Open Access

Mutations of the cystic fibrosis transmembrane conductance regulator gene in males with congenital bilateral absence of the vas deferens: Reproductive implications and genetic counseling (Review)

  • Authors:
    • Xiangrong Cui
    • Xueqing Wu
    • Qiang Li
    • Xuan Jing
  • View Affiliations

  • Published online on: August 24, 2020     https://doi.org/10.3892/mmr.2020.11456
  • Pages: 3587-3596
  • Copyright: © Cui et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Congenital bilateral absence of the vas deferens (CBAVD) is predominantly caused by mutations in the cystic fibrosis transmembrane conductance regulator (CFTR) gene. CBAVD accounts for 2‑6% of male infertility cases and up to 25% of cases of obstructive azoospermia. With the use of pre‑implantation genetic diagnosis, testicular or epididymal sperm aspiration, intracytoplasmic sperm injection and in vitro fertilization, patients affected by CBAVD are able to have children who do not carry CFTR gene mutations, thereby preventing disease. Therefore, genetic counseling should be provided to couples receiving assisted reproductive techniques to discuss the impact of CFTR gene mutations on reproductive health. In the present article, the current literature concerning the CFTR gene and its association with CBAVD is reviewed.

Introduction

Congenital bilateral absence of the vas deferens (CBAVD) accounts for 2–6% of male infertility cases and up to 25% of cases of obstructive azoospermia (1,2). In addition to CBAVD, the clinical symptoms and characteristics of this condition generally include normal or slightly small testicles, atrophy or absence of seminal vesicle and cauda epididymis and normal serum follicle-stimulating hormone (3). The hallmarks of CBAVD include azoospermia, reduced semen volume (<1 ml), pH value (≤7.0) and seminal fructose, as well as decreased production of spermatozoa in the testicles (4).

CBAVD may occur as an isolated manifestation or as an atypical symptom of cystic fibrosis (CF), one of the most common autosomal recessive genetic disorders in the Caucasian population (5). CF has an incidence of ~1:2,500 and is usually caused by mutations in the CF transmembrane conductance regulator (CFTR) gene (5). Since almost 97% of male patients with CF are infertile due to CBAVD with resulting obstructive azoospermia, it may be hypothesized that isolated CBAVD and CF have a common genetic origin (1). The anatomical defects associated with CBAVD arise at the embryonic stage, and normal development of the male reproductive tract may require CFTR or CFTR-mediated anion secretion (3,6). In addition, 60–70% of patients with CBAVD carry pathogenic CFTR mutations (Online Mendelian Inheritance in Man no. 602421), usually one severe and one mild, in compound heterozygosity (7,8). However, the exact role of CFTR in male reproductive physiology remains to be fully elucidated. The present review summarized recent findings linking CFTR gene mutations to CBAVD and highlighted the assisted reproductive techniques (ART) and genetic counseling provided clear diagnosis and treatment options for such patients. Literature databases, such as Pubmed (pubmed.ncbi.nlm.nih.gov/) and Medline (de.medline.eu/), and Shanxi Medical University library (library.sxmu.edu.cn/) were searched for articles on CFTR and male fertility with relevant keywords such as cystic fibrosis transmembrane conductance regulator gene (CFTR), Congenital bilateral absence of the vas deferens (CBAVD), variation, pre-implantation genetic diagnosis (PGD), and assisted reproductive techniques (ART).

Overview of CFTR structure and function

The CFTR gene was first cloned and identified by Riordan et al (9) in 1989. In humans, it is located on chromosome 7q31.2 and spans a total length of 230 kb comprising 27 exons; the composition of the gene is outlined in Table I (10). CFTR encodes a protein containing 1,480 amino acid residues, with a molecular weight of 168,138 Da (11). The CFTR protein is an ATP-bound glycosylated transmembrane protein located in the apical membrane of several types of exocrine epithelial cells, which provides a selective channel for chloride ions across the epithelium (12). The structure of CFTR consists of five domains, including two membrane-spanning domains (MSD-1 and −2) together forming a selective channel, two nucleotide-binding domains (NBD-1 and −2) and a regulatory domain (R domain) (Fig. 1). Each MSD is composed of six transmembrane helices, of whichTM1-6 constitute MSD-1 and TM7-12 constitute MSD-2, followed by a cytoplasmic NBD (13). The MSD-2 and NBD-1 modules are connected by the R domain, a regulatory region with disordered structure that is unique to the CFTR gene and not found in other transporters (14). The two MSD domains form a selective chloride channel and the two NBDs regulate the gating of this channel (15). Furthermore, phosphorylation of the R domain controls channel activity (15). The R domain contains several phosphorylation sites for protein kinase A, a cyclic AMP-dependent kinase that is highly conserved between species (16).

Table I.

Description of the structure of the cystic fibrosis transmembrane conductance regulator gene.

Table I.

Description of the structure of the cystic fibrosis transmembrane conductance regulator gene.

A, Exons

NameHistorical nameStart, base positionEnd, base positionLength, basepairs
Exon 1Exon 1117,120,017117,120,201185
Exon 2Exon 2117,144,307117,144,417111
Exon 3Exon 3117,149,088117,149,196109
Exon 3Exon 3117,170,953117,171,168216
Exon 5Exon 5117,174,330117,174,41990
Exon 6Exon 6a117,175,302117,175,465164
Exon 7Exon 6b117,176,602117,176,727126
Exon 8Exon 7117,180,154117,180,400247
Exon 9Exon 8117,182,070117,182,16293
Exon 10Exon 9117,188,695117,188,877183
Exon 11Exon 10117,199,518117,199,709192
Exon 12Exon 11117,227,793117,227,88795
Exon 13Exon 12117,230,407117,230,49387
Exon 14Exon 13117,231,988117,232,711724
Exon 15Exon 14a117,234,984117,235,112129
Exon 16Exon 14b117,242,880117,242,91738
Exon 17Exon 15117,243,586117,243,836251
Exon 18Exon 16117,246,728117,246,80780
Exon 19Exon 17a117,250,573117,250,723151
Exon 20Exon 17b117,251,635117,251,862228
Exon 21Exon 18117,254,667117,254,767101
Exon 22Exon 19117,267,576117,267,824249
Exon 23Exon 20117,282,492117,282,647156
Exon 24Exon 21117,292,896117,292,98590
Exon 25Exon 22117,304,742117,304,914173
Exon 26Exon 23117,305,513117,305,618106
Exon 27Exon 24117,306,962117,308,7151,754

B, Introns

NameHistorical nameStartEndLength, basepairs

Intron 1–2  −117,120,202117,144,30624,105
Intron 2–3117,144,418117,149,0874,670
Intron 3–4117,149,197117,170,95221,756
Intron 4–5117,171,169117,174,3293,161
Intron 5–6117,174,420117,175,301882
Intron 6–7117,175,466117,176,6011,136
Intron 7–8117,176,728117,180,1533,426
Intron 8–9117,180,401117,182,0691,669
Intron 9–10117,182,163117,188,6946,532
Intron 10–11117,188,878117,199,5171,064
Intron 11–12117,199,710117,227,79228,083
Intron 12–13117,227,888117,230,4062,519
Intron 13–14117,230,494117,231,9871,494
Intron 14–15117,232,712117,234,9832,272
Intron 15–16117,235,113117,242,8797,767
Intron 16–17117,242,918117,243,585668
Intron 17–18117,243,837117,246,7272,891
Intron 18–19117,246,808117,250,5723,765
Intron 19–20117,250,724117,251,634911
Intron 20–21117,251,863117,254,6662,804
Intron 21–22117,254,768117,267,57512,808
Intron 22–23117,267,825117,282,49114,667
Intron 23–24117,282,648117,292,89510,248
Intron 24–25117,292,986117,304,74111,756
Intron 25–26117,304,915117,305,512598
Intron 26–27117,305,619117,306,9611,343

CFTR is expressed in epithelial tissues of various organs, including the pancreas, intestine, sweat glands and vas deferens, and participates in the secretion of substances and fluids and maintains the balance of electrolyte and homeostasis in the lumen (17). In the steady state, the chloride channel is blocked by anions or negative charges in the regulatory R domain. Phosphorylation of the R domain is required for channel activity (18). As the concentration of intracellular chloride ion increases, Protein Kinase A (PKA) is activated, causing the channel to open (19). Similarly, the extracellular chloride concentration also regulates channel gating, and increased concentrations of extracellular chloride promote the opening of channels (20). Thus, CFTR facilitates two-way permeability of chloride ions (20). Previous studies have suggested that, in addition to regulating chloride ion transport through epithelial cells, CFTR is involved in the regulation of various ion transporters, including epithelial sodium channels, chloride/bicarbonate exchangers, sodium/proton exchangers and water channels (2022). Therefore, CFTR-dependent physiological processes have a key role in maintaining homeostasis of pH, ions and water levels in the fluids of secretory epithelium (21).

The CFTR gene and development of the vas deferens

The majority of male patients with CF and CBAVD have primary infertility (22), suggesting that the CFTR protein has an important role in the development of the male reproductive tract (23). CFTR mutations among patients with CBAVD (58.1%) are common but not as frequent as in patients with unilateral absence of vas deferens (75%), and a normal amount of functional CFTR protein may be required to ensure the normal development of the vas deferens (24). Mutations in the CFTR gene impair the normal function of chloride channels and prevent them from regulating the flow of chloride ions and water across the cell membrane, resulting in the production of highly viscous mucus by epithelial cells of the male genital tract (25). This mucus, in turn, causes blockage of the vas deferens during embryonic development, causing it to denature or even deteriorate.

In 12–18 week-old abortive fetuses with CFTR gene mutations, the vas deferens is obstructed and denatured by mucus, and the vas deferens may be further aggravated in embryonic development (26). Furthermore, Gaillard et al (26) demonstrated that the expression of the CFTR in epididymal epithelium in human fetuses carrying CFTR mutations at 10–33 weeks of gestation was low, resulting in the production of highly viscous mucus by epithelial cells of the male genital tract. In addition, CFTR mutations are absent in cases of CBAVD with renal abnormalities, suggesting that the mesonephric duct has an important role in the development of the vas deferens (26). It has also been demonstrated that fluid secretion is necessary for the normal development of the mesonephric duct (25). Abnormal fluid secretion impairs normal development, leading to hypoplasia of the mesonephric duct and degeneration during early embryonic development (25,26). Therefore, CFTR may have an important role in the development of the vas deferens and mutations of CFTR may be associated with the occurrence of CBAVD.

The CFTR gene in patients with CBAVD

Numerous studies have investigated the genetic link between CFTR mutations and the risk of CBAVD (2730). A large number of CFTR mutant alleles have been found in patients with CBAVD. CFTR mutations display high heterogeneity in terms of spectrum and frequency (27,28). A total of 50,022 CFTR gene variants have been identified to date (National Center for Biotechnology Information; ncbi.nlm.nih.gov/variation/view/?q=1080%5Bgeneid%5D&assm=GCF_000001405.25), most minor allele frequency (MAF) of which are <0.01%. Only 605 variants are >1% and 347 variants are >5%. A total of 1.68% of CFTR gene variants are considered to be potentially pathogenic, while 1.16% are not known to cause functional changes and another 97.16% have not been characterized (27). In addition, 90.43% of the mutations involve sequences that are only a few bases long. Of these, missense mutations were the most common, accounting for 2.97% of all mutations, while gene rearrangement was rare, accounting for only 1.13% of mutations (27).

To date, >2,000 CF-causing CFTR mutations have been identified (Table II; genet.sickkids.on.ca/StatisticsPage.html). CFTR gene mutations are divided into six classes, according to the extent of CFTR protein downregulation (2830). Class I mutations refer to the production of non-functional protein or non-functional mRNA products that are degraded by nonsense-mediated mRNA decay (31). Class I mutations usually result from splice site mutations, nonsense mutations, frameshift mutations or large fragment insertion-deletion, the most common ones being these three types (p.G542X, p.R533X and p.W1282X) of missense mutations (31). Class II mutations result in defects of protein maturity (32). A common example is the p.F508del mutation, which leads to abnormal folding of the CFTR protein, resulting in partial function and low CFTR protein concentration (32). Class III mutations, such as p.G551D, refer to defects in chloride channel gating function despite normal protein structure, which are usually caused by impaired binding of the NBD domain with ATP and hydrolysis (33). Class IV mutations in the CFTR gene, such as p.R117, p.HR334W and p.R347P, lead to abnormal regulation of chloride channel function, usually due to abnormal structure of MSDs (34). With class V mutations, protein function is normal, but CFTR protein synthesis is reduced (35). For instance, the c.3717+12191 C>T mutation in intron 19 causes leakage in transcription of exon 19, resulting in a decrease in mRNA to only 5–10% of the normal mRNA levels (35).

Table II.

Mutations in the cystic fibrosis transmembrane conductance regulator.

Table II.

Mutations in the cystic fibrosis transmembrane conductance regulator.

Mutation typeNumber of known mutationsFrequency
Missense81038.87
Frameshift33616.12
Splicing22910.99
Nonsense1758.40
In-frame indel432.06
Large indel592.83
Promoter170.82
Sequence variation26912.91
Unknown1467.01

[i] Mutation frequency refers to the frequency at which mutations in the CFTR detected make up this locus in all of the DNA. Indel, insertion-deletion.

A novel class of mutations, Class VI, has recently been identified, which may be further divided into subclasses A and B (36). Class VI-A mutations lead to instability of CFTR protein, with a half-life reduced to only about one-fifth of normal proteins. Examples of Class VI-A mutation include p.Q1412X, c.4323delTC and c.4279insA, which result in the loss of 70–100 bp of mRNA and abnormal folding of the C-terminal end of the CFTR protein (36,37). Class VI-B mutations, such as the p.G551D missense mutation, usually affect the interaction between MSDs of the CFTR protein (38).

Class I, II and VI mutations are generally considered severe, with near-complete loss of gene function. The transcription level of the CFTR gene is <3%, which is easy to cause CF and high clinical mortality (31,32,36). By contrast, class IV and class V mutations are mild mutations with slightly higher residual function, accounting for 3–10% of all CFTR mutations, and lead to CF-associated diseases, such as p.R117H-T7, lead to CBAVD (36).

A meta-analysis by Yu et al (8) suggested that 78% of patients with CBAVD had at least one type of CFTR gene mutation, among which 48% had two simultaneous mutations. Most of these mutations in patients with CBAVD were point mutations (class IV or V) with a mild phenotype, with few (<1%) gene deletions or rearrangements. In addition, 15–25% of patients with CF displayed gene rearrangements (Fig. 2). The mutation frequency and genomic region of CFTR in which mutations occur vary by region and ethnicity. For instance, the CFTR mutation rate in Caucasian populations is significantly higher than in Asians. Furthermore, CF is the most common lethal autosomal recessive genetic disease in Caucasians (8,39). The most common mutant alleles in Caucasians with CBAVD are p.F508del, c.IVS9-10 T5 and p.R117H, with frequencies of 13–21, 22–29 and 2–4%, respectively (8,39).

A severe heterozygous mutation on one allele combined with a heterozygous mutation on the other (88%), or two mild heterozygous mutations (12%) usually lead to the occurrence of CBAVD. However, a severe homozygous mutation (88%) or two severe heterozygous mutations (12%) may lead to the occurrence of CF (40). This suggests that CFTR gene mutations are ethnically specific and their non-coding regions may lead to mutations in CBAVD disease, while these types of mutations do not lead to CF.

Hotspot mutation types in patients with CBAVD

M470V

The p.M470V mutation is an amino acid substitution caused by A to G polymorphism of the nucleotide at position 1,540 in exon 11 of the CFTR gene. It is a missense mutation and affects the chloride channel activity of CFTR (41). Residue 470 codes for part of the first NBD domain, and both M470 and V470 may lead to the synthesis of a completely glycosylated CFTR protein. However, the chloride channel activity of the M470 variant is twice as high as that of V470 (41,42). Residue 470 also interacts with other sites on the CFTR gene, thereby influencing the expression levels of the CFTR protein (43). The M470V mutation is frequently combined with the (TGm) and Tn polymorphisms. The expression of the V470 CFTR variant decreases with increasing numbers of TG repeats and shorter T repeats, leading to an increased risk of CBAVD (43).

The ratio of M470 to V470 varies among different populations. This ratio is normally ~1:1 in the healthy Caucasian population. However, the ratio in the normal and infertile population remains to be determined. Furthermore, in Chinese patients with CBAVD, the ratio of M470 to V470 is 0.7:1, in line with the Hardy-Weinberg equilibrium, and it is thought to be unrelated to the pathogenesis of CBAVD (44).

p.F508del

p.F508del refers to the deletion of three bases (CTT) at positions 1,653-1,655 in exon 11. The resulting deletion of residue F508 of the CFTR protein leads to abnormal protein folding and a significant reduction in CFTR protein levels on the cell membrane of exocrine gland epithelial cells. Failure to form effective chloride channels on the epithelium, in turn, results in severe impairment of small molecule transport (45). p.F508del is the most common mutation in Caucasian patients with CBAVD, with a mutation frequency ranging from 35 to 74.6% (45). However, the incidence of this mutation is low in Asians (46,47).

p.D1152H

p.D1152H mutation is an amino acid substitution caused by G to C polymorphism at position 3,586 in exon 21 of the CFTR gene. This mutation is a class IV mild mutation, resulting in partial CFTR protein function and mild clinical manifestations of CF disease. In males, the major clinical manifestation is CBAVD (48). The incidence of the p.D1152H mutation is 12% in Ashkenazi Jewish (49) and 6.27% in Hispanic and African American individuals (50), which included pre-marital and pre-natal screening in certain areas with a high incidence of CF (48).

p.R117H

p.R117H results from a G to A substitution at position 4,482 of exon 4; the resulting change from arginine to histidine is considered a mild mutation. However, pR117H combined with the intervening sequence (IVS) 9–10 T5 variant leads to severe clinical manifestations of CF (51). In addition, p.R117H combined with IVS9-10 T9 does not cause disease. Furthermore, when combined with IVS9-10 T7, this mutation frequently leads to CBAVD and results in CF when combined with IVS9-10 T5 (52). p.R117H has a high incidence (~3%) in Caucasian patients with CF.

Mutations in the promoter region

Promoters regulate gene transcription levels by interacting with trans-acting factors (17,53). The CFTR gene promoter is a housekeeping-type, GC-rich (<65% GC content) promoter lacking the typical TATA box, and its specific cis-acting element location and regulatory mechanism are poorly characterized (53). Yoshimura et al (54) suggested that the core sequence of the CFTR gene promoter may be located ~3.8 kb upstream of the ATG transcription initiation site and contains conserved sequences acting as binding sites for several important transcription proteins, such as CTCC-binding factor (CTCF), activator protein-1 (AP-1), specificity protein 1 (SP1), greater curvature (GRE), carbapenem-resistant (CRE), CCAAT/enhancer-binding protein (C/EBP) and Y-box proteins (55). Giordano et al (56) examined the CFTR gene promoter in patients with CF and identified 23 mutations. All of these were point mutations that affected transcription, resulting in lack of regulation, upregulation and downregulation of the CFTR gene (56). Lopez et al (67) demonstrated that the forkhead box I1 (FOXI1) transcription factor was able to downregulate CFTR gene transcription. Indeed, the c.-165G>A promoter mutation enhanced FOXI1 binding, thereby inhibiting transcription of the CFTR gene. Feng et al (53) recently demonstrated that the c.-195C>A promoter mutation significantly enhanced the inhibitory effect of specificity protein (SP) on CFTR expression.

Intron 9–10 (TG)m and Tn polymorphism

(TG)m and Tn polymorphism inIVS 9–10 of the CFTR gene are associated with the occurrence of CBAVD. Tn refers to the length of a noncoding thymidine sequence. The Tn sequence is usually 7 or 9 bases long (T7 and T9, respectively) in the normal population, while the incidence of 5T is 5.2% (57,58). The shorter the Tn sequence, the lower the transcription level of the CFTR gene.

In patients who are homozygous for the IVS9-10 5T (c.1210-7_1210-6delTT variant, formerly known as IVS8 5T), a sequence of 5 consecutive thymidines in intron 9 may cause alternative splicing of exon 10 mRNA, resulting in missing transcription of the CFTR gene. This, in turn, causes a reduction in the synthesis of CFTR protein with normal functions (only 10% of the normal protein); these patients with CBAVD do not exhibit other symptoms of CF (Fig. 3A) (30). Exon 10 encodes 60 amino acids of the CFTR protein NBD-1. Thus, a mutant CFTR protein lacking this amino acid sequence loses chloride channel activity, thus reducing the permeability of chloride to the epithelial cell membrane of the exocrine gland and increasing the reabsorption of sodium ions in the exocrine fluid, eventually leading to CF and associated diseases including chronic obstructive pulmonary disease, pancreatic exocrine dysfunction and meconium ileus (59).

Radpour et al (60) examined the distribution of Tn alleles in 106 patients with CBAVD and demonstrated that the incidence of T5 was 25.94%, although T5 distribution was not investigated among males with normal fertility. However, the incidence of Tn polymorphism in patients with CBAVD varies with ethnicity, ranging from 13 (Asian) to 43.7% (Caucasian) (61). Disset et al (62) identified a rare IVS9 T3 mutation in one case. Functional experiments demonstrated that this mutation led to severe cystic fibrosis (62), consistent with the hypothesis that leakage in exon 10 transcription becomes more serious with the decrease in length of the Tn sequence (59). The T5 variant is currently considered to be a mild CFTR gene mutation rather than a polymorphism.

Furthermore, the (TG)m (c.1210-34) polymorphism also affects the transcription of exon 10 (Fig. 3B). In association with the T7 variant, loss of exon 10 transcription caused by TG11 and TG12 is 2.8 and 6.0 times more likely than with TG10, respectively (63). Thus, (TG)m and Tn synergistically affect the transcription of exon 10. Indeed, the greater the number of TG repeats, and the shorter the number of T repeats and the higher the probability of transcription loss of exon 10 is (63). IVS9-10 (TG)m Tn is a common mutation type of the CFTR gene. Furthermore, the frequency of this mutation in Chinese patients with CBAVD is slightly higher than that in Caucasian patients (44).

Intron 10–11 (TAAA)n short-tandem repeats

Short-tandem repeats are widely distributed in the genome and the length of repeats is closely associated with the occurrence of diseases, such as neuromuscular disorders (64) and gynecomastia (65). Previous studies have also demonstrated that short-tandem repeats in promoters and introns may affect gene transcription (6466). A large number of dimer and tetramer short-tandem repeats have been identified in the noncoding region of the CFTR gene that affect its expression. (TAAA)n is a short-tandem repeat composed of four bases in IVS10-11 that may impair the transcription of exon 10. However, it does not affect the transcription of exon 10 as Tn in intron 9–10 does. Lopez et al (67) suggested that FOXI1 transcription factor binds to IVS10-11 (TAAA)n short-tandem repeats to negatively regulate the expression of the CFTR gene. With decreasing numbers of repeats, binding of FOXI1 transcription factor and promoter was enhanced, thus reducing the transcriptional levels of the CFTR gene. Furthermore, combination with the c.-165G>A mild promoter mutation or IVS9-10 (TG)mTn may lead to the occurrence of CBAVD (67).

Although the CFTR gene was identified >20 years ago, the relationship between its genotype and clinical phenotype remains controversial, mostly due to the wide variety of CFTR gene mutations and differences across ethnicities. Furthermore, other factors may also be involved, including environmental factors, viruses or bacteria (27,50,61).

In addition, it has been observed that 6–15% of patients with CBAVD did not carry any CFTR gene mutations (1), further highlighting possible limits of detection during diagnosis or the possibility that other genetic mutations are present in certain patients.

Assisted reproductive techniques and genetic counselling

Compared with patients with CF, patients CBAVD are more likely to have mild CFTR gene mutations (class IV and V) associated with CFTR protein dysfunction. Most patients with CBAVD have normal spermatogenic function, and their spouses may be impregnated via testicular tissue sperm extraction (TESE) or intracytoplasmic sperm injection (ICSI), a procedure, through which a single sperm is injected into the cytoplasm of a mature oocyte to obtain a viable embryo (68). The first pregnancy for a couple in whom the male had CBAVD was reported in 1987 (68). A recent meta-analysis suggested that the results of ICSI were not affected by the source of sperm (fresh, frozen, epididymal or testicular), but a lower rate of fertilization and a higher rate of miscarriage were observed in patients with CBAVD as compared with those with acquired aspermia (69).

Due to possible inheritance of the CFTR mutation and risk of CF occurrence, genetic counseling prior to the use of assisted reproductive technology (ART) is particularly important, especially for patients with a CBAVD familial history. To date, CFTR gene mutation detection has been applied in pre-implantation genetic diagnosis (PGD) in the USA and certain European countries such as the Netherlands (70), Russia (71) and France (72). CFTR mutation screening as part of a PGD screening program may select embryos carrying a healthy CFTR gene, which has certain clinical significance for couples affected by CBAVD and their potential children. Liu et al (70) reported the first successful PGD in a couple affected by CBAVD, in which both partners were heterozygotes with p.F508del mutations. In this instance, three carrier embryos were implanted and a healthy male infant was born. The presence of CF or CBAVD-associated mutations in the male partner does not currently appear to significantly affect in vitro oocyte fertilization, embryo implantation rates or successful delivery of asymptomatic offspring after PGD (73,74).

For a couple affected by CBAVD with CFTR deficiency planning to have their own children, the risk for both male and female offspring to have CF or associated diseases, and for male offspring to have CBAVD depends on whether the female partner is a carrier, as one mutant allele is always inherited from the male. If the female does not carry the CFTR gene mutation, the probability of CBAVD in the male offspring is low, estimated to be <1%. The inheritance of CBAVD is more complex than that of CF, as genetic analysis may include but not exclude the diagnosis of reproductive CF. Furthermore, when rare mutations are detected in males or females, the risk of CF or CBAVD in the offspring may be unpredictable. The couple should be informed that tests cannot detect all mutations in the gene; therefore, negative mutation screening may reduce but not eliminate the risk of being a carrier.

Clinical examination and follow-up of children born to couples affected by CBAVD are of great significance, as it may provide information regarding the variability in phenotypes associated with CFTR mutations. Identifying a CFTR mutation in a male affected by CBAVD is important not only for the patient himself, but also for his family. Healthy siblings of a male child born to a parent with CBAVD have a 50% chance of being carriers of CF. Therefore, both the patient and their partner should be screened for CFTR gene mutations. If no mutation is detected in the male partner, it is not certain that the unborn child will be free of CF, and the risk may be <1:1,000 (27). Among male infertile heterosexual couples not affected by CF, the risk of CF in the offspring was no different from that of couples in the general population. The risk of infertility in their male offspring remains elusive. In addition, for all couples with positive results, PGD and prenatal diagnosis (PND) are recommended, together with genetic counseling (Fig. 4).

Conclusions

CBAVD is an atypical manifestation of CF caused by mutations in the CFTR gene, which frequently leads to male infertility. The present review summarized recent findings linking the CFTR gene to male fertility and highlighted the progress made in understanding the role of CFTR in reproductive events pertinent to male fertility. Genetic counseling for couples affected by CBAVD remains challenging. Characterizing the mutation spectrum of the CFTR gene and its association with the pathogenesis of CBAVD may prove conducive to disease diagnosis and genetic risk assessment prior to IVF, thereby reducing the incidence of this disease or the rate of CFTR mutations in infants resulting from ART. Consistent guidelines are also required concerning the extent of mutational analysis for the CFTR gene during the diagnosis of CBAVD, as well as options available for couples affected by CBAVD in association with PND and PGD.

Acknowledgements

Not applicable.

Funding

The current study was supported by the Scientific Research Project of Shanxi Provincial Department of Health (grant no. 201601070), the Initial Scientific Research Fund of PhD at Shanxi Provincial People's Hospital (grant no. b201635), the Natural Science Foundation of Shanxi (grant nos. 201901D211519 and 201901D211546),the Research Project Supported by Shanxi Scholarship Council of China (grant no. HGKY2019092) and China Postdoctoral Science Foundation (grant no. 2020M670703).

Availability of data and materials

Data sharing is not applicable to this article, as no datasets were generated or analyzed during the present study.

Authors' contributions

XC and XJ designed the study. XW and QL interpreted the data. All authors wrote, read and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Radpour R, Gourabi H, Dizaj AV, Holzgreve W and Zhong XY: Genetic investigations of CFTR mutations in congenital absence of vas deferens, uterus, and vagina as a cause of infertility. J Androl. 29:506–513. 2008. View Article : Google Scholar : PubMed/NCBI

2 

Gajbhiye R, Kadam K, Khole A, Gaikwad A, Kadam S, Shah R, Kumaraswamy R and Khole V: Cystic fibrosis transmembrane conductance regulator (CFTR) gene abnormalities in Indian males with congenital bilateral absence of vas deferens & renal anomalies. Indian J Med Res. 143:616–623. 2016. View Article : Google Scholar : PubMed/NCBI

3 

Li CY, Jiang LY, Chen WY, Li K, Sheng HQ, Ni Y, Lu JX, Xu WX, Zhang SY and Shi QX: CFTR is essential for sperm fertilizing capacity and is correlated with sperm quality in humans. Hum Reprod. 25:317–327. 2010. View Article : Google Scholar : PubMed/NCBI

4 

Ferlin A and Stuppia L: Diagnostics of CFTR-negative patients with congenital bilateral absence of vas deferens: Which mutations are of most interest? Expert Rev Mol Diagn. 20:265–267. 2020. View Article : Google Scholar : PubMed/NCBI

5 

Gallego A, Rogel R, Perez-Ardavin J, Lorenzo L, Lujan S, Oltra S, Molina I and Broseta E: Congenital bilateral absence of the vas deferens (CBAVD): Do genetic disorders modify assisted reproductive technologies outcomes? Arch Esp Urol. 72:1038–1042. 2019.(In Spanish). PubMed/NCBI

6 

Diao R, Fok KL, Zhao L, Chen H, Tang H, Chen J, Zheng A, Zhang X, Gui Y, Chan HC and Cai Z: Decreased expression of cystic fibrosis transmembrane conductance regulator impairs sperm quality in aged men. Reproduction. 146:637–645. 2013. View Article : Google Scholar : PubMed/NCBI

7 

Mak V, Zielenski J, Tsui LC, Durie P, Zini A, Martin S, Longley TB and Jarvi KA: Proportion of cystic fibrosis gene mutations not detected by routine testing in men with obstructive azoospermia. JAMA. 281:2217–2224. 1999. View Article : Google Scholar : PubMed/NCBI

8 

Yu J, Chen Z, Ni Y and Li Z: CFTR mutations in men with congenital bilateral absence of the vas deferens (CBAVD): A systemic review and meta-analysis. Hum Reprod. 27:25–35. 2012. View Article : Google Scholar : PubMed/NCBI

9 

Riordan JR, Rommens JM, Kerem B, Alon N, Rozmahel R, Grzelczak Z, Zielenski J, Lok S, Plavsic N, Chou JL, et al: Identification of the cystic fibrosis gene: Cloning and characterization of complementary DNA. Science. 245:1066–1073. 1989. View Article : Google Scholar : PubMed/NCBI

10 

Tsui LC and Dorfman R: The cystic fibrosis gene: A molecular genetic perspective. Cold Spring Harb Perspect Med. 3:a0094722013. View Article : Google Scholar : PubMed/NCBI

11 

Guillot L, Beucher J, Tabary O, Le Rouzic P, Clement A and Corvol H: Lung disease modifier genes in cystic fibrosis. Int J Biochem Cell Biol. 52:83–93. 2014. View Article : Google Scholar : PubMed/NCBI

12 

Linsdell P: Cystic fibrosis transmembrane conductance regulator chloride channel blockers: Pharmacological, biophysical and physiological relevance. World J Biol Chem. 5:26–39. 2014. View Article : Google Scholar : PubMed/NCBI

13 

Poroca DR, Amer N, Li A, Hanrahan JW and Chappe VM: Changes in the R-region interactions depend on phosphorylation and contribute to PKA and PKC regulation of the cystic fibrosis transmembrane conductance regulator chloride channel. FASEB Bioadv. 2:33–48. 2019. View Article : Google Scholar : PubMed/NCBI

14 

Jarosz-Griffiths HH, Scambler T, Wong CH, Lara-Reyna S, Holbrook J, Martinon F, Savic S, Whitaker P, Etherington C, Spoletini G, et al: Different CFTR modulator combinations downregulate inflammation differently in cystic fibrosis. Elife. 9:e545562020. View Article : Google Scholar : PubMed/NCBI

15 

Jaworska J, Marach-Mocarska A and Sands D: Uncommon clinical presentation of cystic fibrosis in a patient homozygous for a rare CFTR mutation: A case report. BMC Pediatr. 20:902020. View Article : Google Scholar : PubMed/NCBI

16 

Chen JH: Protein kinase A phosphorylation potentiates cystic fibrosis transmembrane conductance regulator gating by relieving autoinhibition on the stimulatory C terminus of the regulatory domain. J Biol Chem. 295:4577–4590. 2020. View Article : Google Scholar : PubMed/NCBI

17 

NandyMazumdar M, Yin S, Paranjapye A, Kerschner JL, Swahn H, Ge A, Leir SH and Harris A: Looping of upstream cis-regulatory elements is required for CFTR expression in human airway epithelial cells. Nucleic Acids Res. 48:3513–3524. 2020. View Article : Google Scholar : PubMed/NCBI

18 

Laselva O, Stone TA, Bear CE and Deber CM: Anti-infectives restore ORKAMBI((R)) rescue of F508del-CFTR function in human bronchial epithelial cells infected with clinical strains of P. aeruginosa. Biomolecules. 10:3342020. View Article : Google Scholar

19 

McCarron A, Cmielewski P, Reyne N, McIntyre C, Finnie J, Craig F, Rout-Pitt N, Delhove J, Schjenken JE, Chan HY, et al: Phenotypic characterization and comparison of cystic fibrosis Rat models generated using CRISPR/Cas9 gene editing. Am J Pathol. 190:977–993. 2020. View Article : Google Scholar : PubMed/NCBI

20 

Froux L, Elbahnsi A, Boucherle B, Billet A, Baatallah N, Hoffmann B, Alliot J, Zelli R, Zeinyeh W, Haudecoeur R, et al: Targeting different binding sites in the CFTR structures allows to synergistically potentiate channel activity. Eur J Med Chem. 190:1121162020. View Article : Google Scholar : PubMed/NCBI

21 

Choi JY, Muallem D, Kiselyov K, Lee MG, Thomas PJ and Muallem S: Aberrant CFTR-dependent HCO3- transport in mutations associated with cystic fibrosis. Nature. 410:94–97. 2001. View Article : Google Scholar : PubMed/NCBI

22 

Van Mourik P, van Haaren P, Kruisselbrink E, Korkmaz C, Janssens HM, de Winter-de Groot KM, van der Ent CK, Hagemeijer MC and Beekman JM: R117H-CFTR function and response to VX-770 correlate with mRNA and protein expression in intestinal organoids. J Cyst Fibros. 2020.(Epub ahead of print). View Article : Google Scholar

23 

De Santi C, Fernandez Fernandez E, Gaul R, Vencken S, Glasgow A, Oglesby IK, Hurley K, Hawkins F, Mitash N, Mu F, et al: Precise targeting of miRNA sites restores CFTR activity in CF bronchial epithelial cells. Mol Ther. 28:1190–1199. 2020. View Article : Google Scholar : PubMed/NCBI

24 

Jain R, Middleton PG and Rowe SM: Triple therapy for cystic fibrosis with a phe508del CFTR mutation. Reply. N Engl J Med. 382:6842020. View Article : Google Scholar : PubMed/NCBI

25 

Morris-Rosendahl DJ, Edwards M, McDonnell MJ, John S, Alton EWFW, Davies JC and Simmonds NJ: Whole gene sequencing of CFTR reveals a high prevalence of the intronic variant c.3874-4522A>G in cystic fibrosis. Am J Respir Crit Care Med. 201:1438–1441. 2020. View Article : Google Scholar : PubMed/NCBI

26 

Gaillard DA, Carre-Pigeon F and Lallemand A: Normal vas deferens in fetuses with cystic fibrosis. J Urol. 158:1549–1552. 1997. View Article : Google Scholar : PubMed/NCBI

27 

Ferec C and Cutting GR: Assessing the disease-liability of mutations in CFTR. Cold Spring Harb Perspect Med. 2:a0094802012. View Article : Google Scholar : PubMed/NCBI

28 

Tamburino L, Guglielmino A, Venti E and Chamayou S: Molecular analysis of mutations and polymorphisms in the CFTR gene in male infertility. Reprod Biomed Online. 17:27–35. 2008. View Article : Google Scholar : PubMed/NCBI

29 

Patel B, Parets S, Akana M, Kellogg G, Jansen M, Chang C, Cai Y, Fox R, Niknazar M, Shraga R, et al: Comprehensive genetic testing for female and male infertility using next-generation sequencing. J Assist Reprod Genet. 35:1489–1496. 2018. View Article : Google Scholar : PubMed/NCBI

30 

Cuppens H and Cassiman JJ: CFTR mutations and polymorphisms in male infertility. Int J Androl. 27:251–256. 2004. View Article : Google Scholar : PubMed/NCBI

31 

Hinzpeter A, Aissat A, Sondo E, Costa C, Arous N, Gameiro C, Martin N, Tarze A, Weiss L, de Becdelièvre A, et al: Alternative splicing at a NAGNAG acceptor site as a novel phenotype modifier. PLoS Genet. 6:e10011532010. View Article : Google Scholar : PubMed/NCBI

32 

Kopito RR: Biosynthesis and degradation of CFTR. Physiol Rev. 79 (1 Suppl):S167–S173. 1999. View Article : Google Scholar : PubMed/NCBI

33 

Schwiebert EM, Egan ME, Hwang TH, Fulmer SB, Allen SS, Cutting GR and Guggino WB: CFTR regulates outwardly rectifying chloride channels through an autocrine mechanism involving ATP. Cell. 81:1063–1073. 1995. View Article : Google Scholar : PubMed/NCBI

34 

Seibert FS, Linsdell P, Loo TW, Hanrahan JW, Riordan JR and Clarke DM: Cytoplasmic loop three of cystic fibrosis transmembrane conductance regulator contributes to regulation of chloride channel activity. J Biol Chem. 271:27493–27499. 1996. View Article : Google Scholar : PubMed/NCBI

35 

Highsmith WE, Burch LH, Zhou Z, Olsen JC, Boat TE, Spock A, Gorvoy JD, Quittel L, Friedman KJ, Silverman LM, et al: A novel mutation in the cystic fibrosis gene in patients with pulmonary disease but normal sweat chloride concentrations. N Engl J Med. 331:974–980. 1994. View Article : Google Scholar : PubMed/NCBI

36 

Green DM, McDougal KE, Blackman SM, Sosnay PR, Henderson LB, Naughton KM, Collaco JM and Cutting GR: Mutations that permit residual CFTR function delay acquisition of multiple respiratory pathogens in CF patients. Respir Res. 11:1402010. View Article : Google Scholar : PubMed/NCBI

37 

Haardt M, Benharouga M, Lechardeur D, Kartner N and Lukacs GL: C-terminal truncations destabilize the cystic fibrosis transmembrane conductance regulator without impairing its biogenesis. A novel class of mutation. J Biol Chem. 274:21873–21877. 1999. View Article : Google Scholar : PubMed/NCBI

38 

Fulmer SB, Schwiebert EM, Morales MM, Guggino WB and Cutting GR: Two cystic fibrosis transmembrane conductance regulator mutations have different effects on both pulmonary phenotype and regulation of outwardly rectified chloride currents. Proc Natl Acad Sci USA. 92:6832–6836. 1995. View Article : Google Scholar : PubMed/NCBI

39 

Bombieri C, Claustres M, De Boeck K, Derichs N, Dodge J, Girodon E, Sermet I, Schwarz M, Tzetis M, Wilschanski M, et al: Recommendations for the classification of diseases as CFTR-related disorders. J Cyst Fibros. 10 (Suppl 2):S86–S102. 2011. View Article : Google Scholar : PubMed/NCBI

40 

Claustres M, Guittard C, Bozon D, Chevalier F, Verlingue C, Ferec C, Girodon E, Cazeneuve C, Bienvenu T, Lalau G, et al: Spectrum of CFTR mutations in cystic fibrosis and in congenital absence of the vas deferens in France. Hum Mutat. 16:143–156. 2000. View Article : Google Scholar : PubMed/NCBI

41 

Cuppens H, Lin W, Jaspers M, Costes B, Teng H, Vankeerberghen A, Jorissen M, Droogmans G, Reynaert I, Goossens M, et al: Polyvariant mutant cystic fibrosis transmembrane conductance regulator genes. The polymorphic (Tg)m locus explains the partial penetrance of the T5 polymorphism as a disease mutation. J Clin Invest. 101:487–496. 1998. View Article : Google Scholar : PubMed/NCBI

42 

Kosova G, Pickrell JK, Kelley JL, McArdle PF, Shuldiner AR, Abney M and Ober C: The CFTR Met 470 allele is associated with lower birth rates in fertile men from a population isolate. PLoS Genet. 6:e10009742010. View Article : Google Scholar : PubMed/NCBI

43 

Pompei F, Ciminelli BM, Bombieri C, Ciccacci C, Koudova M, Giorgi S, Belpinati F, Begnini A, Cerny M, Des Georges M, et al: Haplotype block structure study of the CFTR gene. Most variants are associated with the M470 allele in several European populations. Eur J Hum Genet. 14:85–93. 2006. View Article : Google Scholar : PubMed/NCBI

44 

Du Q, Li Z, Pan Y, Liu X, Pan B and Wu B: The CFTR M470V, intron 8 poly-T, and 8 TG-repeats detection in Chinese males with congenital bilateral absence of the vas deferens. Biomed Res Int. 2014:6891852014. View Article : Google Scholar : PubMed/NCBI

45 

Kumar S, Tana A and Shankar A: Cystic fibrosis--what are the prospects for a cure? Eur J Intern Med. 25:803–807. 2014. View Article : Google Scholar : PubMed/NCBI

46 

Leonardi S, Pratico AD, Rotolo N, Di Dio G, Lionetti E and La Rosa M: Early acute pancreatitis in a child with compound heterozygosis F508/R1438W/Y1032C cystic fibrosis: A case report. J Med Case Rep. 7:1882013. View Article : Google Scholar : PubMed/NCBI

47 

Wu CC, Hsieh-Li HM, Lin YM and Chiang HS: Cystic fibrosis transmembrane conductance regulator gene screening and clinical correlation in Taiwanese males with congenital bilateral absence of the vas deferens. Hum Reprod. 19:250–253. 2004. View Article : Google Scholar : PubMed/NCBI

48 

Zvereff VV, Faruki H, Edwards M and Friedman KJ: Cystic fibrosis carrier screening in a North American population. Genet Med. 16:539–546. 2014. View Article : Google Scholar : PubMed/NCBI

49 

Kornreich R, Ekstein J, Edelmann L and Desnick RJ: Premarital and prenatal screening for cystic fibrosis: Experience in the Ashkenazi Jewish population. Genet Med. 6:415–420. 2004. View Article : Google Scholar : PubMed/NCBI

50 

Sugarman EA, Rohlfs EM, Silverman LM and Allitto BA: CFTR mutation distribution among U.S. Hispanic and African American individuals: Evaluation in cystic fibrosis patient and carrier screening populations. Genet Med. 6:392–399. 2004. View Article : Google Scholar : PubMed/NCBI

51 

Massie RJ, Poplawski N, Wilcken B, Goldblatt J, Byrnes C and Robertson C: Intron-8 polythymidine sequence in Australasian individuals with CF mutations R117H and R117C. Eur Respir J. 17:1195–1200. 2001. View Article : Google Scholar : PubMed/NCBI

52 

Kiesewetter S, Macek M Jr, Davis C, Curristin SM, Chu CS, Graham C, Shrimpton AE, Cashman SM, Tsui LC, Mickle J, et al: A mutation in CFTR produces different phenotypes depending on chromosomal background. Nat Genet. 5:274–278. 1993. View Article : Google Scholar : PubMed/NCBI

53 

Feng J, Wu X and Zhang Y, Yang X, Ma G, Chen S, Luo S and Zhang Y: A novel mutation (−195C>A) in the promoter region of CFTR gene is associated with Chinese congenital bilateral absence of vas deferens (CBAVD). Gene. 719:1440072019. View Article : Google Scholar : PubMed/NCBI

54 

Yoshimura K, Nakamura H, Trapnell BC, Dalemans W, Pavirani A, Lecocq JP and Crystal RG: The cystic fibrosis gene has a ‘housekeeping’-type promoter and is expressed at low levels in cells of epithelial origin. J Biol Chem. 266:9140–9144. 1991.PubMed/NCBI

55 

McCarthy VA and Harris A: The CFTR gene and regulation of its expression. Pediatr Pulmonol. 40:1–8. 2005. View Article : Google Scholar : PubMed/NCBI

56 

Giordano S, Amato F, Elce A, Monti M, Iannone C, Pucci P, Seia M, Angioni A, Zarrilli F, Castaldo G and Tomaiuolo R: Molecular and functional analysis of the large 5′promoter region of CFTR gene revealed pathogenic mutations in CF and CFTR-related disorders. J Mol Diagn. 15:331–340. 2013. View Article : Google Scholar : PubMed/NCBI

57 

Chillon M, Casals T, Mercier B, Bassas L, Lissens W, Silber S, Romey MC, Ruiz-Romero J, Verlingue C, Claustres M, et al: Mutations in the cystic fibrosis gene in patients with congenital absence of the vas deferens. N Engl J Med. 332:1475–1480. 1995. View Article : Google Scholar : PubMed/NCBI

58 

Mak V, Jarvi KA, Zielenski J, Durie P and Tsui LC: Higher proportion of intact exon 9 CFTR mRNA in nasal epithelium compared with vas deferens. Hum Mol Genet. 6:2099–2107. 1997. View Article : Google Scholar : PubMed/NCBI

59 

Hwang TC and Sheppard DN: Gating of the CFTR Cl- channel by ATP-driven nucleotide-binding domain dimerisation. J Physiol. 587:2151–2161. 2009. View Article : Google Scholar : PubMed/NCBI

60 

Radpour R, Gilani MA, Gourabi H, Dizaj AV and Mollamohamadi S: Molecular analysis of the IVS8-T splice variant 5T and M470V exon 10 missense polymorphism in Iranian males with congenital bilateral absence of the vas deferens. Mol Hum Reprod. 12:469–473. 2006. View Article : Google Scholar : PubMed/NCBI

61 

Anzai C, Morokawa N, Okada H, Kamidono S, Eto Y and Yoshimura K: CFTR gene mutations in Japanese individuals with congenital bilateral absence of the vas deferens. J Cyst Fibros. 2:14–18. 2003. View Article : Google Scholar : PubMed/NCBI

62 

Disset A, Michot C, Harris A, Buratti E, Claustres M and Tuffery-Giraud S: A T3 allele in the CFTR gene exacerbates exon 9 skipping in vas deferens and epididymal cell lines and is associated with congenital bilateral absence of vas deferens (CBAVD). Hum Mutat. 25:72–81. 2005. View Article : Google Scholar : PubMed/NCBI

63 

Nam MH, Hijikata M, Tuan LA, Lien LT, Shojima J, Horie T, Nakata K, Matsushita I, Ohashi J, Tokunaga K and Keicho N: Variations of the CFTR gene in the Hanoi-Vietnamese. Am J Med Genet A. 136:249–253. 2005. View Article : Google Scholar : PubMed/NCBI

64 

Ranum LP and Cooper TA: RNA-mediated neuromuscular disorders. Annu Rev Neurosci. 29:259–277. 2006. View Article : Google Scholar : PubMed/NCBI

65 

Czajka-Oraniec I, Zgliczynski W, Kurylowicz A, Mikula M and Ostrowski J: Association between gynecomastia and aromatase (CYP19) polymorphisms. Eur J Endocrinol. 158:721–727. 2008. View Article : Google Scholar : PubMed/NCBI

66 

Martin P, Makepeace K, Hill SA, Hood DW and Moxon ER: Microsatellite instability regulates transcription factor binding and gene expression. Proc Natl Acad Sci USA. 102:3800–3804. 2005. View Article : Google Scholar : PubMed/NCBI

67 

Lopez E, Viart V, Guittard C, Templin C, René C, Méchin D, Des Georges M, Claustres M, Romey-Chatelain MC and Taulan M: Variants in CFTR untranslated regions are associated with congenital bilateral absence of the vas deferens. J Med Genet. 48:152–159. 2011. View Article : Google Scholar : PubMed/NCBI

68 

Silber SJ, Balmaceda J, Borrero C, Ord T and Asch R: Pregnancy with sperm aspiration from the proximal head of the epididymis: A new treatment for congenital absence of the vas deferens. Fertil Steril. 50:525–528. 1988. View Article : Google Scholar : PubMed/NCBI

69 

Nicopoullos JD, Gilling-Smith C, Almeida PA and Ramsay JW: The results of 154 ICSI cycles using surgically retrieved sperm from azoospermic men. Hum Reprod. 19:579–585. 2004. View Article : Google Scholar : PubMed/NCBI

70 

Liu J, Lissens W, Silber SJ, Devroey P, Liebaers I and Van Steirteghem A: Birth after preimplantation diagnosis of the cystic fibrosis delta F508 mutation by polymerase chain reaction in human embryos resulting from intracytoplasmic sperm injection with epididymal sperm. JAMA. 272:1858–1860. 1994. View Article : Google Scholar : PubMed/NCBI

71 

Rechitsky S, Verlinsky O and Kuliev A: PGD for cystic fibrosis patients and couples at risk of an additional genetic disorder combined with 24-chromosome aneuploidy testing. Reprod Biomed Online. 26:420–430. 2013. View Article : Google Scholar : PubMed/NCBI

72 

Girardet A, Ishmukhametova A, Willems M, Coubes C, Hamamah S, Anahory T, Des Georges M and Claustres M: Preimplantation genetic diagnosis for cystic fibrosis: The montpellier center's 10-year experience. Clin Genet. 87:124–132. 2015. View Article : Google Scholar : PubMed/NCBI

73 

McCallum TJ, Milunsky JM, Cunningham DL, Harris DH, Maher TA and Oates RD: Fertility in men with cystic fibrosis: An update on current surgical practices and outcomes. Chest. 118:1059–1062. 2000. View Article : Google Scholar : PubMed/NCBI

74 

Chiang HS, Wu CC, Wu YN, Lu JF, Lin GH and Hwang JL: CFTR mutation analysis of a Caucasian father with congenital bilateral absence of vas deferens, a Taiwanese mother, and twins resulting from ICSI procedure. J Formos Med Assoc. 107:736–740. 2008. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

November-2020
Volume 22 Issue 5

Print ISSN: 1791-2997
Online ISSN:1791-3004

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Cui X, Wu X, Li Q and Jing X: Mutations of the cystic fibrosis transmembrane conductance regulator gene in males with congenital bilateral absence of the vas deferens: Reproductive implications and genetic counseling (Review). Mol Med Rep 22: 3587-3596, 2020
APA
Cui, X., Wu, X., Li, Q., & Jing, X. (2020). Mutations of the cystic fibrosis transmembrane conductance regulator gene in males with congenital bilateral absence of the vas deferens: Reproductive implications and genetic counseling (Review). Molecular Medicine Reports, 22, 3587-3596. https://doi.org/10.3892/mmr.2020.11456
MLA
Cui, X., Wu, X., Li, Q., Jing, X."Mutations of the cystic fibrosis transmembrane conductance regulator gene in males with congenital bilateral absence of the vas deferens: Reproductive implications and genetic counseling (Review)". Molecular Medicine Reports 22.5 (2020): 3587-3596.
Chicago
Cui, X., Wu, X., Li, Q., Jing, X."Mutations of the cystic fibrosis transmembrane conductance regulator gene in males with congenital bilateral absence of the vas deferens: Reproductive implications and genetic counseling (Review)". Molecular Medicine Reports 22, no. 5 (2020): 3587-3596. https://doi.org/10.3892/mmr.2020.11456