Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Sep 4.
Published in final edited form as: Nat Genet. 2012 Sep 23;44(11):1243–1248. doi: 10.1038/ng.2414

Mutations in ADAR1 cause Aicardi-Goutières syndrome associated with a type I interferon signature

Gillian I Rice 1, Paul R Kasher 1, Gabriella M A Forte 1, Niamh M Mannion 2, Sam M Greenwood 2, Marcin Szynkiewicz 1, Jonathan E Dickerson 1, Sanjeev S Bhaskar 1, Massimiliano Zampini 1, Tracy A Briggs 1, Emma M Jenkinson 1, Carlos A Bacino 3, Roberta Battini 4, Enrico Bertini 5, Paul A Brogan 6, Louise A Brueton 7, Marialuisa Carpanelli 8, Corinne De Laet 9, Pascale de Lonlay 10, Mireia del Toro 11, Isabelle Desguerre 12, Elisa Fazzi 13, Àngels Garcia-Cazorla 14,15, Arvid Heiberg 16, Masakazu Kawaguchi 17, Ram Kumar 18, Jean-Pierre S-M Lin 19, Charles M Lourenco 20, Alison M Male 21, Wilson Marques Jr 20, Cyril Mignot 22,24, Ivana Olivieri 25, Simona Orcesi 25, Prab Prabhakar 26, Magnhild Rasmussen 27, Robert A Robinson 26, Flore Rozenberg 28, Johanna L Schmidt 29, Katharina Steindl 30, Tiong Y Tan 31, William G van der Merwe 32, Adeline Vanderver 29, Grace Vassallo 33, Emma L Wakeling 34, Evangeline Wassmer 35, Elizabeth Whittaker 36, John H Livingston 37, Pierre Lebon 28, Tamio Suzuki 17, Paul J McLaughlin 38, Liam P Keegan 2, Mary A O’Connell 2, Simon C Lovell 39, Yanick J Crow 1
PMCID: PMC4154508  NIHMSID: NIHMS624328  PMID: 23001123

Abstract

Adenosine deaminases acting on RNA (ADARs) catalyze the hydrolytic deamination of adenosine to inosine in double-stranded RNA (dsRNA) and thereby potentially alter the information content and structure of cellular RNAs. Notably, although the overwhelming majority of such editing events occur in transcripts derived from Alu repeat elements, the biological function of non-coding RNA editing remains uncertain. Here, we show that mutations in ADAR1 (also known as ADAR) cause the autoimmune disorder Aicardi-Goutières syndrome (AGS). As in Adar1-null mice, the human disease state is associated with upregulation of interferon-stimulated genes, indicating a possible role for ADAR1 as a suppressor of type I interferon signaling. Considering recent insights derived from the study of other AGS-related proteins, we speculate that ADAR1 may limit the cytoplasmic accumulation of the dsRNA generated from genomic repetitive elements.


Aicardi-Goutières syndrome (MIM 225750) is a genetically determined inflammatory disorder particularly affecting the brain and skin. In its most characteristic form, AGS is a clinical mimic of in utero–acquired infection1 and, like congenital infection, is associated with increased production of the antiviral cytokine interferon α (IFN-α)2. AGS can result from mutations in any one of the genes encoding the DNA exonuclease TREX1 (AGS1)3, the three non-allelic components of the RNase H2 endonuclease complex (RNASEH2B, AGS2; RNASEH2C, AGS3; and RNASEH2A, AGS4)4 and the deoxynucleoside triphosphate triphosphohydrolase SAMHD1 (AGS5)5,6. Although AGS is most typically inherited as an autosomal recessive trait7, rare examples of disease due to de novo dominant mutations in TREX1 have been reported8-10.

Studies of the function of TREX1 have delineated a cell-intrinsic mechanism for the initiation of an autoimmune response by interferon (IFN)-stimulatory nucleic acid11,12, begging the question of the source of nucleic acid inducing the type I IFN–mediated immune disturbance in AGS. In this regard, it has been shown that TREX1 can metabolize reverse-transcribed DNA and that single-stranded DNA derived from endogenous retroelements accumulates in TREX1-deficient cells11. On a related note, TREX1 (ref. 13), SAMHD1 (refs. 14–16) and RNase H2 (ref. 17) have been implicated in the metabolism of the (exogenous) retrovirus HIV-1. Perhaps most notably, a recent study showed rescue of the lethal inflammatory TREX1-null mouse phenotype by a combination of reverse transcriptase inhibitors (antiretroviral therapy as used to treat HIV-1)18, suggesting that the accumulation of cytosolic DNA in TREX1-null cells can be ameliorated by inhibiting endogenous retroelement cycling.

To define other genes relevant to the AGS phenotype, we undertook whole-exome sequencing in four individuals with a clinical diagnosis of AGS, all of whom screened negative for mutations in TREX1, RNASEH2A, RNASEH2B, RNASEH2C and SAMHD1. Using in-solution hybridization followed by massively parallel sequencing, we derived over 2 Gb of mapped sequence for each subject, such that an average of 56-fold coverage was achieved across the exome for all the samples (Supplementary Table 1). We performed an analysis of the called nonsynonymous, splice-site, substitution and coding insertion and/or deletion exome variants under a model of a rare autosomal recessive disorder. Visual inspection of the generated data identified two affected individuals, AGS81_P1 and AGS219, who each had two nonsynonymous coding alterations in ADAR1, a gene we had already highlighted as a candidate for AGS in view of its known role as a suppressor of type I IFN signaling19,20. Sanger sequencing confirmed the variants in these individuals, as well as in two further affected siblings from family AGS81. In light of these data, we proceeded to sequence ADAR1 in other individuals lacking mutations in TREX1, RNASEH2B, RNASEH2C, RNASEH2A and SAMHD1 (AGS1AGS5, respectively) from our AGS cohort.

In total, 12 affected individuals from 8 families harbored biallelic ADAR1 variants (Fig. 1, Table 1 and Supplementary Table 2), which were considered likely pathogenic on the basis of species conservation (Supplementary Figs. 1 and 2) and the output of pathogenicity prediction packages (Supplementary Table 3). In these families, all parents tested were heterozygous for one putative mutation. Two further unrelated affected individuals, AGS150 and AGS474, harbored a single mutation encoding p.Gly1007Arg that was not present in either parent. Genotyping of microsatellite markers was consistent with stated paternity, indicating that this variant had arisen de novo in both cases (Supplementary Table 4). Of the nine distinct ADAR1 mutations we identified, the c.577C>G (p.Pro193Ala) transversion was seen in the compound heterozygous state in five families of European ancestry. This same variant was observed in 41 subjects (32 of 4,350 European-Americans and 9 of 2,203 African-Americans) annotated on the Exome Variant Server database, whereas none of the other ADAR1 variants present in our AGS cohort were found in more than 12,000 control alleles.

Figure 1.

Figure 1

Schematic of the human ADAR1 gene. (a) ADAR1 spans 26,191 bp of genomic sequence on chromosome 1q21.3 (154,554,533–154,580,724). Neighboring genes are also shown. Cen, centromeric; tel, telomeric. (b) Position of identified mutations within the genomic sequence of the ADAR1 long isoform (p150). The number of alleles with each mutation is shown in parentheses. *, the mutation encoding p.Gly1007Arg identified as a single heterozygous de novo change in two families; †, the same mutation identified in identical twins (therefore counted once). Numbers given above the gene indicate the relevant exons (only exons with mutations are numbered). The shorter isoform (p110) of ADAR1 starts at c.886 of the p150 isoform. (c) Position of identified variants within the ADAR1 p150 1,226-amino-acid protein. Numbers above the protein are the amino-acid count at the exon boundaries. The shorter isoform starts at amino acid 296 of p150, giving rise to a 931-amino-acid protein. (d) Schematic of the position of protein domains and their amino-acid boundaries in the p150 isoform of ADAR1. Note that the p110 isoform does not include the Zα DNA/RNA-binding domain and nuclear export signal.

Table 1.

Ancestry, pedigree structure, consanguinity status and sequence alterations in ADAR1 mutation–positive families in an AGS cohort

Family Ancestry Individuals tested Consanguinity Nucleotide alteration Exon Amino-acid alteration
AGS81 Norwegian 3A, M, F c.[577C>G]+[2675G>A] 2, 9 p.[Pro193Ala]+[Arg892His]
AGS93 Italian 1A, M, F c.[577C>G]+[2608G>A] 2, 8 p.[Pro193Ala]+[Ala870Thr]
AGS107 Pakistani 2A, M, F + c.3337G>C (hom) 14 p.Asp1113His
AGS150 Brazilian 1A, M, F c.3019G>A (het, de novo) 11 p.Gly1007Arg
AGS219 Pakistani 1A + c.3335A>T (hom) 14 p.Tyr1112Phe
AGS228 Indian 1A, M, F c.2997G>T (hom) 11 p.Lys999Asn
AGS251 White British 1A, M c.[577C>G]+[2615T>C] 2, 8 p.[Pro193Ala]+[Ile872Thr]
AGS327 Italian 1A c.[577C>G]+[1076_
1080del]
2, 2 p.[Pro193Ala]+[Lys359Argfs*14]
AGS430 Spanish 2Aa, M, F c.[577C>G]+[2675G>A] 2, 9 p.[Pro193Ala]+[Arg892His]
AGS474 European-American 1A, M, F c.3019G>A (het, de novo) 11 p.Gly1007Arg

A, affected individual; M, mother; F, father; het, heterozygous; hom, homozygous.

a

Identical twins.

ADARs catalyze the hydrolytic deamination of adenosine to inosine in dsRNA21. Four ADARs have been described in mammals (ADAR1, ADAR2, ADAR3 and TENR), although only ADAR1 and ADAR2 are known to have catalytic activity. ADAR1 is encoded by a single-copy gene that maps to human chromosome 1q21. Two main isoforms of ADAR1 are present in mammalian cells: a truncated ADAR1 protein (p110; nucleotide, NM_001025107.2; protein, NP_001020278.1) is constitutively expressed, whereas a full-length form of ADAR1 (p150; nucleotide, NM_001111.4; protein, NP_001102.2) is IFN inducible22. Both isoforms have been shown to shuttle between the nucleus and the cytoplasm. ADAR1 is a modular protein with a C-terminal deaminase catalytic domain, three centrally located dsRNA-binding domains (dsRBDs) and one or two N-terminal Z-DNA–binding domains; compared with p110, the p150 isoform of human ADAR1 possesses an additional 295 N-terminal amino acids containing a nuclear export signal and an extra Z-DNA/Z-RNA–binding domain (designated Zα) (Fig. 1)23.

Of the eight amino-acid substitutions identified in our AGS cohort, seven involve residues situated in the catalytic domain of ADAR1 (Fig. 2). Five of these seven catalytic domain residues (Arg892, Lys999, Gly1007, Tyr1112 and Asp1113) lie along the surface of the protein that interacts with dsRNA (Fig. 2a), and the two others (Ala870 and Ile872) lie internal to the domain structure and are predicted to destabilize the protein (Fig. 2b,c). In contrast, Pro193 is positioned within the Z-DNA/Z-RNA–binding domain. In the wild-type protein, Pro193 makes direct contact with the nucleic acid, and substitution of this residue with alanine removes important atomic interactions between the protein and DNA/RNA (Fig. 2d,e).

Figure 2.

Figure 2

Structural context of ADAR1 protein substitutions. (a) The surface of the ADAR domain (dark pink) with surface substitutions highlighted in bright pink. The active site contains a zinc ion (black) in the center. Arg892His, Lys999Asn, Gly1007Arg, Tyr1112Phe and Asp1113His are all on the same side of the domain as the active site, and all have the potential to alter charge and/or hydrogen bonding characteristics of the surface in the region that is likely to be responsible for RNA binding. (b,c) Models of wild-type (b) and mutant (c) residues at position 870 in the ADAR domain. Interactions between the residue and the surrounding protein structure are indicated by all-atom contact dots (blue). Green dots represent energetically favorable van der Waals interactions, whereas red and pink spikes indicate unfavorable van der Waals overlaps. The Ile872Thr substitution introduces an unsatisfied hydrogen bond donor/acceptor group, which is destabilizing but not easily depicted. (d,e) Interactions of the proline residue at position 193 (Pro193) in the Z-DNA–binding domain (d). Contact dots (blue, green, yellow) indicate favorable interactions between Pro193 and the DNA backbone (white). These interactions are absent in the mutant form (e). (f) Modeling of the deaminase domain of ADAR2 suggests contact with dsRBD2 close to Gly1007, highlighting the possibility for an arginine residue to make functionally important polyvalent interactions.

More than 130 different ADAR1 mutations (Supplementary Table 5) have already been documented in individuals with dyschromatosis symmetrica hereditaria 1 (DSH), an autosomal dominant disorder characterized by the childhood onset of hypo- and hyperpigmented macules on the face and dorsal aspects of the extremities24,25. The frequent observation of stop and frameshift ADAR1 variants in individuals with DSH indicates haploinsufficiency as the likely molecular pathology. However, ADAR1 missense variants, spread throughout the gene, are also commonly seen in association with the DSH phenotype. All except one of the ADAR1 mutations recorded in our AGS cohort were missense variants, and protein blotting of lymphoblastoid cells from affected individuals showed normal levels of both the constitutive and IFN-inducible isoforms of ADAR1 (Fig. 3). We predict that the proteins containing the amino-acid alterations seen in individuals with AGS act as hypomorphs and that, as in the Adar1-null mouse, complete loss of ADAR1 protein activity is embryonic lethal19.

Figure 3.

Figure 3

Protein blot of lymphoblastoid cell lines (LCLs). Protein blot analysis of ADAR1 expression in Epstein-Barr virus (EBV)-transformed LCLs from one unrelated control and two affected individuals (AGS93, AGS107). Whole-cell lysates were derived from 1 × 107 cells per sample, and 10 μg of total protein was loaded per lane. To test antibody specificity and confirm IFN induction of the p150 isoform of ADAR1, unstimulated cells were compared to IFN-stimulated cells. The antibody to ADAR1 recognizes both the p110 and p150 isoforms. Immunoblotting of tubulin (50–55 kDa) was used as a loading control.

Of the nine discrete ADAR1 mutations observed in our AGS cases, only the allele encoding p.Gly1007Arg has been reported previously26,27. Uniquely, this mutation was described in two individuals with DSH also demonstrating neurodegeneration with dystonia and intracranial calcification (Supplementary Table 6). Using a known ADAR1 editing substrate, miR376-a2, we found that, of six ADAR1 p110 mutants (Ala870Thr, Ile872Thr, Arg892His, Lys999Asn, Gly1007Arg and Asp1113His) expressed following transfection of plasmid constructs into HEK293 cells, only the Gly1007Arg variant showed a significant effect on editing, with levels of editing equivalent to those seen with inactive protein (Fig. 4a). Modeling to position the deaminase domain active site at the target adenosine along the dsRNA substrate (Fig. 2f) predicts that the deaminase domain also contacts dsRBD2 close to Gly1007. The proximity of Gly1007 to the RNA backbone, and the possibility for an arginine residue to make polyvalent interactions there, suggests a mechanism whereby Arg1007 might confer a dominant-negative effect: by binding more tightly to RNA, the mutant protein could act as a competitive inhibitor of wild-type protein, while being itself catalytically inactive. In keeping with this model, a plasmid expressing Gly1007Arg ADAR1 showed stronger inhibition of wild-type ADAR1 than equivalent amounts of a plasmid expressing catalytically inactive ADAR1 (Fig. 4b). Although these observations might explain the seemingly unique nature of the Gly1007Arg variant, how the other mutations that were tested in our system cause an aberrant phenotype remains unclear at this time. Possible explanations might relate, for example, to non-editing functions of ADAR1 (ref. 28) or to editing substrate29 or cell type specificity19.

Figure 4.

Figure 4

Site-specific and competition editing assays. (a) HEK293 cells were co-transfected with 500 ng of a plasmid expressing miR376-a2 and 500 ng of a plasmid expressing wild-type (WT) ADAR1 or ADAR1 mutants. Background editing in HEK293 cells with 500 ng of substrate plasmid is approximately 20%. As previously observed, only one monomer in an ADAR1 dimer is required to be enzymatically active, such that addition of inactive protein initially increases editing40. Editing activity is expressed as a proportion of WT ADAR1 editing activity, which is 1 (y axis). Error bars, s.e.m. ***P = 0.0005. (b) Competition between ADAR1 and inactive ADAR1 mutants. HEK293 cells were co-transfected with 200 ng of a plasmid expressing miR376-a2 and 200 ng of a plasmid expressing ADAR1 p110 in the presence of increasing amounts of a plasmid expressing a catalytically inactive form of ADAR1 (Glu912Ala)28 or Gly1007Arg. Error bars, s.e.m. **P = 0.0026.

The recurrent p.Pro193Ala alteration lies in the Zα DNA/RNA-binding domain, thus implicating the IFN-inducible p150 isoform of ADAR1 in the AGS phenotype. Mice lacking ADAR1 die by around embryonic day 12.5 owing to defective hematopoiesis and wide-spread apoptosis, which are associated with global upregulation of IFN-stimulated genes (ISGs), indicating that ADAR1 acts as a suppressor of type I IFN signaling19. In light of these observations, using whole blood from 8 ADAR1 mutation–positive individuals, we performed quantitative RT-PCR to analyze the mRNA levels of 15 ISGs. Compared to nine controls, all tested individuals with mutations in ADAR1, including AGS150 and AGS474 harboring a de novo heterozygous allele encoding p.Gly1007Arg, showed a consistent pattern of ISG upregulation (Supplementary Fig. 3). We then undertook an analysis of a subset of the 6 most highly expressed ISGs in 10 ADAR1 mutation–positive AGS cases, 6 sets of parents with heterozygous mutations in ADAR1 and 18 ADAR1 mutation–positive individuals with DSH (Fig. 5 and Supplementary Table 7). For the six ISGs assayed, expression was variably higher in AGS heterozygous parents and DSH cases versus controls, whereas individuals with a clinical diagnosis of AGS (due either to biallelic mutations in ADAR1 or a heterozygous mutation resulting in a p.Gly1007Arg amino-acid substitution) had even higher levels of expression.

Figure 5.

Figure 5

Quantitative RT-PCR of a panel of six ISGs in whole blood measured in individuals with AGS, their parents and individuals with DSH. Scatter plots showing log10-transformed RQ values for a panel of 6 ISGs measured in whole blood from 10 AGS cases with mutations in ADAR1, 6 sets of parents heterozygous for mutations in ADAR1, 18 individuals with ADAR1 mutation–positive DSH and 20 healthy controls. All genes were significantly upregulated in AGS cases (P < 0.001) compared to controls. RQ is equal to 2−ΔΔCT, with −ΔΔCT ± s.d., that is the normalized fold change relative to a calibrator. ***P ≤ 0.001; **P ≤ 0.01; *P < 0.05.

The skin lesions typical of DSH have not been described in AGS (they are distinct from AGS-related chilblains)30, and none of the ADAR1 mutation–positive AGS cases included in our study had obvious features of DSH. Of note, DSH has only very rarely been reported outside of Japan and China. Moreover, even within known families segregating the DSH phenotype, a marked variability in expression is well recognized31. Although our findings do not necessarily implicate dysregulation of type I IFN in the DSH phenotype, they suggest that missense and null heterozygous mutations in ADAR1 are consistent with both DSH (depending on, for example, ancestry) and with carrier or, in the case of the p.Gly1007Arg alteration, affected status for AGS. Thus, we believe that the lack of DSH skin features in our AGS cases and their heterozygous parents most likely relates to non-penetrance and/or variable expressivity due to ancestry or other genetic or non-genetic factors, and we predict that two individuals with DSH would have a 1 in 4 risk of having a child with AGS.

Because inosine is recognized as guanosine by the translation and splicing machineries, editing of adenosine to inosine can alter the protein-coding information of messenger RNA and the structural stability of dsRNA. The first identified ADAR-edited substrates were in codons, and ADARs were presumed to function primarily in proteome diversification. However, although codon editing is clearly important, it represents only a small fraction of editing events in the transcriptome, with editing sites in RNA derived from non-coding regions, most particularly Alu elements, being vastly more prevalent32-35. The biological function of such repetitive element editing is uncertain36. Recent studies have highlighted an antagonistic interaction between ADARs and the RNA interference machinery, suggesting that a pool of common RNA substrates is capable of engaging both pathways37 and specifically implicating the binding properties of the p150 isoform of ADAR1, beyond its editing activity28,38, in this relationship. How, then, does loss of ADAR1 activity lead to upregulation of type I IFN signaling? Theoretically, wild-type ADAR1 might edit specific currently undefined transcripts that are important in IFN regulation. Alternatively, and perhaps more likely, lack of editing in ADAR1-deficient cells may lead to an increase in immunoreactive dsRNA37 and/or result in failure to generate inosine:uracil (IU)-dsRNA with an intrinsic function in suppressing IFN induction39. Considering insights derived from the study of TREX1, SAMHD1 and RNase H2, we speculate that ADAR1 has a role in the metabolism of retroelements and that a major contribution of ADARs to evolutionary fitness may be to regulate the accumulation of dsRNA generated from basally transcribed repetitive sequences within the genome.

URLs. UCSC Human Genome Browser, http://genome.ucsc.edu/; Ensembl, http://www.ensembl.org/; dbSNP, http://www.ncbi.nlm.nih.gov/projects/SNP/; Exome Variant Server, National Heart, Lung, and Blood Institute (NHLBI) Exome Sequencing Project (ESP) (accessed 21 March 2012), http://snp.gs.washington.edu/EVS/; PolyPhen, http://genetics.bwh.harvard.edu/pph/; SIFT, http://sift.jcvi.org/www/SIFT_enst_submit.html; Align GVGD, http://agvgd.iarc.fr/index.php; Clustal Omega, http://www.ebi.ac.uk/Tools/msa/clustalo/; Protein Data Bank (PDB), http://www.pdb.org/; Alamut, http://www.interactive-biosoftware.com/.

ONLINE METHODS

Affected individuals and families

All affected individuals included in this study had a clinical diagnosis of AGS that was based on the presence of early-onset encephalopathy (at <18 months of age), negative investigations for common prenatal infections, intracranial calcification with or without white matter changes in a typical distribution, elevated levels of IFN-α with or without pterins in the cerebrospinal fluid and/or chilblains. Clinical information and samples were obtained with informed consent. The study was approved by the Leeds (East) Research Ethics Committee (reference 10/H1307/132).

Exome sequencing

Genomic DNA was extracted from lymphocytes from affected individuals and parents by standard techniques. For whole-exome analysis, targeted enrichment and sequencing were performed on 3 μg of DNA extracted from the peripheral blood of four individuals (AGS81_P1, AGS125, AGS163 and AGS219). Enrichment was undertaken using the 38 Mb SureSelect Human All Exon kit (Agilent) following the manufacturer’s protocol, and samples were paired-end sequenced on an Illumina HiSeq 2000. Sequence data were mapped using BWA (Burrows-Wheeler Aligner) against the hg18 (NCBI Build 36) human genome as a reference. Variants were called using SOAPsnp and SOAPindel (from the Short Oligonucleotide Analysis Package) with medium stringency and were then filtered for those with greater than 5× coverage.

Sanger sequencing

Primers were designed to amplify the coding exons of ADAR1 (Supplementary Table 8). Purified PCR amplification products were sequenced using BigDye terminator chemistry and an ABI 3130 DNA sequencer. Mutations were annotated on the basis of the reference cDNA sequence NM_001111.4, with nucleotide numbering beginning from the first A in the initiating ATG codon.

Gene expression analysis

The expression of 15 genes known to be interferon stimulated was assessed in whole blood. Total RNA was extracted from whole blood using the PAXgene RNA isolation kit (PreAnalytix). RNA concentration was assessed using a spectrophotometer (FLUOstar Omega, Labtech). Quantitative RT-PCR analysis was performed using TaqMan Universal PCR Master Mix (Applied Biosystems) and cDNA derived from 40 ng of total RNA. The relative abundance of target transcripts, measured using TaqMan probes for Ly6E (Hs00158942_m1), MX1 (Hs00895598_m1), USP18 (Hs00276441_m1), RSAD2 (Hs01057264_m1), OAS1 (Hs00973637_m1), IFI44L (Hs00199115_m1), IFI27 (Hs01086370_m1), ISG15 (Hs00192713_m1), IFIT1 (Hs00356631_g1), IFI44 (Hs00197427_m1), IFI6 (Hs00242571_m1), SIGLEC1 (Hs00988063_m1), IFIT3 (Hs00155468_m1), IRF7 (Hs00185375_m1) and STAT1 (Hs01013989_m1), was normalized to the expression level of HPRT1 (Hs03929096_g1) and 18s (Hs999999001_s1) and assessed with Applied Biosystems StepOne Software v2.1. Statistical significance between groups was determined by t tests using DataAssist v2.0 (Applied Biosystems). Data from affected individuals are expressed relative to the average of nine normal controls. A subset of the six most highly expressed ISGs (RSAD2, IFI44L, IFI27, ISG15, IFIT1 and SIGLEC1) were measured in 20 controls, 10 ADAR1 mutation–positive cases, 6 sets of ADAR1 heterozygous parents and 18 individuals with ADAR1 mutation–positive DSH. RNA from individuals with DSH was extracted using the QIAamp RNA Blood Mini kit (Qiagen), and cDNA was derived as above. Statistical significance between groups was determined by Kruskall Wallis tests using GraphPad Prism 5. The median fold change of the 6 ISGs compared to the median of the 20 healthy controls was used to create a score for each affected individual, similarly to previously described methods41,42.

Interferon stimulation

EBV-transformed lymphoblastoid cells from affected individuals and controls were counted using a Bright-Line Hemacytometer (Sigma). We then transferred 10 ml of cells at a concentration of 1 × 106 cells/ml to each of two T25 flasks per cell line. One flask for each cell line was stimulated with 1,000 international units (IU)/ml of human IFN-α (human Intron A, Shering-Plough) for 24 h.

Protein analysis

Whole-cell lysates were prepared from lymphoblastoid cells (1 × 107 cells per sample) using 10 mM EDTA–RIPA buffer containing protease inhibitors (Roche). For protein blot analysis, 10 μg of total protein was loaded onto 8% SDS-PAGE gels, and electrophoresis was performed using the Mini-PROTEAN system (Bio-Rad Laboratories). Following wet-blotting transfer of the proteins onto PVDF membrane (Amersham), non-specific antibody binding was blocked using 5% normal goat serum (Vector Laboratories) and 2% BSA (Sigma) in 0.1% TBS-Tween (Sigma) overnight at 4 °C with gentle agitation. Rabbit primary antibody to ADAR (Sigma, Prestige Antibodies, HPA003890) was incubated with the membranes for 2 h at room temperature using a dilution of 1:800. As a loading control, membranes were incubated with a 1:4,000 dilution of rabbit primary antibody to α/β-tubulin (Cell Signaling Technology, 2148). Horseradish peroxidase (HRP)-labeled goat secondary antibody to rabbit (1:10,000; Cell Signaling Technology, 7074) was incubated with the membranes for 1 h at room temperature to detect both ADAR1 and tubulin antibodies. Signal was detected using a 1:10 dilution of Enhanced Chemiluminescence reagents (Lumigen).

Microsatellite genotyping

To exclude non-paternity, informative polymorphic microsatellite markers on chromosomes 3 (D3S3640 and D3S3560), 11 (D11S913 and D11S1889) and 20 (D20S847, D20S896 and D20S843) were genotyped using DNA from AGS150, AGS475, their parents and an unrelated control sample. DNA samples were amplified by standard PCR (primer sequences available upon request). Each amplicon was mixed with Hi-Di Formamide (Applied Biosystems) and 500 ROX Size Standard (Applied Biosystems) and run on the Genetic Analyzer 3010 capillary electrophoresis system. Results were analyzed with GeneMapper v4.1 software (Applied Biosystems).

Editing assays

Transient transfections were performed with slight adaptations to a previously described protocol28 in 24-well plates seeded the day before transfection with 0.14 × 106 cells using Lipofectamine 2000 and OptiMEM I reduced serum medium. Cells were transfected with 500 ng of plasmid expressing a known ADAR1 editing substrate, miR376-a2, together with 500 ng of plasmid expressing wild-type ADAR1 p110 protein or the mutants Ala870Thr, Ile872Thr, Arg892His, Lys999Asn, Gly1007Arg and Asp1113His, for which coding sequences had been subcloned into the pcDNA3.1 expression vector. Reduced serum medium was replaced after 6 h with DMEM, and cells were harvested after 48 h. Total RNA was extracted with TRIzol reagent (Invitrogen), purified and treated with TURBO DNase (Ambion) and RNasin Plus RNase Inhibitor (Promega). Reverse transcription was carried out using SuperScript II and random hexamers, and cDNA was amplified by PCR with Platinum Taq DNA polymerase and specific primers for pri-miR376-a2 (Supplementary Table 8). The 300-bp PCR product was purified with exonuclease in combination with shrimp alkaline phosphatase (SAP) before sequencing. RNA editing of the pri-miR376-a2 transcript was identified as an adenosine-to-guanine change within the cDNA sequence, and the editing level was expressed as a percentage. Peak heights were measures for edited (G) and unedited (A), the editing ratio percentage was calculated by (G / (A + G)) × 100. Results are the average of two separate experiments. Statistical significance between groups was determined by two-tailed t tests using Excel.

Protein modeling

The ADAR1 substitutions p.Ala870Thr, p.Ile872Thr, p.Arg892His, p.Lys999Asn, p.Gly1007Arg, p.Tyr1112Phe and p.Asp1113His all fall within the adenosine deaminase domain. No crystal structure is available for this domain in human ADAR1; thus, a comparative model was constructed using the deaminase domain of human ADAR2 (Protein Data Bank (PDB) 1zy7)43. The sequences of ADAR1 and ADAR2 were aligned using ClustalW44, and 25 models were built using Modeller45. The model with the lowest discrete optimized protein energy (DOPE) score was selected (representing the most probable model judged by the fit to the Modeller statistical potential). Residues 975–996 could not be modeled accurately, as they have no equivalent in ADAR2; for this reason, these residues were not analyzed further. Models of mutations were built using KiNG46, hydrogen atoms were added with Reduce47, and all-atom contacts were calculated with Probe48. In each case, all low-energy rotamers49 were considered, and the rotamer with the best Probe score was used. We assumed a similar binding mode for ADAR2 to that described for the tRNA deaminase–tRNA anticodon complex50, allowing us to orientate the deaminase domain of ADAR2 on dsRNA and position it relative to the ADAR2-dsRBD2 complex51. The crystal structure of the ADAR1 Z-DNA–binding domain52 (PDB 1qbj) was used to analyze the likely structural effect of the p.Pro193Ala substitution. As the side chain of alanine has no degrees of freedom from rotatable dihedral angles, the position of the side chain is determined by the conformation of the protein backbone.

Supplementary Material

Vanderver_SD_Fig2
Vanderver_SD_Fig3-5
vanderver_SD_Fig1

ACKNOWLEDGMENTS

We sincerely thank the participating families for the use of genetic samples and clinical information. We thank B. Hamel, H. Brunner and other clinical collaborators for contributing samples not included in the current manuscript, A.P. Jackson for highlighting the presence of intracranial calcification in individuals with DSH with a p.Gly1007Arg alteration in ADAR1 and D.B. Stetson and D.T. Bonthron for critical reading of the manuscript. D.B. Stetson (University of Washington) provided the ADAR1 constructs used in the editing assays. We thank the NHLBI GO Exome Sequencing Project and its ongoing studies that produced and provided exome variant calls for comparison: the Lung GO Sequencing Project (HL-102923), the Women’s Health Initiative (WHI) Sequencing Project (HL-102924), the Broad GO Sequencing Project (HL-102925), the Seattle GO Sequencing Project (HL-102926) and the Heart GO Sequencing Project (HL-103010). Y.J.C. acknowledges the Manchester National Institute for Health Research (NIHR) Biomedical Research Centre. The research leading to these results has received funding from the European Union’s Seventh Framework Programme (FP7/2007-2013) under grant agreement 241779 and from the Great Ormond Street Hospital Children’s Charity.

Footnotes

Note: Supplementary information is available in the online version of the paper.

AUTHOR CONTRIBUTIONS

G.I.R. performed quantitative PCR analysis. P.R.K. performed protein blot analysis. G.M.A.F. performed cell culture and interferon stimulation experiments with assistance from G.I.R. and T.A.B. M.S. performed Sanger sequencing with assistance from M.Z., G.M.A.F. and E.M.J. G.I.R. analyzed sequence data. M.S. and G.M.A.F. undertook microsatellite genotyping. J.E.D. and S.S.B. undertook analysis of the exome sequence data. J.H.L. was responsible for neuroradiological phenotyping. P.L. and F.R. measured interferon activity in affected individuals. M.A.O., L.P.K., S.M.G. and N.M.M. carried out ADAR1 editing assays. T.S. and M.K. provided the DSH RNA samples. S.C.L. and P.J.M. carried out ADAR1 structural analysis. Y.J.C. designed and supervised the project and wrote the manuscript with support from G.I.R. C.A.B., R.B., E.B., P.A.B., L.A.B., M.C., C.D.L., P.d.L., M.d.T., I.D., E.F., A.G.-C., A.H., R.K., J.-P.S.-M.L., C.M.L., A.M.M., W.M., C.M., I.O., S.O., P.P., M.R., R.A.R., J.L.S., K.S., T.Y.T., W.G.v.d.M., A.V., G.V., E.L.W., E. Wassmer and E. Whittaker identified subjects with AGS or assisted with related clinical and laboratory studies.

COMPETING FINANCIAL INTERESTS

The authors declare no competing financial interests.

References

  • 1.Crow YJ, Livingston JH. Aicardi-Goutières syndrome: an important Mendelian mimic of congenital infection. Dev. Med. Child Neurol. 2008;50:410–416. doi: 10.1111/j.1469-8749.2008.02062.x. [DOI] [PubMed] [Google Scholar]
  • 2.Lebon P, et al. Intrathecal synthesis of interferon-α in infants with progressive familial encephalopathy. J. Neurol. Sci. 1988;84:201–208. doi: 10.1016/0022-510x(88)90125-6. [DOI] [PubMed] [Google Scholar]
  • 3.Crow YJ, et al. Mutations in the gene encoding the 3′-5′ DNA exonuclease TREX1 cause Aicardi-Goutières syndrome at the AGS1 locus. Nat. Genet. 2006;38:917–920. doi: 10.1038/ng1845. [DOI] [PubMed] [Google Scholar]
  • 4.Crow YJ, et al. Mutations in genes encoding ribonuclease H2 subunits cause Aicardi-Goutières syndrome and mimic congenital viral brain infection. Nat. Genet. 2006;38:910–916. doi: 10.1038/ng1842. [DOI] [PubMed] [Google Scholar]
  • 5.Rice GI, et al. Mutations involved in Aicardi-Goutières syndrome implicate SAMHD1 as regulator of the innate immune response. Nat. Genet. 2009;41:829–832. doi: 10.1038/ng.373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Goldstone DC, et al. HIV-1 restriction factor SAMHD1 is a deoxynucleoside triphosphate triphosphohydrolase. Nature. 2011;480:379–382. doi: 10.1038/nature10623. [DOI] [PubMed] [Google Scholar]
  • 7.Rice G, et al. Clinical and molecular phenotype of Aicardi-Goutières syndrome. Am. J. Hum. Genet. 2007;81:713–725. doi: 10.1086/521373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Rice G, et al. Heterozygous mutations in TREX1 cause familial chilblain lupus and dominant Aicardi-Goutières syndrome. Am. J. Hum. Genet. 2007;80:811–815. doi: 10.1086/513443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Ramantani G, et al. Expanding the phenotypic spectrum of lupus erythematosus in Aicardi-Goutières syndrome. Arthritis Rheum. 2010;62:1469–1477. doi: 10.1002/art.27367. [DOI] [PubMed] [Google Scholar]
  • 10.Haaxma CA, et al. A de novo p.Asp18Asn mutation in TREX1 in a patient with Aicardi-Goutières syndrome. Am. J. Med. Genet. A. 2010;152A:2612–2617. doi: 10.1002/ajmg.a.33620. [DOI] [PubMed] [Google Scholar]
  • 11.Stetson DB, Ko JS, Heidmann T, Medzhitov R. Trex1 prevents cell-intrinsic initiation of autoimmunity. Cell. 2008;134:587–598. doi: 10.1016/j.cell.2008.06.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Gall A, et al. Autoimmunity initiates in nonhematopoietic cells and progresses via lymphocytes in an interferon-dependent autoimmune disease. Immunity. 2012;36:120–131. doi: 10.1016/j.immuni.2011.11.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Yan N, et al. The cytosolic exonuclease TREX1 inhibits the innate immune response to human immunodeficiency virus type 1. Nat. Immunol. 2010;11:1005–1013. doi: 10.1038/ni.1941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Laguette N, et al. SAMHD1 is the dendritic- and myeloid-cell-specific HIV-1 restriction factor counteracted by Vpx. Nature. 2011;474:654–657. doi: 10.1038/nature10117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Hrecka K, et al. Vpx relieves inhibition of HIV-1 infection of macrophages mediated by the SAMHD1 protein. Nature. 2011;474:658–661. doi: 10.1038/nature10195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Lahouassa H, et al. SAMHD1 restricts the replication of human immunodeficiency virus type 1 by depleting the intracellular pool of deoxynucleoside triphosphates. Nat. Immunol. 2012;13:223–228. doi: 10.1038/ni.2236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Genovesio A, et al. Automated genome-wide visual profiling of cellular proteins involved in HIV infection. J. Biomol. Screen. 2011;16:945–958. doi: 10.1177/1087057111415521. [DOI] [PubMed] [Google Scholar]
  • 18.Beck-Engeser GB, Eilat D, Wabl M. An autoimmune disease prevented by anti-retroviral drugs. Retrovirology. 2011;8:91. doi: 10.1186/1742-4690-8-91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Hartner JC, Walkley CR, Lu J, Orkin SH. ADAR1 is essential for the maintenance of hematopoiesis and suppression of interferon signaling. Nat. Immunol. 2009;10:109–115. doi: 10.1038/ni.1680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Crow YJ. Type I interferonopathies: a novel set of inborn errors of immunity. Ann. NY Acad. Sci. 2011;1238:91–98. doi: 10.1111/j.1749-6632.2011.06220.x. [DOI] [PubMed] [Google Scholar]
  • 21.Hogg M, Paro S, Keegan LP, O’Connell MA. RNA editing by mammalian ADARs. Adv. Genet. 2011;73:87–120. doi: 10.1016/B978-0-12-380860-8.00003-3. [DOI] [PubMed] [Google Scholar]
  • 22.George CX, Samuel CE. Human RNA-specific adenosine deaminase ADAR1 transcripts possess alternative exon 1 structures that initiate from different promoters, one constitutively active and the other interferon inducible. Proc. Natl. Acad. Sci. USA. 1999;96:4621–4626. doi: 10.1073/pnas.96.8.4621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Herbert A, et al. A Z-DNA binding domain present in the human editing enzyme, double-stranded RNA adenosine deaminase. Proc. Natl. Acad. Sci. USA. 1997;94:8421–8426. doi: 10.1073/pnas.94.16.8421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Miyamura Y, et al. Mutations of the RNA-specific adenosine deaminase gene (DSRAD) are involved in dyschromatosis symmetrica hereditaria. Am. J. Hum. Genet. 2003;73:693–699. doi: 10.1086/378209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Li M, et al. Mutational spectrum of the ADAR1 gene in dyschromatosis symmetrica hereditaria. Arch. Dermatol. Res. 2010;302:469–476. doi: 10.1007/s00403-010-1039-2. [DOI] [PubMed] [Google Scholar]
  • 26.Tojo K, et al. Dystonia, mental deterioration, and dyschromatosis symmetrica hereditaria in a family with ADAR1 mutation. Mov. Disord. 2006;21:1510–1513. doi: 10.1002/mds.21011. [DOI] [PubMed] [Google Scholar]
  • 27.Kondo T, et al. Dyschromatosis symmetrica hereditaria associated with neurological disorders. J. Dermatol. 2008;35:662–666. doi: 10.1111/j.1346-8138.2008.00540.x. [DOI] [PubMed] [Google Scholar]
  • 28.Heale BS, et al. Editing independent effects of ADARs on the miRNA/siRNA pathways. EMBO J. 2009;28:3145–3156. doi: 10.1038/emboj.2009.244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Cho DS, et al. Requirement of dimerization for RNA editing activity of adenosine deaminases acting on RNA. J. Biol. Chem. 2003;278:17093–17102. doi: 10.1074/jbc.M213127200. [DOI] [PubMed] [Google Scholar]
  • 30.Abdel-Salam GM, et al. Chilblains as a diagnostic sign of Aicardi-Goutières syndrome. Neuropediatrics. 2010;41:18–23. doi: 10.1055/s-0030-1255059. [DOI] [PubMed] [Google Scholar]
  • 31.Kondo T, et al. Six novel mutations of the ADAR1 gene in patients with dyschromatosis symmetrica hereditaria: histological observation and comparison of genotypes and clinical phenotypes. J. Dermatol. 2008;35:395–406. doi: 10.1111/j.1346-8138.2008.00493.x. [DOI] [PubMed] [Google Scholar]
  • 32.Levanon EY, et al. Systematic identification of abundant A-to-I editing sites in the human transcriptome. Nat. Biotechnol. 2004;22:1001–1005. doi: 10.1038/nbt996. [DOI] [PubMed] [Google Scholar]
  • 33.Athanasiadis A, Rich A, Maas S. Widespread A-to-I RNA editing of Alu-containing mRNAs in the human transcriptome. PLoS Biol. 2004;2:e391. doi: 10.1371/journal.pbio.0020391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Blow M, Futreal PA, Wooster R, Stratton MR. A survey of RNA editing in human brain. Genome Res. 2004;14:2379–2387. doi: 10.1101/gr.2951204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Kim DD, et al. Widespread RNA editing of embedded Alu elements in the human transcriptome. Genome Res. 2004;14:1719–1725. doi: 10.1101/gr.2855504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Hundley HA, Bass BL. ADAR editing in double-stranded UTRs and other noncoding RNA sequences. Trends Biochem. Sci. 2010;35:377–383. doi: 10.1016/j.tibs.2010.02.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Wu D, Lamm AT, Fire AZ. Competition between ADAR and RNAi pathways for an extensive class of RNA targets. Nat. Struct. Mol. Biol. 2011;18:1094–1101. doi: 10.1038/nsmb.2129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Yang W, et al. ADAR1 RNA deaminase limits short interfering RNA efficacy in mammalian cells. J. Biol. Chem. 2005;280:3946–3953. doi: 10.1074/jbc.M407876200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Vitali P, Scadden AD. Double-stranded RNAs containing multiple IU pairs are sufficient to suppress interferon induction and apoptosis. Nat. Struct. Mol. Biol. 2010;17:1043–1050. doi: 10.1038/nsmb.1864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Gallo A, Keegan LP, Ring GM, O’Connell MA. An ADAR that edits transcripts encoding ion channel subunits functions as a dimer. EMBO J. 2003;22:3421–3430. doi: 10.1093/emboj/cdg327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Yao Y, et al. Development of potential pharmacodynamic and diagnostic markers for anti–IFN-α monoclonal antibody trials in systemic lupus erythematosus. Hum. Genomics Proteomics. 2009;2009:374312. doi: 10.4061/2009/374312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Higgs BW, et al. Patients with systemic lupus erythematosus, myositis, rheumatoid arthritis and scleroderma share activation of a common type I interferon pathway. Ann. Rheum. Dis. 2011;70:2029–2036. doi: 10.1136/ard.2011.150326. [DOI] [PubMed] [Google Scholar]
  • 43.Macbeth MR, et al. Inositol hexakisphosphate is bound in the ADAR2 core and required for RNA editing. Science. 2005;309:1534–1539. doi: 10.1126/science.1113150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Larkin MA, et al. Clustal W and Clustal X version 2.0. Bioinformatics. 2007;23:2947–2948. doi: 10.1093/bioinformatics/btm404. [DOI] [PubMed] [Google Scholar]
  • 45.Sali A, Blundell TL. Comparative protein modelling by satisfaction of spatial restraints. J. Mol. Biol. 1993;234:779–815. doi: 10.1006/jmbi.1993.1626. [DOI] [PubMed] [Google Scholar]
  • 46.Chen VB, Davis IW, Richardson DC. KING (Kinemage, Next Generation): a versatile interactive molecular and scientific visualization program. Protein Sci. 2009;18:2403–2409. doi: 10.1002/pro.250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Word JM, Lovell SC, Richardson JS, Richardson DC. Asparagine and glutamine: using hydrogen atom contacts in the choice of side-chain amide orientation. J. Mol. Biol. 1999;285:1735–1747. doi: 10.1006/jmbi.1998.2401. [DOI] [PubMed] [Google Scholar]
  • 48.Word JM, et al. Visualizing and quantifying molecular goodness-of-fit: small-probe contact dots with explicit hydrogen atoms. J. Mol. Biol. 1999;285:1711–1733. doi: 10.1006/jmbi.1998.2400. [DOI] [PubMed] [Google Scholar]
  • 49.Lovell SC, Word JM, Richardson JS, Richardson DC. The penultimate rotamer library. Proteins. 2000;40:389–408. [PubMed] [Google Scholar]
  • 50.Losey HC, Ruthenburg AJ, Verdine GL. Crystal structure of Staphylococcus aureus tRNA adenosine deaminase TadA in complex with RNA. Nat. Struct. Mol. Biol. 2006;13:153–159. doi: 10.1038/nsmb1047. [DOI] [PubMed] [Google Scholar]
  • 51.Stefl R, et al. The solution structure of the ADAR2 dsRBM-RNA complex reveals a sequence-specific readout of the minor groove. Cell. 2010;143:225–237. doi: 10.1016/j.cell.2010.09.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Schwartz T, et al. Crystal structure of the Zα domain of the human editing enzyme ADAR1 bound to left-handed Z-DNA. Science. 1999;284:1841–1845. doi: 10.1126/science.284.5421.1841. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Vanderver_SD_Fig2
Vanderver_SD_Fig3-5
vanderver_SD_Fig1

RESOURCES