Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Novel mutations in COL4A3, COL4A4, and COL4A5 in Chinese patients with Alport Syndrome

  • Jian-Hong Liu ,

    Contributed equally to this work with: Jian-Hong Liu, Xiu-Xiu Wei, Ang Li

    Affiliation Institute of Clinical Laboratory Science, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  • Xiu-Xiu Wei ,

    Contributed equally to this work with: Jian-Hong Liu, Xiu-Xiu Wei, Ang Li

    Affiliation Binhai Genomics Institute, BGI-Tianjin, BGI-shenzhen, Tianjin, China

  • Ang Li ,

    Contributed equally to this work with: Jian-Hong Liu, Xiu-Xiu Wei, Ang Li

    Affiliation Institute of Clinical Laboratory Science, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  • Ying-Xia Cui,

    Affiliation Institute of Clinical Laboratory Science, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  • Xin-Yi Xia,

    Affiliation Institute of Clinical Laboratory Science, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  • Wei-Song Qin,

    Affiliation National Clinical Research Center of Kidney Diseases, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  • Ming-Chao Zhang,

    Affiliation National Clinical Research Center of Kidney Diseases, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  • Er-Zhi Gao,

    Affiliation National Clinical Research Center of Kidney Diseases, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  • Jun Sun,

    Affiliation Binhai Genomics Institute, BGI-Tianjin, BGI-shenzhen, Tianjin, China

  • Chun-Lin Gao,

    Affiliation Department of Pediatric Nephrology, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  • Feng-Xia Liu,

    Affiliation Binhai Genomics Institute, BGI-Tianjin, BGI-shenzhen, Tianjin, China

  • Qiu-Yue Wu,

    Affiliation Institute of Clinical Laboratory Science, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  • Wei-Wei Li,

    Affiliation Institute of Clinical Laboratory Science, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  • Asan ,

    xiaojunli62@126.com (XJL); asan@genomics.cn (A); zhliunj@vip.163.com (ZHL)

    Affiliation Binhai Genomics Institute, BGI-Tianjin, BGI-shenzhen, Tianjin, China

  • Zhi-Hong Liu ,

    xiaojunli62@126.com (XJL); asan@genomics.cn (A); zhliunj@vip.163.com (ZHL)

    Affiliation National Clinical Research Center of Kidney Diseases, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China

  •  [ ... ],
  • Xiao-Jun Li

    xiaojunli62@126.com (XJL); asan@genomics.cn (A); zhliunj@vip.163.com (ZHL)

    Affiliations Institute of Clinical Laboratory Science, Jinling Hospital, Nanjing University School of Medicine, Nanjing, China, State Key Laboratory of Analytical Chemistry for Life Science, Department of Chemistry, Nanjing University, Nanjing, China

  • [ view all ]
  • [ view less ]

Abstract

Alport syndrome (AS) is a clinically and genetically heterogeneous, progressive nephropathy caused by mutations in COL4A3, COL4A4, and COL4A5, which encode type IV collagen. The large sizes of these genes and the absence of mutation hot spots have complicated mutational analysis by routine polymerase chain reaction (PCR)-based approaches. Here, in order to design a rapid and effective method for the genetic diagnosis of AS, we developed a strategy by utilizing targeted capture associated with next-generation sequencing (NGS) to analyze COL4A3, COL4A4, and COL4A5 simultaneously in 20 AS patients. All the coding exons and flanking sequences of COL4A3, COL4A4, and COL4A5 from the probands were captured followed by HiSeq 2500 sequencing. Candidate mutations were validated by classic Sanger sequencing and quantitative (q)PCR. Sixteen patients (16/20, 75%) showed X-linked inheritance, and four patients (4/20, 20%) showed autosomal recessive inheritance. None of the individuals had autosomal-dominant AS. Fifteen novel mutations, 6 known mutations, and 2 novel fragment deletions were detected by targeted capture and NGS. Of these novel mutations, 12, 3, and 2 mutations were detected in COL4A5, COL4A4, and COL4A3, respectively. A comparison of the clinical manifestations caused by different types of mutations in COL4A5 suggested that nonsense mutations and glycine substitution by an acidic amino acid are more severe than the other missense mutations. Pathogenic mutations were detected in 20 patients. These novel mutations can expand the genotypic spectrum of AS. Our results demonstrated that targeted capture and NGS technology are effective in the genetic diagnosis of AS.

Introduction

Alport Syndrome (AS) is a type of inherited nephropathy characterized by hematuria, proteinuria, and progressive renal failure, often associated with extrarenal manifestations such as sensorineural hypoacusis and ocular abnormalities [1]. The pathogenesis of AS is genetically heterogeneous and is caused by mutations in the genes encoding type IV collagen [2, 3]. The α3, α4 and α5 chains of type IV collagen are encoded by COL4A3, COL4A4, and COL4A5, respectively [4]. Type IV collagen is a major constituent of the glomerular basement membrane (GBM), and is composed of six kinds of homologous α chains, which assemble into three different heterotrimers (α1α1α2, α3α4α5, and α5α5α6) [3]. Among them, the α3α4α5 heterotrimer is essential for the structure and function of the basement membrane in the glomeruli of the kidney, cochlea, and eye [3, 5].

The majority (about 85%) of AS patients show an X-linked dominant inheritance pattern (XLAS, OMIM no. 301050) due to mutations in COL4A5, which is located in the Xq22 region [6]; while the other AS patients (about 15%) show autosomal recessive inheritance (ARAS, OMIM no. 104200) and autosomal dominant inheritance (ADAS OMIM no.203780) caused by mutations in COL4A3 or COL4A4 located in 2q36.3 [711]. COL4A3, COL4A4, and COL4A5 are large genes comprising 52, 48, and 53 exons, respectively. Thus far, 225 mutations in COL4A3, 170 mutations in COL4A4, and 909 mutations in COL4A5 have been described (The Human Gene Mutation Database at the Institute of Medical Genetics in Cardiff, 2016.4. http://www.hgmd.cf.ac.uk/ac/gene.php?gene=COL4A3; http://www.hgmd.cf.ac.uk/ac/gene.php?gene=COL4A4; http://www.hgmd.cf.ac.uk/ac/gene.php?gene=COL4A5). These mutations are spread throughout the genes without any identified mutational hot spots.

Clinically, AS is heterogeneous, and patients always present a wide variability of clinical manifestations: the rate of progression to end-stage renal disease (ESRD) and the presence or absence of sensorineural deafness and ocular changes [12] depends on the mutation they carry [1315]. In the case of XLAS, hemizygous males exhibit more severe symptoms than heterozygous female patients and usually reach ESRD before the age of 30 [13]. However, the female patients have more variable symptoms, from isolated hematuria to ESRD [16]. Individuals homozygous for COL4A3 and COL4A4 genes resulting in ARAS have similar clinical symptoms and prognosis as males showing XLAS, reaching ESRD in the first or second decade of life [17]. However, early diagnosis and therapy can improve the prognosis. Life expectancy can be increased via early effective and safe therapy for patients with AS [4]. The clinical phenotypes of heterozygous mutations in COL4A3 and COL4A4 are heterogeneous, and can cause autosomal dominant AS (ADAS), thin basement membrane nephropathy (TBMN), focal segmental glomerulosclerosis (FSGS) and benign familial hematuria (BFH) [18, 19]. The early clinical manifestations of these diseases are similar, and therefore, genetic classification is critical. However, the large size of the genes and the absence of mutational hot spots have significantly hindered mutational analysis with Sanger sequencing. Recent advances in targeted gene capture associated with next-generation sequencing (NGS) and validated bioinformatics tools have facilitated research and diagnostics in the field of nephrogenetics [20]. Therefore, in this study, we developed a strategy to analyze COL4A3, COL4A4, and COL4A5 simultaneously using targeted capture and NGS in 20 patients with AS.

Materials and methods

Patients

Twenty patients (18 males and two females) from unrelated Chinese families were elected from 400 patients from 2011 to 2014 diagnosed with AS according to standard criteria, and were enrolled in this study [21]. All these patients met Gregory’s criteria of AS, including family history of nephritis, persistent hematuria and/or proteinuria, immunohistochemical evidence of complete or partial lack of epitope of type IV collagen in glomerular or epidermal basement membranes or both, widespread GBM ultrastructural abnormalities, and presence or absence of ear and eye abnormalities. A brief clinical summary of the patients is shown in Table 1.

thumbnail
Table 1. Clinical and pathological features of all the patients.

https://doi.org/10.1371/journal.pone.0177685.t001

Peripheral blood samples were collected from all the participants. This study was approved by the Ethics Committee of Jinling Hospital and written informed consent was obtained from patients and both parents.

Targeted capture and next-generation sequencing

A custom capture array (NimbleGen, Roche, USA) was designed to capture all exons (1,625), splice sites, and the flanking intron sequences of 78 genes known to be associated with genetic kidney disease, including AS. The total size of the target regions of the capture array was 0.5 Mb. The pipeline described in previous studies [22, 23] was followed to capture and enrich the targeted sequences, prepare the sequencing library, and for NGS. Genomic DNA was extracted from peripheral blood using the QIAamp DNA BloodMiNi Kit (Qiagen, Hilden, Germany) and then sheared into fragments ranging from 200–300 bp using an ultrasonoscope (Covaris S2, Massachusetts, USA). Oligonucleotide adapters from Illumina (single reads) were ligated to the fragments. After the ligation was complete, successful adapter ligation was confirmed by four-cycle PCR using a high-fidelity polymerase with PCR primers containing a custom-synthesized barcode sequence (8 bp) as a sample index signature. PCR was used to generate a library for further analysis, and the DNA adapter-ligated and indexed fragments from 10 libraries were pooled and hybridized to customized oligonucleotide probes. After hybridization of the sequencing primer, base incorporation was carried out on an Illumina Hiseq2500 platform (Illumina, San Diego, CA), following the manufacturer’s standard cluster generation and sequencing protocols, for 90 cycles of sequencing per read, to generate paired-end reads including 90 bps at each end and 8 bps of the index tag by BGI, Tianjin, China.

Validation of candidate mutations by Sanger sequencing and quantitative (q)PCR

To validate the variations identified with NGS, the corresponding gene regions surrounding the variations were amplified by PCR and sequenced by Sanger sequencing. Purified PCR products were sequenced using the same primers and PCR conditions in both directions on an ABI3730xl DNA sequencer and were analyzed with the sequencer software (Sequencing Analysis 5.2). The two fragment deletions were identified by comparing the normalized sequencing depth of each exon in the same batch. Exons with a depth ratio of specific sample to others > 1.4 were considered to have duplication, while those with values <0.6 were considered to have fragment deletion. The cut-off value for the Z-score was 2.58. qPCR was used to validate the fragment deletions using Step One Plus (Applied Biosystems, CA, USA). The primers used for the amplification of COL4A3, COL4A4, and COL4A5 are listed in S1 Table.

Function prediction and alignment analysis of missense variations

Image analyses, error estimation, and base calling were processed by using the Illumina pipeline (version 1.3.4) after the entire run was completed to generate primary data. Indexed primers were used to identify the different reads from different samples in the primary data, and only reads that were perfectly matched to the theoretical adapter indexed sequences and those that matched the theoretical primer indexed sequences with a maximum of three mismatches were considered acceptable reads. Then, we removed a few unqualified sequences from the primary data using a local dynamic programming algorithm, which included low quality reads, defined as reads that contained >10% Ns in the read length, 50% reads with a quality value of <5 and with an average quality of <10, and adapter sequences including indexed sequences. The remaining sequences were termed as clean reads for further analysis. Clean reads were aligned to the reference human genomic sequence (NCBI37/gh19) of COL4A3 (NM_000091.4), COL4A4 (NM_000092.4), and COL4A5 (NM_033380.2) by using the BWA software package (Burrows Wheeler Aligner) [24]. SNPs and INDELs were identified via the SOAPsnp software and GATK IndelGenotyper (http://www.broadinstitute.org/gsa/wiki/index.php/), respectively [25]. The control database used in the pipeline included the 1000 genome database (http://www.1000genomes.org), dbSNP database, and a BGI in-house database, which included 2,087 normal subjects. Phylop (phyloP46wayPlacental) was used to calculate the conservation of each SNP. The functional impact of missense variation was analyzed with the PolyPhen-2 (Polymorphism Phenotyping v2.2.5) [26] and SIFT algorithm (Sorting Intolerant From Tolerant v5.1) [27]. According to the criteria of the American College of Medical Genetics, we divided the mutation sites into the following five categories: pathogenic, likely pathogenic, benign, likely benign and uncertain significance [28].

Results

Identification of candidate mutations in COL4A3, COL4A4, and COL4A5

A total of 272 variants and 2 fragment deletions were identified in COL4A3, COL4A4, and COL4A5 from the 20 patients diagnosed with AS using targeted capture and NGS. Among these variants, 90.8% (247/272) were termed as “polymorphisms” with high frequency (>0.01 in control databases). We identified 16 missense mutations, one nonsense mutation, three frameshift mutations, and one 5′-untranslated region (UTR) mutation. Among these, six were known mutations that had been reported previously and 15 were novel ones (Table 2); 66.67% (14/21) were new amino acid substitutions of glycine and 9.52% (2/21) were new amino acid substitutions of proline. These candidate mutations were validated by Sanger sequencing. The frequencies and prediction scores of these mutations are listed in Table 2.

thumbnail
Table 2. Mutations detected in COL4A3, COL4A4, and COL4A5.

https://doi.org/10.1371/journal.pone.0177685.t002

In AS patient IID16, in addition to a glycine substitution (G572A) in COL4A4 in the heterozygous state, there was another heterozygous mutation c.-23T>G in the 5′-UTR of this gene. This heterozygous mutation c.-23T>G was inherited by his two sons (S1 Fig, III1, and III2) who presented persistent isolated hematuria and normal renal function. The clinical features of these two isolated hematuria patients accorded to the diagnosis of familial benign hematuria.

Identification of fragment deletions in COL4A5

IID3 and IID10 carried fragment deletions in COL4A5 which were validated by qPCR (Fig 1). IID10 had no sensorineural hearing loss and the mildest glomerular injury. However, IID10 carried a heterozygous c.3946G>A (p.G1316S) mutation in COL4A3 (S2 Fig & Table 2) and a hemizygous deletion of exon 44 COL4A5. These two mutations were inherited from each of his parents, who were healthy (S2 Fig). A hemizygous deletion of exon 29 was identified in IID3, and the deletion was inherited from his mother, who had hematuria. The sequencing depths of COL4A5 exons of patients IID3 and IID10 are listed in S2 Table; the PCR quantification results are shown in Fig 1.

thumbnail
Fig 1. The PCR quantification detected in IID3 and IID10.

A, the PCR quantification results of IID3. The comparison of quantification of exon 29 of COL4A5 between patient IID3 and male controls. RQ, real-time quantitative PCR. B, the PCR quantification results of IID10. The comparison of quantification of exon 44 of COL4A5 between patient IID10 and male controls. RQ, real-time quantitative PCR.

https://doi.org/10.1371/journal.pone.0177685.g001

Discussion

Of the 20 AS patients, 16 (16/20, 75%) showed X-linked inheritance (mutations in COL4A5) and four (4/20, 20%) showed autosomal recessive inheritance (mutations in COL4A3 and COL4A4). This ratio is similar to that reported in the literature [4]. In our study, we did not find any individuals with ADAS caused by mutations in COL4A3 or COL4A4. Twenty-one mutations and two fragment deletions were identified by targeted capture and NGS. Sixteen missense mutations, one nonsense mutation, three frameshift mutations, and one 5′-UTR mutation were identified. Among these, 15 were novel mutations and six were known mutations that had been reported previously. According to the criteria of the American College of Medical Genetics [28], the 5′-UTR mutation is an uncertain significant mutation, while the others are pathogenic mutations. Our results showed eight glycine substitutions and two proline substitutions in COL4A5, two glycine substitutions in COL4A4, and four glycine substitutions in COL4A3. Analyzing the correlations of phenotype and genotype, we found that the clinical phenotypes caused by glycine substitution by an acidic amino acid seem more severe than by a neutral amino acid in COL4A5. Patients IID1 (with p.G1060E), IID4 (with p.G675D) and IID15 (with p.G878E) presented relatively more severe symptoms, such as grosser hematuria or proteinuria, higher level of serum creatinine, more rapid progression of disease and with extra renal manifestations (Table 1) than other patients IID7 (with p.G1229S), IID11 (with p.G669S), IID12 (with p.G1229S) and IID13 (with p.G1170V). Similar severity was observed in the clinical phenotypes caused by truncate mutations in COL4A5. All male patients IID6 (with p.R373*), IID18 (with p.S1371*) and IID20 (with p.P809Wfs*9) presented extra renal manifestations (Table 1). In addition, IID6 and IID18 had gross hematuria or proteinuria and high-level serum creatinine (Table 1). IID20 exhibited modest clinical manifestations (Table 1), possibly because of his relatively young age. However, as the sample number is not big enough to make a meaningful conclusion, it is necessary to enlarge the sample size for getting further verification in future.

The patient (IID10) with double gene mutations (p.G1316S in COL4A3 and a hemizygous deletion in COL4A5) which were inherited from each of his healthy parents (S2 Fig) had only moderate proteinuria, microscopic hematuria, normal renal function, and no extra renal manifestations. However, Mencarelli et al reported that double heterozygotes have more severe phenotypes, such as progression to ESRD by age 44 and sensorineural hearing loss, than individuals with heterozygous mutations in COL4A5 (in women) or COL4A4 [33]. As the modest clinical symptoms of IID10, the phenotype of this patient could be caused mainly by the COL4A5 deletion while COL4A3 mutation, as an incidental finding, probably played an insignificant role in the patient. However, considering his young age, further follow-up is necessitated for his clinical symptoms.

The individuals with heterozygous mutations of COL4A3 or COL4A4 exhibit a broad range of phenotypes. Recent studies have shown that about 40% patients with FBH [34], 10% of familial FSGS patients [35], and <5% AS patients carry heterozygous mutations in COL4A3 or COL4A4 [18]. In our study, compound heterozygous mutations p.G997E and p.G1167R in COL4A3 were identified in the ARAS patient IID2. Interestingly, Xie et al reported p.G997E heterozygous mutations in an FSGS patient, and the pathological manifestations showed typical focal segmental glomerulosclerosis [29]. However, the patient IID2, whose renal biopsy showed lamellation changes of the GBM and significantly decreased expression of the α5 chain in GBM, was diagnosed with ARAS. A similar situation was observed in the family of patient IID16. The proband carried two mutations in COL4A4 (p.G572A and c. -23T>G) and was diagnosed with AS. Both his sons (S1 Fig, III1, and III2) with heterozygous mutations (c. -23T>G) inherited from their father, presented isolated hematuria and accorded to the clinical features of FBH. As the mutations in COL4A3 or COL4A4 are heterozygous, half of the α3 or α4 chains in the α3α4α5 heterotrimers of type IV collagen have a normal structure. Therefore, carriers of heterozygous mutations presented mild phenotype clinically. Thus, these carriers are different from the asymptomatic heterozygous carriers of some recessive metabolic diseases such as phenylketonuria. On the basis of these two cases, we proposed that FBH, AS and partial FSGS are type IV collagen-related disorders, and patients who carry a single heterozygous mutation or compound heterozygous mutations may present independent diseases.

Genetic testing is of great and increasing importance for diagnosing AS, and it has some advantages over conventional methods. Compared with renal biopsy, using DNA extracted from peripheral blood for genetic analysis is much less invasive for patients, especially for children. The sequencing results are independent of the stage of disease and the age of the proband. In addition, genetic testing is helpful in clinical diagnosis, especially in sporadic patients or patients whose renal biopsy data not available.

In conclusion, based on targeted capture and NGS, 17 new mutations in genes encoding type IV collagen were identified in 20 AS patients which will be added to the mutation spectrums of COL4A3, COL4A4 and COL4A5 related to AS. Therefore, combining genetic test with clinical and pathological phenotype, comprehensive diagnosis of AS can be performed at the individual, tissue and molecular levels.

Supporting information

S1 Fig. Pedigree structure of the IID16 family.

Squares indicate males and circles indicate females. White symbols indicate individuals without clinical features of the AS disease. Filled gray symbols denote individuals with isolated hematuria or slightly abnormal urine analysis. Filled black symbols denote individuals with hematuria and/or proteinuria, renal failure enven urinaemia. IID, individual ID. The arrows indicate the proband of the family. The horizontal line stand for marriage and double horizontal line stands for consanguineous marriage.

https://doi.org/10.1371/journal.pone.0177685.s001

(TIF)

S2 Fig. Pedigree structure of the IID3 and IID10 family.

https://doi.org/10.1371/journal.pone.0177685.s002

(TIF)

S2 Table. The value of sequencing depth of COL4A5 exons of patient IID3 and patient IID10.

https://doi.org/10.1371/journal.pone.0177685.s004

(DOC)

Acknowledgments

We thank the patients and their families for permitting us to do this study. This work is supported by the grants from Science and Technology Program of Jiangsu Province (BL2014072), Science and Technology Development Project of Nanjing (No. 201402008), (No. 201503010), and National Key Clinical Program of China (No. 2014ZDZK003).

Author Contributions

  1. Conceptualization: XJL ZHL A.
  2. Data curation: JHL XXW AL YXC.
  3. Formal analysis: YXC XYX.
  4. Funding acquisition: XJL YXC XYX.
  5. Investigation: AL WSQ MCZ.
  6. Methodology: XJL EZG.
  7. Project administration: JS CLG.
  8. Resources: XXW FXL.
  9. Software: JHL QYW.
  10. Supervision: XJL YXC ZHL.
  11. Validation: JHL XXW.
  12. Visualization: JHL AL WWL.
  13. Writing – original draft: JHL AL.
  14. Writing – review & editing: YXC AL.

References

  1. 1. Alport AC. Hereditary Familial Congenital Haemorrhagic Nephritis. Br Med J. 1927;1(3454):504–6. pmid:20773074
  2. 2. Kashtan CE, Michael AF. Alport syndrome: from bedside to genome to bedside. American journal of kidney diseases: the official journal of the National Kidney Foundation. 1993;22(5):627–40.
  3. 3. Hudson BG. The molecular basis of Goodpasture and Alport syndromes: beacons for the discovery of the collagen IV family. Journal of the American Society of Nephrology: JASN. 2004;15(10):2514–27. pmid:15466256
  4. 4. Kruegel J, Rubel D, Gross O. Alport syndrome—insights from basic and clinical research. Nature reviews Nephrology. 2013;9(3):170–8. pmid:23165304
  5. 5. Khoshnoodi J, Pedchenko V, Hudson BG. Mammalian collagen IV. Microsc Res Tech. 2008;71(5):357–70. pmid:18219669
  6. 6. Pirson Y. Making the diagnosis of Alport's syndrome. Kidney international. 1999;56(2):760–75. pmid:10432421
  7. 7. Longo I, Porcedda P, Mari F, Giachino D, Meloni I, Deplano C, et al. COL4A3/COL4A4 mutations: from familial hematuria to autosomal-dominant or recessive Alport syndrome. Kidney international. 2002;61(6):1947–56. pmid:12028435
  8. 8. Marcocci E, Uliana V, Bruttini M, Artuso R, Silengo MC, Zerial M, et al. Autosomal dominant Alport syndrome: molecular analysis of the COL4A4 gene and clinical outcome. Nephrology, dialysis, transplantation: official publication of the European Dialysis and Transplant Association—European Renal Association. 2009;24(5):1464–71.
  9. 9. Nagel M, Nagorka S, Gross O. Novel COL4A5, COL4A4, and COL4A3 mutations in Alport syndrome. Hum Mutat. 2005;26(1):60.
  10. 10. Savige J, Gregory M, Gross O, Kashtan C, Ding J, Flinter F. Expert guidelines for the management of Alport syndrome and thin basement membrane nephropathy. Journal of the American Society of Nephrology: JASN. 2013;24(3):364–75. pmid:23349312
  11. 11. Jefferson JA, Lemmink HH, Hughes AE, Hill CM, Smeets HJ, Doherty CC, et al. Autosomal dominant Alport syndrome linked to the type IV collage alpha 3 and alpha 4 genes (COL4A3 and COL4A4). Nephrology, dialysis, transplantation: official publication of the European Dialysis and Transplant Association—European Renal Association. 1997;12(8):1595–9.
  12. 12. Wei G, Zhihong L, Huiping C, Caihong Z, Zhaohong C, Leishi L. Spectrum of clinical features and type IV collagen alpha-chain distribution in Chinese patients with Alport syndrome. Nephrology, dialysis, transplantation: official publication of the European Dialysis and Transplant Association—European Renal Association. 2006;21(11):3146–54.
  13. 13. Jais JP, Knebelmann B, Giatras I, De Marchi M, Rizzoni G, Renieri A, et al. X-linked Alport syndrome: natural history and genotype-phenotype correlations in girls and women belonging to 195 families: a "European Community Alport Syndrome Concerted Action" study. Journal of the American Society of Nephrology: JASN. 2003;14(10):2603–10. pmid:14514738
  14. 14. Gross O, Netzer KO, Lambrecht R, Seibold S, Weber M. Meta-analysis of genotype-phenotype correlation in X-linked Alport syndrome: impact on clinical counselling. Nephrology, dialysis, transplantation: official publication of the European Dialysis and Transplant Association—European Renal Association. 2002;17(7):1218–27.
  15. 15. Bekheirnia MR, Reed B, Gregory MC, McFann K, Shamshirsaz AA, Masoumi A, et al. Genotype-phenotype correlation in X-linked Alport syndrome. Journal of the American Society of Nephrology: JASN. 2010;21(5):876–83. pmid:20378821
  16. 16. Kashtan CE. Alport syndromes: phenotypic heterogeneity of progressive hereditary nephritis. Pediatric nephrology. 2000;14(6):502–12. pmid:10872195
  17. 17. Mochizuki T, Lemmink HH, Mariyama M, Antignac C, Gubler MC, Pirson Y, et al. Identification of mutations in the alpha 3(IV) and alpha 4(IV) collagen genes in autosomal recessive Alport syndrome. Nature genetics. 1994;8(1):77–81. pmid:7987396
  18. 18. Kamiyoshi N, Nozu K, Fu XJ, Morisada N, Nozu Y, Ye MJ, et al. Genetic, Clinical, and Pathologic Backgrounds of Patients with Autosomal Dominant Alport Syndrome. Clinical journal of the American Society of Nephrology: CJASN. 2016;11(8):1441–9. pmid:27281700
  19. 19. Heidet L, Arrondel C, Forestier L, Cohen-Solal L, Mollet G, Gutierrez B, et al. Structure of the human type IV collagen gene COL4A3 and mutations in autosomal Alport syndrome. Journal of the American Society of Nephrology: JASN. 2001;12(1):97–106. pmid:11134255
  20. 20. Renkema KY, Stokman MF, Giles RH, Knoers NV. Next-generation sequencing for research and diagnostics in kidney disease. Nature reviews Nephrology. 2014;10(8):433–44. pmid:24914583
  21. 21. Gregory MC, Terreros DA, Barker DF, Fain PN, Denison JC, Atkin CL. Alport syndrome—clinical phenotypes, incidence, and pathology. Contrib Nephrol. 1996;117:1–28. pmid:8801040
  22. 22. Liu G, Wei X, Chen R, Zhou H, Li X, Sun Y, et al. A novel mutation of the SLC25A13 gene in a Chinese patient with citrin deficiency detected by target next-generation sequencing. Gene. 2014;533(2):547–53. pmid:24161253
  23. 23. Wei X, Ju X, Yi X, Zhu Q, Qu N, Liu T, et al. Identification of sequence variants in genetic disease-causing genes using targeted next-generation sequencing. PLoS One. 2011;6(12):e29500. pmid:22216297
  24. 24. Li H, Durbin R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics. 2009;25(14):1754–60. pmid:19451168
  25. 25. McKenna A, Hanna M, Banks E, Sivachenko A, Cibulskis K, Kernytsky A, et al. The Genome Analysis Toolkit: a MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 2010;20(9):1297–303. pmid:20644199
  26. 26. Adzhubei IA, Schmidt S, Peshkin L, Ramensky VE, Gerasimova A, Bork P, et al. A method and server for predicting damaging missense mutations. Nat Methods. 2010;7(4):248–9. pmid:20354512
  27. 27. Ng PC, Henikoff S. Predicting deleterious amino acid substitutions. Genome Res. 2001;11(5):863–74. pmid:11337480
  28. 28. Richards S, Aziz N, Bale S, Bick D, Das S, Gastier-Foster J, et al. Standards and guidelines for the interpretation of sequence variants: a joint consensus recommendation of the American College of Medical Genetics and Genomics and the Association for Molecular Pathology. Genet Med. 2015;17(5):405–24. pmid:25741868
  29. 29. Xie J, Wu X, Ren H, Wang W, Wang Z, Pan X, et al. COL4A3 mutations cause focal segmental glomerulosclerosis. J Mol Cell Biol. 2015;7(2):184. pmid:25888712
  30. 30. Pochet JM, Bobrie G, Landais P, Goldfarb B, Grunfeld JP. Renal prognosis in Alport's and related syndromes: influence of the mode of inheritance. Nephrology, dialysis, transplantation: official publication of the European Dialysis and Transplant Association—European Renal Association. 1989;4(12):1016–21.
  31. 31. Beirowski B, Weber M, Gross O. Chronic renal failure and shortened lifespan in COL4A3+/- mice: an animal model for thin basement membrane nephropathy. Journal of the American Society of Nephrology: JASN. 2006;17(7):1986–94. pmid:16775036
  32. 32. Longo I, Scala E, Mari F, Caselli R, Pescucci C, Mencarelli MA, et al. Autosomal recessive Alport syndrome: an in-depth clinical and molecular analysis of five families. Nephrology, dialysis, transplantation: official publication of the European Dialysis and Transplant Association—European Renal Association. 2006;21(3):665–71.
  33. 33. Mencarelli MA, Heidet L, Storey H, van Geel M, Knebelmann B, Fallerini C, et al. Evidence of digenic inheritance in Alport syndrome. J Med Genet. 2015;52(3):163–74. pmid:25575550
  34. 34. Papazachariou L, Demosthenous P, Pieri M, Papagregoriou G, Savva I, Stavrou C, et al. Frequency of COL4A3/COL4A4 mutations amongst families segregating glomerular microscopic hematuria and evidence for activation of the unfolded protein response. Focal and segmental glomerulosclerosis is a frequent development during ageing. PLoS One. 2014;9(12):e115015. pmid:25514610
  35. 35. Malone AF, Phelan PJ, Hall G, Cetincelik U, Homstad A, Alonso AS, et al. Rare hereditary COL4A3/COL4A4 variants may be mistaken for familial focal segmental glomerulosclerosis. Kidney international. 2014;86(6):1253–9. pmid:25229338