. Corbi, D.; Sunder, S.; Weinreich, M.; Skokotas, A.; Johnson, E.S.; Winter, E. Multisite Phosphorylation of the Sum1 Transcriptional Repressor by S-Phase Kinases Controls Exit from Meiotic Prophase in Yeast. Mol. Cell. Biol. 2014, 34, 2249–2263. [Google Scholar] [CrossRef] [PubMed]
  72. Benjamin, K.R.; Zhang, C.; Shokat, K.M.; Herskowitz, I. Control of landmark events in meiosis by the CDK Cdc28 and the meiosis-specific kinase Ime2. Gene Dev. 2003, 17, 1524–1539. [Google Scholar] [CrossRef] [PubMed]
  73. Brar, G.A.; Yassour, M.; Friedman, N.; Regev, A.; Ingolia, N.T.; Weissman, J.S. High-Resolution View of the Yeast Meiotic Program Revealed by Ribosome Profiling. Science 2012, 335, 552–557. [Google Scholar] [CrossRef] [PubMed]
  74. Jin, L.; Zhang, K.; Sternglanz, R.; Neiman, A.M. Predicted RNA Binding Proteins Pes4 and Mip6 Regulate mRNA Levels, Translation, and Localization during Sporulation in Budding Yeast. Mol. Cell. Biol. 2017, 37, e00408-16. [Google Scholar] [CrossRef] [PubMed]
  75. Nunez, G.; Zhang, K.; Mogbheli, K.; Hollingsworth, N.M.; Neiman, A.M. Recruitment of the lipid kinase Mss4 to the meiotic spindle pole promotes prospore membrane formation in Saccharomyces cerevisiae. Mol. Biol. Cell 2023, 34, ar33. [Google Scholar] [CrossRef] [PubMed]
  76. Park, J.-S.; Neiman, A.M. VPS13 regulates membrane morphogenesis during sporulation in Saccharomyces cerevisiae. J. Cell Sci. 2012, 125, 3004–3011. [Google Scholar] [CrossRef] [PubMed]
  77. Park, J.-S.; Okumura, Y.; Tachikawa, H.; Neiman, A.M. SPO71 Encodes a Developmental Stage-Specific Partner for Vps13 in Saccharomyces cerevisiae. Eukaryot. Cell 2013, 12, 1530–1537. [Google Scholar] [CrossRef] [PubMed]
  78. Parodi, E.M.; Roesner, J.M.; Huang, L.S. SPO73 and SPO71 Function Cooperatively in Prospore Membrane Elongation during Sporulation in Saccharomyces cerevisiae. PLoS ONE 2015, 10, e0143571. [Google Scholar] [CrossRef]
  79. Okumura, Y.; Nakamura, T.S.; Tanaka, T.; Inoue, I.; Suda, Y.; Takahashi, T.; Nakanishi, H.; Nakamura, S.; Gao, X.-D.; Tachikawa, H. The Dysferlin Domain-Only Protein, Spo73, Is Required for Prospore Membrane Extension in Saccharomyces cerevisiae. mSphere 2015, 1, e00038-15. [Google Scholar] [CrossRef]
  80. Park, J.-S.; Thorsness, M.K.; Policastro, R.; McGoldrick, L.L.; Hollingsworth, N.M.; Thorsness, P.E.; Neiman, A.M. Yeast Vps13 promotes mitochondrial function and is localized at membrane contact sites. Mol. Biol. Cell 2016, 27, 2435–2449. [Google Scholar] [CrossRef]
  81. Nakamura, T.S.; Suda, Y.; Muneshige, K.; Fujieda, Y.; Okumura, Y.; Inoue, I.; Tanaka, T.; Takahashi, T.; Nakanishi, H.; Gao, X.-D.; et al. Suppression of Vps13 adaptor protein mutants reveals a central role for PI4P in regulating prospore membrane extension. PLoS Genet. 2021, 17, e1009727. [Google Scholar] [CrossRef]
  82. Rabitsch, K.P.; Tóth, A.; Gálová, M.; Schleiffer, A.; Schaffner, G.; Aigner, E.; Rupp, C.; Penkner, A.M.; Moreno-Borchart, A.C.; Primig, M.; et al. A screen for genes required for meiosis and spore formation based on whole-genome expression. Curr. Biol. 2001, 11, 1001–1009. [Google Scholar] [CrossRef] [PubMed]
  83. Yu, J.W.; Mendrola, J.M.; Audhya, A.; Singh, S.; Keleti, D.; DeWald, D.B.; Murray, D.; Emr, S.D.; Lemmon, M.A. Genome-Wide Analysis of Membrane Targeting by S. cerevisiae Pleckstrin Homology Domains. Mol. Cell 2004, 13, 677–688. [Google Scholar] [CrossRef] [PubMed]
  84. Dziurdzik, S.K.; Conibear, E. The Vps13 Family of Lipid Transporters and Its Role at Membrane Contact Sites. Int. J. Mol. Sci. 2021, 22, 2905. [Google Scholar] [CrossRef] [PubMed]
  85. Leonzino, M.; Reinisch, K.M.; Camilli, P.D. Insights into VPS13 properties and function reveal a new mechanism of eukaryotic lipid transport. Biochim. Biophys. Acta (BBA) Mol. Cell Biol. Lipids 2021, 1866, 159003. [Google Scholar] [CrossRef] [PubMed]
  86. Lang, A.B.; Peter, A.T.J.; Walter, P.; Kornmann, B. ER–mitochondrial junctions can be bypassed by dominant mutations in the endosomal protein Vps13. J. Cell Biol. 2015, 210, 883–890. [Google Scholar] [CrossRef] [PubMed]
  87. Bankaitis, V.A.; Johnson, L.M.; Emr, S.D. Isolation of yeast mutants defective in protein targeting to the vacuole. Proc. Natl. Acad. Sci. USA 1986, 83, 9075–9079. [Google Scholar] [CrossRef] [PubMed]
  88. De, M.; Oleskie, A.N.; Ayyash, M.; Dutta, S.; Mancour, L.; Abazeed, M.E.; Brace, E.J.; Skiniotis, G.; Fuller, R.S. The Vps13p–Cdc31p complex is directly required for TGN late endosome transport and TGN homotypic fusion. J. Cell Biol. 2017, 216, 425–439. [Google Scholar] [CrossRef]
  89. Li, P.; Lees, J.A.; Lusk, C.P.; Reinisch, K.M. Cryo-EM reconstruction of a VPS13 fragment reveals a long groove to channel lipids between membranes. J. Cell Biol. 2020, 219, e202001161. [Google Scholar] [CrossRef]
  90. Kumar, N.; Leonzino, M.; Hancock-Cerutti, W.; Horenkamp, F.A.; Li, P.; Lees, J.A.; Wheeler, H.; Reinisch, K.M.; Camilli, P.D. VPS13A and VPS13C are lipid transport proteins differentially localized at ER contact sites. J. Cell Biol. 2018, 217, 3625–3639. [Google Scholar] [CrossRef]
  91. Lees, J.A.; Reinisch, K.M. Inter-organelle lipid transfer: A channel model for Vps13 and chorein-N motif proteins. Curr. Opin. Cell Biol. 2020, 65, 66–71. [Google Scholar] [CrossRef]
  92. Fares, H.; Goetsch, L.; Pringle, J.R. Identification of a developmentally regulated septin and involvement of the septins in spore formation in Saccharomyces cerevisiae. J. Cell Biol. 1996, 132, 399–411. [Google Scholar] [CrossRef]
  93. Taxis, C.; Maeder, C.; Reber, S.; Rathfelder, N.; Miura, K.; Greger, K.; Stelzer, E.H.K.; Knop, M. Dynamic Organization of the Actin Cytoskeleton During Meiosis and Spore Formation in Budding Yeast. Traffic 2006, 7, 1628–1642. [Google Scholar] [CrossRef] [PubMed]
  94. Chu, S.; DeRisi, J.; Eisen, M.; Mulholland, J.; Botstein, D.; Brown, P.O.; Herskowitz, I. The Transcriptional Program of Sporulation in Budding Yeast. Science 1998, 282, 699–705. [Google Scholar] [CrossRef]
  95. McMurray, M.A.; Thorner, J. Septin Stability and Recycling during Dynamic Structural Transitions in Cell Division and Development. Curr. Biol. 2008, 18, 1203–1208. [Google Scholar] [CrossRef]
  96. Virgilio, C.D.; DeMarini, D.J.; Pringle, J.R. SPR28, a sixth member of the septin gene family in Saccharomyces cerevisiae that is expressed specifically in sporulating cells. Microbiology 1996, 142, 2897–2905. [Google Scholar] [CrossRef] [PubMed]
  97. Garcia, G.; Finnigan, G.C.; Heasley, L.R.; Sterling, S.M.; Aggarwal, A.; Pearson, C.G.; Nogales, E.; McMurray, M.A.; Thorner, J. Assembly, molecular organization, and membrane-binding properties of development-specific septins. J. Cell Biol. 2016, 212, 515–529. [Google Scholar] [CrossRef]
  98. Hussain, A.; Nguyen, V.T.; Reigan, P.; McMurray, M. Evolutionary degeneration of septins into pseudoGTPases: Impacts on a hetero-oligomeric assembly interface. Front. Cell Dev. Biol. 2023, 11, 1296657. [Google Scholar] [CrossRef] [PubMed]
  99. Heasley, L.R.; McMurray, M.A. Roles of septins in prospore membrane morphogenesis and spore wall assembly in Saccharomyces cerevisiae. Mol. Biol. Cell 2016, 27, 442–450. [Google Scholar] [CrossRef]
  100. Padmore, R.; Cao, L.; Kleckner, N. Temporal comparison of recombination and synaptonemal complex formation during meiosis in S. cerevisiae. Cell 1991, 66, 1239–1256. [Google Scholar] [CrossRef]
  101. Heasley, L.R.; Singer, E.; Cooperman, B.J.; McMurray, M.A. Saccharomyces spores are born prepolarized to outgrow away from spore–spore connections and penetrate the ascus wall. Yeast 2021, 38, 90–101. [Google Scholar] [CrossRef]
  102. Joseph-Strauss, D.; Zenvirth, D.; Simchen, G.; Barkai, N. Spore germination in Saccharomyces cerevisiae: Global gene expression patterns and cell cycle landmarks. Genome Biol. 2007, 8, R241. [Google Scholar] [CrossRef]
  103. Nakamura, T.S.; Numajiri, Y.; Okumura, Y.; Hidaka, J.; Tanaka, T.; Inoue, I.; Suda, Y.; Takahashi, T.; Nakanishi, H.; Gao, X.-D.; et al. Dynamic localization of a yeast development–specific PP1 complex during prospore membrane formation is dependent on multiple localization signals and complex formation. Mol. Biol. Cell 2017, 28, 3881–3895. [Google Scholar] [CrossRef]
  104. Suda, Y.; Tachikawa, H.; Suda, T.; Kurokawa, K.; Nakano, A.; Irie, K. Remodeling of the secretory pathway is coordinated with de novo membrane formation in budding yeast gametogenesis. bioRxiv 2023. [Google Scholar] [CrossRef]
  105. Lam, C.; Santore, E.; Lavoie, E.; Needleman, L.; Fiacco, N.; Kim, C.; Neiman, A.M. A Visual Screen of Protein Localization during Sporulation Identifies New Components of Prospore Membrane-Associated Complexes in Budding Yeast. Eukaryot. Cell 2014, 13, 383–391. [Google Scholar] [CrossRef]
  106. Maier, P.; Rathfelder, N.; Finkbeiner, M.G.; Taxis, C.; Mazza, M.; Panse, S.L.; Haguenauer-Tsapis, R.; Knop, M. Cytokinesis in yeast meiosis depends on the regulated removal of Ssp1p from the prospore membrane. EMBO J. 2007, 26, 1843–1852. [Google Scholar] [CrossRef] [PubMed]
  107. Maier, P.; Rathfelder, N.; Maeder, C.I.; Colombelli, J.; Stelzer, E.H.; Knop, M. The SpoMBe pathway drives membrane bending necessary for cytokinesis and spore formation in yeast meiosis. EMBO J. 2008, 27, 2363–2374. [Google Scholar] [CrossRef] [PubMed]
  108. Suda, Y.; Nakanishi, H.; Mathieson, E.M.; Neiman, A.M. Alternative Modes of Organellar Segregation during Sporulation in Saccharomyces cerevisiae. Eukaryot. Cell 2007, 6, 2009–2017. [Google Scholar] [CrossRef] [PubMed]
  109. Diamond, A.E.; Park, J.-S.; Inoue, I.; Tachikawa, H.; Neiman, A.M. The Anaphase Promoting Complex Targeting Subunit Ama1 Links Meiotic Exit to Cytokinesis during Sporulation in Saccharomyces cerevisiae. Mol. Biol. Cell 2009, 20, 134–145. [Google Scholar] [CrossRef] [PubMed]
  110. Paulissen, S.M.; Slubowski, C.J.; Roesner, J.M.; Huang, L.S. Timely Closure of the Prospore Membrane Requires SPS1 and SPO77 in Saccharomyces cerevisiae. Genetics 2016, 203, 1203–1216. [Google Scholar] [CrossRef] [PubMed]
  111. Sarkar, P.; Florczyk, M.; McDonough, K.; Nag, D. SSP2, a sporulation-specific gene necessary for outer spore wall assembly in the yeast Saccharomyces cerevisiae. Mol. Genet. Genom. 2002, 267, 348–358. [Google Scholar] [CrossRef] [PubMed]
  112. McDonald, C.M.; Wagner, M.; Dunham, M.J.; Shin, M.E.; Ahmed, N.T.; Winter, E. The Ras/cAMP Pathway and the CDK-like Kinase Ime2 Regulate the MAPK Smk1 and Spore Morphogenesis in Saccharomyces cerevisiae. Genetics 2009, 181, 511–523. [Google Scholar] [CrossRef]
  113. Whinston, E.; Omerza, G.; Singh, A.; Tio, C.W.; Winter, E. Activation of the Smk1 Mitogen-Activated Protein Kinase by Developmentally Regulated Autophosphorylation. Mol. Cell. Biol. 2013, 33, 688–700. [Google Scholar] [CrossRef]
  114. Tio, C.W.; Omerza, G.; Sunder, S.; Winter, E. Autophosphorylation of the Smk1 MAPK is spatially and temporally regulated by Ssp2 during meiotic development in yeast. Mol. Biol. Cell 2015, 26, 3546–3555. [Google Scholar] [CrossRef]
  115. Tio, C.W.; Omerza, G.; Phillips, T.; Lou, H.J.; Turk, B.E.; Winter, E. Ssp2 Binding Activates the Smk1 Mitogen-Activated Protein Kinase. Mol. Cell. Biol. 2017, 37, e00607-16. [Google Scholar] [CrossRef]
  116. Phillips, T.; Tio, C.W.; Omerza, G.; Rimal, A.; Lokareddy, R.K.; Cingolani, G.; Winter, E. RNA Recognition-like Motifs Activate a Mitogen-Activated Protein Kinase. Biochemistry 2018, 57, 6878–6887. [Google Scholar] [CrossRef] [PubMed]
  117. Grandin, N.; Reed, S.I. Differential Function and Expression of Saccharomyces cerevisiae B-type Cyclins in Mitosis and Meiosis. Mol. Cell. Biol. 1993, 13, 2113–2125. [Google Scholar] [CrossRef] [PubMed]
  118. Dahmann, C.; Futcher, B. Specialization of B-type cyclins for mitosis or meiosis in S. cerevisiae. Genetics 1995, 140, 957–963. [Google Scholar] [CrossRef] [PubMed]
  119. Dirick, L.; Goetsch, L.; Ammerer, G.; Byers, B. Regulation of Meiotic S Phase by Ime2 and a Clb5,6-Associated Kinase in Saccharomyces cerevisiae. Science 1998, 281, 1854–1857. [Google Scholar] [CrossRef] [PubMed]
  120. Smith, H.E.; Mitchell, A.P. A Transcriptional Cascade Governs Entry into Meiosis in Saccharomyces cerevisiae. Mol. Cell. Biol. 1989, 9, 2142–2152. [Google Scholar] [CrossRef] [PubMed]
  121. Schindler, K.; Winter, E. Phosphorylation of Ime2 Regulates Meiotic Progression in Saccharomyces cerevisiae. J. Biol. Chem. 2006, 281, 18307–18316. [Google Scholar] [CrossRef] [PubMed]
  122. MacKenzie, A.M.; Lacefield, S. CDK Regulation of Meiosis: Lessons from S. cerevisiae and S. pombe. Genes 2020, 11, 723. [Google Scholar] [CrossRef]
  123. Slubowski, C.J.; Paulissen, S.M.; Huang, L.S. The GCKIII Kinase Sps1 and the 14-3-3 Isoforms, Bmh1 and Bmh2, Cooperate to Ensure Proper Sporulation in Saccharomyces cerevisiae. PLoS ONE 2014, 9, e113528. [Google Scholar] [CrossRef]
  124. Weiss, E.L. Mitotic Exit and Separation of Mother and Daughter Cells. Genetics 2012, 192, 1165–1202. [Google Scholar] [CrossRef] [PubMed]
  125. Matellán, L.; Monje-Casas, F. Regulation of Mitotic Exit by Cell Cycle Checkpoints: Lessons from Saccharomyces cerevisiae. Genes 2020, 11, 195. [Google Scholar] [CrossRef] [PubMed]
  126. Mah, A.S.; Jang, J.; Deshaies, R.J. Protein kinase Cdc15 activates the Dbf2-Mob1 kinase complex. Proc. Natl. Acad. Sci. USA 2001, 98, 7325–7330. [Google Scholar] [CrossRef] [PubMed]
  127. Visintin, R.; Amon, A. Regulation of the Mitotic Exit Protein Kinases Cdc15 and Dbf2. Mol. Biol. Cell 2001, 12, 2961–2974. [Google Scholar] [CrossRef]
  128. D’Aquino, K.E.; Monje-Casas, F.; Paulson, J.; Reiser, V.; Charles, G.M.; Lai, L.; Shokat, K.M.; Amon, A. The Protein Kinase Kin4 Inhibits Exit from Mitosis in Response to Spindle Position Defects. Mol. Cell 2005, 19, 223–234. [Google Scholar] [CrossRef] [PubMed]
  129. Pereira, G.; Schiebel, E. Kin4 Kinase Delays Mitotic Exit in Response to Spindle Alignment Defects. Mol. Cell 2005, 19, 209–221. [Google Scholar] [CrossRef]
  130. Maekawa, H.; Priest, C.; Lechner, J.; Pereira, G.; Schiebel, E. The yeast centrosome translates the positional information of the anaphase spindle into a cell cycle signal. J. Cell Biol. 2007, 179, 423–436. [Google Scholar] [CrossRef]
  131. Chan, L.Y.; Amon, A. Spindle Position Is Coordinated with Cell-Cycle Progression through Establishment of Mitotic Exit-Activating and -Inhibitory Zones. Mol. Cell 2010, 39, 444–454. [Google Scholar] [CrossRef]
  132. Bertazzi, D.T.; Kurtulmus, B.; Pereira, G. The cortical protein Lte1 promotes mitotic exit by inhibiting the spindle position checkpoint kinase Kin4. J. Cell Biol. 2011, 193, 1033–1048. [Google Scholar] [CrossRef]
  133. Rock, J.M.; Amon, A. Cdc15 integrates Tem1 GTPase-mediated spatial signals with Polo kinase-mediated temporal cues to activate mitotic exit. Genes Dev. 2011, 25, 1943–1954. [Google Scholar] [CrossRef] [PubMed]
  134. Falk, J.E.; Campbell, I.W.; Joyce, K.; Whalen, J.; Seshan, A.; Amon, A. LTE1 promotes exit from mitosis by multiple mechanisms. Mol. Biol. Cell 2016, 27, 3991–4001. [Google Scholar] [CrossRef] [PubMed]
  135. Gruneberg, U.; Campbell, K.; Simpson, C.; Grindlay, J.; Schiebel, E. Nud1p links astral microtubule organization and the control of exit from mitosis. EMBO J. 2000, 19, 6475–6488. [Google Scholar] [CrossRef] [PubMed]
  136. Luca, F.C.; Mody, M.; Kurischko, C.; Roof, D.M.; Giddings, T.H.; Winey, M. Saccharomyces cerevisiae Mob1p Is Required for Cytokinesis and Mitotic Exit. Mol. Cell. Biol. 2001, 21, 6972–6983. [Google Scholar] [CrossRef] [PubMed]
  137. Rock, J.M.; Lim, D.; Stach, L.; Ogrodowicz, R.W.; Keck, J.M.; Jones, M.H.; Wong, C.C.L.; Yates, J.R., III; Winey, M.; Smerdon, S.J.; et al. Activation of the Yeast Hippo Pathway by Phosphorylation-Dependent Assembly of Signaling Complexes. Science 2013, 340, 871–875. [Google Scholar] [CrossRef] [PubMed]
  138. Visintin, R.; Craig, K.; Hwang, E.S.; Prinz, S.; Tyers, M.; Amon, A. The Phosphatase Cdc14 Triggers Mitotic Exit by Reversal of Cdk-Dependent Phosphorylation. Mol. Cell 1998, 2, 709–718. [Google Scholar] [CrossRef]
  139. Shou, W.; Seol, J.H.; Shevchenko, A.; Baskerville, C.; Moazed, D.; Chen, Z.W.S.; Jang, J.; Shevchenko, A.; Charbonneau, H.; Deshaies, R.J. Exit from Mitosis Is Triggered by Tem1-Dependent Release of the Protein Phosphatase Cdc14 from Nucleolar RENT Complex. Cell 1999, 97, 233–244. [Google Scholar] [CrossRef]
  140. Mohl, D.A.; Huddleston, M.J.; Collingwood, T.S.; Annan, R.S.; Deshaies, R.J. Dbf2–Mob1 drives relocalization of protein phosphatase Cdc14 to the cytoplasm during exit from mitosis. J. Cell Biol. 2009, 184, 527–539. [Google Scholar] [CrossRef]
  141. Manzoni, R.; Montani, F.; Visintin, C.; Caudron, F.; Ciliberto, A.; Visintin, R. Oscillations in Cdc14 release and sequestration reveal a circuit underlying mitotic exit. J. Cell Biol. 2010, 190, 209–222. [Google Scholar] [CrossRef]
  142. Stegmeier, F.; Amon, A. Closing mitosis: The Functions of the Cdc14 Phosphatase and Its Regulation. Annu. Rev. Genet. 2004, 38, 203–232. [Google Scholar] [CrossRef]
  143. Visintin, R.; Hwang, E.S.; Amon, A. Cfi1 prevents premature exit from mitosis by anchoring Cdc14 phosphatase in the nucleolus. Nature 1999, 398, 818–823. [Google Scholar] [CrossRef]
  144. Rock, J.M.; Amon, A. The FEAR network. Curr. Biol. 2009, 19, R1063–R1068. [Google Scholar] [CrossRef]
  145. Marston, A.L.; Lee, B.H.; Amon, A. The Cdc14 Phosphatase and the FEAR Network Control Meiotic Spindle Disassembly and Chromosome Segregation. Dev. Cell 2003, 4, 711–726. [Google Scholar] [CrossRef]
  146. Azzam, R.; Chen, S.L.; Shou, W.; Mah, A.S.; Alexandru, G.; Nasmyth, K.; Annan, R.S.; Carr, S.A.; Deshaies, R.J. Phosphorylation by Cyclin B-Cdk Underlies Release of Mitotic Exit Activator Cdc14 from the Nucleolus. Science 2004, 305, 516–519. [Google Scholar] [CrossRef]
  147. Queralt, E.; Lehane, C.; Novak, B.; Uhlmann, F. Downregulation of PP2ACdc55 Phosphatase by Separase Initiates Mitotic Exit in Budding Yeast. Cell 2006, 125, 719–732. [Google Scholar] [CrossRef]
  148. D’Amours, D.; Stegmeier, F.; Amon, A. Cdc14 and Condensin Control the Dissolution of Cohesin-Independent Chromosome Linkages at Repeated DNA. Cell 2004, 117, 455–469. [Google Scholar] [CrossRef] [PubMed]
  149. Sullivan, M.; Higuchi, T.; Katis, V.L.; Uhlmann, F. Cdc14 Phosphatase Induces rDNA Condensation and Resolves Cohesin-Independent Cohesion during Budding Yeast Anaphase. Cell 2004, 117, 471–482. [Google Scholar] [CrossRef] [PubMed]
  150. Torres-Rosell, J.; Machín, F.; Jarmuz, A.; Aragón, L. Nucleolar segregation lags behind the rest of the genome and requires Cdc14p activation by the FEAR network. Cell Cycle 2004, 3, 496–502. [Google Scholar] [CrossRef] [PubMed]
  151. Campbell, I.W.; Zhou, X.; Amon, A. The Mitotic Exit Network integrates temporal and spatial signals by distributing regulation across multiple components. eLife 2019, 8, e41139. [Google Scholar] [CrossRef] [PubMed]
  152. Jaspersen, S.L.; Morgan, D.O. Cdc14 activates Cdc15 to promote mitotic exit in budding yeast. Curr. Biol. 2000, 10, 615–618. [Google Scholar] [CrossRef] [PubMed]
  153. König, C.; Maekawa, H.; Schiebel, E. Mutual regulation of cyclin-dependent kinase and the mitotic exit network. J. Cell Biol. 2010, 188, 351–368. [Google Scholar] [CrossRef] [PubMed]
  154. Zhou, X.; Li, W.; Liu, Y.; Amon, A. Cross-compartment signal propagation in the mitotic exit network. eLife 2021, 10, e63645. [Google Scholar] [CrossRef] [PubMed]
  155. Ptacek, J.; Devgan, G.; Michaud, G.; Zhu, H.; Zhu, X.; Fasolo, J.; Guo, H.; Jona, G.; Breitkreutz, A.; Sopko, R.; et al. Global analysis of protein phosphorylation in yeast. Nature 2005, 438, 679–684. [Google Scholar] [CrossRef] [PubMed]
  156. Gihana, G.M.; Cross-Najafi, A.A.; Lacefield, S. The mitotic exit network regulates the spatiotemporal activity of Cdc42 to maintain cell size. J. Cell Biol. 2020, 220, e202001016. [Google Scholar] [CrossRef]
  157. Stegmeier, F.; Visintin, R.; Amon, A. Separase, Polo Kinase, the Kinetochore Protein Slk19, and Spo12 Function in a Network that Controls Cdc14 Localization during Early Anaphase. Cell 2002, 108, 207–220. [Google Scholar] [CrossRef]
  158. Jaspersen, S.L.; Charles, J.F.; Morgan, D.O. Inhibitory phosphorylation of the APC regulator Hct1 is controlled by the kinase Cdc28 and the phosphatase Cdc14. Curr. Biol. 1999, 9, 227–236. [Google Scholar] [CrossRef]
  159. Zachariae, W.; Schwab, M.; Nasmyth, K.; Seufert, W. Control of Cyclin Ubiquitination by CDK-Regulated Binding of Hct1 to the Anaphase Promoting Complex. Science 1998, 282, 1721–1724. [Google Scholar] [CrossRef]
  160. Nasmyth, K. Segregating Sister Genomes: The Molecular Biology of Chromosome Separation. Science 2002, 297, 559–565. [Google Scholar] [CrossRef]
  161. Uhlmann, F.; Wernic, D.; Poupart, M.-A.; Koonin, E.V.; Nasmyth, K. Cleavage of Cohesin by the CD Clan Protease Separin Triggers Anaphase in Yeast. Cell 2000, 103, 375–386. [Google Scholar] [CrossRef]
  162. Wäsch, R.; Cross, F.R. APC-dependent proteolysis of the mitotic cyclin Clb2 is essential for mitotic exit. Nature 2002, 418, 556–562. [Google Scholar] [CrossRef] [PubMed]
  163. Yeong, F.M.; Lim, H.H.; Padmashree, C.G.; Surana, U. Exit from Mitosis in Budding Yeast Biphasic Inactivation of the Cdc28-Clb2 Mitotic Kinase and the Role of Cdc20. Mol. Cell 2000, 5, 501–511. [Google Scholar] [CrossRef] [PubMed]
  164. Pablo-Hernando, M.E.; Arnaiz-Pita, Y.; Nakanishi, H.; Dawson, D.; del Rey, F.; Neiman, A.M.; de Aldana, C.R.V. Cdc15 Is Required for Spore Morphogenesis Independently of Cdc14 in Saccharomyces cerevisiae. Genetics 2007, 177, 281–293. [Google Scholar] [CrossRef] [PubMed]
  165. Attner, M.A.; Amon, A. Control of the mitotic exit network during meiosis. Mol. Biol. Cell 2012, 23, 3122–3132. [Google Scholar] [CrossRef] [PubMed]
  166. Gordon, O.; Taxis, C.; Keller, P.J.; Benjak, A.; Stelzer, E.H.; Simchen, G.; Knop, M. Nud1p, the yeast homolog of Centriolin, regulates spindle pole body inheritance in meiosis. EMBO J. 2006, 25, 3856–3868. [Google Scholar] [CrossRef] [PubMed]
  167. Argüello-Miranda, O.; Zagoriy, I.; Mengoli, V.; Rojas, J.; Jonak, K.; Oz, T.; Graf, P.; Zachariae, W. Casein Kinase 1 Coordinates Cohesin Cleavage, Gametogenesis, and Exit from M Phase in Meiosis II. Dev. Cell 2017, 40, 37–52. [Google Scholar] [CrossRef] [PubMed]
  168. Tan, G.S.; Magurno, J.; Cooper, K.F. Ama1p-activated anaphase-promoting complex regulates the destruction of Cdc20p during meiosis II. Mol. Biol. Cell 2011, 22, 315–326. [Google Scholar] [CrossRef]
  169. Cooper, K.F.; Mallory, M.J.; Egeland, D.B.; Jarnik, M.; Strich, R. Ama1p is a meiosis-specific regulator of the anaphase promoting complex/cyclosome in yeast. Proc. Natl. Acad. Sci. USA 2000, 97, 14548–14553. [Google Scholar] [CrossRef]
  170. Holt, L.J.; Hutti, J.E.; Cantley, L.C.; Morgan, D.O. Evolution of Ime2 Phosphorylation Sites on Cdk1 Substrates Provides a Mechanism to Limit the Effects of the Phosphatase Cdc14 in Meiosis. Mol. Cell 2007, 25, 689–702. [Google Scholar] [CrossRef]
  171. Okaz, E.; Argüello-Miranda, O.; Bogdanova, A.; Vinod, P.K.; Lipp, J.J.; Markova, Z.; Zagoriy, I.; Novak, B.; Zachariae, W. Meiotic Prophase Requires Proteolysis of M Phase Regulators Mediated by the Meiosis-Specific APC/CAma1. Cell 2012, 151, 603–618. [Google Scholar] [CrossRef]
  172. Oelschlaegel, T.; Schwickart, M.; Matos, J.; Bogdanova, A.; Camasses, A.; Havlis, J.; Shevchenko, A.; Zachariae, W. The Yeast APC/C Subunit Mnd2 Prevents Premature Sister Chromatid Separation Triggered by the Meiosis-Specific APC/C-Ama1. Cell 2005, 120, 773–788. [Google Scholar] [CrossRef] [PubMed]
  173. Penkner, A.M.; Prinz, S.; Ferscha, S.; Klein, F. Mnd2, an Essential Antagonist of the Anaphase-Promoting Complex during Meiotic Prophase. Cell 2005, 120, 789–801. [Google Scholar] [CrossRef] [PubMed]
  174. McDonald, C.M.; Cooper, K.F.; Winter, E. The Ama1-Directed Anaphase-Promoting Complex Regulates the Smk1 Mitogen-Activated Protein Kinase during Meiosis in Yeast. Genetics 2005, 171, 901–911. [Google Scholar] [CrossRef] [PubMed]
  175. Omerza, G.; Tio, C.W.; Philips, T.; Diamond, A.; Neiman, A.M.; Winter, E. The meiosis-specific Cdc20 family-member Ama1 promotes binding of the Ssp2 activator to the Smk1 MAP kinase. Mol. Biol. Cell 2018, 29, 66–74. [Google Scholar] [CrossRef] [PubMed]
  176. Rimal, A.; Kamdar, Z.P.; Tio, C.W.; Winter, E. Isc10, an Inhibitor That Links the Anaphase-Promoting Complex to a Meiosis-Specific Mitogen-Activated Protein Kinase. Mol. Cell. Biol. 2020, 40, e00097-20. [Google Scholar] [CrossRef] [PubMed]
  177. Seitz, B.C.; Mucelli, X.; Majano, M.; Wallis, Z.; Dodge, A.C.; Carmona, C.; Durant, M.; Maynard, S.; Huang, L.S. Meiosis II spindle disassembly requires two distinct pathways. Mol. Biol. Cell 2023, 34, ar98. [Google Scholar] [CrossRef] [PubMed]
  178. King, G.A.; Goodman, J.S.; Schick, J.G.; Chetlapalli, K.; Jorgens, D.M.; McDonald, K.L.; Ünal, E. Meiotic cellular rejuvenation is coupled to nuclear remodeling in budding yeast. eLife 2019, 8, e47156. [Google Scholar] [CrossRef]
  179. King, G.A.; Ünal, E. The dynamic nuclear periphery as a facilitator of gamete health and rejuvenation. Curr. Genet. 2020, 66, 487–493. [Google Scholar] [CrossRef]
  180. King, G.A.; Wettstein, R.; Varberg, J.M.; Chetlapalli, K.; Walsh, M.E.; Gillet, L.C.J.; Hernández-Armenta, C.; Beltrao, P.; Aebersold, R.; Jaspersen, S.L.; et al. Meiotic nuclear pore complex remodeling provides key insights into nuclear basket organization. J. Cell Biol. 2022, 222, e202204039. [Google Scholar] [CrossRef]
  181. Otto, G.M.; Cheunkarndee, T.; Leslie, J.M.; Brar, G.A. Programmed cortical ER collapse drives selective ER degradation and inheritance in yeast meiosis. J. Cell Biol. 2021, 220, e202108105. [Google Scholar] [CrossRef]
  182. Sawyer, E.M.; Joshi, P.R.; Jorgensen, V.; Yunus, J.; Berchowitz, L.E.; Ünal, E. Developmental regulation of an organelle tether coordinates mitochondrial remodeling in meiosis. J. Cell Biol. 2019, 218, 559–579. [Google Scholar] [CrossRef]
  183. McCartney, B.; Dudin, O. Cellularization across eukaryotes: Conserved mechanisms and novel strategies. Curr. Opin. Cell Biol. 2023, 80, 102157. [Google Scholar] [CrossRef] [PubMed]
  184. Hoffman, C.S.; Wood, V.; Fantes, P.A. An Ancient Yeast for Young Geneticists: A Primer on the Schizosaccharomyces pombe Model System. Genetics 2015, 201, 403–423. [Google Scholar] [CrossRef] [PubMed]
  185. Shimoda, C. Forespore membrane assembly in yeast: Coordinating SPBs and membrane trafficking. J. Cell Sci. 2003, 117, 389–396. [Google Scholar] [CrossRef] [PubMed]
  186. Ohtsuka, H.; Imada, K.; Shimasaki, T.; Aiba, H. Sporulation: A response to starvation in the fission yeast Schizosaccharomyces pombe. MicrobiologyOpen 2022, 11, e1303. [Google Scholar] [CrossRef]
  187. Yan, H.; Balasubramanian, M.K. Meiotic actin rings are essential for proper sporulation in fission yeast. J. Cell Sci. 2012, 125, 1429–1439. [Google Scholar] [CrossRef]
  188. Sokac, A.M.; Biel, N.; Renzis, S.D. Membrane-actin interactions in morphogenesis: Lessons learned from Drosophila cellularization. Semin. Cell Dev. Biol. 2023, 133, 107–122. [Google Scholar] [CrossRef]
  189. Haglund, K.; Nezis, I.P.; Stenmark, H. Structure and functions of stable intercellular bridges formed by incomplete cytokinesis during development. Commun. Integr. Biol. 2010, 4, 1–9. [Google Scholar] [CrossRef]
Figure 1. Meiosis occurs during sporulation in S. cerevisiae. Cartoon depicting sporulation in Saccharomyces cerevisiae. Progression through sporulation is diagrammed from left to right. Nuclei are labelled in yellow, spindle pole bodies in black, and prospore membranes in magenta. The mannan, β-glucan, chitosan, and dityrosine layers of the spore wall are labelled pink, turquoise, dark blue, and green, respectively.
Figure 1. Meiosis occurs during sporulation in S. cerevisiae. Cartoon depicting sporulation in Saccharomyces cerevisiae. Progression through sporulation is diagrammed from left to right. Nuclei are labelled in yellow, spindle pole bodies in black, and prospore membranes in magenta. The mannan, β-glucan, chitosan, and dityrosine layers of the spore wall are labelled pink, turquoise, dark blue, and green, respectively.
Jof 10 00132 g001
Figure 2. Prospore membranes undergo various morphological changes during meiosis II. (A) Cartoon depicting the stages of a single growing prospore membrane during sporulation. The prospore membrane is labelled magenta, nucleus in yellow, meiotic outer plaque in black, and leading-edge protein complex (LEP) in blue. Pathways controlling closure of the prospore membrane are indicated. Details in text. (B) Prospore membrane development in wild-type SK1 cells (LH1146), which contain the genomically integrated prospore membrane marker E20 (his3::SPO2051–91-GFPENVY:HIS3 [25,26]) shown in magenta and histones (HTB2-mCherry:TRP1 [27]) shown in yellow. Scale bar = 2 microns. Fluorescent images are maximum intensity projections of 3 µm z-stacks merges taken on a Zeiss Axioskop Mot2 wide-field microscope, using a 100× objective (NA 1.45).
Figure 2. Prospore membranes undergo various morphological changes during meiosis II. (A) Cartoon depicting the stages of a single growing prospore membrane during sporulation. The prospore membrane is labelled magenta, nucleus in yellow, meiotic outer plaque in black, and leading-edge protein complex (LEP) in blue. Pathways controlling closure of the prospore membrane are indicated. Details in text. (B) Prospore membrane development in wild-type SK1 cells (LH1146), which contain the genomically integrated prospore membrane marker E20 (his3::SPO2051–91-GFPENVY:HIS3 [25,26]) shown in magenta and histones (HTB2-mCherry:TRP1 [27]) shown in yellow. Scale bar = 2 microns. Fluorescent images are maximum intensity projections of 3 µm z-stacks merges taken on a Zeiss Axioskop Mot2 wide-field microscope, using a 100× objective (NA 1.45).
Jof 10 00132 g002
Figure 3. Meiotic exit pathway compared to mitotic exit. Sporulation-specific proteins are pink. Proteins used only in mitotic exit are blue or green. Proteins used in both are gray. Details in text.
Figure 3. Meiotic exit pathway compared to mitotic exit. Sporulation-specific proteins are pink. Proteins used only in mitotic exit are blue or green. Proteins used in both are gray. Details in text.
Jof 10 00132 g003
Figure 4. Spindle disassembly and prospore membrane closure during exit from meiosis II. (A) Cartoon depicting the stages of a cell during anaphase, spindle disassembly, and prospore membrane closure with time progression from left to right. Prospore membranes are labelled magenta, leading-edge protein complex (LEP) blue, and spindles green. (B) Prospore membrane development was examined in SK1 cells (LH1147), which contain the genomically integrated prospore membrane marker K20 (his3::SPO2051−91-mKATE:HIS3 [103]) shown in magenta, leading-edge marker Don1 (DON1-mTagBFP2::KANMX [25]) shown in blue, and spindle marker Tub1 (alpha-tubulin; GFPENVY-TUB1+3′UTR:LEU2 [177]) shown in green. Scale bar = 2 microns. Fluorescent images are maximum intensity projections of 3 µm z-stacks merged taken on a Zeiss Axioskop Mot2 wide-field microscope, using a 100× objective (NA 1.45).
Figure 4. Spindle disassembly and prospore membrane closure during exit from meiosis II. (A) Cartoon depicting the stages of a cell during anaphase, spindle disassembly, and prospore membrane closure with time progression from left to right. Prospore membranes are labelled magenta, leading-edge protein complex (LEP) blue, and spindles green. (B) Prospore membrane development was examined in SK1 cells (LH1147), which contain the genomically integrated prospore membrane marker K20 (his3::SPO2051−91-mKATE:HIS3 [103]) shown in magenta, leading-edge marker Don1 (DON1-mTagBFP2::KANMX [25]) shown in blue, and spindle marker Tub1 (alpha-tubulin; GFPENVY-TUB1+3′UTR:LEU2 [177]) shown in green. Scale bar = 2 microns. Fluorescent images are maximum intensity projections of 3 µm z-stacks merged taken on a Zeiss Axioskop Mot2 wide-field microscope, using a 100× objective (NA 1.45).
Jof 10 00132 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Durant, M.; Mucelli, X.; Huang, L.S. Meiotic Cytokinesis in Saccharomyces cerevisiae: Spores That Just Need Closure. J. Fungi 2024, 10, 132. https://doi.org/10.3390/jof10020132

AMA Style

Durant M, Mucelli X, Huang LS. Meiotic Cytokinesis in Saccharomyces cerevisiae: Spores That Just Need Closure. Journal of Fungi. 2024; 10(2):132. https://doi.org/10.3390/jof10020132

Chicago/Turabian Style

Durant, Matthew, Xheni Mucelli, and Linda S. Huang. 2024. "Meiotic Cytokinesis in Saccharomyces cerevisiae: Spores That Just Need Closure" Journal of Fungi 10, no. 2: 132. https://doi.org/10.3390/jof10020132

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