Summary
Telomere maintenance requires extension of the G-rich telomeric repeat strand by telomerase and fill-in synthesis of the C-rich strand by Polα/primase. At telomeres, Polα/primase is bound to Ctc1/Stn1/Ten1 (CST), a single-stranded DNA-binding complex. Like mutations in telomerase, mutations affecting CST–Polα/primase result in pathological telomere shortening and cause a telomere biology disorder, Coats plus (CP). We determined cryogenic electron microscopy structures of human CST bound to the shelterin heterodimer POT1/TPP1 that reveal how CST is recruited to telomeres by POT1. Our findings suggest that POT1 hinge phosphorylation is required for CST recruitment, and the complex is formed through conserved interactions involving several residues mutated in CP. Our structural and biochemical data suggest that phosphorylated POT1 holds CST–Polα/primase in an inactive auto-inhibited state until telomerase has extended the telomere ends. We propose that dephosphorylation of POT1 releases CST–Polα/primase into an active state that completes telomere replication through fill-in synthesis.
Keywords: telomere, DNA replication, CST, Polymerase α/primase, shelterin, POT1, cryo-EM, DNA-protein complex, phospho-regulation
Graphical Abstract
In Brief
Cryo-EM structures of CST–POT1/TPP1 complexes reveal the molecular basis of how the shelterin subunit POT1 recruits CST–Polα/primase to telomeres to solve the end-replication problem. POT1 phosphorylation regulates the activity of CST–Polα/primase, providing a mechanism for the temporal control of telomeric C-strand maintenance in humans.
Introduction
Telomeres prevent the inappropriate activation of the DNA damage response at the ends of linear chromosomes. Human telomeric DNA is comprised of double-stranded (ds) 5’-TTAGGG-3’ repeats that terminate in a single-stranded (ss) 3’ overhang. This DNA protects chromosome ends through its interaction with the six-subunit shelterin complex (reviewed by de Lange1). With each cell cycle, incomplete DNA synthesis and nucleolytic resection at telomere ends results in attrition of the telomeric DNA that can lead to excessive telomere shortening if unchecked (reviewed by Cai and de Lange2). Telomere maintenance is critical for the long-term proliferation of human cells and involves two distinct DNA synthesizing enzymes: telomerase, which elongates the G-rich strand, and CST–Polα/primase, which maintains the C-rich repeats through fill-in synthesis (reviewed by Cai and de Lange2 and Lim and Cech3).
Loss of CST leads to progressive shortening of the C-rich strand at telomeres duplicated by lagging-strand DNA synthesis. This loss of telomeric DNA is caused by the inability of the replisome to generate Okazaki fragments close to the end of a linear DNA4. This lagging-end replication problem is solved by CST–Polα/primase, not telomerase. CST–Polα/primase also counteracts C-strand shortening at telomeres duplicated by leading-strand DNA synthesis, where the blunt-ended leading-strand products are resected to recreate the 3’ overhang. In this context, CST–Polα/primase guards against C-strand loss resulting from hyper-resection5–8. Whereas much is known about how telomerase is recruited and regulated at telomere ends9, 10, the mechanism of CST–Polα/primase recruitment, regulation, and action is not well understood.
CST is a heterotrimeric ssDNA-binding protein that shares structural homology with the major eukaryotic ssDNA-binding protein, Replication Protein A (RPA)11–13. However, it is likely that CST evolved independently in an archaeal ancestor and became specialized for telomere maintenance in eukaryotes14 (reviewed by Cai and de Lange2). In metazoans, the large Ctc1 subunit of CST contains two sets of tandem oligosaccharide/oligonucleotide-binding (OB) folds13 that are connected by a three-helix bundle, termed the Acidic Rpa1 OB-binding Domain-Like three-helix bundle (Ctc1ARODL) based on similarity to its archaeal counterpart2, 14. Ctc1OB-A/B/C/D and the Ctc1ARODL comprise the metazoan-specific N-terminal domain of the protein, whereas Ctc1OB-E/F/G are ancestral and mediate ssDNA binding and formation of the Ctc1/Stn1/Ten1 heterotrimer. Stn1, comprising an N-terminal OB fold (Stn1N) and C-terminal tandem winged helix-turn-helix (wHTH) domains (Stn1C), connects the single OB fold of Ten1 to Ctc113.
Recent cryo-EM structures of CST–Polα/primase revealed how CST interacts with Polα/primase in two distinct conformations. In these structures, Polα/primase either adopts a compact, auto-inhibited state15 or an extended state compatible with enzymatic activity16. CST binds the auto-inhibited Polα/primase using the Ctc1 N-terminal domain, and this interaction is conserved in metazoans15. The inactive complex was proposed to represent a Recruitment Complex (RC) that describes the conformation of CST–Polα/primase as it is brought to the telomere prior to fill-in. In the Pre-Initiation Complex (PIC) in which Polα/primase is primed for activation, CST uses its ancestral RPA-like regions (Ctc1OB-E/F/G, Stn1, and Ten1) to stabilize and stimulate the extended Polα/primase16. The cryo-EM structures indicate that a significant conformational change is required to transition from the RC to PIC, but how this change is regulated is unclear.
The importance of CST–Polα/primase in the maintenance of telomeres is illustrated by CP, a rare pediatric condition characterized by the ophthalmological disorder Coats disease plus associated systemic disorders primarily affecting the brain, bone, and gastrointestinal tract17–25. CP is caused by hypomorphic mutations in Ctc1 and Stn1 that reveal their phenotype only when the second copy of the gene is non-functional. In addition, CP can be caused by homozygosity for a hypomorphic mutation in Stn1 or POT123, 26.
It has remained unclear how CST–Polα/primase is recruited to telomeres. Human TPP1 readily interacts with Ctc1 and was therefore thought to play a primary role in recruitment of human CST27. However, at mouse telomeres, it is one of the two POT1 proteins (mPOT1b) that interacts with CST and mutations in mPOT1b diminish CST recruitment and lead to lack of C-strand fill-in at telomeres5. Whereas human POT1 has not been shown to interact with CST in co-immunoprecipitation (co-IP) experiments, it does interact with Ctc1 in a yeast two-hybrid assay28 (but also see Miyake et al.12). Furthermore, the finding that the POT1 CP mutation (S322L) confers a telomere phenotype consistent with CST deficiency argues that POT1 contributes to CST recruitment23. An initially confusing result indicated that removal of TPP1 or POT1 from human telomeres led to an increase, rather than a decrease, of CST loading28. However, this result is explained by the activation of ATR signaling at deprotected telomeres lacking POT1, which leads to 53BP1- and Shieldin-dependent loading of CST29, 30.
Here, we present structural and biochemical data revealing that human CST primarily interacts with POT1. This interaction is dependent on POT1 phosphorylation, reconciling the negative co-IP data. We identified four amino acids in mPOT1b that, when inserted into the human protein, result in POT1 phosphorylation in both human cells and insect cells. With the phosphorylated form of POT1, we reconstituted CST–POT1/TPP1 complexes in the presence and absence of telomeric ssDNA and determined their structures using cryo-EM. The structures suggest that phosphorylation of the POT1 hinge region drives charge complementarity between the two proteins. Two other interfaces form independently from the mPOT1b insertion and confirm that human POT1 is the primary recruiter of CST. Several CP mutations map to or near the primary interface between Ctc1OB-D, Ctc1ARODL, and POT1OB-3. The POT1 N-terminal OB folds interact with Ctc1’s ssDNA-binding site, directly competing with CST binding to ssDNA and Polα/primase in the active PIC conformation, whereas the major interface on Ctc1 for Polα/primase binding in the auto-inhibited RC conformation is unobstructed. In vitro C-strand synthesis assays reveal that POT1/TPP1 inhibits CST–Polα/primase activity when bound to CST, supporting the structural analysis. Together, our data point toward a phosphorylation-dependent switch in POT1 that controls its interaction with CST and can regulate the activity of CST–Polα/primase at the telomere. This provides a molecular basis for how shelterin recruits and regulates CST–Polα/primase and reveals the mechanism of mutations in CST and POT1 that cause telomere dysfunction in CP.
Results
Reconstitution of a stable CST–POT1/TPP1 complex
To understand how POT1/TPP1 recruits CST, we examined how TPP1, the established CST interactor, binds. Co-IP data indicated that a region within the C-terminal 20 residues of TPP1 is necessary for its interaction with CST (Fig. S1A). This region of TPP1 constitutes its TIN2-Interaction Domain (TID) and is required to anchor the POT1/TPP1 heterodimer to the telomeric dsDNA-binding proteins in shelterin31. Indeed, co-IP experiments showed that TIN2 competes with CST for TPP1 binding (Fig. S1B). AlphaFold-Multimer32 modeling of the TID bound to Ctc1 predicted that TPP1 binds to OB-B of Ctc1 using the same amino acids used for binding to TIN2, consistent with the finding that TPP1 can interact with either TIN2 or Ctc1 but not both (Fig. S1B–D). Since its association with TIN2 is required for TPP1/POT1 to localize to telomeres33, 34, we consider it unlikely that the TPP1TID is involved in the recruitment of CST. As predicted, a recent report confirmed that loss of the interaction between TPP1TID and Ctc1OB-B did not disrupt fill-in by CST–Polα/primase35. We therefore omitted the C-terminus (aa 252–458), including the TID, from TPP1 to limit sample heterogeneity from this largely disordered region.
Interestingly, although human POT1 does not form a stable complex with CST in co-IP experiments, mPOT1b readily bound to human CST (Fig. 1A–C). This finding suggested that mPOT1b can constitutively bind CST whereas the binding of human POT1 to CST may be regulated, perhaps by post-translational modifications. To determine which region in mPOT1b promotes the CST interaction, chimeric proteins containing swaps between mPOT1b and human POT1 regions were assayed for CST binding in co-IP experiments (Fig. 1B). When swapped with the human POT1 hinge (aa 299–344), the mPOT1b hinge (aa 300–350) conferred CST interaction (Fig. 1A, C). Further truncations and sequence comparison identified four residues in mPOT1b (ESDL, aa 323–326, Fig. 1B–C) that, when inserted into the human POT1 hinge (between aa 320–321, Fig. 1B–C), conferred robust interaction with CST (Fig. 1A–C). Furthermore, whereas there was no detectable interaction between purified wild-type (WT) POT1/TPP1 and CST, POT1(ESDL)/TPP1 formed a complex with CST that was detectable by fluorescence-detection size-exclusion chromatography (FSEC, Fig. 1D–E). POT1(ESDL)/TPP1 formed a complex with CST in the presence or absence of ssDNA, but addition of a telomeric [GGTTAG]3 ssDNA allowed complex formation at lower protein concentrations (Fig. 1E; Fig. S2A).
Figure 1. Identification of residues promoting the POT1–CST interaction.
See also Figure S1 and Figure S2.
(A) Domain schematics and design of constructs containing chimeric swaps between human POT1 (red) and mouse mPOT1b (salmon). HMb: Human/Mouse POT1b swap; OB: oligosaccharide/oligonucleotide-binding domain; HJRL: Holliday junction resolvase-like domain. The interaction with human CST is indicated as shown in (C).
(B) Sequence alignment of the region within the POT1 hinge in POT1, mPOT1b, and POT1(ESDL) showing the location of the ESDL insertion.
(C) Co-IPs of Myc-tagged POT1/mPOT1b chimeras and FLAG-tagged Ctc1 from 293T cells transfected with Myc-POT1 (and variants), FLAG-TPP1, FLAG-Ctc1, FLAG-Stn1, and FLAG-Ten1. Western blots of Myc IPs were probed with anti-Myc and anti-FLAG antibodies to detect POT1 and Ctc1, respectively. POT1 construct details and names are provided in (A).
(D) Coomassie-stained SDS-PAGE gels and cartoon schematics of the purified POT1/TPP1 and CST proteins used for fluorescent-detection size-exclusion chromatography (FSEC) analysis.
(E) FSEC analysis of the interaction between CST and POT1(WT)/GFP-TPP1 (black) or POT1(ESDL)/GFP-TPP1 (blue) in the absence (top) or presence (bottom) of telomeric ssDNA. Traces without CST are shown as dashed lines and color key is the same for both panels. RFU: relative fluorescence units.
We reconstituted two complexes of CST–POT1(ESDL)/TPP1. With the addition of the 18-nt [GGTTAG]3 ssDNA, we could reconstitute a stable, native complex containing full-length CST, full-length POT1(ESDL), and TPP1 (aa 1–251; Fig. 2A; Fig. S2B). Without ssDNA, the complex dissociated at the low concentrations (< 200 nM) used for negative-stain and cryo-EM (Fig. S2A). To increase the local concentration of CST relative to POT1(ESDL)/TPP1 for structural studies, we fused aa 1–403 of TPP1 to the N-terminus of Stn1, retaining 162 aa of the disordered serine-rich linker of TPP1 to allow for flexibility between the two polypeptides (Fig. S2C). This fusion allowed for purification of an apo CST–POT1(ESDL)/TPP1 complex in the absence of ssDNA (Fig. S2D).
Figure 2. Cryo-EM structures of CST–POT1(ESDL)/TPP1 complexes.
See also Figure S3 and Figure S4.
(A) Domain organization of CST and the shelterin subunits POT1(ESDL) and TPP1. Regions of TPP1 not observed in the cryo-EM maps are shown at 50% opacity. Dashed lines indicate regions not included in the expression construct. *The TPP1 serine-rich linker was included in the apo complex but not the ssDNA-bound complex. See also Fig. S2C. ARODL: Acidic Rpa1 OB-binding Domain-Like 3-helix bundle; wH: winged helix-turn-helix domain; RD: recruitment domain, TID: TIN2-interacting domain. For other abbreviations see Fig. 1.
(B) Cryo-EM density map (contour threshold 0.15) of apo CST–POT1(ESDL)/TPP1 at 3.9-Å resolution (see also Fig. S3).
(C) Cryo-EM density map (contour threshold 0.15) of ssDNA-bound CST–POT1(ESDL)/TPP1 at 4.3-Å resolution (see also Fig. S4).
(D-E) Atomic models for the apo (D) and ssDNA-bound (E) CST–POT1(ESDL)/TPP1 complexes, respectively. See also Video S1.
Cryo-EM structures of CST–POT1/TPP1
Negative-stain EM analysis showed additional density bound to CST, attributable to the addition of POT1(ESDL)/TPP1 in the two complexes (Fig. S2E). We determined cryo-EM structures of apo and ssDNA-bound CST–POT1(ESDL)/TPP1 at overall resolutions of 3.9 Å and 4.3 Å, respectively (Fig. 2; Fig. S3–4; Video S1; Table S1). Into these density maps, we could unambiguously place models of the individual subunits predicted by AlphaFold236, 37 or previously determined by X-ray crystallography or cryo-EM as a starting point to build the models (see Table S1). The map quality of the apo CST–POT1(ESDL)/TPP1 reconstruction was higher (Fig. S3H), so we primarily used that map for model building. The resulting atomic model from the apo complex was then used as a starting model to build the ssDNA-bound complex. We observed continuous density for the POT1 flexible hinge and were able to build its Cα backbone de novo, generating an experimental structure of full-length POT1(ESDL) (Fig. 2D–E, Fig. S3–4; Video S1). The POT1 C-terminus, consisting of POT1OB-3 and POT1HJRL (Fig. 2A), is bound by the POT1-recruitment domain of TPP1 (TPP1RD), consistent with previous crystal structures (PDB 5H6538 and PDB 5UN739) (Fig. 2D–E). TPP1RD does not interact directly with CST in either structure. The N-terminal OB fold of TPP1 (TPP1OB) was not resolved, presumably due to flexible attachment, and TPP1OB was indeed dispensable in the interaction of POT1(ESDL)/TPP1 with CST (Fig. S5A).
POT1(ESDL) is held in a single conformation stretched along the entire length of Ctc1 and buries a total of 5,246 Å2 of solvent-accessible surface area (Fig. 2D–E). The structures of the apo and ssDNA-bound complexes had identical overall interactions between POT1(ESDL) and CST, but the addition of ssDNA allowed visualization of the POT1OB-1 domain, which binds ssDNA in conjunction with POT1OB-2 (Fig. 2C, E). Because the region of the map containing the POT1 N-terminal OB-folds was at lower resolution (Fig. S4H), we could only dock in the existing crystal structure of POT1 bound to 5’-TTAGGGTTAG-3’ (PDB 1XJV40) and could not assign a register to the DNA or observe additional nucleotides of the [GGTTAG]3 oligonucleotide. Without ssDNA, POT1OB-1 is flexibly attached and thus remains unresolved in the map of the apo complex (Fig. 2B; Fig. S3).
Phosphorylation of POT1 mediates the interaction with CST
The hinge region of POT1(ESDL) containing the ESDL insertion sits in a positively charged cleft of Ctc1OB-D (Fig. 3A; Fig. S5B). We term this region of POT1 the Ctc1 Cleft-Interacting Region (CCIR, aa 309–321). Although there is connectivity in the cryo-EM map, we could not unambiguously resolve the side chain positions, suggesting this interface is formed by weaker interactions (Fig. 3A; Fig. S5B). Notably, the ESDL insertion does not make specific contacts with Ctc1, suggesting that the observed complex is not the result of an artificial interaction of this mouse/human chimera. The POT1 CCIR contains aspartate residues that complement the basic cleft of Ctc1 as well as an abundance of serine residues, suggesting that this interaction may be regulated by phosphorylation of the CCIR to drive charge complementarity between the two proteins.
Figure 3. Phosphorylation-dependent interaction of POT1(ESDL) with CST.
See also Figure S5.
(A) Interaction between the POT1 CCIR and Ctc1. (Left) The black box indicates the region of the apo CST–POT1(ESDL)/TPP1 model shown in the right panels. (Right) Zoomed-in view showing surface electrostatic potential analysis of the POT1(ESDL) CCIR interacting with Ctc1OB-D. The POT1(ESDL) CCIR is shown in cartoon (left) and surface (right) representation with electrostatic potential colored from red (−10 kBT/e) to white (0 kBT/e) to blue (+10 kBT/e) (see also Fig. S5B). The ESDL insertion is colored salmon as in Fig. 1A, and residue numbering corresponds to the WT human POT1 sequence.
(B) FSEC analysis of the phosphorylation-dependent interaction between CST and POT1(WT)/GFP-TPP1 (left, black) and POT1(ESDL)/GFP-TPP1 (right, blue) in the presence of telomeric ssDNA (see also Fig. S5C–D). Traces without CST are shown as dashed lines, and traces containing POT1/TPP1 that have been treated with λ protein phosphatase (λPP) are shown in red. RFU: relative fluorescence units. The chromatograms for the WT mock and ESDL mock samples are the same as in Fig. 1E. They are reproduced here as the controls for the phosphatase experiment, which was performed simultaneously.
(C) Representative MP histograms showing the phosphorylation-dependent interaction between CST and POT1(WT)/GFP-TPP1 (left, black) and POT1(ESDL)/GFP-TPP1 (right, blue) in the presence of telomeric ssDNA (see also Fig. S5E). Experiments with λPP-treated POT1/TPP1 experiments are shown in translucent red. Peak maxima are indicated along with cartoons (peak 1: free POT1/TPP1, peak 2: free CST, peak 3: CST–POT1/TPP1) that correspond to complexes of that molecular mass to within 10%.
(D) Kinase assay with HeLa nuclear extract and filter-bound peptides corresponding to the CCIR regions of POT1, mPOT1b, mPOT1a, and POT1(ESDL). Phosphorylation is monitored using incorporation of γ-32P-ATP.
(E) Quantification of MP results (Fig. S5E). Proportion bound POT1/TPP1 corresponds to the ratio of total counts in peak 3 (CST–POT1/TPP1 complex) divided by total counts of POT1/TPP1 (peak 1 and peak 3). Error bars represent the S.D. from three technical replicate experiments. POT1 construct details and names are provided in (H).
(F - G) Co-IPs of Myc-tagged POT1 protein variants and FLAG-tagged Ctc1 from 293T cells co-transfected with Myc-POT1, FLAG-TPP1, FLAG-Ctc1, FLAG-Stn1, and FLAG-Ten1. Western blots of Myc IPs were probed with anti-Myc and anti-FLAG antibodies to detect POT1 and Ctc1, respectively. Short and long exposures of the anti-FLAG probing for Ctc1 are shown. POT1 construct details and names are provided in (H).
(H) Sequences of phosphomimetic, alanine, and CP CCIR mutants. The interaction with CST is indicated as shown in (E-G). The “+” signifies and interaction matching that of the positive control, POT1(ESDL), and “−” signifies no interaction. “++” represents greater interaction than the positive control and “+/−“ and “−/+” indicate slightly weaker interactions in decreasing strength, respectively.
To test the role of phosphorylation, we treated the POT1/TPP1 proteins with λ protein phosphatase (λPP; Fig. S5C). The CST–POT1(ESDL)/TPP1 interaction was severely diminished by dephosphorylation of POT1(ESDL)/TPP1 in both FSEC and mass photometry (MP) analysis of complex formation (Fig. 3B–C; Fig. S5D–E). The propensity of the ESDL insertion to enhance phosphorylation was shown in a kinase assay in which HeLa cell nuclear extract was incubated with nitrocellulose membrane-bound synthetic peptides (Fig. 3D). Human POT1 and mPOT1b peptides containing the analogous CCIRs were phosphorylated more than the corresponding peptide from mPOT1a, consistent with the inability of mPOT1a to bind CST5, and the POT1(ESDL) peptide was phosphorylated to the greatest degree of the four (Fig. 3D). Together, this suggests that the ESDL insertion primarily functions to enhance the phosphorylation of human POT1 rather than to directly interact with CST.
Mass spectrometry efforts to determine the sites of phosphorylation in the Ser-rich CCIR were not informative, as the relevant peptide was not well-represented despite multiple trials with different proteases. To validate the structure and further characterize the phosphorylation events involved in complex formation, we generated phosphomimetic CCIR mutants (S-to-D) in POT1 and POT1(ESDL) as well as S-to-A mutants (Fig. 3E–H). MP was used to quantify binding efficiency by measuring the proportion of CST-bound POT1/TPP1 (MP peak at ~390 kDa) to free POT1/TPP1 (MP peak at ~140 kDa; Fig. 3C, E; Fig. S5E). In contrast to the sharp complex peak (σ ≤ 40) observed with POT1(ESDL)/TPP1, the CST–POT1/TPP1 peaks with CCIR phosphomimetic changes were broader (σ > 40) but still symmetric and centered at the same mass value (Fig. S5E). The broader peaks suggest that the phosphomimetic mutant complexes may be slightly more heterogeneous and less stable than the phosphorylated POT1(ESDL)/TPP1. The quantification by total peak counts (area under the Gaussian curve) masks this subtle difference, but importantly, the phosphomimetic CCIR mutants conferred λPP-resistant CST binding compared to both wild-type POT1/TPP1 and dephosphorylated POT1(ESDL)/TPP1 (Fig. 3E; Fig. S5E). Phosphomimetic substitutions at Ser317, Ser318, Ser320, and Ser322 increased the proportion of CST-bound POT1/TPP1 to almost the same level as observed with the ESDL insertion, and the interactions with the phosphomimetic proteins were resistant to phosphatase treatment (Fig. 3E; Fig. S5E). Similarly, in co-IP experiments that queried the interaction of POT1 with Ctc1, phosphomimetic CCIR substitutions of Ser317, Ser318, and Ser320 led to a robust interaction close to that seen with the ESDL version of POT1 (Fig. 3F). Greater interaction over that of POT1(ESDL) was observed with a combination of the phosphomimetic substitutions and the ESDL, suggesting that the insertion adds additional charge to provide the increased stability seen with the purified proteins (Fig. 3E–F, H). However, the in vitro and in vivo assays gave disparate results regarding the Ser322 position. In the co-IPs, the S322D substitution diminished the interaction despite the phosphomimetic substitutions at the other positions whereas in vitro, the S322D substitution was required for optimal binding of POT1/TPP1 to CST. The reason for this discrepancy remains to be determined. Nonetheless, both assays point to the critical role of serine phosphorylation in the CCIR of POT1 for Ctc1 binding. This conclusion was further substantiated by converting CCIR residues Ser317, Ser318, and Ser320 to alanine in POT1(ESDL), which abolished the ability of POT1(ESDL) to form a complex with Ctc1 in co-IP experiments (Fig. 3G–H).
POT1OB-3 forms the primary interface with Ctc1
Human POT1 directly interacts with Ctc1 at two sites independent from the POT1CCIR and the ESDL insertion (Fig. 2D–E). The primary interface contains POT1OB-3, Ctc1ARODL, and Ctc1OB-D and buries 2,012 Å2 of solvent-accessible surface area (Fig. 4A–B; Fig. S6A). This interface is formed by the C-terminal region of the POT1 hinge (aa 322–347), which folds back on POT1OB-3 to create the surface to which Ctc1 binds. Part of this region (aa 332–347) has previously been crystallized39, but in another crystal structure it was too flexible to be resolved38 (Fig. 4A). In our structure, this region of the hinge is pinned to POT1OB-3 at two sites (sites (i) and (ii)) of intramolecular interactions that limit its flexibility (Fig. 4A–B). POT1 Ser322 makes the furthest N-terminal self-interaction with POT1OB-3 with Arg367 at site (i) to create a linchpin that locks aa 322–341 in place (Fig. 4B; Fig. S6A). Ser322 is mutated to leucine in CP23, which would disrupt the hydrogen-bond interaction. Indeed, mutation of the corresponding Ser326 in POT1(ESDL) to alanine or leucine severely attenuates CST binding (Fig. 4C–D). This interaction is conserved, as the corresponding S328L mutation in mPOT1b abrogates CST binding (Fig. S6B). The S322A substitution also abolishes binding of the phosphomimetic CCIR mutants (Fig. 4C), consistent with the linchpin role of Ser322. At site (ii), aa 335–337 of the hinge interact with the C-terminus of POT1 (aa 627–629) in an anti-parallel β-strand configuration, as previously observed39 (Fig. 4B; Fig. S6A). In the context of the full complex, these two points of contact secure the otherwise flexible hinge region against POT1OB-3 (Fig. 4A–B).
Figure 4. CP mutations map to the primary Ctc1–POT1 interface.
See also Figure S6.
(A) Comparison of CST-bound POT1OB-3/TPP1 (this study, colored) to unbound POT1OB-3/TPP1 (PDB 5H65/5UN738, 39, grayscale).
(B) Close-up views of sites of interest at or near the interface between POT1 and Ctc1 (see also Fig. S6A). Residues colored in bright red are mutated in CP. (i) Self-interaction between POT1 Ser322 and Arg367. Hydrogen bonds are shown as yellow dashed lines. CP mutation S322L is predicted to disrupt this linchpin interaction. (ii) Self-interaction between POT1 hinge and C-terminus. (iii) Position of Gly503 in the hydrophobic core of Ctc1ARODL predicts a destabilizing effect of the G503R CP mutation. (iv) Primary hydrophobic interface between POT1 hinge, POT1OB-3, and Ctc1ARODL. CP mutation H484P is predicted to disrupt the hydrophobic stacking interactions with POT1 Pro603 and Ctc1 His488 and Pro483. (v) Salt bridge between POT1 hinge residue Glu325 and Ctc1OB-D residue Arg624.
(C) Co-IPs of Myc-tagged POT1 protein variants and FLAG-tagged Ctc1 from 293T cells transfected with Myc-POT1, FLAG-TPP1, FLAG-Ctc1, FLAG-Stn1, and FLAG-Ten1. Western blots of Myc IPs were probed with anti-Myc and anti-FLAG antibodies to detect POT1 and Ctc1, respectively. Short and long exposures of the anti-FLAG probing for Ctc1 are shown.
(D) Co-IPs of Myc-tagged POT1(ESDL) and FLAG-tagged Ctc1 WT and CP variants from 293T cells transfected with Myc-POT1, FLAG-TPP1, FLAG-Ctc1, FLAG-Stn1, and FLAG-Ten1. POT1(ESDL) CP contains the S326L mutation, analogous to S322L in POT1CP (see (E)). Western blots of Myc IPs were probed with anti-Myc and anti-FLAG antibodies to detect POT1 and Ctc1, respectively. Short and long exposures of the anti-FLAG probing for Ctc1 are shown.
(E) Sequences of CCIR mutants and their interaction with CST as shown in (C-D).
Our structure also provides insight into the Ctc1 G503R CP mutation17, 41, 42. Although Gly503 is not at the Ctc1–POT1 interface, it is in the hydrophobic core of Ctc1ARODL at site (iii). A G-to-R substitution would destabilize Ctc1ARODL, which forms the primary interaction with POT1 (Fig. 4B; Fig. S6A). A loop in Ctc1ARODL, which interacts with POT1, is the location of another CP mutation, H484P24. At site (iv), His484 participates in a network of hydrophobic stacking interactions with Ctc1 Pro483, Ctc1 His488, and POT1 Pro603. This interface is stabilized by hydrophobic packing of the Ctc1ARODL loop against the POT1 tail (aa 630–634), POT1OB-3, and its hinge (Fig. 4B). Both the G503R and H484P substitutions in Ctc1 abrogate binding to POT1(ESDL) (Fig. 4D).
The end of the POT1 tail (632-DVI-634 in humans) was previously implicated in CST binding by mPOT1b5. Substitution of the sequence in mPOT1b (DII) to the mPOT1a sequence (NVV) diminished the interaction with CST5, and our structure explains the importance of the bulky hydrophobic residues at this site. The structure also shows that POT1 Asp632 is at an appropriate distance to hydrogen bond with Ctc1 His489, which would further stabilize the complex (Fig. 4B; Fig. S6A). Finally, there is an additional salt bridge between POT1 Glu325 and Ctc1 Arg624 at site (v), which are in the POT1 hinge and Ctc1OB-D, respectively (Fig. 4B; Fig. S6A).
POT1 blocks the CST ssDNA anchor site
The second OB fold of POT1, POT1OB-2, also interacts with CST (Fig. 2C, E). The resolution of the CST–POT1OB-2 interface was in the 6-Å range (Fig. S3H; Fig. S4H) but still allowed for visualization of the docked structures. POT1OB-2 sits between Ctc1 and the C-terminal lobe of Stn1 (Stn1C) at Ctc1’s ssDNA anchor site13 (Fig. 5A). The POT1 ssDNA-binding interface is facing outwards and is bound to the [GGTTAG]3 ssDNA (Fig. 2E). Helix αD of POT1 directly superposes with the ssDNA observed in the cryo-EM structure of CST13 and the N-terminal part of the POT1 hinge runs through the ssDNA anchor site towards the basic cleft in Ctc1OB-D (Fig. 5A; Fig. 3A). Stn1C, which can bind Ctc1 in multiple configurations, is in a position consistent with the structure of monomeric CST in the absence of ssDNA13 (Fig. 5A). There was additional low-resolution density of a flexible loop in Ctc1 (aa 909–927), which has not been resolved in other structures of CST, and we used poly-alanine stubs to indicate its presence. The low resolution indicates that POT1OB-2 may only weakly interact with this loop (Fig. 5A).
Figure 5. POT1OB-2 interacts with the CST ssDNA anchor site, precluding formation of the CST–Polα/primase PIC.
See also Figure S7.
(A) Interaction between POT1OB-2 and Ctc1 at the CST ssDNA anchor site. (Left) The black box indicates the region of the apo CST–POT1(ESDL)/TPP1 model shown in the right panels. (Right) Zoomed-in views of the interface. The first panel shows the ssDNA-bound CST structure (PDB 6W6W13, gray with DNA colored cyan). The second panel shows the same view of the CST–POT1(ESDL)/TPP1 structure (this study, colored). Ctc1 aa 909–927 (density map shown; contour threshold 0.15) are modeled as poly-alanine stubs, and no ssDNA is bound to Ctc1. The two structures are superposed in panel three.
(B) Negative-stain EM 2D averages of ssDNA-bound CST–POT1(ESDL)/TPP1 (left) showing three major conformations depicted by the cartoon schematics (right). 2D averages are flipped or rotated by 90° increments from Fig. S2E into similar orientations for ease of comparison and are sorted by number of particles per class from the most populous class at top left to the least populous class at bottom right. The 2D averages that correspond to each cartoon state are indicated with black, gray, or white circles. The scale bar represents 333 Å (See also Fig. S2E).
(C) The CST–POT1(ESDL)/TPP1–ssDNA complex (left) and its superposition with the CST–Polα/primase recruitment complex (RC, PDB 7U5C15, middle) and pre-initiation complex (PIC, PDB 8D0B16, right). The CST–POT1(ESDL)/TPP1–ssDNA complex is shown as an opaque surface, and CST–Polα/primase complexes are shown in cartoon representation with transparent surfaces. Orthogonal views show that POT1(ESDL)/TPP1 binding does not obstruct the major interface of the RC, but POT1OB-1/2 and Stn1C would interfere with binding of the POLA1 catalytic core to the CST ssDNA anchor site in the PIC (see also Fig. S7A–B).
(D) Cartoon schematics showing how CST–POT1/TPP1 is incompatible with PIC formation but could form RC-like complexes in both monomeric and dimeric forms.
The particles of the ssDNA-bound CST–POT1(ESDL)/TPP1 dataset were more heterogenous than those in the apo dataset; a subset of particles appeared to be missing density for the POT1 N-terminal OB-folds and Stn1C (Fig. S2E; Fig. S4D). Loss of this density suggests that ssDNA may be bound to Ctc1, in contrast to the major conformation described above. POT1 has a greater affinity for telomeric ssDNA than CST43, 44 (Fig. S6C) and is thus expected to outcompete CST for access to the ssDNA in limiting conditions. The CST–POT1(ESDL)/TPP1 fusion complex bound the telomeric [GGTTAG]3 ssDNA with the same high affinity (KD,app ~100 pM) as POT1(ESDL)/TPP1, indicating that the presence of CST did not change the ssDNA-binding affinity of POT1 (Fig. S6C). However, the reconstitution was performed with an excess of [GGTTAG]3 that could result in both proteins being bound to DNA. We could not obtain a high-resolution map for the conformation in which both CST and POT1 were bound to ssDNA, likely because monomeric ssDNA-bound CST is flexible13. The negative-stain EM data raised the possibility that this conformation is an intermediate in the dimerization of CST–POT1/TPP1 (Fig. 5B). CST can dimerize in the presence of ssDNA13, and CST dimers bound by POT1/TPP1 are observed in negative-stain EM 2D averages (Fig. 5B; Fig. S2E). CST dimerization requires displacement of POT1OB-2 and Stn1C13, but the POT1 C-terminus appears to be still attached by the POT1OB-3 and CCIR interactions. The dimer interface observed for CST–POT1/TPP1 appears similar to that of CST alone, though we note that the position of the POT1 C-terminus interaction would block the previously observed downstream tetramerization and decamerization of CST13. The CST–POT1(ESDL)/TPP1 dimers, which were a minor fraction of the particles observed by negative-stain EM (~6%), could not be found in the cryo-EM dataset (Fig. S4), indicating that they may not withstand the vitrification process.
POT1/TPP1 can recruit CST–Polα/primase in the RC but not in the PIC state
The ssDNA anchor site of CST is critical for formation of the CST–Polα/primase PIC in which the POLA1 catalytic core contacts the telomeric ssDNA template and is stabilized by Stn1C, which is in a different position relative to the monomeric and DNA-bound states of CST16. POT1/TPP1 binding to CST is incompatible with Polα/primase binding in a PIC-like conformation when POT1OB-1, POT1OB-2, and Stn1C are engaged (Fig. 5C; Fig. S7A–B; Video S2). Therefore, we propose that POT1/TPP1 does not bind CST–Polα/primase in the PIC conformation. In contrast, the major interface between Ctc1 and Polα/primase in the auto-inhibited RC-like conformation is orthogonal to the POT1/TPP1 interface and is unobstructed, thus allowing for the formation of a POT1/TPP1-bound RC-like complex (Fig. 5C–D; Fig. S7A–B; Video S3). There is a minor clash between PRIM2 and POT1HJRL, but the cryo-EM structure of the RC was determined in cross-linking conditions, limiting the flexibility of the complex15. This clash could be alleviated by the flexibility of CST relative to Polα/primase observed in multi-body refinement analysis45 of the CST–Polα/primase RC complex (Fig. S7C; Videos S4). Structural heterogeneity of CST–Polα/primase in an RC-like state was previously observed and proposed to be partially attributed to binding of the POLA2 N-terminal domain (POLA2NTD), which is attached by a flexible linker to POLA215. Indeed, AlphaFold-Multimer predicts an interaction between the POLA2NTD and Ctc1 (Fig. S7D–E), supporting the existence of flexible RC-like states observed in low-resolution cryo-EM maps15.
We propose that the OB-2-disengaged state, when telomeric ssDNA is not limiting, represents a transient intermediate to the formation of a dimer of CST–POT1/TPP1 (Fig. 5B). A CST–POT1/TPP1 dimer is likely to form at the telomere in the context of shelterin, which is fully dimeric and hence contains two copies of POT1/TPP1 each capable of binding to CST46. The high local concentrations of two CST molecules brought together by a shelterin dimer would favor dimerization. CST dimerization would also prevent PIC formation16 and is compatible with the RC15 (Fig. 5D).
POT1 bound to CST inhibits CST–Polα/primase C-strand synthesis in vitro
Our structural model predicts that the purified POT1(ESDL), which is competent for CST binding compared to POT1(WT), will inhibit C-strand synthesis by CST–Polα/primase by limiting formation of an active PIC. Previous experiments showed that POT1(WT)/TPP1 produced in E. coli did not have a strong inhibitory effect on CST–Polα/primase activity47, but our data indicate that this protein complex is unable to interact with CST. Thus, we used our insect cell-derived POT1(WT)/TPP1 and POT1(ESDL)/TPP1 to directly assay the effect of POT1 binding on inhibition of active CST–Polα/primase. As previously described, when CST–Polα/primase co-purified from HEK293T cells was incubated with a synthetic telomeric G-strand template (9xTEL or 15xTEL, representing 9x or 15x of ss TTAGGG repeats), a characteristic 6-nt repeat pattern of products was produced47 (Fig. 6A–B). Extension on the 9xTEL template was only inhibited by POT1(WT)/TPP1 at high concentrations (10-fold excess over template and CST–Polα/primase) as previously observed47. On the other hand, POT1(ESDL)/TPP1 was a strong inhibitor of the C-strand synthesis reaction with an IC50 7-fold lower than that of the WT protein. The preferential inhibition of C-strand synthesis by POT1(ESDL)/TPP1 was also clear on the 15xTEL template, although higher concentrations were required on the longer template containing more binding sites (Fig. 6A–B).
Figure 6. CST binding enhances POT1/TPP1 inhibition of telomeric C-strand synthesis.
See also Figure S8.
(A) DNA synthesis by 50 nM CST–Polα/primase on 50 nM ssDNA templates consisting of 9 or 15 telomeric TTAGGG repeats (9xTEL and 15xTEL, respectively). Products labeled with α-32P-dATP. LC3: an oligonucleotide loading control.
(B) Quantification of the data in (A), normalized to reactions without added POT1/TPP1. IC50 +/− uncertainty of the fit to the data points. An independent repeat of the 9xTEL experiment gave equivalent results (Fig. S8A–C).
(C) FSEC analysis of the interaction between CST and POT1(ESDL ΔOB1)/GFP-TPP1 in the absence of telomeric ssDNA. The trace without CST is shown as a dashed line. RFU: relative fluorescence units.
(D) Design of pre-primed templates to monitor primer extension by Polα. RNA sequence is shown in red.
(E) CST–Polα extension of pre-primed templates. M: telomerase reaction products as size markers. The two panels have equal exposure, but one-ninth as much dT42–2xTEL product as 9xTEL product was loaded to compensate for the higher incorporation with the former template. LC1, LC2, LC3: three oligonucleotide loading controls.
(F) Quantification of the experiment shown in (E) (left). IC50 values include replicate experiments on separate days, with errors encompassing the range of values obtained.
(G) Quantification of the experiment shown in (E) (right). Only partial inhibition occurred at the highest POT1/TPP1 concentrations, so IC50 values were not calculated.
To test the role of the ssDNA-binding activity of POT1, we generated a POT1(ESDL ΔOB1) construct missing POT1OB-1 (aa 1–145), rendering it unable to bind DNA48. Consistent with our cryo-EM structures showing that POT1OB-1 does not directly interact with CST (Fig. 2), deletion of this domain had no effect on complex formation between POT1(ESDL ΔOB1)/TPP1 and CST (Fig. 6C). POT1(ESDL ΔOB1)/TPP1 only weakly inhibited C-strand synthesis (Fig. S8A–C).
To isolate the effect of POT1 on CST and Polα in a primase-independent setting, we designed pre-primed DNA templates (Fig. 6D; Fig. S8D). The RNA primer length and sequence were as previously determined47, and the RNA primer was made as a chimera with a 12-nt DNA to specify its binding site on the DNA template by base pairing. Extension by CST–Polα/primase no longer required ribonucleotides, confirming that the reaction was now independent of primase activity (Fig. S8E). Polα extension of the primer on the pre-primed telomeric DNA template was again inhibited most strongly by POT1(ESDL)/TPP1, with an IC50 4–5-fold lower than for POT1(WT)/TPP1 (Fig. 6E–F). Notably, the IC50 for Polα extension was considerably higher (weaker inhibition) than for reactions initiated by primase (Fig. 6B), consistent with greater inhibition occurring at the initiation step. POT1(ESDL ΔOB1)/TPP1 was too weak of an inhibitor to measure IC50 (Fig. 6E–F), suggesting that telomeric DNA binding is important for inhibition by POT1/TPP1.
To confirm that DNA binding was important for POT1/TPP1 inhibition of C-strand synthesis, we substituted the telomeric repeats of the pre-primed template with poly(dT) (Fig. 6D). This pre-primed template resulted in much stronger synthesis due to the lack of inhibitory G-quadruplex folding47. On the poly(dT) template, which is poorly bound by both POT1/TPP1 and CST40, 44, 49, all three POT1/TPP1 complexes were similarly weak inhibitors of DNA synthesis (Fig. 6E; Fig. 6G). Thus, telomeric DNA binding, in addition to CST binding, contributes strongly to the inhibition of C-strand synthesis by POT1/TPP1. The enhanced inhibition of POT1(ESDL)/TPP1 also occurred with 10-fold higher DNA template to CST–Polα/primase ratios, where re-priming is repressed and the extension is limited by the inherent processivity of the polymerase (Fig. S8F–G). In summary, the markedly greater inhibition of C-strand synthesis by POT1(ESDL)/TPP1 in vitro provides functional support for the structure-based model where POT1 prevents formation of the CST–Polα/primase PIC.
Discussion
Decades of research have revealed how shelterin prevents inappropriate DNA damage signaling and DNA repair at telomeres1. Similarly, extensive studies have established the role of shelterin in the recruitment and regulation of telomerase, the enzyme that solves part of the end-replication problem by extending the G-rich telomeric repeat strand9. The data presented here illuminate the third function of shelterin, which is to promote and regulate the synthesis of the C-rich telomeric repeats by CST–Polα/primase, the second enzyme needed to solve the end-replication problem4 (Fig. 7A).
Figure 7. Model for telomere maintenance by shelterin and CST–Polα/primase.
(A) Shelterin-mediated telomere protection and maintenance. Shelterin recruits and regulates two telomere maintenance enzymes: telomerase and CST–Polα/primase. Adapted from2.
(B) Model for the recruitment and regulation of CST–Polα/primase by POT1. CST–Polα/primase is recruited to telomeres in the auto-inhibited RC-like state by phosphorylated POT1, preventing activation of CST–Polα/primase during 5’ end resection and telomerase action following canonical DNA replication. After telomerase has elongated telomeres, a proposed switch results in dephosphorylation of POT1 and subsequent release of CST–Polα/primase to the telomeric ssDNA. When released, CST–Polα/primase can readily form the PIC and execute fill-in synthesis, the final step of telomere replication.
The structures of CST bound to POT1/TPP1 provide insights into the mechanism by which POT1 recruits and regulates CST–Polα/primase (Fig. 7B). The complex formed by POT1/TPP1 and CST is compatible with the binding of CST to Polα/primase in the auto-inhibited state, which we previously proposed to be the recruitment state of CST–Polα/primase. In contrast, the structure of CST bound to POT1/TPP1 is not compatible with the active state of the fill-in machinery, and our biochemical data confirm that POT1/TPP1-binding inhibits the ability of CST–Polα/primase to form an active fill-in enzyme. Therefore, CST–Polα/primase must undergo a major conformational switch at telomeres and its release from the POT1/TPP1 subunits of shelterin is required for this switch. The data further reveal the critical role of phosphorylation of human POT1 in regulating the recruitment and release of CST–Polα/primase. Our structural and biochemical data suggest that phosphorylation in a short Ser-rich part of the POT1 hinge (the CCIR) promotes insertion of the CCIR into a basic cleft of Ctc1, thereby allowing complex formation and recruitment of CST–Polα/primase to telomeres (Fig. 7B).
We propose that phosphorylated POT1 positions CST–Polα/primase at telomeres and holds the complex in an inactive state until the replication machinery has completed DNA synthesis, 5’ end-resection has taken place, and telomerase has extended the leading- and lagging-end telomere termini. At that point, we imagine that dephosphorylation of POT1 releases CST–Polα/primase so that the enzyme can transition into the active state and execute fill-in synthesis as the last event in the telomere end processing reactions (Fig. 7B). This scenario is compatible with data obtained on the order of processing events at human telomeres (6, 50, 51; reviewed by Cai and de Lange2).
The ability of POT1 to hold CST–Polα/primase in an inactive state until telomerase has initiated G-strand extension may be needed to ensure that CST does not prematurely inhibit telomerase. CST can block telomerase from reinitiating G-strand synthesis in vitro52 and loss of CST leads to inappropriate telomere extension by telomerase in vivo28. This ability of CST to terminate telomerase-mediated telomere extension requires its ssDNA-binding activity52. Since POT1 blocks the CST ssDNA anchor site, we predict that POT1-bound CST is an ineffective telomerase inhibitor. Release of CST–Polα/primase from POT1 would therefore both initiate fill-in synthesis and terminate further telomere extension by telomerase.
Human POT1 has at least two critical functions. First, it represses the activation of ATR signaling at telomeres by blocking the binding of RPA to the ssDNA53–55 and by preventing the Rad17/9-1-1 complex from engaging the telomeric 5’ end56. Second, as shown here, POT1 promotes the maintenance of telomeric DNA through its ability to recruit CST–Polα/primase. In most metazoan shelterin complexes, a single POT1 must execute both functions57. In contrast, a rodent lineage that includes mice and rats has duplicated the POT1 locus and evolved two distinct POT1 proteins, one of which is dedicated to ATR repression (POT1a) and the other to CST–Polα/primase recruitment (POT1b). Interestingly, mouse POT1b brings CST–Polα/primase to telomeres throughout the cell cycle5 whereas in human cells, CST peaks at telomeres in late S/G228. Perhaps the presence of POT1a allows POT1b to permanently associate with CST whereas the ancestral POT1 needs to be regulated by phosphorylation to toggle between a CST-bound form and a form that can block ATR signaling.
Our results show how a subset of CP mutations affect telomeric CST recruitment, explaining the POT1 S322L genetic data and providing further evidence for the causative link between telomere maintenance defects and CP pathogenesis. Four CP mutations map to regions involved in CST–Polα/primase recruitment: Ctc1 V665G maps to the primary interface observed in the RC, and POT1 S322L, Ctc1 H484P, and Ctc1 G503R affect CST–POT1 complex formation. Furthermore, CP mutations in Ctc1OB-B are numerous and potentially affect the predicted interactions between Ctc1 and the POLA2NTD, though further experiments are needed to confirm the AlphaFold-Multimer predictions. Together, these CP mutations highlight CST–Polα/primase recruitment as a critical step in normal telomere maintenance. Once the kinase(s) responsible for regulation of this step is identified, it is possible that additional patient mutations will be discovered within the kinase or its regulatory factors.
These findings open another avenue for research on the regulation of telomere maintenance. Whereas telomere elongation by telomerase has been studied for more than 30 years, examination of the equally important synthesis of the telomeric C-strand has lagged (pun intended). We expect that the identification of the kinases/phosphatases that control CST recruitment and its subsequent release will reveal how the cell cycle controls the critical last step in telomere maintenance.
Limitations of the study
Here, we propose a model for the regulated recruitment of CST–Polα/primase by the shelterin subunit POT1 based on our structural and biochemical data. As discussed above, further in vivo work will be required to determine the kinase, phosphatase, and exact cell cycle timing of this process. Because the kinase was not known, we did not formally show native phosphorylation of human POT1 in this study, though our findings are highly suggestive. Instead, we used a modified POT1(ESDL) construct to increase the stability of the CST–POT1/TPP1 complex for structural studies. The phosphomimetic mutations in POT1 confer similar binding, but they are slightly weaker interactors. We speculate that this could be a feature, not a bug, of the human system compared to that in the mouse, as discussed above. To confirm this, additional biochemical analysis using a natively phosphorylated human POT1 is required (once the kinase is identified). Similarly, the TPP1TID was truncated to limit structural heterogeneity, but given the proximity of many CP mutations to the Ctc1OB-B–TPP1TID interface, it will be interesting to continue to explore the functional role of this interaction. Finally, although the in vitro C-strand synthesis data are highly suggestive of our model, the structural information provided here are limited to CST and POT1. Additional study is needed to define the structure of the proposed POT1–CST–Polα/primase ternary recruitment complex.
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, Titia de Lange (delange@rockefeller.edu).
Materials Availability
Unique reagents generated in this study are available from the lead contact with a completed Materials Transfer Agreement.
Data and Code Availability
The cryo-EM maps have been deposited at the Electron Microscopy Data Bank under accession codes EMD-40659 (apo CST–POT1/TPP1 complex) and EMD-40660 (ssDNA–CST–POT1/TPP1 complex), and the coordinates have been deposited in the Protein Data Bank under accession codes PDB 8SOJ and PDB 8SOK, respectively. Motion-corrected micrographs have been deposited in the Electron Microscopy Public Image Archive59 under accession codes EMPIAR-12046 and EMPIAR-12047, respectively. This study also uses data previously reported2 and deposited under the accession code EMPIAR-11131. The data are publicly available as of the date of publication.
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
All cell lines and source materials used in this study are listed in the key resources table.
KEY RESOURCES TABLE.
REAGENT or RESOURCE | SOURCE | IDENTIFIER |
---|---|---|
Antibodies | ||
Rabbit monoclonal Anti-DYKDDDDK Tag (D6W5B) antibody | Cell Signaling Technology | Cat#14793; RRID:AB_2217020 |
Mouse monoclonal Anti-c-Myc (9E10) antibody | Sigma-Aldrich | Cat#M4439; RRID:AB_439694 |
Mouse monoclonal Anti-Strep-tag antibody | IBA Lifesciences | Cat#2-1507-001; RRID:AB_513133 |
Mouse monoclonal Anti-HA.11 antibody (16B12) | Covance | Cat#MMS-101R; RRID:AB_291262 |
Bacterial and Virus Strains | ||
E. coli: MAX Efficiency™ DH10Bac Competent Cells | Thermo Fisher | Cat#10361012 |
E. coli: One Shot™ TOP10 Chemically Competent E. coli | Thermo Fisher | Cat#C404003 |
Chemicals, Peptides, and Recombinant Proteins | ||
Cellfectin® II Reagent | Thermo Fisher | Cat#10362100 |
DMEM | Corning | Cat#10-013-CV |
HyClone Bovine Calf Serum (BCS) | Cytiva | Cat#SH30073.03 |
L-Glutamine (200 mM) | Gibco | Cat#25030081 |
Penicillin-Streptomycin (10,000 U/mL) | Gibco | Cat#15140122 |
MEM Non-Essential Amino Acids Solution (100X) | Gibco | Cat#11140050 |
Phosphate-Buffered Saline, 1X without calcium and magnesium, pH 7.4 ± 0.1 | Corning | Cat#21-040-CM |
cOmplete™, EDTA-free Protease Inhibitor Cocktail | Roche | Cat#11873580001 |
PhosSTOP™ | Roche | Cat#4906845001 |
Pierce™ Anti-c-Myc Magnetic Beads | Thermo Fisher | Cat#88842 |
Benzonase® Nuclease | Millipore Sigma | Cat#E1014 |
NuPAGE™ LDS Sample Buffer (4X) | Thermo Fisher | Cat#NP0007 |
Biotinylated peptides | Kinexus Bioinformatics Co. | N/A |
ATP, [γ-32P] | Perkin Elmer | Cat#BLU502Z |
Strep-Tactin® Superflow® high capacity resin | IBA Lifesciences | Cat#2-1208-025 |
Lambda protein phosphatase (expressed and purified in-house from E. coli) | N/A | UNIPROT: P03772 |
Fos-Choline-8, Fluorinated, Anagrade | Anatrace | Cat#F300F |
T4 Polynucleotide Kinase | New England Biolabs | Cat#M0201 |
Hybond®-XL hybridization membranes | GE Healthcare | Cat#GERPN2020S |
Deposited Data | ||
Cryo-EM data of CST–Polα/primase recruitment complex | Cai et al., 2022 | EMPIAR-11131 |
Cryo-EM map of apo CST–POT1/TPP1 complex | This paper | EMD-40659 |
Cryo-EM map of ssDNA–CST–POT1/TPP1 complex | This paper | EMD-40660 |
Coordinates for apo CST–POT1/TPP1 complex | This paper | PDB8SOJ |
Coordinates for ssDNA–CST–POT1/TPP1 complex | This paper | PDB8SOK |
Motion-corrected micrographs for apo CST–POT1/TPP1 complex dataset | This paper | EMPIAR-12046 |
Motion-corrected micrographs for ssDNA–CST–POT1/TPP1 complex dataset | This paper | EMPIAR-12047 |
Experimental Models: Cell Lines | ||
HEK 293T/17 | ATCC | Cat#CRL-11268; RRID:CVCL_1926 |
Sf9 | Thermo Fisher | Cat#11496015; RRID:CVCL_0549 |
Tni | Expression Systems | Cat#94-002F; RRID:CVCL_RY32 |
HEK 293T | ATCC | Cat#CRL-3216; RRID:CVCL_0063 |
HeLa1.3 | Takai et al., 2010 | RRID:CVCL_B5DA |
Oligonucleotides | ||
d[GGTTAG]3 oligonucleotide | Thermo Fisher | N/A |
d[TTAGGG]9 9xTEL oligonucleotide | IDT | N/A |
d[TTAGGG]15 15xTEL oligonucleotide | IDT | N/A |
d[T]42[TTAGGG]2CTGGAACCTAGT top strand for dT42-2xTEL pre-primed template | IDT | N/A |
[TTAGGG]9CTGGAACCTAGT top strand for 9xTEL pre-primed template | IDT | N/A |
d[ACTAGGTTCCAGCCCT]r[AACCCUAA] bottom strand for pre-primed template | IDT | N/A |
Recombinant DNA | ||
See Table S1 for a list of recombinant DNA constructs | N/A | N/A |
Software and Algorithms | ||
EMAN2 v2.06 | Tang et al., 2007 | https://blake.bcm.edu/emanwiki/EMAN2 RRID:SCR_016867 |
RELION-3.1 | Zivanov et al., 2018 | http://www2.mrc-lmb.cam.ac.uk/relion RRID:SCR_016274 |
SerialEM | Mastronarde, 2005 | http://bio3d.colorado.edu/SerialEM/ RRID:SCR_017293 |
cryoSPARC v4.1.1 | Punjani et al., 2017 | https://cryosparc.com/ RRID:SCR_016501 |
PHENIX v1.20.1 | Adams et al., 2010 Liebschner et al., 2019 | https://www.phenix-online.org/ RRID:SCR_014224 |
3DFSC | Tan et al., 2017 | https://3dfsc.salk.edu/ |
UCSF Chimera v1.13.1 | Pettersen et al., 2004 | https://www.cgl.ucsf.edu/chimera/ RRID:SCR_004097 |
UCSF ChimeraX v1.5 | Pettersen et al., 2021 | https://www.cgl.ucsf.edu/chimerax/ RRID:SCR_015872 |
AlphaFold Protein Structure Database | Jumper et al., 2021 Varadi et al., 2021 | https://alphafold.ebi.ac.uk/ RRID:SCR_023662 |
Coot v0.9.8.7 | Emsley et al., 2010 | http://www2.mrc-lmb.cam.ac.uk/personal/pemsley/coot/ RRID:SCR_014222 |
ISOLDE v1.5 | Croll, 2018 | https://tristanic.github.io/isolde/ |
AlphaFold-Multimer v1.0 | Evans et al., 2022 | https://github.com/google-deepmind/alphafold |
COSMIC2 | Cianfrocco et al., 2017 | https://cosmic-cryoem.org/ |
AcquireMP v2022R1 | Refeyn | https://www.refeyn.com/ |
DiscoverMP v2022R1 | Refeyn | https://www.refeyn.com/ |
ImageJ | Schneider et al., 2012 | https://imagej.net/ RRID:SCR_003070 |
GraphPad Prism v9.5.1 | GraphPad | http://www.graphpad.com/ RRID:SCR_002798 |
ImageQuant | GE Life Sciences | http://www.gelifesciences.com/webapp/wcs/stores/servlet/catalog/en/GELifeSciences-us/products/AlternativeProductStructure_16016/29000605 RRID:SCR_014246 |
KaleidaGraph | Synergy Software | http://www.synergy.com/ RRID:SCR_014980 |
PyMOL v2.5.4 | Schrödinger | http://www.pymol.org/ RRID:SCR_000305 |
Adobe Illustrator | Adobe | http://www.adobe.com/products/illustrator.html RRID:SCR_010279 |
Gautomatch v0.53 | MRC LMB (Kai Zhang) | https://www2.mrc-lmb.cam.ac.uk/download/gautomatch-053/ |
CTFFIND-4 | Rohou and Grigorieff, 2015 | http://grigoriefflab.janelia.org/ctffind4 RRID:SCR_016732 |
MotionCor2 | Zheng et al., 2017 | https://emcore.ucsf.edu/cryoem-software RRID:SCR_016499 |
Other | ||
Graphene Oxide on Quantifoil® R1.2/1.3 Cu, 400 mesh TEM grid | Electron Microscopy Sciences | Cat#GOQ400R1213Cu100 |
Amersham™ Typhoon™ NIR | Cytiva | Cat#29238583 |
HiLoad Superdex 200 16/600 PG | Cytiva | Cat#28989335 |
Superose 6 Increase 10/300 GL | Cytiva | Cat#29091596 |
HiTrap Heparin HP 1 mL | Cytiva | Cat#17040601 |
SW 41 Ti Swinging-Bucket Rotor | Beckman Coulter | Cat#331362 |
METHOD DETAILS
DNA construct generation
The DNA constructs used in this study are listed in Table S2. For recombinant insect cell expression, the biGBac60 vector system was used. All protein sequences are numbered according to the canonical sequence entry in UniProt. Recombinant bacmids were generated from the plasmids in Table S2 using MAX Efficiency DH10Bac competent cells (Thermo Fisher Scientific) and transfected into Sf9 insect cells (Thermo Fisher Scientific) with Cellfectin II Reagent (Thermo Fisher Scientific) to generate a P1 baculovirus stock. P1 baculovirus was amplified in adherent Sf9 insect cells to generate P2 and P3 stocks, and the P3 virus was used to infect Tni suspension insect cell culture (Expression Systems) for protein expression.
Co-immunoprecipitation
293T cells were grown in DMEM with 10% (v/v) bovine calf serum (HyClone), 2 mM L-glutamine (Gibco), 100 U/mL penicillin (Gibco), 0.1 mg/mL streptomycin (Gibco), and 0.1 mM nonessential amino acids (Gibco). 2×106 293T cells were plated in a 10-cm dish at 20–24 h prior to transfection by calcium-phosphate precipitation. Transfections were done using plasmid DNA as indicated such that total DNA did not exceed 20 μg per plate. Each experiment combined two identical 10-cm plates. At 16 h post-transfection, the medium was changed, and 24 h later, cells were harvested by flushing with 1x phosphate-buffered saline (PBS; Corning), frozen in liquid nitrogen (LN2), and resuspended in 1 mL lysis buffer (50 mM Tris-HCl pH 7.5, 150 mM NaCl, 1 mM EDTA, 10% (v/v) glycerol, 0.1% (v/v) NP-40, cOmplete EDTA-free protease inhibitor cocktail (Roche), and PhosSTOP phosphatase inhibitor mix (Roche)), and incubated on ice for 30 m. After centrifugation at 16,000 rcf for 10 m at 4°C, input samples were taken from the supernatant. 20 μL of equilibrated Pierce Anti-c-Myc magnetic beads slurry (Thermo Fisher) and 1 μL benzonase (Millipore Sigma) were added to the supernatant, and samples were incubated with end-over-end rotation for 2–3 h at 4°C. Beads were washed 3 times at 4°C with lysis buffer and immunoprecipitated proteins were eluted with 100 μL of 1x NuPAGE LDS buffer (Thermo Fisher). Samples were boiled for 5 m before separation on SDS-PAGE.
Peptide phosphorylation assay
The phosphorylation assay was performed as previously described61 with some modifications. Biotinylated peptides of the POT1 protein CCIR regions (POT1 aa 313–327; mPOT1b aa 319–333; mPOT1a aa 319–333; POT1(ESDL) aa 313–331) were produced by Kinexus Bioinformatics Co. HeLa cells cultured in the same medium as the 293T cells were used for isolation of nuclear extract61. 6×107 cells were trypsinized, rinsed with cold PBS, and washed twice with 2 mL cytoplasm buffer (10 mM Tris-HCl pH 7, 10 mM NaCl, 3 mM MgCl2, 30 mM sucrose, and 0.5% NP-40) on ice. After centrifugation, the pellet was rinsed twice with 2 mL of CaCl2 buffer (10 mM Tris-HCl pH 7, 10 mM NaCl, 3 mM MgCl2, 30 mM sucrose, and 100 μM CaCl2). The pellet was resuspended in 1 mL buffer C (20 mM Tris-HCl pH 7.9, 20% glycerol, 100 mM KCl, and 0.2 mM EDTA), and centrifuged at 20,000 rcf for 30 m at 4°C. The supernatant was collected, stored at −80°C, and used as nuclear extract for the kinase assay. The nitrocellulose membrane containing the spotted peptide array was rinsed once with reaction buffer (20 mM Tris-HCl pH 7.9, 50 mM KCl, 10 mM MgCl2, 1 mM DTT, cOmplete EDTA-free protease inhibitor cocktail, PhosSTOP) and subsequently incubated in 4 mL of reaction buffer containing 50 μM ATP, 50 μCi/μmol [γ-32P] ATP, and nuclear extract (16 μg protein) for 45 m at 30°C. EDTA (15 mM final concentration) was added to stop the reaction, and the membrane was rinsed 10 times with cold 1 M NaCl. The membrane was rinsed once with wash buffer (4 M guanidine, 1% (v/v) SDS, 0.5% 2-mercaptoethanol (β-ME)) and exposed for radiography (Typhoon Biomolecular Imager, GE Healthcare).
Protein expression and purification
50 mL of P3 baculovirus was used per 500 mL of Tni culture, infected at a cell density of 2×106 cells/mL. Infected cells were grown in spinner flasks at 150 rpm for 72 h at 27°C. Cells were harvested by centrifugation (500 rcf) and transferred to a syringe before flash freezing droplets in LN2.
CST was purified as previously described15. For CST, frozen pellets were lysed by cryogenic milling (Retsch) and the cryo-milled powder was resuspended in a buffer containing 20 mM Tris pH 8.0, 500 mM NaCl, 15 mM β-ME, 20 mM imidazole, 10% glycerol, and 1 mM phenylmethylsulfonyl fluoride (PMSF), supplemented with cOmplete EDTA-free protease inhibitor cocktail (Roche). The lysate was cleared by centrifugation at 40,000 rcf for 1 h at 4°C. Supernatants were incubated with end-over-end rotation for 1 h at 4°C with Ni-NTA resin (Invitrogen) equilibrated with a buffer containing 20 mM Tris pH 8.0, 500 mM NaCl, 15 mM β-ME, 20 mM imidazole, and 5% glycerol and subsequently washed with 20–50 column volumes (CV) of the same buffer. Bound protein was eluted in a buffer containing 20 mM HEPES (pH 7.5), 150 mM NaCl, 0.5 mM tris(2-carboxyethyl)phosphine (TCEP), 250 mM imidazole, and 5% glycerol. After elution from the Ni-NTA resin, His6-MBP-PreSc-Ctc1/Stn1/Ten1 was loaded onto a HiTrap Heparin HP column (Cytiva) equilibrated with a buffer containing 20 mM HEPES pH 7.5, 150 mM NaCl, 0.5 mM TCEP, and 5% glycerol and eluted with a linear gradient of NaCl concentration to 1 M. Protein-containing fractions were loaded onto a HiLoad Superdex 200 16/600 PG column (Cytiva) equilibrated with a buffer containing 20 mM HEPES pH 7.5, 300 mM NaCl, 0.5 mM TCEP, and 5% glycerol. The His6-MBP tag of Ctc1 was only cleaved off for the negative-stain EM analysis of the CST–POT1(ESDL)/TPP1–ssDNA complex (Fig. S2E) but left uncleaved for all other experiments to issues with aggregation and poor vitrification. For tag cleavage, the eluate from the Ni-NTA resin was incubated with rhinovirus 3C protease overnight at 4°C to remove the His6-MBP tag prior to Heparin-affinity and size-exclusion chromatography as described above.
For the POT1/TPP1 proteins, frozen pellets were lysed by cryogenic milling (Retsch) and the cryo-milled powder was resuspended in a buffer containing 50 mM HEPES pH 7.5, 350 mM NaCl, 5 mM β-ME, 5% glycerol, 0.05% (v/v) Tween-20, and 1 mM PMSF, supplemented with cOmplete EDTA-free protease inhibitor cocktail (Roche). The lysate was cleared by centrifugation at 40,000 rcf for 1 h at 4°C. Supernatants were incubated with end-over-end rotation for 1 h at 4°C with Strep-Tactin Superflow high-capacity resin (IBA Lifesciences) equilibrated with gel-filtration buffer (20 mM HEPES pH 7.5, 300 mM NaCl, 0.1 mM TCEP, 0.05% Tween-20, and 5% glycerol) and subsequently washed with 20–50 CV of the same buffer. Bound protein was eluted in the same buffer supplemented with 10 mM d-desthiobiotin, concentrated, and loaded onto a HiLoad Superdex 200 16/600 PG column (Cytiva) equilibrated with the same gel-filtration buffer.
The fusion CST–POT1/TPP1 complex was purified similarly, but after elution from the Strep-Tactin resin, it was concentrated to 500 μL and loaded on an 11-mL linear 10–30% glycerol gradient. Ultracentrifugation was carried out at 41,000 rpm in an SW 41 Ti rotor for 18 h at 4°C. 500 μL fractions were manually collected and protein-containing fractions were identified by SDS-PAGE.
Protein-containing fractions were concentrated, flash frozen in LN2, and stored as aliquots at −80°C. Protein concentrations were measured on a NanoDrop-1000 spectrophotometer.
Fluorescence-detection size-exclusion chromatography analysis
Protein–protein interaction experiments were performed on a Superose 6 Increase gel-filtration column (Cytiva) equilibrated in buffer containing 20 mM HEPES pH 7.5, 150 mM NaCl, and 0.1 mM TCEP. Eluate from the column was passed through an RF-20A fluorescence detector (Shimadzu) operated with excitation and emission wavelengths of 280 nm and 340 nm, respectively. Purified proteins were mixed to a final volume of 120 μL and incubated on ice for at least 30 mutes prior to injection of 100 μL on the column. GFP-tagged POT1/TPP1 complexes were added at the specified concentrations, and [GGTTAG]3 ssDNA (when present) and CST were added at a twofold molar excess. Complex formation was evaluated by comparing the mobility on the gel-filtration column of POT1/TPP1 alone versus when mixed with CST.
For the dephosphorylation experiments, 10 μM POT1/TPP1 was incubated with 10 μM λ protein phosphatase (λPP, expressed and purified in-house from E. coli) and 1 mM MnCl2 in the FSEC gel-filtration buffer. For mock treated samples, the λPP was omitted and replaced with buffer. Samples were incubated for 8–16 h at 4°C. An aliquot from each sample was taken for SDS-PAGE analysis before proteins were diluted for the FSEC experiments described above.
Reconstitution of the ssDNA–CST–POT1/TPP1 complex
Purified His6-MBP-Ctc1/Stn1/Ten1 and His6-POT1(ESDL)/TwinStrep-GFP-TPP1 were mixed with [GGTTAG]3 ssDNA in 150 μL at final concentrations of 6 μM, 6 μM, and 9 μM, respectively, in a buffer containing 20 mM HEPES pH 7.5, 0.5 mM TCEP pH 7.5, and 1% glycerol. Based on the volume of the input proteins (which were purified at 300 mM NaCl), the NaCl was adjusted to a final concentration of 150 mM. The protein components were mixed first and incubated on ice for 1 h prior to addition of the ssDNA. The protein and ssDNA were incubated on ice for 2 h prior to loading on an 11-mL linear 10–30% glycerol gradient. Ultracentrifugation was carried out at 41,000 rpm in an SW 41 Ti rotor for 18 h at 4°C. 500 μL fractions were manually collected and protein- and DNA-containing fractions were identified by SDS-PAGE and native PAGE.
Negative-stain EM sample preparation, data collection, and image processing
Protein samples for negative-stain EM (3.5 μL drops, in a concentration range of 0.01–0.05 mg/mL) were adsorbed to glow-discharged carbon-coated copper grids with a collodion film, washed with three drops of deionized water, and stained with two drops of freshly prepared 0.7% (w/v) uranyl formate. Samples were imaged at room temperature using a Phillips CM10 electron microscope equipped with a tungsten filament and operated at an acceleration voltage of 80 kV. The magnification used for the CST-only samples corresponds to a calibrated pixel size of 3.5 Å. The magnification used for the CST–POT1(ESDL)/TPP1 samples corresponds to a calibrated pixel size of 2.6 Å. Particles were auto-picked using the Swarm picker (CST-only) or Gauss picker (CST–POT1(ESDL)/TPP1) in EMAN262. Particle extraction and 2D classification were performed in RELION-3.163.
Cryo-EM sample preparation and data collection
Apo CST–POT1(ESDL)/TPP1 complex was frozen at a concentration of 0.075 mg/mL, corresponding to 0.2 μM of GFP-tagged protein, in buffer containing 20 mM HEPES pH 7.5, 150 mM NaCl, and 0.1 mM TCEP. The CST–POT1(ESDL)/TPP1–ssDNA complex was prone to aggregation following vitrification at the concentration used for the apo complex and faced challenges with inconsistent ice thickness across the grid. Thus, the CST–POT1(ESDL)/TPP1–ssDNA complex was diluted to 0.02 mg/mL, or 0.05 μM of GFP-tagged protein. The addition of 0.75 mM fluorinated Fos-choline-8 (Anatrace) resulted in even ice thickness.
3.5 μL of the samples were applied to Quantifoil R1.2/1.3 mesh Cu400 holey carbon grids covered with a graphene oxide support layer (EMS) that were glow-discharged for 5 s at 40 mA in an EMS100X glow discharge unit (EMS) and then blotted for 0.5–1 s (Blot Force −2; Wait Time 20 s; Drain Time 0 s) and plunge frozen in liquid ethane using a Vitrobot Mark IV (Thermo Fisher Scientific) operated at 4°C and 100% humidity. Cryo-EM imaging was performed in the Evelyn Gruss-Lipper Cryo-EM Resource Center at The Rockefeller University using SerialEM64. Data collection parameters are summarized in Table S1.
For both complexes, data were collected on a 300-kV Titan Krios electron microscope equipped with a Cs corrector at a nominal magnification of ×105,000, corresponding to a calibrated pixel size of 0.86 Å (image dimensions of 5,760×4,092 px) on the specimen level. Images were collected using a slit width of 20 eV on the GIF Quantum energy filter (Gatan) and a defocus range from −1 to −2.5 μm with a K3 direct electron detector (Gatan) in super-resolution counting mode. Three movie stacks were recorded per hole in a 3×3 matrix of holes acquired using beam-image shift with a maximum shift of 3.5 μm. Exposures of 1.2 s were dose-fractionated into 40 frames (30 ms per frame) with an exposure rate of 31 electrons/pixel/s (approximately 1.25 electrons per Å2 per frame), resulting in a total electron exposure of 50.3 electrons per Å2.
Cryo-EM data processing
For all datasets, movie stacks were motion-corrected with the RELION-3.163 implementation of MotionCor265, and motion-corrected micrographs were manually inspected and curated (graphene oxide coverage of grids was inconsistent) prior to CTF parameter estimation with CTFFIND-466 implemented in RELION-3.1.
To generate templates and an initial model for the apo complex, particles were automatically picked without a template using Gautomatch (v0.53) and extracted (352 px box, binned 4-fold to 88 px) in RELION-3.1. After 2D classification and cleanup to remove contaminating ice particles, the cleaned particle stack was imported into cryoSPARC (v4.1.1) and 2D classification was performed. The best 2D classes (containing ~150,000 particles) were used to generate an initial model that was used as a 3D reference. The motion-corrected micrographs were then imported into cryoSPARC67, and CTF parameter estimation was performed with CTFFIND-466 implemented in cryoSPARC. Particles were automatically picked using the Template Picker with the initial model as a 3D reference. Particles were extracted, Fourier-cropped from a 352 px box to a 256 px box to speed up image processing and processed using the described image-processing pipeline depicted in Fig. S3. Three rounds of heterogeneous refinement were performed with one good reference and three noise “decoy” references to filter out junk particles. After the number of particles in the junk classes dropped to 15% or below, the good particle stack was further classified using heterogeneous refinement into four classes using a single reference. The 372,930 particles in the best class were re-extracted at full-size (352 px box). There was additional heterogeneity, particularly in the region of Stn1C and POT1OB-1/2, so further unsupervised heterogeneous refinement was used to identify particles with fully intact Stn1C and POT1OB-1/2. The final stack of 132,356 particles was refined and sharpened to a nominal resolution of 3.6 Å using the Non-uniform Refinement job in cryoSPARC and the local resolution of the map was estimated in cryoSPARC. The reported 3.9-Å resolution was calculated with Phenix (phenix.validation_cryoem) using the gold-standard Fourier shell correlation (FSC) between half-maps (Fig. S3I). FSC plots and sphericity values58 of the reconstruction were calculated using the 3D-FSC server (https://3dfsc.salk.edu/).
For the ssDNA-bound complex, the motion-corrected micrographs were imported into cryoSPARC, and CTF parameter estimation was performed with the Patch CTF Estimation job. The apo complex map was used as a 3D template for particle picking with the Template Picker job. Particles were extracted, Fourier-cropped from a 352 px box to a 128 px box to speed up image processing and processed using the described image-processing pipeline depicted in Fig. S4. The 2D-class averages of the extracted particles and negative-stain EM analysis of the complex (Fig. 4) revealed heterogeneity. This heterogeneity could be divided into three categories of particles: fully intact CST–POT1/TPP1 complexes that resembled the apo complex and contained bound Stn1C and POT1OB-1/2, half complexes in which Stn1C and POT1OB-1/2 were disengaged but the POT1 C-terminus was bound, and particles containing CST without POT1/TPP1. Thus, three 3D references were generated from the apo complex by segmenting the map in UCSF Chimera68. The references are shown in Fig. S4. Two noise decoy references were used in addition to the three real references to perform iterative heterogeneous refinement until the noise classes contained fewer than 10% of the particles. An additional heterogeneous refinement with the three different references was performed prior to re-extracting the particles with Stn1C and POT1OB-1/2 bound at full size. There was still some heterogeneity in that region, so the map was segmented to generate a mask around the Ctc1 C-terminus, Stn1, Ten1, and POT1OB-1/2. This mask was used to perform focused 3D classification into four classes without alignment in cryoSPARC. Two classes contained density for both Stn1C and POT1OB-1/2 and were pooled into a final stack of 76,359 particles. The other two classes were missing density for one of the components. The final particle stack was refined and sharpened to a nominal resolution of 4.0 Å using the Non-uniform Refinement job in cryoSPARC and the local resolution of the map was estimated in cryoSPARC. The reported 4.3-Å resolution was calculated with Phenix (phenix.validation_cryoem) using the gold-standard FSC between half-maps (Fig. S4I). FSC plots and sphericity values58 of the reconstruction were calculated using the 3D-FSC server (https://3dfsc.salk.edu/).
We also attempted to continue processing the complex in which ssDNA was presumably bound to CST where only the C-terminus of POT1 is bound to CST (Fig. S4D). Focused 3D classification revealed that there were few particles truly missing the Stn1C and POT1OB-1/2 as cartooned in Fig. 5B, but the association was weak. Neither heterogeneous refinement nor focused 3D classification without alignment were able to further classify the particles or align to higher-resolution features. This is likely due to the intrinsic flexibility of CST, for which a high-resolution structure of a monomeric complex without other interacting factors could not be determined13. It appears that when bound, the Stn1C and POT1OB-1/2 interactions stabilized CST to allow for high-resolution structure determination.
Model building and refinement
The map of the apo complex was used for atomic model building, as it had a higher overall resolution and better quality. The AlphaFold236, 37 database model of Ctc1 (AF-Q2NJK3) was merged with the cryo-EM structure of CST (PDB 6W6W13) as the starting model for CST. The X-ray crystal structures of POT1OB-3/HJRL/TPP1RD (PDB 5H6538) and the POT1 OB-folds bound to ssDNA (PDB 1XJV40) were used as starting models for the POT1/TPP1 components. CST, POT1OB-2, and POT1OB-3/HJRL/TPP1RD were first rigid-body docked into the map using UCSF Chimera. The model was then manually inspected in Coot69, where additional density of the hinge region of POT1 was observed. There was clear density for aa 322–347 that allowed unambiguous manual modeling of the hinge self-interactions with POT1OB-3. The N-terminal region of the hinge (aa 299–321 and the ESDL insertion) showed connected Cα backbone density, but many side-chain placements were ambiguous. Because this region of the hinge makes a weak interaction that appears to be mediated primarily by charge complementarity rather than specific interactions, the exact side-chain positions were not critical to the analysis. Thus, we first modeled the hinge Cα backbone into the density and observed that it spanned the distance between the preceding and following regions of the chain. Then, we assigned the sequence in register with the rest of the protein. The key reason for modeling the side-chains for residues 308–321 was to show the abundance of serine and aspartate residues in this region. Residues 301–307 were modeled as poly-alanine stubs, as this part of the chain runs in a groove of Ctc1 and we did not want to over- or mis-interpret specific interactions. The resolution of POT1OB-2 was not high, so POT1OB-1 was only flexibly fitted into the density using ISOLDE70 after rigid-body docking. For Ctc1, the AlphaFold2 model fit well overall into the density and required only few manual adjustments. One region of Ctc1 (aa 909–927) appeared to contact POT1 weakly. This region of Ctc1 has not been resolved in previous experimental structures and AlphaFold2 predicts this loop with low confidence. There was clear density near the predicted position of the loop, so it was approximated with poly-alanine stubs. The map regions containing Stn1 and Ten1 were also at lower resolution, so they were docked in and fitted similarly to POT1OB-2. Iterative real-space refinement in Phenix (phenix.real_space_refine) and manual model adjustment in ISOLDE was used to correct the most egregious geometry errors, though some errors could not be fixed in a way supported by the experimental map. The model and maps were validated in Phenix (phenix.validation_cryoem)71, 72.
For the ssDNA-bound complex, the apo complex model was used as the starting model. POT1OB-2 from the apo complex was replaced with a rigid-body docking of POT1OB-1/2 and a 5’-TTAGGGTTAG-3’ DNA from the X-ray crystal structure (PDB 1XJV40). Because the overall resolution of the map of the ssDNA-bound complex was lower, but the overall interaction and map density was similar between the apo and ssDNA-bound complex, we treated the apo model as relatively correct. The apo model was fitted to the ssDNA-bound complex map using iterative real-space refinement in Phenix (phenix.real_space_refine) and some manual model adjustment in ISOLDE. The resolution of POT1OB-1/2 was low, so it was modeled by flexible fitting of the crystal structure and some attempts were made to correct its geometry in ISOLDE. ssDNA was clearly present, but the resolution was too low to assign the register or see if more than 10 nt were bound. Thus, only the 5’-TTAGGGTTAG-3’ from the crystal structure was retained in the final model. The model and maps were validated in Phenix (phenix.validation_cryoem).
For PDB deposition, the POT1 ESDL insertion was notated using insertion code as an insertion after residue 320 (E – 320A; S – 320B; D – 320C; L – 320D) to maintain the canonical numbering of the human POT1 sequence. The TPP1–Stn1 fusion in the apo complex required deposition as a single chain, so the numbering corresponds to the fusion. However, individual mappings to TPP1 and Stn1 within the chain are annotated in the deposition.
Multi-body refinement analysis of CST–Polα/primase RC flexibility
The dataset used for RC analysis was the same as previously published15 (EMPIAR-11131) and reprocessed in RELION-3.1 with the same particle coordinates and particles extracted. The final stack used for multi-body refinement contained 203,045 particles (300 px box; 1.08 Å/px). Masks for Polα/primase (body 1) and CST (body 2) were generated by segmentation of the RC map in UCSF Chimera and mask creation in RELION-3.1. Multi-body refinement was performed with default parameters45. Angular/translational priors of 10°/2px and 20°/5px were used for Polα/primase and CST, respectively. Volume series movies were generated in UCSF ChimeraX73.
AlphaFold-Multimer complex structure prediction
AlphaFold-Multimer (v1.0)32 was used on the COSMIC2 server74 with default settings (full_dbs). For the Ctc1OB-A/B/C–TPP1TID complex, the input sequence file contained Ctc1 aa 1–410 and TPP1 aa 361–458. For the Ctc1–POLA2 complex, the input sequence file contained Ctc1 aa 1–1217 and POLA2 aa 1–158. Flexible regions with low predicted confidence were not shown in the figures.
Mass photometry
Purified POT1/TPP1 (5 μM) and CST (7.5 μM) proteins were mixed with 7.5 μM [GGTTAG]3 in MP buffer (20 mM HEPES pH 7.5, 150 mM NaCl, and 0.1 mM TCEP pH 7.5, filtered 3x through 0.22 μm syringe filter) and incubated for 1 h at 4°C. Where indicated, λPP treatment of POT1/TPP1 proteins was performed as described above prior to incubation with CST. Immediately prior to measurement, samples were diluted to 100 nM POT1/TPP1 in MP buffer, then 2 μL was added to 8 μL MP buffer in a sealed well on a glass slide for a final concentration of 20 nM POT1/TPP1 and 1.5-fold excess CST and ssDNA. Data were collected using a OneMP mass photometer (Refeyn) calibrated with bovine serum albumin (66 kDa), beta amylase (224 kDa), and thyroglobulin (660 kDa). Movies were acquired for 6,000 frames (60 s) using AcquireMP software (v2022R1) and default settings. Final protein concentrations were empirically determined to achieve ~50 binding events per second. Raw event measurements were converted to frequency distributions using DiscoverMP software (v2022R1) or Prism 9 (GraphPad) and a bin size of 10 kDa. Gaussian curves were autofitted with DiscoverMP to provide total peak counts for quantification.
Double-filter DNA-binding assays
A modified double-filter DNA-binding assay was used to calculate apparent binding constants (KD,app)44. Binding buffer contained 50 mM HEPES-KOH pH 7.5, 150 mM KCl, 0.5 mM TCEP, 0.1 mg/mL bovine serum albumin, and 15% glycerol. The [GGTTAG]3 ssDNA oligonucleotide (Thermo Fisher) was 5’-end-labeled using [γ-32P]ATP with T4 polynucleotide kinase (NEB). Free nucleotides were removed using Micro Bio-Spin 6 columns (BioRad) and the labeled oligonucleotide was diluted in binding buffer to achieve a final concentration of <10 pM in the experiment. The maximum concentration of 10 pM was calculated with the assumption of no product loss on the spin column. Experiments were performed in a 96-well vacuum manifold dot blot apparatus (GE Healthcare) with nitrocellulose, Hybond-XL filters (charged nylon; GE Healthcare), and 3MM chromatography filter paper (GE Healthcare). All filters were presoaked in 50 mM HEPES pH 7.5 prior to assembly in the apparatus. Proteins were serially diluted to reach the concentrations indicated, mixed with the final <10 pM labeled oligonucleotide in a final volume of 220 μL, and incubated in a 96-well plate on a cold block in ice for 1 h. 100 μL binding buffer was applied to the apparatus to wash the membranes and briefly vacuumed without allowing the membrane to completely dry. The vacuum line was removed prior to loading 100 μL protein-ssDNA mix per well with a multi-channel pipette. Two technical replicates to control for pipetting error were performed per experiment by loading two wells per condition from the same protein-ssDNA mix. After incubation for 1 m, vacuum was applied and the wells were washed three times with 100 μL binding buffer. Filters were air dried and wrapped in plastic wrap before exposing to phosphor imaging screens for 10 days (GE Healthcare). Screens were scanned on an Amersham Typhoon imager and data were quantified using ImageJ75 and GraphPad Prism 9.
C-Strand synthesis assay
CST–Polα/primase was expressed in HEK293T/17 cells (ATCC, CRL-11268, LOT: 63696280) transfected with plasmids encoding the three subunits of CST and purified by sequential pull-down with antibodies to the FLAG tag on Ctc1 and the HA tag on Ten1, as described47. The reported protein concentrations refer to CST; the endogenous Polα/primase, which co-immunopurifies with the CST, is present at about 20% stoichiometry. DNA templates and chimeric RNA-DNA primers were purchased from IDT. C-strand synthesis reactions proceeded for 60 m at 30°C in 50 mM Tris-HCl pH 8.0, 1 mM MgCl2, 1 mM spermidine, 5 mM β-ME, 50 mM KCl, 50 mM NaCl, 0.2 mM each ATP, CTP and UTP, 0.5 mM each dCTP and TTP, 0.01 mM dATP, and 0.17 mM 32P-dATP (3,000 Ci/mmole, Revvity, formerly Perkin Elmer). Products were separated by electrophoresis on a 10% acrylamide/7 M urea/1x TBE gel. 32P-labeled loading controls L1 (18 nt), L2 (16 nt) and L3 (15 nt) were added with the stop buffer (3.6 M NH4Ac, 0.2 mg/mL glycogen). Gels were dried and scanned with a phosphorimager (Cytiva), signal intensities were determined (ImageQuant), normalized to the signal intensity in the absence of added POT1/TPP1, and data fit to a sigmoidal binding curve (KaleidaGraph). Reactions with pre-primed templates were performed similarly, with equal concentrations of the template and primer first pre-incubated for 10 m at room temperature and then stored at −20°C. Prior to the reaction the pre-primed templates were thawed at room temperature and added to the same reaction mixture as above except with no ribonucleotides. Products from the pre-primed substrates were analyzed as described above for C-strand synthesis.
Structure analysis and visualization
The apo CST–POT1(ESDL)/TPP1 complex was used for all interaction analysis, as the model and map quality was superior to that of the ssDNA-bound complex. The additional POT1OB-1 and telomeric DNA in the ssDNA-bound complex did not interact with CST. UCSF Chimera, UCSF ChimeraX, and PyMOL (Schrödinger) were used to analyze the maps and models. Figures were prepared using UCSF ChimeraX and Adobe Illustrator. Maps in ChimeraX were prepared with the “hide dust” limit set to 10 and the contour thresholds as specified in the figure legends.
QUANTIFICATION AND STATISTICAL ANALYSIS
Sample sizes for the cryo-EM datasets were determined by the need to obtain meaningful structures and the availability of cryo-EM time. For the apo CST-POT1/TPP1 dataset, 15,201 movie stacks were collected, and 132,356 particles were used for the final reconstruction, which was sufficient to yield a 3.9-Å resolution structure. For the ssDNA-CST-POT1/TPP1 dataset, 31,914 movie stacks were collected, and 76,359 particles were used for the final reconstruction, which was sufficient to yield a 4.3-Å resolution structure. Micrographs clearly suffering from astigmatism, image drift, poor graphene oxide coverage, ice contamination, and/or cubic ice formation were excluded during the micrograph curation step in all cryo-EM datasets analyzed. Particles in 2D classes showing no secondary structural features and in 3D classes showing unsatisfactory structural features were excluded from the final reconstructions in all datasets analyzed.
For the high-resolution structures obtained in this study, small negative-stain EM datasets were first collected on an 80-kV Phillips CM10 electron microscope. Cryo-EM data collection on a 300-kV Titan Krios electron microscope resulted in structures that matched the negative-stain EM 2D averages obtained. Each biochemical reconstitution was performed at least three times and was reproducible.
All other statistical analyses and technical replicates are described in the methods and figure legends for the corresponding experiments.
Supplementary Material
(A) Co-IPs of Myc-tagged TPP1 proteins with the indicated C-terminal truncations with FLAG-tagged Ctc1, Stn1, and Ten1 from co-transfected 293T cells showing that the C-terminus of TPP1 is required for the CST–TPP1 interaction. Western blots of Myc IPs were probed with anti-Myc and anti-FLAG antibodies.
(B) Co-IPs of Myc-tagged TPP1 with HA-tagged TIN2 and FLAG-tagged Ctc1, Stn1, and Ten1 from co-transfected 293T cells showing that TIN2 competes with CST for TPP1 interaction. Western blots of Myc IPs were probed with anti-Myc, anti-FLAG, and anti-HA antibodies.
(C) Predicted aligned error (PAE) plots of the top four (of 5) ranked AlphaFold-Multimer32 models for Ctc1–TPP1. Green arrowheads indicate high confidence in the position prediction of TPP1 relative to Ctc1. All 5 top-ranked models predicted the interaction with high confidence.
(D) (Left) AlphaFold-Multimer model of Ctc1 OB-folds A, B, and C bound to TPP1TID compared to the crystal structure of TRF2/TIN2/TPP1TID (PDB 5XYF31). TPP1TID is predicted to bind to Ctc1 and TIN2 using the same peptide, providing a structural basis for their competition. (Right) Domain schematics of Ctc1, TPP1, and TIN2 with gray shadows showing the overlapping interaction site on TPP1TID. OB: oligosaccharide/oligonucleotide-binding domain; ARODL: Acidic Rpa1 OB-binding Domain-Like 3-helix bundle; RD: recruitment domain, TID: TIN2-interacting domain; TRFH: telomere repeat factor homology domain.
(A) FSEC analysis of the binding between increasing concentrations of CST and POT1(ESDL)/TPP1 in the absence of telomeric ssDNA showing the concentration dependence of the interaction. RFU: relative fluorescence units.
(B) Native glycerol gradient analysis of ssDNA-bound CST–POT1(ESDL)/TPP1. Pooled fractions are indicated with an asterisk. (Top) Coomassie-stained SDS-PAGE gel (4–12% Bis-Tris gel run in MOPS-SDS buffer, Invitrogen) of glycerol gradient fractions. (Bottom) SYBR Gold-stained native PAGE (4–20% TBE gel run in 0.5x TB buffer, Invitrogen).
(C) Domain architecture of components used to reconstitute an apo CST–POT1(ESDL)/TPP1 complex. TPP1 is fused to the N-terminus of Stn1 and retains part of the TPP1 serine-rich linker.
(D) Coomassie-stained SDS-PAGE gel (4–12% Bis-Tris gel run in MOPS-SDS buffer, Invitrogen) showing the purified CST–POT1(ESDL)/TPP1 fusion complex used for structural analysis.
(E) Negative-stain EM analysis of CST and CST–POT1(ESDL)/TPP1 complexes. (Left) Representative negative-stain EM micrographs. (Middle) Top 25 reference-free 2D-class averages (sorted by number of particles per class from most populous class at top left to least populous class at bottom right) of each complex. (Right) Enlarged views of selected 2D averages to show CST features. Additional density attributable to the addition of POT1(ESDL)/TPP1 is indicated with red arrowheads.
(A) Representative motion-corrected micrograph.
(B) Enlarged view of the area marked in (A) with selected particles circled.
(C) Representative 2D-class averages show high-resolution features and different orientations.
(D) Cryo-EM image-processing pipeline used for the apo CST–POT1(ESDL)/TPP1 complex, including supervised 3D classification with noise decoy classes. Several rounds of heterogeneous refinement were used to select for particles with well-resolved Stn1C and POT1OB1/2 (Table S1).
(E) Final map (contour threshold 0.2) of the apo CST–POT1(ESDL)/TPP1 complex with POT1(ESDL)/TPP1 colored as in Fig. 2A for reference.
(F) Directional FSC plots and sphericity values58 of the reconstruction calculated using the 3D-FSC server (https://3dfsc.salk.edu/).
(G) Plot of the angular distribution of particles in the final reconstruction.
(H) Local resolution estimates for the map (contour threshold 0.2) of apo CST–POT1(ESDL)/TPP1.
(I) Gold-standard (blue) and model-vs-map (red) FSC curves for the reconstruction of apo CST–POT1(ESDL)/TPP1. The map resolution was estimated using the gold-standard FSC and a cut-off criterion of 0.143. The similarity between the resolution based on the gold-standard FSC (0.143 cut-off) and the resolution estimate based on the model-vs-map FSC (0.5 cut-off) suggests no substantial over-fitting.
(J) Cryo-EM map densities for each subunit indicating quality of fit for the apo CST–POT1(ESDL)/TPP1 model. Contour thresholds for Ctc1, Stn1, Ten1, POT1, and TPP1 are 0.1, 0.065, 0.15, 0.1, and 0.05, respectively.
(A) Representative motion-corrected micrograph.
(B) Enlarged view of the area marked in (A) with selected particles circled.
(C) Representative 2D-class averages show high-resolution features and different orientations.
(D) Cryo-EM image-processing pipeline used for the ssDNA-bound CST–POT1(ESDL)/TPP1 complex, including supervised 3D classification with noise decoy classes. Focused 3D classification with a mask was used to select for particles with well-resolved Stn1C and POT1OB1/2 (Table S1).
(E) Final map (contour threshold 0.15) of the ssDNA-bound CST–POT1(ESDL)/TPP1 complex with ssDNA–POT1(ESDL)/TPP1 colored as in Fig. 2A for reference.
(F) Directional FSC plots and sphericity values58 of the reconstruction calculated using the 3D-FSC server (https://3dfsc.salk.edu/).
(G) Plot of the angular distribution of particles in the final reconstruction.
(H) Local resolution estimates for the map (contour threshold 0.15) of the ssDNA-bound CST–POT1(ESDL)/TPP1.
(I) Gold-standard (blue) and model-vs-map (red) FSC curves for the reconstruction of CST–POT1(ESDL)/TPP1. The map resolution was estimated using the gold-standard FSC and a cut-off criterion of 0.143. The similarity between the resolution estimate based on the gold-standard FSC (0.143 cut-off) and the resolution estimate based on the model-vs-map FSC (0.5 cut-off) suggests no substantial over-fitting.
(J) Cryo-EM map densities for each subunit indicating quality of fit for the model of ssDNA-bound CST–POT1(ESDL)/TPP1. Contour thresholds for Ctc1, Stn1, Ten1, POT1, and TPP1 are 0.075, 0.14, 0.1, 0.1, and 0.1, respectively.
(A) Deletion of the TPP1 OB-fold does not affect the CST–POT1(ESDL)/TPP1 interaction. Protein used and FSEC analysis of CST–POT1(ESDL)/TPP1(ΔOB) interaction in the absence (top) and presence (bottom) of telomeric ssDNA. Traces without CST are shown as dashed lines. RFU: relative fluorescence units.
(B) Electrostatic surface from Fig. 3A with map density (contour threshold 0.1) of hinge shown as mesh.
(C) Coomassie-stained SDS-PAGE gel (4–12% Bis-Tris gel run in MOPS-SDS buffer, Invitrogen) showing the purified POT1/TPP1 proteins (untreated or λ-treated) used for FSEC and MP analysis. Descriptions of the POT1 variant sequences are given in Fig. 3E.
(D) FSEC analysis of the phosphorylation-dependent interaction between CST and POT1(WT)/GFP-TPP1 (top, black) and POT1(ESDL)/GFP-TPP1 (bottom, blue) in the absence of telomeric ssDNA (see also Fig. 3B). Traces without CST are shown as dashed lines, and traces containing POT1/TPP1 that have been treated with λ protein phosphatase (λPP) are shown in red. RFU: relative fluorescence units. The chromatograms for the WT and ESDL mock samples are shown in Fig. 1E and are reproduced here as the controls for the phosphatase experiment, which was performed simultaneously.
(E) MP histograms used for the quantification shown in Fig. 3F. Each panel shows histograms from three technical replicates with Gaussian curves auto-fitted by DiscoverMP software. Inset tables show σ (standard deviation of values within peak) and and total counts per peak for each replicate. Total peak counts were used to quantify the proportion of bound POT1/TPP1, the ratio of total counts in peak 3 (CST–POT1/TPP1 complex) divided by total counts of POT1/TPP1 (peak 1 and peak 3).
(A) As in Fig. 4B but individual panels include the cryo-EM map density for the apo CST–POT1(ESDL)/TPP1 complex shown as mesh (contour thresholds 0.3, 0.3, 0.7, 0.5, and 0.5, respectively).
(B) Co-IPs of Strep-tagged POT1 (or mPOT1b) protein variants and FLAG-tagged Ctc1, Stn1, and Ten1 from co-transfected 293T cells showing that the corresponding S328L CP mutation in mPOT1b disrupts the CST interaction. Western blots of Strep IPs were probed with anti-Strep and anti-FLAG antibodies.
(C) Double-filter DNA-binding assays showing high affinity [GGTTAG]3 binding by POT1/TPP1 alone and in complex with CST. Left: Representative membrane scans. The nitrocellulose membrane is labeled as “Bound” and the Hybond-XL membrane is labeled as “Free.” Right: Quantitation of binding data. Data were fitted using the “one site – specific binding” model on GraphPad Prism 9. Error bars represent SEM from two technical replicate experiments.
(A) Multiple views of CST–Polα/primase in RC (left) and PIC (right) conformations. CST–Polα/primase structures are shown in cartoon representation with a transparent surface.
(B) Superposition of the ssDNA-bound CST–POT1(ESDL)/TPP1 structure with the structures of the RC (left, see also Video S3) and PIC (right, see also Video S2) complexes showing additional views compared to Fig. 5C. The clash between the POT1HJRL and PRIM2 is indicated.
(C) Multi-body analysis of CST–Polα/primase in the RC conformation. Polα/primase was designated body 1 and CST was designated body 2 with the corresponding masks shown. Histograms of the projections of the relative orientations onto the corresponding components show a unimodal distribution, consistent with continuous flexibility rather than discrete states. The first three principal components account for 61% of the variance in the data. Reconstructed maps from the extreme ends are shown in red and blue for each of the first three principal components with arrows indicating the direction of motion (see also Video S4).
(D) PAE plots of the top two (of 5) ranked AlphaFold-Multimer32 models for Ctc1/POLA2. Green arrowheads indicate high confidence in the position prediction of POLA2 relative to Ctc1. All 5 top-ranked models predicted the interaction with high confidence.
(E) AlphaFold-Multimer model of Ctc1 bound to POLA2NTD. Residues with known CP mutations are shown as spheres. Red spheres indicate mutations previously shown to disrupt Polα/primase association41; black spheres indicate mutations with unknown mechanism.
(A) Inhibition of C-strand synthesis by POT1(ESDL)/TPP1 is enhanced by its DNA-binding domain. Reactions of 50 nM CST–Polα/primase with 50 nM single-stranded 9xTEL DNA template. LC1, LC2, LC3: oligonucleotide loading controls.
(B) Quantification of the initial extension products, bands 21–23.
(C) Quantification of all products, bands 21–40, gave similar IC50 values as in (B).
(D) Design of pre-primed templates to monitor primer extension by Polα. Red sequence is RNA.
(E) Validation of the pre-primed template reactions by performing reactions in the absence of ribonucleotides. Most reaction products were dependent on both template and primer. The 9xTEL template products (which were formed in the absence of the primer strand) were longer than the template, indicating that they resulted from self-priming of the DNA template.
(F) CST–Polα extension of pre-primed templates. M: telomerase reaction products as size markers. Stronger inhibition of C-strand synthesis by POT1(ESDL)/TPP1 persists under conditions that limit re-priming (50 nM template and 5 nM CST–Polα/primase, a ten-fold higher template/enzyme ratio than used in Fig. 6E).
(G) Quantification of the data in (F) (left), normalized to reactions without added POT1/TPP1. IC50 values with error bars indicating uncertainty of the fit to the data points.
(H) Quantification of the data in (F) (right), normalized to reactions without added POT1/TPP1. IC50 values were not calculated because inhibition was incomplete at highest POT1/TPP1 concentrations.
360° rotation of the cryo-EM map and model from apo and ssDNA-bound CST–POT1(ESDL)/TPP1. Colors are the same as in Fig. 2.
Rotation movie showing superposition of CST–POT1(ESDL)/TPP1–ssDNA complex (surface representation; solid colors as in Fig. 2) with CST–Polα/primase PIC (surface representation at 50% opacity; cartoon model shown underneath). Colors of Polα/primase are the same as in15. PRIM1-salmon; PRIM2-light orange; POLΑ1-light green (Ctc1-recognition loop-lime green); POLA2-yellow.
Rotation movie showing superposition of CST–POT1(ESDL)/TPP1–ssDNA complex (surface representation; solid colors as in Fig. 2) with CST–Polα/primase RC (surface representation at 50% opacity; cartoon model shown underneath). Colors of Polα/primase are the same as in15. PRIM1-salmon; PRIM2-light orange; POLΑ1-light green (Ctc1-recognition loop-lime green); POLA2-yellow.
Principal components 1, 2, and 3, respectively.
Highlights.
Cryo-EM structures reveal how the shelterin subunit POT1 binds to Ctc1 in CST
The structures reveal the mechanism of Coats plus syndrome mutations in Ctc1 and POT1
CST recruitment to shelterin requires phosphorylation of POT1
Activation of CST–Polα/primase for C-strand synthesis involves its release from POT1
Acknowledgements
We thank J. Zinder for generating and optimizing the expression strategy of the wild-type POT1/TPP1 expression construct and for providing the λ protein phosphatase. We thank T. Bakker for designing the TPP1-Stn1 fusion gene under the supervision of J. Zinder, L. Vostal and T. Kapoor for training on and use of the mass photometry instrument, H. Molina at the Proteomics Resource Center of Rockefeller University for assistance with mass spectrometry, and D. Winkler at Kinexus Bioinformatics for advice on filter-bound peptide assay. We thank M. Ebrahim, J. Sotiris, and H. Ng at the Evelyn Gruss Lipper Cryo-EM Resource Center of The Rockefeller University for assistance with cryo-EM data collection. Some grids were also frozen at the Electron Microscopy Resource Center of The Rockefeller University. SWC is supported by the David Rockefeller Graduate Program and an NSF Graduate Research Fellowship (1946429). HT is supported by an R50 grant from the NIH (5 R50 CA243771). This work is supported by grants to TdL from the NIH (5 R35 CA210036) and the Breast Cancer Research Foundation (BCRF-19–036). TRC is an investigator of the Howard Hughes Medical Institute.
Footnotes
Competing interests
TRC is a member of the SAB of Storm Therapeutics, Eikon Therapeutics, and SomaLogic, Inc. The other authors have no conflicts to declare.
Additional Information
Supplementary Information is available for this paper. Correspondence and requests for materials should be addressed to Titia de Lange (delange@rockefeller.edu).
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.
References
- 1.de Lange T (2018). Shelterin-Mediated Telomere Protection. Annu Rev Genet 52, 223–247. [DOI] [PubMed] [Google Scholar]
- 2.Cai SW, and de Lange T (2023). CST-Polα/Primase: the second telomere maintenance machine. Genes Dev 37, 555–569. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Lim CJ, and Cech TR (2021). Shaping human telomeres: from shelterin and CST complexes to telomeric chromatin organization. Nat Rev Mol Cell Biol 22, 283–298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Takai H, Aria V, Borges P, Yeeles JTP, and de Lange T (2024). CST-polymerase α-primase solves a second telomere end-replication problem. Nature 627, 664–670. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Wu P, Takai H, and de Lange T (2012). Telomeric 3’ overhangs derive from resection by Exo1 and Apollo and fill-in by POT1b-associated CST. Cell 150, 39–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Wang F, Stewart JA, Kasbek C, Zhao Y, Wright WE, and Price CM (2012). Human CST has independent functions during telomere duplex replication and C-strand fill-in. Cell Rep 2, 1096–1103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Stewart JA, Wang F, Chaiken MF, Kasbek C, Chastain PD, Wright WE, and Price CM (2012). Human CST promotes telomere duplex replication and general replication restart after fork stalling. EMBO J 31, 3537–3549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Feng X, Hsu SJ, Kasbek C, Chaiken M, and Price CM (2017). CTC1-mediated C-strand fill-in is an essential step in telomere length maintenance. Nucleic Acids Res 45, 4281–4293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Hockemeyer D, and Collins K (2015). Control of telomerase action at human telomeres. Nat Struct Mol Biol 22, 848–852. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.He Y, and Feigon J (2022). Telomerase structural biology comes of age. Curr Opin Struct Biol 76, 102446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Casteel DE, Zhuang S, Zeng Y, Perrino FW, Boss GR, Goulian M, and Pilz RB (2009). A DNA polymerase-{alpha}{middle dot}primase cofactor with homology to replication protein A-32 regulates DNA replication in mammalian cells. J Biol Chem 284, 5807–5818. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Miyake Y, Nakamura M, Nabetani A, Shimamura S, Tamura M, Yonehara S, Saito M, and Ishikawa F (2009). RPA-like mammalian Ctc1-Stn1-Ten1 complex binds to single-stranded DNA and protects telomeres independently of the Pot1 pathway. Mol Cell 36, 193–206. [DOI] [PubMed] [Google Scholar]
- 13.Lim CJ, Barbour AT, Zaug AJ, Goodrich KJ, McKay AE, Wuttke DS, and Cech TR (2020). The structure of human CST reveals a decameric assembly bound to telomeric DNA. Science 368, 1081–1085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Madru C, Martínez-Carranza M, Laurent S, Alberti AC, Chevreuil M, Raynal B, Haouz A, Le Meur RA, Delarue M, Henneke G, et al. (2023). DNA-binding mechanism and evolution of replication protein A. Nat Commun 14, 2326. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Cai SW, Zinder JC, Svetlov V, Bush MW, Nudler E, Walz T, and de Lange T (2022). Cryo-EM structure of the human CST-Polα/primase complex in a recruitment state. Nat Struct Mol Biol 29, 813–819. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.He Q, Lin X, Chavez BL, Agrawal S, Lusk BL, and Lim CJ (2022). Structures of the human CST-Polα-primase complex bound to telomere templates. Nature 608, 826–832. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Anderson BH, Kasher PR, Mayer J, Szynkiewicz M, Jenkinson EM, Bhaskar SS, Urquhart JE, Daly SB, Dickerson JE, O’Sullivan J, et al. (2012). Mutations in CTC1, encoding conserved telomere maintenance component 1, cause Coats plus. Nat Genet 44, 338–342. [DOI] [PubMed] [Google Scholar]
- 18.Keller RB, Gagne KE, Usmani GN, Asdourian GK, Williams DA, Hofmann I, and Agarwal S (2012). CTC1 Mutations in a patient with dyskeratosis congenita. Pediatr Blood Cancer 59, 311–314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Polvi A, Linnankivi T, Kivelä T, Herva R, Keating JP, Mäkitie O, Pareyson D, Vainionpää L, Lahtinen J, Hovatta I, et al. (2012). Mutations in CTC1, encoding the CTS telomere maintenance complex component 1, cause cerebroretinal microangiopathy with calcifications and cysts. Am J Hum Genet 90, 540–549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Walne AJ, Bhagat T, Kirwan M, Gitiaux C, Desguerre I, Leonard N, Nogales E, Vulliamy T, and Dokal IS (2013). Mutations in the telomere capping complex in bone marrow failure and related syndromes. Haematologica 98, 334–338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Hidalgo-Sanz J, Perez-Delgado R, Garcia-Jimenez I, Lopez-Pison J, Castillo-Castejon O, Poch ML, and Izquierdo-Alvarez S (2019). [Infant with intracranial calcifications and retinopathy]. Rev Neurol 69, 289–292. [DOI] [PubMed] [Google Scholar]
- 22.Passi GR, Shamim U, Rathore S, Joshi A, Mathur A, Parveen S, Sharma P, Crow YJ, and Faruq M (2020). An Indian child with Coats plus syndrome due to mutations in STN1. Am J Med Genet A 182, 2139–2144. [DOI] [PubMed] [Google Scholar]
- 23.Takai H, Jenkinson E, Kabir S, Babul-Hirji R, Najm-Tehrani N, Chitayat DA, Crow YJ, and de Lange T (2016). A POT1 mutation implicates defective telomere end fill-in and telomere truncations in Coats plus. Genes Dev 30, 812–826. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Netravathi M, Kumari R, Kapoor S, Dakle P, Dwivedi MK, Roy SD, Pandey P, Saini J, Ramakrishna A, Navalli D, et al. (2015). Whole exome sequencing in an Indian family links Coats plus syndrome and dextrocardia with a homozygous novel CTC1 and a rare HES7 variation. BMC Med Genet 16, 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Lin H, Gong L, Zhan S, Wang Y, and Liu A (2017). Novel biallelic missense mutations in CTC1 gene identified in a Chinese family with Coats plus syndrome. J Neurol Sci 382, 142–145. [DOI] [PubMed] [Google Scholar]
- 26.Simon AJ, Lev A, Zhang Y, Weiss B, Rylova A, Eyal E, Kol N, Barel O, Cesarkas K, Soudack M, et al. (2016). Mutations in STN1 cause Coats plus syndrome and are associated with genomic and telomere defects. J Exp Med 213, 1429–1440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Wan M, Qin J, Songyang Z, and Liu D (2009). OB fold-containing protein 1 (OBFC1), a human homolog of yeast Stn1, associates with TPP1 and is implicated in telomere length regulation. J Biol Chem 284, 26725–26731. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Chen LY, Redon S, and Lingner J (2012). The human CST complex is a terminator of telomerase activity. Nature 488, 540–544. [DOI] [PubMed] [Google Scholar]
- 29.Mirman Z, Lottersberger F, Takai H, Kibe T, Gong Y, Takai K, Bianchi A, Zimmermann M, Durocher D, and de Lange T (2018). 53BP1-RIF1-shieldin counteracts DSB resection through CST- and Polα-dependent fill-in. Nature 560, 112–116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Mirman Z, Cai S, and de Lange T (2023). CST/Polα/primase-mediated fill-in synthesis at DSBs. Cell Cycle 22, 379–389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Hu C, Rai R, Huang C, Broton C, Long J, Xu Y, Xue J, Lei M, Chang S, and Chen Y (2017). Structural and functional analyses of the mammalian TIN2-TPP1-TRF2 telomeric complex. Cell Res 27, 1485–1502. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Evans R, O’Neill M, Pritzel A, Antropova N, Senior A, Green T, Žídek A, Bates R, Blackwell S, Yim J, et al. (2021). Protein complex prediction with AlphaFold-Multimer. bioRxiv [Google Scholar]
- 33.Takai KK, Kibe T, Donigian JR, Frescas D, and de Lange T (2011). Telomere protection by TPP1/POT1 requires tethering to TIN2. Mol Cell 44, 647–659. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Frescas D, and de Lange T (2014). Binding of TPP1 protein to TIN2 protein is required for POT1a,b protein-mediated telomere protection. J Biol Chem 289, 24180–24187. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Wang H, Ma T, Zhang X, Chen W, Lan Y, Kuang G, Hsu SJ, He Z, Chen Y, Stewart J, et al. (2023). CTC1 OB-B interaction with TPP1 terminates telomerase and prevents telomere overextension. Nucleic Acids Res 51, 4914–4928. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, Tunyasuvunakool K, Bates R, Žídek A, Potapenko A, et al. (2021). Applying and improving AlphaFold at CASP14. Proteins 89, 1711–1721. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Varadi M, Anyango S, Deshpande M, Nair S, Natassia C, Yordanova G, Yuan D, Stroe O, Wood G, Laydon A, et al. (2022). AlphaFold Protein Structure Database: massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res 50, D439–D444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Chen C, Gu P, Wu J, Chen X, Niu S, Sun H, Wu L, Li N, Peng J, Shi S, et al. (2017). Structural insights into POT1-TPP1 interaction and POT1 C-terminal mutations in human cancer. Nat Commun 8, 14929. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Rice C, Shastrula PK, Kossenkov AV, Hills R, Baird DM, Showe LC, Doukov T, Janicki S, and Skordalakes E (2017). Structural and functional analysis of the human POT1-TPP1 telomeric complex. Nat Commun 8, 14928. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Lei M, Podell ER, and Cech TR (2004). Structure of human POT1 bound to telomeric single-stranded DNA provides a model for chromosome end-protection. Nat Struct Mol Biol 11, 1223–1229. [DOI] [PubMed] [Google Scholar]
- 41.Chen LY, Majerská J, and Lingner J (2013). Molecular basis of telomere syndrome caused by CTC1 mutations. Genes Dev 27, 2099–2108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Gu P, and Chang S (2013). Functional characterization of human CTC1 mutations reveals novel mechanisms responsible for the pathogenesis of the telomere disease Coats plus. Aging Cell 12, 1100–1109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Wang F, Podell ER, Zaug AJ, Yang Y, Baciu P, Cech TR, and Lei M (2007). The POT1-TPP1 telomere complex is a telomerase processivity factor. Nature 445, 506–510. [DOI] [PubMed] [Google Scholar]
- 44.Hom RA, and Wuttke DS (2017). Human CST Prefers G-Rich but Not Necessarily Telomeric Sequences. Biochemistry 56, 4210–4218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Nakane T, and Scheres SHW (2021). Multi-body Refinement of Cryo-EM Images in RELION. Methods Mol Biol 2215, 145–160. [DOI] [PubMed] [Google Scholar]
- 46.Zinder JC, Olinares PDB, Svetlov V, Bush MW, Nudler E, Chait BT, Walz T, and de Lange T (2022). Shelterin is a dimeric complex with extensive structural heterogeneity. Proc Natl Acad Sci U S A 119, e2201662119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Zaug AJ, Goodrich KJ, Song JJ, Sullivan AE, and Cech TR (2022). Reconstitution of a telomeric replicon organized by CST. Nature 608, 819–825. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Loayza D, and De Lange T (2003). POT1 as a terminal transducer of TRF1 telomere length control. Nature 423, 1013–1018. [DOI] [PubMed] [Google Scholar]
- 49.Loayza D, Parsons H, Donigian J, Hoke K, and de Lange T (2004). DNA binding features of human POT1: a nonamer 5’-TAGGGTTAG-3’ minimal binding site, sequence specificity, and internal binding to multimeric sites. J Biol Chem 279, 13241–13248. [DOI] [PubMed] [Google Scholar]
- 50.Zhao Y, Sfeir AJ, Zou Y, Buseman CM, Chow TT, Shay JW, and Wright WE (2009). Telomere extension occurs at most chromosome ends and is uncoupled from fill-in in human cancer cells. Cell 138, 463–475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Hirai Y, Masutomi K, and Ishikawa F (2012). Kinetics of DNA replication and telomerase reaction at a single-seeded telomere in human cells. Genes Cells 17, 186–204. [DOI] [PubMed] [Google Scholar]
- 52.Zaug AJ, Lim CJ, Olson CL, Carilli MT, Goodrich KJ, Wuttke DS, and Cech TR (2021). CST does not evict elongating telomerase but prevents initiation by ssDNA binding. Nucleic Acids Res 49, 11653–11665. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Denchi EL, and de Lange T (2007). Protection of telomeres through independent control of ATM and ATR by TRF2 and POT1. Nature 448, 1068–1071. [DOI] [PubMed] [Google Scholar]
- 54.Gong Y, and de Lange T (2010). A Shld1-controlled POT1a provides support for repression of ATR signaling at telomeres through RPA exclusion. Mol Cell 40, 377–387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Kratz K, and de Lange T (2018). Protection of telomeres 1 proteins POT1a and POT1b can repress ATR signaling by RPA exclusion, but binding to CST limits ATR repression by POT1b. J Biol Chem 293, 14384–14392. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Tesmer VM, Brenner KA, and Nandakumar J (2023). Human POT1 protects the telomeric ds-ss DNA junction by capping the 5’ end of the chromosome. Science 381, 771–778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Myler LR, Kinzig CG, Sasi NK, Zakusilo G, Cai SW, and de Lange T (2021). The evolution of metazoan shelterin. Genes Dev 35, 1625–1641. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Tan YZ, Baldwin PR, Davis JH, Williamson JR, Potter CS, Carragher B, and Lyumkis D (2017). Addressing preferred specimen orientation in single-particle cryo-EM through tilting. Nat Methods 14, 793–796. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Iudin A, Korir PK, Somasundharam S, Weyand S, Cattavitello C, Fonseca N, Salih O, Kleywegt GJ, and Patwardhan A (2023). EMPIAR: the Electron Microscopy Public Image Archive. Nucleic Acids Res 51, D1503–D1511. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Weissmann F, Petzold G, VanderLinden R, Huis In ‘t Veld PJ, Brown NG, Lampert F, Westermann S, Stark H, Schulman BA, and Peters JM (2016). biGBac enables rapid gene assembly for the expression of large multisubunit protein complexes. Proc Natl Acad Sci U S A 113, E2564–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Horejsí Z, Takai H, Adelman CA, Collis SJ, Flynn H, Maslen S, Skehel JM, de Lange T, and Boulton SJ (2010). CK2 phospho-dependent binding of R2TP complex to TEL2 is essential for mTOR and SMG1 stability. Mol Cell 39, 839–850. [DOI] [PubMed] [Google Scholar]
- 62.Tang G, Peng L, Baldwin PR, Mann DS, Jiang W, Rees I, and Ludtke SJ (2007). EMAN2: an extensible image processing suite for electron microscopy. J Struct Biol 157, 38–46. [DOI] [PubMed] [Google Scholar]
- 63.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, e42166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Mastronarde DN (2005). Automated electron microscope tomography using robust prediction of specimen movements. J Struct Biol 152, 36–51. [DOI] [PubMed] [Google Scholar]
- 65.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. Nat Methods 14, 331–332. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.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]
- 67.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]
- 68.Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, and Ferrin TE (2004). UCSF Chimera--a visualization system for exploratory research and analysis. J Comput Chem 25, 1605–1612. [DOI] [PubMed] [Google Scholar]
- 69.Emsley P, Lohkamp B, Scott WG, and Cowtan K (2010). Features and development of Coot. Acta Crystallogr D Biol Crystallogr 66, 486–501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Croll TI (2018). ISOLDE: a physically realistic environment for model building into low-resolution electron-density maps. Acta Crystallogr D Struct Biol 74, 519–530. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Adams PD, Afonine PV, Bunkóczi 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 Crystallogr D Biol Crystallogr 66, 213–221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Liebschner D, Afonine PV, Baker ML, Bunkóczi G, Chen VB, Croll TI, Hintze B, Hung LW, Jain S, McCoy AJ, et al. (2019). Macromolecular structure determination using X-rays, neutrons and electrons: recent developments in Phenix. Acta Crystallogr D Struct Biol 75, 861–877. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.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. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Cianfrocco MA, Wong-Barnum M, Youn C, Wagner R, and Leschziner A (2017). COSMIC2. Proceedings of the Practice and Experience in Advanced Research Computing 2017 on Sustainability, Success and Impact 1–5. [Google Scholar]
- 75.Schneider CA, Rasband WS, and Eliceiri KW (2012). NIH Image to ImageJ: 25 years of image analysis. Nat Methods 9, 671–675. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
(A) Co-IPs of Myc-tagged TPP1 proteins with the indicated C-terminal truncations with FLAG-tagged Ctc1, Stn1, and Ten1 from co-transfected 293T cells showing that the C-terminus of TPP1 is required for the CST–TPP1 interaction. Western blots of Myc IPs were probed with anti-Myc and anti-FLAG antibodies.
(B) Co-IPs of Myc-tagged TPP1 with HA-tagged TIN2 and FLAG-tagged Ctc1, Stn1, and Ten1 from co-transfected 293T cells showing that TIN2 competes with CST for TPP1 interaction. Western blots of Myc IPs were probed with anti-Myc, anti-FLAG, and anti-HA antibodies.
(C) Predicted aligned error (PAE) plots of the top four (of 5) ranked AlphaFold-Multimer32 models for Ctc1–TPP1. Green arrowheads indicate high confidence in the position prediction of TPP1 relative to Ctc1. All 5 top-ranked models predicted the interaction with high confidence.
(D) (Left) AlphaFold-Multimer model of Ctc1 OB-folds A, B, and C bound to TPP1TID compared to the crystal structure of TRF2/TIN2/TPP1TID (PDB 5XYF31). TPP1TID is predicted to bind to Ctc1 and TIN2 using the same peptide, providing a structural basis for their competition. (Right) Domain schematics of Ctc1, TPP1, and TIN2 with gray shadows showing the overlapping interaction site on TPP1TID. OB: oligosaccharide/oligonucleotide-binding domain; ARODL: Acidic Rpa1 OB-binding Domain-Like 3-helix bundle; RD: recruitment domain, TID: TIN2-interacting domain; TRFH: telomere repeat factor homology domain.
(A) FSEC analysis of the binding between increasing concentrations of CST and POT1(ESDL)/TPP1 in the absence of telomeric ssDNA showing the concentration dependence of the interaction. RFU: relative fluorescence units.
(B) Native glycerol gradient analysis of ssDNA-bound CST–POT1(ESDL)/TPP1. Pooled fractions are indicated with an asterisk. (Top) Coomassie-stained SDS-PAGE gel (4–12% Bis-Tris gel run in MOPS-SDS buffer, Invitrogen) of glycerol gradient fractions. (Bottom) SYBR Gold-stained native PAGE (4–20% TBE gel run in 0.5x TB buffer, Invitrogen).
(C) Domain architecture of components used to reconstitute an apo CST–POT1(ESDL)/TPP1 complex. TPP1 is fused to the N-terminus of Stn1 and retains part of the TPP1 serine-rich linker.
(D) Coomassie-stained SDS-PAGE gel (4–12% Bis-Tris gel run in MOPS-SDS buffer, Invitrogen) showing the purified CST–POT1(ESDL)/TPP1 fusion complex used for structural analysis.
(E) Negative-stain EM analysis of CST and CST–POT1(ESDL)/TPP1 complexes. (Left) Representative negative-stain EM micrographs. (Middle) Top 25 reference-free 2D-class averages (sorted by number of particles per class from most populous class at top left to least populous class at bottom right) of each complex. (Right) Enlarged views of selected 2D averages to show CST features. Additional density attributable to the addition of POT1(ESDL)/TPP1 is indicated with red arrowheads.
(A) Representative motion-corrected micrograph.
(B) Enlarged view of the area marked in (A) with selected particles circled.
(C) Representative 2D-class averages show high-resolution features and different orientations.
(D) Cryo-EM image-processing pipeline used for the apo CST–POT1(ESDL)/TPP1 complex, including supervised 3D classification with noise decoy classes. Several rounds of heterogeneous refinement were used to select for particles with well-resolved Stn1C and POT1OB1/2 (Table S1).
(E) Final map (contour threshold 0.2) of the apo CST–POT1(ESDL)/TPP1 complex with POT1(ESDL)/TPP1 colored as in Fig. 2A for reference.
(F) Directional FSC plots and sphericity values58 of the reconstruction calculated using the 3D-FSC server (https://3dfsc.salk.edu/).
(G) Plot of the angular distribution of particles in the final reconstruction.
(H) Local resolution estimates for the map (contour threshold 0.2) of apo CST–POT1(ESDL)/TPP1.
(I) Gold-standard (blue) and model-vs-map (red) FSC curves for the reconstruction of apo CST–POT1(ESDL)/TPP1. The map resolution was estimated using the gold-standard FSC and a cut-off criterion of 0.143. The similarity between the resolution based on the gold-standard FSC (0.143 cut-off) and the resolution estimate based on the model-vs-map FSC (0.5 cut-off) suggests no substantial over-fitting.
(J) Cryo-EM map densities for each subunit indicating quality of fit for the apo CST–POT1(ESDL)/TPP1 model. Contour thresholds for Ctc1, Stn1, Ten1, POT1, and TPP1 are 0.1, 0.065, 0.15, 0.1, and 0.05, respectively.
(A) Representative motion-corrected micrograph.
(B) Enlarged view of the area marked in (A) with selected particles circled.
(C) Representative 2D-class averages show high-resolution features and different orientations.
(D) Cryo-EM image-processing pipeline used for the ssDNA-bound CST–POT1(ESDL)/TPP1 complex, including supervised 3D classification with noise decoy classes. Focused 3D classification with a mask was used to select for particles with well-resolved Stn1C and POT1OB1/2 (Table S1).
(E) Final map (contour threshold 0.15) of the ssDNA-bound CST–POT1(ESDL)/TPP1 complex with ssDNA–POT1(ESDL)/TPP1 colored as in Fig. 2A for reference.
(F) Directional FSC plots and sphericity values58 of the reconstruction calculated using the 3D-FSC server (https://3dfsc.salk.edu/).
(G) Plot of the angular distribution of particles in the final reconstruction.
(H) Local resolution estimates for the map (contour threshold 0.15) of the ssDNA-bound CST–POT1(ESDL)/TPP1.
(I) Gold-standard (blue) and model-vs-map (red) FSC curves for the reconstruction of CST–POT1(ESDL)/TPP1. The map resolution was estimated using the gold-standard FSC and a cut-off criterion of 0.143. The similarity between the resolution estimate based on the gold-standard FSC (0.143 cut-off) and the resolution estimate based on the model-vs-map FSC (0.5 cut-off) suggests no substantial over-fitting.
(J) Cryo-EM map densities for each subunit indicating quality of fit for the model of ssDNA-bound CST–POT1(ESDL)/TPP1. Contour thresholds for Ctc1, Stn1, Ten1, POT1, and TPP1 are 0.075, 0.14, 0.1, 0.1, and 0.1, respectively.
(A) Deletion of the TPP1 OB-fold does not affect the CST–POT1(ESDL)/TPP1 interaction. Protein used and FSEC analysis of CST–POT1(ESDL)/TPP1(ΔOB) interaction in the absence (top) and presence (bottom) of telomeric ssDNA. Traces without CST are shown as dashed lines. RFU: relative fluorescence units.
(B) Electrostatic surface from Fig. 3A with map density (contour threshold 0.1) of hinge shown as mesh.
(C) Coomassie-stained SDS-PAGE gel (4–12% Bis-Tris gel run in MOPS-SDS buffer, Invitrogen) showing the purified POT1/TPP1 proteins (untreated or λ-treated) used for FSEC and MP analysis. Descriptions of the POT1 variant sequences are given in Fig. 3E.
(D) FSEC analysis of the phosphorylation-dependent interaction between CST and POT1(WT)/GFP-TPP1 (top, black) and POT1(ESDL)/GFP-TPP1 (bottom, blue) in the absence of telomeric ssDNA (see also Fig. 3B). Traces without CST are shown as dashed lines, and traces containing POT1/TPP1 that have been treated with λ protein phosphatase (λPP) are shown in red. RFU: relative fluorescence units. The chromatograms for the WT and ESDL mock samples are shown in Fig. 1E and are reproduced here as the controls for the phosphatase experiment, which was performed simultaneously.
(E) MP histograms used for the quantification shown in Fig. 3F. Each panel shows histograms from three technical replicates with Gaussian curves auto-fitted by DiscoverMP software. Inset tables show σ (standard deviation of values within peak) and and total counts per peak for each replicate. Total peak counts were used to quantify the proportion of bound POT1/TPP1, the ratio of total counts in peak 3 (CST–POT1/TPP1 complex) divided by total counts of POT1/TPP1 (peak 1 and peak 3).
(A) As in Fig. 4B but individual panels include the cryo-EM map density for the apo CST–POT1(ESDL)/TPP1 complex shown as mesh (contour thresholds 0.3, 0.3, 0.7, 0.5, and 0.5, respectively).
(B) Co-IPs of Strep-tagged POT1 (or mPOT1b) protein variants and FLAG-tagged Ctc1, Stn1, and Ten1 from co-transfected 293T cells showing that the corresponding S328L CP mutation in mPOT1b disrupts the CST interaction. Western blots of Strep IPs were probed with anti-Strep and anti-FLAG antibodies.
(C) Double-filter DNA-binding assays showing high affinity [GGTTAG]3 binding by POT1/TPP1 alone and in complex with CST. Left: Representative membrane scans. The nitrocellulose membrane is labeled as “Bound” and the Hybond-XL membrane is labeled as “Free.” Right: Quantitation of binding data. Data were fitted using the “one site – specific binding” model on GraphPad Prism 9. Error bars represent SEM from two technical replicate experiments.
(A) Multiple views of CST–Polα/primase in RC (left) and PIC (right) conformations. CST–Polα/primase structures are shown in cartoon representation with a transparent surface.
(B) Superposition of the ssDNA-bound CST–POT1(ESDL)/TPP1 structure with the structures of the RC (left, see also Video S3) and PIC (right, see also Video S2) complexes showing additional views compared to Fig. 5C. The clash between the POT1HJRL and PRIM2 is indicated.
(C) Multi-body analysis of CST–Polα/primase in the RC conformation. Polα/primase was designated body 1 and CST was designated body 2 with the corresponding masks shown. Histograms of the projections of the relative orientations onto the corresponding components show a unimodal distribution, consistent with continuous flexibility rather than discrete states. The first three principal components account for 61% of the variance in the data. Reconstructed maps from the extreme ends are shown in red and blue for each of the first three principal components with arrows indicating the direction of motion (see also Video S4).
(D) PAE plots of the top two (of 5) ranked AlphaFold-Multimer32 models for Ctc1/POLA2. Green arrowheads indicate high confidence in the position prediction of POLA2 relative to Ctc1. All 5 top-ranked models predicted the interaction with high confidence.
(E) AlphaFold-Multimer model of Ctc1 bound to POLA2NTD. Residues with known CP mutations are shown as spheres. Red spheres indicate mutations previously shown to disrupt Polα/primase association41; black spheres indicate mutations with unknown mechanism.
(A) Inhibition of C-strand synthesis by POT1(ESDL)/TPP1 is enhanced by its DNA-binding domain. Reactions of 50 nM CST–Polα/primase with 50 nM single-stranded 9xTEL DNA template. LC1, LC2, LC3: oligonucleotide loading controls.
(B) Quantification of the initial extension products, bands 21–23.
(C) Quantification of all products, bands 21–40, gave similar IC50 values as in (B).
(D) Design of pre-primed templates to monitor primer extension by Polα. Red sequence is RNA.
(E) Validation of the pre-primed template reactions by performing reactions in the absence of ribonucleotides. Most reaction products were dependent on both template and primer. The 9xTEL template products (which were formed in the absence of the primer strand) were longer than the template, indicating that they resulted from self-priming of the DNA template.
(F) CST–Polα extension of pre-primed templates. M: telomerase reaction products as size markers. Stronger inhibition of C-strand synthesis by POT1(ESDL)/TPP1 persists under conditions that limit re-priming (50 nM template and 5 nM CST–Polα/primase, a ten-fold higher template/enzyme ratio than used in Fig. 6E).
(G) Quantification of the data in (F) (left), normalized to reactions without added POT1/TPP1. IC50 values with error bars indicating uncertainty of the fit to the data points.
(H) Quantification of the data in (F) (right), normalized to reactions without added POT1/TPP1. IC50 values were not calculated because inhibition was incomplete at highest POT1/TPP1 concentrations.
360° rotation of the cryo-EM map and model from apo and ssDNA-bound CST–POT1(ESDL)/TPP1. Colors are the same as in Fig. 2.
Rotation movie showing superposition of CST–POT1(ESDL)/TPP1–ssDNA complex (surface representation; solid colors as in Fig. 2) with CST–Polα/primase PIC (surface representation at 50% opacity; cartoon model shown underneath). Colors of Polα/primase are the same as in15. PRIM1-salmon; PRIM2-light orange; POLΑ1-light green (Ctc1-recognition loop-lime green); POLA2-yellow.
Rotation movie showing superposition of CST–POT1(ESDL)/TPP1–ssDNA complex (surface representation; solid colors as in Fig. 2) with CST–Polα/primase RC (surface representation at 50% opacity; cartoon model shown underneath). Colors of Polα/primase are the same as in15. PRIM1-salmon; PRIM2-light orange; POLΑ1-light green (Ctc1-recognition loop-lime green); POLA2-yellow.
Principal components 1, 2, and 3, respectively.
Data Availability Statement
The cryo-EM maps have been deposited at the Electron Microscopy Data Bank under accession codes EMD-40659 (apo CST–POT1/TPP1 complex) and EMD-40660 (ssDNA–CST–POT1/TPP1 complex), and the coordinates have been deposited in the Protein Data Bank under accession codes PDB 8SOJ and PDB 8SOK, respectively. Motion-corrected micrographs have been deposited in the Electron Microscopy Public Image Archive59 under accession codes EMPIAR-12046 and EMPIAR-12047, respectively. This study also uses data previously reported2 and deposited under the accession code EMPIAR-11131. The data are publicly available as of the date of publication.
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.