Skip to main content
eLife logoLink to eLife
. 2014 Oct 3;3:e03779. doi: 10.7554/eLife.03779

TORC2-dependent protein kinase Ypk1 phosphorylates ceramide synthase to stimulate synthesis of complex sphingolipids

Alexander Muir 1,2, Subramaniam Ramachandran 1, Françoise M Roelants 1, Garrett Timmons 3, Jeremy Thorner 1,*
Editor: Tony Hunter4
PMCID: PMC4217029  PMID: 25279700

Abstract

Plasma membrane lipid composition must be maintained during growth and under environmental insult. In yeast, signaling mediated by TOR Complex 2 (TORC2)-dependent protein kinase Ypk1 controls lipid abundance and distribution in response to membrane stress. Ypk1, among other actions, alleviates negative regulation of L-serine:palmitoyl-CoA acyltransferase, upregulating production of long-chain base precursors to sphingolipids. To explore other roles for TORC2-Ypk1 signaling in membrane homeostasis, we devised a three-tiered genome-wide screen to identify additional Ypk1 substrates, which pinpointed both catalytic subunits of the ceramide synthase complex. Ypk1-dependent phosphorylation of both proteins increased upon either sphingolipid depletion or heat shock and was important for cell survival. Sphingolipidomics, other biochemical measurements and genetic analysis demonstrated that these modifications of ceramide synthase increased its specific activity and stimulated channeling of long-chain base precursors into sphingolipid end-products. Control at this branch point also prevents accumulation of intermediates that could compromise cell growth by stimulating autophagy.

DOI: http://dx.doi.org/10.7554/eLife.03779.001

Research organism: S. cerevisiae

eLife digest

Cells are enclosed by a plasma membrane that separates and protects each cell from its environment. These membranes are made of a variety of proteins and fatty molecules called lipids, which are carefully organized throughout the membrane. When cells experience stresses such as heat or excessive pressure, the plasma membrane changes to help protect the cell. In particular, more of a group of lipids called sphingolipids are incorporated into the membrane under stress conditions.

In yeast cells, a protein called Ypk1 plays an important role in protecting the cell from stress. Ypk1 controls the activity of a number of proteins that are responsible for balancing the amounts of different types of lipids in cell membranes. The combined action of these Ypk1-dependent proteins leads to the remodelling of the cell membrane to protect against stress. While several proteins that work with Ypk1 are known, some of the changes that serve to protect the plasma membrane cannot be explained by the action of these proteins alone.

To provide a more comprehensive picture of how Ypk1 helps cells to respond to changes in the environment, Muir et al. developed a new approach that combines biochemical, genetic and bioinformatics techniques to survey the yeast genome for proteins that could be Ypk1 targets. Muir et al. first produced a list of potential candidate proteins by searching for proteins with features similar to known Ypk1 targets, and then considered those that are known to be involved in processes that also involve Ypk1. To filter the potential targets further, Muir et al. performed experiments in yeast cells to see which proteins prevented normal cell growth if they were over-produced. Further experiments investigating which of these proteins interact with Ypk1 when purified identified 12 new proteins that are most likely targets of the Ypk1 protein.

Two of these newly identified Ypk1 target proteins form part of an enzyme complex called ceramide synthase, which produces a family of waxy lipid molecules from which more complex sphingolipids are built. Muir et al. discovered that during stress, Ypk1 enhances the activity of the ceramide synthase enzyme, which increases lipid production and the amount of sphingolipid deposited in the cell membrane. If this process is interrupted at any stage, cells struggle to survive under stress conditions.

The other candidate proteins identified by Muir et al. remain to be validated and characterized as Ypk1 targets. Nevertheless, the techniques used have conclusively identified some new Ypk1 targets and could also be applied to similar searches for proteins targeted in other biological processes.

DOI: http://dx.doi.org/10.7554/eLife.03779.002

Introduction

A eukaryotic plasma membrane (PM) is a complex structure composed of many protein (Sachs and Engelman, 2006) and lipid (Simons and Sampaio, 2011) species arranged with a high degree of compositional and spatial organization. The levels and distributions of lipids are clearly important for PM processes—from solute transport (Divito and Amara, 2009), to endocytosis (Platta and Stenmark, 2011), to receptor function and signal transduction (Groves and Kuriyan, 2010). It is essential, therefore, for a cell to maintain proper lipid balance. How cells sense alterations in PM organization in response to developmental cues or environmental stresses and adjust the rates of the reactions that maintain lipid homeostasis are vital mechanisms to understand.

We (Roelants et al., 2011) and others (Berchtold et al., 2012; Niles and Powers, 2012) have shown that dynamic changes in the function of Target of Rapamycin (TOR) Complex 2 (TORC2) and in a downstream protein kinase that it stimulates, Ypk1 (and its paralog Ypk2), are important for cell survival in response to membrane stress. Ypk1 (and Ypk2) are two members of the AGC kinase family (Pearce et al., 2010) and orthologs of mammalian SGK1 (Casamayor et al., 1999). In Saccharomyces cerevisiae, Ypk1 (and Ypk2) are activated at a basal level by phosphorylation of a Thr in the activation loop by the eisosome-associated protein kinase Pkh1 (and its paralog Pkh2) (Casamayor et al., 1999; Roelants et al., 2002). The ortholog responsible for this reaction in mammalian cells is PDK1 (Mora et al., 2004). However, this level of activity is not sufficient for Ypk1 to permit survival when cells are challenged with certain membrane perturbants (Roelants et al., 2004). Under such conditions of membrane stress, TORC2 stimulates Ypk1 and Ypk2 function by phosphorylating at least two C-terminal sites (Kamada et al., 2005; Roelants et al., 2011; Niles et al., 2012). Constitutively-active Ypk1 and Ypk2 alleles bypass the need for functional TORC2, indicating that Ypk1 and Ypk2 are solely responsible for executing all the essential downstream functions of TORC2 (Kamada et al., 2005; Roelants et al., 2011; Niles et al., 2012).

How membrane stress is communicated to TORC2 is a question of on-going research. It has been reported that membrane stress caused by inhibition of sphingolipid synthesis or membrane stretch (induced by hypotonic shock) causes two PH domain-containing proteins (Slm1 and Slm2) to relocalize from eisosomes to a separate PM region that contains TORC2 and leads to increased TORC2 activation of Ypk1 and Ypk2 (Berchtold et al., 2012), purportedly because Slm1 and Slm2 recruit Ypk1 and Ypk2 to TORC2 (Niles et al., 2012). However, other evidence indicates that Avo1 (ortholog in other organisms is Sin1) is the subunit of the TORC2 complex primarily responsible for substrate recognition of Ypk1 and its orthologs, including association of S. cerevisiae Ypk2 with Avo1 (Liao and Chen, 2012), Gad8 with Sin1 in fission yeast (Ikeda et al., 2008; Kataoka et al., 2014), and SGK1 with mSin1 in mammalian cells (Jacinto et al., 2006; Yang et al., 2006; Lu et al., 2011; Liu et al., 2013). Regardless of the actual mechanism by which the activity of the TORC2-Ypk1 signaling module is affected by these assaults on the PM, it clearly sets in motion processes that allow the cells to cope appropriately with these stresses.

Several targets of Ypk1 have already been identified that shed light on how TORC2-Ypk1 signaling reprograms cellular processes to cope with PM stress. Ypk1 phosphorylates and negatively regulates Fpk1, a protein kinase responsible for activation of PM-localized aminophospholipid flippases (Roelants et al., 2010), thereby fine-tuning the phosphatidylethanolamine content of the inner leaflet of the PM bilayer. Ypk1 also phosphorylates and alleviates the inhibitory function of two endoplasmic reticulum (ER)-localized proteins (Orm1 and Orm2) that impede the function of the first enzyme unique to sphingolipid biosynthesis, L-serine:palmitoyl-CoA acyltransferase (SPT) (Roelants et al., 2011; Berchtold et al., 2012; Sun et al., 2012), thereby increasing the rate of formation of the long-chain base precursor (phytosphingosine) to yeast sphingolipids. Ypk1 also phosphorylates and inhibits one of the two isoforms (Gpd1) of glycerol-3-phosphate dehydrogenase, a primary source of glycerol-3P for production of glycerophospholipids (Lee et al., 2012b). In response to hyperosmotic shock, and in contrast to other PM stressors, TORC2 activity is markedly decreased, preventing Ypk1-mediated inhibition of Gpd1 [which also gets upregulated transcriptionally (Ansell et al., 1997)], thereby greatly increasing synthesis of glycerol-3P, which is dephosphorylated to produce glycerol, an innocuous osmolyte that yeast cells accumulate as a means to counteract the increase in external osmotic pressure (Lee et al., 2012b). Control of all these reactions already made it clear that TORC2-Ypk1 signaling is a central regulator of PM lipid homeostasis.

To gain further insight into how the TORC2-Ypk1 signaling axis contributes to PM homeostasis and potentially other cellular processes, we devised a three-step procedure to screen in an unbiased and genome-wide manner for additional candidate Ypk1 substrates whose physiological relevance we could then evaluate. As described here, we first used bioinformatics to search the yeast proteome for presumptive targets, then applied a genetic method to narrow down the hits to likely, functionally important substrates, and then used biochemical analysis both in vitro and in vivo to validate the best prospects. Although we report here the general outcomes of this screen, we focus mainly on our discovery and demonstration that both of the catalytic subunits of the ER-localized ceramide synthase complex (Schorling et al., 2001; Vallée and Riezman, 2005) are bona fide targets of Ypk1. These findings provide new insight into how TORC2-initiated signaling regulates flux through the sphingolipid pathway and how this specific control mechanism is important for survival when sphingolipid synthesis is compromised. In addition, we show that this regulation of the ceramide synthase complex is important for preventing accumulation of pathway intermediates that would otherwise compromise cell growth by triggering an inappropriate autophagic response.

Results

A three-tiered screen to identify new Ypk1 substrates

We devised a three-step strategy (Figure 1A) to pinpoint bona fide cellular targets of Ypk1, utilizing bioinformatics to predict potential Ypk1 substrates, then an in vivo genetic test involving a novel variation of the synthetic dosage lethality method to winnow the list to likely candidates, and finally biochemical analysis in vitro to confirm whether the identified gene product serves as a direct substrate of Ypk1. The physiological relevance of Ypk1-dependent modification of each protein on the resulting final list could then been evaluated.

Figure 1. A three-part screen to identify likely Ypk1 substrates.

Figure 1.

(A) The three-part screening strategy to identify Ypk1 substrates is shown schematically as a flow chart. Numbers indicate the number of hits/considered genes at each step in the screen. (B) The bioinformatic approach towards identifying Ypk1 substrates is schematized as a flowchart with each filter as a box. Genes were first filtered by MOTIPS on the basis of having likely phosphorylatable Ypk1 motifs. Subsequently, substrates were filtered by having many Ypk1 motifs or having a Ypk1 site known to be phosphorylated in published data sets. Lastly, genes were filtered by requiring the gene to have a published chemical sensitivity like Ypk1 does, or a published interaction with Ypk1, Ypk1 regulators (TORC2 or PP2A) or sphingolipid biosynthetic machinery. (C) A possible explanation for Ypk1 synthetic dosage lethality interactions is shown. Normally, the cell has enough kinase activity to buffer overexpression of a substrate (Substrate 2), so that essential substrates are regulated and normal growth is unperturbed. However, concurrent decrease in kinase activity coupled with substrate overexpression causes loss of regulation of essential substrate(s) (Substrate 1) leading to observable growth defects. (D) ypk1-as ypk2Δ (yAM135–A) cells were transformed with PGAL1-GFP (negative control), PGAL1-Orm1 or PGAL1-Orm2 (known Ypk1 substrates, positive SDL controls) plasmids. Overnight cultures were then serially diluted onto either dextrose (to repress substrate overexpression) or galatose (to induce substrate overexpression) containing media with increasing concentrations of the Ypk1-as inhibitor 3-MB-PP1. (E) GST-Orm1(1–85) (pFR203) and GST-Fps1(1–255) (pBT6) were purified from E. coli and incubated with [γ-32P]ATP and Ypk1-as, purified from S. cerevisiae, in the absence or presence of 3-MB-PP1. The products were then resolved by SDS/PAGE and analyzed as described in ‘Materials and methods’.

DOI: http://dx.doi.org/10.7554/eLife.03779.003

For initial bioinformatic search of the yeast proteome, we developed a position-weighted consensus sequence logo (Figure 1B) for the preferred Ypk1 phospho-acceptor site based on two primary criteria: (a) the known Ypk1 sites in five, validated in vivo targets (Fpk1, Fpk2, Orm1, Orm2, and Gpd1) (Roelants et al., 2010, 2011; Lee et al., 2012b; Niles et al., 2012; Sun et al., 2012); and, (b) the sequence preference displayed by Ypk1 for phosphorylation of synthetic peptides in vitro (Casamayor et al., 1999; Mok et al., 2010). All demonstrated substrates either in vivo or in vitro contain Arg at positions −5 and −3 with respect to the phosphorylated Ser (or Thr); thus, these positions were invariant in the search motif. Given that nearly all the verified sites within known in vivo targets possess a hydrophobic residue (V, I, F) at position +1, the search motif gave preference to sites with a hydrophobic residue at the +1 position. We then took advantage of the existing MOTIPS motif analysis package (Lam et al., 2010) to identify those S. cerevisiae gene products that contain occurrences of the search logo. Several authentic Ypk1 substrates (e.g., Fpk1, Orm1, and Orm2) contain multiple Ypk1 phosphorylation sites. Thus, we filtered our search further by prioritizing candidates containing multiple matches to the search logo, as predicted by MOTIPS. However, to avoid disregarding potential Ypk1 substrates with a single predicted match to the sequence logo, we also considered MOTIPS hits wherein there was existing evidence in the PhosphoGRID database (Sadowski et al., 2013) indicating that a predicted site is phosphorylated in vivo. To narrow down the list of potential substrates further, we chose to pursue those gene products containing matches to the sequence logo for which there was existing information in the literature suggesting a phenotypic relationship to known Ypk1-dependent processes: (a) a loss-of-function mutation in the candidate gene exhibits elevated sensitivity to agents (aureobasidin A, caspofungin, and/or myriocin) toward which a ypk1Δ mutant is also sensitive (Hillenmeyer et al., 2008); (b) the candidate gene product is reported to be involved in a genetic or biochemical interaction with Ypk1, as curated in YeastMine (Balakrishnan et al., 2012); (c) the candidate gene product is connected in some way to known Ypk1 regulators (e.g., TORC2, PP2A); and/or, (d) the candidate gene product is involved in a known Ypk1-regulated process (e.g., sphingolipid metabolism) (for further details, see ‘Materials and methods’). Reassuringly, our approach identified three known Ypk1 substrates (Fpk1, Orm1 and Orm2); absence of Fpk2 and Gpd1 from the list generated solely by the MOTIPS search criteria arose from the fact that these substrates contain only a single predicted site that is not presently recorded in PhosphoGRID. For this reason, we also restored for consideration additional gene products that contain a single match to the consensus sequence logo that YeastMine indicated are involved in processes in which Ypk1 is implicated. The resulting candidates, grouped via cellular process on the basis of current GO Slim terminology (http://www.geneontology.org/GO.slims.shtml), are cataloged in Table 1, and represent fewer than 100 gene products out of the approximately 6000 protein-coding genes in the yeast genome (Lin et al., 2013) [although the number of authentic open-reading-frames undergoes constant revision (http://www.yeastgenome.org/cache/genomeSnapshot.html)].

Table 1.

Known Ypk1 substrates and potential substrates predicted by MOTIPS listed under GO Slim terms*

DOI: http://dx.doi.org/10.7554/eLife.03779.004

Gene MOTIPS sites Chemical Sensitivity/YeastMine Interaction(s) SDL score Ypk1 dosage rescue In vitro substrate
Known Ypk1 Substrates
 GPD1/YDL022W* 24P YeastMine +++ N/A +
 FPK1/YNR047W 37, 200, 244, 436, 481 YeastMine + N/A +
 FPK1(D621A) [Kinase-dead mutant] 37, 200, 244, 436, 481 YeastMine ++ N/A +
 ORM1/YGR038W 52, 53 YeastMine +++ N/A +
 ORM2/YLR350W 47, 48 YeastMine ++++ N/A +
Cytoskeleton Organization
 AVO1/YOL078W 552P,597,1078 YeastMine N/A N/A
 AVO2/YMR068W 273P, 305, 407 YeastMine N/A N/A
 BEM2/YER155C 83, 168, 1810, 1813 Myr, YeastMine N/A N/A
 BNI1/YNL271C 119, 1138P, 1533 AbA, Casp, YeastMine N/A N/A
 CDH1/YGL003C 51, 195, 213P AbA, YeastMine TOXIC N/A
 ENT1/YDL161W 160P YeastMine N/A N/A
 GIC2/YDR309C 90, 312, 345P Myr, AbA N/A N/A
 LSB3/YFR024C-A 262P YeastMine N/A N/A
 PAL1/YDR348C 49P, 391, 436 AbA, Casp ++++ N/A
 SLA1/YBL007C 445, 447P, 449P, 477 YeastMine N/A N/A
 TSC11/YER093C 19P, 97, 188 YeastMine N/A N/A N/A
YHR097C 58, 288P, 294P Myr +++ N/A +/−
 YSC84/YHR016C 274, 374P Myr, YeastMine N/A N/A
Biological Process Unknown
 COM2/YER130C 251, 370, 380 Myr, YeastMine N/A N/A
 ECM3/YOR092W* 312, 350 Myr, AbA, YeastMine N/A N/A
 ICS2/YBR157C 14, 95, 136, 172P Myr N/A N/A
 JIP4/YDR475C 348, 352, 592, 649 Myr, AbA N/A N/A N/A
 KKQ8/YKL168C 83, 113, 144, 146, 212, 293 YeastMine N/A N/A
 RTS3/YGR161C* 30, 238P YeastMine N/A N/A
 SEG1/YMR086W 56, 118P, 217, 634, 752P Myr + N/A N/A
 YDR186C 334P, 540P, 542P, 620, 715P YeastMine N/A N/A
 YHR080C 401, 513, 667 YeastMine N/A N/A
YNR014W 54, 115, 156, 197 YeastMine +++ N/A +
 YPK3/YBR028C 72, 73, 90P Myr, Casp N/A N/A
Transcription from RNA Polymerase II Promoter
 EPL1/YFL024C 24, 28, 61 YeastMine N/A N/A
 FKH1/YIL131C 404 Myr, YeastMine TOXIC
 GAL11/YOL051W 1003P Myr, YeastMine N/A N/A
 HCM1/YCR065W* 80 AbA, YeastMine N/A N/A
 HOT1/YMR172W* 387, 520, 586 Myr N/A N/A
 RLM1/YPL089C* 20 Myr TOXIC N/A
SMP1/YBR182C* 20, 107 YeastMine TOXIC + +
 SSN2/YDR443C 608P Myr N/A N/A
 YHP1/YDR451C* 180, 182 Myr N/A N/A
Mitotic Cell Cycle
 BCK2/YER167W 12, 38, 373, 430 YeastMine TOXIC N/A
 ESC2/YDR363W 114, 143, 145 YeastMine N/A N/A
 PTK2/YJR059W 59P, 82, 91, 171, 275 Myr, YeastMine + N/A N/A
 SET3/YKR029C 236, 405, 428 YeastMine N/A N/A
 SWI4/YER111C 816P Myr, YeastMine N/A N/A
 VHS2/YIL135C 316, 318, 325P Myr, Casp N/A N/A
 ZDS1/YMR273C* 78, 370 AbA, YeastMine N/A N/A
 ZDS2/YML109W 183, 267, 345 YeastMine N/A N/A
Protein Phosphorylation
 HAL5/YJL165C 17P, 217P, 233 YeastMine N/A N/A
 KIN1/YDR122W 652, 791P, 879, 986P YeastMine + N/A N/A
 KIN2/YLR096W 665P, 818, 1020 Myr N/A N/A
 KSP1/YHR082C 594, 827P, 884P Myr, AbA, YeastMine TOXIC N/A
 NPR1/YNL183C 125P, 255P, 257P, 317P YeastMine ++++ N/A
 SKY1/YMR216C 383P Myr, YeastMine N/A N/A
 YAK1/YJL141C 128P, 206, 240 Myr, Casp, YeastMine N/A N/A
 YPL150W 371P, 452, 890 YeastMine N/A N/A
Lipid Metabolic Process
 ADR1/YDR216W 180, 230P, 756 Myr, AbA N/A N/A
 CDC1/YDR182W* 9 N/A N/A +
 CKI1/YLR133W 14P, 25P, 30P Myr, AbA, YeastMine N/A N/A
GPT2/YKR067W 27, 652 Myr +++ N/A +
LAC1/YKL008C* 23, 24 Myr, YeastMine +++ N/A +
LAG1/YHL003C 24P Myr, YeastMine +++ N/A +
 LCB3/YJL134W 16P Myr, YeastMine N/A +
Cellular Ion Homeostasis and Transport
 AVT3/YKL146W 55, 59P, 172, 174 Myr, AbA, YeastMine N/A N/A
 CCH1/YGR217W 146, 148, 347 YeastMine N/A +
FPS1/YLL043W 147, 181, 185, 570P Myr, YeastMine +++++ N/A +
 MNR2/YKL064W 165, 620, 621, 826 AbA N/A N/A
 NHA1/YLR138W* 544, 830 Myr, YeastMine N/A N/A
 PPZ1/YML016C 122, 203, 250P Myr, YeastMine TOXIC N/A
Translation
 DED1/YOR204W 84, 576P YeastMine N/A N/A
 HCR1/YLR192C* 223 Myr, AbA, YeastMine N/A N/A
 HEF3/YNL014W* 898 Myr, AbA, YeastMine N/A N/A
 RPL3/YOR063W 24P, 337 Myr, AbA, YeastMine N/A N/A
 SUI2/YJR007W* 58 Myr N/A N/A
 TEF1/YPR080W* 72P Myr N/A N/A
Cell Wall Organization or Biogenesis
 BPH1/YCR032W 1334, 1336, 1963 Casp N/A N/A
 CSR2/YPR030W 61, 103, 525, 987 Myr N/A N/A
 ROM2/YLR371W 76P, 193p, 396 YeastMine N/A N/A
 SSD1/YDR293C 164P, 482P, 503 Myr, AbA, YeastMine TOXIC N/A
Golgi Vesicle Transport
 BRE5/YNR051C 398P Myr, YeastMine +++ N/A
 EXO84/YBR102C 76, 313, 494, 554 YeastMine N/A N/A
MUK1/YPL070W 173, 184P, 185P Myr +++ N/A +/−
 RGP1/YDR137W 220, 364P, 450, 452 YeastMine N/A N/A
Signaling
 IRA2/YOL081W 882, 884, 1578, 1745, 3069 YeastMine N/A N/A N/A
 GIS3/YLR094C* 249, 333 Myr, AbA N/A N/A
 MDS3/YGL197W 757, 824, 842, 851, 1204 Myr, AbA, YeastMine +++ N/A
 SYT1/YPR095C 277P, 410, 728 YeastMine N/A N/A N/A
DNA Replication
 CDC13/YDL220C 314, 333P YeastMine TOXIC N/A
 CTI6/YPL181W 155, 216P Myr, YeastMine N/A N/A
 RIM4/YHL024W 93, 429, 525, 607 YeastMine N/A N/A
Endocytosis
 ALY2/YJL084C 166P, 201, 225, 803 Myr N/A N/A
ROD1/YOR018W 563, 617, 807, 823 Myr +++ N/A +/−
 ROG3/YFR022W 425, 584, 718 YeastMine N/A N/A
Other
 FRT1/YOR324C 167, 201, 203, 228P, 385 Myr, YeastMine N/A N/A
 HER1/YOR227W 28P, 102p, 157P Myr, AbA TOXIC +
YSP2/YDR326C 326, 518, 1237 Myr, YeastMine +++ N/A +
RNA Catabolic Process
 JSN1/YJR091C 174, 275P, 600 YeastMine N/A N/A
 PUF2/YPR042C 55, 143, 246, 902 Myr N/A N/A N/A
Cytokinesis
 CYK3/YDL117W 159, 207P, 746 AbA, YeastMine +++ N/A
Chromosome Segregation
 DSN1/YIR010W 240, 250P YeastMine N/A N/A
Peroxisome Organization
PEX31/YGR004W 432P YeastMine ++ N/A +
Pseudohyphal Growth
 PAM1/YDR251W 471, 553P, 625 Myr, AbA, Casp, YeastMine N/A N/A
Response to Starvation
ATG21/YPL100W 191, 237P Myr +++ N/A +
*

Genes that are not bioinformatically predicted Ypk1 substrates, but contain Ypk1 motifs and were included in this study are marked with an asterisk. SDL assay results are listed for each bioinformatically predicted Ypk1 substrate. The scoring system reports growth phenotypes of the ypk1-as ypk2Δ strain transformed with the indicated PGAL1-SUBSTRATE plasmid upon overexpression on galactose with varying levels of 3-MB-PP1-imposed Ypk1-as inhibition. A growth phenotype is defined as at least 1 serial dilution spot less growth than YCpLG-GFP control at the given 3-MB-PP1 concentration. A strong growth phenotype is defined as no growth at the given 3-MB-PP1 concentration. (+++++) indicates a growth phenotype with no 3-MB-PP1. (++++) is a strong growth phenotype on 1 μM 3-MB-PP1. (+++) indicates a growth phenotype on 1 μM 3-MB-PP1. (++) is defined as no phenotype on 1 μM 3-MB-PP1, but a strong growth phenotype on 2 μM 3-MB-PP1. (+) indicates no phenotype on 1 μM 3-MB-PP1, but a detectable growth phenotype on 2 μM 3-MB-PP1. (−) indicates no growth phenotype at any concentration of 3-MB-PP1 tested. TOXIC indicates overexpression of the putative substrate on galactose-containing medium was deleterious to growth even in the wild-type control strain (BY4741). These toxic genes were then tested for Ypk1 dosage rescue (for details, see ‘Materials and methods’); here, (+) indicates that Ypk1 overexpression could at least partially rescue the overexpression toxicity of the indicated gene and (−) indicates that Ypk1 overexpression could not rescue the overexpression toxicity. Lastly, the results of testing the indicated purified predicted Ypk1 target as a substrtate in the in vitro protein kinase assay with purified Ypk1-as; here, (+) indicates that Ypk1-as- dependent (3-MB-PP1 inhibitable) incorporation was detectable for the substrate at a level comparable to incorporation into the positive control [the known Ypk1 substrate, GST-Orm1(1–85)]; (+/−) indicates that readily detectable Ypk1-as-dependent incorporation was found, but at a level lower than that seen for an equivalent amount of GST-Orm1(1–85) protein. (N/A) indicates that the indicated gene product was not tested in the indicated assay.

The SDL assay was performed with a plasmid constitutively overexpressing CCH1 under the TDH3 promoter [pBCT-CCH1H, (Iida et al., 2007)], as our efforts to generate a PGAL1-CCH1 vector failed.

As the secondary filter for the candidates recognized bioinformatically, we developed an in vivo approach to identify those gene products that manifested an expected hallmark of protein-substrate interaction. We reasoned that under conditions where the level of activity of an essential kinase, like Ypk1, is near-limiting for normal growth, high-level over-expression of an authentic substrate might tie up the available pool of active enzyme and prevent efficient phosphorylation of other cellular substrates necessary for growth and/or viability (Figure 1C). This scheme is a novel variation on a genetic approach referred to as synthetic dosage lethality (SDL) (Sopko et al., 2006; Sharifpoor et al., 2012). To limit Ypk1 activity, we used ypk1-as ypk2Δ cells, which express from the YPK1 locus an analog-sensitive allele, Ypk1(L424A) (Roelants et al., 2011; Niles et al., 2012), and titrated down its activity by addition of a low concentration of an efficacious inhibitor,1-(tert-butyl)-3-(3-methylbenzyl)-1H-pyrazolo[3,4-d]pyrimidin-4-amine (3-MB-PP1) (Burkard et al., 2007) that has no effect on wild-type cells (see Figure 2C). To achieve high-level over-expression, each bioinformatic hit was expressed from the galactose-inducible GAL1 promoter on a CEN plasmid. As proof of concept, we used two known Ypk1 substrates, Orm1 and Orm2, as positive controls and GFP, which is not a Ypk1 substrate (data not shown), as a negative control. In the absence of limiting the activity of Ypk1(L424A) with inhibitor, overexpression of neither Orm proteins nor GFP was deleterious to cell growth (Figure 1D, left panels, compare lower to upper). However, in the presence of a low dose of 3-MB-PP1, overexpression of Orm1 and Orm2 on galactose medium prevented cell growth, whereas overexpression of GFP did not (Figure 1D, middle and right panels, compare lower to upper). We were able to test the majority of the candidates (90/96) that arose in the bioinformatic search in this same fashion [however, 10/90 caused toxicity upon over-expression even in wild-type cells and, hence, could not be scored].

Figure 2. Lac1 and Lag1, subunits of ceramide synthase, were identified by the screen.

Figure 2.

(A) A diagram of the membrane topology of Lac1 and Lag1 derived from (Kageyama-Yahara and Riezman, 2006); Lac1 and Lag1 are experimentally determined to have eight transmembrane domains. The N terminus of these proteins is cytosolic and therefore accessible for Ypk1 phosphorylation. (B) Diagram of yeast de novo sphingolipid biosynthesis derived from (Dickson, 2008). Metabolites appear as boxes and enzymes as ovals. Metabolites in green are those directly produced or derived from ceramide synthase while those in red are alternative products at the ceramide synthase branch point. (C) SDL results for Lac1 and Lag1. ypk1-as ypk2Δ (yAM135–A) or wildtype (BY4741) cells were transformed with PGAL1-GFP (negative control), PGAL1-Lac1 or PGAL1-Lag1 plasmids. The SDL assay was then performed as described in ‘Materials and methods’. (D) GST-Lac1(1–76) (pAX131), GST-Lag1(1–80)(pFR29), GST-Lac1(1–76)(S23A S24A)(pAX132) and GST-Lag1(1–80)(S23A S24A)(pAX133) were purified from E. coli and Ypk1-as kinase assays were performed as in Figure 1D.

DOI: http://dx.doi.org/10.7554/eLife.03779.005

Those candidates that, like Orm1 and Orm2, exhibited toxicity only on galactose medium and only when Ypk1(L424A) activity was limited in the presence of 3-MB-PP1, but not when inhibitor was absent, were designated SDL hits (Table 1, column 4). Moreover, use of a series of 3-MB-PP1 concentrations allowed for quantification of the strength of the SDL effect (from + to ++++). In one case (Fps1), a marked SDL effect was observed upon overexpression in the ypk1-as ypk2Δ cells in the absence of chemical inhibition; we considered this a valid SDL hit because GAL promoter-driven over-expression of Fps1 was not growth inhibitory in wild-type (YPK1+ YPK2+) cells. Thus, as summarized in Table 1 (column 4), a significant fraction (20/90) of the candidates identified bioinformatically that were tested in this fashion, but far from all, displayed an SDL phenotype. In this regard, it is important to note that all known Ypk1 substrates tested (Gpd1, Fpk1, Orm1 and Orm2) displayed an SDL phenotype, whereas nearly 80% of the bioinformatic hits, like GFP, did not. Consistent with the view that the SDL phenotype could arise from the over-expressed target serving as a decoy substrate that titers a limited pool of active Ypk1 away from acting on its essential substrates, we observed that over-expressed catalytically-inactive Fpk1 caused an SDL phenotype equivalent to or stronger than wild-type Fpk1 (Table 1). If such SDL phenotypes reflect occlusion of a limited pool of enzyme by over-expressed substrate, then, conversely, co-overexpression of Ypk1 or even of a kinase-dead allele Ypk1(K376A) (driven from the MET25 promoter) might rescue the toxicity. Indeed, the deleterious effect of Smp1 over-expression was rescued by co-overexpression of either Ypk1 or Ypk1(K376A) (Table 1), suggesting that the SDL phenotype of over-expressed Smp1 also arises from titration of a limited amount of Ypk1 away from essential substrates.

Lastly, to determine whether the gene products that displayed an SDL phenotype are indeed substrates for Ypk1, we incubated those (17/20) that we were able to successfully express and purify as recombinant proteins or protein fragments (as GST fusions) from Escherichia coli with [γ-32P]ATP and Ypk1(L424A), which was highly purified from yeast cells as described in ‘Materials and methods’. We chose to use the analog-sensitive allele, even though it is only about 50% as active as wild-type Ypk1 (Roelants et al., 2011), because ablation of activity in the presence of 3-MB-PP1 allowed us to confirm that any 32P incorporation into substrate observed was due to phosphorylation by Ypk1(L424A) itself (and not due to some other protein kinase that might be present in the preparation). All known in vivo substrates of Ypk1 (Fpk1, Fpk2, Gpd1, Orm1 and Orm2) display robust incorporation (as judged by autoradiography) in this in vitro assay (Roelants et al., 2010, 2011; Lee et al., 2012b). Therefore, in testing each candidate, an appropriate positive control, Orm1(1–85) was included (Figure 1E), which also allowed comparison between independent assays. Gratifyingly, 12/17 (70%) of the SDL hits tested in this manner displayed readily detectable and Ypk1-specific phosphorylation (Table 1, right column). Moreover, the results of such assays clearly demonstrated that Ypk1 is not a ‘promiscuous’ enzyme. For example, Fps1 has three possible N-terminal Ypk1 sites (RPRGQT147T, RRRSRS181R and RSRATS185N) and one C-terminal site. The C-terminal site is a Ypk1 target (data not shown); however, none of the N-terminal motifs serves as a Ypk1 phospho-acceptor site (Figure 1E), most likely because each lacks a hydrophobic residue at +1.

Thus, we considered it very likely that the dozen candidates identified bioinformatically that also displayed an SDL phenotype and served as Ypk1 substrates in vitro (highlighted in bold in Table 1) would be functionally important Ypk1 substrates in vivo. To validate this conclusion and confirm that these candidates are indeed physiologically relevant Ypk1 targets, we chose to characterize Lac1 and Lag1, two of the dozen candidates (Table 1), because they are the catalytic subunits of the ceramide synthase complex and might further our understanding about how sphingolipid production is regulated by the TORC2-Ypk1 signaling axis.

Ceramide synthase components Lac1 and Lag1 are identified as Ypk1 substrates

Lac1 (418 residues) and Lag1 (411 residues) are apparent paralogs at the primary sequence level (69% identity, 77% similarity) (Byrne and Wolfe, 2005) and are polytopic integral proteins in the ER membrane with the predicted Ypk1 site in each protein residing near its N-terminus (Figure 2A). It has been shown that the N-termini of these proteins are exposed to the cytosol (Kageyama-Yahara and Riezman, 2006). Along with a small accessory subunit Lip1 (150 residues), Lac1 and Lag1 are demonstrated constituents of the ceramide synthase complex (Schorling et al., 2001; Vallée and Riezman, 2005), which catalyzes N-acylation of the free amino group on the long-chain base (LCB), mainly phytosphingosine in yeast, using fatty acyl-CoA as the acyl donor, thereby forming phytoceramide (Figure 2B). Genetically, Lac1 and Lag1 appear to play overlapping functional roles; lac1Δ or lag1Δ single mutants are viable, whereas a lac1Δ lag1Δ double mutant is reportedly either inviable (Jiang et al., 1998) or extremely slow growing (Barz and Walter, 1999; Schorling et al., 2001; Vallée and Riezman, 2005). The ceramide synthase reaction lies at an important branch point in the sphingolipid metabolic network (Figure 2B) because de novo synthesis of ceramides both consumes LCBs and prevents conversion of LCBs to their 1-phosphorylated derivatives (LCBPs). Thus, the rate of ceramide synthesis is tightly coupled to the levels of both LCBs and LCBPs (Kobayashi and Nagiec, 2003); and, moreover, the balance between ceramides and total LCBs and LCBPs affects growth rate in both fungi (Kobayashi and Nagiec, 2003) and mammalian cells (Spiegel and Milstien, 2003). By virtue of their position in the pathway, Lac1 and Lag1 are situated to be important regulators of this balance.

Among the bioinformatically predicted substrates, both Lac1 and Lag1 displayed a readily detectable SDL phenotype (Figure 2C) and both GST-Lac1(1–76) and GST-Lag1(1–80) served as in vitro substrates for Ypk1, albeit with the phosphorylation of Lac1 being reproducibly much more robust than Lag1 (Figure 2D). Site-directed mutagenesis confirmed that the Ypk1-mediated phosphorylation of both substrates occurred exclusively at the predicted phospho-acceptor site(s), specifically Ser23 and Ser24 in both proteins (Figure 2D).

Lac1 and Lag1 are phosphorylated by Ypk1 in vivo

To determine whether both Lac1 and Lag1 are phosphorylated in vivo in a Ypk1-dependent manner and at their Ypk1 consensus sites, integrated 3xHA- or 3xFLAG-tagged versions of each protein and its corresponding S23A S24A mutant were expressed in yeast and extracts of the cells were analyzed by phosphate-affinity SDS-PAGE (Phos-tag gels) (Kinoshita et al., 2009). In this separation technique, the more highly phosphorylated the protein, the more its migration is retarded. Both Lac1 (Figure 3A, left) and Lag1 (Figure 3A, right) migrated as two species, and the slower of the two could be attributed to phosphorylation because it was eliminated if the sample was pre-treated with calf intestinal phosphatase. This slower mobility species represented phosphorylation at S23 and S24 because the band was also eliminated in Lac1(S23A S24A) and Lag1(S23A S24A) mutants (Figure 3A). We noted that phosphatase treatment, even of the Lac1(S23A S24A) and Lag1(S23A S24A) mutants, resulted in appearance of a third, even faster migrating species, presumably due to removal of a phosphorylation(s) elsewhere in these proteins, consistent with indirect evidence that Lac1 and Lag1 might be subject to casein kinase II (yeast Cka2)-dependent modification (Kobayashi and Nagiec, 2003).

Figure 3. Ypk1 phosphorylates Lac1 and Lag1 at S23 and S24 in vivo.

Figure 3.

(A) 3xHA-Lac1 (yAM165–A) 3xHA-Lac1(S23A S24A) (yAM166–A), 3xFLAG-Lag1 (yAM159–A) and 3xFLAG-Lag1(S23A S24A) (yAM163–A) strains were grown to mid-exponential phase in YPD. Cells were then harvested and whole-cell extracts were prepared. Extracts were split and one fraction was then treated with calf intestinal phosphatase. The extracts resolved by Phos-tag SDS-PAGE and immunoblotted with anti-HA, -FLAG or -Pgk1 (loading control) antibodies. P-Lac1 and P-Lag1 indicate the band corresponding to S23 S24 phosphorylation. * indicates a non-specific band that appears in HA blots of yeast whole cell extracts. (B) Wildtype (BY4741), ypk1Δ (JTY6142) and ypk2Δ (yAM120–A) strains were transformed with a plasmid centromeric plasmid encoding 3xHA-Lac1 expressed under control of its endogenous promoter (pAX136). Cells were grown to mid-exponential phase and then treated with a sublethal dose of myriocin (0.625 μM) or methanol (vehicle) for 2 hr. Cell extracts were then prepared, resolved by Phos-tag SDS-PAGE and blotted as above in (A). (C) TOR2 (yKL04) or tor2-as (yKL05) strains were transformed with a 3xHA-Lac1 expressing plasmid (pAX136). Cells were grown to mid-exponential phase and treated with 1 μM BEZ-235 for the indicated times. Cell extracts were then prepared, resolved by Phostag SDS-PAGE and blotted as above in (A).

DOI: http://dx.doi.org/10.7554/eLife.03779.006

In agreement with their relative efficacies as in vitro substrates (Figure 2D), we found that, reproducibly, the majority of Lac1 was present in the cell as the slower mobility isoform, whereas the opposite was true for Lag1 (Figure 3A). Because the behavior of Lac1 gave us greater sensitivity of detection, and for the sake of conciseness, some of our subsequent findings are illustrated with data for Lac1 only. However, all experiments were repeated with both proteins with virtually identical results and conclusions.

To further confirm that Ypk1 is the protein kinase responsible for phosphorylation at Ser23 Ser24 in vivo, plasmids encoding 3xHA-tagged Lac1 and Lag1 were introduced by DNA-mediated transformation into wild-type, ypk1Δ, and ypk2Δ strains. Cultures of the resulting cells were grown in the absence or presence of a sub-lethal dose of myriocin, a condition which several previous studies have shown activates TORC2- and Ypk1-mediated signaling (Roelants et al., 2011; Berchtold et al., 2012; Sun et al., 2012), and the resulting extracts were analyzed on Phos-tag gels. As observed previously for two other bona fide substrates, Orm1 and Orm2 (Roelants et al., 2011), absence of Ypk1 totally abrogated the appearance of phosphorylated Lac1 (Figure 3B) and phosphorylated Lag1 (data not shown), whereas elimination of Ypk2 had no effect. Thus, Ypk1 is the paralog solely responsible for the observed in vivo phosphorylation of Lac1 and Lag1 at Ser23 and Ser24. Furthermore, under conditions that stimulate TORC2 activity, nearly all of the Lac1 (Figure 3B) and much more of the Lag1 (data not shown) were converted to the phosphorylated isoform indicating that TORC2 activation is relayed to ceramide synthase via Ypk1. To further confirm that TORC2 function is essential for Ypk1-mediated phosphorylation of ceramide synthase, Lac1 phosphorylation was monitored in TOR2 cells and in a tor2-as mutant after addition of a specific tor2-as inhibitor, BEZ-235 (Kliegman et al., 2013). TORC2 inhibition markedly reduced Lac1 phosphorylation within 20 min in tor2-as cells, but not in the otherwise isogenic control cells (Figure 3C), confirming that TORC2 activity is necessary for Ypk1-mediated ceramide synthase phosphorylation. Thus, our screening approach was successful in revealing two, previously uncharacterized, cellular targets of the TORC2-Ypk1 signaling axis.

Increased Lac1 and Lag1 phosphorylation is required for cell survival under stress

The fact that impeding LCB production with a sub-lethal dose of the SPT inhibitor myriocin stimulated Ypk1-mediated Lac1 and Lag1 phosphorylation suggested that this modification is important for their physiological function. As one means to confirm that reduction in sphingolipid biosynthesis capacity results in up-regulation of Lac1 and Lag1 phosphorylation, we subjected the pathway to blockade near its end by treating the cells expressing integrated 3xHA-tagged Lac1 or Lag1 with aureobasidin A (Heidler and Radding, 1995), an antibiotic that prevents formation of complex sphingolipids in yeast by inhibiting Aur1 (phosphatidylinositol:ceramide phosphoinositol transferase) (Nagiec et al., 1997). As observed for treatment with myriocin, the amount of phosphorylated Lac1 (Figure 4A) and Lag1 (data not shown) was markedly increased in response to treatment with aureobasidin A.

Figure 4. Enhanced Ypk1 phosphorylation of Lac1 and Lag1 under sphingolipid and heat stress.

Figure 4.

(A) 3xHA-Lac1 (yAM165–A) cells were grown to early exponential phase in YPD. Cultures were then treated with sublethal doses of myriocin (0.625 μM) or methanol (vehicle) or aureobasidin A (1.8 μM) or ethanol (vehicle) for 2 hr. (B) 3xHA-Lac1 (yAM165–A) cells were grown to exponential phase in YPD at 30°C. A sample of this culture was then harvested. The remaining culture was then moved to 42°C to initiate heat shock and samples were harvested at the indicated time points. Whole cell extracts were prepared from each sample, resolved by Phos-tag SDS-PAGE, and immunoblotted as in Figure 3. P-Lac1 indicates the band corresponding to S23 S24 phosphorylation. * indicates a non-specific band that appears in HA blots of yeast whole cell extracts. (C) LAC1 LAG1 (yAM205–A), Lac1SSAA Lag1SSAA (yAM207B) and Lac1SSEE Lag1SSEE (yAM210) were grown to exponential phase in YPD. Serial dilutions of each culture were made and spotted on YPD plates containing vehicle or the indicated concentration of myriocin or aureobasidin A. Cells were allowed to grow for 3 days at 30°C prior to imaging. (D) 3xHA-Lac1::HIS3 3xFLAG-Lag1::LEU2 (yAM168) 3xHA-Lac1(S23A S24A)::HIS3 3xFLAG-Lag1(S23A S24A)::LEU2 (Lac1SSAA Lag1SSAA) (yAM184) and 3xHA-Lac1(S23E S24E)::HIS3 3xFLAG-Lag(S23E S24E)1::LEU2 (Lac1SSEE Lag1SSEE) (yAM192–A) cells were grown to mid-exponential phase and then treated with 1.0 μM myriocin or 18.2 nM aureobasidin A for 2 hr. Whole cell extracts were prepared from each sample, resolved by SDS-PAGE, and immunoblotted as indicated.

DOI: http://dx.doi.org/10.7554/eLife.03779.007

Another perturbation that has been shown to transiently up-regulate TORC2-Ypk1-mediated signaling is heat shock (Sun et al., 2012). Consistent with that response, it has been shown previously that heat shock leads to a transient increase in sphingolipid production and that sphingolipid production is important for heat shock survival (Jenkins et al., 1997; Cowart et al., 2003). Moreover, measurement of pathway intermediates and mathematical modeling also suggested that a sharp spike of increased ceramide synthase activity may occur after heat shock (Chen et al., 2013). Therefore, we subjected the same cells to heat shock and monitored Lac1 and Lag1 phosphorylation at various times thereafter. Within 5 min, the amount of phosphorylated Lac1 (Figure 4B) and Lag1 (data not shown) increased markedly, but was back to the resting level by 30 min.

If these changes in phosphorylation state at Ser23 and Ser24 in Lac1 and Lag1 are important for the metabolic adjustments that the cell needs to adapt appropriately, then preventing phosphorylation at these sites should impair cell survival. To test this prediction, we integrated as the sole source of ceramide synthase, mutant versions of Lac1 and Lag1 in which Ser23 and Ser24 were mutated to Ala and, hence, cannot be phosphorylated under any circumstances. As a control, we also generated integrated versions of Lac1 and Lag1 in which Ser23 and Ser24 were mutated to Glu, to mimic conversion of the entire population to the phosphorylated state, a response that we showed can be achieved for the wild-type protein (see, for example, Figure 3B). Indeed, we found that the cells co-expressing Lac1(S23A S24A) and Lag1(23A S24A) grew detectably less well when challenged with either myriocin or aureobasidin A than either otherwise isogenic wild-type cells or cells co-expressing Lac1(S23E S24E) and Lag1(S23E S24E) (Figure 4C). These growth differences could not be attributed to differences in the level of expression of these proteins, as immunoblot analysis demonstrated the wild-type and mutant ceramide synthase subunits were present in equivalent amounts under the conditions tested (Figure 4D). Thus, Ypk1-dependent phosphorylation of these sites in Lac1 and Lag1 is functionally important for cell survival in response to the stress of limiting sphingolipid biosynthesis. Furthermore, the fact that the Lac1(S23E S24E) Lag1(S23E S24E) strain phenocopied a LAC1+ LAG1+ strain under conditions that promote phosphorylation of Lac1 and Lag1 indicates that these mutations generated effective phosphomimetic alleles.

Calcineurin down-regulates Lac1 and Lag1 phosphorylation at Ser23 and Ser24

As a means to delineate what cellular phosphatase is responsible for counteracting the Ypk1-mediated phospho-regulation of Lac1 and Lag1, plasmid-borne 3xHA-tagged Lac1 was expressed in a collection of deletion strains lacking each of the non-essential protein phosphatase genes or their associated factors, and the phosphorylation state of Lac1 was assessed using Phos-tag gels. By this approach, we were unable to find any phosphatase-deficient mutant that exhibited a significant increase in the amount of phosphorylated Lac1 compared to control cells (data not shown). However, considerable genetic evidence indicates a strong connection between Ca2+ signaling and sphingolipid biosynthesis (Beeler et al., 1998). Moreover, it has been reported previously that TORC2-Ypk-dependent regulation of sphinglipid biosynthesis is antagonized by the action of the Ca2+/calmodulin-dependent protein phosphatase calcineurin (also known as phosphoprotein phosphatase 2B), although the level at which the phosphatase acted was unknown (Aronova et al., 2008). Hence, we conducted the converse experiment by stimulating the cells expressing 3xHA-tagged Lac1 acutely with 0.2 M CaCl2, a condition known to robustly activate calcineurin in yeast (Stathopoulos-Gerontides et al., 1999). Within <10 min, we found total abrogation of phospho-Lac1 (Figure 5A, left) and total abrogation of phospho-Lag1 (data not shown) in wild-type cells, whereas in otherwise isogenic cna1Δ cna2Δ mutants (which lack both calcineurin catalytic subunit paralogs) a substantial portion of the phosphorylated species remained (Figure 5A, right).

Figure 5. Activation of calcineurin leads to rapid dephosphorylation of Lac1 and Lag1 without affecting Ypk1 function.

Figure 5.

(A) Wildtype (BY4741) or cna1Δ cna2Δ (JTY5574) strains were transformed with a centromeric plasmid encoding 3xHA-Lac1 expressed under control of its endogenous promoter (pAX136). Cultures were grown to mid-exponential phase in selective media and then treated with 200 mM CaCl2 for 10 min to activate calcineurin. Cultures were then harvested and whole cell extracts were prepared from each sample, resolved by Phos-tag SDS-PAGE, and immunoblotted as in Figure 3. (B) 3xHA-Lac1 (yAM165-A) cells were transformed with a centromeric plasmid encoding the hyperactive TORC2 independent Ypk1D242A allele expressed under its own promoter (pFR273) or vector (pRS316). Cultures were grown in selective media to mid-exponential phase before treatment with 200 mM CaCl2 for 10 min. Cultures were then harvested and whole cell extracts were prepared from each sample, resolved by Phos-tag SDS-PAGE, and immunoblotted as above. (C) 3xFLAG-Ypk1 (YDB379) cells were grown to mid-exponential phase in YPD. Cultures were treated with or without 200 mM CaCl2 for 10 min. Cells were then harvested and whole extracts prepared. Extracts were resolved by SDS-PAGE and blotted with anti-pSGK(T256), which recognizes Ypk1 T504 activation loop phosphorylation (Roelants et al., 2010) and anti-FLAG antibody (Ypk1 loading control). The blot is representative of triplicate samples and the quantitation of the ratio of pT504/FLAG from these replicates is shown below the blot.

DOI: http://dx.doi.org/10.7554/eLife.03779.008

These results suggested that calcineurin may directly reverse the phoshorylations introduced into Lac1 and Lag1 by Ypk1. However, it is also the case that two of the ancillary subunits associated with TORC2, Slm1 and Slm2, are demonstrated calcineurin-binding proteins (Bultynck et al., 2006; Tabuchi et al., 2006) and that calcineurin action appears to oppose TORC2-dependent signaling (Mulet et al., 2006; Daquinag et al., 2007; Berchtold et al., 2012). Hence, it was possible, therefore, that the observed loss of phospho-Lac1 and -Lag1 might be due to a Ca2+-stimulated and calcineurin-mediated reduction in their TORC2-Ypk1-dependent phosphorylation, rather than to direct action of calcineurin on phospho-Lac1 and -Lag1. To rule out the former possibility, we examined Lac1 and Lag1 phosphorylation in cells expressing constitutively-active Ypk1(D242A), which we have demonstrated bypasses the need for its TORC2-dependent activation (Roelants et al., 2011). We found that Ca2+ addition still led to nearly complete Lac1 and Lag1 dephosphorylation in these cells (Figure 5B). Furthermore, as judged by immunoblotting with a phospho-site specific antibody, Pkh1- and Pkh2-dependent phosphorylation of the activation loop in Ypk1 (Figure 5C) was not diminished in cells stimulated with Ca2+, confirming that there was no calcineurin-mediated decrease in the amount of active Ypk1 present. Therefore, it seems clear that calcineurin dephosphorylates the sites in Lac1 and Lag1 phosphorylated by Ypk1 directly, rather than through down-regulation of Ypk1 function.

Ypk1-mediated phosphorylation of Lac and Lag1 stimulates ceramide synthase activity

Given the phenotypic evidence that TORC2-Ypk1-dependent modulation of Lac1 and Lag1 is physiologically important (Figure 4C), we next conducted biochemical analysis to determine how Ypk1-mediated phosphorylation affects the function of Lac1 and Lag1 in ceramide synthesis. Toward that end, we first used LC-MS to monitor the levels of LCBs and LCBPs extracted from equivalent numbers of cells from exponentially-growing cultures of wild-type cells or otherwise isogenic cells expressing as the sole source of Lac1 and Lag1 either the non-phosphorylatable Lac1(S23A S24A) and Lag1(S23A S24A) mutants or the phosphomimetic Lac1(S23E S24E) and Lag1(S23E S24E) mutants. Relative to wild-type cells, the Lac1(S23A S24A) Lag1(S23A S24A) cells accumulated significantly more PHS (Figure 6A, top left), as well as more dihydrosphingosine (DHS) (Figure 6A, top right), which is a more minor LCB in yeast (note the difference in scale of the ordinate), whereas the Lac1(S23E S24E) Lag1(S23E S24E) cells displayed a level of both PHS and DHS that was somewhat lower, and a similar trend was observed even for PHS-1P, an even less abundant metabolite (Figure 6A, bottom left). The level of DHS-1P (Figure 6A, bottom right) was so low as to make reliable measurement difficult, but no differences between strains could be detected. Given that the ceramide synthase complex is responsible for the conversion of LCBs into ceramides, these findings indicate that, in the absence of Ypk1-mediated phosphorylation, the rate of LCB utilization is significantly reduced, consistent with the conclusion that Ypk1-dependent modification of Lac1 and Lag1 promotes ceramide synthase function.

Figure 6. Ypk1 phosphorylation of Lac1 and Lag1 stimulates ceramide synthase activity.

Figure 6.

(A) Cultures of LAC1 LAG1 (yAM205–A), Lac1SSAA Lag1SSAA (yAM207B) and Lac1SSEE Lag1SSEE (yAM210) strains were grown to mid-exponential phase and then harvested. Sphingolipids were extracted and analyzed as described in ‘Materials and methods’. Values represent the mean of three independent experiments (each performed in triplicate) and error bars represent SEM. (B) Triplicate exponentially-growing cultures of LAC1 LAG1 (yAM205–A), Lac1SSAA Lag1SSAA (yAM207–B) and Lac1SSEE Lag1SSEE (yAM210) were grown overnight and duplicate cultures were diluted to OD600 = 1.0. Complex sphingolipids were labeled and analyzed by thin layer chromatography (TLC) as in ‘Materials and methods’. A representative TLC plate is shown with the origin at the bottom of the image. The assigned identity of species as IPCs and MIPCs was confirmed by pharmacological or genetic inhibition of the production of these species in control cultures (data not shown). Quantification of total complex sphingolipids was performed in ImageJ by integrating the Phosphorimager screen intensity across the lane for each sample and normalized to 1 for the wild-type ceramide synthase samples. (C) Upper, ceramide synthase was purified by FLAG immunopurification from 3xHA-Lac1::HIS3 3xFLAG-Lag1::LEU2 (yAM168) 3xHA-Lac1(S23A S24A)::HIS3 3xFLAG-Lag1(S23A S24A)::LEU2 (Lac1SSAA Lag1SSAA) (yAM184) and 3xHA-Lac1(S23E S24E)::HIS3 3xFLAG-Lag(S23E S24E)1::LEU2 (Lac1SSEE Lag1SSEE) (yAM192–A) cells. Immunoprecipitates were then split into three fractions and in vitro ceramide synthase assays (60 min reactions) were performed in triplicate. A small sample of each ceramide synthase assay was resolved by SDS-PAGE and immunoblotted. The signal intensity quantified from the immunoblot was then used to normalize ceramide synthase activity in each sample. Lower, ceramide synthase was immunopurified from 3xHA-Lac1::HIS3 3xFLAG-Lag1::LEU2 (yAM168) 3xHA-Lac1(S23A S24A)::HIS3 or 3xFLAG-Lag1(S23A S24A)::LEU2 (Lac1SSAA Lag1SSAA) (yAM184) cells as above except cultures were treated with 1.0 μM myriocin or methanol (vehicle) prior to harvesting. Values represent the mean of three independent experiments (each performed in triplicate) and error bars represent SEM. Statistical significance of values (Student's t test): *p = <0.05, **p = <0.009; and, ***p < 0.0009.

DOI: http://dx.doi.org/10.7554/eLife.03779.009

As an independent means to measure flux through the sphingolipid pathway, and because all complex sphingolipids in yeast contain inositol-phosphate, an equivalent number of cells of the same three strains in mid-exponential phase were pulse-labeled, in triplicate, with [32P]PO4−3, and the acidic sphingolipids extracted and analyzed by thin-layer chromatography. Strikingly, the amount of complex sphingolipids generated during the pulse was reproducibly higher in the Lac1(S23E S24E) Lag1(S23E S24E) cells than in the wild-type controls, and Lac1(S23A S24A) Lag1(S23A S24A) generated levels of complex sphingolipids lower than the wild-type controls (Figure 6B). These findings are again consistent with the conclusion that Ypk1-mediated phosphorylation of Lac1 and Lag1 stimulates the production of the ceramide precursors to complex sphingolipids.

Ypk1-mediated phosphorylation could stimulate the ceramide synthase reaction in vivo by stabilizing Lac1 and Lag1 thereby increasing their steady-state level, by enhancing their association with the small ancillary subunit Lip1, and/or by direct activation. Immunoblotting of exponentially-growing cultures expressing the 3xHA-tagged versions of wild-type Lac1 and Lag1 and the Lac1(S23A S24A) Lag1(S23A S24A) and Lac1(S23E S24E) Lag1(S23E S24E) indicated no discernible difference in their steady-state level (see Figure 4D). Likewise, in cells co-expressing the same proteins and FLAG-tagged Lip1 (gift of Howard Riezman, Univ. of Geneva), we observed no difference in the efficiency of Lip1 co-immunoprecipitation between wild-type Lac1 and Lag1 and either the Lac1(S23A S24A) Lag1(S23A S24A) or Lac1(S23E S24E) Lag1(S23E S24E) mutants (data not shown). These results suggested that Ypk1-mediated phosphorylation may directly enhance the catalytic efficiency of Lac1 and Lag1. To test this possibility directly, 3xFLAG-tagged versions of Lac1 and Lag1 were immunopurified from detergent-solubilized microsomes isolated from exponentially-growing cells and equivalent amounts of the resulting protein assayed in vitro, monitoring the formation of ceramide from PHS and steroyl-CoA by LC-MS. No product was observed in the absence of added steroyl-CoA (data not shown), and product formation was reduced by 85% in the presence of PHS and steroyl-CoA if 1 μM australifungin, a demonstrated and specific ceramide synthase inhibitor (Mandala and Harris, 2000), was added (data not shown), confirming that the reaction measured was catalyzed by ceramide synthase. We found that the specific activity of the Lac1(S23E S24E) Lag1(S23E S24E) complex was reproducibly ∼2 higher than either the Lac1(S23A S24A) Lag1(S23A S24A) mutant or the wild-type complex (Figure 6C, upper panel), similar to the degree of difference in sphingolipid metabolites between these same strains that we measured by other means (Figure 6A,B). In two independent trials (each performed in triplicate), the identical trend was found when microsomes from these same cells were assayed directly (i.e. without detergent solubilization and enzyme enrichment by immunoprecipitation [data not shown]). Additionally, we found that immunopurified ceramide synthase from wild-type cultures treated with myriocin had higher activity than ceramide synthase prepared from untreated cells (Figure 6C, lower panel). Furthermore, this increase in ceramide synthase activity in response to myriocin treatment was not observed in Lac1(S23A S24A) Lag1(S23A S24A) cells (Figure 6C, lower panel), consistent with TORC2-Ypk1 signaling increasing ceramide synthase activity by phosphorylation at these residues. These findings are also in agreement with a reported ∼two-fold decrease in the rate of ceramide production by ceramide synthase complex isolated from TORC2-deficient yeast (Aronova et al., 2008). Hence, we conclude that Ypk1 phosphorylation directly increases the catalytic activity of ceramide synthase.

Ypk1-mediated stimulation of Lac1 and Lag1 prevents autophagy induction during TORC2-driven up-regulation of sphingolipid biosynthesis

The observed increase in ceramide synthase activity in response to Ypk1-mediated phosphorylation could serve two roles that are not mutually exclusive: (i) to produce more ceramide and the derived complex sphingolipids; and, (ii) to prevent inadvertent accumulation of LCBs and the derived LCBPs when TORC2-driven Ypk1 activation stimulates metabolic flow into the sphingolipid pathway by alleviating Orm1- and Orm2-imposed inhibition of SPT (Roelants et al., 2011). We reasoned that if the latter were one of the important physiological functions of Ypk1-dependent stimulation of ceramide synthase, then TORC2 activation would be detrimental to cells in which the Lac1(S23A S24A) Lag1(S23A S24A) mutant is the sole source of this enzyme. To mimic TORC2-stimulated elevation of Ypk1 activity, the constitutively-active Ypk1(D242A) allele (hereafter referred to as Ypk1*) was expressed from the YPK1 promoter on a CEN plasmid in either wild-type cells or the Lac1(S23A S24A) Lag1(S23A S24A) mutant. Indeed, compared to the empty vector control, expression of Ypk1* was well tolerated by cells containing wild-type Lac1 and Lag1, but deleterious to the growth of the Lac1(S23A S24A) Lag1(S23A S24A) mutant cells, whether measured on agar plates (Figure 7A, top) or in liquid culture (Figure 7A, bottom).

Figure 7. Failure of Ypk1 to upregulate ceramide synthase causes LCBP accumulation that triggers autophagy.

Figure 7.

(A) LAC1 LAG1 (yAM205–A) or Lac1SSAA Lag1SSAA (yAM207–B) were transformed with PYPK1-Ypk1D242A (shown as Ypk1*) (pFR273) or empty vector pRS316 (EV). Transformants were grown to exponential phase in synthetic complete medium and then diluted to OD600 = 0.1 and grown in microtiter plates (lower) or on agar plates (upper). For liquid cultures, each was grown in at least quadruplicate and the error bars indicate the SEM of replicates at each time point. (B) Cells from (A) were grown to mid-exponential phase in selective synthetic complete media and then harvested. Sphingolipids were extracted and analyzed as described in ‘Materials and methods’. Values represent the mean of three independent experiments (each performed in triplicate) and error bars represent SEM. (C) LAC1 LAG1 lcb4Δ (yAM237) or Lac1SSAA Lag1SSAA lcb4Δ (yAM238–A) were transformed with Ypk1* (pFR273) and growth experiments performed as in (A). (D) Liquid growth assays were performed as in (A) for LAC1 LAG1 (yAM205–A), Lac1SSAA Lag1SSAA (yAM207–B), Lac1SSEE Lag1SSEE (yAM210), LAC1 LAG1 lcb3Δ (yGT12), Lac1SSAA Lag1SSAA lcb3Δ (yGT13) and Lac1SSEE Lag1SSEE lcb3Δ (yGT14) strains. (E) LAC1 LAG1 (yAM205–A) or Lac1SSAA Lag1SSAA (yAM207–B) or LAC1 LAG1 lcb3Δ (yGT12) strains were transformed with Ypk1* (pFR273) or pRS316 (EV) and additionally PTPI1-GFP-Atg8. Growing cultures treated with vehicle or 2 μg/ml rapamycin for 2 hr and then harvested and whole extracts prepared. Extracts were resolved by SDS-PAGE and blotted with anti-GFP to detect GFP-Atg8 and free GFP arising from GFP-Atg8 autophagic processing and anti-Pgk1 antibody. The blot is representative of triplicate samples and the quantitation of the ratio of free GFP/Pgk1 from these replicates is shown below the blot. (F) LAC1 LAG1 atg1Δ (yAM239–A) or Lac1SSAA Lag1SSAA atg1Δ (yAM240–A) sensitivity to Ypk1* was measured as in (A). (G) Cells from (F) were grown to mid-exponential phase in selective synthetic complete media and then harvested. Sphingolipids were extracted and analyzed as described in (B).

DOI: http://dx.doi.org/10.7554/eLife.03779.010

Given that accumulation of LCBPs has been shown to be toxic to yeast cell growth (Kim et al., 2000), we reasoned that the most likely metabolic perturbation responsible for the observed decrease in growth in the Lac1(S23A S24A) Lag1(S23A S24A) cells expressing Ypk1* was the build-up of LCBs and derived LCBPs. Consistent with this conclusion, we found a reproducible and statistically significant increase in the LCBP level in Lac1(S23A S24A) Lag1(S23A S24A) cells, compared to LAC1+ LAG1+ controls, and a further increase in Lac1(S23A S24A) Lag1(S23A S24A) cells expressing Ypk1* (Figure 7B). If accumulation of LCBPs in Lac1(S23A S24A) Lag1(S23A S24A) co-expressing Ypk1* is indeed responsible for the poor growth, then reduction in LCBP synthesis by elimination of the gene LCB4, which encodes the major LCB kinase (Nagiec et al., 1998), should alleviate the growth inhibition. As expected, introduction of an lcb4Δ null mutation into the Lac1(S23A S24A) Lag1(S23A S24A) mutant suppressed the growth inhibitory effect of Ypk1* (Figure 7C). Conversely, and consistent with toxicity arising from accumulation of LCBPs when Lac1 and Lag1 cannot be stimulated by Ypk1, we found that the poor growth phenotype of cells lacking the gene (LCB3) encoding the major LCBP phosphatase (Mandala et al., 1998), was markedly exacerbated by introduction of the Lac1(S23A S24A) Lag1(S23A S24A) alleles, but not by introduction of the Lac1(S23E S24E) Lag1(S23E S24E) alleles (Figure 7D). In fact, and strikingly, presence of the Lac1(S23E S24E) Lag1(S23E S24E) allele afforded nearly complete rescue of the slow-growth phenotype of the lcb3Δ mutation (Figure 7D). Consistent with the toxicity of LCBPs when LCB utilization by ceramide synthase is inefficient, an elo3 mutation, which prevents efficient formation of C26-CoA (an acyl chain found in yeast ceramides), was synthetically lethal with lcb3Δ (Kobayashi and Nagiec, 2003).

Finally, it has been reported that aberrant increases in LCBP level impede growth by triggering autophagy even under nutrient-rich conditions (Zimmermann et al., 2013). In agreement with the poor growth arising from LCBP-evoked induction of autophagy, we found, first, that Lac1(S23A S24A) Lag1(S23A S24A) cells expressing Ypk1* exhibited a readily detectable increase in basal GFP-Atg8 processing, comparable to that in lcb3Δ cells (but, of course, much less than that occurring when cells were treated with the starvation mimetic rapamycin) (Figure 7E). Second, we found that preventing autophagy by ablating the gene ATG1, which encodes a protein kinase necessary for induction of autophagophore formation and its elongation (Papinski and Kraft, 2014), provided substantial rescue of the growth debilitating effect of Ypk1* expression in Lac1(S23A S24A) Lag1(S23A S24A) cells (Figure 7F, compare to Figure 7A). This rescue was not due to an indirect effect of the absence of Atg1 on LCBP level because introduction of the atg1Δ mutation did not prevent the observed hyper-accumulation of LCBP in Lac1(S23A S24A) Lag1(S23A S24A) cells expressing Ypk1* (Figure 7G). Collectively, these results indicate that, aside from stimulating ceramide synthesis per se, another physiologically important role of Ypk1-dependent ceramide synthase activation is, at least in part, to prevent hyper-accumulation of LCBPs and thereby avoid inadvertent induction of autophagy under nutrient-sufficient conditions (Figure 8).

Figure 8. TORC2-Ypk1 signaling globally activates sphingolipid synthesis, selectively directs flux toward ceramide metabolites, and prevents LCBP cross-talk to the autophagy pathway.

Figure 8.

(A) Diagram of yeast de novo sphingolipid biosynthesis shown is derived from (Dickson, 2008). Enzymes are in ovals. Metabolites are in boxes. Increased color intensity indicates level of metabolite increase in response TORC2-Ypk1 activation. TORC2-Ypk1 signaling globally activates de novo sphingolipid biosynthesis via derepression of the SPT complex (Roelants et al., 2011; Berchtold et al., 2012; Sun et al., 2012), potentially increasing levels of all metabolites. However, Ypk1 also upregulates ceramide synthesis via phosphorylation of Lac1 and Lag1, thus primarily directing this increased flux towards ceramides and away from LCBs and LCBPs. (B) In the absence of Ypk1 mediated ceramide biosynthesis regulation, increased sphingolipid flux raises LCB and LCBP levels. This slows cell growth by activating autophagy. Thus, TORC2-Ypk1 signaling not only activates sphingolipid biosynthesis in response to stress, but also insulates this flux towards ceramides to prevent metabolite mediated crosstalk to the autophagy machinery.

DOI: http://dx.doi.org/10.7554/eLife.03779.011

Discussion

Identification of new TORC2-Ypk1 substrates

Our approach identified Lac1 and Lag1 as potential Ypk1 targets, and our subsequent characterization demonstrated unequivocally that both Lac1 and Lag1 are bona fide Ypk1 substrates in vivo and that their Ypk1-dependent modification is biologically important for optimal modulation of sphingolipid metabolism. These findings validate the ability of our methods for discovery of physiologically relevant protein kinase substrates. Phospho-acceptor site motif pattern-matching alone, although useful in identifying kinase substrates in some cases (Yaffe et al., 2001; Manning et al., 2002; Mah et al., 2005; Holt et al., 2007; Linding et al., 2007; Rennefahrt et al., 2007; Gwinn et al., 2008; Hutti et al., 2009), can yield a large number of false positives. This possibility was a significant concern for us because certain features of the basophilic Ypk1 motif are shared with other protein kinases (Mok et al., 2010). As a means to avoid this problem, we devised a novel SDL-based genetic approach to apply as a secondary filter to parse the bioinformatically selected candidates further.

Although not the only possible explanation for detecting an SDL hit, one mechanism our genetic method should be able to assess is whether the candidate gene product displays a primary characteristic of a true substrate, namely the ability to compete with other substrates for association with Ypk1. A genuine Ypk1 substrate should, when highly over-expressed, sequester a large fraction of the available pool of active kinase, and thus impede its actions on its essential cellular targets. In the absence of Ypk2, over-expression of an authentic Ypk1 substrate should therefore be deleterious for growth when the activity of an analog-sensitive Ypk1 allele is reduced with a selective inhibitor. The fact that 70% of the SDL hits were indeed substrates for Ypk1-mediated phosphorylation in vitro verified that our use of this genetic method as a secondary filter to pick out true substrates from the list of bioinformatically identified candidates was well justified. In this regard, overexpression of Fpk1(KD), a catalytically-inactive non-essential Ypk1 substrate, yielded a readily detectable SDL phenotype. Thus, using the protocol we devised, SDL may be a generally useful way to identify substrates and perhaps other binding partners of protein kinases, not just those directly connected to any output phenotype being measured.

Other genetic approaches have also been useful in identifying physiologically relevant targets of Ypk1. For example, a transposon insertion that suppressed the growth defect of a ypk1-ts ypk2Δ strain at an otherwise non-permissive temperature initially identifed Orm2 as a potential Ypk1 substrate (Roelants et al., 2002), a finding later corroborated by us (Roelants et al., 2011) and others (Berchtold et al., 2012; Liu et al., 2012; Niles et al., 2012; Sun et al., 2012). Similarly, Smp1, which we confirmed here is a likely Ypk1 substrate, was first identified as a potential Ypk1 target because it was isolated as a dosage suppressor of the temperature-sensitive phenotype of ypk1-ts ypk2Δ cells (Roelants et al., 2002). Smp1 is a transcription factor (de Nadal et al., 2003) that mediates iron toxicity (Lee et al., 2012a) and, similarly, Ypk1 is required for iron toxicity (Lee et al., 2012a). Thus, TORC2-Ypk1 signaling might be mechanistically coupled to iron metabolism by modulation of Smp1 transcriptional output.

Likewise, other situations where chemical genetics can be applied have proven useful in gleaning what aspects of cell function are controlled by a protein kinase. For example, mutagenesis of yeast Tor2 to confer susceptibility to a chemical inhibitor and thereby selectively inhibit TORC2 action has been achieved recently (Kliegman et al., 2013). This tool was combined with a coilection of deletion mutants to identify what processes, when eliminated, are especially deleterious to cell growth and survival when TORC2 action (and presumably Ypk1 activity) is limiting. This analysis suggested some connection between TORC2 action and the pentose-phosphate pathway (Kliegman et al., 2013), in keeping with the growth-promoting roles of both TORC1 and TORC2 and the demand for NADPH in many cellular anabolic reactions. Similarly, use of TOR inhibitors has implicated TORC2-Ypk1 signaling in regulation of actin filament formation that is somehow required for yeast cell survival in response to low levels of DNA damage (Shimada et al., 2013).

By contrast, the chemical genetic approach in our SDL method is quite different, in that it scores the deleterious effect arising from overexpression of a gene product (increased protein dosage) rather than from the total absence of a gene product, when activity of the kinase of interest is limited by inhibitor. Theoretically, under our conditions, we should also have been able to observe synthetic dosage rescue (‘SDR’); however, we found no such examples. In any event, our method independently identified nearly all of the previously known in vivo substrates of Ypk1, as well as nearly a dozen genuine Ypk1 targets, including Lac1 and Lag1, that have only been pinpointed by our three-tiered method. Thus, our SDL screening technique provides a complementary approach for identifying substrates of a protein kinase above and beyond those accessible through genetic interactions between the kinase and single-gene deletions or other genetic schemes.

Many of the gene products identified by our screen are involved in processes that the TORC2-Ypk1 signaling axis is already known to regulate. For example, Ypk1 regulates glycerol-3-phosphate production via phosphorylation and inhibition of glycerol-3-phosphate dehydrogenase Gpd1 (Lee et al., 2012b). Interesting, a very likely Ypk1 target that met all of the criteria in our screen is Gpt2, sn-glycerol-3-phosphate 1-acyltransferase (Zheng and Zou, 2001), an enzyme that esterifies glycerol-3-phosphate as the first step in glycerolipid formation. In this same regard, as another very likely Ypk1 target we also identified Fps1, a membrane channel (aquaglyceroporin) that regulates efflux of glycerol (a glycerol-3-phosphate-derived metabolite) (Luyten et al., 1995). These results strengthen the conclusion that TORC2-Ypk1 signaling is intimately involved in modulating the level of the precursor to both glycerophospholipids and the osmolyte glycerol (Lee et al., 2012b).

Ypk1 action has been implicated in regulation of both fluid phase and receptor-mediated endocytosis (deHart et al., 2002). In this regard, we identified as a very likely Ypk1 target an endocytic adaptor, the α-arrestin Rod1, which is necessary for ubiquitinylation-triggered internalization of nutrient permeases (Lin et al., 2008; Becuwe et al., 2012) and the pheromone receptor Ste2 (Alvaro et al., 2014). Thus, as it does for PM lipids, TORC2-Ypk1 signaling may modulate PM protein composition via this α-arrestin, a possibility we are pursuing.

Ypk1 function has also been implicated in regulating production of reactive oxygen species (ROS) (Niles et al., 2014), but an as yet undefined mechanism. In our screen, we found Ysp2, a protein that regulates mitochondrial morphology and ROS levels (Sokolov et al., 2006), as a very likely Ypk1 substrate, possibly providing insight into the molecular basis of the connection between TORC2-Ypk1 signaling and ROS levels.

Similarly, several other prospects that were identified by our screen as very likely Ypk1 substrates remain to be validated. Such candidates include Muk1, a GEF for yeast Rab 5-type small GTPases (Vps21, Ypt52, and Ypt53) involved in vesicle-mediated Golgi body-to-endosome trafficking (Paulsel et al., 2013), suggesting that Ypk1 may also control switches that direct the flow of lipids. Muk1 is also intriguing for another potential reason. In Schizosaccharomyces pombe, a Rab 5-like GTPase (Ryh1) and its Muk1-like GEF were identified in a screen for TORC2 activators (Tatebe et al., 2010). Thus, if Ypk1-mediated phosphorylation inhibits Muk1 function, it could represent a negative feedback mechanism exerted on TORC2; conversely, if Ypk1-mediated phosphorylation stimulates Muk1 function, it could represent a mechanism for self-reinforcing maintenance of TORC2 activity and, thus, a high level of activated Ypk1. Clearly, by further investigating the physiological relevance of these and other remaining candidates much new biology may be learned. Indeed, several gene products of totally unknown function, such as Yhr097c and Ynr014w, as well as gene products (e.g., Atg21 and Pex31) not previous linked to either TORC2 or Ypk1, if validated, may provide new mechanistic insight into additional cellular processes regulated by TORC2-Ypk1 signaling.

Mechanism of phosphoregulation of ceramide synthase

As we have demonstrated here, TORC2-dependent Ypk1-mediated phosphorylation of Lac1 and Lag1 stimulates the function of the ceramide synthase complex. Consistent with our findings, a previous study found a consistent decrease in ceramide synthase activity in microsomal fractions isolated from cells in which TORC2 had been inactivated (and thus Ypk1 activity was presumably reduced) (Aronova et al., 2008), although indirect effects of the loss of TORC2 function on ceramide synthase activity could not be ruled out. Our findings make it clear that the role of TORC2 is to promote the Ypk1-dependent phosphorylation of Lac1 and Lag1 subunits of this enzyme. However, the precise molecular mechanism by which this post-translational modification stimulates this enzyme is still not completely clear. In this regard, it has been shown that mammalian ceramide synthase activity increases upon heterodimerization of different catalytic subunit isoforms (Laviad et al., 2012). However, as judged by co-immunoprecipitation, we found no difference in the state of Lac1-Lag1 association between the wild-type proteins and either our Lac1(S23A S24A) Lag1(S23A S24A) or Lac1(S23E S24E) Lag1(S23E S24E) mutants (data not shown). As mentioned in Results, we found no difference in the steady-state level of these same complexes or in their content of Lip1, a non-catalytic component of the complexes also essential for ceramide synthase activity (Vallée and Riezman, 2005). Thus, understanding of how phosphorylation of Ser23 and Ser24 in Lac1 and Lag1 stimulate ceramide synthase activity may require detailed structural information, which will be challenging to obtain for these polytopic integral membrane proteins.

TORC2-Ypk1 control of ceramide synthesis is conserved

Although the enzymic steps that carry out sphingolipid biosynthesis have been largely elucidated, much less was known, until recently, about regulation of these enzymes (Breslow and Weissman, 2010; Breslow, 2013). The first insight came when it was demonstrated (Roelants et al., 2011; Berchtold et al., 2012; Sun et al., 2012) that, in response to PM stresses, including treatment with myriocin and aureobasidin A, TORC2-Ypk1 signaling is activated and alleviates inhibition of the SPT complex by phosphorylating the negative regulatory proteins Orm1 and Orm2 (Figure 8). As a direct consequence, the rate of de novo production of the LCB precursor to sphingolipids is increased. Although TORC2 signaling had been implicated in promoting synthesis of ceramide, the product of LCB N-acylation (Aronova et al., 2008), it was unknown whether that role was simply the result of TORC2-Ypk1-dependent stimulation of SPT function and the resulting increase in LCB supply. As we demonstrated here, Ypk1-mediated phosphorylation of Lac1 and Lag1 also increases in response to both myriocin and aureobasidin A, suggesting that ceramide synthesis per se, and not simply general elevation of LCB levels, is important for allowing the cells to cope with the effects of these antibiotics. Indeed, collectively, the findings we describe here demonstrate unequivocally that, in addition to up-regulation of SPT, TORC2-Ypk1 exerts direct control on the ceramide synthase step of the sphingolipid biosynthetic pathway by phosphorylating and stimulating the function of the Lac1 and Lag1 subunits of the ceramide synthase complex. Moreover, as we also demonstrated, the ceramide synthase reaction represents an important branch point in sphingolipid biosynthesis (Figures 2B and 8). LCBs produced by the SPT reaction can either be converted to ceramides or become phosphorylated by LCB kinase Lcb4 (and its paralog Lcb5) to form LCBPs (Nagiec et al., 1998). Accumulation of LCBPs has been shown to be toxic to yeast cell growth (Kim et al., 2000), as we have also confirmed here, at least in large part because, as is now known, these metabolites trigger inappropriate induction of autophagy (Zimmermann et al., 2013). Thus, the rate of ceramide production must be properly adjusted to maintain the pool of LCBs and derived LCBPs at a non-deleterious level, in agreement with evidence in yeast and other organisms that ceramides and LCBPs generally play antagonist roles and must be maintained in the proper dynamic balance (Kobayashi and Nagiec, 2003; Spiegel and Milstien, 2003; Kihara et al., 2007; Dickson, 2008; Breslow and Weissman, 2010; Bikman and Summers, 2011). Hence, the function we have discovered and described here for TORC2-Ypk1 in stimulating ceramide synthase promotes utilization of the increased LCB generated upon TORC2-Ypk1-mediated up-regulation of SPT. This metabolic control has multiple clear-cut physiological benefits to the cell: (a) directing flow in the sphingolipid pathway toward complex sphingolipids to populate the PM barrier; (b) reduction of the level of potentially toxic LCBPs; and, (c) avoidance of inappropriate induction of autophagy under nutrient-sufficient conditions (Figure 8). Indeed, our results indicate that a significant role for the coordination exerted by TORC2-Ypk1 between the level of SPT activity and the level of ceramide synthase activity is to prevent metabolic ‘cross-talk’ to the autophagy pathway.

If this function of TORC2-Ypk1 signaling is important, then it should be conserved. Indeed, alignments of the primary structures of Lac1 and Lag1 homologs predicted from sequenced fungal genomes, from Saccharomyces sensu stricto species to the very distantly related Ustilago maydis, nearly all contain a basophilic sequence that matches the Ypk1 phospho-acceptor site consensus and is located, as in S. cerevisiae Lac1 and Lag1, in their N-terminal cytosolic extensions, suggests that Ypk1-related protein kinases may regulate ceramide synthase across essentially all fungal species. Consistent with this suggestion, a genetic interaction between a Ypk1 homolog (YpkA) and a Lac1/Lag1-like ceramide synthase component (BarA1) has been reported in Aspergillus nidulans (Colabardini et al., 2013).

There is also evidence that Orm1 and Orm2, when phosphorylated at unique sites by protein kinase Npr1, promotes steps in the sphingolipid pathway that lead to more complex sphingolipids (Shimobayashi et al., 2013). In contrast to Ypk1, which is activated by TORC2, Npr1 is inhibited by TORC1 (MacGurn et al., 2011). Thus, this control mechanism will only be exerted under conditions that inactivate TORC1, such as amino acid starvation (Loewith and Hall, 2011), a condition that presumably requires adjustment of both PM lipid and protein composition to maximize the cell's ability to scavenge and assimilate nutrients. Under the same condition, autophagy is induced because, like Npr1, TORC1 also negatively regulates the autophagy-inducing protein kinase Atg1-Atg13 complex (Alers et al., 2014). Conversely, under nutrient sufficient conditions, TORC1 is active, and phosphorylates and stimulates protein kinase Sch9. Interestingly, Sch9 action should act in concert with TORC2-Ypk1 signaling to help keep the levels of LCBs and LCBPs low and ceramides high. This is likely because Sch9 promotes transcriptional repression of genes (YDC1 and YPC1) that encode ceramidases and inhibits a phosphosphingolipid phospholipase C (Isc1) that hydrolyzes complex sphingolipids (Swinnen et al., 2014). Such complex multi-component controls may be a general feature of signaling modalities that interface with biosynthetic pathways that have intermediates, like LCBPs, that are not inocuous, but are themselves bioactive metabolites.

Calcineurin negatively regulates ceramide synthase

Prior studies had suggested that calcineurin negatively regulates sphingolipid production via effects on the function of the ancillary TORC2 subunits, Slm1 and Slm2 (Bultynck et al., 2006; Mulet et al., 2006; Tabuchi et al., 2006; Daquinag et al., 2007), although the molecular connection between Slm1 and Slm2 and sphingolipid biosynthesis was unclear. Subsequently, it was observed that, in cells lacking the regulatory subunit (Cnb1), there was an increase in C26-containing ceramides, suggesting that calcineurin somehow antagonizes TORC2-dependent signaling (Aronova et al., 2008). We found that calcineurin negatively regulates ceramide synthesis by directing the dephosphorylation of Lac1 and Lag1. This Ca2+-activated calcineurin-dependent dephosphorylation occurred even in cells expressing a TORC2-independent constitutively-active Ypk1 allele. Moreover, we showed here that calcineurin does not affect Pkh1- (and Pkh2-) mediated phosphorylation of the activation loop of Ypk1, and we demonstrated previously that presence or absence of calcineurin does not alter TORC2-mediated phosphorylation of Ypk1 or cause any substantial change in Ypk1 specific activity (Roelants et al., 2011). Thus, direct down-modulation of either TORC2 or Ypk1 by calcineurin cannot account for the negative regulation it exerts on sphingolipid biosynthesis. What our findings now make clear is that calcineurin negatively regulates the sphingolipid pathway at the level of ceramide synthesis, at least in large part, by direct dephosphorylation of the stimulatory phosphorylations in the ceramide synthase subunits Lac1 and Lag1 that are installed by TORC2-Ypk1 signaling. Calcineurin recognizes substrates via a docking motif (PxIxIT or variants thereof), typically also accompanied quite a distance upstream by a secondary docking site (LxVP) (Roy and Cyert, 2009). However, these sites can be quite degenerate; for example, the more hydrophobic variant PVIVIT is much more potent in recruiting calcineurin when it is used to replace the native ‘PxIxIT’ sequences in either the transcription factor Crz1 (PIISIQ) (Roy et al., 2007) or the endocytic adaptor Aly1 (PILKIN) (O'Donnell et al., 2013). In both Lac1 and Lag1, there is a similar hydrophobic sequence located at the identical position in both proteins (355PIVFVL360). Whether this or any other degenerate match represents a calcineurin-binding site remains to be determined.

In conclusion, our screening approach has provided a number of new insights into how the TORC2-Ypk1 axis modulates sphingolipid homeostasis and has uncovered a significant number of candidate substrates that will very likely shed further light on other aspects of cellular function that are regulated by TORC2-Ypk1 signaling.

Materials and methods

Construction of yeast strains

All S. cerevisiae strains used in this study are listed in Table 2. Strains were constructed using standard yeast genetic manipulations (Burke et al., 2005). For all strains constructed, integration of the desired DNA fragment into the correct genomic loci was confirmed by PCR using an oligonucleotide complementary to the integrated DNA fragment and an oligonucleotide complementary to genomic sequence at least 150 bases away from the integration site.

Table 2.

Saccharomyces cerevisiae strains used in this study

DOI: http://dx.doi.org/10.7554/eLife.03779.012

Strain Genotype Source/reference
BY4741 MATa his3Δ1 leu2Δ0 met15Δ0 ura3Δ0 Research Genetics, Inc.
BY4742 MATα his3Δ1 leu2Δ0 lys2Δ0 ura3Δ0 Research Genetics, Inc.
yAM135-A BY4741 Ypk1(L424A)::URA3-ypk2Δ::KanMX4 This study
JTY6142 BY4741 ypk1Δ::KanMX4 Research Genetics, Inc.
yAM120-A BY4741 ypk2Δ::KanMX4 This study
yAM159-A BY4741 3xFLAG-Lag1::LEU2 This study
yAM163-A BY4741 3xFLAG-Lag1(S23A S24A)::LEU2 This study
yAM165-A BY4742 3xHA-Lac1::HIS3 This study
yAM166-A BY4742 3xHA-Lac1 (S23A S24A)::HIS3 This study
JTY5574 BY4741 cna1Δ::KanMX4 cna2Δ::KanMX4 M.S. Cyert, Stanford Univ.
YDB379 BY4741 Ypk1-3xFLAG::natNT2 J.S. Weissman, Univ. of California, San Francisco
yAM205-A BY4742 Lac1::LEU2 Lag1::LEU2 This study
yAM207-B BY4742 Lac1(S23A S24A)::LEU2 Lag1(S23A S24A)::LEU2 This study
yAM210 BY4742 Lac1(S23E S24E)::LEU2 Lag1(S23E S24E)::LEU2 This study
yGT12 BY4742 LYS2+ Lac1::LEU2 Lag1::LEU2 lcb3Δ::natNT2 This study
yGT13 BY4742 LYS2+ Lac1(S23A S24A)::LEU2 Lag1(S23A S24A)::LEU2 lcb3Δ::natNT2 This study
yGT14 BY4742 LYS2+ Lac1(S23E S24E)::LEU2 Lag1(S23E S24E)::LEU2 lcb3Δ::natNT2 This study
yAM168 BY4741 3xHA-Lac1::HIS3 3xFLAG-Lag1::LEU2 This study
yAM184 BY4741 3xHA-Lac1(S23A S24A)::HIS3 3xFLAG-Lag1(S23A S24A)::LEU2 This study
yAM192-A BY4741 MET15+ 3xHA-Lac1(S23E S24E)::HIS3 3xFLAG-Lag1(S23E S24E)::LEU2 This study
yKL4 BY4741 TOR2+::Hygr Kristin Leskoske, this lab
yKL5 BY4741 Tor2(L2178A)::Hygr Kristin Leskoske, this lab

Plasmids and recombinant DNA methods

All plasmids used in this study (except the library of PGAL1 based overexpression plasmids for synthetic dosage lethality [SDL] screening) are listed in Table 3. All plasmids were constructed and maintained in E. coli using standard laboratory methods (Green and Sambrook, 2012). For SDL screening, the entire open reading frame of each predicted and known Ypk1 substrate was amplified by PCR from BY4741 genomic DNA and ligated into the multiple cloning site of YCpLG (CEN, PGAL1, LEU2), generating a vector allowing galactose inducible overexpression of each substrate. All constructs generated in this study were confirmed by sequence analysis covering all promoter and coding regions in the construct.

Table 3.

Plasmids used in this study

DOI: http://dx.doi.org/10.7554/eLife.03779.013

Plasmid Description Source/reference
pGEX6P-1 GST tag, bacterial expression vector GE Healthcare, Inc.
pGEX4T-1 GST tag, bacterial expression vector GE Healthcare, Inc.
YCpLG CEN, LEU2, PGAL1 vector (Bardwell et al., 1998)
BG1805 2 µm, URA3, PGAL1, C-terminal tandem affinity (TAP) tag vector Open Biosystems, Inc.
pRS313 CEN, HIS3, vector (Sikorski and Hieter, 1989)
pRS316 CEN, URA3, vector (Sikorski and Hieter, 1989)
pRS416 CEN, URA3, vector (Sikorski and Hieter, 1989)
pBC111 CEN, LEU2, vector (Iida et al., 2007)
CHp282 pRS416 PMET25-GFP Chau Huynh, this laboratory
pLB215 pRS416 PMET25-Ypk1 (Niles et al., 2012)
pAX53 pRS416 PMET25-Ypk1(K376A) This study
pAX50 BG1805 Ypk1(L424A) This study
pFR203 pGEX4T-1 Orm1(1-85) (Roelants et al., 2011)
pBT6 pGEX6P-1 Fps1(1-255) This study
pBT7 pGEX6P-1 Fps1(531-669) This study
pBT12 pGEX6P-1 Smp1 This study
pAX55 pGEX6P-1 Lcb3(1-79) This study
pAX56 pGEX6P-1 Cdc1(1-41) This study
pAX58 pGEX6P-1 Her1(1-224) This study
pAX59 pGEX6P-1 Rts3 This study
pAX62 pGEX6P-1 Fkh1 This study
pAX63 pGEX6P-1 Yhp1 This study
pAX66 pGEX6P-1 YNR014W This study
pAX67 pGEX6P-1 YHR097C This study
pAX94 pGEX6P-1 Mds3(545-1016) This study
pAX131 pGEX4T-1 Lac1(1-76) This study
pAX132 pGEX4T-1 Lac1(1-76)(S23A S24A) This study
pFR291 pGEX4T-1 Lag1(1-80) This study
pAX133 pGEX4T-1 Lag1(1-80)(S23A S24A) This study
pAX134 pGEX6P-1 Muk1(1-305) This study
pAX215 pGEX6P-1 Cyk3 This study
pAX223 pGEX6P-1 Gpt2(1-35) This study
pAX224 pGEX6P-1 Gpt2(570-743) This study
pAX225 pGEX6P-1 Bre5 This study
pAX226 pGEX6P-1 Npr1(1-437) This study
pAX227 pGEX6P-1 Pal1 This study
pAX228 pGEX6P-1 Ysp2(97-665) This study
pAX229 pGEX6P-1 Ysp2(1072-1282) This study
pAX230 pGEX6P-1 Atg21 This study
pAX231 pGEX6P-1 Pex31(250-462) This study
pAX136 pRS313 PLAC1-3xHA-Lac1 This study
pFR273 pRS316 PYPK1-Ypk1(D242A) (Roelants et al., 2011)
pAX250 pRS313 PTPI1-GFP-Atg8 This study
pBCT-CCH1H pBC111 PTDH3-Cch1 (Iida et al., 2007)

Bioinformatic prediction of Ypk1 substrates

A Nx20 position weight matrix defining Ypk1 phosphoacceptor site specificity was made merging previously published data sets defining Ypk1 primary sequence specificity (Casamayor et al., 1999; Mok et al., 2010). This position weight matrix was then used with MOTIPS (Lam et al., 2010) to identify proteins with likely phosphorylated occurrences of this motif. Yeastmine (Balakrishnan et al., 2012) was used to identify all S. cerevisiae genes that showed genetic interactions with Ypk1, Ypk2, any component of the sphingolipid biosynthetic pathway or any component of known Ypk1 regulators (TORC2 and PP2A). Genes that showed significant growth phenotypes on myriocin, aureobasidin A and caspifungin (all compounds that cause severe growth phenotypes in ypk1Δ strains) were identified from the literature (Hillenmeyer et al., 2008). Lastly, MOTIPS predicted Ypk1 phosphorylation sites were compared to Phosphogrid (Sadowski et al., 2013) to identify those sites known to be phosphorylated in vivo. To be considered a potential Ypk1 substrate in this study, a protein had to have a myriocin, aureobasidin or caspofungin phenotype or a Yeastmine identified genetic interaction and: (a) 4 or more MOTIPS predicted sites, (b) 3 sites with at least one above a MOTIPS likelihood score of 0.7 or a Phosphogrid identified site or (c) 1–2 sites with a MOTIPS likelihood score of 0.7 and a Phosphogrid identified site. We also considered for further analysis a limited number of proteins that did not meet these criteria, but which still contained Ypk1 motifs. These are indicated in Table 3.

Yeast growth assays and synthetic dosage lethality screening

For SDL screening, yAM135-A and BY4741 strains were both transformed with each SDL plasmid. Transformants were then cultured overnight in SC media containing 2% raffinose and 0.2% sucrose. 10-fold serial dilutions of overnight cultures starting from OD600 = 1.0 were then made in sterile water and spotted onto SC solid media with 2% galactose (to induce protein expression) or 2% dextrose (no protein expression). These plates also contained 1:1000 DMSO, 1 μM 3-MB-PP1 or 2 μM 3-MB-PP1 to inhibit to varying degrees Ypk1-as kinase activity in yAM135-A. Serially spotted cultures were allowed to grow in the dark at 30°C for 3 days. Plates were then scanned on a flatbed scanner and growth phenotypes were assessed and scored.

For all SDL overexpression constructs that caused toxicity upon overexpression in WT BY4741 strains, we also performed Ypk1 dosage rescue growth assays. PMET25-Ypk1 or PMET25-Ypk1(K367A) (kinase dead) plasmids were co-transformed into BY4741 with the toxic SDL plasmid. Transformants were then cultured overnight in SC media containing 2% raffinose and 0.2% sucrose. 10-fold serial dilutions of overnight cultures starting from OD600 = 1.0 were then made in sterile water and spotted onto SC solid media with 2% galactose (to induce protein expression) or 2% dextrose (no protein expression), in the presence of absence of methionine to drive Ypk1 overexpression. Serially spotted cultures were allowed to grow in the dark at 30°C for 3 days. Plates were then scanned on a flatbed scanner and growth phenotypes were assessed to determine if Ypk1 overexpression could rescue toxicity of overexpression of the SDL protein.

For broth growth assays, exponential phase cultures growing in rich YP (Burke et al., 2005) media with 2% dextrose were diluted to OD600 = 0.1. 100 μl of each culture was placed in a well in a 96 well plate with vehicle or drug at the indicated concentration. Cultures were grown with orbital shaking at 30°C in a Tecan Infinite M-1000 PRO plate reader (Tecan Systems Inc., San Jose, CA) for 24 hr. Absorbance measurements were taken every 15 min. Absorbance values were converted to OD600 values using a standard curve of absorbance values of cultures at known OD600 taken on the same plate reader.

Protein purification and in vitro Ypk1 kinase assay

To purify Ypk1-as kinase, pAX50 (2 μ, PGAL1-Ypk1-as-TAP, URA3) transformed yAM135-A yeasts were diluted to OD600 = 0.125 in 3 l of SC 2% raffinose 0.2% sucrose and grown shaking at 30°C to mid-exponential phase. Expression was induced for ∼18 hr by the addition of 2% galactose. Cells were harvested by centrifugation and frozen in liquid nitrogen. The cells were then lysed cryogenically using Mixer Mill MM301 (Retsch, Düsseldorf, Germany). The lysate was resuspended at 2 ml/g in TAP-B (50 mM Tris-Cl pH 7.5, 200 mM NaCl, 1.5 mM MgOAc, 1 mM DTT, 2 mM NaVO4, 10 mM NaF, 10 mM Na-PPi, 10 mM β-glycerol phosphate, 1× complete protease inhibitor [Roche, Basel, Switzerland]). The lysate was clarified by centrifugation at 15×kg for 20 min. Clarified lysate was then further centrifuged at 100×kg for 1 hr and then brought to 0.15% NP-40 using 10% NP-40 detergent stock. Ypk1-as-TAP fusion was then affinity purified from the lysate using IgG-agarose resin (GE Healthcare). The resin was extensively washed with Protease 3C Buffer (50 mM Tris-Cl pH 7.5, 200 mM NaCl, 1.5 mM MgOAc, 1 mM DTT, 0.01% NP-40, 10% Glycerol, 2 mM NaVO4, 10 mM NaF, 10 mM Na-PPi, 10 mM β-glycerol phosphate) and then resuspended in 1 ml Protease 3C Buffer. Ypk1-as was eluted by the addition of 80 U Prescission Protease (GE Healthcare, Little Chalfont, UK) for 5 hr at 4°C. Protease 3C was removed by the addition of glutathione-agarose (GE Healthcare).

Putative Ypk1 substrates were expressed as N terminal GST fusions in BL21 E. coli. 1 l cultures were grown at 37°C to mid-exponential phase and then induced at room temperature for 4 hr with 0.5 mM IPTG. Cells were harvested by centrifugation and the fusion proteins were purified by affinity chromatography using glutathione-agarose and standard procedures.

For kinase assays 0.25 μg of Ypk1-as kinase was incubated with purified GST-substrate fusion in Kinase Assay Buffer (50 mM Tris-Cl pH 7.5, 200 mM NaCl, 10 mM MgCl2, 0.1 mM EDTA) with 2 μCi [γ-32P]ATP at 30°C in the presence of absence of 10 μM 3-MB-PP1 for 30 min. Reactions were terminated by the addition of SDS/PAGE sample buffer containing 6% SDS followed by boiling for 5 min. Labeled proteins were resolved by SDS/PAGE and analyzed Coomassie blue staining and autoradiography with Phosphorimager plates (Molecular Dynamics, Sunnyvale, CA) on a Typhoon imaging system (GE Healthcare).

Preparation of cell extracts and immunoblotting

Cell extracts were made by alkaline lysis followed by trichloroacetic acid precipitation as previously described (Westfall et al., 2008). To resolve Lag1 and Lac1 phosphorylated species, 15 μl of TCA extract was resolved by SDS-PAGE (8% acrylamide, 35 µM Phos-tag [Wako Chemicals USA, Inc., Richmond, VA], 35 µM MnCl2 at 160 V). The gel was then transferred to nitrocellulose and incubated with primary antibody in Odyssey buffer (Licor Biosciences Inc., Lincoln, NE), washed, and incubated with IRDye680-conjugated anti-mouse IgG (Licor Biosciences) in Odyssey buffer with 0.1% Tween-20 and 0.02% SDS. Blots were imaged using an Odyssey infrared scanner (Licor Biosciences).

Primary antibodies and dilutions used in this study were: 1:1000 mouse anti-HA (Covance Inc., Princeton, NJ), 1:10000 mouse anti-FLAG (Sigma–Aldrich, St. Louis, MO),1:500 mouse anti-GFP (Roche), 1:500 rabbit anti-pSGK (T256) (to recognize Ypk1 phosphorylated at T504) (Santa Cruz Biotechnology, Dallas, TX ) and 1:10000 rabbit anti-Pgk1 (our laboratory).

For phosphatase treatment of cell extracts, TCA extracts were made as above and the precipitated proteins were solubilized in 100 µl Solubilization buffer (50 mM Tris-Cl pH 8.0, 150 mM NaCl, 2% β-mercaptoethanol, 2% SDS). These extracts were then diluted with 900 µl CIP Dilution buffer (50 mM Tris-Cl pH 8.0, 150 mM NaCl, 11.1 mM MgCl2). 150 U calf intestinal phosphatase (New England Biolabs, Ipswich, MA) was then added and incubated for 2 hr at 37°C. Proteins were then TCA precipitated again and resolved by SDS-PAGE as above.

Analysis of sphingolipid species

Complex sphingolipids were analyzed by thin layer chromatography. Cultures of strains in mid-exponential phase were adjusted to OD600 = 1.0 and 2 ml cultures were labeled with 100 µCi of [32P] PO4–3 and cells allowed to grow for 3 hr. Lipids were extracted and resolved as previously described (Hanson and Lester, 1980; Momoi et al., 2004) with minor modifications. The cell pellet was washed twice with 2 ml water and treated with 5% trichloroacetic acid for 20 min on ice. Pellets were extracted twice with 0.75 ml of ethanol/water/diethyl-ether/pyridine/NH4OH (15:15:5:1:0.018) at 60°C for 1 hr. Glycerophospholipids in the extract were hydrolyzed by treating with 0.1 M monomethylamine at 50°C for 1 hr after which the base was neutralized by addition of 12 µl of glacial acetic acid. Lipids were extracted with 1 ml chloroform, 0.5 ml methanol and phases separated with addition of 1 ml water. For some samples 1 ml of 4 N NaCl was used, in order to facilitate the separation of the aqueous and organic phases. The organic layer was dried under vacuum and resuspended in 50 µl of chloroform/methanol/water (16:16:5) and resolved on a silica gel TLC plate with chloform/methanol/4.2 N NH4OH (9:7:2). Radioactivity on the TLC plate was visualized with a Phosphorimager screen and Typhoon imaging system.

LCBs and LCB-1Ps levels were monitored by liquid chromatography-mass spectrometry- Overnight cultures were adjusted to OD600 = 0.2 and allowed to grow to an OD600 = 1.0. Cell pellets from 10 ml of this culture equivalent to 10 OD600 units was used for analysis. C17-sphingosine (Avanti Polar Lipids, Alabaster, AL), (5 nmol) was added to all samples as an internal standard. Lipids were extracted as described above and the final dried lipid extract was dissolved in methanol/water/formic acid (79:20:1), centrifuged to remove insoluble material and an aliquot of this material was injected into the HPLC for analysis.

Lipid extracts were analyzed using an Agilent 1200 liquid chromatograph (LC; Santa Clara, CA) that was connected in-line with an LTQ Orbitrap XL mass spectrometer equipped with an electrospray ionization source (ESI; Thermo Fisher Scientific, Waltham, MA).

The LC was equipped with a C4 analytical column (Viva C4, 150 mm length × 1.0 mm inner diameter, 5 µm particles, 300 Å pores, Restek, Bellefonte, PA) and a 100 µl sample loop. Solvent A was 99.8% water/0.2% formic acid and solvent B was 99.8% methanol/0.2% formic acid (vol/vol). Solvents A and B both contained 5 mM ammonium formate. The sample injection volume was 85 µl (partial loop). The elution program consisted of isocratic flow at 30% B for 2 min, a linear gradient to 65% B over 0.1 min, a linear gradient to 100% B over 4.9 min, isocratic flow at 100% B for 4 min, a linear gradient to 30% B over 0.1 min, and isocratic flow at 30% B for 18.9 min, at a flow rate of 170 µl/min. The column and sample compartments were maintained at 40°C and 4°C, respectively. The injection needle was rinsed with a 1:1 methanol/water (vol/vol) solution after each injection.

The column exit was connected to the ESI probe of the mass spectrometer using PEEK tubing (0.005ʺ inner diameter × 1/16ʺ outer diameter, Agilent). Mass spectra were acquired in the positive ion mode over the range m/z = 250 to 1200 using the Orbitrap mass analyzer, in profile format, with a mass resolution setting of 100,000 (at m/z = 400, measured at full width at half-maximum peak height). In the data-dependent mode, the five most intense ions exceeding an intensity threshold of 30,000 counts were selected from each full-scan mass spectrum for tandem mass spectrometry (MS/MS) analysis using pulsed-Q dissociation (PQD). MS/MS spectra were acquired in the positive ion mode using the linear ion trap, in centroid format, with the following parameters: isolation width 3 m/z units, normalized collision energy 40%, and default charge state 1+. A parent mass list was used to preferentially select ions of interest for targeted MS/MS analysis. To avoid the occurrence of redundant MS/MS measurements, real-time dynamic exclusion was enabled to preclude re-selection of previously analyzed precursor ions, using the following parameters: repeat count 2, repeat duration 30 s, exclusion list size 500, exclusion duration 180 s, and exclusion width 0.1 m/z unit. Data were analyzed using Xcalibur software (version 2.0.7 SP1, Thermo) and LIPID MAPS Online Tools (Fahy et al., 2007).

Exact masses of precursor ions in positive ion (M + H+) state obtained from LIPID MAPS Online Tools were as follows- phytosphingosine- 318.3003, dihydrosphingosine-302.3053, phytosphingosine-1 phosphate-398.2666 and dihydrosphingosine-1 phosphate- 382.2717. The ions were further confirmed by tandem MS and identification of the fragments; product ions for the phytosphingosine headgroup was 282.3 and dihydrosphigosine headgroup was 266.4.

In vitro ceramide synthase assay

Yeast expressing 3xFLAG-Lag1 (wild-type and phospho-site mutations) were grown at 30°C to mid-exponential phase and microsomes were prepared as described previously (Schorling et al., 2001) and resuspended in B88 buffer (20 mM HEPES-KOH pH 6.8, 150 mM KAc, 5 mM MgOAc, and 250 mM sorbitol). Ceramide synthase was then immunopurified from these microsomes using anti-FLAG agarose (Sigma) as described previously (Vallée and Riezman, 2005). Prior to assembling ceramide synthase reactions, a small alioquot of each (15 μl) immunoprecipitate was taken and proteins were resolved by SDS-PAGE and immunoblotted with anti-FLAG to determine the relative amount of immunopurified ceramide synthase in each reaction. Recovered ceramide levels for each reaction were normalized to the amount of ceramide synthase determined by this procedure.

Ceramide synthesis reaction was assembled in B88 buffer with 5 mg/ml BSA, in a total reaction volume of 0.1 ml containing, 47.5 µl anti-FLAG agarose immunoprecipitates, 50 µM phytosphingosine and 0.1 mM of steroyl (C18)- CoA. Reactions were incubated at 30°C for 1 hr. Reactions were terminated by addition of 0.4 ml methanol/CHCl3 (2:1) and further extracted with 0.4 ml CHCl3 and 0.4 ml water. The CHCl3 phase was separated by centrifugation and removed and dried under vacuum. The dried lipids resuspended in methanol/water/formic acid (79/20/1) and an aliquot analyzed by LC-MS as described above. The product from the reaction, C18-phytoceramide was identified by the exact mass of its precursor ion at 584.5612 and its product ion upon fragmentation at 282.3. Additionally, authentic C18-phytoceramide (Matreya, LLC, Pleasant Gap, PA) standard was used to obtain a standard curve in order to establish that values from ceramide synthase reaction were in the linear range of estimation.

Acknowledgements

This work was supported by an NIH Predoctoral Traineeship GM07232 and a Predoctoral Fellowship from the UC Systemwide Cancer Research Coordinating Committee (to AM) and by NIH R01 Research Grant GM21841 and Senior Investigator Award 11-0118 from the American Asthma Foundation (to JT). We thank Tony Iavarone in the QB3/Chemistry Mass Spectrometry Facility for expert assistance with the analysis of sphingolipids by mass spectrometry, Howard Riezman for the gift of the FLAG-Lip1 construct, Jonathan Weissman, Martha Cyert, Benjamin Turk, Daniel Klionsky, and Hidetoshi Iida for other strains or plasmids, Jeffrey Liu and Brian Ting for technical assistance, and other members of the Thorner Lab for various reagents, plasmids, and many helpful discussions.

Funding Statement

The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.

Funding Information

This paper was supported by the following grants:

  • National Institute of General Medical Sciences FundRef identification ID: http://dx.doi.org/10.13039/100000057 GM21841 to Jeremy Thorner.

  • American Asthma Foundation Senior Investigator Award 11-0118 to Jeremy Thorner.

  • University of California FundRef identification ID: http://dx.doi.org/10.13039/100005595 Cancer Research Coordinating Committee (CRCC) to Alexander Muir.

Additional information

Competing interests

The authors declare that no competing interests exist.

Author contributions

AM, Conception and design, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

SR, Conception and design, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

FMR, Conception and design, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

GT, Contributed in a major way to screening all of the bioinformatic hits using the synthetic dosage lethal method devised in this paper, Acquisition of data.

JT, Conception and design, Analysis and interpretation of data, Drafting or revising the article.

References

  1. Alers S, Wesselborg S, Stork B. 2014. ATG13: just a companion, or an executor of the autophagic program? Autophagy 10:944–956 [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Alvaro CG, O'Donnell AF, Prosser DC, Augustine AA, Goldman A, Brodsky JL, Cyert MS, Wendland B, Thorner J. 2014. Specific alpha-arrestins negatively regulate Saccharomyces cerevisiae pheromone response by down-modulating the G-protein coupled receptor Ste2. Molecular and Cellular Biology 34:2660–2681. doi: 10.1128/MCB.00230-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Ansell R, Granath K, Hohmann S, Thevelein JM, Adler L. 1997. The two isoenzymes for yeast NAD+-dependent glycerol 3-phosphate dehydrogenase encoded by GPD1 and GPD2 have distinct roles in osmoadaptation and redox regulation. The EMBO Journal 16:2179–2187. doi: 10.1093/emboj/16.9.2179 [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Aronova S, Wedaman K, Aronov PA, Fontes K, Ramos K, Hammock BD, Powers T. 2008. Regulation of ceramide biosynthesis by TOR complex 2. Cell Metabolism 7:148–158. doi: 10.1016/j.cmet.2007.11.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Balakrishnan R, Park J, Karra K, Hitz BC, Binkley G, Hong EL, Sullivan J, Micklem G, Cherry JM. 2012. YeastMine–an integrated data warehouse for Saccharomyces cerevisiae data as a multipurpose tool-kit. Database 2012:bar062. doi: 10.1093/database/bar062 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bardwell L, Cook JG, Zhu-Shimoni JX, Voora D, Thorner J. 1998. Differential regulation of transcription: repression by unactivated mitogen-activated protein kinase Kss1 requires the Dig1 and Dig2 proteins. Proceedings of the National Academy of Sciences of USA 95:15400–15405. doi: 10.1073/pnas.95.26.15400 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Barz WP, Walter P. 1999. Two endoplasmic reticulum (ER) membrane proteins that facilitate ER-to-Golgi transport of glycosylphosphatidylinositol-anchored proteins. Molecular Biology of the Cell 10:1043–1059. doi: 10.1091/mbc.10.4.1043 [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Becuwe M, Vieira N, Lara D, Gomes-Rezende J, Soares-Cunha C, Casal M, Haguenauer-Tsapis R, Vincent O, Paiva S, Léon S. 2012. A molecular switch on an arrestin-like protein relays glucose signaling to transporter endocytosis. The Journal of Cell Biology 196:247–259. doi: 10.1083/jcb.201109113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Beeler T, Bacikova D, Gable K, Hopkins L, Johnson C, Slife H, Dunn T. 1998. The Saccharomyces cerevisiae TSC10/YBR265w gene encoding 3-ketosphinganine reductase is identified in a screen for temperature-sensitive suppressors of the Ca2+-sensitive csg2Δ mutant. The Journal of Biological Chemistry 273:30688–30694. doi: 10.1074/jbc.273.46.30688 [DOI] [PubMed] [Google Scholar]
  10. Berchtold D, Piccolis M, Chiaruttini N, Riezman I, Riezman H, Roux A, Walther TC, Loewith R. 2012. Plasma membrane stress induces relocalization of Slm proteins and activation of TORC2 to promote sphingolipid synthesis. Nature Cell Biology 14:542–547. doi: 10.1038/ncb2480 [DOI] [PubMed] [Google Scholar]
  11. Bikman BT, Summers SA. 2011. Ceramides as modulators of cellular and whole-body metabolism. The Journal of Clinical Investigation 121:4222–4230. doi: 10.1172/JCI57144 [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Breslow DK. 2013. Sphingolipid homeostasis in the endoplasmic reticulum and beyond. Cold Spring Harbor Perspectives in Biology 5:a013326. doi: 10.1101/cshperspect.a013326 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Breslow DK, Weissman JS. 2010. Membranes in balance: mechanisms of sphingolipid homeostasis. Molecular Cell 40:267–279. doi: 10.1016/j.molcel.2010.10.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Bultynck G, Heath VL, Majeed AP, Galan JM, Haguenauer-Tsapis R, Cyert MS. 2006. Slm1 and slm2 are novel substrates of the calcineurin phosphatase required for heat stress-induced endocytosis of the yeast uracil permease. Molecular and Cellular Biology 26:4729–4745. doi: 10.1128/MCB.01973-05 [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Burkard ME, Randall CL, Larochelle S, Zhang C, Shokat KM, Fisher RP, Jallepalli PV. 2007. Chemical genetics reveals the requirement for Polo-like kinase 1 activity in positioning RhoA and triggering cytokinesis in human cells. Proceedings of the National Academy of Sciences of USA 104:4383–4388. doi: 10.1073/pnas.0701140104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Burke D, Amberg D, Strathern J. 2005. Methods in yeast genetics: a cold spring harbor laboratory course manual. Cold Spring Harbor Laboratory Press [Google Scholar]
  17. Byrne KP, Wolfe KH. 2005. The yeast gene order browser: combining curated homology and syntenic context reveals gene fate in polyploid species. Genome Research 15:1456–1461. doi: 10.1101/gr.3672305 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Casamayor A, Torrance PD, Kobayashi T, Thorner J, Alessi DR. 1999. Functional counterparts of mammalian protein kinases PDK1 and SGK in budding yeast. Current Biology 9:186–197. doi: 10.1016/S0960-9822(99)80088-8 [DOI] [PubMed] [Google Scholar]
  19. Chen PW, Fonseca LL, Hannun YA, Voit EO. 2013. Coordination of rapid sphingolipid responses to heat stress in yeast. PLOS Computational Biology 9:e1003078. doi: 10.1371/journal.pcbi.1003078 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Colabardini AC, Brown NA, Savoldi M, Goldman MH, Goldman GH. 2013. Functional characterization of Aspergillus nidulans ypkA, a homologue of the mammalian kinase SGK. PLOS ONE 8:e57630. doi: 10.1371/journal.pone.0057630 [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Cowart LA, Okamoto Y, Pinto FR, Gandy JL, Almeida JS, Hannun YA. 2003. Roles for sphingolipid biosynthesis in mediation of specific programs of the heat stress response determined through gene expression profiling. The Journal of Biological Chemistry 278:30328–30338. doi: 10.1074/jbc.M300656200 [DOI] [PubMed] [Google Scholar]
  22. Daquinag A, Fadri M, Jung SY, Qin J, Kunz J. 2007. The yeast PH domain proteins Slm1 and Slm2 are targets of sphingolipid signaling during the response to heat stress. Molecular and Cellular Biology 27:633–650. doi: 10.1128/MCB.00461-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. de Nadal E, Casadomé L, Posas F. 2003. Targeting the MEF2-like transcription factor Smp1 by the stress-activated Hog1 mitogen-activated protein kinase. Molecular and Cellular Biology 23:229–237. doi: 10.1128/MCB.23.1.229-237.2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. deHart AK, Schnell JD, Allen DA, Hicke L. 2002. The conserved Pkh-Ypk kinase cascade is required for endocytosis in yeast. The Journal of Cell Biology 156:241–248. doi: 10.1083/jcb.200107135 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Dickson RC. 2008. Thematic review series: sphingolipids. New insights into sphingolipid metabolism and function in budding yeast. Journal of Lipid Research 49:909–921. doi: 10.1194/jlr.R800003-JLR200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Divito CB, Amara SG. 2009. Close encounters of the oily kind: regulation of transporters by lipids. Molecular Interventions 9:252–262. doi: 10.1124/mi.9.5.8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Fahy E, Sud M, Cotter D, Subramaniam S. 2007. LIPID MAPS online tools for lipid research. Nucleic Acids Research 35:W606–W612. doi: 10.1093/nar/gkm324 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Green MR, Sambrook J. 2012. Molecular cloning: a laboratory manual. Cold Spring Harbor, N.Y: Cold Spring Harbor Laboratory Press [Google Scholar]
  29. Groves JT, Kuriyan J. 2010. Molecular mechanisms in signal transduction at the membrane. Nature Structural & Molecular Biology 17:659–665. doi: 10.1038/nsmb.1844 [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Gwinn DM, Shackelford DB, Egan DF, Mihaylova MM, Mery A, Vasquez DS, Turk BE, Shaw RJ. 2008. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Molecular Cell 30:214–226. doi: 10.1016/j.molcel.2008.03.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Hanson BA, Lester RL. 1980. The extraction of inositol-containing phospholipids and phosphatidylcholine from Saccharomyces cerevisiae and Neurospora crassa. Journal of Lipid Research 21:309–315 [PubMed] [Google Scholar]
  32. Heidler SA, Radding JA. 1995. The AUR1 gene in Saccharomyces cerevisiae encodes dominant resistance to the antifungal agent aureobasidin A (LY295337). Antimicrobial Agents and Chemotherapy 39:2765–2769. doi: 10.1128/AAC.39.12.2765 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Hillenmeyer ME, Fung E, Wildenhain J, Pierce SE, Hoon S, Lee W, Proctor M, St Onge RP, Tyers M, Koller D, Altman RB, Davis RW, Nislow C, Giaever G. 2008. The chemical genomic portrait of yeast: uncovering a phenotype for all genes. Science 320:362–365. doi: 10.1126/science.1150021 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Holt LJ, Hutti JE, Cantley LC, Morgan DO. 2007. Evolution of Ime2 phosphorylation sites on Cdk1 substrates provides a mechanism to limit the effects of the phosphatase Cdc14 in meiosis. Molecular Cell 25:689–702. doi: 10.1016/j.molcel.2007.02.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Hutti JE, Shen RR, Abbott DW, Zhou AY, Sprott KM, Asara JM, Hahn WC, Cantley LC. 2009. Phosphorylation of the tumor suppressor CYLD by the breast cancer oncogene IKKepsilon promotes cell transformation. Molecular Cell 34:461–472. doi: 10.1016/j.molcel.2009.04.031 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Iida K, Teng J, Tada T, Saka A, Tamai M, Izumi-Nakaseko H, Adachi-Akahane S, Iida H. 2007. Essential, completely conserved glycine residue in the domain III S2-S3 linker of voltage-gated calcium channel alpha1 subunits in yeast and mammals. The Journal of Biological Chemistry 282:25659–25667. doi: 10.1074/jbc.M703757200 [DOI] [PubMed] [Google Scholar]
  37. Ikeda K, Morigasaki S, Tatebe H, Tamanoi F, Shiozaki K. 2008. Fission yeast TOR complex 2 activates the AGC-family Gad8 kinase essential for stress resistance and cell cycle control. Cell Cycle 7:358–364. doi: 10.4161/cc.7.3.5245 [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Jacinto E, Facchinetti V, Liu D, Soto N, Wei S, Jung SY, Huang Q, Qin J, Su B. 2006. SIN1/MIP1 maintains rictor-mTOR complex integrity and regulates Akt phosphorylation and substrate specificity. Cell 127:125–137. doi: 10.1016/j.cell.2006.08.033 [DOI] [PubMed] [Google Scholar]
  39. Jenkins GM, Richards A, Wahl T, Mao C, Obeid L, Hannun Y. 1997. Involvement of yeast sphingolipids in the heat stress response of Saccharomyces cerevisiae. The Journal of Biological Chemistry 272:32566–32572. doi: 10.1074/jbc.272.51.32566 [DOI] [PubMed] [Google Scholar]
  40. Jiang JC, Kirchman PA, Zagulski M, Hunt J, Jazwinski SM. 1998. Homologs of the yeast longevity gene LAG1 in Caenorhabditis elegans and human. Genome Research 8:1259–1272. doi: 10.1101/gr.8.3.268 [DOI] [PubMed] [Google Scholar]
  41. Kageyama-Yahara N, Riezman H. 2006. Transmembrane topology of ceramide synthase in yeast. The Biochemical Journal 398:585–593. doi: 10.1042/BJ20060697 [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Kamada Y, Fujioka Y, Suzuki NN, Inagaki F, Wullschleger S, Loewith R, Hall MN, Ohsumi Y. 2005. Tor2 directly phosphorylates the AGC kinase Ypk2 to regulate actin polarization. Molecular and Cellular Biology 25:7239–7248. doi: 10.1128/MCB.25.16.7239-7248.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Kataoka S, Furuita K, Hattori Y, Kobayashi N, Ikegami T, Shiozaki K, Fujiwara T, Kojima C. 2014. 1H, 15N and 13C resonance assignments of the conserved region in the middle domain of S. pombe Sin1 protein. Biomolecular NMR Assignments [Epub ahead of print, 8 March 2014]. [DOI] [PubMed] [Google Scholar]
  44. Kihara A, Mitsutake S, Mizutani Y, Igarashi Y. 2007. Metabolism and biological functions of two phosphorylated sphingolipids, sphingosine 1-phosphate and ceramide 1-phosphate. Progress in Lipid Research 46:126–144. doi: 10.1016/j.plipres.2007.03.001 [DOI] [PubMed] [Google Scholar]
  45. Kim S, Fyrst H, Saba J. 2000. Accumulation of phosphorylated sphingoid long chain bases results in cell growth inhibition in Saccharomyces cerevisiae. Genetics 156:1519–1529 [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Kinoshita E, Kinoshita-Kikuta E, Koike T. 2009. Separation and detection of large phospho-proteins using Phos-tag SDS-PAGE. Nature Protocols 4:1513–1521. doi: 10.1038/nprot.2009.154 [DOI] [PubMed] [Google Scholar]
  47. Kliegman JI, Fiedler D, Ryan CJ, Xu YF, Su XY, Thomas D, Caccese MC, Cheng A, Shales M, Rabinowitz JD, Krogan NJ, Shokat KM. 2013. Chemical genetics of rapamycin-insensitive TORC2 in S. cerevisiae. Cell Reports 5:1725–1736. doi: 10.1016/j.celrep.2013.11.040 [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Kobayashi SD, Nagiec MM. 2003. Ceramide/long-chain base phosphate rheostat in Saccharomyces cerevisiae: regulation of ceramide synthesis by Elo3p and Cka2p. Eukaryotic Cell 2:284–294. doi: 10.1128/EC.2.2.284-294.2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Lam HY, Kim PM, Mok J, Tonikian R, Sidhu SS, Turk BE, Snyder M, Gerstein MB. 2010. MOTIPS: automated motif analysis for predicting targets of modular protein domains. BMC Bioinformatics 11:243. doi: 10.1186/1471-2105-11-243 [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Laviad EL, Kelly S, Merrill AH, Futerman AH. 2012. Modulation of ceramide synthase activity via dimerization. The Journal of Biological Chemistry 287:21025–21033. doi: 10.1074/jbc.M112.363580 [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Lee YJ, Huang X, Kropat J, Henras A, Merchant SS, Dickson RC, Chanfreau GF. 2012a. Sphingolipid signaling mediates iron toxicity. Cell Metabolism 16:90–96. doi: 10.1016/j.cmet.2012.06.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Lee YJ, Jeschke GR, Roelants FM, Thorner J, Turk BE. 2012b. Reciprocal phosphorylation of yeast glycerol-3-phosphate dehydrogenases in adaptation to distinct types of stress. Molecular and Cellular Biology 32:4705–4717. doi: 10.1128/MCB.00897-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Liao HC, Chen MY. 2012. Target of rapamycin complex 2 signals to downstream effector yeast protein kinase 2 (Ypk2) through adheres-voraciously-to-target-of-rapamycin-2 protein 1 (Avo1) in Saccharomyces cerevisiae. The Journal of Biological Chemistry 287:6089–6099. doi: 10.1074/jbc.M111.303701 [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Lin CH, MacGurn JA, Chu T, Stefan CJ, Emr SD. 2008. Arrestin-related ubiquitin-ligase adaptors regulate endocytosis and protein turnover at the cell surface. Cell 135:714–725. doi: 10.1016/j.cell.2008.09.025 [DOI] [PubMed] [Google Scholar]
  55. Lin D, Yin X, Wang X, Zhou P, Guo FB. 2013. Re-annotation of protein-coding genes in the genome of Saccharomyces cerevisiae based on support vector machines. PLOS ONE 8:e64477. doi: 10.1371/journal.pone.0064477 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Linding R, Jensen LJ, Ostheimer GJ, van Vugt MA, Jørgensen C, Miron IM, Diella F, Colwill K, Taylor L, Elder K, Metalnikov P, Nguyen V, Pasculescu A, Jin J, Park JG, Samson LD, Woodgett JR, Russell RB, Bork P, Yaffe MB, Pawson T. 2007. Systematic discovery of in vivo phosphorylation networks. Cell 129:1415–1426. doi: 10.1016/j.cell.2007.05.052 [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Liu M, Huang C, Polu SR, Schneiter R, Chang A. 2012. Regulation of sphingolipid synthesis through Orm1 and Orm2 in yeast. Journal of Cell Science 125:2428–2435. doi: 10.1242/jcs.100578 [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Liu P, Gan W, Inuzuka H, Lazorchak AS, Gao D, Arojo O, Liu D, Wan L, Zhai B, Yu Y, Yuan M, Kim BM, Shaik S, Menon S, Gygi SP, Lee TH, Asara JM, Manning BD, Blenis J, Su B, Wei W. 2013. Sin1 phosphorylation impairs mTORC2 complex integrity and inhibits downstream Akt signalling to suppress tumorigenesis. Nature Cell Biology 15:1340–1350. doi: 10.1038/ncb2860 [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Loewith R, Hall MN. 2011. Target of rapamycin (TOR) in nutrient signaling and growth control. Genetics 189:1177–1201. doi: 10.1534/genetics.111.133363 [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Lu M, Wang J, Ives HE, Pearce D. 2011. mSIN1 protein mediates SGK1 protein interaction with mTORC2 protein complex and is required for selective activation of the epithelial sodium channel. The Journal of Biological Chemistry 286:30647–30654. doi: 10.1074/jbc.M111.257592 [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Luyten K, Albertyn J, Skibbe WF, Prior BA, Ramos J, Thevelein JM, Hohmann S. 1995. Fps1, a yeast member of the MIP family of channel proteins, is a facilitator for glycerol uptake and efflux and is inactive under osmotic stress. The EMBO Journal 14:1360–1371 [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. MacGurn JA, Hsu PC, Smolka MB, Emr SD. 2011. TORC1 regulates endocytosis via Npr1-mediated phosphoinhibition of a ubiquitin ligase adaptor. Cell 147:1104–1117. doi: 10.1016/j.cell.2011.09.054 [DOI] [PubMed] [Google Scholar]
  63. Mah AS, Elia AE, Devgan G, Ptacek J, Schutkowski M, Snyder M, Yaffe MB, Deshaies RJ. 2005. Substrate specificity analysis of protein kinase complex Dbf2-Mob1 by peptide library and proteome array screening. BMC Biochemistry 6:22. doi: 10.1186/1471-2091-6-22 [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Mandala SM, Harris GH. 2000. Isolation and characterization of novel inhibitors of sphingolipid synthesis: australifungin, viridofungins, rustimicin and khafrefungin. Methods in Enzymology 311:335–348. doi: 10.1016/S0076-6879(00)11094-8 [DOI] [PubMed] [Google Scholar]
  65. Mandala SM, Thornton R, Tu Z, Kurtz MB, Nickels J, Broach J, Menzeleev R, Spiegel S. 1998. Sphingoid base 1-phosphate phosphatase: a key regulator of sphingolipid metabolism and stress response. Proceedings of the National Academy of Sciences of USA 95:150–155. doi: 10.1073/pnas.95.1.150 [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Manning BD, Tee AR, Logsdon MN, Blenis J, Cantley LC. 2002. Identification of the tuberous sclerosis complex-2 tumor suppressor gene product tuberin as a target of the phosphoinositide 3-kinase/akt pathway. Molecular Cell 10:151–162. doi: 10.1016/S1097-2765(02)00568-3 [DOI] [PubMed] [Google Scholar]
  67. Mok J, Kim PM, Lam HY, Piccirillo S, Zhou X, Jeschke GR, Sheridan DL, Parker SA, Desai V, Jwa M, Cameroni E, Niu H, Good M, Remenyi A, Ma JL, Sheu YJ, Sassi HE, Sopko R, Chan CS, De Virgilio C, Hollingsworth NM, Lim WA, Stern DF, Stillman B, Andrews BJ, Gerstein MB, Snyder M, Turk BE. 2010. Deciphering protein kinase specificity through large-scale analysis of yeast phosphorylation site motifs. Science Signaling 3:ra12. doi: 10.1126/scisignal.2000482 [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Momoi M, Tanoue D, Sun Y, Takematsu H, Suzuki Y, Suzuki M, Suzuki A, Fujita T, Kozutsumi Y. 2004. SLI1 (YGR212W) is a major gene conferring resistance to the sphingolipid biosynthesis inhibitor ISP-1, and encodes an ISP-1 N-acetyltransferase in yeast. The Biochemical Journal 381:321–328. doi: 10.1042/BJ20040108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Mora A, Komander D, van Aalten DM, Alessi DR. 2004. PDK1, the master regulator of AGC kinase signal transduction. Seminars in Cell & Developmental Biology 15:161–170. doi: 10.1016/j.semcdb.2003.12.022 [DOI] [PubMed] [Google Scholar]
  70. Mulet JM, Martin DE, Loewith R, Hall MN. 2006. Mutual antagonism of target of rapamycin and calcineurin signaling. The Journal of Biological Chemistry 281:33000–33007. doi: 10.1074/jbc.M604244200 [DOI] [PubMed] [Google Scholar]
  71. Nagiec MM, Nagiec EE, Baltisberger JA, Wells G, Lester RL, Dickson RC. 1997. Sphingolipid synthesis as a target for antifungal drugs. Complementation of the inositol phosphorylceramide synthase defect in a mutant strain of Saccharomyces cerevisiae by the AUR1 gene. The Journal of Biological Chemistry 272:9809–9817. doi: 10.1074/jbc.272.15.9809 [DOI] [PubMed] [Google Scholar]
  72. Nagiec MM, Skrzypek M, Nagiec EE, Lester RL, Dickson RC. 1998. The LCB4 (YOR171c) and LCB5 (YLR260w) genes of Saccharomyces encode sphingoid long chain base kinases. The Journal of Biological Chemistry 273:19437–19442. doi: 10.1074/jbc.273.31.19437 [DOI] [PubMed] [Google Scholar]
  73. Niles BJ, Joslin AC, Fresques T, Powers T. 2014. TOR complex 2-Ypk1 signaling maintains sphingolipid homeostasis by sensing and regulating ROS accumulation. Cell Reports 6:541–552. doi: 10.1016/j.celrep.2013.12.040 [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Niles BJ, Mogri H, Hill A, Vlahakis A, Powers T. 2012. Plasma membrane recruitment and activation of the AGC kinase Ypk1 is mediated by target of rapamycin complex 2 (TORC2) and its effector proteins Slm1 and Slm2. Proceedings of the National Academy of Sciences of USA 109:1536–1541. doi: 10.1073/pnas.1117563109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Niles BJ, Powers T. 2012. Plasma membrane proteins Slm1 and Slm2 mediate activation of the AGC kinase Ypk1 by TORC2 and sphingolipids in S. cerevisiae. Cell Cycle 11:3745–3749. doi: 10.4161/cc.21752 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. O'Donnell AF, Huang L, Thorner J, Cyert MS. 2013. A calcineurin-dependent switch controls the trafficking function of alpha-arrestin Aly1/Art6. The Journal of Biological Chemistry 288:24063–24080. doi: 10.1074/jbc.M113.478511 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Papinski D, Kraft C. 2014. Atg1 kinase organizes autophagosome formation by phosphorylating Atg9. Autophagy 10:1338–1340. doi: 10.4161/auto.28971. [Epub ahead of print, 29 April 2014]. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Paulsel AL, Merz AJ, Nickerson DP. 2013. Vps9 family protein Muk1 is the second Rab5 guanosine nucleotide exchange factor in budding yeast. The Journal of Biological Chemistry 288:18162–18171. doi: 10.1074/jbc.M113.457069 [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Pearce L, Komander D, Alessi DR. 2010. The nuts and bolts of AGC protein kinases. Nature Reviews Molecular Cell Biology 11:9–22. doi: 10.1038/nrm2822 [DOI] [PubMed] [Google Scholar]
  80. Platta HW, Stenmark H. 2011. Endocytosis and signaling. Current Opinion in Cell Biology 23:393–403. doi: 10.1016/j.ceb.2011.03.008 [DOI] [PubMed] [Google Scholar]
  81. Rennefahrt UE, Deacon SW, Parker SA, Devarajan K, Beeser A, Chernoff J, Knapp S, Turk BE, Peterson JR. 2007. Specificity profiling of Pak kinases allows identification of novel phosphorylation sites. The Journal of Biological Chemistry 282:15667–15678. doi: 10.1074/jbc.M700253200 [DOI] [PubMed] [Google Scholar]
  82. Roelants FM, Baltz AG, Trott AE, Fereres S, Thorner J. 2010. A protein kinase network regulates the function of aminophospholipid flippases. Proceedings of the National Academy of Sciences of USA 107:34–39. doi: 10.1073/pnas.0912497106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Roelants FM, Breslow DK, Muir A, Weissman JS, Thorner J. 2011. Protein kinase Ypk1 phosphorylates regulatory proteins Orm1 and Orm2 to control sphingolipid homeostasis in Saccharomyces cerevisiae. Proceedings of the National Academy of Sciences of USA 108:19222–19227. doi: 10.1073/pnas.1116948108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Roelants FM, Torrance PD, Bezman N, Thorner J. 2002. Pkh1 and Pkh2 differentially phosphorylate and activate Ypk1 and Ykr2 and define protein kinase modules required for maintenance of cell wall integrity. Molecular Biology of the Cell 13:3005–3028. doi: 10.1091/mbc.E02-04-0201 [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Roelants FM, Torrance PD, Thorner J. 2004. Differential roles of PDK1- and PDK2-phosphorylation sites in the yeast AGC kinases Ypk1, Pkc1 and Sch9. Nature Reviews Microbiology 150:3289–3304. doi: 10.1099/mic.0.27286-0 [DOI] [PubMed] [Google Scholar]
  86. Roy J, Cyert MS. 2009. Cracking the phosphatase code: docking interactions determine substrate specificity. Science Signaling 2:re9. doi: 10.1126/scisignal.2100re9 [DOI] [PubMed] [Google Scholar]
  87. Roy J, Li H, Hogan PG, Cyert MS. 2007. A conserved docking site modulates substrate affinity for calcineurin, signaling output, and in vivo function. Molecular Cell 25:889–901. doi: 10.1016/j.molcel.2007.02.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Sachs JN, Engelman DM. 2006. Introduction to the membrane protein reviews: the interplay of structure, dynamics, and environment in membrane protein function. Annual Review of Biochemistry 75:707–712. doi: 10.1146/annurev.biochem.75.110105.142336 [DOI] [PubMed] [Google Scholar]
  89. Sadowski I, Breitkreutz BJ, Stark C, Su TC, Dahabieh M, Raithatha S, Bernhard W, Oughtred R, Dolinski K, Barreto K, Tyers M. 2013. The PhosphoGRID Saccharomyces cerevisiae protein phosphorylation site database: version 2.0 update. Database 2013:bat026. doi: 10.1093/database/bat026 [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Schorling S, Vallée B, Barz WP, Riezman H, Oesterhelt D. 2001. Lag1p and Lac1p are essential for the acyl-CoA-dependent ceramide synthase reaction in Saccharomyces cerevisiae. Molecular Biology of the Cell 12:3417–3427. doi: 10.1091/mbc.12.11.3417 [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Sharifpoor S, van Dyk D, Costanzo M, Baryshnikova A, Friesen H, Douglas AC, Youn JY, VanderSluis B, Myers CL, Papp B, Boone C, Andrews BJ. 2012. Functional wiring of the yeast kinome revealed by global analysis of genetic network motifs. Genome Research 22:791–801. doi: 10.1101/gr.129213.111 [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Shimada K, Filipuzzi I, Stahl M, Helliwell SB, Studer C, Hoepfner D, Seeber A, Loewith R, Movva NR, Gasser SM. 2013. TORC2 signaling pathway guarantees genome stability in the face of DNA strand breaks. Molecular Cell 51:829–839. doi: 10.1016/j.molcel.2013.08.019 [DOI] [PubMed] [Google Scholar]
  93. Shimobayashi M, Oppliger W, Moes S, Jenö P, Hall MN. 2013. TORC1-regulated protein kinase Npr1 phosphorylates Orm to stimulate complex sphingolipid synthesis. Molecular Biology of the Cell 24:870–881. doi: 10.1091/mbc.E12-10-0753 [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Sikorski RS, Hieter P. 1989. A system of shuttle vectors and yeast host strains designed for efficient manipulation of DNA in Saccharomyces cerevisiae. Genetics 122:19–27 [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Simons K, Sampaio JL. 2011. Membrane organization and lipid rafts. Cold Spring Harbor Perspectives in Biology 3:a004697–1.. doi: 10.1101/cshperspect.a004697 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Sokolov S, Knorre D, Smirnova E, Markova O, Pozniakovsky A, Skulachev V, Severin F. 2006. Ysp2 mediates death of yeast induced by amiodarone or intracellular acidification. Biochimica et Biophysica Acta 1757:1366–1370. doi: 10.1016/j.bbabio.2006.07.005 [DOI] [PubMed] [Google Scholar]
  97. Sopko R, Huang D, Preston N, Chua G, Papp B, Kafadar K, Snyder M, Oliver SG, Cyert M, Hughes TR, Boone C, Andrews B. 2006. Mapping pathways and phenotypes by systematic gene overexpression. Molecular Cell 21:319–330. doi: 10.1016/j.molcel.2005.12.011 [DOI] [PubMed] [Google Scholar]
  98. Spiegel S, Milstien S. 2003. Sphingosine-1-phosphate: an enigmatic signalling lipid. Nature Reviews Molecular Cell Biology 4:397–407. doi: 10.1038/nrm1103 [DOI] [PubMed] [Google Scholar]
  99. Stathopoulos-Gerontides A, Guo JJ, Cyert MS. 1999. Yeast calcineurin regulates nuclear localization of the Crz1p transcription factor through dephosphorylation. Genes & Development 13:798–803. doi: 10.1101/gad.13.7.798 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Sun Y, Miao Y, Yamane Y, Zhang C, Shokat KM, Takematsu H, Kozutsumi Y, Drubin DG. 2012. Orm protein phosphoregulation mediates transient sphingolipid biosynthesis response to heat stress via the Pkh-Ypk and Cdc55-PP2A pathways. Molecular Biology of the Cell 23:2388–2398. doi: 10.1091/mbc.E12-03-0209 [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Swinnen E, Wilms T, Idkowiak-Baldys J, Smets B, De Snijder P, Accardo S, Ghillebert R, Thevissen K, Cammue B, De Vos D, Bielawski J, Hannun YA, Winderickx J. 2014. The protein kinase Sch9 is a key regulator of sphingolipid metabolism in Saccharomyces cerevisiae. Molecular Biology of the Cell 25:196–211. doi: 10.1091/mbc.E13-06-0340 [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Tabuchi M, Audhya A, Parsons AB, Boone C, Emr SD. 2006. The phosphatidylinositol 4,5-biphosphate and TORC2 binding proteins Slm1 and Slm2 function in sphingolipid regulation. Molecular and Cellular Biology 26:5861–5875. doi: 10.1128/MCB.02403-05 [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Tatebe H, Morigasaki S, Murayama S, Zeng CT, Shiozaki K. 2010. Rab-family GTPase regulates TOR complex 2 signaling in fission yeast. Current Biology 20:1975–1982. doi: 10.1016/j.cub.2010.10.026 [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Vallée B, Riezman H. 2005. Lip1p: a novel subunit of acyl-CoA ceramide synthase. The EMBO Journal 24:730–741. doi: 10.1038/sj.emboj.7600562 [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Westfall PJ, Patterson JC, Chen RE, Thorner J. 2008. Stress resistance and signal fidelity independent of nuclear MAPK function. Proceedings of the National Academy of Sciences of USA 105:12212–12217. doi: 10.1073/pnas.0805797105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Yaffe MB, Leparc GG, Lai J, Obata T, Volinia S, Cantley LC. 2001. A motif-based profile scanning approach for genome-wide prediction of signaling pathways. Nature Biotechnology 19:348–353. doi: 10.1038/86737 [DOI] [PubMed] [Google Scholar]
  107. Yang Q, Inoki K, Ikenoue T, Guan KL. 2006. Identification of Sin1 as an essential TORC2 component required for complex formation and kinase activity. Genes & Development 20:2820–2832. doi: 10.1101/gad.1461206 [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Zheng Z, Zou J. 2001. The initial step of the glycerolipid pathway: identification of glycerol 3-phosphate/dihydroxyacetone phosphate dual substrate acyltransferases in Saccharomyces cerevisiae. The Journal of Biological Chemistry 276:41710–41716. doi: 10.1074/jbc.M104749200 [DOI] [PubMed] [Google Scholar]
  109. Zimmermann C, Santos A, Gable K, Epstein S, Gururaj C, Chymkowitch P, Pultz D, Rødkær SV, Clay L, Bjørås M, Barral Y, Chang A, Færgeman NJ, Dunn TM, Riezman H, Enserink JM. 2013. TORC1 inhibits GSK3-mediated Elo2 phosphorylation to regulate very long chain fatty acid synthesis and autophagy. Cell Reports 5:1036–1046. doi: 10.1016/j.celrep.2013.10.024 [DOI] [PubMed] [Google Scholar]
eLife. 2014 Oct 3;3:e03779. doi: 10.7554/eLife.03779.014

Decision letter

Editor: Tony Hunter1

eLife posts the editorial decision letter and author response on a selection of the published articles (subject to the approval of the authors). An edited version of the letter sent to the authors after peer review is shown, indicating the substantive concerns or comments; minor concerns are not usually shown. Reviewers have the opportunity to discuss the decision before the letter is sent (see review process). Similarly, the author response typically shows only responses to the major concerns raised by the reviewers.

Thank you for sending your work entitled “TORC2-dependent protein kinase Ypk1 phosphorylates ceramide synthase to stimulate synthesis of complex sphingolipids” for consideration at eLife. Your article has been favorably evaluated by Tony Hunter (Senior Editor), who acted as the Reviewing Editor, and 3 reviewers.

The Reviewing Editor and the other reviewers discussed their comments before we reached a decision, and the Reviewing Editor has assembled the following comments to help you prepare a revised submission.

They all agreed that this is a rather complete story that illustrates the power of using a combination of protein kinase consensus sequence based bioinformatics and synthetic dosage lethality in yeast to identify protein kinase targets in key cellular pathways. Your finding that stress-activated Ypk1, an orthologue of the mammalian Sgk1 protein kinase, phosphorylates and stimulates the Lag1/Lac1 catalytic subunits of the ceramide synthase complex thus coordinating regulation of the sphingolipid biosynthesis pathway in response to stress is a significant advance and constitutes a very nice piece of work that merits publication in eLife. As you may be aware, eLife editorial practice is to assemble a decision letter that summarizes the main points that need to be addressed before the paper can be accepted. In this case, the only major point is that better evidence is needed to establish your conclusion that constitutively active Ypk1 is deleterious to growth because it results in induction of autophagy (Reviewers 2 and 3). Otherwise, the reviewers' comments are relatively minor, but appear to be eminently doable and worth addressing in revision, and they have been included below (together with the complimentary prefatory remarks!).

Reviewer 1:

The paper describes a unique, well conceived and well described candidate genetic screen to identify phosphorylation substrates of yeast, Ypk1, a TORC dependent protein kinase. Beyond the screen itself, the paper focuses on characterization of the functional significance of two clearly identified targets, Lag1 and Lac1. Most significantly, the study demonstrates that phosphorylation of ceramide synthase by Ypk1 leads to up-regulation of its activity. Moreover, they show that calcineurin dephosphorylates ceramide synthase. These results significantly expand the known targets of Ypk1 and provide significant new details of regulation of sphingolipid synthesis by a conserved protein kinase. This paper constitutes a significant advance in understanding regulation of crucial aspects of sphingolipid lipid synthesis in the yeast model system.

The only significant criticism concerns the data in Figure 6B. This part of the Figure is not clearly labeled. Are the two side-by-side lanes, under each strain, repeats? They do not look very reproducible and there is no quantification given in the Table or elsewhere in the paper. Furthermore, there is an unlabeled band below the MIPCs. Is this an unknown lipid or is it M(IP)2C? Where is M(IP)2C? Some quantification of relative amounts of these sphingolipids and their relative proportions in the three strains should be presented and/or discussed in more detail.

Reviewer 2:

This is an interesting study combining bioinformatics, genetics, and biochemistry to identify novel substrates of the yeast kinase YPK1 (ortholog of mammalian SGK isoforms). Two highly related ER transmembrane proteins that comprise the ceramide synthase complex, Lac1 and Lag1, are identified as bona fide YPK1 substrates whose phosphorylation is required for survival following inhibition of sphingolipid synthesis, a condition which activates YPK1. Metabolite profiles and enzymatic assays suggest that the phosphorylation stimulates ceramide synthase activity. However, links back to the stresses that activate TORC2 are not sufficiently made. Genetic evidence is provided that the YPK1-mediated stimulation of Lac1/Lag1 protects cells from the growth inhibitory effects of long chain bases and autophagy. This is a particularly interesting aspect of the proposed model, but the link to autophagy is not well established by the data shown.

Specific comments:

1) The expression levels of the various phospho-mutant and mimetic alleles relative to wild-type Lac1 and Lag1 need to be shown under the different experimental conditions. These are stated as “data not shown”, but are essential controls for interpreting the data presented. This is particularly important for the IP-enzymatic assay in Figure 6C. The levels of each in the IPs should be shown, with the reactions normalized to differences in enzyme levels in the IP. It is unclear from the current description how these assays were normalized (e.g., what is the “protein unit” on the Y-axis?).

2) To make a clearer link between the stresses that activate YPK1 and the phosphorylation-dependent activation of ceramide synthase, the assays in Figure 6C should be repeated, comparing wild-type to ypk1-deleted strains in the presence or absence of myriocin. Furthermore, the ability of myriocin to stimulate ceramide synthase should be assessed comparing wild-type to the Lac1/Lag1-SSAA mutant.

3) Figure 7: It is proposed that the growth inhibitory effects of YPK1 activation and lc3b-deletion on the Lac1/Lag1-SSAA mutant are through autophagy. This is only supported by an atg1-deletion having very modest rescue effects. It is reasonable that some measurement of autophagy be undertaken in these settings to support such a conclusion.

Reviewer 3:

Muir et al. identified a number of novel substrates for Ypk1 kinase using a very elegant genome-wide screen combined with an in vitro Ypk1 kinase assay. Among the hits, the authors demonstrate that Ypk1 directly phosphorylates and activates the ceramide synthases Lag1 and Lac1. This finding answers the long-standing question of how the TORC2-Ypk signaling pathway controls ceramide synthesis, which was previously shown by the Ted Powers group (Aronova et al., 2008). Experiments are properly controlled, the manuscript is well written, and the finding significantly advances our understanding of TORC2-Ypk signaling in sphingolipid metabolism. Specific comments, all relatively minor, are below.

Specific comments:

1) The title “TORC2-dependent protein kinase Ypk1 phosphorylates ceramide synthase to stimulate synthesis of complex sphingolipids” is not fully supported by the experiments. It would be more accurate if the authors demonstrated that Lag1 and Lac1 phosphorylation is regulated also by TORC2, using tor2 or ypk1-T662A (mutated TORC2 target site) mutant cells. Both strains were used in their previous publication (Roelants et al., 2011).

2) Figure 6. The authors demonstrate that Ypk1-mediated phosphorylation of Lag1 and Lac1 stimulates ceramide synthase activity by measuring levels of complex sphingolipids, downstream products of ceramide, rather than by measuring ceramide levels per se. Is there a reason for this?

3) Figure 7. The authors show that lag1/lac1 phospho-deficient mutant (AA) cells exhibit a slow growth phenotype when a constitutively active Ypk1 is expressed. Since an observed upregulation of LCBPs in the AA cells was shown to impede cell growth by triggering autophagy (Zimmermann et al., 2013), the authors conclude the expression of constitutive Ypk1 is deleterious to the growth of the lag1/lac1 AA cells most likely due to hyper accumulation of LCBPs inducing autophagy. The authors should measure if expression of Ypk1 in AA cells indeed increases LCBPs.

4) Figure 7. Related to the point 3 above, it would strengthen the authors' conclusion that Ypk1-mediated stimulation of Lag1 and Lac1 prevents autophagy if they could show autophagy is indeed upregulated upon expression of constitutively active Ypk1 in AA cells.

eLife. 2014 Oct 3;3:e03779. doi: 10.7554/eLife.03779.015

Author response


[…] In this case, the only major point is that better evidence is needed to establish your conclusion that constitutively active Ypk1 is deleterious to growth because it results in induction of autophagy (Reviewers 2 and 3). Otherwise, the reviewers' comments are relatively minor, but appear to be eminently doable and worth addressing in revision, and they have been included below (together with the complimentary prefatory remarks!).

We appreciate the several suggestions for further improving the impact of this work. As requested by the editor (and reviewers 2 and 3), we now provide new data (by examining processing of GFP-Atg8) demonstrating that basal autophagy is indeed increased when constitutively-activated Ypk1 (which elevates flux into long-chain base synthesis by alleviating Orm inhibition of SPT) is combined with alleles of Lag1 and Lac1 that cannot concomitantly undergo Ypk1-dependent phosphorylation (and thus cannot efficiently divert the long-chain bases generated into ceramides). Moreover, we appreciate 'the complimentary prefatory remarks' from all three referees and have attended to the additional ('relatively minor') issues raised by them, which we now have addressed in all instances by direct experiment, where warranted.

Reviewer 1:

The paper describes a unique, well conceived and well described candidate genetic screen to identify phosphorylation substrates of yeast, Ypk1, a TORC dependent protein kinase. Beyond the screen itself, the paper focuses on characterization of the functional significance of two clearly identified targets, Lag1 and Lac1. Most significantly, the study demonstrates that phosphorylation of ceramide synthase by Ypk1 leads to up-regulation of its activity. Moreover, they show that calcineurin dephosphorylates ceramide synthase. These results significantly expand the known targets of Ypk1 and provide significant new details of regulation of sphingolipid synthesis by a conserved protein kinase. This paper constitutes a significant advance in understanding regulation of crucial aspects of sphingolipid lipid synthesis in the yeast model system.

We thank this referee for these extremely laudatory comments.

The only significant criticism concerns the data in Figure 6B. This part of the Figure is not clearly labeled. Are the two side-by-side lanes, under each strain, repeats? They do not look very reproducible and there is no quantification given in the Table or elsewhere in the paper. Furthermore, there is an unlabeled band below the MIPCs. Is this an unknown lipid or is it M(IP)2C? Where is M(IP)2C? Some quantification of relative amounts of these sphingolipids and their relative proportions in the three strains should be presented and/or discussed in more detail.

In the original Figure 6B, and as clearly stated in the legend to original Figure 6B, duplicate cultures of each genotype grown to saturation were analyzed; hence, the two side-byside lanes were, in essence, independent repeats. We have done this analysis many times and always see the same trend. However, to address this concern, we repeated this analysis several more times using cultures grown to mid-exponential phase and refined our TLC separation solvent system to better resolve the IPC and MIPC species [we did not resolve MIPCs from M(IP)2Cs]. The IPCs and MIPCs each separate into a family of bands due to differences in the length of the amide-linked fatty acid attached in these ceramides. We now show new representative data for such separations with accompanying quantitation, which fully support our original conclusions.

Reviewer 2:

This is an interesting study combining bioinformatics, genetics, and biochemistry to identify novel substrates of the yeast kinase YPK1 (ortholog of mammalian SGK isoforms). Two highly related ER transmembrane proteins that comprise the ceramide synthase complex, Lac1 and Lag1, are identified as bona fide YPK1 substrates whose phosphorylation is required for survival following inhibition of sphingolipid synthesis, a condition which activates YPK1. Metabolite profiles and enzymatic assays suggest that the phosphorylation stimulates ceramide synthase activity. However, links back to the stresses that activate TORC2 are not sufficiently made. Genetic evidence is provided that the YPK1-mediated stimulation of Lac1/Lag1 protects cells from the growth inhibitory effects of long chain bases and autophagy. This is a particularly interesting aspect of the proposed model, but the link to autophagy is not well established by the data shown.

We thank this referee for the generally favorable comments. However, this referee felt that we needed to better document two points: (i) that stimulation of Ypk1 function is linked to TORC2 action; and, (ii) that increased autophagy is linked to elevated long-chain base formation. We now have addressed both by direct experiment. First, using an analog-sensitive Tor2 allele [in cells containing Tor1, Tor2 is only found in the TORC2 complex] and a TORdirected inhibitor (BEZ-235) at 1 μM concentration, we demonstrate that Ypk1-dependent phosphorylation of Lac1 is obliterated almost completely, even in the absence of any stress (see Figure 3, new Panel C). Second, as discussed above in our response to the Editor's comments, we observe a statistically significant increase in the level of GFP-Atg8 processing in Lac1(S23A S24A) Lag1(S23A S24A) cells expressing constitutively-active Ypk1, compared to control cells (see Figure 7, new Panel E).

Specific comments:

1) The expression levels of the various phospho-mutant and mimetic alleles relative to wild-type Lac1 and Lag1 need to be shown under the different experimental conditions. These are stated as “data not shown”, but are essential controls for interpreting the data presented. This is particularly important for the IP-enzymatic assay in Figure 6C. The levels of each in the IPs should be shown, with the reactions normalized to differences in enzyme levels in the IP. It is unclear from the current description how these assays were normalized (e.g., what is the “protein unit” on the Y-axis?).

The referee felt that it was important to show, rather than simply state as 'data not shown,' an immunoblot documenting equivalent expression of wild-type Lac1 and Lag1, and the derived Lac1(S23A S24A) Lag1(S23A S24A) and Lac1(S23E S24E) Lag1(S23E S24E) mutants, in the cells in which our genetic and biochemical analyses were performed. A representative example of such data is now provided for one such experiment (see Figure 4, new Panel D), where the control for equal total sample loading is a cytosolic marker protein (Pgk1). In the case of the ceramide synthase assays, the procedure for a comparable immunoblotting analysis to quantify the amount of Lac1 and Lag1 present, and for normalizing to obtain the specific activity, now is given in detail in the revised Materials and Methods and in the accompanying figure legend.

2) To make a clearer link between the stresses that activate YPK1 and the phosphorylation-dependent activation of ceramide synthase, the assays in Figure 6C should be repeated, comparing wild-type to ypk1-deleted strains in the presence or absence of myriocin. Furthermore, the ability of myriocin to stimulate ceramide synthase should be assessed comparing wild-type to the Lac1/Lag1-SSAA mutant.

As requested by this referee, we now provide new data (Figure 6C, new lower panel) that directly address this issue by showing that, unlike WT cells, which display an increase in ceramide synthase activity in response to challenge with myriocin, otherwise isogenic Lac1(S23A S24A) Lag1(S23A S24A) do not. Moreover, we would like to remind the referee that we already demonstrated that Lac1 (and Lag1) are phospho-proteins in vivo (Figure 3A), that their phosphorylation in vivo requires Ypk1 (Figure 3B), that membrane stresses, including myriocin-treatment (Figure 4B) and heat treatment (Figure 4B), stimulate the Ypk1-mediated phosphorylation of Lac1 and Lag1, and that the inability to phosphorylate the Ypk1 sites in Lac1 and Lag1 causes an increase in LCBs and decrease in ceramides, compared to wild-type cells, determined by two independent methods (Figure 6A & 6B).

3) Figure 7: It is proposed that the growth inhibitory effects of YPK1 activation and lc3b-deletion on the Lac1/Lag1-SSAA mutant are through autophagy. This is only supported by an atg1-deletion having very modest rescue effects. It is reasonable that some measurement of autophagy be undertaken in these settings to support such a conclusion.

As requested by this referee, and as mentioned above in response to the editor, we now provide new data (by examining processing of GFP-Atg8) demonstrating that basal autophagy is indeed elevated both when Lcb3 is absent (so LCB-Ps accumulate) and when constitutively-activated Ypk1 (which elevates flux into long-chain base synthesis by alleviating Orm inhibition of SPT) is combined with alleles of Lag1 and Lac1 that cannot concomitantly undergo Ypk1-dependent phosphorylation (and thus cannot efficiently divert the long-chain bases generated into ceramides, so that LCB-Ps also accumulate) (see Figure 7, new Panel E).

Given the recent report of Zimmermann C et al. (2013) that we cited in the original manuscript, in which it is rather convincingly shown that accumulation of LCB-Ps is a condition that stimulates autophagy, I also think it is fair to state here that the genetic evidence we provided for this conclusion in our case was already quite strong because not only did loss of Lcb3 (phosphatase) make the growth of Lac1(S23A S24A) Lag1(S23A S24A) cells decidedly worse (Figure 7D), but conversely loss of Lcb4 (kinase) distinctly ameliorated the growth of Lac1(S23A S24A) Lag1(S23A S24A) cells (Figure 7C), in addition to the detectable growth rescue conferred by loss of the essential autophagy inducer, Atg1.

Reviewer 3:

Muir et al. identified a number of novel substrates for Ypk1 kinase using a very elegant genome-wide screen combined with an in vitro Ypk1 kinase assay. Among the hits, the authors demonstrate that Ypk1 directly phosphorylates and activates the ceramide synthases Lag1 and Lac1. This finding answers the long-standing question of how the TORC2-Ypk signaling pathway controls ceramide synthesis, which was previously shown by the Ted Powers group (Aronova et al., 2008). Experiments are properly controlled, the manuscript is well written, and the finding significantly advances our understanding of TORC2-Ypk signaling in sphingolipid metabolism. Specific comments, all relatively minor, are below.

We are very gratified that this referee feels that our work 'answers a longstanding question' and also 'significantly advances our understanding' and presents our findings in a manner that is 'properly controlled' and 'well written,' and has concerns categorized as only 'relatively minor.'

Specific comments:

1) The title “TORC2-dependent protein kinase Ypk1 phosphorylates ceramide synthase to stimulate synthesis of complex sphingolipids” is not fully supported by the experiments. It would be more accurate if the authors demonstrated that Lag1 and Lac1 phosphorylation is regulated also by TORC2, using tor2 or ypk1-T662A (mutated TORC2 target site) mutant cells. Both strains were used in their previous publication (Roelants et al., 2011).

As recommended by this referee, and already discussed above, we took a different approach, but one highly related to the referee's suggestions, to document that the Ypk1-mediated phosphorylation of ceramide synthase is indeed TORC2-dependent (see Figure 3, new Panel C).

2) Figure 6. The authors demonstrate that Ypk1-mediated phosphorylation of Lag1 and Lac1 stimulates ceramide synthase activity by measuring levels of complex sphingolipids, downstream products of ceramide, rather than by measuring ceramide levels per se. Is there a reason for this?

In yeast, all of the ultimate, ceramide-derived, sphingolipid pathway products ('complex sphingolipids') contain one or more phosphate groups. Hence, it is experimentally very convenient to measure the rate of de novo flux through the entire pathway by measuring incorporation of [32P]PO4 3- into these end-products, which is why we chose to do so.

3) Figure 7. The authors show that lag1/lac1 phospho-deficient mutant (AA) cells exhibit a slow growth phenotype when a constitutively active Ypk1 is expressed. Since an observed upregulation of LCBPs in the AA cells was shown to impede cell growth by triggering autophagy (Zimmermann et al., 2013), the authors conclude the expression of constitutive Ypk1 is deleterious to the growth of the lag1/lac1 AA cells most likely due to hyper accumulation of LCBPs inducing autophagy. The authors should measure if expression of Ypk1 in AA cells indeed increases LCBPs.

The experiment requested by this referee has now been performed by mass spectrometry, and expression of constitutively-active Ypk1 in Lac1(S23A S24A) Lag1(S23A S24A) cells does indeed lead to a significance increase in the level of LCB-Ps, compared to the control cells (see Figure 7, new Panel B).

4) Figure 7. Related to the point 3 above, it would strengthen the authors' conclusion that Ypk1-mediated stimulation of Lag1 and Lac1 prevents autophagy if they could show autophagy is indeed upregulated upon expression of constitutively active Ypk1 in AA cells.

To corroborate our genetic findings (Figure 7), we now provide biochemical evidence (level of processing of GFP-Atg8) that supports the conclusion that the level of basal autophagy is elevated in Lac1(S23A S24A) Lag1(S23A S24A) cells expressing constitutivelyactive Ypk1, compared to control cells (Figure 7, new Panel E), in agreement with the fact that the same cells exhibit a significant increase in the level of LCB-Ps, compared to control cells (see Figure 7, new Panel B).


Articles from eLife are provided here courtesy of eLife Sciences Publications, Ltd

RESOURCES