Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

DNA polymerase-α regulates the activation of type I interferons through cytosolic RNA:DNA synthesis

Abstract

Aberrant nucleic acids generated during viral replication are the main trigger for antiviral immunity, and mutations that disrupt nucleic acid metabolism can lead to autoinflammatory disorders. Here we investigated the etiology of X-linked reticulate pigmentary disorder (XLPDR), a primary immunodeficiency with autoinflammatory features. We discovered that XLPDR is caused by an intronic mutation that disrupts the expression of POLA1, which encodes the catalytic subunit of DNA polymerase-α. Unexpectedly, POLA1 deficiency resulted in increased production of type I interferons. This enzyme is necessary for the synthesis of RNA:DNA primers during DNA replication and, strikingly, we found that POLA1 is also required for the synthesis of cytosolic RNA:DNA, which directly modulates interferon activation. Together this work identifies POLA1 as a critical regulator of the type I interferon response.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: WGS identifies a recurrent intronic mutation as the cause of XLPDR.
Figure 2: XLPDR is due to an intronic mutation that disrupts POLA1 expression.
Figure 3: POLA1 deficiency results in type-I interferon activation.
Figure 4: POLA1 deficiency leads to excessive activation of the IRF and NF-κB pathways.
Figure 5: POLA1 deficiency is associated with reduced levels of cytosolic RNA:DNA.
Figure 6: Cytosolic RNA:DNA attenuates the activation of IRFs.
Figure 7: The generation of cytosolic RNA:DNA via POLA1 modulates nucleic acid–sensor pathways.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

  1. Sancho-Shimizu, V. et al. Herpes simplex encephalitis in children with autosomal recessive and dominant TRIF deficiency. J. Clin. Invest. 121, 4889–4902 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Hambleton, S. et al. STAT2 deficiency and susceptibility to viral illness in humans. Proc. Natl. Acad. Sci. USA 110, 3053–3058 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Casrouge, A. et al. Herpes simplex virus encephalitis in human UNC-93B deficiency. Science 314, 308–312 (2006).

    Article  CAS  PubMed  Google Scholar 

  4. Yang, Y.G., Lindahl, T. & Barnes, D.E. Trex1 exonuclease degrades ssDNA to prevent chronic checkpoint activation and autoimmune disease. Cell 131, 873–886 (2007).

    CAS  PubMed  Google Scholar 

  5. Mannion, N.M. et al. The RNA-editing enzyme ADAR1 controls innate immune responses to RNA. Cell Rep. 9, 1482–1494 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Crow, Y.J. & Manel, N. Aicardi-Goutières syndrome and the type I interferonopathies. Nat. Rev. Immunol. 15, 429–440 (2015).

    CAS  PubMed  Google Scholar 

  7. Crow, Y.J. et al. Characterization of human disease phenotypes associated with mutations in TREX1, RNASEH2A, RNASEH2B, RNASEH2C, SAMHD1, ADAR, and IFIH1. Am. J. Med. Genet. A. 167A, 296–312 (2015).

    PubMed  Google Scholar 

  8. Reijns, M.A. & Jackson, A.P. Ribonuclease H2 in health and disease. Biochem. Soc. Trans. 42, 717–725 (2014).

    CAS  PubMed  Google Scholar 

  9. Günther, C. et al. Defective removal of ribonucleotides from DNA promotes systemic autoimmunity. J. Clin. Invest. 125, 413–424 (2015).

    PubMed  Google Scholar 

  10. Partington, M.W., Marriott, P.J., Prentice, R.S., Cavaglia, A. & Simpson, N.E. Familial cutaneous amyloidosis with systemic manifestations in males. Am. J. Med. Genet. 10, 65–75 (1981).

    CAS  PubMed  Google Scholar 

  11. Partington, M.W. & Prentice, R.S. X-linked cutaneous amyloidosis: further clinical and pathological observations. Am. J. Med. Genet. 32, 115–119 (1989).

    CAS  PubMed  Google Scholar 

  12. Adès, L.C., Rogers, M. & Sillence, D.O. An X-linked reticulate pigmentary disorder with systemic manifestations: report of a second family. Pediatr. Dermatol. 10, 344–351 (1993).

    PubMed  Google Scholar 

  13. Anderson, R.C., Zinn, A.R., Kim, J. & Carder, K.R. X-linked reticulate pigmentary disorder with systemic manifestations: report of a third family and literature review. Pediatr. Dermatol. 22, 122–126 (2005).

    PubMed  Google Scholar 

  14. Fernandez-Guarino, M. et al. X-linked reticulate pigmentary disorder: report of a new family. Eur. J. Dermatol. 18, 102–103 (2008).

    PubMed  Google Scholar 

  15. Pezzani, L., Brena, M., Callea, M., Colombi, M. & Tadini, G. X-linked reticulate pigmentary disorder with systemic manifestations: a new family and review of the literature. Am. J. Med. Genet. A. 161A, 1414–1420 (2013).

    PubMed  Google Scholar 

  16. Kim, B.S., Seo, S.H., Jung, H.D., Kwon, K.S. & Kim, M.B. X-Linked reticulate pigmentary disorder in a female patient. Int. J. Dermatol. 49, 421–425 (2010).

    PubMed  Google Scholar 

  17. Mégarbané, H. et al. X-linked reticulate pigmentary layer. Report of a new patient and demonstration of a skewed X-inactivation. Genet. Couns. 16, 85–89 (2005).

    PubMed  Google Scholar 

  18. Fraile, G., Norman, F., Reguero, M.E., Defargues, V. & Redondo, C. Cryptogenic multifocal ulcerous stenosing enteritis (CMUSE) in a man with a diagnosis of X-linked reticulate pigmentary disorder (PDR). Scand. J. Gastroenterol. 43, 506–510 (2008).

    PubMed  Google Scholar 

  19. Picard, C., Casanova, J.L. & Puel, A. Infectious diseases in patients with IRAK-4, MyD88, NEMO, or IκBα deficiency. Clin. Microbiol. Rev. 24, 490–497 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Jaeckle Santos, L.J. et al. Refined mapping of X-linked reticulate pigmentary disorder and sequencing of candidate genes. Hum. Genet. 123, 469–476 (2008).

    CAS  PubMed  Google Scholar 

  21. Foley, S.B. et al. Use of whole genome sequencing for diagnosis and discovery in the cancer genetics clinic. EBioMedicine 2, 74–81 (2015).

    PubMed  Google Scholar 

  22. O'Donnell, M., Langston, L. & Stillman, B. Principles and concepts of DNA replication in bacteria, archaea, and eukarya. Cold Spring Harb. Perspect. Biol. 5, a010108 (2013).

    PubMed  PubMed Central  Google Scholar 

  23. Dong, Q., Copeland, W.C. & Wang, T.S. Mutational studies of human DNA polymerase α. Identification of residues critical for deoxynucleotide binding and misinsertion fidelity of DNA synthesis. J. Biol. Chem. 268, 24163–24174 (1993).

    CAS  PubMed  Google Scholar 

  24. Kunkel, T.A. & Burgers, P.M. Dividing the workload at a eukaryotic replication fork. Trends Cell Biol. 18, 521–527 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Murakami, Y., Yasuda, H., Miyazawa, H., Hanaoka, F. & Yamada, M. Characterization of a temperature-sensitive mutant of mouse FM3A cells defective in DNA replication. Proc. Natl. Acad. Sci. USA 82, 1761–1765 (1985).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Victor, R.G. et al. The Dallas Heart Study: a population-based probability sample for the multidisciplinary study of ethnic differences in cardiovascular health. Am. J. Cardiol. 93, 1473–1480 (2004).

    PubMed  Google Scholar 

  27. Petrovski, S., Wang, Q., Heinzen, E.L., Allen, A.S. & Goldstein, D.B. Genic intolerance to functional variation and the interpretation of personal genomes. PLoS Genet. 9, e1003709 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Huff, C.D. et al. Maximum-likelihood estimation of recent shared ancestry (ERSA). Genome Res. 21, 768–774 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Browning, S.R. & Browning, B.L. High-resolution detection of identity by descent in unrelated individuals. Am. J. Hum. Genet. 86, 526–539 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Dogan, R.I., Getoor, L., Wilbur, W.J. & Mount, S.M. SplicePort–an interactive splice-site analysis tool. Nucleic Acids Res. 35, W285–W291 (2007).

    PubMed  PubMed Central  Google Scholar 

  31. Kishore, S., Khanna, A. & Stamm, S. Rapid generation of splicing reporters with pSpliceExpress. Gene 427, 104–110 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Bogunovic, D. et al. Mycobacterial disease and impaired IFN-γ immunity in humans with inherited ISG15 deficiency. Science 337, 1684–1688 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Zhang, X. et al. Human intracellular ISG15 prevents interferon-α/β over-amplification and auto-inflammation. Nature 517, 89–93 (2015).

    CAS  PubMed  Google Scholar 

  34. Puel, A. et al. Chronic mucocutaneous candidiasis in humans with inborn errors of interleukin-17 immunity. Science 332, 65–68 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Lee, M.S. & Kim, Y.J. Signaling pathways downstream of pattern-recognition receptors and their cross talk. Annu. Rev. Biochem. 76, 447–480 (2007).

    CAS  PubMed  Google Scholar 

  36. Hayden, M.S. & Ghosh, S. NF-κB, the first quarter-century: remarkable progress and outstanding questions. Genes Dev. 26, 203–234 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Ysebrant de Lendonck, L. et al. Interferon regulatory factor 3 controls interleukin-17 expression in CD8 T lymphocytes. Proc. Natl. Acad. Sci. USA 110, E3189–E3197 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Asplund, A., Edqvist, P.H., Schwenk, J.M. & Pontén, F. Antibodies for profiling the human proteome-The Human Protein Atlas as a resource for cancer research. Proteomics 12, 2067–2077 (2012).

    CAS  PubMed  Google Scholar 

  39. Brown, M., Bollum, F.J. & Chang, L.M. Intracellular localization of DNA polymerase α. Proc. Natl. Acad. Sci. USA 78, 3049–3052 (1981).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Mankan, A.K. et al. Cytosolic RNA:DNA hybrids activate the cGAS-STING axis. EMBO J. 33, 2937–2946 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Dobbs, N. et al. STING activation by translocation from the ER is associated with infection and autoinflammatory disease. Cell Host Microbe 18, 157–168 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Neitzel, H. A routine method for the establishment of permanent growing lymphoblastoid cell lines. Hum. Genet. 73, 320–326 (1986).

    CAS  PubMed  Google Scholar 

  43. Uphoff, C.C., Denkmann, S.A. & Drexler, H.G. Treatment of mycoplasma contamination in cell cultures with Plasmocin. J. Biomed. Biotechnol. 2012, 267678 (2012).

    PubMed  PubMed Central  Google Scholar 

  44. Duckett, C.S., Gedrich, R.W., Gilfillan, M.C. & Thompson, C.B. Induction of nuclear factor κB by the CD30 receptor is mediated by TRAF1 and TRAF2. Mol. Cell. Biol. 17, 1535–1542 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Lois, C., Hong, E.J., Pease, S., Brown, E.J. & Baltimore, D. Germline transmission and tissue-specific expression of transgenes delivered by lentiviral vectors. Science 295, 868–872 (2002).

    CAS  PubMed  Google Scholar 

  46. Starokadomskyy, P. et al. CCDC22 deficiency in humans blunts activation of proinflammatory NF-κB signaling. J. Clin. Invest. 123, 2244–2256 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Dozmorov, I. & Lefkovits, I. Internal standard-based analysis of microarray data. Part 1: analysis of differential gene expressions. Nucleic Acids Res. 37, 6323–6339 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Dozmorov, M.G., Guthridge, J.M., Hurst, R.E. & Dozmorov, I.M. A comprehensive and universal method for assessing the performance of differential gene expression analyses. PLoS One 5, e12657 (2010).

    PubMed  PubMed Central  Google Scholar 

  49. Li, Q.Z. et al. Protein array autoantibody profiles for insights into systemic lupus erythematosus and incomplete lupus syndromes. Clin. Exp. Immunol. 147, 60–70 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Li, H., Starokadomskyy, P. & Burstein, E. Methodology to study NF-κB/RelA ubiquitination in vivo. Methods Mol. Biol. 1280, 371–381 (2015).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the patients and their families for their participation in this research project and, in particular, the XLPDR International Association for support for the study of this disease; T. Hyatt for technical assistance with allelic discrimination assays; D. Baltimore (Caltech) for the lentiviral packaging plasmids pHCMV-VSV-G, pMDLg/pRRE and pRSV-Rev; Y. Pavlov (University of Nebraska) for the plasmid pcDNA3-POLA1; L. Schmitz (Justus-Liebig-University) for the HeLa fluorescent ubiquitination-based cell cycle indicator; J. Shay (University of Texas Southwestern Medical Center) for the gene encoding human telomerase; and H. Hobbs, J. Cohen, M. Attanasio and Z. (J.) Chen for comments and suggestions. Supported by the US National Institutes of Health (R01DK073639 to Ez.B.; R56AI113274 to Ez.B. and A.R.Z.; UL1TR001105 to J.J.R.; T32AI005284 to V.P.; R01AI098569 to N.Y.; and P30CA142543 for the UT Southwestern Live Cell Imaging Facility), the Children's Medical Center Foundation (A.R.Z.) and the National Natural Science Foundation of China (81271744 to Y.Y.).

Author information

Authors and Affiliations

Authors

Contributions

P.S. performed most of the cellular and biochemical experiments; T.G. analyzed human DNA samples and performed splicing studies; J.J.R. and C.X. analyzed WGS data; R.C.W. obtained skin biopsies and assisted with the hTERT immortalization of fibroblasts; H.L., S.K., N.M., P.R. and K.K. did some biochemical analysis; V.P. and N.Y. assisted with experiments on knockout mouse embryonic fibroblast cell lines; I.D. analyzed RNA-sequencing data; G.F., Zh.X., Zi.X., L.M., Z.L., H.W., Y.Y., D.B.-A., N.O., H.M., Eu.B., G.T., E.G. and A.S. contributed subjects; E.K.W. assisted with the analysis of cytoplasmic nucleic acids; M.T.d.I.M. performed clinical immunologic studies; A.R.Z. oversaw human genetic and splicing studies; Ez.B. oversaw cellular and biochemical studies; and P.S., A.R.Z. and Ez.B. wrote the manuscript.

Corresponding authors

Correspondence to Andrew R Zinn or Ezra Burstein.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Clinical characteristics of patients with XLPDR.

(a) Photographs of the new probands depicting the typical pigmentary changes as well as facial features that are characteristic of XLPDR. (b) Histopathology depicting skin changes. Black arrows point to melanophage accumulation in the dermis; open arrows point to accumulation of amyloid. (c) A representative example of the allelic discrimination assay for the POLA1 intronic variant in XLPDR probands, heterozygous carriers, and unaffected subjects from the Dallas Heart Study. (d) Immunoblotting for POLA1 and β-actin in lymphoblastoid cell lines. (e) Quantitative RT-PCR analysis of POLA1 mRNA in lymphoblastoid cell lines with normalization to ACTB. The probands analyzed are indicated in Fig. 1. (f) Densitometry-based quantification analysis of POLA1 protein levels in lymphoblastoid cell lines from (d) with normalization to β-actin. Data are pooled from 1 independent experiment (c), or representative from 2 (e,f) or 6 (d) independent experiments (individual values (c) or mean and individual sample values (e,f)).

Supplementary Figure 2 POLA1 deficiency does not affect cell proliferation.

(a) Proliferation rate of primary fibroblasts derived from an unaffected individual (WT, F1) and his XLPDR-affected son (P1). (b) Flow cytometric analysis of cell cycle distribution in XLPDR primary fibroblasts using propidium iodide staining. Numbers indicate the proportions of cells in G1, S, G2, and M stages.(c) Mitogen-activated lymphocyte proliferation was determined in XLPDR patients (n=2) and compared to unrelated healthy controls (n=2). (d) Flow cytometric analysis of cell cycle distribution in HeLa FUCCI cells transfected with control or POLA1 siRNA. Using the stably-expressed cell cycle fluorescent markers, the proportion of cells in G1, S, G2, and M stages was determined. POLA1 expression was determined by immunoblotting and normalized to β-actin (bottom panel). (e) RT-PCR for POLA1 mRNA to detect WT and aberrant transcripts using primer pairs A+B and X+B (see Fig. 2d) in LCL samples from patients and controls. The amplified products were analyzed by agarose electrophoresis and ethidium bromide staining. (f) Immunoblotting analysis of XLPDR cell lysates with four different commercial anti-POLA1. The specific epitopes of each antibody are depicted as well. Data are pooled from 3 (a), 2 (b,d,e) or 1 (c,f) independent experiments (mean and s.e.m.(a,c), individual sample values (b,d)).

Supplementary Figure 3 POLA1 deficiency results in increased responsiveness to diverse immunological stimuli.

(a) Quantitative RT-PCR analysis of ISG activation following in XLPDR-derived fibroblasts and normal controls following stimulation with various ligands, as indicated. For activation of dsDNA cytosolic sensors, cells were transfected with 1 μg/mL poly(dA:dT) for 16 h, or 2 μg/mL ISD or HSV-60 for 24 h. For dsRNA sensors, cells were transfected with 1 μg/mL poly(I:C) for 16 h (cyto poly(I:C)), or 0.5 μg/mL 5'-ppp-dsRNA for 24 h. To activate NF-κB-dependent gene expression, cells were stimulated with TNF (1000 U/mL) or IL-1β (1 μg/mL) for 0, 2, and 12 hours. To activte TLRs, cells were stimulated by adding to the culture media the following TLR agonists: poly(I:C) (250 μg/mL) for 2 and 7 hours, Pam3CSK4 (250 ng/mL) for 24 h, or LPS (10 μg/mL) for 24 h. (b) Quantitative RT-PCR analysis of ISG activation in cells with induced POLA1 deficiency following siRNA. After 48 h, cells were stimulated similarly to (a). Data are pooled from 4 independent experiments (a,b) (mean and s.e.m.(a,b)).

Supplementary Figure 4 Analyses of autoantibody production and leukocyte ultrastructure in patients with XLPDR.

(a) ELISA for anti-nuclear antibodies (ANA) in plasma from unaffected individuals (WT, n=16) and XLPDR probands (XLPDR, n=6). (b) Heatmap of the result of the large screen on presence of DNA-, RNA-, and protein-specific IgG and IgM autoantibodies in plasma of unaffected control individuals (WT, n=16) and XLPDR patients (XLPDR, n=6). The first two rows depict positive controls (total and control IgG or IgM, correspondingly). Full list of antigens is presented in Supplementary Table 4. Blue – low abundance, red – high abundance. (c) Electron microscopy of leukocytes from an XLPDR patient and a WT control. Magnification bar – 2 μM. *p=0.102 (unpaired Student's t-test). Experiments a-c were performed one time (mean and individual sample values (a), individual values (b)).

Supplementary Figure 5 Effects of POLA1 deficiency on the NF-κB pathway.

(a) Immunofluorescence analysis of TNF-induced nuclear accumulation of RelA (red) in XLPDR fibroblasts. Scale bar – 10 m. (b) Immunoblotting analysis of TNF-induced nuclear accumulation of RelA in XLPDR fibroblasts. (c) Immunoblotting analysis of total and phosphorylated forms of RelA in POLA1-deficient cells after TNF stimulation (left panel). Quantification of relative levels of two indicated forms of phospho-RelA was performed by quantitative fluorescence imaging (Li-COR, right panels). Data are representative from 2 (a) or 1 (b,c) independent experiment (individual values in (c)).

Supplementary Figure 6 Detection of cytosolic RNA:DNA.

(a) For the experiment shown in Fig. 5c, immunoblotting of the cytosolic and nuclear markers (RelA and p84, respectively) is shown (input fractions). In addition, the RNA:DNA immunoprecipitation was coupled to a control precipitation (RelA) to ensure equal lysate loading and bead preparation (IP). (b) Immunoprecipitation of RNA:DNA from three cell lines. Representative images of multiple beads stained with Picogreen from the experiment shown in Fig. 5c are shown here. Data are representative of 3 (a) and 4 (b) independent experiments.

Supplementary Figure 7 POLA1 localizes to the cytosol and co-localizes with RNA:DNA

(a) Immunofluorescence staining for POLA1 in the indicated cell lines. Cells were fixed and permeabilized with 100% methanol; POLA1 is visualized in green, and the nuclei were stained with DAPI. Scale bar – 10 μm. (b) Immunoblotting for POLA1 in nuclear and cytoplasmic fractions of the indicated cell lines. C - cytoplasm, N - nucleus. (c) Immunofluoresce staining for POLA1 (green) and RNA:DNA (red) in dermal fibroblasts. Scale bar – 10 μm. Red squares depict the original position of the enlarged area. Data are representative of 2 independent experiments (a,b,c).

Supplementary Figure 8 Quality control for the cell-fractionation procedure in Figure 6f.

Immunoblotting for POLA1 and the fraction marker proteins GAPDH (cytosol) and p84 (nucleus) was performed for the experiment shown in Fig. 6f. The experiment was performed one time.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8 and Supplementary Tables 2, 5 and 6 (PDF 1533 kb)

Supplementary Table 1

Normalized gene expression data for RNA-seq experiments presented in Fig. 3b,c,d (XLSX 71 kb)

Supplementary Table 3

Summary table of clinical immunological data (XLSX 23 kb)

Supplementary Table 4

Extended data of screen on presence of IgG and IgM autoantibodies in plasma from XLPDR probands and unaffected indivduals, presented in Supplementary Fig. 4b (XLSX 132 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Starokadomskyy, P., Gemelli, T., Rios, J. et al. DNA polymerase-α regulates the activation of type I interferons through cytosolic RNA:DNA synthesis. Nat Immunol 17, 495–504 (2016). https://doi.org/10.1038/ni.3409

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ni.3409

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing