Summary:
In mammalian cells, two phosphatidylserine (PS) synthases drive PS synthesis. Gain-of-function mutations in the Ptdss1 gene lead to heightened PS production, causing Lenz-Majewski syndrome (LMS). Recently, pharmacological inhibition of PSS1 has been shown to suppress tumorigenesis. Here, we report the cryo-EM structures of wild-type human PSS1 (PSS1WT), the LMS-causing Pro269Ser mutant (PSS1P269S) and PSS1WT in complex with its inhibitor DS55980254. PSS1 contains 10 transmembrane-helices (TMs), with TMs 4–8 forming a catalytic core in the luminal leaflet. These structures revealed a working mechanism of PSS1 akin to the postulated mechanisms of the membrane-bound O-acyltransferase family. Additionally, we showed that both PS and DS55980254 can allosterically inhibit PSS1, and that inhibition by DS55980254 activates the SREBP pathways, thus enhancing the expression of LDL receptors and increasing cellular LDL uptake. This work uncovers a mechanism of mammalian PS synthesis and suggests that selective PSS1 inhibitors have the potential to lower blood cholesterol levels.
Keywords: Phosphatidylserine, cholesterol trafficking, PSS1, MBOAT, LDL, LDL receptors, DS55980254
Graphical Abstract

In Brief:
PSS1 employs its catalytic histidine and a calcium ion in the transmembrane core to replace the polar head of PC or PE for PS synthesis and specific PSS1 inhibitors can block this synthesis to activate the SREBP pathway, increasing LDL receptor expression, and promoting LDL uptake in cells.
CELL-D-24-00876R3
Structural Insights into Human Phosphatidylserine Synthase 1 and How Its Inhibition Triggers SREBP Activation
Introduction
Phosphatidylserine (PS), the most abundant negatively charged glycerophospholipid in eukaryotic membranes, functions as a component of the cell membrane and plays vital roles in intracellular signaling, cell death, blood coagulation, cholesterol trafficking, and viral infection1–5. It has also been shown to protect the health of nerve cell membranes and myelin6,7. PS is preferentially distributed in the cytosolic leaflet of the plasma membrane and is also found in late endocytic compartments. The movement of PS from the inner to the outer leaflet of the plasma membrane takes place during apoptosis, which is thought to be pivotal in triggering cell death8,9. PS is also exploited by many pathogens for their cell entry and propagation, for example, HIV and Ebola viruses can attack host cells by binding to exposed PS10–12. On the surface of activated platelets, the exposed PS engages coagulation factors, leading to an increased production of thrombin, indicating its crucial role in significantly promoting blood clotting13,14. Moreover, PS has been shown to modulate the activity of many proteins, such as protein kinase C, Raf-1, Na+/K+ ATPase, and the AMPA glutamate receptor15–17.
The mechanisms of phospholipid biosynthesis from yeast to mammalian cells are typically conserved. In general, cells make Phosphatidylcholine (PC) or Phosphatidylethanolamine (PE) via the CDP-choline pathway or the CDP-ethanolamine pathway, but the mechanism of PS synthesis is distinct from that of the other phospholipids18. In yeast, PS is generated by the replacement of cytidine monophosphate of CDP-diacylglycerol with l-serine, whereas in mammalian cells, the de novo synthesis of PS occurs at the endoplasmic reticulum (ER) in a calcium-dependent manner19. Two membrane enzymes, named phosphatidylserine synthases 1 (PSS1) and phosphatidylserine synthases 2 (PSS2), are responsible for exchanging the polar head group of PC or PE by l-serine19,20 (Figure 1A). PSS2 shows a preference for PE as its substrate, whereas PSS1 can utilize either PE or PC19. Animal studies found that mice with either Ptdss1 or Ptdss2 gene deficiency are viable, however, the knockout of both genes is lethal21. While PSS1 and PSS2 are well conserved across different species (Figure S1A) with an amino acid sequence similarity of over 50%, their expression profiles vary depending on the tissue type22,23.
Figure 1. Functional characterization and overall structure of human PSS1.

(A) Schematic of PS synthesis. The structures of serine, choline and ethanolamine are shown.
(B) The activity of PSS1WT and PSS1P269S. Data are mean ± S.D. (n=3). P values (two-sided) were calculated by Student’s t tests using GraphPad Prism v.9; **, P<0.01; ****, P<0.0001.
(C) DS55980254 (termed as “DS”) inhibits the activity of PSS1WT and PSS1P269S in vitro. Data are mean ± S.D. (n=3). The 2D chemical structure of DS is shown. The IC50 of DS to PSS1WT is 1.65 μM (the deviation range is 1.22–2.26 μM). The IC50 of DS to PSS1P269S is 1.27 μM (the deviation range is 0.87–1.87 μM).
(D) Overall structure showing PSS1WT dimer viewed from the side of the membrane. The cryo-EM map of the putative PE is shown.
(E) View of the dimer from the cytosol. The cryo-EM map of the putative PI is shown.
(F) View of the dimer from the lumen. TMs1-3 and TM9 are located at the interface and are involved in dimer assembly. TM5 and TM6 contribute to engage the calcium (green sphere).
(G) Structural details of the calcium binding site.
(H) Structural details of PI-mediated dimeric interface. The putative PC (gray sticks), PS (yellow sticks), PI (red sticks) and PE (cyan sticks) are shown. The interactions between calcium and residues are indicated by dashed lines. R349 from another protomer is underlined.
See also Figures S1 and S2.
Systematic alanine scanning revealed that the catalytic site of PS synthesis is in the luminal leaflet of PSS124,25. In addition, SERINC proteins, which incorporate serine into membranes, were identified to facilitate the synthesis of PS26 by providing ample amounts of accessible l-serine. However, higher concentrations of PS in the cell membrane can inhibit the activity of PSS1, indicating a feedback regulation of PS production19. Despite these previous studies, the details of the catalytic and regulatory mechanisms of PS synthesis remain unclear.
Membrane concentration of PS is highly correlated to cholesterol metabolism. Low-density lipoprotein (LDL)-derived cholesterol first moves from the lysosome to the plasma membrane and then to the ER. The movement from the plasma membrane to the ER has been shown to be highly dependent on the presence of PSS127,28. PS also serves as a negative regulator of acyl-CoA cholesterol acyltransferases (ACATs), which convert cholesterol to cholesteryl ester for cholesterol storage29. These findings indicate an important role of PSS1 in the intracellular cholesterol transport and homeostasis. Moreover, gain-of-function mutations on the Ptdss1 gene cause Lenz-Majewski syndrome (LMS), which is characterized by a combination of intellectual disability, sclerosing bone dysplasia, and distinct appearance anomalies30. These mutants have been shown to disrupt the pattern formation of the actin cytoskeleton in osteoclasts31.
Recently, two studies demonstrated that pharmacological inhibition of PSS1 can suppress the growth of certain tumors and contribute to current therapeutic approaches for treating Ptdss2-deficient cancer and B cell lymphomas32,33. Thus, structural information of PSS1 will help the design of its potent inhibitors. Here, we report cryo-electron microscopy (cryo-EM) structures of wild-type human PSS1 (PSS1WT), its P269S mutant (PSS1P269S) and inhibitor-bound PSS1WT, which along with functional assays and molecular dynamics (MD) simulations uncover the working mechanism of PSS1. Moreover, our functional studies demonstrate that pharmacological inhibition of PSS1 may contribute to stimulate LDL uptake in cells.
Results
Overall Structure of PSS1WT
We overexpressed Flag-tagged human PSS1WT and PSS1P269S in HEK293 cells and purified the recombinant proteins by an anti-Flag M2 resin with CaCl2 at a final concentration of 5 mM. Gel filtration reveals that the resulting proteins migrate as a single peak (Figure S1B). To further validate the activity of recombinant PSS1, we used liquid chromatography–mass spectrometry (LC-MS/MS) to examine the synthesized PS. We supplemented l-serine into liposomes containing 2 mM 1-palmitoyl-2-oleoyl-glycero-3-phosphocholine (POPC) at different concentrations and incubated with PSS1WT or PSS1P269S at 37 °C for 90 minutes. The lipids were extracted and analyzed by LC-MS/MS. Our results showed that PSS1P269S exhibits more robust activity than PSS1WT, confirming P269S as a gain-of-function mutation (Figure 1B). The IC50 of DS55980254 to PSS1WT and PSS1P269S is 1.65 and 1.27 μM respectively in the presence of 200 μM serine, showing proper enzymatic activity of recombinant PSS1WT and PSS1P269S in vitro (Figure 1C).
We determined the structure of PSS1WT by cryo-EM at 2.7-Å resolution (Figure 1D–F, Figure S2, and Table S1). Two PSS1 molecules form a homodimer, and the overall dimensions of the dimer are ~100 Å × 55 Å × 50 Å. Each monomer contains four peripheral α-helices (PH) and 10 transmembrane helices (TMs) with the dimensions of ~60 Å × 55 Å × 50 Å, consistent with the previous prediction of its topology25. C2 symmetry was applied during data processing. The local resolution of the transmembrane helix core is about ~2.5-Å, which allows us to assign the side chains of most residues and most of associated endogenous lipids (Figure S2E). A homology search using Dali34 shows that the overall structure of PSS1 does not share any similar fold with membrane proteins whose structures have been determined before. TMs 4–8 form the catalytic cavity in the luminal leaflet, which opens to both the lipid bilayer and lumen (Figure 1F). A phospholipid density was found between the luminal halves of TM4 and TM8 in the cryo-EM map of PSS1WT (Figure 1D). Based on the morphology of this density and the chemical environment of the binding site, we tentatively assigned this lipid as a PE. We calculated the binding free energy of either POPC or POPE to PSS1WT using the molecular mechanics-Poisson-Boltzmann surface area (MM-PBSA) method with the implicit membrane model35. The results showed that POPC has a slightly stronger binding affinity compared to POPE (Figure S3A). Previous findings indicated that PSS1 may associate with calcium19. Consistent with this result, a calcium ion is coordinated by residues Glu197, Glu200, Gln217 and Asp221 in TM5 and TM6 (Figure 1G).
The dimer interface between the two PSS1 molecules comprises about 2,500 Å2. TMs 1–3 and TM9 are involved in the assembly of the dimer (Figure 1F). Interestingly, two phospholipids were found in the dimeric interface. Owing to the electron density of these lipids in the cryo-EM map, we tentatively assigned them as phosphoinositides (PIs) (Figure 1E). TMs 1–2 and TM9 of the other protomer generate the PI binding site. The main-chain carbonyls of Leu86 and Phe88 and the guanidino group of Arg102 along with the guanidino group of Arg349 in the other protomer engage the polar head of PI (Figure 1H). The structural analysis shows that PS association and the engagement of phospholipid substrates in PSS1 do not depend on its dimeric state. However, it is possible that the dimeric assembly plays an essential role in stabilizing the enzyme, thus indirectly facilitating PS production. The rationale behind the dimeric assembly for PSS1 function remains unclear and warrants further investigation.
PS-binding sites in the cytosolic leaflet
Two phospholipid-binding pockets are found around PH4 in the cytosolic leaflet of each PSS1 protomer (Figure 2A). A phospholipid density was observed in each positively charged cavity (Figure 2B). Based on the morphology of the density and the chemical environment of the binding site, we assigned each lipid as PS. Owing to the flexible nature, the lipid tails of PS1 were not completely observed in the cryo-EM map. The first PS binding pocket (termed “Site 1”) consists of the linker between PH1 and TM1, the linker between TM2 and TM3, and PH4, while the second pocket (termed “Site 2”) is formed by TMs 3, 4, 8 and PH4. The binding free energy of PS to PSS1WT is −70.2 kcal/mol for the PS in Site 1 (PS1) and −83.4 kcal/mol for the PS in Site 2 (PS2) from the MD simulations (Figure S3B).
Figure 2. The cytosolic PS-binding sites.

(A) Luminal view of two PS-binding sites. One PSS1 monomer is colored in blue, and the other is colored in cyan. The PS molecules are shown by yellow sticks. The residues that contribute to the rigidity of the sites are indicated. Pro269 is highlighted in red. PSs are only labeled in one subunit, but present in both subunits with C2 symmetry.
(B) Electrostatic surface representation of the cytosolic PS-binding sites.
(C) Interaction details of the PS1-binding site. The cryo-EM map of PS1 is shown.
(D) Interaction details of the PS2-binding site. Residues are represented as sticks; dashed lines represent hydrophilic interactions with the distance. The cryo-EM map of PS2 is shown.
See also Figure S2.
Two π-cation interactions formed by Arg95/Trp247 and Arg336/Trp272 stabilize both cavities (Figure 2A). In Site 1, residues Arg95, Arg262, and Gln266 bind to PS1 via hydrophilic interactions, while residues Phe37, Phe250, Ala263, and Phe267 are responsible for accommodating the fatty acid chains of PS1 (Figure 2C). In Site 2, there are more extensive interactions between PS and PSS1 than that in Site 1, which is consistent with the stronger binding in Site 2 from the MD simulations. Residues Thr273 and Arg336 and the main-chain carbonyl of Arg95 and Phe267 interact with the polar head of PS2, while residues Phe93, Leu100, Trp101, Met178, Phe267, and Leu325 generate a hydrophobic environment to host the fatty acid chains of PS2 (Figure 2D). Alanine scanning demonstrated that mutating Arg95, Arg262, Gln266, or Arg336 to alanine increased the production of PS24, probably owing to the disruption of these two sites. Moreover, two LMS-causing gain-of-function mutations, Leu265Pro and Pro269Ser, are also located in this area (Figure 2A). A previous functional analysis using patient’s samples revealed higher activity of PSS1L265P and PSS1P269S compared to that of PSS1WT 30. Therefore, it is most likely that PSS1 may sense cytoplasmic levels of PS via these two PS-binding sites to regulate its activity. Since PS can inhibit the activity of PSS119,30 and two PS molecules bind to the cytosolic pockets of PSS1WT, we speculate the structure of PSS1WT represents the inactive state.
Overall Structure of PSS1P269S
The structure of PSS1P269S was determined by cryo-EM at 3.3-Å resolution (Figure 3A, Figure S4 and Table S1). In this structure there are also some lipid-like densities between the luminal halves of TM4 and TM8, between TM7 and TM9 and near cytosolic leaflet in the cryo-EM map of PSS1P269S (Figure S4F), however owing to the resolution limitation of the map, we could not unambiguously model them. Although the overall structure of PSS1P269S shares a similar fold to that of PSS1WT, there are several marked conformational changes (Figure 3B). PH1, as well as the linkers that form the PS-binding sites (PH1 to TM1 and TM6 to TM7), are invisible in the cryo-EM map of PSS1P269S. However, these structural elements are still notable when the cryo-EM map of PSS1WT has been filtered to 3.3-Å (Figure S4G). Our MD simulations showed that PH4 (Figure 3C, lower panel) and the linker between PH1 and TM1 (Figure 3C, upper panel) exhibit higher root mean square fluctuations (RMSF) in PSS1WT without PS on a 200 ns time scale compared to in the presence of PS. This indicates that these regions, which are key structural features of the PS-binding sites, are more flexible without PS in the cytosolic sites.
Figure 3. Overall structure of PSS1P269S.

(A) Overall structure showing PSS1P269S dimer viewed from the side of the membrane.
(B) The comparison of TMs in PSS1P269S (yellow, dark green ball) and PSS1WT (blue, light green ball).
(C) R.M.S.F. of linker between PH1 and TM1 (upper) and PH4 (lower) in PSS1WT with (blue) and without PS binding (green). Data of each residue is an average of two monomers in three parallel groups:
(D) Luminal view of PSS1P269S compared to PSS1WT. The conformational changes are indicated by arrows.
(E) Structural comparison of the entrances of the phospholipid substrate.
(F) Structural comparison of the linker between TM5 and TM6.
(G) Changes in the distance between representative amino acid residues (Cα) and the calcium ion with and without PS binding (from 200 ns MD). ΔDistance=Distance in PSSWT without PS binding – Distance in PSSWT with PS binding. Data is an average of three parallel groups. Each group contains two PSS1 molecules in one homodimer.
See also Figures S1, S2 and S4.
TMs 2, 3 and 9 that contribute to the dimeric interface in both structures are well aligned; however, TMs 4–6, 10, and PH3 undergo a notable shift toward to the edge of PSS1 (Figure 3B and D). Structural analysis shows that Phe168, Gly171, Gly175, and Met178 in TM4 and Trp321, Leu325, and Gly328 in TM8 create the gate for phospholipid access (Figure 3E). Although the conformations of TM8 in both structures appear similar, there is a ~5-Å movement of Phe168 at the luminal end of TM4 in the structure of PSS1P269S, causing an opening of the phospholipid-binding site in the luminal leaflet that is absent in the structure of PSS1WT (Figure 3E).
In the structure of PSS1P269S, Glu212 in the linker between TM5 and TM6 as well as Glu197, Glu200, and Asp221 directly bind to the calcium ion (Figure 3F), which is consistent with the AlphaFold predicted model of PSS1WT 36 (Figure S4H), but different from the PSS1WT(Figure 1G). These conformational changes lead to the catalytic core shifting towards to the edge of the enzyme. The MD simulations showed the backbones of the linker between TM5 and TM6 including residues 207–213 move closer to the calcium ion in the absence of PS-binding on a 200 ns time scale; in contrast, residues 214–219 move away from the calcium ion (Figure 3G). Notably, Glu197, Glu200, and Asp221 remain at similar positions regardless of the presence of PS (Figure 3G and Figure S3C). This finding is consistent with our structural observations on the conformational changes in both structures (Figure 3F). A previous study showed that Asn209 plays a crucial role in the l-serine exchange during PS production24 and our structural analysis reveals Asn209 in PSS1P269S is 4-Å closer to calcium than that in PSS1WT (Figure 3F). Thus, it is tempting to speculate that the structure of PSS1P269S represents an active state of PSS1WT without PS binding. In the absence of PS binding, the arrangement of the calcium binding site may facilitate the access of serine to the catalytic core, which would promote the PS synthesis reaction. Furthermore, the binding free energy of POPC or POPE to PSS1WT without PS in the cytosolic sites is lower than it is with PS, indicating a favorable recruitment of phospholipid substrates in the active state (Figure S3A).
Catalytic Mechanism
TMs 4–8 form the catalytic cavity in the luminal leaflet, which opens to both the lipid bilayer and lumen (Figure 4A and B). The cavities in both structures are negatively charged, which is favorable for engaging the positively charged phospholipid substrates (PC and PE), while the calcium ion would facilitate the recruitment of serine into the cavity. Structural analysis reveals that sufficient space is available around the luminal halves of TMs 4–6, particularly in proximity to calcium and His172 within the catalytic center, and a serine can be docked into the calcium site (Figure 4C). Our MD simulations showed that the free energy of binding of serine to PSS1P269S monomer is −8.1 kcal/mol (Figure S3D). His172 may function as a catalytic base to subtract the hydroxyl proton of the serine, which remains trapped by the calcium ion, during the nucleophilic attack of the hydroxyl group on the ester bond between the phosphate group and the polar head of PC or PE to form PS (Figure 4D). After the reaction, the negatively charged PS would be released due to the repulsive electrostatic forces between PS and the cavity (Figure 4A and B).
Figure 4. The catalytic mechanism of PSS1.

(A) and (B) Electrostatic surface representation of catalytic cavity of PSS1WT (A in blue) and PSS1P269S (B in yellow).
(C) The docking model of serine (gray sticks) in the calcium binding site of PSS1P269S.
(D) A working model of PSS1-meidated PS synthesis.
(E) Functional validation of the catalytic residue and lipid entrance. Data are mean ± S.D. (n=3).
(F) Salt bridge between Lys179 and Glu301 induces a shift of Phe168. Residues are represented as sticks; dashed lines represent hydrophilic interactions.
See also Figures S1, S2 and S7.
To validate our model, our functional assay showed that the PSS1H172A presents no activity compared to that of PSS1WT, although the purification results show that PSS1H172A is well behaved in the detergent solution (Figure 4E and Figure S1B). When we mutated Phe168, Gly171 and Gly175 individually to tryptophan to introduce steric hindrance in the lipid gate, PSS1F168W, PSS1G171W and PSS1G175W showed minimal enzymatic activity (Figure 4E and Figure S1B). It has also been shown that Glu197Ala, Glu200Ala, Glu212Ala, and Asp221Ala mutations almost completely abolished PSS1 activity24, supporting their critical roles in coordinating calcium, which in turn is essential for PS synthesis.
In the structure of PSS1P269S, a salt bridge forms between Lys179 and Glu301, causing the cytosolic leaflet of TM4 to be drawn towards TM7. Consequently, this salt bridge pushes the luminal leaflet of TM4 away from TM8 (Figure 4F). Further MD analysis indicates when PS is not bound, the distance between the Cα atoms of Phe168 and Trp321 increases, leading to the opening of the catalytic core facilitating the access of the phospholipid substrates (Figure S3E). Our functional analysis showed that PSS1F168A has almost no activity (Figure 4E and Figure S1B), suggesting that Phe168 plays an important role in engaging the phospholipid substrates, as observed from the structure (Figure 3E). The conformational changes of Phe168 and Asn209 in both structures along with the MD simulations suggest that the catalytic core may be easily accessed by phospholipids and serine in the absence of PS. This sheds light on the mechanism underlying how PS in the cytosolic pockets modulates the activity of PSS1.
Inhibitor-bound State
Most recently, several PSS1 specific inhibitors, all of which share a similar backbone, have been identified32,33. To elucidate the binding mode of these inhibitors, we expressed PSS1WT in the presence of DS55980254 at a final concentration of 1 μM and purified the complex with DS55980254 at 10 μM, determining its structure at 2.87-Å (Figure 5A, Figure S5, and Table S1). DS55980254 binds to the cavity that is generated by TM3, PH2, TM7 and TM8 in the luminal leaflet (Figure 5B). Hydrophilic interactions with DS55980254 are provided by residues Asp134, Ser320, Arg323 and the main chain of Phe313. Phe313 forms a π-π interaction with the inhibitor while His317 forms a π-cation interaction. Importantly, residues Phe313 and Ser320 are not conserved in PSS2, underscoring the specificity of DS55980254 (Figure S1). Our functional analysis also showed that DS55980254 at a final concentration of 100 μM does not inhibit the base-exchange activity of PSS1F313M/S320V (Figure 5C). Comparison with the structure of apo PSS1WT reveals that DS55980254 triggers a conformational change of TM8 and fixes the conformation of TM7 with the loop between TM7 and TM8 via a hydrophilic bond between Lys308 and Phe313 (Figure 5D). To further dissect the inhibitory mechanism, we performed MD simulations to analyze the dynamics of PSS1 with or without DS55980254 binding in the absence of PS (Figure 5E). The results show the dynamics of the cytosolic PS-binding sites (residues 245–275, magenta region in Figure 5A) are reduced when DS55980254 binds to PSS1, which is akin to our structure in the presence of PS. Our finding concludes that PS and DS55980254 both act allosterically, although they bind to different sites (Figures 2A and 5A).
Figure 5. The inhibition of PSS1 by its specific inhibitor DS55980254.

(A) Overall structure of DS55980254 (labeled as “DS”) bound PSS1WT dimer viewed from the side of the membrane. The cryo-EM map of DS (sticks) is shown. The residues 245–275 were colored in magenta.
(B) Interaction details between DS and PSS1. Dashed lines represent hydrophilic interactions.
(C) Functional validation of the DS55980254-binding residues. Data are mean ± S.D. (n=3 independent experiments). P value (two-sided) was calculated by Student’s t tests using GraphPad Prism v.9; ns, P > 0.05.
(D) The comparison of TM7 and TM8 between the inhibitor-bound PSS1WT (light blue) and apo PSS1WT (blue). The conformational changes are indicated by red arrows.
(E) R.M.S.F. of residues 165–300 in PSS1WT with DS (yellow), PS (blue) and without DS and PS (green) during 200 ns MD simulations. Data of each residue is an average of two monomers in three parallel groups:
See also Figures S1, S2 and S5.
PSS1 inhibitor activates SREBP pathway and stimulates LDL uptake
Two studies have revealed that the movement of accessible cholesterol from the PM to the ER depends on GRAMD1/Aster proteins37,38. PS binds to the Aster proteins, which has been shown to be indispensable for Aster-mediated cholesterol transport37. Additionally, PS is also required for an Aster-independent pathway for cholesterol transport28. These studies demonstrate the necessity of PS for cholesterol transport to the ER. Moreover, depletion of Ptdss1 in CHO-K1 cells led to a deficiency of intracellular PS and triggered the cleavage of sterol regulatory element binding protein-2 (SREBP-2) to turn on the SREBP pathway28.
To test the potential link between PSS1 enzymatic activity and the activation of the SREBP pathway, we conducted a SREBP-2 cleavage assay in Ptdss1−/− CHO-K1 cells27 by observing the ratio between the nuclear fragment and the membrane-bound precursor of SREBP-2 by SDS-PAGE. A previous study showed that in wild-type CHO-K1 cells, LDL can inhibit the cleavage of SREBP-2 by ~80%28. Our results show that the LDL-induced inhibition of SREBP-2 cleavage has been abolished in Ptdss1−/− cells (Figure 6A, lanes 1 and 2), while inhibition was restored to 79% when Flag-tagged PSS1WT was overexpressed in Ptdss1−/− cells (Figure 6A, lanes 3 and 4). This is consistent with the previous finding28. Notably, adding LDL into the medium did not prevent the proteolytic process of SREBP-2 in the cells that overexpressed Flag-tagged PSS1WT with DS55980254 at a final concentration of 1 μM (Figure 6A, lanes 5 and 6). When we overexpressed Flag-tagged PSS1H172A, a catalytic mutant (Figure 4E), LDL inhibited the proteolytic processing of SREBP-2 only by 20% (Figure 6A, lanes 7 and 8). These findings indicate a direct correlation between the activity of PSS1 and the cleavage of SREBP-2.
Figure 6. PSS1 deficiency triggers the SREBP-2 cleavage and reduces the ER cholesterol concentration.

(A) The SREBP-2 processing with and without DS55980254 (labeled as “DS”) in Ptdss1−/− CHO-K1 cells. On day 0, cells were set up in 6-well plates at a density of 25 × 104 cells/well. On day 1, cells were switched to 1 ml medium A containing 10% FCS and 100 μl baculovirus expressed hPSS1WT or hPSS1H172A were supplemented into the medium. The inhibitor (DS) was supplemented into the medium at a final concentration of 1 μM at the time of infection and was kept at a constant concentration in each buffer throughout the experiment. On day 2, cells were switched to cholesterol-depletion medium A containing 1% hydroxypropyl-β-cyclodextrin (HPCD), which can deplete the free cholesterol in cells. After incubation for 1 hour, cells received cholesterol-depletion medium A in the absence or presence of 100 μg protein/ml LDL. After 6 hours, cells were harvested for immunoblotting of SREBP-2, Flag and Histone H3. P, precursor. N, nuclear. SREBP cleavage was quantified using ImageJ. For each lane, the ratio of nuclear to total SREBP-2 (nuclear + precursor) was calculated.
(B) Diagram of ER membrane fractionation scheme.
(C) and (D) denote major fractions recovered and analyzed by immunoblot analysis. CHO-K1 cells (C) and Ptdss1−/− CHO-K1 cells (D) were treated according to the fractionation scheme as described in METHOD DETAILS. Aliquots representing equal volumes of each fraction (A–F) were subjected to immunoblot analysis for the indicated organelle markers.
(E) Cholesterol content of the purified ER membranes in CHO-K1 cells and Ptdss1−/− CHO-K1 cells. Lipids were extracted from the purified ER, and the amounts of cholesterol and phospholipids were quantified as described in METHOD DETAILS. Mole % of cholesterol in the ER from different cells was calculated as the ratio of cholesterol to phospholipids plus cholesterol. Data are mean ± S.D. (n=4). P value (two-sided) was calculated by Student’s t tests using GraphPad Prism v.9; **, P<0.01.
(F) Working model of PSS1 inhibitors in simulating LDL uptake.
To investigate the direct relationship between PSS1 and cholesterol concentration in the ER, we isolated the ER fractions from regular CHO-K1 cells and Ptdss1−/− CHO-K1 cells according to a previously established protocol39 (Figure 6B–D). Lipidomics analysis showed that the mole ratio of cholesterol to phospholipids plus cholesterol in the ER of Ptdss1−/− cells is ~35% lower than that in the regular cells (Figure 6E). This result supports our findings from the SREBP-2 cleavage assay, which implies that there is a relatively low cholesterol concentration in the ER of Ptdss1−/− cells (Figure 6A). Since PSS1 deficiency triggers the activation of SREBP, we hypothesize that PSS1 inhibitors are promising compounds to increase the uptake of LDL-derived cholesterol (Figure 6F).
To validate this hypothesis, we conducted the SREBP-2 cleavage assay in CHO-K1 cells treated with DS55980254 from 0.01 μM to 1 μM. The proteolytic processing of SREBP-2 increased as did the mRNA level of LDLRs (Figure 7A–B). When SV589j cells, a line of SV40-immortalized human fibroblasts27, were treated with DS55980254 at a final concentration of 1 μM, the cleavage of SREBP-2 increased (Figure 7C). More importantly, the surface expression of LDLRs in treated cells was three times higher than in untreated cells (Figure 7D and Figure S6A), and after a two-hour treatment with fluorescence labeled LDL, the uptake of LDL doubled in the presence of DS55980254 compared to untreated cells (Figure 7E and Figure S6B). The cholesterol staining assay showed that inhibiting PSS1 activity does not lead to cholesterol accumulation in lysosomes, in contrast, inhibiting NPC1 by its inhibitor U18666A40 causes the substantial accumulation of cholesterol in lysosomes (Figure 7F).
Figure 7. DS55980254 stimulates LDLR expression and LDL uptake without causing cholesterol accumulation in lysosomes or overexpression of HMGCR in the ER.

(A) The SREBP-2 processing and (B) LDLR mRNA level with and without DS55980254 (labeled as “DS”) in CHO-K1 cells. On day 0, cells were set up in 6-well plates at a density of 15 × 104 cells/well. On day 1, DS was supplemented into the medium at different concentration as indicated and was kept at a constant concentration in each buffer throughout the experiment. On day 2, cells were switched to cholesterol-depletion medium A containing 1% hydroxypropyl-β-cyclodextrin (HPCD). After incubation for 1 hour, the cells received cholesterol-depletion medium A in presence of 100 μg protein/ml LDL. After 18 hours, cells were harvested for immunoblotting of SREBP-2, Flag and Histone H3 and measuring LDLR mRNA level. P, precursor. N, nuclear. SREBP cleavage was quantified using ImageJ. For each lane, the ratio of nuclear to total SREBP-2 (nuclear + precursor) was calculated. Data of LDL mRNA level is presented as mean ± S.D. (n=3 biological replicates, with each point representing the average of three technical replicates).
(C) The SREBP-2 processing, (D) surface LDLR expression level and (E) BODIPY FL-LDL uptake with and without DS in SV589j cells. On day 0, cells were set up in 6-well plates at a density of 25 × 104 cells/well. After plating for 6 hours, cells were supplemented with DMSO (solvent control) or 1μM DS, which was kept at a constant concentration in each buffer throughout the experiment. On day 1, cells were switched to cholesterol-depletion medium B. After incubation for 16 hours, the cells received cholesterol-depletion medium B in presence of 50 μg protein/ml LDL. After 24 hours, cells were harvested for immunoblotting of SREBP-2, Flag and Histone H3 or by incubation with EDTA, washed, incubated with PE-anti-LDLR, and subjected to flow cytometry. The rest of cells were supplemented with 5 μg protein/ml BODIPY FL-LDL. After 2 hours, cells were harvested for flow cytometry. Data are mean ± S.D. (n=3). P value (two-sided) was calculated by Student’s t tests using GraphPad Prism v.9; **, P<0.01.
(F) Cholesterol distribution in CHO-K1 cells with different compounds treatment. On day 0, CHO-K1 cells were plated on 12-mm glass coverslips in medium A with 5% FCS. On day 1, cells were incubated with DMSO (solvent control), 1μM DS or 1μM U18666A. Both compounds were kept at a constant concentration in each buffer throughout the experiment. On day 2, cells were switched to cholesterol-depletion medium A. After a 12-hour incubation, cells were switched to medium B containing 5% LPDS, 10 μM compactin, and 200 μM mevalonate. After 6 h, cells were imaged using filipin (Scale bar: 20 μm.).
(G) Measurement of sterol synthesis. On day 0, CHO-K1 cells were set up in 6 cm plates at a density of 4 × 105 cells/plate. On day 1, cells were incubated with DMSO (solvent control), 1 μM DS or 1 μM compactin (CPN). Compounds were kept at a constant concentration in each buffer throughout the experiment. On day 2, cells were switched to medium A supplemented with 5% FCS and 3 μCi/ml [14C] acetate as well as cold acetate at a final concentration of 0.1 mM. After 4 hours, cells were harvested for the metabolic labeling study and immunoblotting of HMG-CoA reductase, and Histone H3. Data are mean ± S.D. (n=3). P values (two-sided) were calculated by Student’s t tests using GraphPad Prism v.9; *, P<0.05.
See also Figure S6.
To explore the effect of DS55980254 on cholesterol synthesis, we measured the de novo cellular sterol amount after a 4-hour incubation with radiolabeled acetate. Our results showed that DS55980254 increases sterol levels by 10% compared to the solvent control; in contrast, compactin, which inhibits the activity of HMG-CoA reductase (HMGCR, the rate-limiting enzyme in cholesterol synthesis), reduces sterol levels by 10% compared to the solvent control (Figure 7G). Moreover, compactin induces a considerable overexpression of HMGCR, whereas DS55980254 does not (Figure 7G).
Our results suggest that abolishing the enzymatic activity of PSS1 leads to lower cholesterol levels in the ER. Furthermore, DS55980254 can trigger SREBP-2 cleavage and promote LDL uptake in cells while only slightly increasing cholesterol synthesis, but does not induce a notable overexpression of HMGCR. Since selective inhibition of PSS1 does not induce ER stress in the presence of PSS232, it is worthwhile to test whether PSS1-specific inhibitors can reduce blood cholesterol levels.
Discussion
The generation of PS in mammalian cells exclusively occurs through PS synthases. The phospholipid substrate gains entry to the enzyme through a gate formed between the luminal halves Notably, three glycine residues (Gly171, Gly175, and Gly328) that face the gate reduce steric hindrance, facilitating the unimpeded access of the phospholipid (Figure 3E). l-serine, the other substrate, can be incorporated into the membrane from the cytosol by SERINC proteins. A previous study demonstrated that the overexpression of mammalian SERINC1, SERINC2, or SERINC5 in COS cells increases the enzymatic activity of PSS126. Prior to the reaction, l-serine can be sequestered by the calcium in the core. This interaction ensures that l-serine attains an optimal proximity to the catalytic histidine, facilitating the deprotonation of its hydroxyl group. The activated serine is poised to initiate an attack on the ester bond linking the phosphate group and the polar head of PC or PE. Since PS appears to be more abundant in the cytoplasmic leaflet than in the luminal leaflet in the ER41, the newly synthesized PS must be flipped to accumulate on the cytosolic leaflet. It is highly possible that without PS in the cytosolic sites, PSS1 represents an active state. In turn, excessive PS molecules in the cytosolic leaflet become negative regulators to reduce the enzymatic activity and preventing the overproduction of PS.
If the structure of PSS1WT represents the inactive state, why could we detect its activity through the LC-MS/MS analysis? It is conceivable that POPC-rich liposomes serve as mimics for PS-deficient membranes, prompting PSS1 to release its endogenously bound PS into the lipid environment. This release converts PSS1 from an inactive to an active state. Since the orientation of PSS1 reconstituted into the liposome can be random, as PS is synthesized, the PS-binding site on both leaflets becomes occupied, leading to a decrease in the activity of PSS1WT. In contrast, the PS-binding sites were disrupted in PSS1P269S, preventing its activity from being regulated by the newly synthesized PS, resulting in a higher activity compared to PSS1WT (Figure 1B).
Intriguingly, the mechanism of PSS1 bears resemblance to the previously proposed mechanisms of membrane-bound O-acyl transferases (MBOATs)42–48, although the architecture of PSS1 does not align with any known MBOATs and there is no evolutionary correlation between them (Figure S7A). As ER-membrane enzymes, MBOATs catalyze the transfer of an acyl group from an acyl-CoA to protein or lipid substrates49 (Figure S7B). For instance, Porcupine employs a catalytic histidine to trigger the activation of a serine residue on Wnt for its lipidation47. l-serine and Wnt can access these enzymes via their luminal side and a catalytic histidine triggers serine activation by deprotonating its hydroxyl group to facilitate nucleophilic attack on the ester bond of their substrates (thioester bond of acyl-CoA and phosphoester bond of PC or PE). In contrast, the fatty acid donors traverse distinct pathways to reach the catalytic core. Acyl-CoA approaches the catalytic histidine of MBOATs from the cytosol; conversely, PC or PE gains access to the catalytic histidine of PSS1 from the luminal leaflet. Following product formation, most lipidated Wnts are transported to their trafficking carrier, Wntless, through a transit complex50. It remains unclear whether the newly synthesized PS diffuses directly into the luminal leaflet or if PSS1 forms a transient complex with a flippase, which can then shuttle the PS to the cytosolic leaflet.
Given that PS plays a crucial role as a signaling molecule, the synthesis of PS must be tightly regulated by cells. Our structural findings, combined with previous discoveries24, suggest that cells employ three strategies to mediate the appropriate production of PS within the ER membrane. First, the activity of SERINC proteins may control the abundance and therefore the accessibility of l-serine to the reaction center. Second, the newly synthesized PS must be flipped from the luminal leaflet to cytosolic leaflet. PS flippases couple the energy from the hydrolysis of ATP to the transport of PS from the outer to the inner leaflet thereby maintaining an asymmetric distribution of PS in the ER membrane. A previous study showed that ATPase activity of phospholipid-transporting ATPase IA (ATP8A1) is stimulated by PS in vitro51. Third, excessive PS can bind to the cytosolic pockets of PSS1 to regulate its enzymatic activity, inhibiting the production of PS. The concentration of PS in the cytosolic leaflet will be reduced after PS is distributed to other intracellular organelles. Then, the reduced concentration of PS in the membrane will cause the PS in the binding pockets to diffuse into the lipid bilayer, thereby re-activating its PS synthesis.
Limitation of the study
We determined our structures using detergent micelles, and it is conceivable that PSS1 might exhibit distinct conformations within different types of nanodiscs or the native cell membrane. Another interesting question is how PS is flipped after production and whether PSS1 can form a complex with one of the PS flippases. Future in vitro biochemical and cell biological work may uncover this puzzle. Moreover, in vivo studies are needed to determine whether targeting PSS1 could either serve as a statin alternative or be used in conjunction with statins to lower blood cholesterol levels.
STAR★METHODS
RESOURCE AVAILABILITY
Lead contact
Further information and requests for resources and reagents should be directed to and will be fulfilled by the lead contact, Xiaochun Li (xiaochun.li@utsouthwestern.edu).
Materials availability
Plasmids generated in this study and the raw cryo-EM dataset are available upon request from the lead contact.
Data and code availability
The 3D cryo-EM density maps have been deposited in the Electron Microscopy Data Bank (EMDB) under the accession numbers EMD-44178, EMD-44179 and EMD-44180. Atomic coordinates for the atomic model have been deposited in the Protein Data Bank (PDB) under the accession numbers 9B4E, 9B4F and 9B4G.
This paper does not report original code.
Any additional information required to reanalyze the data reported in this paper is available from the lead contact upon request.
EXPERIMENTAL MODEL AND STUDY PARTICIPANT DETAILS
Cell lines
Spodoptera frugiperda Sf9 cells (ATCC) were used to generate baculovirus and maintained in Sf-900 III SFM medium (Gibco) at 27 °C with shaking. HEK 293S GnTI− cells (ATCC) were used for protein expression and maintained in Freestyle 293 expression medium (Gibco) with 2% FCS at 37 °C and 8% CO2 with shaking. Medium A is a 1:1 mixture of Ham’s F-12 medium and DMEM containing 2 mM L-glutamine. Medium B is Dulbecco’s modified Eagle’s medium (DMEM)-low glucose (1000 mg/L). Cholesterol-depletion medium A consists of medium A supplemented with 5% newborn calf lipoprotein-deficient serum (LPDS)52, 10 μM compactin, and 200 μM mevalonate. Cholesterol-depletion medium B consists of medium A supplemented with 5% LPDS, 30 μM compactin, and 200 μM mevalonate. All media except Sf-900 III SFM medium are supplemented with 100 U/ml of penicillin and 100 μg/ml of streptomycin sulfate. CHO-K1 cells (Chinese hamster) were maintained in medium A supplemented with 5% FCS. CHO-K1 cells lacking PSS1 (designated as Ptdss1−/− cells) were generated using CRISPR-Cas9 by the Brown/Goldstein lab27 and grown in medium A containing 10% FCS. SV589j cells, a clonal line of SV40-immortalized human fibroblasts derived from SV589 cells by the Brown/Goldstein lab27, were grown in medium A containing 5% FCS. Buffer B is Hank’s Balanced Salt Solution (HBSS) containing 2% FCS and 5 mM sodium EDTA, pH 8.0. All cells were routinely monitored for mycoplasma contamination.
METHOD DETAILS
Protein expression and purification
The complementary DNA (cDNA) of human PSS1 (Protein ID NP_055569.1) was cloned into pEG-BacMam with a C-terminal Flag-tag. Point mutations were introduced into the coding region of PSS1 by site-directed mutagenesis using the QuikChange II XL Site-Directed Mutagenesis Kit (Agilent Technologies). The coding region of each plasmid was sequenced to ensure the integrity of the construct. The protein was expressed in HEK-293S GnTI− cells using the baculovirus system. To generate baculovirus, 5 μg bacmid was transfected into 1 × 107 sf9 cells. After total 72 h, the P1 virus was harvested. For the P2 virus, 1 ml P1 was added to 50 ml 3 × 106 sf9 cells, and then 50 ml medium was added the next day. After total 72 h, the P2 virus was harvested for protein expression.
Protein purification was performed as previously described29. Briefly, 800 ml of HEK-293S GnTI− cells at a density of 3 × 106 were infected with 50 ml of P2 virus. After 8 hours, 10 mM sodium butyrate was added, and cells were then incubated at 37 °C to express proteins. For the PSS1WT-inhibitor complex, DS55980254 (ProbeChem) was added to the culture at a final concentration of 1 μM upon infection and maintained at a final concentration of 10 μM in all subsequent purification steps. The cells were collected after 48 h of infection and were disrupted by sonication in buffer A (20 mM HEPES, pH 7.5, 150 mM NaCl, 5mM CaCl2) supplemented with 1 mM phenylmethylsulfonyl fluoride and 5 μg/ml leupeptin. After low-speed centrifugation (4,000 r.p.m, rotor: A-4–62, Eppendorf, 4 °C, 5 min), the resulting supernatant was incubated in buffer A with 1% (w/v) lauryl maltose neopentyl glycol (LMNG) (Anatrace) for 1 h at 4 °C. The lysate was then centrifuged at high-speed (18,000 r.p.m, rotor: F21–8x50y, Thermo scientific, 4 °C, 30 min) to remove insoluble components, and the supernatant was loaded onto the Flag M2 affinity resin (Sigma-Aldrich). After two tandem washes, the protein was eluted in buffer A supplemented with 100 μg/ml 3×Flag peptide, 0.01% (w/v) LMNG and concentrated by Amicon Centrifugal Filters (Millipore). The concentrated protein was purified by a Superose 6 Increase 10/300 size-exclusion chromatography column (GE Healthcare) in buffer A and 0.06% (w/v) Digitonin. Finally, the peak fractions were pooled, concentrated, then directly flash frozen in liquid nitrogen and stored at −80 °C. Typically, 800 ml cells can produce 400 μg protein.
EM sample preparation and imaging
3 μl of protein sample at ~8 mg ml−1 (~150 μM) was added to Quantifoil R1.2/1.3 400 mesh Au holey carbon grids (Quantifoil), blotted using a Vitrobot Mark IV (FEI) (15 force, 5 s blotting time and 100% humidity at 22 °C), and frozen in liquid ethane. For the PSS1WT-inhibitor complex, DS55980254 was added into the sample at the final concentration of 150 μM before grid freezing. The grids were imaged in a 300 kV Titan Krios (FEI). For PSS1WT and PSS1WT-inhibitor complex, raw movie stacks were collected with a Gatan K3 Summit direct electron detector (Gatan). Data were collected using SerialEM at 0.834 Å per pixel and a nominal defocus range of −0.8 to −1.8 μm. Images were recorded for 5 s exposures in 50 subframes to give a total dose of roughly 60 electrons per Å2. For PSS1P269S, raw movie stacks were collected with a Falcon 4i camera at 0.738 Å per pixel and a nominal defocus range of −0.8 to −1.8 μm. The exposure time for each micrograph was 4 s with a total dose of ~60 electrons per Å2.
Imaging processing and 3D reconstruction
For PSS1WT and DS55980254-bound PSS1WT, dark subtracted images were first normalized by gain reference (0.834 Å per pixel). Drift correction was performed using MotionCor253 and performed in RELION3.154. The contrast transfer function (CTF) was estimated using CTFFIND455. For PSS1P269S, dark subtracted images were first normalized by gain reference (0.738 Å per pixel). Drift correction and CTF were performed and estimated using CryoSPARC v4.4.156. Auto picking for all datasets was performed with crYOLO-v1.7.6 or 1.9.357 using the general model with the particle threshold of 0.1. Particle extraction was performed in RELION3.1 or CryoSPARC v4.4.1 with a box size of 300 pixels for PSS1WT and DS55980254-bound PSS1WT, and 340 pixels for PSS1P269S. Subsequent two-dimensional classification, multi-class ab initio modeling, heterogenous three-dimensional (3D) refinement, nonuniform refinement, CTF refinement and local refinement of the best class were performed in CryoSPARC v.3.3.0 or v4.4.1 (Figures. S2, S4 and S5). All the masks for the refinement were automatically generated by CryoSPARC during the nonuniform refinement and local refinement.
The initial models were built de novo based on AlphaFold2 predicted structures and then manually adjusted and refined using COOT58. Due to the limited resolution, residues 1–11, 143–151 and 410–473 of PSS1WT, residues 1–11, 142–151 and 410–473 of DS55980254-bound PSS1WT, and residues 1–39, 249–283 and 410–473 of PSS1P269S were not built. Residues 154–162 of PSS1WT, and residues 154, 156–158, 210 and 212 of DS55980254-bound PSS1WT were built as alanine. The models were refined in real space using PHENIX59. For cross-validations, MolProbity60 was used to validate the final model. Local resolutions were estimated using CryoSPARC. Structure figures were generated using COOT58, PyMOL (http://www.pymol.org) and ChimeraX61.
ER isolation
On day 0, both CHO-K1cells and Ptdss1−/− CHO-K1 cells were set up in medium A with 10% FCS at 7 × 105 cells/10 cm dish. On day 3, the cells were washed with phosphate-buffered saline, and then scraped into 15 ml tubes. All further operations were carried out at 4 °C. The cell pellets were collected after centrifugation at 1000 g for 10 min and resuspended in 1 ml of ice-cold Buffer C (50 mM Tris-HCl, pH 7.5, 150 mM NaCl, and protease inhibitors: 50 μg/ml leupeptin, 25 μg/ml pepstatin A, 10 μg/ml aprotinin, 25 μg/ml phenylmethyl sulfonylchloride) with 15% (w/v) sucrose. Cells were then disrupted by passaging them 10 times through a ball-bearing homogenizer with a 10 μm clearance. The homogenized cells (Fraction A, see fractionation scheme in Figure 6B) were centrifuged at 7000 g for 15 min to yield a pellet and a supernatant. The supernatant was diluted to a total volume of 3 ml using Buffer C containing 15% sucrose. A discontinuous sucrose gradient was generated in an ULTRA CLEAR tube (Beckman Instruments) by overlaying the following sucrose solutions all in Buffer C: 2 ml 45%, 4 ml 30%, 3 ml of the diluted supernatant in 15% sucrose, and 1 ml 7.5%. The gradient was centrifuged at 100,000 g in a TH641 rotor (Thermo Scientific) for 1 h and allowed to slow down without braking, after which two bands of membranes were clearly visible and collected using a Pasteur pipet. The sucrose concentration in the light membrane ranged from 14%–18%, and 34%–38% in the heavy membrane fraction (Fraction C, about 800 μl). A discontinuous iodixanol gradient was generated by underlaying, in succession, 2.25 ml of 19%, 21%, 23%, and 25% (v/v) iodixanol, all in Buffer C. This discontinuous gradient was allowed to stand for 1–2 h, allowing for diffusion across the interfaces to form a continuous gradient. The heavy membrane fraction from above was loaded at the bottom of this gradient, which was then centrifuged for 2 h at 110,000 g in a TH641 rotor and allowed to slow down without braking. Fractions (0.8 ml each) were collected from the bottom. Lipids were extracted from ER fraction (Fraction E, in Figure 6B) by adding an equal volume of chloroform/methanol (1:1, v/v). Bottom organic layer was collected in a glass tube after vortex, spun at 2000 rpm (rotor: A-4–62, Eppendorf) for 10 minutes and evaporated under vacuum for overnight.
Cholesterol was further extracted using a modified MTBE method (2:1:1 MTBE:MeOH:PBS v/v) mixture in a 16 × 100 mm screw cap test tube with PTFE-lined caps62. The samples were vortexed and centrifuged to enhance phase separation. The upper layer (organic) was transferred to a separate screw cap vial with a Pasteur Pipette. The extraction process was repeated by first adding an additional 1-ml aliquot of MTBE to the sample and following the procedure described above. Extracts were dried under a gentle stream of nitrogen at 40 °C and reconstituted in methanol. Cholesterol levels were measured with isotope dilution mass spectrometry using a deuterated analog of cholesterol (d7) (Avanti Polar Lipids) added to the sample prior to extraction as a surrogate standard. Cholesterol was measured using high performance liquid chromatography (HPLC, Shimadzu LC30 coupled to a triple quadrupole mass spectrometer (MS, SCIEX API 5000) through an atmospheric pressure chemical ionization source63. Total phospholipid levels were determined by a colorimetric assay64 that measures inorganic phosphate released by Malachite Green Phosphate Assay Kit after acidic digestion. Cholesterol content in the ER is expressed as the mole percentage of total lipids (phospholipids + cholesterol).
Assaying of base-exchange activities by LC-MS/MS
POPC liposomes were prepared by sonication in a buffer containing 50 mM HEPES, pH 7.5 and 5 mM CaCl2. Serine-exchange activity was measured by supplementation of 500 nM PSS1 or PSS1 variants into liposomes and indicated concentration of serine. Assays were performed in a 100 μl volume for 90 min at 37 °C and were terminated by adding 10 μl of 0.5 M EDTA. For the inhibition assay, DS55980254 was preincubated with PSS1 or PSS1 variants at indicated concentration for 10 min at room temperature.
Lipids were extracted from the liposomes using a modified methyl-tert-butyl ether (mTBE) method62. Briefly, 1 ml Milli-Q water, 1 ml of methanol and 2 ml of mTBE were added to glass tubes containing the sample, vortexed, and centrifuged. The organic phase was collected, spiked with 20 μL of a 1:5 diluted Splash Lipidomix standard solution (Avanti lipids, Alabaster, USA), dried under N2 and resuspended in hexane. Lipids were analyzed by LC-MS/MS using a SCIEX QTRAP 6500+ (SCIEX, Framingham, USA) equipped with a Shimadzu LC-30AD (Shimadzu, Columbia, USA) HPLC system and a 150x2.1 mm, 5 μm silica column (Supelco, Bellefonte, USA). Samples were injected at a flow rate of 0.3 ml/min starting at 97.5% solvent A (hexane) a 2.5% solvent B (mTBE). Solvent B was increased to 5% over 3 min and then to 60% over 6 min.
Following this, solvent B was decreased to 0% during 30 s while Solvent C (90:10 (v/v) isopropanol-water) was set at 20% and increased to 40% over the next 11 min. Solvent C is increased to 44% over 6 min and then to 60% over 50 s. The system was held at 60% solvent C for 1 min prior to re-equilibration at 2.5% of solvent B for 5 min at a 1.2 ml/min flow rate. Solvent D (95:5 (v/v) acetonitrile-water with 10 mM Ammonium acetate) was infused post-column at 0.03 ml/min. The column oven temperature was maintained at 25 °C. Data was acquired in negative ionization mode using multiple reaction monitoring (MRM) and analyzed using MultiQuant software (SCIEX). The identified lipid species were normalized to its corresponding internal standard.
SREBP-2 processing and western blot
The details of cell plating and treatment are described in figure legends. Cells were treated with 25 μg/ml of N-acetyl-Leu-Leu-norLeucinal (ALLN) 1 h prior to harvesting for western blot of SREBP-2 and Histone H3 as described below.
Cells were lysed by the addition of PBS containing 0.3% (w/v) SDS and 150 U/ml of Benzonase Nuclease. Aliquots were mixed with 4x loading dye, after which 20 μl protein were applied to 4–12% or 4–20% gradient gels. After electrophoresis, the proteins were transferred to nitrocellulose filters, which were then incubated overnight at 4 °C with one of the following antibodies dissolved in 5% (w/v) nonfat milk: 5 μg/ml anti-SREBP-2, anti-Flag (1:1500), anti-Histone H3 (1:5000), anti-TOM20 (1:1000), anti-CREB (1:1000), anti-EEA1 (1:1000), anti-GM130 (1:1000), anti-Na+/K+ ATPase (1:1000), anti-Calnexin (1:1000), anti-NPC1 (1:1000), anti-ABCD3 (1:1000) and 5 μg/ml anti-HMGCR. Bound antibodies were visualized by chemiluminescence using Supersignal West Pico Chemiluminescent Substrate (Thermo Scientific) after incubation for 0.5 h at room temperature with goat anti-rabbit IgG conjugated to horseradish peroxidase (1:2500) or horse anti-mouse IgG conjugated to horseradish peroxidase (1:2500). The images were scanned using an Odyssey FC Imager (Dual-Mode Imaging System; 2 min integration time) and analyzed using Image Studio lite v5.2.
Quantitative real-time PCR
Total RNA was isolated from cells using the RNeasy Plus Universal Mini Kit (QIAGEN) according to the manufacturer’s instructions and then subjected to reverse transcription. Three replicate samples were subjected to real-time PCR analysis as described previously65. The real time PCR contained, in a final volume of 20 μl, 20 ng of reverse transcribed total RNA, 167 nM forward and reverse primers, and 10 μl of 2× SYBR Green PCR Master Mix (PerkinElmer Life Sciences). PCR reactions were carried out in 384-well plates using the ViiA7 Real-Time PCR System (Applied Biosystems). All reactions were done in triplicate. mRNA for acidic ribosomal phosphoprotein 36B4 served as an invariant control. Relative amounts of mRNAs were calculated using the comparative CT method. The primer sequences used for PCR are listed here:
36B4_forward: TCTGCACTCACGCTTCCTAGAG;
36B4_reverse: GGCAACAGTCGGGTAACCAA;
LDLR_forward: AGACACATGCGACAGGAATGAG;
LDLR_reverse: GACCCACTTGCTGGCGATA.
Surface LDLR expression
Anti-LDLR, a mouse monoclonal antibody that binds to human LDLR66, is produced by the Brown/Goldstein lab. Phycoerythrin-conjugated anti-LDLR (PE-anti-LDLR) was generated using PE/R-Phycoerythrin Conjugation Kit (Abcam) according to the manufacturer’s instructions. After plating and incubation as described in the Figure legends, cells were released from monolayers by incubation with buffer B, after which ~1 × 106 cells were suspended in buffer B and centrifuged for 5 min at 1000 rpm (rotor: TX-750, Thermo scientific). Each cell pellet was resuspended in 0.1 ml buffer B containing 2 μg/ml PE-anti-LDLR. After incubation for 15 min at room temperature, cells were washed once with buffer B and then resuspended in buffer B to a concentration of ~1 × 106 cells/ml for flow cytometry.
BODIPY FL-LDL uptake
After plating and 2-hour incubation with BODIPY FL-LDL as described in the Figure legends, cells were released from monolayers by treatment with buffer B and then then centrifuged at 4 °C for 5 min at 1000 rpm (rotor: TX-750, Thermo scientific). The pelleted cells were washed once by 1 ml buffer B, resuspended in buffer B, and subjected to flow cytometry.
Metabolic labeling study
After plating and incubation as described in the figure legends, cells were lysed by 0.1 N NaOH and then supplemented with ethanol and KOH at final concentrations of 40% and 15%, respectively. The samples were saponified at 75 °C for 60 min. After saponification, the total sterols were extracted by petroleum ether. The organic phase was evaporated until completely dry and then solubilized in Complete Counting Cocktail for the determination of de novo cholesterol synthesis via liquid scintillation counting. To establish background radioactivity, the cells were chilled to 4 °C, re-fed with radiolabeled medium and immediately washed, saponified and extracted. An aliquot of each sample was taken for protein concentration determination using the BCA protein assay reagent (Thermo Scientific).
Flow cytometry
Resuspended cells were adjusted to a concentration of ~1 × 106 cells/ml with buffer B in a single U-bottom tube and loaded into a MACSQuant Analyzer 16 flow cytometer. Dead cells and cell doublets were excluded by gating so that >20,000 cells were available for analysis in each condition. The data were analyzed using MACSQuantify v2.13.3 software.
Visualization of cholesterol distribution by fluorescence microscopy
After plating and incubation as described in the Figure legends, cells were fixed in 3.7% (w/v) formaldehyde in PBS for 15 min at room temperature. After washing with PBS, cells were labelled with 25 μg/ml filipin in PBS for 30 min at room temperature, then washed 3 times in PBS, and mounted in Shandon Immu-mount (Thermo Scientific). Fluorescence images were acquired using a Plan-Apochromat ×63/1.4 oil DIC objective (Zeiss, Oberkochen, Germany), a Zeiss LSM 800 microscope (Zeiss), and ZEN imaging software (Zeiss). Images of different conditions were captured with the same parameters.
Construction of protein models for MD simulations
The structure of PSS1WT from cryo-EM was modelled in a lipid bilayer. The structure contains the homodimer of PSS1. Within each homodimer, two product PS molecules are in the allosteric binding site. Each monomer also binds two 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) molecules and each active site binds a Ca2+ ion. Two dipalmitoyl phosphatidylinositol (PI) molecules are built at the contact interface of the two monomers. The structural model of PSS1 without PS-binding was constructed by removing the coordinate of PS at the allosteric site. Hydrogen atoms were added using PROPKA367. A disulfide bond was constructed between Cys153 and Cys213. The transmembrane domain was then placed into an intact lipid bilayer by replacing overlapping lipid molecules using the CHARMM-GUI68. The lipid bilayer consists of upper (cytosol) and lower (extracellular space) leaflets, both of which contain 100% POPC. The protein-membrane system was next solvated in 22.5 Å thickness of water with 0.15 M KCl, resulting in 127 K+ ions, 125 Cl− ions, and 48282 water molecules for PSS1 with PS and 130 K+ ions, 132 Cl− ions, and 49889 water molecules for PSS1 without PS. The PSS1 model with PS-binding involves 233990 atoms in a tetragonal box of dimension 150.1 Å × 150.1 Å × 111.6 Å. The PSS1 without PS-binding involves 239487 atoms in a tetragonal box of dimension 150.1 Å × 150.1 Å × 114.0 Å.
Molecular dynamics simulations
MD simulations were conducted using AMBER22 software suit69. We used CHARMM-GUI to generate required input files70,71. The AMBER ff14SB72 force field was used for the protein part, and the generalized AMBER force field73,74 was used for the bound molecules PS, PC, and PI with AM1-BCC model to determine the atomic charges75. The lipid bilayer and solvent water were modeled with Lipid2176 and TIP3P force fields, respectively. The SHAKE algorithm77 was used to constrain all hydrogen-containing bonds. The solvated protein-membrane complex was first optimized using the steepest descent method for 2500 steps followed by the conjugate gradient method for another 2500 steps. A constraint of 250 kcal·mol−1·rad−2 was applied to the dihedral angle defined by C1-C3-C2-O2 and the double bond of each POPC molecule during the geometry optimization. Positional restraints relative to the initial structure were defined for the protein, bound molecules (except Ca2+), and all phosphorus atoms of the POPC molecule. The force constants were 10.0 kcal·mol−1·Å−2 for proteins and bound molecules and 2.5 kcal·mol−1· Å−2 for phosphorus. Next, the system was equilibrated with molecular dynamics at 300 K. The equilibration followed the default of CHARMM-GUI. The constraints gradually decreased to none during the equilibration. The system was heated from 0 to 300 K and equilibrated for 250 ps under constant volume at 300 K, and further equilibrated at 300 K and 1 atm for 1625 ns. The whole equilibration cost 1875 ps. After equilibration, the production runs for 200 ns and output the trajectories every 100 ps, resulting in a total of 2000 snapshots for each production run. The production was performed with a time step of 2 fs. The Langevin thermostat78 and Berendsen barostat79 were used to control the temperature and pressure, respectively. Three parallel simulations were conducted for both PSS1 with and without PS. The input topology and structure file, structural restraint files, and parameter files for each MD step were included in the supporting data. All MD trajectories with membrane and water stripped were also included. All MD-related files can be found at https://doi.org/10.5281/zenodo.11404258.
Trajectory analyses
All atomic distance and conformational fluctuation analyses were conducted using the cpptraj utility of AMBER80. To quantify the binding conformation at the active site, we calculated the distance between key residues and the Ca2+. For Glu197, Glu200, Glu212, and Asp221, we calculated the distance between their carboxylate group (represented by the middle point of the two sidechain carboxylic oxygen atoms) and the Ca2+. For residues 197 to 221, we also calculated the distance between each Cα to Ca2+. We calculated the distance between the Cα atoms of Phe168 and Trp321 to evaluate the opening of the catalytic core. To investigate the conformational dynamics of PSS1, we calculated the root-mean-square fluctuation (RMSF) for each residue. We included the backbone heavy atoms in the calculation. Conformations of each monomer in each trajectory were first overlayed to generate an average structure. This average structure was then used in the corresponding trajectory to calculate the RMSF for each residue.
The binding free energy analyses were based on the MM-PBSA method enabled by the MMPBSA.py utility of AMBER81. We used a uniform membrane dielectric constant in a slab-like implicit membrane to model the membrane environment35. The PSS1 dimer with Ca2+ ions in the active sites were treated as the receptor. Ligands except the ones of interest, if any, were treated as solvent together with the double-layer membrane, K+, Cl−, and water molecules. To reduce the computational cost, we evenly sampled 50 out of 2000 snapshots in each trajectory. Other specifics were included in an example input file at https://doi.org/10.5281/zenodo.11404258.
Reproducibility
All enzymatic activity, metabolic and imaging experiments were repeated at least two times on different days. Similar results were obtained.
QUANTIFICATION AND STATISTICAL ANALYSIS
Statistical analyses of data were performed using GraphPad Prism 9. Quantification methods and tools used are described in each relevant section of the METHOD DETAILS or figure legends.
Supplementary Material
(A) Sequence alignment of human PSS1 and PSS2 with mouse, chicken, xenopus, and zebrafish PSS1. The transmembrane helices and the residue numbers of human PSS1 are indicated above the protein sequence. The specific residues necessary for calcium binding are indicated by circles, the catalytic histidine is indicated by an asterisk and the cysteines for disulfate bond are indicated by purple squares. The residues necessary for inhibitor binding are indicated by dark blue squares.
(B) Purification of human PSS1WT and its variants for cryo-EM study and activity assays. Representative Superose 6 increase 10/30 gel-filtration chromatogram of PSS1WT, PSS1P269S, PSS1H172A, PSS1F168A, PSS1F168W, PSS1G171W, PSS1G175W and PSS1F313M/S320V in buffer containing 20 mM HEPES pH 7.5, 150 mM NaCl, 5 mM CaCl2 and 0.06% Digitonin. The peak fractions used in this study are indicated by red arrows.
(A) Representative cryo-EM image of PSS1WT. Scale bar, 40 nm.
(B) Selected 2D class averages of PSS1WT.
(C) Summary of cryo-EM data processing procedures. FSC curves between two half maps are shown. The mask, which was generated by CryoSPARC for the final refinement, is shown in gray.
(D) Angular distribution of PSS1WT particles in the final round of 3D refinement using CryoSPARC.
(E) Local resolution of cryo-EM map. Maps are colored according to local resolution, estimated using CryoSPARC.
(F) Cryo-EM map of the transmembrane helices.
(A) The binding free energy of PE and PC to PSS1WT monomer with and without PS from three MD simulations. This is also the case for (B) and (D). Unit: kcal/mol.
(B) The binding free energy of PS to PSS1WT monomer from three MD simulations. Unit: kcal/mol.
(C) The representative curves showing the distance between the side chain center of the corresponding residue and calcium ions in the structure of PSS1 with (blue) or without PS (green).
(D) The binding free energy of serine to PSS1P269S monomer from three MD simulations. Unit: kcal/mol.
(E) Distance between Cα atoms of Phe168 and Trp321 in the PSS1WT two monomers with (blue) or without PS (green).
(A) Summary of cryo-EM data processing procedures. The mask, which was generated by CryoSPARC for the final refinement, is shown in gray.
(B) Angular distribution of PSS1P269S particles in the final round of 3D refinement using CryoSPARC. FSC curves between two half maps are shown.
(C) Local resolution of cryo-EM map. Maps are colored according to local resolution, estimated using CryoSPARC.
(D) Cryo-EM map of the transmembrane helices.
(E) The extra density at the lipid entrance in the cryo-EM map of PSS1P269S.
(F) The extra lipid-like density in the cryo-EM map of PSS1P269S.
(G) Comparison of the cryo-EM maps of PSS1P269S and PSS1WT indicates the flexibility of PH4 in PSS1P269S. The cryo-EM map of PSS1WT has been filtered to 3.3-Å.
(H) Comparison of calcium binding sites of PSS1P269S and AF model (gray). The calcium in the structure of PSS1P269S is shown as a green sphere. Dashed line represents hydrophilic interactions.
(A) Summary of cryo-EM data processing procedures. FSC curves between two half maps are shown. The mask, which was generated by CryoSPARC for the final refinement, is shown in gray.
(B) Angular distribution of DS55980254 bound PSS1WT particles in the final round of 3D refinement using CryoSPARC.
(C) Local resolution of cryo-EM map. Maps are colored according to local resolution, estimated using CryoSPARC.
(D) Cryo-EM map of the transmembrane helices.
(A) Measurement of LDLRs by flow cytometry in SV589j cells incubated with LDL. On day 0, the indicated cells were set up in medium A with 5% FCS. After plating for 6 hours, cells were supplemented with DMSO (solvent control) or 1 μM DS, which was kept at a constant concentration in each buffer throughout the experiment. On day 1, cells were switched to cholesterol-depletion medium B. After 16 h, cells then received the above cholesterol-depletion medium containing 50 μg protein/ml of LDL. After 24 h, the cells were harvested by incubation with EDTA, washed, incubated with PE-anti-LDLR, and subjected to flow cytometry (METHOD DETAILS).
(B) Measurement of BODIPY FL-LDL uptake by flow cytometry in SV589j cells. On day 0, the indicated cells were set up in medium A with 5% FCS. After plating for 6 hours, cells were supplemented with DMSO (solvent control) or 1 μM DS, which was kept at a constant concentration in each buffer throughout the experiment. On day 1, cells were switched to cholesterol-depletion medium B. After 16 h, cells then received the above cholesterol-depletion medium containing 50 μg protein/ml of LDL. After 24 h, the cells were treated by 5 μg protein/ml of BODIPY FL-LDL for 2 hours, then the cells were harvested by incubation with EDTA, washed and subjected to flow cytometry (METHOD DETAILS).
(A) Phylogenetic tree of PSSs and MBOATs. Human, mouse, frog, zebrafish, Arabidopsis, yeast and bacteria proteins were used for the comparison. Numbers at the nodes indicates the bootstrap values on neighbor joining analysis. The calculation was performed in MEGA X.
(B) Working mechanisms of PORCN and ACAT1. The catalytic histidine activates the hydroxyl group of the serine of Wnt (PORCN) or cholesterol (ACAT1), which would attack the thioester bond of acyl-CoA substrate, to form the product.
KEY RESOURCES TABLE
| REAGENT or RESOURCE | SOURCE | IDENTIFIER |
|---|---|---|
| Antibodies | ||
| Rabbit polyclonal Anti-Flag antibody | Sigma-Aldrich | Cat# F7425 |
| Rabbit monoclonal Anti-SREBP-2 | Sigma-Aldrich | Cat# MABS1988 |
| HRP Rabbit polyclonal to Histone H3 antibody | Abcam | Cat# ab21054 |
| Rabbit Anti-TOM20 antibody | Protein Tech | Cat# 11802-1-AP |
| Mouse Anti-CREB antibody | Thermo Fisher | Cat# 35-0900 |
| Mouse Anti-EEA1 antibody | Sigma-Aldrich | Cat# E7659 |
| Mouse Anti-GM130 antibody | BD Biosciences | Cat# 610822 |
| Rabbit Anti-Na+/K+ ATPase antibody | Abcam | Cat# ab76020 |
| Rabbit Anti-Calnexin antibody | Bio-techne | Cat# NB100-1965 |
| Mouse Anti-NPC1 | Abcam | Cat# 134113 |
| Rabbit Anti-ABCD3 | Affinity Biosciences | Cat# DF8315 |
| Mouse Anti-HMGCR | Sigma-Aldrich | Cat# MABS1233 |
| Anti-Mouse IgG, HRP-linked antibody | Cell Signaling Technology | Cat# 7076S |
| Anti-Rabbit IgG, HRP-linked antibody | Cell Signaling Technology | Cat# 7074S |
| Bacterial and virus strains | ||
| E. coli DH5α Competent Cells | GoldBio | Cat# CC-101-TR |
| E. coli DH10Bac Competent Cells | Thermo Fisher | Cat# 10361012 |
| Chemicals, peptides, and recombinant proteins | ||
| LB broth | Fisher Scientific | Cat# BP1426-500 |
| Sf-900 III SFM medium | Gibco | Cat# 12658027 |
| Freestyle 293 expression medium | Gibco | Cat# 12338018 |
| DMEM/F12 50/50 Mix medium | Corning | Cat# 10-090-CV |
| DMEM-low glucose (1000 mg/l) | Sigma-Aldrich | Cat# D6046 |
| Fetal Bovine Serum (FBS) | GeminiBio | Cat# 100-500 |
| Filipin III from Streptomyces filipinensis | Sigma-Aldrich | Cat# F4767-1 MG |
| Newborn calf lipoprotein-deficient serum (LPDS, d<1.215 g/ml) | Goldstein et al.52 | N/A |
| Penicillin-Streptomycin solution (100x) | Corning | Cat# 30-002-CI |
| HEPES | Gibco | Cat# 11344041 |
| NaCl | Fisher Scientific | Cat# S271 |
| Sodium butyrate | Millipore-Sigma | Cat# 303410 |
| Phenylmethylsulfonyl fluoride (PMSF) | GoldBio | Cat# P-470-25 |
| Leupeptin | Biosynth | Cat# ILP-4041 |
| Lauryl maltose neopentyl glycol (LMNG) | Anatrace | Cat# NG310 |
| Digitonin | Thermo Scientific | Cat# 407560050 |
| Anti-Flag M2 resin | Millipore-Sigma | Cat# A2220 |
| 3x Flag peptide | ApexBio | Cat# A6001 |
| 14C-Acetate | American Radiolabeled Chemicals | Cat# ARC 0173A |
| Petroleum Ether | Fisher Chemical | Cat# E139-1 |
| DS55980254 | ProbeChem | Cat# PC-49548 |
| Cellfectin II reagent | Gibco | Cat# 10362-100 |
| Phosphate buffered saline with Tween 20 (PBST) | Millipore-Sigma | Cat# 08057-100TAB-F |
| Benzonase nuclease | Millipore-Sigma | Cat# 70746-3 |
| Mevalonate | Sigma-Aldrich | Cat# M4667 |
| Compactin | Sigma-Aldrich | Cat# M2147 |
| Dimethyl Sulfoxide (DMSO) | Sigma-Aldrich | Cat# D8418 |
| EDTA, sodium, pH 8.0 | Thermo Fisher | Cat# 15575020 |
| Hydroxypropyl-β-cyclodextrin | Cyclodextrin Technologies Development, Inc. | Cat# THPB-P |
| Hank’s Balanced Salt Solution (HBSS) | Gibco | Cat# 14025-092 |
| LDL, human (d1.019-1.063 g/ml) | Goldstein et al.52 | N/A |
| Lipoprotein, low density from human plasma | Sigma-Aldrich | Cat# SAE0053-10MG |
| BODIPY FL-LDL | Invitrogen | Cat# L3483 |
| PE/R-Phycoerythrin Conjugation Kit | Abcam | Cat# ab102918 |
| N-Acetyl-Leu-Leu-norLeucinal (ALLN) | Calbiochem | Cat# 208719 |
| 1-palmitoyl-2-oleoyl-glycero-3-phosphocholine (POPC) | Avanti Polar Lipids | Cat# 850457C-25mg |
| Phosphate Buffered Saline (PBS) | Corning | Cat# 21-031-CV |
| QuikChange II XL Site-Directed Mutagenesis Kit | Agilent | Cat# 200522 |
| Sodium Dodecyl Sulfate (SDS) | Sigma-Aldrich | Cat# 71736 |
| Bolt™ Bis-Tris Plus Mini Protein Gels, 4–12%, 1.0 mm, WedgeWell™ format | Invitrogen | Cat# NW04127B0X |
| Novex™ Tris-Glycine Mini Protein Gels, 4–20%, 1.0 mm, WedgeWell™ format | Invitrogen | Cat# XP04205B0X |
| U18666A | Sigma-Aldrich | Cat# U3633 |
| Hexane | Sigma-Aldrich | Cat# 293253 |
| Acetonitrile | Sigma-Aldrich | Cat# 34998 |
| Ammonium Acetate | Sigma-Aldrich | Cat# 421311 |
| Splash Lipidomix standard solution | Avanti Polar Lipids | Cat# 330707 |
| Critical commercial assays | ||
| Malachite Green Phosphate Assay Kit | Millipore-Sigma | Cat# MAK307-1KT |
| Deposited data | ||
| Cryo-EM structure of PSS1WT | This paper | PDB: 9B4E; EMDB: EMD-44178 |
| Cryo-EM structure of PSS1P269S | This paper | PDB: 9B4F; EMDB: EMD-44179 |
| Cryo-EM structure of DS55980254-bound PSS1WT | This paper | PDB: 9B4G; EMDB: EMD-44180 |
| Experimental models: Cell lines | ||
| Sf9 insect cell | ATCC | Cat# CRL-1711; RRID: CVCL_0549 |
| HEK-293S GnTI− cell | ATCC | Cat# CRL-3022; RRID: CVCL_A785 |
| CHO-K1 cell | ATCC | Cat# CCL-61; RRID: CVCL_0214 |
| Ptdss1−/− CHO-K1 cell | Trinh et al.27 | N/A |
| SV589j | Trinh et al.27 | N/A |
| Recombinant DNA | ||
| pEG-BacMam-PSS1 | This paper | N/A |
| Software and algorithms | ||
| Image Studio Lite v5.2 | LI-COR | N/A |
| MACSQuantify Software 2.13 | Meltenyi Biotec | N/A |
| MotionCor2 | Zheng et al.53 | https://emcore.ucsf.edu/ucsf-motioncor2 |
| RELION 3.1 | Zivanov et al.54 | https://www3.mrc-lmb.cam.ac.uk/relion/ |
| CTFFIND4 | Rohou and Grigorieff55 | http://grigoriefflab.janelia.org/ctffind4 |
| CryoSPARC v3.3.0 or v4.4.1 | Punjani et al.56 | https://cryosparc.com/ |
| SPHIRE-crYOLO v1.7.6 or 1.9.3 | Wagner et al.57 | http://sphire.mpg.de/ |
| Coot 0.8.8 | Emsley et al.58 | https://www2.mrc-lmb.cam.ac.uk/personal/pemsley/coot/ |
| Phenix 1.17 | Adams et al.59 | http://www.phenix-online.org/ |
| UCSF ChimeraX 1.5 | Pettersen et al61 | https://www.cgl.ucsf.edu/chimerax/ |
| PyMOL 2.3.4 | Schrodinger | https://pymol.org/2/ |
| Prism 9 | GraphPad | https://www.graphpad.com/ |
| PROPKA3 | Olsson et al.67 | https://github.com/jensengroup/propka |
| CHARMM | Jo et al.68 | https://www.charmm.org/ |
| AMBER22 | The Amber Project | https://ambermd.org/AmberMD.php |
| Other | ||
| R1.2/1.3 400 mesh Au holey carbon grids | Quantifoil | Cat# 1210627 |
| Superose 6 Increase 10/300 GL column | Cytiva | Cat# 29091596 |
| Milli-Q IQ7000 Ultrapure Water System | Sigma-Aldrich | N/A |
Highlights:
The catalytic mechanism of PSS1 is akin to the working principle of MBOATs.
Phosphatidylserine and DS55980254 are allosteric negative modulators of PSS1.
PSS1 deficiency lowers ER cholesterol levels, activating the SREBP pathway.
Specific PSS1 inhibitors may promote LDLR expression and LDL uptake in cells.
Acknowledgements
We thank M. Brown and J. Goldstein for their invaluable support and discussion throughout the project. Cryo-EM Data were collected at the UT Southwestern Medical Center Cryo-EM Facility (funded in part by the CPRIT Award RP220582). We thank L. Beatty, L. Esparza and Y. Qin for technical support, G. Liang, F. Lu, J. Horton, S. Rong, M. Schumacher and B. Wang for discussion, N. Heard for figure preparation and E. Debler for editing the manuscript. This work was supported by NIH R35GM146982 (to Z.Y. and Y.J.), NIH UL1TR003163 and 1P30DK127984-01A1 (to J.G.M), NIH P01HL160487, R35GM149533, Welch Foundation (I-1957) and American Heart Association (23EIA1038669) (to X.L.).
Inclusion and diversity
We support inclusive, diverse, and equitable conduct of research.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Declaration of interests
The authors declare no competing interests.
References:
- 1.Leventis PA, and Grinstein S (2010). The distribution and function of phosphatidylserine in cellular membranes. Annu Rev Biophys 39, 407–427. 10.1146/annurev.biophys.093008.131234. [DOI] [PubMed] [Google Scholar]
- 2.Bevers EM, and Williamson PL (2016). Getting to the Outer Leaflet: Physiology of Phosphatidylserine Exposure at the Plasma Membrane. Physiological reviews 96, 605–645. 10.1152/physrev.00020.2015. [DOI] [PubMed] [Google Scholar]
- 3.Vance JE (2018). Historical perspective: phosphatidylserine and phosphatidylethanolamine from the 1800s to the present. Journal of lipid research 59, 923–944. 10.1194/jlr.R084004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Kimura AK, and Kimura T (2021). Phosphatidylserine biosynthesis pathways in lipid homeostasis: Toward resolution of the pending central issue for decades. FASEB J 35, e21177. 10.1096/fj.202001802R. [DOI] [PubMed] [Google Scholar]
- 5.Yeung T, Gilbert GE, Shi J, Silvius J, Kapus A, and Grinstein S (2008). Membrane phosphatidylserine regulates surface charge and protein localization. Science 319, 210–213. 10.1126/science.1152066. [DOI] [PubMed] [Google Scholar]
- 6.Glade MJ, and Smith K (2015). Phosphatidylserine and the human brain. Nutrition 31, 781–786. 10.1016/j.nut.2014.10.014. [DOI] [PubMed] [Google Scholar]
- 7.Kim HY, Huang BX, and Spector AA (2014). Phosphatidylserine in the brain: metabolism and function. Progress in lipid research 56, 1–18. 10.1016/j.plipres.2014.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Copic A, Dieudonne T, and Lenoir G (2023). Phosphatidylserine transport in cell life and death. Current opinion in cell biology 83, 102192. 10.1016/j.ceb.2023.102192. [DOI] [PubMed] [Google Scholar]
- 9.Sakuragi T, and Nagata S (2023). Regulation of phospholipid distribution in the lipid bilayer by flippases and scramblases. Nature reviews. Molecular cell biology 24, 576–596. 10.1038/s41580-023-00604-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Yasen A, Herrera R, Rosbe K, Lien K, and Tugizov SM (2018). HIV internalization into oral and genital epithelial cells by endocytosis and macropinocytosis leads to viral sequestration in the vesicles. Virology 515, 92–107. 10.1016/j.virol.2017.12.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Shimojima M, Takada A, Ebihara H, Neumann G, Fujioka K, Irimura T, Jones S, Feldmann H, and Kawaoka Y (2006). Tyro3 family-mediated cell entry of Ebola and Marburg viruses. J Virol 80, 10109–10116. 10.1128/JVI.01157-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Bhattacharyya S, Zagorska A, Lew ED, Shrestha B, Rothlin CV, Naughton J, Diamond MS, Lemke G, and Young JA (2013). Enveloped viruses disable innate immune responses in dendritic cells by direct activation of TAM receptors. Cell Host Microbe 14, 136–147. 10.1016/j.chom.2013.07.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Wang J, Yu C, Zhuang J, Qi W, Jiang J, Liu X, Zhao W, Cao Y, Wu H, Qi J, and Zhao RC (2022). The role of phosphatidylserine on the membrane in immunity and blood coagulation. Biomark Res 10, 4. 10.1186/s40364-021-00346-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Zwaal RF, Comfurius P, and Bevers EM (2004). Scott syndrome, a bleeding disorder caused by defective scrambling of membrane phospholipids. Biochimica et biophysica acta 1636, 119–128. 10.1016/j.bbalip.2003.07.003. [DOI] [PubMed] [Google Scholar]
- 15.Nishizuka Y (1992). Intracellular signaling by hydrolysis of phospholipids and activation of protein kinase C. Science 258, 607–614. 10.1126/science.1411571. [DOI] [PubMed] [Google Scholar]
- 16.Baudry M, Massicotte G, and Hauge S (1991). Phosphatidylserine increases the affinity of the AMPA/quisqualate receptor in rat brain membranes. Behav Neural Biol 55, 137–140. 10.1016/0163-1047(91)80134-z. [DOI] [PubMed] [Google Scholar]
- 17.Stace CL, and Ktistakis NT (2006). Phosphatidic acid- and phosphatidylserine-binding proteins. Biochimica et biophysica acta 1761, 913–926. 10.1016/j.bbalip.2006.03.006. [DOI] [PubMed] [Google Scholar]
- 18.Vance JE (2015). Phospholipid synthesis and transport in mammalian cells. Traffic 16, 1–18. 10.1111/tra.12230. [DOI] [PubMed] [Google Scholar]
- 19.Tomohiro S, Kawaguti A, Kawabe Y, Kitada S, and Kuge O (2009). Purification and characterization of human phosphatidylserine synthases 1 and 2. The Biochemical journal 418, 421–429. 10.1042/BJ20081597. [DOI] [PubMed] [Google Scholar]
- 20.Saito K, Nishijima M, and Kuge O (1998). Genetic evidence that phosphatidylserine synthase II catalyzes the conversion of phosphatidylethanolamine to phosphatidylserine in Chinese hamster ovary cells. The Journal of biological chemistry 273, 17199–17205. 10.1074/jbc.273.27.17199. [DOI] [PubMed] [Google Scholar]
- 21.Arikketh D, Nelson R, and Vance JE (2008). Defining the importance of phosphatidylserine synthase-1 (PSS1): unexpected viability of PSS1-deficient mice. The Journal of biological chemistry 283, 12888–12897. 10.1074/jbc.M800714200. [DOI] [PubMed] [Google Scholar]
- 22.Sturbois-Balcerzak B, Stone SJ, Sreenivas A, and Vance JE (2001). Structure and expression of the murine phosphatidylserine synthase-1 gene. The Journal of biological chemistry 276, 8205–8212. 10.1074/jbc.M009776200. [DOI] [PubMed] [Google Scholar]
- 23.Bergo MO, Gavino BJ, Steenbergen R, Sturbois B, Parlow AF, Sanan DA, Skarnes WC, Vance JE, and Young SG (2002). Defining the importance of phosphatidylserine synthase 2 in mice. The Journal of biological chemistry 277, 47701–47708. 10.1074/jbc.M207734200. [DOI] [PubMed] [Google Scholar]
- 24.Ohsawa T, Nishijima M, and Kuge O (2004). Functional analysis of Chinese hamster phosphatidylserine synthase 1 through systematic alanine mutagenesis. The Biochemical journal 381, 853–859. 10.1042/BJ20040443. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Miyata N, and Kuge O (2021). Topology of phosphatidylserine synthase 1 in the endoplasmic reticulum membrane. Protein Sci 30, 2346–2353. 10.1002/pro.4182. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Inuzuka M, Hayakawa M, and Ingi T (2005). Serinc, an activity-regulated protein family, incorporates serine into membrane lipid synthesis. The Journal of biological chemistry 280, 35776–35783. 10.1074/jbc.M505712200. [DOI] [PubMed] [Google Scholar]
- 27.Trinh MN, Brown MS, Goldstein JL, Han J, Vale G, McDonald JG, Seemann J, Mendell JT, and Lu F (2020). Last step in the path of LDL cholesterol from lysosome to plasma membrane to ER is governed by phosphatidylserine. Proceedings of the National Academy of Sciences of the United States of America 117, 18521–18529. 10.1073/pnas.2010682117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Trinh MN, Brown MS, Seemann J, Vale G, McDonald JG, Goldstein JL, and Lu F (2022). Interplay between Asters/GRAMD1s and phosphatidylserine in intermembrane transport of LDL cholesterol. Proceedings of the National Academy of Sciences of the United States of America 119. 10.1073/pnas.2120411119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Doolittle GM, and Chang TY (1982). Solubilization, partial purification, and reconstitution in phosphatidylcholine-cholesterol liposomes of acyl-CoA:cholesterol acyltransferase. Biochemistry 21, 674–679. 10.1021/bi00533a014. [DOI] [PubMed] [Google Scholar]
- 30.Sousa SB, Jenkins D, Chanudet E, Tasseva G, Ishida M, Anderson G, Docker J, Ryten M, Sa J, Saraiva JM, et al. (2014). Gain-of-function mutations in the phosphatidylserine synthase 1 (PTDSS1) gene cause Lenz-Majewski syndrome. Nature genetics 46, 70–76. 10.1038/ng.2829. [DOI] [PubMed] [Google Scholar]
- 31.Sugahara S, Ishino Y, Sawada K, Iwata T, Shimanaka Y, Aoki J, Arai H, and Kono N (2023). Disease-related PSS1 mutant impedes the formation and function of osteoclasts. Journal of lipid research 64, 100443. 10.1016/j.jlr.2023.100443. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Yoshihama Y, Namiki H, Kato T, Shimazaki N, Takaishi S, Kadoshima-Yamaoka K, Yukinaga H, Maeda N, Shibutani T, Fujimoto K, et al. (2022). Potent and Selective PTDSS1 Inhibitors Induce Collateral Lethality in Cancers with PTDSS2 Deletion. Cancer research 82, 4031–4043. 10.1158/0008-5472.CAN-22-1006. [DOI] [PubMed] [Google Scholar]
- 33.Omi J, Kato T, Yoshihama Y, Sawada K, Kono N, and Aoki J (2024). Phosphatidylserine synthesis controls oncogenic B cell receptor signaling in B cell lymphoma. The Journal of cell biology 223. 10.1083/jcb.202212074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Holm L, and Rosenstrom P (2010). Dali server: conservation mapping in 3D. Nucleic acids research 38, W545–549. 10.1093/nar/gkq366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Botello-Smith WM, Liu X, Cai Q, Li Z, Zhao H, and Luo R (2013). Numerical Poisson-Boltzmann Model for Continuum Membrane Systems. Chem Phys Lett 555, 274–281. 10.1016/j.cplett.2012.10.081. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O,Tunyasuvunakool K, Bates R, Zidek A, Potapenko A, et al. (2021). Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589. 10.1038/s41586-021-03819-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Sandhu J, Li S, Fairall L, Pfisterer SG, Gurnett JE, Xiao X, Weston TA, Vashi D, Ferrari A, Orozco JL, et al. (2018). Aster Proteins Facilitate Nonvesicular Plasma Membrane to ER Cholesterol Transport in Mammalian Cells. Cell 175, 514–529 e520. 10.1016/j.cell.2018.08.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Naito T, Ercan B, Krshnan L, Triebl A, Koh DHZ, Wei FY, Tomizawa K, Torta FT, Wenk MR, and Saheki Y (2019). Movement of accessible plasma membrane cholesterol by the GRAMD1 lipid transfer protein complex. eLife 8. 10.7554/eLife.51401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Radhakrishnan A, Goldstein JL, McDonald JG, and Brown MS (2008). Switch-like control of SREBP-2 transport triggered by small changes in ER cholesterol: a delicate balance. Cell metabolism 8, 512–521. 10.1016/j.cmet.2008.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Lu F, Liang Q, Abi-Mosleh L, Das A, De Brabander JK, Goldstein JL, and Brown MS (2015). Identification of NPC1 as the target of U18666A, an inhibitor of lysosomal cholesterol export and Ebola infection. eLife 4. 10.7554/eLife.12177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Tsuji T, Cheng J, Tatematsu T, Ebata A, Kamikawa H, Fujita A, Gyobu S, Segawa K, Arai H, Taguchi T, et al. (2019). Predominant localization of phosphatidylserine at the cytoplasmic leaflet of the ER, and its TMEM16K-dependent redistribution. Proceedings of the National Academy of Sciences of the United States of America 116, 13368–13373. 10.1073/pnas.1822025116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Long T, Sun Y, Hassan A, Qi X, and Li X (2020). Structure of nevanimibe-bound tetrameric human ACAT1. Nature 581, 339–343. 10.1038/s41586-020-2295-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Sui X, Wang K, Gluchowski NL, Elliott SD, Liao M, Walther TC, and Farese RV Jr. (2020). Structure and catalytic mechanism of a human triacylglycerol-synthesis enzyme. Nature 581, 323–328. 10.1038/s41586-020-2289-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Wang L, Qian H, Nian Y, Han Y, Ren Z, Zhang H, Hu L, Prasad BVV, Laganowsky A, Yan N, and Zhou M (2020). Structure and mechanism of human diacylglycerol O-acyltransferase 1. Nature 581, 329–332. 10.1038/s41586-020-2280-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Qian H, Zhao X, Yan R, Yao X, Gao S, Sun X, Du X, Yang H, Wong CCL, and Yan N (2020). Structural basis for catalysis and substrate specificity of human ACAT1. Nature 581, 333–338. 10.1038/s41586-020-2290-0. [DOI] [PubMed] [Google Scholar]
- 46.Jiang Y, Benz TL, and Long SB (2021). Substrate and product complexes reveal mechanisms of Hedgehog acylation by HHAT. Science 372, 1215–1219. 10.1126/science.abg4998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Liu Y, Qi X, Donnelly L, Elghobashi-Meinhardt N, Long T, Zhou RW, Sun Y, Wang B, and Li X (2022). Mechanisms and inhibition of Porcupine-mediated Wnt acylation. Nature 607, 816–822. 10.1038/s41586-022-04952-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Ma D, Wang Z, Merrikh CN, Lang KS, Lu P, Li X, Merrikh H, Rao Z, and Xu W (2018). Crystal structure of a membrane-bound O-acyltransferase. Nature 562, 286–290. 10.1038/s41586-018-0568-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Coupland CE, Ansell TB, Sansom MSP, and Siebold C (2023). Rocking the MBOAT: Structural insights into the membrane bound O-acyltransferase family. Current opinion in structural biology 80, 102589. 10.1016/j.sbi.2023.102589. [DOI] [PubMed] [Google Scholar]
- 50.Qi X, Hu Q, Elghobashi-Meinhardt N, Long T, Chen H, and Li X (2023). Molecular basis of Wnt biogenesis, secretion, and Wnt7-specific signaling. Cell 186, 5028–5040 e5014. 10.1016/j.cell.2023.09.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Hiraizumi M, Yamashita K, Nishizawa T, and Nureki O (2019). Cryo-EM structures capture the transport cycle of the P4-ATPase flippase. Science 365, 1149–1155. 10.1126/science.aay3353. [DOI] [PubMed] [Google Scholar]
- 52.Goldstein JL, Basu SK, and Brown MS (1983). Receptor-mediated endocytosis of low-density lipoprotein in cultured cells. Methods Enzymol 98, 241–260. [DOI] [PubMed] [Google Scholar]
- 53.Zheng SQ, Palovcak E, Armache JP, Verba KA, Cheng Y, and Agard DA (2017). MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy. Nature methods 14, 331–332. 10.1038/nmeth.4193. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Zivanov J, Nakane T, Forsberg BO, Kimanius D, Hagen WJ, Lindahl E, and Scheres SH (2018). New tools for automated high-resolution cryo-EM structure determination in RELION-3. eLife 7. 10.7554/eLife.42166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Rohou A, and Grigorieff N (2015). CTFFIND4: Fast and accurate defocus estimation from electron micrographs. J. Struct. Biol 192, 216–221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Punjani A, Rubinstein JL, Fleet DJ, and Brubaker MA (2017). cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods 14, 290–296. [DOI] [PubMed] [Google Scholar]
- 57.Wagner T, Merino F, Stabrin M, Moriya T, Antoni C, Apelbaum A, Hagel P, Sitsel O, Raisch T, Prumbaum D, et al. (2019). SPHIRE-crYOLO is a fast and accurate fully automated particle picker for cryo-EM. Commun Biol 2, 218. 10.1038/s42003-019-0437-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Emsley P, and Cowtan K (2004). Coot: model-building tools for molecular graphics. Acta crystallographica. Section D, Biological crystallography 60, 2126–2132. 10.1107/S0907444904019158. [DOI] [PubMed] [Google Scholar]
- 59.Adams PD, Afonine PV, Bunkoczi G, Chen VB, Davis IW, Echols N, Headd JJ, Hung LW, Kapral GJ, Grosse-Kunstleve RW, et al. (2010). PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta crystallographica. Section D, Biological crystallography 66, 213–221. 10.1107/S0907444909052925. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Chen VB, Arendall WB 3rd, Headd JJ, Keedy DA, Immormino RM, Kapral GJ, Murray LW, Richardson JS, and Richardson DC (2010). MolProbity: all-atom structure validation for macromolecular crystallography. Acta crystallographica. Section D, Biological crystallography 66, 12–21. 10.1107/S0907444909042073. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Pettersen EF, Goddard TD, Huang CC, Meng EC, Couch GS, Croll TI, Morris JH, and Ferrin TE (2021). UCSF ChimeraX: Structure visualization for researchers, educators, and developers. Protein Sci 30, 70–82. 10.1002/pro.3943. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Matyash V, Liebisch G, Kurzchalia TV, Shevchenko A, and Schwudke D (2008). Lipid extraction by methyl-tert-butyl ether for high-throughput lipidomics. Journal of lipid research 49, 1137–1146. 10.1194/jlr.D700041-JLR200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.McDonald JG, Smith DD, Stiles AR, and Russell DW (2012). A comprehensive method for extraction and quantitative analysis of sterols and secosteroids from human plasma. Journal of lipid research 53, 1399–1409. 10.1194/jlr.D022285. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Chalvardjian A, and Rudnicki E (1970). Determination of lipid phosphorus in the nanomolar range. Analytical biochemistry 36, 225–226. 10.1016/0003-2697(70)90352-0. [DOI] [PubMed] [Google Scholar]
- 65.Liang G, Yang J, Horton JD, Hammer RE, Goldstein JL, and Brown MS (2002). Diminished hepatic response to fasting/refeeding and liver X receptor agonists in mice with selective deficiency of sterol regulatory element-binding protein-1c. The Journal of biological chemistry 277, 9520–9528. 10.1074/jbc.M111421200. [DOI] [PubMed] [Google Scholar]
- 66.Beisiegel U, Schneider WJ, Goldstein JL, Anderson RG, and Brown MS (1981). Monoclonal antibodies to the low density lipoprotein receptor as probes for study of receptor-mediated endocytosis and the genetics of familial hypercholesterolemia. The Journal of biological chemistry 256, 11923–11931. [PubMed] [Google Scholar]
- 67.Olsson MH, Sondergaard CR, Rostkowski M, and Jensen JH (2011). PROPKA3: Consistent Treatment of Internal and Surface Residues in Empirical pKa Predictions. J Chem Theory Comput 7, 525–537. 10.1021/ct100578z. [DOI] [PubMed] [Google Scholar]
- 68.Jo S, Kim T, Iyer VG, and Im W (2008). CHARMM-GUI: a web-based graphical user interface for CHARMM. Journal of computational chemistry 29, 1859–1865. 10.1002/jcc.20945. [DOI] [PubMed] [Google Scholar]
- 69.Case DA, Aktulga HM, Belfon K, Cerutti DS, Cisneros GA, Cruzeiro VWD, Forouzesh N, Giese TJ, Gotz AW, Gohlke H, et al. (2023). AmberTools. J Chem Inf Model 63, 6183–6191. 10.1021/acs.jcim.3c01153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Lee J, Cheng X, Swails JM, Yeom MS, Eastman PK, Lemkul JA, Wei S, Buckner J, Jeong JC, Qi Y, et al. (2016). CHARMM-GUI Input Generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using the CHARMM36 Additive Force Field. J Chem Theory Comput 12, 405–413. 10.1021/acs.jctc.5b00935. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Lee J, Hitzenberger M, Rieger M, Kern NR, Zacharias M, and Im W (2020). CHARMM-GUI supports the Amber force fields. J Chem Phys 153, 035103. 10.1063/5.0012280. [DOI] [PubMed] [Google Scholar]
- 72.Maier JA, Martinez C, Kasavajhala K, Wickstrom L, Hauser KE, and Simmerling C (2015). ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J Chem Theory Comput 11, 3696–3713. 10.1021/acs.jctc.5b00255. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Wang NX, and Wilson AK (2004). The behavior of density functionals with respect to basis set. I. The correlation consistent basis sets. J Chem Phys 121, 7632–7646. 10.1063/1.1792071. [DOI] [PubMed] [Google Scholar]
- 74.Wang J, Wang W, Kollman PA, and Case DA (2006). Automatic atom type and bond type perception in molecular mechanical calculations. J Mol Graph Model 25, 247–260. 10.1016/j.jmgm.2005.12.005. [DOI] [PubMed] [Google Scholar]
- 75.Jakalian A, Jack DB, and Bayly CI (2002). Fast, efficient generation of high-quality atomic charges. AM1-BCC model: II. Parameterization and validation. Journal of computational chemistry 23, 1623–1641. 10.1002/jcc.10128. [DOI] [PubMed] [Google Scholar]
- 76.Dickson CJ, Walker RC, and Gould IR (2022). Lipid21: Complex Lipid Membrane Simulations with AMBER. J Chem Theory Comput 18, 1726–1736. 10.1021/acs.jctc.1c01217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Coleman TG, Mesick HC, and Darby RL (1977). Numerical integration: a method for improving solution stability in models of the circulation. Ann Biomed Eng 5, 322–328. 10.1007/BF02367312. [DOI] [PubMed] [Google Scholar]
- 78.Loncharich RJ, Brooks BR, and Pastor RW (1992). Langevin dynamics of peptides: the frictional dependence of isomerization rates of N-acetylalanyl-N’-methylamide. Biopolymers 32, 523–535. 10.1002/bip.360320508. [DOI] [PubMed] [Google Scholar]
- 79.Berendsen HJC, Postma JPM, Gunsteren W.F.v., DiNola A, and Haak JR (1984). Molecular dynamics with coupling to an external bath. J. Chem. Phys 81, 3684–3690. [Google Scholar]
- 80.Roe DR, and Cheatham TE 3rd (2013). PTRAJ and CPPTRAJ: Software for Processing and Analysis of Molecular Dynamics Trajectory Data. J Chem Theory Comput 9, 3084–3095. 10.1021/ct400341p. [DOI] [PubMed] [Google Scholar]
- 81.Miller BR 3rd, McGee TD Jr., Swails JM, Homeyer N, Gohlke H, and Roitberg AE (2012). MMPBSA.py: An Efficient Program for End-State Free Energy Calculations. J Chem Theory Comput 8, 3314–3321. 10.1021/ct300418h. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
(A) Sequence alignment of human PSS1 and PSS2 with mouse, chicken, xenopus, and zebrafish PSS1. The transmembrane helices and the residue numbers of human PSS1 are indicated above the protein sequence. The specific residues necessary for calcium binding are indicated by circles, the catalytic histidine is indicated by an asterisk and the cysteines for disulfate bond are indicated by purple squares. The residues necessary for inhibitor binding are indicated by dark blue squares.
(B) Purification of human PSS1WT and its variants for cryo-EM study and activity assays. Representative Superose 6 increase 10/30 gel-filtration chromatogram of PSS1WT, PSS1P269S, PSS1H172A, PSS1F168A, PSS1F168W, PSS1G171W, PSS1G175W and PSS1F313M/S320V in buffer containing 20 mM HEPES pH 7.5, 150 mM NaCl, 5 mM CaCl2 and 0.06% Digitonin. The peak fractions used in this study are indicated by red arrows.
(A) Representative cryo-EM image of PSS1WT. Scale bar, 40 nm.
(B) Selected 2D class averages of PSS1WT.
(C) Summary of cryo-EM data processing procedures. FSC curves between two half maps are shown. The mask, which was generated by CryoSPARC for the final refinement, is shown in gray.
(D) Angular distribution of PSS1WT particles in the final round of 3D refinement using CryoSPARC.
(E) Local resolution of cryo-EM map. Maps are colored according to local resolution, estimated using CryoSPARC.
(F) Cryo-EM map of the transmembrane helices.
(A) The binding free energy of PE and PC to PSS1WT monomer with and without PS from three MD simulations. This is also the case for (B) and (D). Unit: kcal/mol.
(B) The binding free energy of PS to PSS1WT monomer from three MD simulations. Unit: kcal/mol.
(C) The representative curves showing the distance between the side chain center of the corresponding residue and calcium ions in the structure of PSS1 with (blue) or without PS (green).
(D) The binding free energy of serine to PSS1P269S monomer from three MD simulations. Unit: kcal/mol.
(E) Distance between Cα atoms of Phe168 and Trp321 in the PSS1WT two monomers with (blue) or without PS (green).
(A) Summary of cryo-EM data processing procedures. The mask, which was generated by CryoSPARC for the final refinement, is shown in gray.
(B) Angular distribution of PSS1P269S particles in the final round of 3D refinement using CryoSPARC. FSC curves between two half maps are shown.
(C) Local resolution of cryo-EM map. Maps are colored according to local resolution, estimated using CryoSPARC.
(D) Cryo-EM map of the transmembrane helices.
(E) The extra density at the lipid entrance in the cryo-EM map of PSS1P269S.
(F) The extra lipid-like density in the cryo-EM map of PSS1P269S.
(G) Comparison of the cryo-EM maps of PSS1P269S and PSS1WT indicates the flexibility of PH4 in PSS1P269S. The cryo-EM map of PSS1WT has been filtered to 3.3-Å.
(H) Comparison of calcium binding sites of PSS1P269S and AF model (gray). The calcium in the structure of PSS1P269S is shown as a green sphere. Dashed line represents hydrophilic interactions.
(A) Summary of cryo-EM data processing procedures. FSC curves between two half maps are shown. The mask, which was generated by CryoSPARC for the final refinement, is shown in gray.
(B) Angular distribution of DS55980254 bound PSS1WT particles in the final round of 3D refinement using CryoSPARC.
(C) Local resolution of cryo-EM map. Maps are colored according to local resolution, estimated using CryoSPARC.
(D) Cryo-EM map of the transmembrane helices.
(A) Measurement of LDLRs by flow cytometry in SV589j cells incubated with LDL. On day 0, the indicated cells were set up in medium A with 5% FCS. After plating for 6 hours, cells were supplemented with DMSO (solvent control) or 1 μM DS, which was kept at a constant concentration in each buffer throughout the experiment. On day 1, cells were switched to cholesterol-depletion medium B. After 16 h, cells then received the above cholesterol-depletion medium containing 50 μg protein/ml of LDL. After 24 h, the cells were harvested by incubation with EDTA, washed, incubated with PE-anti-LDLR, and subjected to flow cytometry (METHOD DETAILS).
(B) Measurement of BODIPY FL-LDL uptake by flow cytometry in SV589j cells. On day 0, the indicated cells were set up in medium A with 5% FCS. After plating for 6 hours, cells were supplemented with DMSO (solvent control) or 1 μM DS, which was kept at a constant concentration in each buffer throughout the experiment. On day 1, cells were switched to cholesterol-depletion medium B. After 16 h, cells then received the above cholesterol-depletion medium containing 50 μg protein/ml of LDL. After 24 h, the cells were treated by 5 μg protein/ml of BODIPY FL-LDL for 2 hours, then the cells were harvested by incubation with EDTA, washed and subjected to flow cytometry (METHOD DETAILS).
(A) Phylogenetic tree of PSSs and MBOATs. Human, mouse, frog, zebrafish, Arabidopsis, yeast and bacteria proteins were used for the comparison. Numbers at the nodes indicates the bootstrap values on neighbor joining analysis. The calculation was performed in MEGA X.
(B) Working mechanisms of PORCN and ACAT1. The catalytic histidine activates the hydroxyl group of the serine of Wnt (PORCN) or cholesterol (ACAT1), which would attack the thioester bond of acyl-CoA substrate, to form the product.
Data Availability Statement
The 3D cryo-EM density maps have been deposited in the Electron Microscopy Data Bank (EMDB) under the accession numbers EMD-44178, EMD-44179 and EMD-44180. Atomic coordinates for the atomic model have been deposited in the Protein Data Bank (PDB) under the accession numbers 9B4E, 9B4F and 9B4G.
This paper does not report original code.
Any additional information required to reanalyze the data reported in this paper is available from the lead contact upon request.
