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

TLR9 -1486T/C and 2848C/T SNPs Are Associated with Human Cytomegalovirus Infection in Infants

  • Edyta Paradowska ,

    eparadowska@cbm.pan.pl

    Affiliation Laboratory of Molecular Virology and Biological Chemistry, Institute of Medical Biology, Polish Academy of Sciences, Lodz, Poland

  • Agnieszka Jabłońska,

    Affiliation Laboratory of Molecular Virology and Biological Chemistry, Institute of Medical Biology, Polish Academy of Sciences, Lodz, Poland

  • Mirosława Studzińska,

    Affiliation Laboratory of Molecular Virology and Biological Chemistry, Institute of Medical Biology, Polish Academy of Sciences, Lodz, Poland

  • Katarzyna Skowrońska,

    Current address: Department of Neurotoxicology, Mossakowski Medical Research Centre, Polish Academy of Sciences, Warsaw, Poland

    Affiliation Laboratory of Molecular Virology and Biological Chemistry, Institute of Medical Biology, Polish Academy of Sciences, Lodz, Poland

  • Patrycja Suski,

    Affiliation Laboratory of Molecular Virology and Biological Chemistry, Institute of Medical Biology, Polish Academy of Sciences, Lodz, Poland

  • Małgorzata Wiśniewska-Ligier,

    Affiliation Department of Pediatrics, Immunology, and Nephrology, Polish Mother's Memorial Hospital Research Institute, Lodz, Poland

  • Teresa Woźniakowska-Gęsicka,

    Affiliation 3rd Department of Pediatrics, Polish Mother’s Memorial Hospital Research Institute, Lodz, Poland

  • Dorota Nowakowska,

    Affiliation Department of Perinatology and Gynecology, Polish Mother’s Memorial Hospital Research Institute, Lodz, Poland

  • Zuzanna Gaj,

    Affiliations Department of Perinatology and Gynecology, Polish Mother’s Memorial Hospital Research Institute, Lodz, Poland, Scientific Laboratory of the Center of Medical Laboratory Diagnostics, Polish Mother's Memorial Hospital Research Institute, Lodz, Poland

  • Jan Wilczyński,

    Affiliation 2nd Department of Obstetrics and Gynecology, Warsaw Medical University, Warsaw, Poland

  • Zbigniew J. Leśnikowski

    Affiliation Laboratory of Molecular Virology and Biological Chemistry, Institute of Medical Biology, Polish Academy of Sciences, Lodz, Poland

Abstract

Toll-like receptor 9 (TLR9) recognizes non-methylated viral CpG-containing DNA and serves as a pattern recognition receptor that signals the presence of human cytomegalovirus (HCMV). Here, we present the genotype distribution of single-nucleotide polymorphisms (SNPs) of the TLR9 gene in infants and the relationship between TLR9 polymorphisms and HCMV infection. Four polymorphisms (-1237T/C, rs5743836; -1486T/C, rs187084; 1174G/A, rs352139; and 2848C/T, rs352140) in the TLR9 gene were genotyped in 72 infants with symptomatic HCMV infection and 70 healthy individuals. SNP genotyping was performed by using polymerase chain reaction-restriction fragment length polymorphism (PCR-RFLP). Digested fragments were separated and identified by capillary electrophoresis. The HCMV DNA copy number was measured by a quantitative real-time PCR assay. We found an increased frequency of heterozygous genotypes TLR9 -1486T/C and 2848C/T in infants with HCMV infection compared with uninfected cases. Heterozygous variants of these two SNPs increased the risk of HCMV disease in children (P = 0.044 and P = 0.029, respectively). In infants with a mutation present in at least one allele of -1486T/C and 2848C/T SNPs, a trend towards increased risk of cytomegaly was confirmed after Bonferroni’s correction for multiple testing (Pc = 0.063). The rs352139 GG genotype showed a significantly reduced relative risk for HCMV infection (Pc = 0.006). In contrast, the -1237T/C SNP was not related to viral infection. We found no evidence for linkage disequilibrium with the four examined TLR9 SNPs. The findings suggest that the TLR9 -1486T/C and 2848C/T polymorphisms could be a genetic risk factor for the development of HCMV disease.

Introduction

Human cytomegalovirus (HCMV), a member of the Betaherpesvirinae subfamily of herpesviruses, is a ubiquitous pathogen with a seroprevalence of 45–100% [1]. Primary HCMV infection is usually asymptomatic in healthy individuals and is followed by a persistent infection. In contrast, HCMV is a major reason of multiorgan disease in immunocompromised patients and a leading cause of congenital infection, occurring in 0.5–2% of pregnancies in the United States and Europe [2, 3]. Among congenitally infected neonates, approximately 10–15% exhibit signs and symptoms of disease at birth, and these symptomatic infants have an increased risk for sensorineural hearing loss and central nervous system damage [4, 5]. However, the factors that dictate whether HCMV infection is asymptomatic or symptomatic are not clear. Innate immunity plays a crucial role in preventing the acquisition of HCMV infections, whereas its failure may contribute to an increased risk of infection. It is suggested that variations in the genes that modulate innate immune responses, including TLRs genes, may result in distinct clinical presentations of infection.

Toll-like receptors (TLRs) play an important role in the innate immune response to pathogens and have been implicated in infectious and autoimmune processes. There is evidence that interactions between some TLRs with viruses influence both the immune response and outcome of HCMV infection. HCMV activates host cells via multiple TLRs, predominantly those that reside on cell surface or in endosomes. TLR2 is pattern recognition receptor (PRR) for HCMV glycoproteins B (gB) and gH [6]. TLR2 and CD14 participate in HCMV attachment, uptake, and subsequent signaling, leading to the expression of pro-inflammatory cytokine genes [6, 7]. TLR4 induces THP-1 cells signaling via the TLR4/MD2/CD14 complex which initiates and regulates additional downstream signaling [8]. It has been also demonstrated that TLR4-ligand complexes enhance the ability of dendritic cells to present viral antigen [9]. Endosomal TLR3, TLR7, and TLR9 are involved in HCMV-elicited signaling, which leads to a pro-inflammatory and antiviral response, such as type I interferons (IFNs) [6, 7, 10]. TLR3 recognizes dsRNA, also generated during the life cycle of viruses, whereas TLR7 and 8 recognize single-stranded RNA (ssRNA). TLR9, localized intracellularly within endolysosomes, recognizes unmethylated CpG dinucleotide motifs located in viral DNA and plays a central role in host defense against viral infection [11, 12]. HCMV strongly up-regulates TLR9 mRNA levels in human fibroblasts [13]. TLR9 mediates the recognition of murine CMV (MCMV), as evidenced by experiments using a mutated form of TLR9. A missense mutation in the receptor domain of the TLR9 gene (TLR9CpG1) can be induced by N-ethyl-N-nitrosourea and results in unresponsiveness to CpG-containing oligonucleotides. Mice homozygous for the TLR9CpG1 allele are highly susceptible to infection with MCMV and display impaired (infection-induced) IFN-α/ß secretion and NF-κB activation [14]. The TLR9-mediated activation of MyD88 and TLR3-dependent induction of TRIF signaling are activated in vivo upon inoculation of MCMV, leading to type I IFN production. Notably, neither of the pathways alone—in the absence of the other—offers complete protection against MCMV infection, but they instead act in an additive or codependent manner [14].

Little is known about the role of TLR polymorphisms in the pathogenesis of HCMV infection in humans. The R753Q single-nucleotide polymorphism (SNP) in the TLR2 gene was shown to be associated with increased HCMV replication and disease in transplant recipients [15, 16]. In vitro experiments showed that the R753Q SNP abolishes TLR2-mediated immune signaling in response to HCMV [17]. No association between TLR2 Arg753Gln SNP with HCMV infection was detected in both infants and immunocompromised adult patients [18]. Although the relationship between the mutation in the highly polymorphic TLR4 gene and the incidence of HCMV disease has been described in transplant patients [19], no effect of the TLR4 Asp299Gly SNP on viral infection in infancy was found [18]. Few studies have described the association between TLR9 polymorphisms and HCMV infection [2022]. Increasing studies have found that specific TLR SNPs have an association with congenital HCMV infection [23, 24]. Taniguchi et al. [23] found that the homozygous CC genotype of the 1350T/C SNP (rs3804100) in the TLR2 gene was associated with congenital HCMV infection, while no associations between TLR4 and TLR9 SNPs with HCMV infection and disease in infants were found. Recently, TLR4 and TLR9 polymorphisms were found to play a role in the development of congenital HCMV infection in fetuses and neonates [24].

Our study explored the genotype distribution of TLR9 -1237T/C (rs5743836), -1486T/C (rs187084), 1174G/A (rs352139), and 2848C/T (rs352140) SNPs in infants and the correlation between polymorphisms in the TLR9 gene and HCMV infection in infants.

Materials and Methods

Ethics statement

The study protocols were approved by the Bioethics Committee of the Medical University of Lodz (RNN/120/09/KE), and the Ethics Committee of the Polish Mother’s Memorial Hospital Research Institute. Written informed consents were obtained from all parents or guardians on behalf of the children involved in the study.

Study population

Between 2008 and 2011, 72 infants were selected as HCMV-positive at the 3rd Department of Pediatrics of the Polish Mother’s Memorial Hospital Research Institute in Lodz, Poland. HCMV infection was confirmed by HCMV DNA (UL55 gene) detection in whole-blood and/or urine samples after 3 weeks of life and by the presence of HCMV-specific antibodies. Because infants were examined after that time, children were classified as having postnatal or unproven congenital HCMV infection. However, clinical data suggest that the majority of the studied infants were infected congenitally (Table 1). Serum samples from the infants were assessed for anti-HCMV IgG and IgM antibodies with the use of CLIA LIASON CMV IgM and IgG assays (DiaSorin, Sallugia, Italy). HCMV-positive infants were compared with 70 healthy newborn infants (seronegative and aviremic), which were used as controls for the association studies. All of the study subjects were Caucasians and there were no ethnic differences between the cases and control group.

thumbnail
Table 1. Demographic and clinical characteristics of study subjects with HCMV infection.

https://doi.org/10.1371/journal.pone.0154100.t001

All infants with HCMV infection were symptomatic due to the selection bias. Clinical samples were collected from children at diagnosis or after exacerbation of cytomegaly symptoms. The children were classified as having a symptomatic infection based on clinical or laboratory findings, including hematological disorders (anemia, thrombocytopenia), pneumonia, neurological dysfunction, central nervous system damage, psychomotor retardation, liver damage, intrauterine growth restriction, and other symptoms; the classification was made after other causes had been excluded. Of the 72 patients, 30 (41.7%) were female and 42 (58.3%) were male, and their mean age at evaluation was 3.9 ± 2.4 months (median: 3.3 months, range: 1–12 months). The demographic and clinical characteristics of infants with HCMV infection were summarized in Table 1. Laboratory personnel were blinded to the demographic characteristics and clinical findings of the study subjects.

Genotyping of TLR9 polymorphisms

All genotyping was performed on genomic DNA extracted from whole-blood samples using the QIAamp DNA Blood Mini Kit (Qiagen, Hilden, Germany) according to manufacturer’s recommendations. The concentration and purity of DNA were assessed using a NanoDrop 2000c UV-vis Spectrophotometer (Thermo Scientific, Wilmington, DE, USA). The identification of TLR9 SNPs was performed by polymerase chain reaction-restriction fragment length polymorphism (PCR-RFLP) using primers described previously by Hamann et al. [25] (-1237T/C, rs5743836; -1486T/C, rs187084), Fang et al. [26] (1174G/A, rs352139) and Roszak et al. [27] (2848C/T, rs352140). Briefly, PCR was performed in a total volume of 50 μL containing 200 ng of genomic DNA, primers (1 μM each), dNTPs (2.5 mM), 10 x Taq buffer (100 mM Tris-HCl, 500 mM KCl, 0.8% Nonidet P40; pH 8.8), 2 mM MgCl2, and 1 U of Taq DNA Polymerase (Fermentas, Glen Burnie, MD, USA). The thermal cycling conditions for -1237T/C were 4 min at 95°C and 40 cycles each of 30 s at 95°C, 20 s at 50°C, and 30 s at 72°C. The PCR parameters for -1486T/C and 2848C/T were as follows: 4 min at 95°C and 35 cycles each of 30 s at 95°C, 20 s at 60°C, and 30 s at 72°C. The final extension in both PCR profiles was at 72°C for 5 min. The reactions were carried out using a Veriti thermal cycler (Applied Biosystems, Foster City, CA, USA). The PCR-amplified fragments corresponding to the TLR9 -1237T/C, -1486T/C, 1174G/A, and 2848C/T polymorphisms were digested with the restriction enzymes MvaI, AflII, FspBI, and Bsh1236I, respectively (Fermentas, Hanover, MD, USA). The digested fragments were separated using the QIAxcel system (Qiagen). The TLR9 SNPs variants were identified by different fragment lengths (Table 2; Fig 1). The results were confirmed by the direct sequencing of selected samples of each TLR9 genotype using the MiSeq system (Illumina, San Diego, CA, USA). The results of the PCR-RFLP and sequencing methods were in agreement.

thumbnail
Fig 1. Visualization of selected PCR-RFLP products for TLR9 -1486T/C genotyping.

Gel image: 1 and 2, heterozygous TC genotype (192, 327, 490 bp); 3 and 4, TT genotype (192, 327 bp); 5 and 6, CC genotype (490 bp). Alignment markers (15 bp, 1 kbp).

https://doi.org/10.1371/journal.pone.0154100.g001

thumbnail
Table 2. Restriction enzymes and length of the restriction fragments.

https://doi.org/10.1371/journal.pone.0154100.t002

Assessment of HCMV replication

Real-time PCR quantification of the HCMV DNA copy number in DNA isolates from whole-blood and urine samples was performed by using a 7900HT Fast Real-Time PCR System (Applied Biosystems) [28]. For amplification of the HCMV genome, a pair of primers and a TaqMan probe labeled at the 5’ end with FAM (6-carboxyfluoroscein) and at the 3’ end with TAMRA (6-carboxytetramethylrhodamine) targeting the gB sequence were used. The primers and probe used were the following: primer gB1, 5’-GAGGACAACGAAATCCTGTTGGGCA-3’, primer gB2, 5’-GTCGACGGTGGAGATACTGCTGAGG-3’, and probe, 5’-CAATCATGCGTTTGAAGAGGTAGTCCA-3’. The samples for real-time PCR were prepared in a volume of 25 μL containing 5 μL of the DNA extract or calibrated plasmid dilution (standard curve), 0.4 μM of each primer, 0.2 μM of the fluorogenic probe, and the TaqMan Universal PCR Master Mix (Applied Biosystems). The PCR amplification reaction used the condition: 95°C for 10 min, 60 cycles at 95°C for 15 s, followed by 60°C for 1 min. Standard curves for the plasmid DNA were obtained from serial 10-fold dilutions from 105 to 5. The number of target copies was determined from the threshold cycle CT.

Statistical analysis

The calculation of Hardy-Weinberg equilibrium (HWE), linkage disequilibrium (LD) and additional genotype and haplotype associations with HCMV infection status were performed by using the SNPSTATS program (http://bioinfo.iconcologia.net/index.php?module=Snpstats) [29]. The association between the polymorphisms and HCMV infection was estimated using an odds ratio (OR) and 95% confidence intervals (95% CIs) with unadjusted and adjusted multivariate models. P values were corrected (Pc) for multiple testing with the Bonferroni’s correction and the Pc-values of ≤ 0.05 were considered to be significant. Logistic regression models to evaluate the association between the TLR genotype and the symptoms was performed using the SPSS statistical software package for Windows 17.0 (SPSS, Chicago, IL, USA). The distribution of genotypes and alleles in the patient groups was examined by using the Fisher’s exact test. The Mann-Whitney U test was used to study the association between TLR polymorphisms and viral load. The level of significance for all statistical tests was defined as P ≤ 0.05.

Results

Frequency of TLR9 SNPs in infants

TLR9 -1237T/C, -1486T/C, 1174G/A, and 2848C/T SNPs were typed in 142 infants, including the 72 subjects with HCMV infection. Almost all individuals (139/142, 97.9%) possessed the wild-type -1237TT genotype, while the CC genotype was only found in three HCMV-infected cases (3/72, 4.2%; P > 0.05). As the -1237T/C SNP is non-polymorphic in our population, it was excluded from further analyses. The distribution of the -1486T/C, 1174G/A, and 2848C/T SNPs was different between HCMV-infected and uninfected infants (see Table 3). In infants with HCMV infection, the frequencies of the TT, TC, and CC genotypes at the -1486T/C SNP were 22.2%, 62.5%, and 15.3%, respectively. The wild-type TT genotype was detected in 40.0% and the heterozygous TC genotype in 45.7%, while the homozygous CC genotype was detected in 14.3% of healthy individuals. Similar genotype frequencies in HCMV-infected and uninfected infants were observed in the TLR9 2848C/T SNP (Table 3). The frequency of wild-type genetic variants of both -1486T/C and 2848C/T SNPs was significantly higher in uninfected infants than in HCMV-infected cases (P = 0.029; Fisher’s exact test). It should be noted that, the heterozygous variant carriers of the 2848C/T and -1486T/C SNPs were detected more frequently among infants with HCMV infection (P = 0.044 and P = 0.064, respectively; Fisher’s exact test). In case of 1174G/A SNP, the frequencies of the GG, GA, and AA genotypes in children with HCMV infection were 1.4%, 40.3%, and 58.3%, respectively. The GG genotype at this SNP was identified more frequently in uninfected than HCMV-infected infants (14.3% vs 1.4%; P = 0.004; Fisher’s exact test). In the group of infants without HCMV infection, the frequencies of genotypes at all TLR9 SNPs were in HWE (P > 0.05; Table 4).

thumbnail
Table 3. The distribution of genotypes frequencies of TLR9 SNPs in infants and relationship between polymorphisms and the risk of HCMV infection.

https://doi.org/10.1371/journal.pone.0154100.t003

thumbnail
Table 4. TLR9 SNP variance from Hardy-Weinberg equilibrium.

https://doi.org/10.1371/journal.pone.0154100.t004

In the examined infants, the frequencies of the wild-type -1486 T, 1174 G and 2848 C alleles were 58.1%, 28.5% and 56.3%, respectively, while the frequencies of the -1486 C, 1174 A and 2848 T alleles were 41.9%, 71.5% and 43.7%, respectively. In both HCMV-infected and uninfected infant groups, no significant differences in the frequency of TLR9 -1486T/C and 2848C/T alleles were observed (Table 5, P > 0.05). In contrast, the A allele of 1174G/A SNP was detected more frequently in infants with HCMV infection compared with uninfected cases (P = 0.009). Among HCMV-infected infants, in the -1237 locus, the C alleles were significantly more frequent than in uninfected individuals (P = 0.030).

thumbnail
Table 5. The distribution of allele frequencies of TLR9 SNPs in infants with and without HCMV infection.

https://doi.org/10.1371/journal.pone.0154100.t005

TLR9 polymorphisms associated with HCMV infection

Statistically significant differences between infants with and without HCMV infection were observed for the TLR9 -1486T/C, 1174G/A, and 2848C/T polymorphisms (see Table 3). For the -1486T/C, 1174G/A, and 2848C/T SNPs, the wild-type genotypes occurred more frequently in the uninfected infants compared to the HCMV-infected group. When considered separately, TLR9 SNPs were significantly associated with risk of HCMV infection. Heterozygous and homozygous CC genotypes at the -1486 locus were associated with a 2-fold increased risk of HCMV infection (OR 2.33, 95% CI 1.12–4.86, P = 0.021 in dominant model). Similarly, at least a 2-fold increased risk of infection was found for the heterozygous and homozygous TT genotypes at the 2848 locus in almost all genetic models (OR 2.33, 95% CI 1.12–4.86, P = 0.021 and OR 2.10, 95% CI 1.07–4.09, P = 0.029 in the dominant and overdominant models, respectively). These two SNPs showed a trend to association with HCMV infection after Bonferroni’s correction for multiple testing (Pc = 0.063). In infants with double heterozygous -1486T/C and 2848C/T polymorphisms, a trend towards increased risk of cytomegaly was found (OR 1.95, 95% CI 0.92–4.13, P = 0.077). In contrast, the GG genotype of the 1174G/A SNP compared to GA and AA genotypes decreased the risk of HCMV infection (OR 0.08, 95% CI 0.01–0.68, P = 0.002 in recessive model). These P value remained significant even after Bonferroni's correction (Pc = 0.006).

For the TLR9 -1486T/C and 2848C/T SNPs, a mutation present in at least one allele may lead to HCMV infection in infants. No association between specific TLR9 genotypes and the presence of specific symptoms in HCMV-infected children was found.

Associations between TLR9 SNPs and HCMV load

We correlated the SNPs with HCMV DNA concentration in peripheral blood during exacerbation of cytomegaly symptoms. The viremia levels ranged from 0 to 1.41 × 105 copies/mL (mean 2.73 × 103 ± 1.68 × 104 copies/mL) in HCMV-infected infants. However, the mean copy number of HCMV DNA was statistically higher in blood samples from infants with the heterozygous variant of TLR9 -1486T/C than those with the wild-type genotype (Fig 2, P = 0.039). In addition, the children with heterozygous and homozygous recessive genotypes at the -1486 and 2848 locus had almost 4-fold increased risk of HCMV disease in an adjusted model that included the HCMV DNA copy number (OR 3.89, 95% CI 1.45–10.46, P = 0.004 in dominant model; Table 3). This association reach statistical significance after Bonferroni’s correction (Pc = 0.012). No association was observed between the HCMV DNAemia and TLR9 1174G/A SNP (P > 0.05).

thumbnail
Fig 2. Comparison of viremia levels in HCMV-infected infants without or with the TLR9 -1486T/C SNP (N = 72).

Bars in the scatter dot plot represent the mean viral loads. Bars in the box plots represent median viral loads, upper and lower borders represent the 25th and 75th percentiles, and whiskers represent the minimum to maximum values. P ≤ 0.05, Mann-Whitney U test.

https://doi.org/10.1371/journal.pone.0154100.g002

Haplotype analysis

Multiple-SNP analysis showed that the most common haplotype for TLR9 -1237T/C, -1486T/C, 1174G/A and 2848C/T was TCAC (21.62% and 18.46% for HCMV-infected and uninfected cases, respectively). The CCAT, CTGC, and CCAC haplotypes were detected only in infants with HCMV infection, while CTAC haplotype was not detected in any children. We found an association between -1486T/C and 1174G/A SNPs (Pc < 0.001). Linkage disequilibrium analysis revealed that -1237T/C, -1486T/C, 1174G/A, and 2848C/T SNPs were not in LD with each other (correlation coefficient r2 < 0.2). In addition, the haplotypes were not observed to influence the risk for HCMV infection.

An additional multiple-SNP analysis for TLR9 -1237T/C, -1486T/C, and 2848C/T showed that the TCT haplotype was observed at a lower frequency in both infant groups (24.3% and 14.7% for HCMV-infected and uninfected cases, respectively) and it was associated with 2.5-fold increased the risk of HCMV infection (OR 2.47, 95% CI 1.14–5.33, P = 0.023). However, association did not reach statistical significance after Bonferroni’s correction (Pc > 0.05).

Discussion

In this study, we demonstrated that the heterozygous genotype of TLR9 -1486T/C and 2848C/T occurred more frequently in infants with HCMV infection than in uninfected cases. The heterozygous or homozygous recessive genotypes of the TLR9 SNPs were consistently associated with a 2-fold higher risk of HCMV infection and disease development. This study provides the first demonstration that the 2848C/T and -1486T/C genotype might be important markers of HCMV infection. Furthermore, the GG genotype of the 1174G/A locus showed a significantly reduced risk for HCMV infection. No association between a polymorphism in the promoter of the TLR9 gene (-1237T/C) and HCMV infection was found. We speculate that genetic variations in TLR9 lead to a functional deficiency of the receptor signaling pathway and they are partially responsible for the development of cytomegaly.

TLR9 has been implicated in the recognition of different families of DNA viruses, all of which contain genomes rich in CpG DNA motifs, including herpesviruses [3033]. The intracellular localization of TLR9 is critical for the discrimination of self and non-self nucleic acids. TLR9 recognizes unmethylated CpG dinucleotide motifs in DNA, a feature that is enriched in viruses and not present in host genomes. After stimulation with CpG DNA, TLR9 redistributes from the endoplasmic reticulum to lysosomes, where the receptor undergoes proteolytic cleavage of its ectodomain. Only the cleaved receptor is able to recruit MyD88 [34, 35], which signals through a protein complex consisting of TRAF6 and IL-1 receptor-associated kinase 1/4 (IRAK1/4), leading to the activation of IRF7, NF-ĸB, and mitogen-activated protein kinase [3638]. Thus, TLR9 activates the production of type I IFNs and inflammatory cytokines. Because TLR9 signaling leads to the activation of IRF7, the recognition of herpesviruses by dendritic cells leads to the expression of type I IFNs [30, 32, 39].

We studied the -1237T/C and -1486T/C SNPs, which are located within the putative promoter region, 1174G/A SNP located in intron 1, and the 2848C/T SNP located in exon 2 of TLR9 gene. TLR9 polymorphisms that are located in the promoter region most likely alter the functional ability of the receptor, as reported in other studies [4042]. These TLR9 SNPs have been reported to show high heterozygosity in other population [43]. We observed a low prevalence of the -1237T/C SNP in the examined population and an incidence of HCMV disease in individuals carrying the TLR9–1237 C allele. An analysis of the second polymorphism of the TLR9 promoter -1486T/C revealed that the mutation present in at least one allele enhanced risk of HCMV disease. There are no functional data available for the TLR9 -1486T/C polymorphism but it likely alters the function of the promoter. The combination of the C allele at position -1486 with a G allele at position 1174 has the ability to downregulate TLR9 expression, and the C allele predisposes patients to systemic lupus erythematosus [44]. Kikuchi et al. [45] reported that the 2848G/A SNP was associated with alterations in gene expression. This polymorphism does neither results in an amino acid change nor modification of a regulatory site, implying functional linkage with another proximal SNPs [46]. We found that infants with double heterozygous 2848C/T and -1486T/C genotypes had a trend towards increased risk of HCMV infection. Similar to the -1486T/C SNP, a mutation present in at least one allele of 2848C/T was associated with HCMV infection. We speculate that genetic variations of TLR9 that down regulate its expression could reduce the function of the innate immune response against HCMV infection. The previous study showed that the TLR9 2848 heterozygous status predisposes fetuses and newborns to HCMV infection and increases the risk of cytomegaly development [24]. Although the frequency of the heterozygous genotype of TLR9 SNP rs352140 was also higher in Japanese children congenitally infected with HCMV, no significant association between genotype variants and viral infection was found [23]. The present study revealed a significant association between 2848C/T genotype and the higher risk of HCMV infection in infants with postnatal or unproven congenital HCMV infection. Our findings indicate no significant association of TLR9 1174G/A polymorphism with cytomegaly among Polish infants. Nevertheless, we observed that the 1174 GG carriers had a reduced risk of HCMV infection. The genotype and allele distribution of TLR9 SNPs was comparable to that in other European populations [20, 27, 4749] but was different than in Asian and American ethnic groups [48, 5052]. These results indicate the existence of a geographic difference in TLR9 genotypes that may explain the variability to pathogen susceptibility or other conditions under which TLR9 pathways may be involved.

Specific haplotypes for the TLR9 gene might affect host defense mechanisms and influence the susceptibility or resistance to infections. TLR9–1486 C allele carriers are associated with an increased risk and poor prognosis of gastric carcinoma in the Chinese population [53], and the 2848C/T polymorphism may be associated with Hodgkin’s lymphoma [54] and cervical cancer [27, 55]. Previous studies have shown that TLR9 polymorphisms are associated with various viral infections, including HIV-1 [5660], HPV16 [55], and HCMV [23, 24]. Two TLR9 SNPs, 1174G/A (rs352139) and 2848C/T, were linked to viral load and disease progression in HIV-1-infected adults [56, 57, 59]. A significant correlation between these genetic variants and the risk of mother-to-child transmission of HIV-1 infection was found [58]. Moreover, the heterozygous 2848 AG genotype and G haplotype were strongly associated with rapid disease progression in HIV-1-infected children [60]. The role of TLR9 polymorphisms in HCMV infection is only beginning to be appreciated. It was previously reported that the -1237T/C mutated C allele was highly predictive of susceptibility to HCMV infection in stem cell transplant recipients [20]. However, there was no effect of TLR9 -1237T/C and 2848C/T polymorphisms on HCMV infection in renal transplant recipients [22]. Two other SNPs in the TLR9 gene in the donor, 1174G/A and 1635C/T (rs352140), influenced the risk of both acute graft-versus-host disease and HCMV reactivation in allogeneic hematopoietic recipients [21]. In contrast, no correlation was found between these SNPs and congenital HCMV infection or disease [23]. The current and previously described results indicated that TLR9 polymorphisms have an impact on the development of HCMV disease.

Our finding that TLR9 -1486T/C and 2848C/T SNPs are associated with HCMV infection suggests an important role of TLRs and TLR-mediated signaling in the pathogenesis and outcomes of cytomegaly. Studying SNPs among the receptors involved in viral recognition will be essential to define the genetic background associated with risk of HCMV infection and disease. Further studies that focus on gene expression and functional consequences of polymorphisms in cytomegaly are needed to determine the exact consequence of TLR9 SNPs. Increased understanding of how TLR9 polymorphisms affect congenital cytomegaly may provide a means of identifying high risk groups among newborns and their mothers.

In conclusion, it was found that the heterozygous genotypes of TLR9 -1486T/C and 2848C/T SNPs as well as the homozygous AA genotype of the intronic 1174G/A SNP were prevalent in infants with HCMV infection. An association between the heterozygous and homozygous recessive genotypes of the -1486T/C and 2848C/T SNPs and predisposition to viral infection was found. Our observations provide new insight into the role of TLR9 polymorphisms in HCMV infection and may suggest that the presence of the mutation in at least one allele may lead to disease development.

Author Contributions

Conceived and designed the experiments: EP AJ. Performed the experiments: AJ MS KS PS. Analyzed the data: EP AJ. Contributed reagents/materials/analysis tools: EP AJ MS PS ZJL. Wrote the paper: EP AJ MS. Contributed clinical samples: MWL TWG DN ZG JW. Final approval of the version to be published: EP AJ MS KS PS MWL TWG DN ZG JW ZJL.

References

  1. 1. Cannon MJ, Schmid DS, Hyde TB. Review of cytomegalovirus seroprevalence and demographic characteristics associated with infection. Rev Med Virol. 2010;20: 202–213. pmid:20564615
  2. 2. Kenneson A, Cannon MJ. Review and meta-analysis of the epidemiology of congenital cytomegalovirus (CMV) infection. Rev Med Virol. 2007;17: 253–276. pmid:17579921
  3. 3. Wang C, Zhang X, Bialek S, Cannon MJ. Attribution of congenital cytomegalovirus infection to primary versus non-primary maternal infection. Clin Infect Dis. 2011;52(2): e11–13. pmid:21288834
  4. 4. Stagno S. Cytomegalovirus. In: Remington JS, Klein JO, editors. Infectious Diseases of the Fetus and Newborn Infant, 4 th ed. Philadelphia: Saunders; 1995. pp. 312–353.
  5. 5. Demmler GJ. Infectious Diseases Society of America and Centers for Disease Control. Summary of a workshop on surveillance for congenital cytomegalovirus disease. Rev Infect Dis. 1991;13: 315–329. pmid:1645882
  6. 6. Boehme KW, Guerrero M, Compton T. Human cytomegalovirus envelope glycoproteins B and H are necessary for TLR2 activation in permissive cells. J Immunol. 2006;177: 7094–7102. pmid:17082626
  7. 7. Compton T, Kurt-Jones EA, Boehme KW, Belko J, Latz E, Golenbock DT, et al. Human cytomegalovirus activates inflammatory cytokine responses via CD14 and Toll-like receptor 2. J Virol. 2003;77: 4588–4596. pmid:12663765
  8. 8. Yew KH, Carpenter C, Duncan RS, Harrison CJ. Human cytomegalovirus induces TLR4 signaling components in monocytes altering TIRAP, TRAM and downstream interferon-beta and TNF-alpha expression. PLoS One. 2012;7(9): e44500. pmid:22970235
  9. 9. Loré K, Betts MR, Brenchley JM, Kuruppu J, Khojasteh S, Perfetto S, et al. Toll-like receptor ligands modulate dendritic cells to augment cytomegalovirus- and HIV-1-specific T cell responses. J Immunol. 2003;171: 4320–4328. pmid:14530357
  10. 10. Arav-Boger R, Wojcik GL, Duggal P, Ingersoll RG, Beaty T, Pass RF, et al. Polymorphisms in Toll-like receptor genes influence antibody responses to cytomegalovirus glycoprotein B vaccine. BMC Res Notes. 2012;5: 140. pmid:22414065
  11. 11. Latz E, Franko J, Golenbock DT, Schreiber JR. Haemophilus influenzae type b-outer membrane protein complex glycoconjugate vaccine induces cytokine production by engaging human toll-like receptor 2 (TLR2) and requires the presence of TLR2 for optimal immunogenicity. J Immunol. 2004:172: 2431–2438. pmid:14764714
  12. 12. Vaidya SA, Cheng G. Toll-like receptors and innate antiviral responses. Curr Opin Immunol. 2003;15: 402–407. pmid:12900271
  13. 13. Iversen AC, Steinkjer B, Nilsen N, Bohnhorst J, Moen SH, Vik R, et al. A proviral role for CpG in cytomegalovirus infection. J Immunol. 2009;182: 5672–5681. pmid:19380814
  14. 14. Tabeta K, Georgel P, Janssen E, Du X, Hoebe K, Crozat K, et al. Toll-like receptors 9 and 3 as essential components of innate immune defense against mouse cytomegalovirus infection. Proc Natl Acad Sci USA. 2004;101: 3516–3521. pmid:14993594
  15. 15. Kijpittayarit S, Eid AJ, Brown RA, Paya CV, Razonable RR. Relationship between Toll-like receptor 2 polymorphism and cytomegalovirus disease after liver transplantation. Clin Infect Dis. 2007;44: 1315–1320. pmid:17443468
  16. 16. Kang SH, Abdel-Massih RC, Brown RA, Dierkhising RA, Kremers WK, Razonable RR. Homozygosity for the toll-like receptor 2 R753Q single-nucleotide polymorphism is a risk factor for cytomegalovirus disease after liver transplantations. J Infect Dis. 2012;205: 639–646. pmid:22219347
  17. 17. Brown RA, Gralewski JH, Razonable RR. The R753Q polymorphism abrogates toll-like receptor 2 signaling in response to human cytomegalovirus. Clin Infect Dis. 2009;49: e96–99. pmid:19814623
  18. 18. Jabłońska A, Paradowska E, Studzińska M, Suski P, Nowakowska D, Wiśniewska-Ligier M et al. Relationship between toll-like receptor 2 Arg677Trp and Arg753Gln and toll-like receptor 4 Asp299Gly polymorphisms and cytomegalovirus infection. Int J Infect Dis. 2014;25: 11–15. pmid:24813591
  19. 19. Cervera C, Lozano F, Saval N, Gimferrer I, Ibañez A, Suárez B, et al. The influence of innate immunity gene receptors polymorphisms in renal transplant infections. Transplantation. 2007;83: 1493–1500. pmid:17565323
  20. 20. Carvalho A, Cunha C, Carotti A, Aloisi T, Guarrera O, Di Ianni M, et al. Polymorphisms in Toll-like receptor genes and susceptibility to infections in allogeneic stem cell transplantation. Exp Hematol. 2009;37: 1022–1029. pmid:19539691
  21. 21. Xiao HW, Luo Y, Lai XY, Shi JM, Tan YM, He JS, et al. Donor TLR9 gene tagSNPs influence susceptibility to a GVHD and CMV reactivation in the allo-HSCT setting without polymorphisms in the TLR4 and NOD2 genes. Bone Marrow Transplant. 2014;49: 241–247. pmid:24121213
  22. 22. Krüger B, Banas MC, Walberer A, Böger CA, Farkas S, Hoffmann U, et al. A comprehensive genotype-phenotype interaction of different Toll-like receptor variations in a renal transplant cohort. Clin Sci (Lond). 2010;119: 535–544.
  23. 23. Taniguchi R, Koyano S, Suzutani T, Goishi K, Ito Y, Morioka I, et al. Polymorphisms in TLR-2 are associated with congenital cytomegalovirus (CMV) infection but not with congenital CMV disease. Int J Infect Dis. 2013 Dec;17(12):e1092–7. pmid:23906542
  24. 24. Wujcicka W, Paradowska E, Studzińska M, Gaj Z, Wilczyński J, Leśnikowski Z, Nowakowska D. TLR9 2848 GA heterozygotic status possibly predisposes fetuses and newborns to congenital infection with human cytomegalovirus. PLoS One. 2015;10: e0122831. pmid:25844529
  25. 25. Hamann L, Hamprecht A, Gomma A, Schumann RR. Rapid and inexpensive real-time PCR for genotyping functional polymorphisms within the Toll-like receptor -2, -4, and -9 genes. J Immunol Methods. 2004;285: 281–291. pmid:14980441
  26. 26. Fang J, Hu R, Hou S, Ye Z, Xiang Q, Qi J, et al. Association of TLR2 gene polymorphisms with ocular Behçet’s disease in a Chinese Han population. Invest Ophthalmol Vis Sci. 2013;54: 8384–8392. pmid:24255044
  27. 27. Roszak A, Lianeri M, Sowińska A, Jagodziński PP. Involvement of Toll-like receptor 9 polymorphism in cervical cancer development. Mol Biol Rep. 2012;39: 8425–8430. pmid:22714906
  28. 28. Paradowska E, Przepiórkiewicz M, Nowakowska D, Studzińska M, Wilczyński J, Emery VC, et al. Detection of cytomegalovirus in human placental cells by polymerase chain reaction. APMIS. 2006;114: 764–771. pmid:17078856
  29. 29. Sole X, Guino E, Valls J, Iniesta R, Moreno V. SNPStats: a web tool for the analysis of association studies. Bioinformatics. 2006;22: 1928–1929. pmid:16720584
  30. 30. Lund J, Sato A, Akira S, Medzhitov R, Iwasaki A. Toll-like receptor 9-mediated recognition of Herpes simplex virus-2 by plasmacytoid dendritic cells. J Exp Med. 2003;198: 513–520. pmid:12900525
  31. 31. Krug A, Luker GD, Barchet W, Leib DA, Akira S, Colonna M. Herpes simplex virus type 1 activates murine natural interferon-producing cells through toll-like receptor 9. Blood. 2004;103: 1433–1437. pmid:14563635
  32. 32. Varani S, Cederarv M, Feld S, Tammik C, Frascaroli G, Landini MP, et al. Human cytomegalovirus differentially controls B cell and T cell responses through effects on plasmacytoid dendritic cells. J Immunol. 2007;179: 7767–7776. pmid:18025223
  33. 33. Fiola S, Gosselin D, Takada K, Gosselin J. TLR9 contributes to the recognition of EBV by primary monocytes and plasmacytoid dendritic cells. J Immunol. 2010;185: 3620–3631. pmid:20713890
  34. 34. Ewald SE, Lee BL, Lau L, Wickliffe KE, Shi GP, Chapman HA, et al. The ectodomain of Toll-like receptor 9 is cleaved to generate a functional receptor. Nature. 2008;456: 658–662. pmid:18820679
  35. 35. Park B, Brinkmann MM, Spooner E, Lee CC, Kim YM, Ploegh HL. Proteolytic cleavage in an endolysosomal compartment is required for activation of Toll-like receptor 9. Nat Immunol. 2008;9: 1407–1414. pmid:18931679
  36. 36. Medzhitov R, Preston-Hurlburt P, Kopp E, Stadlen A, Chen C, Ghosh S, et al. MyD88 is an adaptor protein in the hToll⁄IL-1 receptor family signaling pathways. Mol Cell. 1998;2: 253–258. pmid:9734363
  37. 37. Burns K, Martinon F, Esslinger C, Pahl H, Schneider P, Bodmer JL, et al. MyD88, an adapter protein involved in interleukin-1 signaling. J Biol Chem. 1998;273: 12203–12209. pmid:9575168
  38. 38. Muzio M, Ni J, Feng P, Dixit VM. IRAK (Pelle) family member IRAK-2 and MyD88 as proximal mediators of IL-1 signaling. Science. 1997;278: 1612–1615. pmid:9374458
  39. 39. Honda K, Yanai H, Negishi H, Asagiri M, Sato M, Mizutani T, et al. IRF-7 is the master regulator of type-I interferon-dependent immune responses. Nature. 2005;434: 772–777. pmid:15800576
  40. 40. Novak N, Yu CF, Bussmann C, Maintz L, Peng WM, Hart J, et al. Putative association of a TLR9 promoter polymorphism with atopic eczema. Allergy. 2007;62: 766–772. pmid:17573724
  41. 41. Yang HY, Lu KC, Lee HS, Huang SM, Lin YF, Wu CC, et al. Role of the functional Toll-Like receptor-9 promoter polymorphism (-1237T/C) in increased risk of end-stage renal disease: a case-control study. PLoS One. 2013;8(3): e58444. pmid:23472199
  42. 42. Mockenhaupt FP, Hamann L, von Gaertner C, Bedu-Addo G, von Kleinsorgen C, Schumann RR, et al. Common polymorphisms of toll-like receptors 4 and 9 are associated with the clinical manifestation of malaria during pregnancy. J Infect Dis. 2006;194: 184–188. pmid:16779724
  43. 43. Lazarus R, Klimecki WT, Raby BA, Vercelli D, Palmer LJ, Kwiatkowski DJ, et al. Single-nucleotide polymorphisms in the Toll-like receptor 9 gene (TLR9): frequencies, pairwise linkage disequilibrium, and haplotypes in three U.S. ethnic groups and exploratory case-control disease association studies. Genomics. 2003;81: 85–91. pmid:12573264
  44. 44. Tao K, Fujii M, Tsukumo S, Maekawa Y, Kishihara K, Kimoto Y, et al. Genetic variations of Toll-like receptor 9 predispose to systemic lupus erythematosus in Japanese population. Ann Rheum Dis. 2007;66: 905–909. pmid:17344245
  45. 45. Kikuchi K, Lian ZX, Kimura Y, Selmi C, Yang GX, Gordon SC, et al. Genetic polymorphisms of toll-like receptor 9 influence the immune response to CpG and contribute to hyper-IgM in primary biliary cirrhosis. J Autoimmun. 2005;24: 347–352. pmid:15878652
  46. 46. Sanders MS, van Well GT, Ouburg S, Morré SA, van Furth AM. Toll-like receptor 9 polymorphisms are associated with severity variables in a cohort of meningococcal meningitis survivors. BMC Infectious Diseases. 2012;12: 112. pmid:22577991
  47. 47. Carvalho A, Cunha C, Almeida AJ, Osório NS, Saraiva M, Teixeira-Coelho M, et al. The rs5743836 polymorphism in TLR9 confers a population-based increased risk of non-Hodgkin lymphoma. Genes Immun. 2012;13: 197–201. pmid:21866115
  48. 48. Piotrowski P, Lianeri M, Wudarski M, Olesińska M, Jagodziński PP. Contribution of toll-like receptor 9 gene single-nucleotide polymorphism to systemic lupus erythematosus. Rheumatol Int. 2013;33: 1121–1125. pmid:22948541
  49. 49. Sanders MS, van Well GT, Ouburg S, Morré SA, van Furth AM. Toll-like receptor 9 polymorphisms are associated with severity variables in a cohort of meningococcal meningitis survivors. BMC Infectious Diseases. 2012;12: 112. pmid:22577991
  50. 50. Ng MW, Lau CS, Chan TM, Wong WH, Lau YL. Polymorphisms of the toll-like receptor 9 (TLR9) gene with systemic lupus erythematosus in Chinese. Rheumatology (Oxford). 2005;44: 1456–1457.
  51. 51. Wu JF, Chen CH, Ni YH, Lin YT, Chen HL, Hsu Hy, et al. Toll-like receptor and hepatitis B virus clearance in chronic infected patients: a long-term prospective cohort study in Taiwan. J Infect Dis. 2012;206: 662–668. pmid:22740716
  52. 52. Velez DR, Wejse C, Stryjewski ME, Abbate E, Hulme WF, Myers JL, et al. Variants in toll-like receptors 2 and 9 influence susceptibility to pulmonary tuberculosis in Caucasians, African-Americans, and West Africans. Hum Genet. 2010;127: 65–73. pmid:19771452
  53. 53. Wang X, Xue L, Yang Y, Xu L, Zhang G. TLR9 promoter polymorphism is associated with both an increased susceptibility to gastric carcinoma and poor prognosis. PLoS One. 2013;8(6): e65731. pmid:23776537
  54. 54. Mollaki V, Georgiadis T, Tassidou A, Ioannou M, Daniil Z, Koutsokera A, et al. Polymorphisms and haplotypes in TLR9 and MYD88 are associated with the development of Hodgkin’s lymphoma: a candidate-gene association study. J Hum Genet. 2009;54: 655–659. pmid:19745833
  55. 55. Lai ZZ, Ni-Zhang , Pan XL, Song L. Toll-like receptor 9 (TLR9) gene polymorphisms associated with increased susceptibility of human papillomavirus-16 infection in patients with cervical cancer. J Int Med Res. 2013;41: 1027–1036. pmid:23816930
  56. 56. Bochud PY, Hersberger M, Taffe P, Bochud M, Stein CM, Rodriques SD, et al. Polymorphisms in Toll-like receptor 9 influence the clinical course of HIV-1 infection. AIDS. 2007;21: 441–446. pmid:17301562
  57. 57. Soriano-Sarabia N, Vallejo A, Ramírez-Lorca R, Rodríguez Mdel M, Salinas A, Pulido I, et al. Influence of the Toll-like receptor 9 1635A/G polymorphism on the CD4 count, HIV viral load, and clinical progression. J Acquir Immune Defic Syndr. 2008;49: 128–135. pmid:18769358
  58. 58. Ricci E, Malacrida S, Zanchetta M, Mosconi I, Montagna M, Giaguinto C, et al. Toll-like receptor 9 polymorphisms influence mother-to-child transmission of human immunodeficiency virus type 1. J Transl Med. 2010;8: 49. pmid:20500814
  59. 59. Pine SO, McElrath MJ, Bochud PY. Polymorphisms in toll-like receptor 4 and toll-like receptor 9 influence viral load in a seroincident cohort of HIV-1-infected individuals. AIDS. 2009;23: 2387–2395. pmid:19855253
  60. 60. Freguja R, Gianesin K, Del Bianco P, Malacrida S, Rampon O, Zanchetta M, et al. Polymorphisms of innate immunity genes influence disease progression in HIV-1-infected children. AIDS. 2012;26: 765–768. pmid:22269973