Telomeres are essential for chromosome integrity and protection, and their maintenance requires the ribonucleoprotein enzyme telomerase. Previously, we have shown that human telomerase reverse transcriptase (hTERT) contains a bipartite nuclear localization signal (NLS; residues 222–240) that is responsible for nuclear import, and that Akt-mediated phosphorylation of residue S227 is important for efficient nuclear import of hTERT. Here, we show that hTERT binds to importin-α proteins through the bipartite NLS and that this heterodimer then forms a complex with importin-β proteins to interact with the nuclear pore complex. Depletion of individual importin-α proteins results in a failure of hTERT nuclear import, and the resulting cytoplasmic hTERT is degraded by ubiquitin-dependent proteolysis. Crystallographic analysis reveals that the bipartite NLS interacts with both the major and minor sites of importin-α proteins. We also show that Akt-mediated phosphorylation of S227 increases the binding affinity for importin-α proteins and promotes nuclear import of hTERT, thereby resulting in increased telomerase activity. These data provide details of a binding mechanism that enables hTERT to interact with the nuclear import receptors and of the control of the dynamic nuclear transport of hTERT through phosphorylation.

Telomeres, the nucleoprotein structures found at the physical ends of eukaryotic chromosomes, are essential for chromosome protection (Blackburn, 2001; Smogorzewska and de Lange, 2004). Mammalian telomeres comprise duplex tandem TTAGGG repeats that terminate in a 3′ single-stranded G overhang (Griffith et al., 1999), and they are coated with the shelterin complex (de Lange, 2005). Functional inhibition of shelterin has been shown to induce telomere loss, chromosome end-to-end fusion and inappropriate DNA repair reactions (Palm and de Lange, 2008); this inhibition also evokes a telomere damage response (Sfeir and de Lange, 2012). Telomere elongation is dependent on telomerase, a ribonucleoprotein enzyme that comprises the telomerase reverse transcriptase (hTERT in humans), the telomerase RNA component (TERC) and several additional proteins (Autexier and Lue, 2006; Bianchi and Shore, 2008). Telomerase offsets telomere erosion that results from a DNA-end replication problem by adding telomere repeats to the chromosome termini (Lingner et al., 1995). In humans, telomerase activity occurs primarily during the early stage of embryogenesis but, later in life, is strongly repressed in most normal somatic cells (Cong et al., 2002). However, telomerase is highly active in a majority of human cancers, indicating that the activation of telomerase is crucial for tumorigenesis (Kim et al., 1994).

Telomerase activity is regulated by hTERT at both the transcriptional and post-translational levels, implying that hTERT is a key component of the telomerase complex (Nakamura and Cech, 1998; Wang et al., 1998). For the assembly of catalytically active telomerase, cytoplasmic hTERT must be translocated to the nucleus through the nuclear pore complex (NPC). Proteins smaller than ∼40 kDa in size can pass through the NPC by simple diffusion, whereas nuclear localization of larger proteins requires an active nuclear transport pathway (Poon and Jans, 2005; Stewart, 2007). In the classic pathway, importin-α proteins recognize a nuclear localization signal (NLS) displayed by a cargo protein. The heterodimeric importin-α–NLS-cargo complex interacts with importin-β proteins and is translocated across the NPC through the interaction with nucleoporins (Cingolani et al., 1999,, 2002; Chook and Süel, 2011). Once translocation through the NPC is complete, RanGTP binds to importin-β proteins and induces dissociation of importin β from the importin-α–NLS complex, subsequently dissociating the cargo from importin α (Forwood et al., 2008; Lonhienne et al., 2009). Importin-α proteins can interact with two major classes of NLS – a monopartite NLS comprising a single cluster of basic amino acids and a bipartite NLS comprising two clusters of basic amino acids (Lange et al., 2007). The architecture of the importin-α protein comprises armadillo repeats, which includes three α-helices. Importin-α proteins contain two NLS-binding pockets – the major site and the minor site (Peifer et al., 1994; Conti et al., 1998). The monopartite NLS primarily binds to the major site, whereas the bipartite NLS interacts simultaneously with both sites (Fontes et al., 2000; Conti and Kuriyan, 2000). The NLS-binding pockets of importin α can also be occupied by the N-terminal importin-β-binding (IBB) domain, which contains basic amino acids (Harreman et al., 2003; Kobe, 1999), suggesting that NLS binding to importin α is regulated by an internal autoinhibitory mechanism.

Because the size of hTERT (∼124 kDa) precludes a passive nuclear transport mechanism, it should undergo active nuclear import, thereby regulating cellular telomerase activity. We have previously reported that hTERT contains a bipartite NLS (residues 222–240) and that Akt-mediated phosphorylation of residue S227 on hTERT plays an important role in the nuclear translocation of hTERT (Chung et al., 2012). Interestingly, S227 is located between two clusters of basic amino acids in the bipartite NLS. Thus, it is of interest to investigate whether phosphorylation of S227 affects the interaction between the hTERT NLS and the import receptors. Recently, importin 7 has been shown to interact with the hTERT NLS and to mediate nuclear transport of hTERT (Frohnert et al., 2014). However, the molecular mechanism regulating the nuclear import of hTERT remains unclear.

In this work, we demonstrate that nuclear import of hTERT is dependent on importin α and importin β proteins, and the cellular GTPase Ran. The hTERT NLS interacts with multiple importin-α-family members, and depletion of individual importin-α proteins inhibits the nuclear transport of hTERT. Crystallography analysis of a complex comprising importin-α5 (also known as KPNA1) and the NLS of hTERT reveals that the bipartite NLS binds to both the major and minor binding sites of importin α5. We also show that Akt-mediated phosphorylation of S227 increases the binding affinity of hTERT for importin-α proteins, resulting in effective nuclear localization of hTERT. These results provide a hTERT-NLS-binding mode in importin α5 and a nuclear transport control of hTERT through Akt activation.

A bipartite NLS is essential for nuclear import of hTERT

We have previously identified a bipartite NLS (222-RRRGGSASRSLPLPKRPRR-240) that is responsible for the nuclear import of hTERT (Chung et al., 2012). Mutations in the two clusters of basic amino acids in the bipartite NLS results in reduced nuclear localization of hTERT. We have also shown that Akt-mediated phosphorylation at S227 is necessary for efficient nuclear translocation of hTERT (Chung et al., 2012). To determine whether nuclear localization of hTERT is affected by different cellular backgrounds, several cell lines – including MCF7, H1299 and U2OS cells – were transfected with Flag–hTERT and various mutant constructs (Flag-7A, Flag-S227A and Flag-S227E, see Fig. 1A), and then subjected to immunofluorescence analyses using an antibody against Flag (supplementary material Fig. S1). The results indicated that both the bipartite NLS and Akt-mediated phosphorylation at S227 are required for efficient nuclear import of hTERT, suggesting that the mechanism of nuclear localization of hTERT is not cell-type specific.

To determine whether the hTERT NLS contains all of the determinants that are sufficient to confer nuclear import, the extended bipartite NLS (residues 212–250) was fused to green fluorescent protein (GFP, 26 kDa) and β-galactosidase (β-gal, 112 kDa) to construct the GFP–NLS–β-gal expression vector (Fig. 1A). Because the expected size of the fusion protein (∼143 kDa) prevents passive diffusion across the NPC, any nuclear accumulation of the protein should be the result of active transport by the soluble receptors. To determine the subcellular localization of the fusion protein, MCF7 cells were transfected with various GFP–NLS–β-gal constructs and subjected to immunofluorescence staining (Fig. 1A). The majority of the GFP–β-gal signal was detected in the cytoplasm (86.7% cytoplasmic), whereas the fusion protein that included the wild-type NLS (GFP–NLS–β-gal) showed a pronounced nuclear localization (88.6% nuclear; Fig. 1B,C). When the two clusters of basic amino acids in the NLS were mutated to alanine residues (7A and S227A/7A), the fusion protein signals were mostly found in the cytoplasm (81.2% cytoplasmic for 7A, and 89.51% cytoplasmic for S227A/7A). When S227 was mutated to alanine, the nuclear fluorescence signals were considerably decreased (49.5% nuclear for S227A). Interestingly, the phosphorylation-mimicking mutant S227E was mostly localized to the nucleus (79.1% nuclear). Taken together, these results demonstrate that the NLS sequence plays a major role in the nuclear translocation of hTERT.

Fig. 1.

See next page for legend.

The bipartite NLS of hTERT is sufficient to confer nuclear import to a heterologous protein. (A) Schematic representation of the hTERT NLS fusion proteins. The extended bipartite NLS (residues 212 and 250) was fused to GFP and β-galactosidase (β-gal). The mutant hTERT NLS constructs in which the basic amino acid residues are mutated to alanine are shown. To generate a phosphorylation-mimicking mutant, S227 was replaced by glutamate (S227E). (B) MCF7 cells were transfected with various GFP–NLS–β-gal constructs and subjected to immunofluorescent staining, followed by observation using fluorescence microscopy. The nuclei were stained with DAPI (blue). (C) After transfection, MCF7 cells were subjected to quantitative analysis of the location of the fusion proteins within the cells. In each experiment, more than 100 cells were evaluated quantitatively for nuclear (N, blue bars), nuclear-cytoplasmic (N+C, red bars) and cytoplasmic (C, green bars) fluorescence. Error bars show the standard deviation from the mean of three independent experiments. (D,E) MCF7 cells that had been transfected with various GFP–NLS–β-gal constructs were left untreated (D) or treated with 10 μM LY294002 (E) for 2 h and subjected to immunofluorescent staining, followed by observation using fluorescence microscopy. The nuclei were stained with DAPI (blue). (F) After transfection, MCF7 cells were subjected to quantitative analysis of the location of the fusion proteins within the cells. Scale bars: 5 μm (B,D,E).

Fig. 1.

See next page for legend.

The bipartite NLS of hTERT is sufficient to confer nuclear import to a heterologous protein. (A) Schematic representation of the hTERT NLS fusion proteins. The extended bipartite NLS (residues 212 and 250) was fused to GFP and β-galactosidase (β-gal). The mutant hTERT NLS constructs in which the basic amino acid residues are mutated to alanine are shown. To generate a phosphorylation-mimicking mutant, S227 was replaced by glutamate (S227E). (B) MCF7 cells were transfected with various GFP–NLS–β-gal constructs and subjected to immunofluorescent staining, followed by observation using fluorescence microscopy. The nuclei were stained with DAPI (blue). (C) After transfection, MCF7 cells were subjected to quantitative analysis of the location of the fusion proteins within the cells. In each experiment, more than 100 cells were evaluated quantitatively for nuclear (N, blue bars), nuclear-cytoplasmic (N+C, red bars) and cytoplasmic (C, green bars) fluorescence. Error bars show the standard deviation from the mean of three independent experiments. (D,E) MCF7 cells that had been transfected with various GFP–NLS–β-gal constructs were left untreated (D) or treated with 10 μM LY294002 (E) for 2 h and subjected to immunofluorescent staining, followed by observation using fluorescence microscopy. The nuclei were stained with DAPI (blue). (F) After transfection, MCF7 cells were subjected to quantitative analysis of the location of the fusion proteins within the cells. Scale bars: 5 μm (B,D,E).

To examine whether Akt-dependent phosphorylation at S227 is required to mediate nuclear import of hTERT, we determined the effect of an Akt inhibitor (LY294002, which inhibits all Akt isoforms) on subcellular localization of the fusion protein. When cells were treated with the inhibitor, nuclear accumulation of GFP–NLS–β-gal was substantially reduced (Fig. 1D,E). By contrast, subcellular localization of GFP–S227A–β-gal was not affected by treatment with the inhibitor (Fig. 1F), highlighting the requirement of Akt-dependent phosphorylation at S227.

Nuclear import of hTERT is mediated by the importin-α–importin-β pathway

It has been well documented that a protein containing a classic basic NLS is imported by a heterodimeric import receptor comprising importin-α and importin-β subunits (Chook and Süel, 2011; Cingolani et al., 2002,, 1999). Seven members of the importin-α family have been described in humans; these are classified into three subfamilies, and each is encoded by a different gene (Kelley et al., 2010). Because the nuclear import of hTERT is mediated by a classic bipartite NLS, we wanted to determine whether hTERT directly interacts with importin-α proteins. Importin α1, importin α3, importin α5 and importin β1 (encoded by KPNA2, KPNA4, KPNA1 and KPNB1, respectively) were fused to glutathione S-transferase (GST); the fusion proteins were purified and then used in an in vitro binding assay. GST–importin-α proteins, but not the control GST, bound to Flag–NLS, indicating that hTERT interacts with importin-α proteins through the NLS (Fig. 2A). We also found that Flag–NLS was precipitated by GST–importin-β1, suggesting that the NLS of hTERT can form a complex with importin-α–importin-β1 dimers. Interestingly, the S227A/7A mutation severely impaired the interactions between hTERT and importin-α proteins (Fig. 2A). Likewise, GST–NLS, but not the control GST or GST–S227A/7A mutant, precipitated hemagglutinin (HA)-tagged importin α1, importin α3 and importin α5 when they were expressed in MCF7 cells (Fig. 2B). Consistently, importin β1 was also found to associate with GST–NLS. To determine whether the NLS of hTERT associates with importin-α proteins in vivo, MCF7 cells were transfected with Flag–NLS and HA-tagged importin-α proteins, and then subjected to immunoprecipitation. All importin proteins were immunoprecipitated by Flag–NLS (Fig. 2C). Importin β1 was also recovered by Flag–NLS.

Fig. 2.

The hTERT NLS interacts physically with importin α. (A) GST–importin-α1, GST–importin-α3, GST–importin-α5 and GST–importin-β1 were affinity-purified and incubated with lysates that had been prepared from cells expressing Flag–NLS or Flag–S227A/7A, followed by immunoblotting with anti-Flag antibody. The purified GST fusion proteins were visualized by Coomassie staining and are indicated with arrowheads. Molecular mass makers are shown in kDa. (B) GST–NLS and GST–S227A/7A were purified and incubated with lysates prepared from cells expressing HA–importin-α1, HA–importin-α3, HA–importin-α5 and HA–importin-β1; the HA-tagged proteins were then detected. HA-tagged importin proteins are indicated with arrowheads. The purified GST fusion proteins were visualized by Coomassie staining. (C) MCF7 cells that had been transfected with Flag–NLS and either HA–importin-α proteins or HA–importin-β1 were subjected to immunoprecipitation (IP) with anti-Flag antibody, followed by immunoblotting with anti-HA antibody. HA-tagged importin proteins are indicated with arrowheads. (D) MCF7 cells transfected with Flag–NLS, HA–importin-α proteins and HA–importin-β1 were subjected to immunoprecipitation with anti-Flag antibody, followed by immunoblotting with anti-HA antibody. (E) MCF7 cells were subjected to immunoprecipitation with an anti-hTERT antibody; the immunoprecipitates were then probed for endogenous importin-α proteins and importin-β1. IgG antibody was used as a negative control. In B,C,D, asterisks mark the position of nonspecific immunoglobulin chains.

Fig. 2.

The hTERT NLS interacts physically with importin α. (A) GST–importin-α1, GST–importin-α3, GST–importin-α5 and GST–importin-β1 were affinity-purified and incubated with lysates that had been prepared from cells expressing Flag–NLS or Flag–S227A/7A, followed by immunoblotting with anti-Flag antibody. The purified GST fusion proteins were visualized by Coomassie staining and are indicated with arrowheads. Molecular mass makers are shown in kDa. (B) GST–NLS and GST–S227A/7A were purified and incubated with lysates prepared from cells expressing HA–importin-α1, HA–importin-α3, HA–importin-α5 and HA–importin-β1; the HA-tagged proteins were then detected. HA-tagged importin proteins are indicated with arrowheads. The purified GST fusion proteins were visualized by Coomassie staining. (C) MCF7 cells that had been transfected with Flag–NLS and either HA–importin-α proteins or HA–importin-β1 were subjected to immunoprecipitation (IP) with anti-Flag antibody, followed by immunoblotting with anti-HA antibody. HA-tagged importin proteins are indicated with arrowheads. (D) MCF7 cells transfected with Flag–NLS, HA–importin-α proteins and HA–importin-β1 were subjected to immunoprecipitation with anti-Flag antibody, followed by immunoblotting with anti-HA antibody. (E) MCF7 cells were subjected to immunoprecipitation with an anti-hTERT antibody; the immunoprecipitates were then probed for endogenous importin-α proteins and importin-β1. IgG antibody was used as a negative control. In B,C,D, asterisks mark the position of nonspecific immunoglobulin chains.

Because the importin-α–NLS complex interacts with importin β1, we further verified the formation of a complex comprising the importin-α–NLS heterodimer and importin β1. As shown in Fig. 2D, HA–importin-β1 was immunoprecipitated by Flag–NLS only when one of the importin-α proteins was co-expressed. To confirm direct interactions with endogenous proteins, we performed immunoprecipitation analyses using MCF7 cells and found that endogenous importin α1, importin α3, importin α5 and importin β1 could be immunoprecipitated by endogenous hTERT (Fig. 2E). Taken together, these results suggest that the nuclear import of hTERT is mediated by the importin-α–importin-β pathway.

Depletion of importin α inhibits nuclear import of hTERT and reduces telomerase activity

To examine the involvement of importin-α proteins in a more physiological setting, endogenous importin-α proteins were depleted using two different small interfering (si)RNA duplexes specific to each importin α. Depletion of importin-α proteins led to a clear reduction in the level of endogenous hTERT (Fig. 3A). This reduction in the level of hTERT was rescued by treatment with MG132, suggesting that degradation of hTERT is mediated by the proteasome. To investigate these findings in more detail, we determined the effect of depleting importin-α proteins on the subcellular localization of hTERT. Nuclear and cytosolic extracts were separately collected from MCF7 cells and subjected to immunoblot analysis to assess the level of endogenous hTERT. In the presence of MG132, the majority of hTERT was localized to the nucleus in the control siRNA cells (Fig. 3B). When importin-α proteins were depleted, substantial amounts of hTERT were detected in the cytoplasm. These results indicate that nuclear import of hTERT is inhibited by the depletion of importin-α proteins. We also examined whether the depletion of importin α is involved in regulating the half-life of hTERT. MCF7 cells were transfected with Flag–hTERT and treated with cycloheximide to inhibit new protein synthesis. The stability of hTERT was monitored by immunoblotting with an antibody against Flag. Depletion of importin-α proteins substantially reduced the half-life of Flag–hTERT, as compared with that in control siRNA cells (supplementary material Fig. S2A,B). The protein levels were rescued when the transfected cells were treated with MG132.

Fig. 3.

Depletion of importin α inhibits nuclear import of hTERT and reduces telomerase activity. (A) MCF7 cells were transfected with control siRNA (siControl), two siRNAs against importin α1 (siImportin α1-1 and siImportin α1-2), importin α3 (siImportin α3-1 and siImportin α3-2) and importin α5 (siImportin α5-1 and siImportin α5-2), and treated with or without 10 μM MG132 for 4 h. The protein levels of endogenous hTERT, importin-α proteins and importin-β1 were measured by immunoblotting as indicated. (B) MCF7 cells were transfected with one of the siRNAs against the indicated importin-α proteins and treated with 10 μM MG132 for 4 h. Nuclear and cytosolic extracts were separately collected, and the protein levels of endogenous hTERT, importin-α proteins and importin-β1 were measured by immunoblotting as indicated. (C) MCF7 cells were transfected with Flag–hTERT, HA–ubiquitin (Ub) and an siRNA against the indicated importin-α proteins, and treated with 10 µM MG132 for 4 h. Immunoprecipitation was performed with an anti-Flag antibody before probing with an anti-HA antibody. Molecular mass markers are shown in kDa. (D) MCF7 cells that had been transfected with an siRNA the indicated importin-α proteins were analyzed for telomerase activity by using the TRAP assay. To test RNA-dependent extension, RNase A (0.25 mg/ml) was added to the extracts before the primer extension reaction. ITAS represents the internal telomerase assay standard. (E) MCF7 cells that had been transfected with an siRNA against the indicated importin-α proteins were treated with or without 10 μM MG132 for 4 h and subjected to immunofluorescent staining with an anti-hTERT antibody (green), which was observed by using fluorescence microscopy. The nuclei were stained with DAPI (blue). Scale bars: 5 μm.

Fig. 3.

Depletion of importin α inhibits nuclear import of hTERT and reduces telomerase activity. (A) MCF7 cells were transfected with control siRNA (siControl), two siRNAs against importin α1 (siImportin α1-1 and siImportin α1-2), importin α3 (siImportin α3-1 and siImportin α3-2) and importin α5 (siImportin α5-1 and siImportin α5-2), and treated with or without 10 μM MG132 for 4 h. The protein levels of endogenous hTERT, importin-α proteins and importin-β1 were measured by immunoblotting as indicated. (B) MCF7 cells were transfected with one of the siRNAs against the indicated importin-α proteins and treated with 10 μM MG132 for 4 h. Nuclear and cytosolic extracts were separately collected, and the protein levels of endogenous hTERT, importin-α proteins and importin-β1 were measured by immunoblotting as indicated. (C) MCF7 cells were transfected with Flag–hTERT, HA–ubiquitin (Ub) and an siRNA against the indicated importin-α proteins, and treated with 10 µM MG132 for 4 h. Immunoprecipitation was performed with an anti-Flag antibody before probing with an anti-HA antibody. Molecular mass markers are shown in kDa. (D) MCF7 cells that had been transfected with an siRNA the indicated importin-α proteins were analyzed for telomerase activity by using the TRAP assay. To test RNA-dependent extension, RNase A (0.25 mg/ml) was added to the extracts before the primer extension reaction. ITAS represents the internal telomerase assay standard. (E) MCF7 cells that had been transfected with an siRNA against the indicated importin-α proteins were treated with or without 10 μM MG132 for 4 h and subjected to immunofluorescent staining with an anti-hTERT antibody (green), which was observed by using fluorescence microscopy. The nuclei were stained with DAPI (blue). Scale bars: 5 μm.

To investigate whether hTERT is ubiquitylated before it is degraded in a proteasome-dependent manner, MCF7 cells were co-transfected with Flag–hTERT and HA–ubiquitin. To visualize ubiquitin-modified hTERT, immunoprecipitations of the Flag tag were evaluated by immunoblotting with an antibody against HA. Ubiquitylated hTERT was barely detected in the control siRNA cells in the presence of MG132, whereas depletion of any of the importin-α proteins resulted in a marked increase in hTERT ubiquitylation (Fig. 3C). Because mutations in the two clusters of basic amino acids and S227 (7A/S227A) inhibit nuclear import of hTERT, the resulting cytoplasmic hTERT could be ubiquitylated before it is degraded by the proteasome. To test this possibility, MCF7 cells were co-transfected with HA–ubiquitin and either Flag–hTERT or Flag–7A/S227A. Ubiquitylation of the 7A/S227A construct was higher than that of wild-type hTERT in the presence MG132 (supplementary material Fig. S2C). These results demonstrate that, when nuclear import is inhibited by depletion of importin-α proteins or by mutations in the NLS, the resulting cytoplasmic hTERT is rapidly ubiquitylated and degraded by the proteasome.

We next examined whether depletion of importin-α proteins influences telomerase activity. As shown in Fig. 3D, telomerase activity was considerably reduced in cells in which importin-α proteins had been knocked down compared with that in control siRNA cells. To determine the effect of importin-α depletion on the nuclear localization of hTERT, we performed an indirect immunofluorescence assay using a confocal laser-scanning microscope. Endogenous hTERT predominantly localized to the nucleus in the control siRNA cells, whereas depletion of importin-α family members resulted in a substantial decrease in the total intensity of hTERT fluorescence (Fig. 3E). This reduction in hTERT fluorescence could be rescued by treatment with MG132. We also found that cytoplasmic accumulation of endogenous hTERT was increased in the presence of MG132 (Fig. 3E). These results support the idea that hTERT fails to be translocated into the nucleus when importin-α proteins are depleted and that hTERT is subsequently degraded by the proteasome. Because the specificity of the antibody against hTERT is crucial to the interpretation of the data presented, we verified that small hairpin (sh)RNA-mediated depletion of hTERT reduced the immunofluorescence staining pattern (supplementary material Fig. S2D). Depletion of importin β1 was also found to inhibit nuclear import of hTERT and to reduce telomerase activity (supplementary material Fig. S4A,B), suggesting that importin β1 is required for nuclear localization of hTERT. By contrast, overexpression of importin-α proteins and importin β1 had no effect on the nuclear import efficiency of hTERT and telomerase activity (data not shown).

Nuclear import of hTERT is dependent on Ran

Conventional importin-α–importin-β-dependent nuclear transport through the NPC is dependent on the nucleotide-bound state of the small GTPase Ran, which affects the disassembly of importin–cargo complexes (Forwood et al., 2008; Lonhienne et al., 2009). To examine the dependence of the subcellular localization of hTERT on Ran, a dominant-negative mutant of Ran (RanQ69L) that is unable to hydrolyze GTP to GDP was used to inhibit importin-α–importin-β-dependent nuclear protein import (Dickmanns et al., 1996; Palacios et al., 1996). MCF7 cells expressing V5-tagged Ran (V5–Ran) or V5-tagged RanQ69L (V5–RanQ69L) were subjected to immunofluorescence staining in order to monitor the localization of endogenous hTERT. A clear accumulation of hTERT in the nucleus was observed in cells that expressed the empty vector or V5–Ran (Fig. 4A). By contrast, when RanQ69L was overexpressed, the hTERT signal was detected in a diffuse pattern throughout the nucleus and cytoplasm in the presence of MG132. These results demonstrate that Ran plays an essential role in the nuclear import of hTERT.

Fig. 4.

Overexpression of a dominant-negative mutant of Ran (RanQ69L) inhibits nuclear localization of hTERT. (A) MCF7 cells that had been transfected with V5–Ran or V5–RanQ69L were treated with 10 μM MG132 for 4 h and subjected to immunofluorescent staining with anti-hTERT (green) or anti-V5 (red) antibodies, which was observed by using fluorescence microscopy. The nuclei were stained with DAPI (blue). Scale bars: 5 μm. (B) MCF7 cells that had been transfected with Flag–hTERT and either V5–Ran or V5–RanQ69L were treated with or without 10 μM MG132 for 4 h. The protein levels were measured by immunoblotting, as indicated. Immunoprecipitation (IP) was performed with anti-Flag antibody before detecting endogenous importin-α and importin-β1 proteins. Asterisks mark the positions of nonspecific immunoglobulin chains. (C) MCF7 cells transfected with V5-Ran or V5-RanQ69L were treated with or without 10 μM MG132 for 4 h and analyzed for telomerase activity by using the TRAP assay. To test RNA-dependent extension, RNase A (0.25 mg/ml) was added to the extracts before the primer extension reaction. ITAS represents the internal telomerase assay standard.

Fig. 4.

Overexpression of a dominant-negative mutant of Ran (RanQ69L) inhibits nuclear localization of hTERT. (A) MCF7 cells that had been transfected with V5–Ran or V5–RanQ69L were treated with 10 μM MG132 for 4 h and subjected to immunofluorescent staining with anti-hTERT (green) or anti-V5 (red) antibodies, which was observed by using fluorescence microscopy. The nuclei were stained with DAPI (blue). Scale bars: 5 μm. (B) MCF7 cells that had been transfected with Flag–hTERT and either V5–Ran or V5–RanQ69L were treated with or without 10 μM MG132 for 4 h. The protein levels were measured by immunoblotting, as indicated. Immunoprecipitation (IP) was performed with anti-Flag antibody before detecting endogenous importin-α and importin-β1 proteins. Asterisks mark the positions of nonspecific immunoglobulin chains. (C) MCF7 cells transfected with V5-Ran or V5-RanQ69L were treated with or without 10 μM MG132 for 4 h and analyzed for telomerase activity by using the TRAP assay. To test RNA-dependent extension, RNase A (0.25 mg/ml) was added to the extracts before the primer extension reaction. ITAS represents the internal telomerase assay standard.

We next determined the effect of RanQ69L on the stability of hTERT. Overexpression of RanQ69L resulted in a clear reduction in the level of Flag–hTERT compared with that in cells expressing the empty vector or wild-type Ran (Fig. 4B). However, the RanQ69L-induced reduction in hTERT was rescued in the presence of MG132. Because the hTERT NLS interacts with importin-α family members and importin β1 for its nuclear import, we determined whether overexpression of RanQ69L affects the formation of heterotrimeric complex. Overexpression of RanQ69L severely impaired the interactions of hTERT with importin-α proteins and importin β1, indicating that Ran is required for the nuclear import of hTERT (Fig. 4B). When cells that expressed RanQ69L were analyzed for the telomerase activity, overexpression of RanQ69L was seen to lead to a substantial reduction in telomerase activity compared with that in the control cells (Fig. 4C). Moreover, this reduction in telomerase activity was not rescued by treatment with MG132, suggesting that cytoplasmic hTERT is catalytically nonfunctional for telomerase activity.

Depletion of importin α reduces telomerase activity by inhibiting the assembly of telomerase in the nucleus

Telomerase undergoes a highly elaborate, stepwise process for the assembly of the telomerase holoenzyme (Venteicher et al., 2008,, 2009; Lee et al., 2014). After translocation to the nucleus, hTERT associates with telomerase RNA component (TERC) and dyskerin to generate catalytically active telomerase. We determined the effect of importin-α depletion on the telomerase ribonucleoprotein (RNP) assembly. The protein level of hTERT was reduced by depletion of importin-α proteins, whereas the protein level and nuclear localization of dyskerin were unaffected (Fig. 5A). We also found that the interaction between hTERT and dyskerin was dramatically reduced by depletion of importin-α proteins (Fig. 5A). We next examined whether dyskerin associates with catalytically active telomerase when importin α is depleted. Importin-α-knockdown cells were immunoprecipitated with a dyskerin-specific antibody and analyzed for telomerase activity. Depletion of importin-α proteins resulted in a substantial reduction in telomerase activity compared with that of the control siRNA cells (Fig. 5B).

Fig. 5.

Depletion of importin α reduces telomerase activity by preventing the assembly of telomerase in the nucleus. (A) MCF7 cells that had been transfected with one siRNA against the indicated importin-α proteins were subjected to immunoprecipitation with an anti-dyskerin antibody; lysates were then immunoblotted to detect endogenous hTERT, dyskerin, importin-α proteins and importin β1. IgG antibody was used as a negative control. (B) MCF7 cells that had been transfected with one siRNA against the indicated importin-α proteins were subjected to immunoprecipitation with an anti-dyskerin antibody. Immunoprecipitates of endogenous dyskerin were analyzed for telomerase activity by using the TRAP assay. IgG was used as a negative control. ITAS represents the internal telomerase assay standard. (C) MCF7 cells that had been transfected with one siRNA against the indicated importin-α proteins were treated with 10 μM MG132 for 4 h. Nuclear and cytosolic extracts were separately collected and subjected to immunoprecipitation with an anti-dyskerin antibody. Immunoprecipitates of endogenous dyskerin were analyzed for telomerase activity by using the TRAP assay.

Fig. 5.

Depletion of importin α reduces telomerase activity by preventing the assembly of telomerase in the nucleus. (A) MCF7 cells that had been transfected with one siRNA against the indicated importin-α proteins were subjected to immunoprecipitation with an anti-dyskerin antibody; lysates were then immunoblotted to detect endogenous hTERT, dyskerin, importin-α proteins and importin β1. IgG antibody was used as a negative control. (B) MCF7 cells that had been transfected with one siRNA against the indicated importin-α proteins were subjected to immunoprecipitation with an anti-dyskerin antibody. Immunoprecipitates of endogenous dyskerin were analyzed for telomerase activity by using the TRAP assay. IgG was used as a negative control. ITAS represents the internal telomerase assay standard. (C) MCF7 cells that had been transfected with one siRNA against the indicated importin-α proteins were treated with 10 μM MG132 for 4 h. Nuclear and cytosolic extracts were separately collected and subjected to immunoprecipitation with an anti-dyskerin antibody. Immunoprecipitates of endogenous dyskerin were analyzed for telomerase activity by using the TRAP assay.

The results presented here demonstrate that depletion of importin-α proteins inhibits the nuclear translocation of hTERT and that the resulting hTERT is degraded through the proteasome but accumulates in the cytoplasm when cells are treated with MG132. To determine whether cytoplasmic hTERT is assembled with the TERC–dyskerin RNP to form catalytically active telomerase, nuclear and cytosolic extracts were collected from importin-α-knockdown cells in the presence of MG132. As shown in Fig. 5C, telomerase activity was exclusively detected in the nuclear extracts, demonstrating that cytoplasmic hTERT is not assembled into active telomerase. When immunoprecipitates of dyskerin were analyzed for telomerase activity, similar results were observed (Fig. 5C). These findings suggest that depletion of importin α causes a reduction in telomerase activity by inhibiting the association of hTERT with dyskerin in the nucleus.

Phosphorylation of hTERT at S227 increases the binding affinity for importin α

We have previously shown that Akt-mediated phosphorylation at S227 is required for efficient nuclear translocation of hTERT (Chung et al., 2012). To explore the molecular mechanism by which phosphorylation of S227 is involved in the nuclear import of hTERT, we used isothermal titration calorimetry (ITC) to determine the binding affinity of importin α1 and importin α5 for an NLS peptide that was phosphorylated at S227 (phospho-S227) compared with that for wild-type NLS peptide. Phosphorylated NLS peptide (at S227, bold, 220GARRRGGpSASRSLPLPKRPRRGA242) was injected into an ITC cell containing 20 μM of the importin-α proteins, and the enthalpy of binding was carefully monitored. As a control experiment, wild-type NLS peptide was injected against the importin-α proteins under identical experiment conditions. The binding affinity of NLS peptide to importin α5 was increased approximately 2-fold when S227 was phosphorylated and decreased about 1.4-fold when S227 was mutated to alanine (Fig. 6A–C). Although the binding affinity of importin α1 is lower than that of importin α5, a similar binding trend was obtained for the binding affinity of NLS peptides to importin α1 (Fig. 6D–F). These findings clearly indicate that the phosphate moiety of S227 functions as a direct binding determinant for importin-α proteins.

Fig. 6.

Measurements of the binding affinity of hTERT NLS peptides for importin α5 and importin α1 by using isothermal titration calorimetry. (A) Injection of wild-type NLS peptide (220GARRRGGSASRSLPLPKRPRRGA242) into a calorimetric cell containing importin α5. (B) Injection of phospho-S227 NLS peptide (220GARRRGGpSASRSLPLPKRPRRGA242) into a calorimetric cell containing importin α5. (C) Injection of S227A NLS peptide (220GARRRGGAASRSLPLPKRPRRGA242) into a calorimetric cell containing importin α5. (D) Injection of wild-type NLS peptide into a calorimetric cell containing importin α1. (E) Injection of phospho-S227 NLS peptide into a calorimetric cell containing importin α1. (F) Injection of S227A NLS peptide into a calorimetric cell containing importin α1. In the sequences above, bold denotes the mutated or modified residue.

Fig. 6.

Measurements of the binding affinity of hTERT NLS peptides for importin α5 and importin α1 by using isothermal titration calorimetry. (A) Injection of wild-type NLS peptide (220GARRRGGSASRSLPLPKRPRRGA242) into a calorimetric cell containing importin α5. (B) Injection of phospho-S227 NLS peptide (220GARRRGGpSASRSLPLPKRPRRGA242) into a calorimetric cell containing importin α5. (C) Injection of S227A NLS peptide (220GARRRGGAASRSLPLPKRPRRGA242) into a calorimetric cell containing importin α5. (D) Injection of wild-type NLS peptide into a calorimetric cell containing importin α1. (E) Injection of phospho-S227 NLS peptide into a calorimetric cell containing importin α1. (F) Injection of S227A NLS peptide into a calorimetric cell containing importin α1. In the sequences above, bold denotes the mutated or modified residue.

Crystal structure of importin α5 bound to the hTERT NLS peptide

To provide further evidence for the molecular mechanism of hTERT nuclear import, we solved the complex structure of importin α5 bound to the hTERT NLS peptide at 2.5-Å resolution (Fig. 7A). The results showed a canonical bipartite binding mode of importin α5 and NLS (Fig. 7B). Major interactions of the side chain of the basic amino acid residues in both the minor (222RRR224) and major (236KRPRR240) sites are within acidic pockets of importin α5 and supplemented by additional hydrogen bonds (Fig. 7B). In the minor site, residue R222 interacts with residue Q399 through a salt bridge and interacts with residue W402 through a hydrophobic interaction (Fig. 7C). Residue R223 interacts with residues G284 and T325 through hydrogen bonds. Residue R224 interacts with residue Q357 through a salt bridge. In the linker region between the major and minor sites, residue R230 interacts with residue N322 through hydrogen bonds. The carbonyl oxygen of residue S231 binds to R318 through hydrogen bonds. Residue P233 interacts with residue W276 through van der Waals interaction, and its carbonyl group interacts with residue R241 through hydrogen bonding. In the major site, the ε-amino group of residue K236 interacts with the carboxyl groups of residues A155 and G157, and the side chains of residues T162 and N199 as a core of the hydrogen bond network (Fig. 7D). The amino and carboxyl groups of residue R237 interacts with residues N195 and W191, respectively, through hydrogen bonds. The side chain of residue R237 stacks on W234. The amino and carboxyl groups of residue R239 interact with residues N153 and W149 through hydrogen bonding, and its side chain interacts with residues Q188 and W191 through hydrogen and hydrophobic interactions, respectively. Residue R240 interacts with residues Q113 and S111 through electrostatic and hydrophobic interactions, respectively. The amino acid residues 225–228 of NLS are not visible in the electron density, presumably owing to flexibility. These interactions have been reported in previous structures of importin-α–NLS complexes (Conti and Kuriyan, 2000; Tarendeau et al., 2007; Fontes et al., 2003).

Fig. 7.

Structure of importin α5 bound to the hTERT NLS. (A) Ribbon representation of importin α5 (cyan) with the hTERT NLS (yellow). The green dashed circles highlight regions of interest that are shown in greater detail in B. (B) Electrostatic surface model showing the bipartite binding mode; minor (left) and major (right) NLS-binding sites of importin α5. (C,D) LIGPLOT diagram for binding of the NLS peptides to importin α5. LIGPLOT diagram shows the bipartite binding mode; minor (C) and major (D) NLS-binding sites of importin α5. The NLS bonds are shown in blue, and the bonds within the NLS environment are shown in brown. Hydrogen bonds are green dashed lines with indicated distances (in Å). The atoms are color-coded (carbon, black; oxygen, red; nitrogen, blue), and residues that are in hydrophobic contact with the NLS are represented by red semicircles with radiating spokes.

Fig. 7.

Structure of importin α5 bound to the hTERT NLS. (A) Ribbon representation of importin α5 (cyan) with the hTERT NLS (yellow). The green dashed circles highlight regions of interest that are shown in greater detail in B. (B) Electrostatic surface model showing the bipartite binding mode; minor (left) and major (right) NLS-binding sites of importin α5. (C,D) LIGPLOT diagram for binding of the NLS peptides to importin α5. LIGPLOT diagram shows the bipartite binding mode; minor (C) and major (D) NLS-binding sites of importin α5. The NLS bonds are shown in blue, and the bonds within the NLS environment are shown in brown. Hydrogen bonds are green dashed lines with indicated distances (in Å). The atoms are color-coded (carbon, black; oxygen, red; nitrogen, blue), and residues that are in hydrophobic contact with the NLS are represented by red semicircles with radiating spokes.

Phospho-S227 in NLS interacts with R395 of importin α5

Although phospho-S227 was not seen in the electron density map because of the disordered region, a possible structural model can be proposed by using molecular dynamics and Autodock tools (Morris et al., 2009). In this structural model, phospho-S227 appears to interact with residue R395 of importin α5 through hydrogen bonding (Fig. 8A–C). To confirm the importance of this interaction, we performed site-directed mutagenesis of importin α5, mutating residue R395 to lysine (R395K) or glutamate (R395E). In the ITC analysis, we found that the binding affinity of the R395K and R395E mutants for phosphorylated NLS peptide decreased approximately 4.3-fold (Kd, 202.9 nM) and approximately 14.1-fold mutant (Kd, 666.7 nM), respectively, compared with that of wild-type importin α5 (Kd, 47.2 nM) (Fig. 8D–F). As would be expected based on our model, the binding affinity for unphosphorylated NLS peptide remained unchanged regardless of mutations in R395 (supplementary material Fig. S3A–C). These findings indicate that the positive charge of R395 is crucial for binding to the phosphate moiety of S227, and that Akt-mediated phosphorylation of S227 increases the binding affinity of hTERT for importin α5 (supplementary material Fig. S3D).

In order for telomeric ends to be elongated by telomerase in vivo, newly synthesized hTERT has to be translocated into the nucleus and assembled with the RNA component to generate the active telomerase holoenzyme (Venteicher et al., 2008,, Venteicher et al., 2009; Lee et al., 2014). The hTERT NLS has been shown to be bipartite, comprising the first basic amino acid residues 222-RRR-224, a 12-residue spacer and the second residues 236-KRPRR-240 (Chung et al., 2012). In this work, we have studied the molecular mechanism by which hTERT enters the nucleus. The extended NLS can confer nuclear import on a heterologous protein through the classical importin-α–importin-β-dependent pathway. In humans, there are at least seven genes that encode importin α proteins (Kelley et al., 2010). These subtypes share the same general structure and exhibit a broad functional redundancy. The bipartite NLS of hTERT is able to interact with importin α1, importin α3 and importin α5. Crystallography analysis of importin α5 in complex with the hTERT NLS reveals the ability of the basic elements to interact with both the major and minor sites of importin α5. We also show that overexpression of a dominant-negative mutant of Ran (RanQ69L) inhibits nuclear import of hTERT, suggesting that nuclear import of hTERT is dependent on the cellular GTPase Ran and soluble import receptors – importin α and importin β1.

Fig. 8.

A structural model for the complex between phospho-S227 NLS and importin α5. (A) Ribbon diagram showing importin α5 (blue) bound to the phospho-S227 NLS (yellow). (B) Surface representation showing the bipartite binding mode (blue dashed circles) at both ends of phospho-S227 NLS peptide. Additional interactions with the phospho-S227 (pS227) NLS peptide occur at the central basic region of importin α5. (C) The phosphate moiety of the phospho-S227 NLS peptide is involved in hydrogen bonds with residue R395 of importin α5, which provide an additional stabilizing force. (D–F) Measurements for the binding affinity of the phospho-S227 NLS peptide to various importin-α5 proteins by using isothermal titration calorimetry. Injection of the phospho-S227 NLS peptide into a calorimetric cell containing wild-type importin α5 (D), importin α5 R395K (E) and importin α5 R395E (F).

Fig. 8.

A structural model for the complex between phospho-S227 NLS and importin α5. (A) Ribbon diagram showing importin α5 (blue) bound to the phospho-S227 NLS (yellow). (B) Surface representation showing the bipartite binding mode (blue dashed circles) at both ends of phospho-S227 NLS peptide. Additional interactions with the phospho-S227 (pS227) NLS peptide occur at the central basic region of importin α5. (C) The phosphate moiety of the phospho-S227 NLS peptide is involved in hydrogen bonds with residue R395 of importin α5, which provide an additional stabilizing force. (D–F) Measurements for the binding affinity of the phospho-S227 NLS peptide to various importin-α5 proteins by using isothermal titration calorimetry. Injection of the phospho-S227 NLS peptide into a calorimetric cell containing wild-type importin α5 (D), importin α5 R395K (E) and importin α5 R395E (F).

Although the classic NLS-receptor complex of importin α can mediate nuclear import of hTERT, importin α cannot be the sole import receptor for hTERT. Importin 7 has been recently shown to interact with hTERT and to mediate nuclear transport of hTERT (Frohnert et al., 2014). To confirm these results in our system, we examined whether importin 7 binds to the hTERT NLS, and whether its depletion affects the levels and subcellular localization of hTERT and telomerase activity. We found that endogenous importin 7 can interact physically with the hTERT NLS (supplementary material Fig. S4C,D). When importin 7 was depleted, telomerase activity was slightly reduced compared with that of the control siRNA cells (supplementary material Fig. S4E). Depletion of importin 7 led to a clear reduction in the nuclear accumulation of hTERT (supplementary material Fig. S4F). We also found that the total intensity hTERT fluorescence was reduced upon importin 7 depletion and that this reduction in hTERT level was rescued by treatment with MG132. Taken together, these results suggest that importin 7 mediates nuclear localization of hTERT. Thus, it is likely that hTERT can be imported into the nucleus through multiple nuclear import pathways. Although redundancy in nuclear import pathways is a usual phenomenon and is observed for many abundant cargo substrates, such as ribosomal proteins (Jakel and Gorlich, 1998) or histones (Muhlhausser et al., 2001), our results offer no clear indication of the comparative transport efficiencies of the different pathways in vivo. Therefore, it will be of interest to investigate how these pathways are specifically regulated and how they are activated under different conditions.

Akt kinase has been shown to facilitate the nuclear import of hTERT through phosphorylation and to enhance telomerase activity (Chung et al., 2012). Two putative Akt phosphorylation sites, S227 and S824, have been identified previously in hTERT (Kang et al., 1999). Although both serine residues have been shown to be phosphorylated by Akt kinase, phosphorylation of S227 appears to be important for nuclear import of hTERT (Chung et al., 2012). Because residue S227 is located between two clusters of basic amino acids in the bipartite NLS sequence, Akt-mediated phosphorylation of S227 might contribute to the interaction of the hTERT NLS with importin-α proteins. Structural characterization of importin α5 bound to the hTERT NLS showed that the phosphate moiety of S227 interacts with R395 of importin α5 through hydrogen bonding. When arginine was replaced by lysine or glutamate at position 395, the binding affinity for phosphorylated NLS peptide was substantially decreased. These findings suggest that the interaction between the phospho-S227 and R395 residues of importin α5 contributes to the efficient nuclear localization of hTERT by increasing the binding affinity of hTERT for importin α5. In addition, biochemical studies using chemically synthesized peptides and ITC showed that phosphorylation of S227 increases the binding affinity for importin α5, compared with that of the unphosphorylated NLS. Thus, the phosphate moiety within the NLS functions as a direct binding determinant for importin α5.

There are several well-documented examples of importin cargos that are regulated by phosphorylation (Nardozzi et al., 2010). One example is the Epstein-Barr virus (EBV) nuclear antigen 1 (EBNA-1) protein (Kitamura et al., 2006). Phosphorylation of residue S385 within the EBVA-1 NLS upregulates the nuclear transport efficiency by increasing its binding affinity for importin α5. In the case of the hepatitis B virus (HBV) core particle, protein kinase C (PKC)-mediated phosphorylation at residue S172 causes a conformational change that exposes the C-terminal NLS on the surface. Upon phosphorylation, the HBV core particle is recognized by importin α1 and is then imported into the nucleus (Kann et al., 1999). Phosphorylation by the protein kinase casein kinase 2 and double-stranded DNA-dependent protein kinase upstream of the classic SV40 NLS promotes the nuclear import of large T-antigen by enhancing its recognition by importin α1 (Hübner et al., 1997; Xiao et al., 1997). Therefore, phosphorylation-mediated regulation of nuclear transport could provide a general mechanism for many more nuclear proteins. Although nuclear accumulation of a cargo is dependent on the NLS, the rate of nuclear import is regulated through phosphorylation within or near the NLS of the cargo (Nardozzi et al., 2010). Thus, it is likely that the nuclear import efficiency of a cargo is determined by the addition and removal of a phosphate moiety, kinases and phosphatases – the activities of which are tightly regulated by diverse biological pathways in different cellular environments. On the basis of a previous report (Chung et al., 2012) and our data presented here, we propose that two functional determinants serve as molecular switches controlling nuclear localization of hTERT – the bipartite NLS and Akt-mediated phosphorylation of S227. The interplay between these two signals is likely to ensure the proper cellular function of telomerase activity by determining the nuclear import efficiency of hTERT.

When importin α is depleted or RanQ69L is overexpressed, hTERT fails to be translocated into the nucleus, and the resulting cytoplasmic hTERT is degraded through ubiquitin-dependent proteolysis. Ubiquitin-dependent degradation of hTERT was efficiently inhibited by treatment with MG132, suggesting that proteasomal degradation of hTERT occurs in a step separate from that of inhibition of its nuclear import. Several E3 ubiquitin ligases, including CHIP (Lee et al., 2010), MKRN1 (Kim et al., 2005) and Hdm2 (Oh et al., 2010), have been identified as targeting hTERT for ubiquitylation. Although all E3 ligases are each capable of promoting ubiquitylation of hTERT, the functional similarity raises many important questions about the physiological significance of multiple pathways that exert negative control over hTERT. The Hsp90 and p23 molecular chaperones have been shown to associate with hTERT to maintain its conformation, enabling nuclear translocation (Holt et al., 1999; Forsythe et al., 2001). By contrast, the binding of CHIP and Hsp70 to hTERT causes a dissociation of p23 from hTERT, resulting in a cytoplasmic accumulation of hTERT that cannot be translocated into the nucleus (Lee et al., 2010). Cytoplasmic hTERT is rapidly ubiquitylated by CHIP and other ubiquitin ligases as well. Based on these findings, we propose a model for chaperone-mediated nuclear import of hTERT – the hTERT–CHIP–Hsp70 complex favors degradation of hTERT in the cytoplasm, whereas the hTERT–Hsp90–p23 complex is preferentially bound by nuclear import receptors and then translocated into the nucleus. Thus, p23 and the CHIP–Hsp70 complex might compete for binding to hTERT in the cytoplasm, and the outcome of this competition is likely to determine the fate of hTERT – nuclear translocation or degradation.

In this study, we show that the nuclear localization of hTERT requires the classic nuclear import machinery, comprising importin α, importin β and GTPase Ran. Crystallography analysis demonstrated that the basic amino acid clusters within the bipartite NLS interact with both the major and minor sites of importin α in a manner that is characteristic of that of other bipartite NLSs (Fontes et al., 2003). Importantly, we show that Akt-mediated phosphorylation of hTERT is an important step that regulates its nuclear availability, which directly affects telomerase activity. Although many important questions about how hTERT phosphorylation is regulated under diverse physiological conditions remain, Akt-mediated phosphorylation of hTERT represents a new pathway for determining proper cellular telomerase activity by regulating its nuclear import efficiency.

Cell culture

The human breast cancer cell line MCF7 was cultured in Dulbecco's modified Eagle's medium (DMEM), the human lung carcinoma cell line H1299 was cultured in RPMI 1640 medium, and the human osteosarcoma cell line U2OS was cultured in McCoy's 5A medium supplemented with 10% fetal bovine serum, 100 units/ml penicillin and 100 μg/ml streptomycin under 5% CO2 at 37°C. The expression vectors were transiently transfected using Lipofectamine-PLUS reagent according to the manufacturer's protocol (Invitrogen).

Plasmid constructs

To construct the GFP–NLS–β-gal expression vectors, the wild-type and mutant NLS fragments were generated by using PCR and subcloned into the XhoI and HindIII sites of the pEGFP-C3-β-gal vector (BD Biosciences). The expression vectors for HA–importin-α proteins, HA–importin-β1, GST–importin-α proteins and GST–importin-β1 were constructed by inserting the whole coding regions into the MluI and NotI sites of the pRK5-HA vector (Stratagene) and the BamHI and NotI sites of the pGEX-5X-1 vector (Amersham Biosciences), respectively. The expression vectors for V5–Ran and V5–RanQ69L were constructed by inserting cDNAs encoding wild-type and mutant Ran into the EcoRI and XhoI sites of pcDNA3.1/V5-His vector (Invitrogen). N-terminally truncated importin α5 containing amino acid residues 66–512 were mutated to generate the R395K and R395E mutants, and subcloned into the EcoRI and XhoI sites of the pET28a-TEV vector.

GST pulldown, immunoprecipitation and immunoblotting

GST pulldown, immunoprecipitation and immunoblot analyses were performed as described previously (Lee et al., 2004). Immunoprecipitation and immunoblotting were performed using anti-Flag (cat. no. F3165, Sigma-Aldrich), anti-HA (cat. no. sc-805, Santa Cruz Biotechnology), anti-GST (cat. no. sc-138, Santa Cruz Biotechnology), anti-Myc (cat. no. sc-40, Santa Cruz Biotechnology), anti-importin-α1 (cat. no. NB 100-79807, Novus Biologicals), anti-importin-α3 (cat. no. NBP1-31260, Novus Biologicals), anti-importin-α5 (cat. no. sc-101292, Santa Cruz Biotechnology), anti-importin-β1 (cat. no. sc-11367, Santa Cruz Biotechnology), anti-importin-7 (cat. no. GTX 106408, GeneTex), anti-hTERT (cat. no. 600-401-252, Rockland), anti-dyskerin (cat. no. sc-48794, Santa Cruz Biotechnology), anti-actin (cat. no. A2066, Sigma-Aldrich), anti-tubulin (cat. no. sc-8035, Santa Cruz Biotechnology) and anti-lamin (cat. no. sc-6215, Santa Cruz Biotechnology) antibodies as specified. All the immunoblots are representatives of at least three experiments that demonstrated the similar results.

Cytoplasmic and nuclear fractionation

After transfection, cells were harvested and resuspended in hypotonic buffer (10 mM HEPES-KOH, pH 7.9, 1.5 mM MgCl2, 10 mM KCl, 0.5 mM DTT, 0.2 mM phenyl methylsulphonyl fluoride) for 10 min on ice, followed by the addition of 0.5% NP-40 for 10 min. After centrifugation at 14,000 g for 5 min at 4°C, the supernatant was used as the cytoplasmic fraction. The nuclear pellets were resuspended in nuclear extraction buffer (20 mM HEPES-KOH, pH 7.9, 25% Glycerol, 420 mM NaCl, 1.5 mM MgCl2, 0.2 mM EDTA, 0.5 mM DTT, 0.2 mM phenyl methylsulphonyl fluoride) for 20 min on ice and centrifuged at 14,000 g for 20 min at 4°C. The supernatants were saved as the nuclear fraction.

Immunofluorescence microscopy

Cells that had been grown on glass coverslips were fixed with 4% paraformaldehyde and permeabilized with 0.5% Triton X-100 in PBS. Cells were then blocked in PBS containing 0.5% bovine serum albumin and 0.2% gelatin, and incubated with mouse anti-Flag (cat. no. F3165, Sigma-Aldrich), rabbit anti-HA (cat. no. sc-805, Santa Cruz Biotechnology), rabbit anti-hTERT (cat. no. 600-401-252, Rockland) or rabbit anti-dyskerin (cat. no. sc-48794, Santa Cruz Biotechnology). After thorough washing with PBS, cells were incubated with Alexa-Fluor-594-conjugated goat anti-mouse immunoglobulin (cat no. A11005, Molecular Probes) and Alexa-Fluor-488-conjugated goat anti-rabbit immunoglobulin (cat no. A11008 , Molecular Probes). DNA was stained with 4,6-diamino-2-phenylindole (DAPI) (Vectashield; Vector Laboratories). Immunofluorescence images were captured using a confocal laser-scanning microscope (Carl Zeiss).

RNA interference

The siRNA duplexes were transfected into MCF7 cells using RNAiMax transfection reagent (Invitrogen). The scrambled sequence (5′-AATCGCATAGCGTATGCCGTT-3′) was used as a control and did not correspond to any known gene in the data bases. Importin α1-1 siRNA, 5′-TAGCATGTGGCTACTTACGTA-3′; importin α1-2 siRNA, 5′-TAGGAGCTTCTGAATTGCCAA-3′; importin α3-1 siRNA, 5′-ATGGGATCTTGCTGCTGCATTA-3′; importin α3-2 siRNA, 5′-TAGGCTGCCCATATAAGTCAA-3′; importin α5-1 siRNA, 5′-ATGGGATCTGCTGCTGCATTA-3′; importin α5-2 siRNA, 5′-AAAGATCTTTGCATGAGGTAA-3′; importin β1 siRNA, 5′-GTGCAACTCTTCAGAATGT-3′; importin 7-1 siRNA, 5′-GATGGAGCCCTGCATATGA-3′; importin 7-2 siRNA, 5′-GGCAGGTGTTATCTATCTG-3′.

In vivo ubiquitylation assay

MCF7 cells were co-transfected with HA–ubiquitin and Flag–hTERT, and treated with 10 μM MG132 to inhibit proteasome function. Lysates were subjected to immunoprecipitation with an anti-Flag antibody, followed by immunoblotting with an anti–HA antibody to show ubiquitin-modified hTERT.

Telomerase assay

The telomeric repeat amplification protocol (TRAP) was used as previously described (Lee et al., 2004).

Isothermal titration calorimetry

ITC experiments were performed at 25°C in a VP-ITC calorimeter (Microcal Software) (Pierce et al., 1999). The hTERT NLS peptides (wild-type, phospho-S227 and S227A) were dissolved in ITC buffer (20 mM Tris-pH 7.5, 150 mM sodium chloride) and injected in 10-μl increments into a calorimetric cell containing 1.8 ml of the importin-α proteins. The interval between injections was 220 s. In separate experiments, the NLS peptides were titrated into the importin-α proteins, as described above. Titration data were analyzed using the Origin 7.0 data analysis software (Microcal Software). Injections were integrated following manual adjustment of the baselines. Heats of dilution were determined from control experiments with the ITC buffer and subtracted prior to curve fitting using a single set of binding site models.

Crystallization and structure determination

For crystallization, purified importin α5 (residues 66–512) was concentrated to 25 mg/ml using a Centricon-30 filter (Amicon) and mixed with phospho-S227 NLS peptide in a 1:5 molar ratio. Crystallization conditions were screened by using micro-bath plates. The best crystals of the importin-α5–NLS complex were obtained with 17% poly(ethylene glycol) 10 K as the precipitant in 100 mM HEPES (pH 7.5) in hanging drop plates. Diffraction data were collected from single crystals with 20% ethylene glycol and flash frozen at 100 K in a nitrogen stream. Data to 2.5-Å resolution were measured on beam line BL5A at the Photon Factory synchrotron using X-rays of wavelength 1.0 Å, and were auto-indexed and processed with the HKL suite. The complex crystals of importin–peptide exhibit monoclinic (space group C2). The complex was located in the asymmetric unit by molecular replacement using PDB 2JDQ (importin α5 with PB2 subunit) as the search model (Tarendeau et al., 2007). Coot was used for visualization of electron density maps and manual rebuilding of atomic models (Emsley and Cowtan, 2004). The initial model was refined using REFMAC5 (Vagin et al., 2004) at 2.5-Å resolution.

We are grateful to Joonyoung Her, Yu Young Jeong, and Juyeong Hong for technical assistance and helpful comments on the manuscript and to the staff scientists at the BL-5A beamline of the Photon Factory.

Author contributions

All authors contributed to the conception and design of experiments, analysis of data and manuscript preparation. S.A.J., K.K, J.H.L., J.S.C., and P.K. performed experiments. S.A.J., H.-S.C. and I.K.C. wrote the manuscript.

Funding

This work was supported in part by a grant from the Bio and Medical Technology Development Program of the National Research Foundation funded by the Korean Ministry of Science, ICT, and Future Planning [grant NRF-2013M3A9B6076413 to I.K.C.]; by Yonsei University Internal Grant [grant 2014-22-0096 to I.K.C.]; and by a grant from National Research Foundation of Korea [grant NRF-2012R1A2A2A01012830 to H.S.C].

Autexier
,
C.
and
Lue
,
N. F.
(
2006
).
The structure and function of telomerase reverse transcriptase
.
Annu. Rev. Biochem.
75
,
493
-
517
.
Bianchi
,
A.
and
Shore
,
D.
(
2008
).
How telomerase reaches its end: mechanism of telomerase regulation by the telomeric complex
.
Mol. Cell
31
,
153
-
165
.
Blackburn
,
E. H.
(
2001
).
Switching and signaling at the telomere
.
Cell
106
,
661
-
673
.
Chook
,
Y. M.
and
Süel
,
K. E.
(
2011
).
Nuclear import by karyopherin-betas: recognition and inhibition
.
Biochim. Biophys. Acta
1813
,
1593
-
1606
.
Chung
,
J.
,
Khadka
,
P.
and
Chung
,
I. K.
(
2012
).
Nuclear import of hTERT requires a bipartite nuclear localization signal and Akt-mediated phosphorylation
.
J. Cell Sci.
125
,
2684
-
2697
.
Cingolani
,
G.
,
Petosa
,
C.
,
Weis
,
K.
and
Muller
,
C. W.
(
1999
).
Structure of importin-beta bound to the IBB domain of importin-alpha
.
Nature
399
,
221
-
229
.
Cingolani
,
G.
,
Bednenko
,
J.
,
Gillespie
,
M. T.
and
Gerace
,
L.
(
2002
).
Molecular basis for the recognition of a nonclassical nuclear localization signal by importin beta
.
Mol. Cell
10
,
1345
-
1353
.
Cong
,
Y.-S.
,
Wright
,
W. E.
and
Shay
,
J. W.
(
2002
).
Human telomerase and its regulation
.
Microbiol. Mol. Biol. Rev.
66
,
407
-
425
.
Conti
,
E.
and
Kuriyan
,
J.
(
2000
).
Crystallographic analysis of the specific yet versatile recognition of distinct nuclear localization signals by karyopherin alpha
.
Structure
8
,
329
-
338
.
Conti
,
E.
,
Uy
,
M.
,
Leighton
,
L.
,
Blobel
,
G.
and
Kuriyan
,
J.
(
1998
).
Crystallographic analysis of the recognition of a nuclear localization signal by the nuclear import factor karyopherin alpha
.
Cell
94
,
193
-
204
.
de Lange
,
T.
(
2005
).
Shelterin: the protein complex that shapes and safeguards human telomeres
.
Genes Dev.
19
,
2100
-
2110
.
Dickmanns
,
A.
,
Bischoff
,
F. R.
,
Marshallsay
,
C.
,
Lührmann
,
R.
,
Ponstingl
,
H.
and
Fanning
,
E.
(
1996
).
The thermolability of nuclear protein import in tsBN2 cells is suppressed by microinjected Ran-GTP or Ran-GDP, but not by RanQ69L or RanT24N
.
J. Cell Sci.
109
,
1449
-
1457
.
Emsley
,
P.
and
Cowtan
,
K.
(
2004
).
Coot: model-building tools for molecular graphics
.
Acta Crystallogr. D. Biol. Crystallogr.
60
,
2126
-
2132
.
Fontes
,
M. R. M.
,
Teh
,
T.
and
Kobe
,
B.
(
2000
).
Structural basis of recognition of monopartite and bipartite nuclear localization sequences by mammalian importin-alpha
.
J. Mol. Biol.
297
,
1183
-
1194
.
Fontes
,
M. R. M.
,
Teh
,
T.
,
Jans
,
D.
,
Brinkworth
,
R. I.
and
Kobe
,
B.
(
2003
).
Structural basis for the specificity of bipartite nuclear localization sequence binding by importin-alpha
.
J. Biol. Chem.
278
,
27981
-
27987
.
Forsythe
,
H. L.
,
Jarvis
,
J. L.
,
Turner
,
J. W.
,
Elmore
,
L. W.
and
Holt
,
S. E.
(
2001
).
Stable association of hsp90 and p23, but Not hsp70, with active human telomerase
.
J. Biol. Chem.
276
,
15571
-
15574
.
Forwood
,
J. K.
,
Lonhienne
,
T. G.
,
Marfori
,
M.
,
Robin
,
G.
,
Meng
,
W.
,
Guncar
,
G.
,
Liu
,
S. M.
,
Stewart
,
M.
,
Carroll
,
B. J.
and
Kobe
,
B.
(
2008
).
Kap95p binding induces the switch loops of RanGDP to adopt the GTP-bound conformation: implications for nuclear import complex assembly dynamics
.
J. Mol. Biol.
383
,
772
-
782
.
Frohnert
,
C.
,
Hutten
,
S.
,
Wälde
,
S.
,
Nath
,
A.
and
Kehlenbach
,
R. H.
(
2014
).
Importin 7 and Nup358 promote nuclear import of the protein component of human telomerase
.
PLoS ONE
9
,
e88887
.
Griffith
,
J. D.
,
Comeau
,
L.
,
Rosenfield
,
S.
,
Stansel
,
R. M.
,
Bianchi
,
A.
,
Moss
,
H.
and
de Lange
,
T.
(
1999
).
Mammalian telomeres end in a large duplex loop
.
Cell
97
,
503
-
514
.
Harreman
,
M. T.
,
Hodel
,
M. R.
,
Fanara
,
P.
,
Hodel
,
A. E.
and
Corbett
,
A. H.
(
2003
).
The auto-inhibitory function of importin alpha is essential in vivo
.
J. Biol. Chem.
278
,
5854
-
5863
.
Holt
,
S. E.
,
Aisner
,
D. L.
,
Baur
,
J.
,
Tesmer
,
V. M.
,
Dy
,
M.
,
Ouellette
,
M.
,
Trager
,
J. B.
,
Morin
,
G. B.
,
Toft
,
D. O.
,
Shay
,
J. W.
, et al. 
(
1999
).
Functional requirement of p23 and Hsp90 in telomerase complexes
.
Genes Dev.
13
,
817
-
826
.
Hübner
,
S.
,
Xiao
,
C.-Y.
and
Jans
,
D. A.
(
1997
).
The protein kinase CK2 site (Ser111/112) enhances recognition of the simian virus 40 large T-antigen nuclear localization sequence by importin
.
J. Biol. Chem.
272
,
17191
-
17195
.
Jakel
,
S.
and
Gorlich
,
D.
(
1998
).
Importin beta, transportin, RanBP5 and RanBP7 mediate nuclear import of ribosomal proteins in mammalian cells
.
EMBO J.
17
,
4491
-
4502
.
Kang
,
S. S.
,
Kwon
,
T.
,
Kwon
,
D. Y.
and
Do
,
S. I.
(
1999
).
Akt protein kinase enhances human telomerase activity through phosphorylation of telomerase reverse transcriptase subunit
.
J. Biol. Chem.
274
,
13085
-
13090
.
Kann
,
M.
,
Sodeik
,
B.
,
Vlachou
,
A.
,
Gerlich
,
W. H.
and
Helenius
,
A.
(
1999
).
Phosphorylation-dependent binding of hepatitis B virus core particles to the nuclear pore complex
.
J. Cell Biol.
145
,
45
-
55
.
Kelley
,
J. B.
,
Talley
,
A. M.
,
Spencer
,
A.
,
Gioeli
,
D.
and
Paschal
,
B. M.
(
2010
).
Karyopherin α7 (KPNA7), a divergent member of the importin α family of nuclear import receptors
.
BMC Cell Biol.
11
,
63
.
Kim
,
N. W.
and
Wu
,
F.
(
1997
).
Advances in quantification and characterization of telomerase activity by the telomeric repeat amplification protocol (TRAP)
.
Nucleic Acids Res.
25
,
2595
-
2597
.
Kim
,
N. W.
,
Piatyszek
,
M. A.
,
Prowse
,
K. R.
,
Harley
,
C. B.
,
West
,
M. D.
,
Ho
,
P. L.
,
Coviello
,
G. M.
,
Wright
,
W. E.
,
Weinrich
,
S. L.
and
Shay
,
J. W.
(
1994
).
Specific association of human telomerase activity with immortal cells and cancer
.
Science
266
,
2011
-
2015
.
Kim
,
J. H.
,
Park
,
S.-M.
,
Kang
,
M. R.
,
Oh
,
S.-Y.
,
Lee
,
T. H.
,
Muller
,
M. T.
and
Chung
,
I. K.
(
2005
).
Ubiquitin ligase MKRN1 modulates telomere length homeostasis through a proteolysis of hTERT
.
Genes Dev.
19
,
776
-
781
.
Kitamura
,
R.
,
Sekimoto
,
T.
,
Ito
,
S.
,
Harada
,
S.
,
Yamagata
,
H.
,
Masai
,
H.
,
Yoneda
,
Y.
and
Yanagi
,
K.
(
2006
).
Nuclear import of Epstein-Barr virus nuclear antigen 1 mediated by NPI-1 (Importin alpha5) is up- and down-regulated by phosphorylation of the nuclear localization signal for which Lys379 and Arg380 are essential
.
J. Virol.
80
,
1979
-
1991
.
Kobe
,
B.
(
1999
).
Autoinhibition by an internal nuclear localization signal revealed by the crystal structure of mammalian importin alpha
.
Nat. Struct. Biol.
6
,
388
-
397
.
Lange
,
A.
,
Mills
,
R. E.
,
Lange
,
C. J.
,
Stewart
,
M.
,
Devine
,
S. E.
and
Corbett
,
A. H.
(
2007
).
Classical nuclear localization signals: definition, function, and interaction with importin alpha
.
J. Biol. Chem.
282
,
5101
-
5105
.
Lee
,
G. E.
,
Yu
,
E. Y.
,
Cho
,
C. H.
,
Lee
,
J.
,
Muller
,
M. T.
and
Chung
,
I. K.
(
2004
).
DNA-protein kinase catalytic subunit-interacting protein KIP binds telomerase by interacting with human telomerase reverse transcriptase
.
J. Biol. Chem.
279
,
34750
-
34755
.
Lee
,
J. H.
,
Khadka
,
P.
,
Baek
,
S. H.
and
Chung
,
I. K.
(
2010
).
CHIP promotes human telomerase reverse transcriptase degradation and negatively regulates telomerase activity
.
J. Biol. Chem.
285
,
42033
-
42045
.
Lee
,
J. H.
,
Lee
,
Y. S.
,
Jeong
,
S. A.
,
Khadka
,
P.
,
Roth
,
J.
and
Chung
,
I. K.
(
2014
).
Catalytically active telomerase holoenzyme is assembled in the dense fibrillar component of the nucleolus during S phase
.
Histochem. Cell Biol.
141
,
137
-
152
.
Lingner
,
J.
,
Cooper
,
J. P.
and
Cech
,
T. R.
(
1995
).
Telomerase and DNA end replication: no longer a lagging strand problem?
Science
269
,
1533
-
1534
.
Lonhienne
,
T. G.
,
Forwood
,
J. K.
,
Marfori
,
M.
,
Robin
,
G.
,
Kobe
,
B.
and
Carroll
,
B. J.
(
2009
).
Importin-beta is a GDP-to-GTP exchange factor of Ran: implications for the mechanism of nuclear import
.
J. Biol. Chem.
284
,
22549
-
22558
.
Morris
,
G. M.
,
Huey
,
R.
,
Lindstrom
,
W.
,
Sanner
,
M. F.
,
Belew
,
R. K.
,
Goodsell
,
D. S.
and
Olson
,
A. J.
(
2009
).
AutoDock4 and AutoDockTools4: automated docking with selective receptor flexibility
.
J. Comput. Chem.
30
,
2785
-
2791
.
Muhlhausser
,
P.
,
Muller
,
E.-C.
,
Otto
,
A.
and
Kutay
,
U.
(
2001
).
Multiple pathways contribute to nuclear import of core histones
.
EMBO Rep.
2
,
690
-
696
.
Nakamura
,
T. M.
and
Cech
,
T. R.
(
1998
).
Reversing time: origin of telomerase
.
Cell
92
,
587
-
590
.
Nardozzi
,
J. D.
,
Lott
,
K.
and
Cingolani
,
G.
(
2010
).
Phosphorylation meets nuclear import: a review
.
Cell Commun. Signal.
8
,
32
.
Oh
,
W.
,
Lee
,
E.-W.
,
Lee
,
D.
,
Yang
,
M.-R.
,
Ko
,
A.
,
Yoon
,
C.-H.
,
Lee
,
H.-W.
,
Bae
,
Y.-S.
,
Choi
,
C. Y.
and
Song
,
J.
(
2010
).
Hdm2 negatively regulates telomerase activity by functioning as an E3 ligase of hTERT
.
Oncogene
29
,
4101
-
4112
.
Palacios
,
I.
,
Weis
,
K.
,
Klebe
,
C.
,
Mattaj
,
I. W.
and
Dingwall
,
C.
(
1996
).
RAN/TC4 mutants identify a common requirement for snRNP and protein import into the nucleus
.
J. Cell Biol.
133
,
485
-
494
.
Palm
,
W.
and
de Lange
,
T.
(
2008
).
How shelterin protects mammalian telomeres
.
Annu. Rev. Genet.
42
,
301
-
334
.
Peifer
,
M.
,
Berg
,
S.
and
Reynolds
,
A. B.
(
1994
).
A repeating amino acid motif shared by proteins with diverse cellular roles
.
Cell
76
,
789
-
791
.
Pierce
,
M. M.
,
Raman
,
C. S.
and
Nall
,
B. T.
(
1999
).
Isothermal titration calorimetry of protein-protein interactions
.
Methods
19
,
213
-
221
.
Poon
,
I. K. H.
and
Jans
,
D. A.
(
2005
).
Regulation of nuclear transport: central role in development and transformation?
Traffic
6
,
173
-
186
.
Sfeir
,
A.
and
de Lange
,
T.
(
2012
).
Removal of shelterin reveals the telomere end-protection problem
.
Science
336
,
593
-
597
.
Smogorzewska
,
A.
and
de Lange
,
T.
(
2004
).
Regulation of telomerase by telomeric proteins
.
Annu. Rev. Biochem.
73
,
177
-
208
.
Stewart
,
M.
(
2007
).
Molecular mechanism of the nuclear protein import cycle
.
Nat. Rev. Mol. Cell. Biol.
8
,
195
-
208
.
Tarendeau
,
F.
,
Boudet
,
J.
,
Guilligay
,
D.
,
Mas
,
P. J.
,
Bougault
,
C. M.
,
Boulo
,
S.
,
Baudin
,
F.
,
Ruigrok
,
R. W. H.
,
Daigle
,
N.
,
Ellenberg
,
J.
, et al. 
(
2007
).
Structure and nuclear import function of the C-terminal domain of influenza virus polymerase PB2 subunit
.
Nat. Struct. Mol. Biol.
14
,
229
-
233
.
Vagin
,
A. A.
,
Steiner
,
R. A.
,
Lebedev
,
A. A.
,
Potterton
,
L.
,
McNicholas
,
S.
,
Long
,
F.
and
Murshudov
,
G. N.
(
2004
).
REFMAC 5 dictionary: organization of prior chemical knowledge and guidelines for its use
.
Acta Crystallogr. D. Biol. Crystallogr.
60
,
2184
-
2195
.
Venteicher
,
A. S.
,
Meng
,
Z.
,
Mason
,
P. J.
,
Veenstra
,
T. D.
and
Artandi
,
S. E.
(
2008
).
Identification of ATPases pontin and reptin as telomerase components essential for holoenzyme assembly
.
Cell
132
,
945
-
957
.
Venteicher
,
A. S.
,
Abreu
,
E. B.
,
Meng
,
Z.
,
McCann
,
K. E.
,
Terns
,
R. M.
,
Veenstra
,
T. D.
,
Terns
,
M. P.
and
Artandi
,
S. E.
(
2009
).
A human telomerase holoenzyme protein required for Cajal body localization and telomere synthesis
.
Science
323
,
644
-
648
.
Wang
,
J.
,
Xie
,
L. Y.
,
Allan
,
S.
,
Beach
,
D.
and
Hannon
,
G. J.
(
1998
).
Myc activates telomerase
.
Genes Dev.
12
,
1769
-
1774
.
Xiao
,
C.-Y.
,
Hübner
,
S.
and
Jans
,
D. A.
(
1997
).
SV40 large tumor antigen nuclear import is regulated by the double-stranded DNA-dependent protein kinase site (serine 120) flanking the nuclear localization sequence
.
J. Biol. Chem.
272
,
22191
-
22198
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information