Next Article in Journal
Associations between Influenza Vaccination and Health Care Access among Adults in the United States
Next Article in Special Issue
Approaches for Selective Vaccinations in Cirrhotic Patients
Previous Article in Journal
Understanding COVID-19 Vaccine Hesitancy among Healthcare Workers in South Africa
Previous Article in Special Issue
Proteomic Analysis of Mucosal and Systemic Responses to SARS-CoV-2 Antigen
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Age and Cytokine Gene Variants Modulate the Immunogenicity and Protective Effect of SARS-CoV-2 mRNA-Based Vaccination

by
Letizia Scola
1,
Donatella Ferraro
2,
Giuseppa Luisa Sanfilippo
2,
Simona De Grazia
2,
Domenico Lio
3,* and
Giovanni Maurizio Giammanco
2
1
Clinical Pathology, Department of Biomedicine, Neuroscience and Advanced Diagnostics (Bi.N.D.), University of Palermo, Corso Tukory, 211, 90134 Palermo, Italy
2
Microbiology, Department of Health Promotion, Mother and Child Care, Internal Medicine and Medical Specialties (PROMISE), University of Palermo, 90133 Palermo, Italy
3
Interdepartmental Research Center “Migrate”, University of Palermo, 90133 Palermo, Italy
*
Author to whom correspondence should be addressed.
Vaccines 2023, 11(2), 413; https://doi.org/10.3390/vaccines11020413
Submission received: 29 December 2022 / Revised: 5 February 2023 / Accepted: 9 February 2023 / Published: 10 February 2023
(This article belongs to the Special Issue Research on Immune Response and Vaccines)

Abstract

:
The introduction of anti-SARS-CoV-2 vaccines in late 2020 substantially changed the pandemic picture, inducing effective protection in the population. However, individual variability was observed with different levels of cellular response and neutralizing antibodies. We report data on the impact of age, gender, and 16 single nucleotide polymorphisms (SNPs) of cytokine genes on the anti-SARS-CoV-2 IgG titers measured 31 and 105 days after administration of the second dose of BNT162b2 vaccine to 122 healthy subjects from the health care staff of the Palermo University Hospital, Italy. The higher titers at 31 days were measured in the younger subjects and in subjects bearing T-positive genotypes of IL-1R1 rs2234650 or the GG homozygous genotype of IL-6 rs1800795 SNP. T-positive genotypes are also significantly more common in subjects with higher titers at day 105. In addition, in this group of subjects, the frequency of the CT genotype of IL-4 rs2243250 is higher among those vaccinated with higher titers. Moreover, these SNPs and TNFA rs1800629 are differently distributed in a group of subjects that were found infected by SARS-CoV-2 at day 105 of evaluation. Finally, subjects that were found to be infected by SARS-CoV-2 at day 105 were significantly older than the uninfected subjects. Taken together, these data seem to suggest that age and polymorphisms of key cytokines, which regulate inflammation and humoral immune response, might influence the magnitude of the antibody response to vaccination with BNT162B2, prompting speculation about the possible benefit of a genetic background-based assessment of a personalized approach to the anti-COVID vaccination schedule.

1. Introduction

The introduction of anti-SARS-CoV-2 vaccines has substantially modified the pandemic picture, inducing effective protection in the population. Several vaccines have been developed and administered, which have shown heterogeneous efficacy and immunogenicity against the different variants of SARS-CoV-2 with wide inter-individual differences, so that booster doses were required. The Pfizer-BioNTech vaccine BNT162b2, based on lipoic nanoparticles containing the mRNA for the spike (S) protein [1], will be introduced at the end of 2020. However, it is well known that the immunogenicity and effectiveness of a vaccine can be impacted by different factors, and differential patterns of vaccine effectiveness have been observed in diverse populations, resulting from the complex interplay among the host, pathogens, and environmental factors [2,3]. Concerning the host factors, gender, aging, and genetic factors, such as the frequencies of specific genetic variants of immune inflammatory genes [4], have been largely documented. As a result, genome-wide association studies (GWAS) and other genetic investigations have focused on large panels of genes, including genetic variants of HLA, PAMPs, cytokines, chemokines, and receptor molecules, as possible causes of variability in the response to antiviral vaccines against hepatitis B, measles, rubella, influenza A, smallpox, anthrax, and mumps [5,6,7,8,9,10,11,12]. Similar results were obtained studying the association of antiviral vaccine response with polymorphic variants in cytokine genes, such as TNF-α, IL-2, IL-4 IL-10, IL-28, IFNγ [13,14,15,16].
However, there is currently limited evidence on factors influencing individual responses to anti-SARS-CoV-2 vaccines. A wide range of sociodemographic, biological, clinical, and nutritional factors have been reported to influence the different production of anti-spike antibodies and their titers after vaccination in people of different ages and ethnicities [17]. Therefore, individual genetic background, which influences the intensity and quality of the immune and inflammatory response, could also be implicated in the regulation of the vaccine-induced anti-SARS-CoV-2 immune response. Based on such evidence, it might be useful to evaluate the impact of cytokine gene variability on vaccine efficacy.
Consistent with this aim, in this study we aimed to detect the biological effect of single nucleotide polymorphisms (SNPs) in cytokine genes involved in the regulation of inflammation and antibody production on the magnitude and duration of the humoral immune response induced by vaccination against SARS-CoV-2. The association of functionally relevant genetic variants in cytokine genes with anti-S1/S2 IgG antibody levels measured at 31 and 105 days after administration of the second dose of Pfizer-BioNTech’s Comirnaty BNT162b2 was evaluated.
We evaluated polymorphisms of the following IL-1 superfamily genes: IL-1A rs1800587 [18], IL-1B rs1143634 [19] and rs16944 [20], IL-18 rs187238 [21] and rs1946518 [22], which influence cytokine levels and production, IL-1RN rs315952, whose alleles modulate efficiency of inflammation control [23], and IL-1R1 rs2234650, whose alleles create two alternative putative binding sites for two different transcription factors and activation of different metabolic pathways [24]. In addition, IL-6 rs1800795 [25] and TNFA rs1800629 [26] polymorphisms, known to be involved in regulating the production of these two key cytokines for the inflammatory response, were typed. Considering the key role of IL-10 in regulating inflammation and antibody production, the IL-10 SNPs rs1800896, rs1800872, and rs3021097, which regulate IL-10 cytokine production [27], were also typed. Finally, the association of functional polymorphisms of key Th1 and Th2 cytokines (IL-4 rs2243250 [28], IL-13 rs1800925 [29], IFNG rs2430561 [30], and IFNGR2 rs2834213 [31]) with vaccination-induced anti-S1/S2 IgG antibody levels was evaluated.

2. Materials and Methods

2.1. Subjects

Blood samples were collected at the Department of Health Promotion, Mother and Child Care, Internal Medicine, and Specialties of Excellence “G. D’Alessandro” (PROSAMI) of the University of Palermo from 122 healthy subjects (66 women and 56 men with a mean age of 49.45 ± 13.41 years) from the health care staff of the University Hospital “P. Giaccone”, professionally in contact with vulnerable people, and vaccinated with Pfizer-BioNTech’s Comirnaty BNT162b2.
All subjects, whose nasopharyngeal swab was negative for the SARS-CoV-2 biomolecular test at the time of vaccination, received the first dose of vaccine between December 2020 and July 2021 and received the second dose within the scheduled time frame (21 to 28 days after the first dose). Enrolled, vaccinated subjects underwent two blood samplings: on days 31 and 105 after the completion of the two-dose vaccination cycle. Exclusion criteria were as follows: (a) positivity for SARS-CoV-2 infection on day 31 after the second dose revealed by antigenic or serologic analyses; (b) acute or chronic diseases; and (c) drug assumption. All participants gave their informed consent. Data were encoded to ensure privacy protection of the subjects. The study was conducted in accordance with the Declaration of Helsinki and approved by the Ethical Committee of the A.O.U.P. “P. Giaccone” University Hospital (Protocol No 0006; Date 24 June 2020)

2.2. Quantitative Determination of Anti-SARS-CoV-2 S1/S2 IgG

Sera from the study population were collected at 31 and 105 days after the second dose of Pfizer vaccination. All samples were analyzed as previously described [32] by chemiluminescent immunoassay (CLIA) technology, (LIAISON® SARS CoV-2, Diasorin, Saluggia (VC)—Italy) according to the manufacturer’s instructions on the LIAISON® XL Analyzer. IgG antibodies against S1/S2 antigens of SARS CoV -2 were detected in a semi-quantitative assay with a lower limit of detection (LoD) of 0.3 AU/mL (arbitrary units/mL) and an upper limit for quantitative evaluation at 400 AU/mL. As suggested by the manufacturer, samples were considered positive when AU/mL (arbitrary unit/mL) was ≥15, and negative when AU/mL was ≤12 AU/mL, while with results between 12 and 15 AU/mL, the samples were considered borderline [33].

2.3. Cytokine SNP Molecular Typing

Blood specimens collected in tripotassium EDTA sterile tubes were stored at −80 °C until used for DNA extraction with the “Magna Pure” 24 System automated extraction method (Roche Diagnostics S.p.A., Monza (MB), Italy). This approach, based on a solid-phase extraction method (capture of silica-coated magnetic microspheres with high DNA affinity followed by DNA elution), allows efficient and high DNA yields. As reported in Table 1, we selected seventeen functional and common SNPs. The dbSNP NCBI database, part of the ENSEMBL project (http://www.ensembl.org/index.html, last access: 8 September 2022), was queried for the selection of SNPs. DNA samples were typed using dedicated competitive allele-specific PCR (polymerase chain reaction) assays (KASPar), based on homogeneous fluorescence resonance energy transfer (FRET) detection, developed by K-Bioscience (K-Bioscience Ltd., Hoddesdon, UK), as previously described [34]. Genotypes were determined using the 7300 system SDS software, version 1.3 (Applera Italia, MONZA (MI), Italy), on each individual sample, on the basis of the detection of fluorescence signals (single for homozygous samples, double for heterozygous samples).

2.4. Identification of SARS-CoV-2 Infection by Molecular Testing of Nasopharyngeal Swabs

Samples from rhino-pharyngeal swabs were analyzed with the Allplextm SARS-CoV-2 Assay (Seegenetm, Arrow Diagnostics S.r.l, Genoa, Italy), which allows qualitative detection of SARS-CoV-2 by multiplex real-time RT-PCR. The assay allows simultaneous amplification and detection of the target nucleic acids of the E gene, RdRP gene, S gene, and N gene of SARS-CoV-2. The presence of specific gene sequences in the reaction is reported as a Ct value using Seegene Viewer analysis software (Seegenetm, Arrow Diagnostics S.r.l, Genoa, Italy). An exogenous gene is used as an internal control (IC) to monitor the entire nucleic acid extraction process and to check for PCR inhibitors.

2.5. Statistics

Allele and genotype frequencies were evaluated by gene count, using an online statistical analysis tool for the evaluation of SNPs (https://www.snpstats.net/start.htm, last access: 25 November 2022). The data were tested for goodness of fit between observed and expected genotype frequencies, according to Hardy-Weinberg equilibrium, by Pearson’s distribution and χ2 tests. Significant differences in allele, homozygous, and heterozygous genotype distributions among groups were calculated using Fisher’s exact test (adjusted for age and gender). Multiple logistic regression models were applied, using dominant (major allele homozygotes versus heterozygotes plus minor allele homozygotes) and recessive (major allele homozygotes plus heterozygotes versus minor allele homozygotes) models. Odds ratios (OR), 95% confidence intervals (95% C.I.), and p values (p-value cutoff < 0.05) were determined using GraphPad InStat software version 3.06 (GraphPad, San Diego, California, USA) and the above-mentioned online statistical analysis tool.

3. Results

3.1. Antibody Titers Stratified for Age and Gender

Table 2 shows descriptive statistics of antibody titers in 122 healthy subjects treated with two doses of recombinant RNA vaccine against the SARS-CoV-2 virus, measured 31 and 105 days after administration of the second dose. In the time interval between the two doses, 39 subjects included in the study were infected with SARS-CoV-2. As reported in the table, a decrease in titers was observed in the remaining 83 uninfected subjects, comparing the values at day 31 with those at day 105.
Since age and gender are factors that can influence the pattern of antibody response, we evaluated whether these two variables were equally distributed in the groups of subjects showing a decrease in titers above or below the median reported in Table 2. Furthermore, it was assessed whether age and gender were equally distributed in our population according to the antibody titers measured on days 31 and 105. As shown in Table 3, at day 31, subjects with titers above the group median were significantly younger than those with titers below the median. The same significant difference in age distribution was observed among uninfected subjects at day 105. Finally, subjects who had suffered through a SARS-CoV-2 infection, by day 105, despite vaccination, were significantly older than the uninfected subjects at the same time point.
Considering the results obtained, we have investigated if titer levels at 31 days were different in subjects infected compared to those that remained uninfected. We found that antibody titer levels at 31 days post-administration of the second BNT162b2 dose were significantly reduced in subjects who then developed infection, compared with those still uninfected at 105 days (Table 4).
The 66 women and 56 men included in the study showed no significant difference in the analysis by gender (Table 5). Both at 31 and 105 days post-vaccine second dose, antibody titer levels above or below the median of the group were equally distributed in men and women. No gender difference was observed at day 105 comparing SARS-CoV-2-infected versus uninfected individuals.

3.2. Correlation of Antibody Titers to Cytokine Gene Polymorphisms

Subjects were classified into two groups according to whether their antibody titers were above or below the median of the titers detected at day 31 (Table 6). The complete evaluation of the effect of alleles and genotypes of the different genes on the degree of antibody response at day 31 is reported in Supplementary Table S1.
No significant results were obtained for IL-1A rs1800587, IL-1B rs1143634 and rs16944, IL-1RN rs315952, TNFA rs1800629, IL-18 rs187238 and rs1946518, IL-10 rs1800896, rs1800872, and rs3021097, IL-4 rs2243250, IL-13 rs1800925, IFNG rs2430561, and IFNGR2 rs2834213. Instead, as shown in Table 6, the CC homozygous genotype of IL-1R1 rs2234650 was less represented among subjects with a higher titer of anti-SARS-CoV-2 antibodies, whereas T-positive genotype (dominant model: C/T-T/T vs C/C, OR 4.14 (1.41–12.19) p 0.0071) frequency was significantly increased among these subjects. Similarly, the GG homozygous genotype of IL-6 rs1800795 SNP was more frequent among subjects with a higher titer of anti-SARS-CoV-2 antibodies, whereas C-positive genotype (associated to a reduced production of the IL-6) was correlated to lower antibody titers.
Among the 122 subjects included in the study, 39 people reported evidence of SARS-CoV-2 infection, detected by antigenic or biomolecular swab, before the second blood sampling, 105 days after the administration of the second dose of vaccine. Therefore, they were excluded from further analyses of the association of genetic polymorphisms with anti-SARS-CoV-2 S1/S2 antibody levels. Data reported in Table 7 were obtained by analyzing the cytokine gene SNP allele and genotype association with anti-SARS-CoV-2 S1/S2 antibody levels in the remaining 83 uninfected subjects. Similarly to the results obtained on days 31and 105, the CC homozygous genotype of IL-1R1 rs2234650 was less represented among subjects with a higher titer of anti-SARS-CoV-2 antibodies, whereas the IL-1R1 rs2234650 T-positive genotype was significantly increased (dominant model: OR 3.67 (1.04–12.99) p 0.033). Moreover, the distribution frequency of the CT genotype of the IL-4 rs2243250 SNP is higher among vaccinated individuals with higher titers. No significant results were obtained for IL-1A rs1800587, IL-1B rs1143634 and rs16944, IL-1RN rs315952, IL-6 rs1800795, TNFA rs1800629, IL-18 rs187238, and rs1946518, IL-10 rs1800896, rs1800872, and rs3021097, IL-13 rs1800925, IFNG rs2430561, and IFNGR2 rs2834213, as shown in Supplementary Table S2.
We also assessed that most of the genotyped SNPs (IL-1A rs1800587, IL-1B rs1143634 and rs16944, IL-1RN rs315952, IL-18 rs187238, and rs1946518, IL-10 rs1800896, rs1800872, and rs3021097, IL-13 rs1800925, IFNG rs2430561, and IFNGR2 rs2834213) were not significantly associated with susceptibility to infection in vaccinated subjects (Supplementary Table S3).
However, as shown in Table 8, the C allele of IL-6 rs1800795 (associated with a reduced production of the cytokine) was increased in the infected subjects, mainly due to the significant rise in the frequency of distribution of the homozygous CC genotype (dominant model: G/C-C/C vs. G/G, OR 2.29 (1.05–4.96), p 0.045). Moreover, the group of subjects who tested positive for SARS-CoV-2 infection before day 105 was characterized by a higher frequency distribution of the heterozygous CT genotype of the IL-1R1 rs2234650 SNP and of the homozygous AA genotype of TNFA rs1800629, which in Caucasians is associated with increased cytokine production. Finally, an increase in the frequency of the T allele of IL-4 rs2243250 SNP, which is associated with a greater production of cytokines, and an increase in T-positive genotypes frequency distribution (dominant model C/T-T/T vs C/C, OR 2.67 (1.21–5.88) p 0.014) was observed in infected subjects.

4. Discussion

SARS-CoV-2 infection can result in considerable variability in the severity of clinical symptoms, from asymptomatic infection to acute respiratory distress syndrome, with a clear age- and gender-dependent trend, with men older than 65 years accounting for approximately 80% of hospitalizations for severe COVID-19 [35]. Accordingly, in early 2021, COVID vaccinations were offered first to people over the age of 80, their caregivers, and sanitary staff professionally in contact with vulnerable people. Considering the impact of age and gender on severe COVID-19 susceptibility, we have evaluated the biological effect of these variables on the antibody response to the Pfizer- BNT162b2 vaccination. Our results indicate that subjects with higher antibody titers were significantly younger than those with lower titers, both at 31 and 105 days after the second vaccine dose. The same significant difference in age distribution was observed among uninfected and infected subjects at day 105. Actually, SARS-CoV-2-infected subjects were older than uninfected individuals. These results are in good agreement with data reported by other groups showing a reduction of anti-SARS-CoV-2 spike-specific IgG and neutralization titers in older subjects [36,37].
About 25% of the subjects recruited for this study were infected by SARS-CoV-2 between 31 and 105 days after administration of the second dose of vaccine, and age seemed to correlate with susceptibility to infection in vaccinated subjects.
The differential gender-related susceptibility to COVID-19 is well known [38,39]. However, our data indicate that gender does not influence antibody titers or the likelihood of vaccinees acquiring a breakthrough SARS-CoV-2 infection. Banki and coworkers [40] had found no gender correlation between SARS-CoV-2-specific T- and B-cell responses, at 35 ± 8 and 215 ± 7 days after the second dose in 600 subjects who participated in a rapid mass vaccination against SARS-CoV-2 with BNT162b2 in Austria.
Although the development of a high titer of anti-S antibodies cannot be considered the only suitable biomarker to predict protection against SARS-CoV-2 after vaccination [41,42], a lower antibody titer could be considered predictive of a higher susceptibility to infection in vaccinees [43]. Our data seem to confirm the latter, as at day 31, the antibody titers of vaccinated subjects who later developed SARS-CoV-2 infection were significantly reduced compared to those of subjects still uninfected at day 105. From our results, reduced anti-S IgG titers in subjects older than 50 years, one month after the second dose of vaccination, might be a marker for assessing the risk of SARS-CoV-2 infection. However, these findings should be confirmed in a larger group of subjects, also evaluating the cell-mediated response to anti-SARS-CoV-2 vaccines.
A key mechanism for inducing an efficient response to RNA based vaccines is the induction of a sustained inflammatory response induced by lipidic envelopment [44]. Some authors indicated that anti-S IgG and neutralizing antibody titers resulting from the second BNT162b2 dose were significantly associated with fever, a reaction in which proinflammatory cytokine activation is a mandatory mechanism [45]. Therefore, a genetically determined increase in proinflammatory cytokine production could be relevant to achieving a satisfactory rate of protection. RNA vaccines activate an inflammatory pathway mediated by IL-1 [44], which is an essential modulator in the natural and induced immune responses. We found no association between the genotypic frequency of IL-1 and an increase or decrease in the antibody titer in response to the BNT162b2 vaccine. However, a significant finding relates to the IL-1R rs2234650 polymorphism, as rs2234650T carrier genotypes correspond to the highest neutralizing Ab titers at days 31 and 105, while the homozygous CC condition is associated with the lowest antibody titers. In a 2015 study, using in silico analyses, Vasilyev and coworkers [24] suggested that rs2234650 alleles create two alternative putative binding sites for two different transcription factors. In silico structure associated to the C allele might bind yin yang 1 transcription factor (YY1), whereas the T allele structural model resulted in a binding site for the activation of protein 1 (AP-1) transcriptional factor. YY1 is involved in inflammation and tumorigenesis [26,46]. In particular, YY1 is known to be involved in viral diseases, playing a role in the activation of HIV-1 replication [47] and in the promotion of oncogenic HPV [48]. AP-1 modulates proliferation, differentiation, apoptosis, and inflammation [49] and has a central role in the induction of IL-1β production and secretion mediated by nuclear translocation of activated AP-1 [50,51]. The conformational change associated with the T allele might facilitate increased production of IL-1, which is essential to obtain protective anti-S IgG and neutralizing antibody titers toward RNA vaccines [44]. Accordingly, our data demonstrate that IL-1R1 rs2234650 T-positive genotypes are significantly associated with higher titers of anti-S IgG at both day 31 and day 105.
Antibody production by B cells is influenced by a concert of signal-specific coactivators, including inflammasome activators and IL-1 linking to the IL-1 receptor, all converging on MyD88 and associated signaling adapters. The latter lead to the activation of the NF-κB and release of trapped NF-κB/Rel transcription factors into the nucleus, resulting in the alteration of the expression of hundreds of target genes, including immunoglobulins genes [52].
IL-6 is involved in both activating inflammation [53] and stimulating immunoglobulin production [54,55] by driving helper lymphocyte differentiation [56]. The best known and most studied polymorphism of IL-6 is located in the promoter, the G/C substitution at position -174 from the transcription start site (rs1800795). The major -174G allele is associated with increased mRNA expression, up to 2.4-fold following IL-1 stimulation, whereas the C allele is associated with genotypes with low production of IL-6 [23,57]. In spite of the well-known role of IL-6 in the cytokine storm induced by SARS-CoV-2 infection [58,59], no clear evidence has been presented on the role of IL-6 SNPs as susceptibility or protective factors in COVID-19. In a recent study, a haplotype common in Asia, C-T-T, represented by variant alleles of rs1800796, rs1524107, and rs2066992 SNPs, associated with a reduced expression of IL-6 following inflammatory stimuli, was identified as a protective genetic background associated with a better outcome of SARS-CoV-2 infection [60]. IL-6 promoter SNPs rs1800795 is polymorphic almost exclusively in Caucasians, and it has been suggested that genotype rs1800795GG, associated with high IL-6 production, might be protective against severe COVID-19 [61,62]. Our data seem to indicate that the IL-6 rs1800795GG genotype is strongly associated with higher anti-spike antibody titers, at least at day 31 after the second dose of the BNT162b2 vaccine. On the other hand, proliferation of B-cells and their differentiation into antibody-secreting plasma cells, the key mechanism for a successful vaccine response, are sustained by type 2 T helper (Th2) cytokines such as IL-4, IL-10, and IL-6 itself [63]. In this view, a genetically determined high IL-6 production might be useful to determine a favorable environment to reach a high rate of antibody protection against SARS-CoV-2 infection, as demonstrated by other groups for vaccination against the H1N1 flu virus [64].
Our observation that the IL-4 rs2243250T positive genotype (heterozygous C/T genotype) is associated with high levels of anti-spike antibodies prompt us to hypothesize that the presence of IL-1R1 rs2234650T, IL-6 rs1800795G, and IL-4 rs2243250T positive genotypes, which appear to be associated with an increased production of the respective cytokines [65,66], tagged the maintenance of optimal antibody production in subjects receiving the second dose of anti-SARS-CoV-2 mRNA vaccine.
As is well known, IFNγ is the key cytokine for IgG isotype switching, whereas IL-4 stimulates the proliferation, maturation, and differentiation of B lymphocytes in plasma cells actively secreting IgE and IgG4 [67]. In addition, IL-4 is essential to maintain naïve B cells and the production of memory B cells after exposure to an antigen or vaccination [68]. Therefore, a genetically determined increased release of IL-4 might be involved in maintaining optimal IgG production after IFNγ-mediated isotype switch.
On the other hand, when the cytokine genotype assets of subjects who had been infected after the second dose of mRNA vaccine were compared to those of uninfected subjects, we found a higher frequency of IL-4 rs2243250 T genotypes (associated with increased cytokine production). As is well known, IL-4, as well as IL-13, is the key cytokine in the induction of Th2- and macrophage 2 (M2)-mediated inflammation [69]. In COVID-19 patients, significantly higher IL-4 lung tissue expression and M2 macrophages were observed [70], and the prevalence of IL-4 Th2-mediated lung damage was a characteristic of the ineffective immune response elicited by SARS-CoV-2 [69]. In addition, it has been reported that in COVID-19 patients, the virus activates apoptosis by stimulating JAK-STAT6 signaling pathway through increased Th2 and IL-4 expression [71]. Therefore, it is possible to speculate that a genetic asset that favors high IL-4 production [28] may have different pleiotropic effects, such as favoring immunoglobulin production after vaccine immune stimulation or, conversely, being a susceptibility factor for COVID-19 in vaccinated subjects.
Interestingly, we found that the frequencies of IL-1R1 rs2234650CT and IL-6 rs1800795C positive genotypes were significantly increased in the group of subjects infected with SARS-CoV-2 after administration of the second dose of vaccine. The opposite genetic asset was observed in the group of subjects with the higher anti-S antibody titer at day 31, and allowed speculating that the simultaneous presence of IL-1R1 rs2234650TT and IL-6 rs1800795GG genotypes may be protective against the SARS-CoV-2 infection in vaccinated patients. Finally, a higher frequency of the homozygous AA genotype, which is characterized by increased transcriptional expression [24] of TNFA rs1800629, was detected in breakthrough infections. The role of TNF-α in worsening the clinical picture of COVID-19, ARDS, and systemic inflammation, is well known [72]. In addition, some research groups have identified TNFA rs1800629A as a marker of susceptibility to COVID-19 [73]. TNFA rs1800629G/A minor alleles seem to be associated with increased risk and severity of other viral respiratory infections, such as respiratory syncytial virus bronchiolitis and pneumonia [74]. Overall, our data highlight the complex role of genetic background in the humoral immune response against SARS CoV-2 and vaccine antigens and suggest further studies to evaluate the role of polymorphic variants following the booster vaccination cycle and in a larger population sample.
A limitation of this study is the lack of data on the cellular immune response to vaccination with BNT162b2 (e.g., evaluation of IFNγ levels by COVID-19-specific Quantiferon assay [75]), so we cannot analyze at this moment the effect of genetic background on the specific cellular immune response. Further studies are warranted to address this important topic.

5. Conclusions

In conclusion, our data seem to suggest that age and polymorphisms of key cytokines that regulate inflammation and humoral immune responses might influence the extent of the antibody response to the anti-SARS-CoV-2 vaccination. These data, to be confirmed in larger population samples and reevaluated at the completion of the vaccination cycle with three or four doses, could be useful for the evaluation of a personalized approach to anti-COVID vaccination scheduling. For subjects with a less effective response to the vaccine, different administration timing or a different vaccine formulation could be considered.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/vaccines11020413/s1. Table S1: Influence of the different cytokine gene polymorphisms on the SARS-CoV-2 S1/S2 antibody titer 31 days after administration of the second dose of recombinant anti-SARS-CoV-2 mRNA vaccine; Table S2: Evaluation of the association of the alleles and genotypes of cytokine gene polymorphisms with the anti-SARS-CoV-2 S1/S2 antibody titers after 105 days from the administration of the second dose of recombinant mRNA vaccine (83 uninfected subjects); Table S3: Evaluation of the association between cytokine gene polymorphisms and the development of SARS-CoV-2 breakthrough infection within day 105 after administration of the second dose of recombinant mRNA vaccine.

Author Contributions

Conceptualization, D.L., G.M.G., L.S. and D.F.; methodology, L.S., G.L.S. and D.F.; validation, G.M.G. and D.L.; formal analysis, D.L.; investigation, L.S., G.L.S. and S.D.G.; resources, G.M.G.; data curation, G.L.S. and D.L.; writing—original draft preparation, L.S.; writing—review and editing, D.L., G.M.G. and D.F.; visualization, S.D.G.; supervision, D.L.; project administration, G.M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

The study was conducted in accordance with the Declaration of Helsinki and approved by the Institutional Ethics Committee of the University Hospital (Comitato Etico Palermo 1, protocol code CET1 0006/2020, date of approval: 24 June 2020).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

All data generated or analyzed during this study are stored in electronic archives that can be supplied on reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Polack, F.P.; Thomas, S.J.; Kitchin, N.; Absalon, J.; Gurtman, A.; Lockhart, S.; Perez, J.L.; Pérez Marc, G.; Moreira, E.D.; Zerbini, C.; et al. Safety and efficacy of the BNT162b2 mRNA COVID-19 vaccine. N. Engl. J. Med. 2020, 383, 2603–2615. [Google Scholar] [CrossRef] [PubMed]
  2. Ellwanger, J.H.; Chies, J.A.B. Host genetic factors can impact vaccine immunogenicity and effectiveness. Lancet Infect. Dis. 2019, 19, 359–360. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, J.; Deng, J.C.; Goldstein, D.R. How aging impacts vaccine efficacy: Known molecular and cellular mechanisms and future directions. Trends Mol. Med. 2022, 28, 1100–1111. [Google Scholar] [CrossRef] [PubMed]
  4. Verbeke, R.; Hogan, M.J.; Loré, K.; Pardi, N. Innate immune mechanisms of mRNA vaccines. Immunity 2022, 55, 1993–2005. [Google Scholar] [CrossRef] [PubMed]
  5. Kennedy, R.B.; Ovsyannikova, I.G.; Haralambieva, I.H.; Lambert, N.D.; Pankratz, V.S.; Poland, G.A. Genome-wide SNP associations with rubella-specific cytokine responses in measles-mumps-rubella vaccine recipients. Immunogenetics 2014, 66, 493–499. [Google Scholar] [CrossRef]
  6. Nishida, N.; Sugiyama, M.; Kawai, Y.; Naka, I.; Iwamoto, N.; Suzuki, T.; Suzuki, M.; Miyazato, Y.; Suzuki, S.; Izumi, S.; et al. Genetic association of IL17 and the importance of ABO blood group antigens in saliva to COVID-19. Sci. Rep. 2022, 12, 3854. [Google Scholar] [CrossRef]
  7. Pajewski, N.M.; Shrestha, S.; Quinn, C.P.; Parker, S.D.; Wiener, H.; Aissani, B.; McKinney, B.A.; Poland, G.A.; Edberg, J.C.; Kimberly, R.P.; et al. A genome-wide association study of host genetic determinants of the antibody response to Anthrax Vaccine Adsorbed. Vaccine 2012, 30, 4778–4784. [Google Scholar] [CrossRef]
  8. Chung, S.; Roh, E.Y.; Park, B.; Lee, Y.; Shin, S.; Yoon, J.H.; Song, E.Y. GWAS identifying HLA-DPB1 gene variants associated with responsiveness to hepatitis B virus vaccination in Koreans: Independent association of HLA-DPB1_04:02 possessing rs1042169G-rs9277355C-rs9277356A. J. Viral Hepat. 2019, 26, 1318–1329. [Google Scholar] [CrossRef]
  9. Pan, L.; Zhang, L.; Zhang, W.; Wu, X.; Li, Y.; Yan, B.; Zhu, X.; Liu, X.; Yang, C.; Xu, J.; et al. A genome-wide association study identifies polymorphisms in the HLA-DR region associated with non-response to hepatitis B vaccination in Chinese Han populations. Hum. Mol. Genet. 2014, 23, 2210–2219. [Google Scholar] [CrossRef]
  10. Png, E.; Thalamuthu, A.; Ong, R.T.; Snippe, H.; Boland, G.J.; Seielstad, M. A genome-wide association study of hepatitis B vaccine response in an Indonesian population reveals multiple independent risk variants in the HLA region. Hum. Mol. Genet. 2011, 20, 3893–3898. [Google Scholar] [CrossRef]
  11. Feenstra, B.; Pasternak, B.; Geller, F.; Carstensen, L.; Wang, T.; Huang, F.; Eitson, J.L.; Hollegaard, M.V.; Svanström, H.; Vestergaard, M.; et al. Common variants associated with general and MMR vaccine–related febrile seizures. Nat. Genet. 2014, 46, 1274–1282. [Google Scholar] [CrossRef] [PubMed]
  12. Hallberg, P.; Smedje, H.; Eriksson, N.; Kohnke, H.; Daniilidou, M.; Öhman, I.; Yue, Q.-Y.; Cavalli, M.; Wadelius, C.; Magnusson, P.K.; et al. Pandemrix-induced narcolepsy is associated with genes related to immunity and neuronal survival. Ebiomedicine 2019, 40, 595–604. [Google Scholar] [CrossRef] [PubMed]
  13. Yao, Y.; Xu, X.; Li, Y.; Wang, X.; Yang, H.; Chen, J.; Liu, S.; Deng, Y.; Zhao, Z.; Yin, Q.; et al. Study of the association of seventeen single nucleotide polymorphisms and their haplotypes in the TNF-α, IL-2, IL-4 and IL-10 genes with the antibody response to inactivated Japanese encephalitis vaccine. Hum. Vaccines Immunother. 2020, 16, 2449–2455. [Google Scholar] [CrossRef]
  14. Hashempour, T.; Dehghani, B.; Musavi, Z.; Moayedi, J.; Hasanshahi, Z.; Sarvari, J.; Hosseini, S.Y.; Hosseini, E.; Moeini, M.; Merat, S. Impact of IL28 Genotypes and Modeling the Interactions of HCV Core Protein on Treatment of Hepatitis C. Interdiscip. Sci. 2020, 12, 424–437. [Google Scholar] [CrossRef] [PubMed]
  15. Córdova-Dávalos, L.E.; Hernández-Mercado, A.; Barrón-García, C.B.; Rojas-Martínez, A.; Jiménez, M.; Salinas, E.; Cervantes-García, D. Impact of genetic polymorphisms related to innate immune response on respiratory syncytial virus infection in children. Virus Genes 2022, 58, 501–514. [Google Scholar] [CrossRef]
  16. Azamor, T.; da Silva, A.; Melgaço, J.; dos Santos, A.; Xavier-Carvalho, C.; Alvarado-Arnez, L.; Batista-Silva, L.; Matos, D.D.S.; Bayma, C.; Missailidis, S.; et al. Activation of an Effective Immune Response after Yellow Fever Vaccination Is Associated with the Genetic Background and Early Response of IFN-γ and CLEC5A. Viruses 2021, 13, 96. [Google Scholar] [CrossRef]
  17. Jolliffe, D.A.; Faustini, S.E.; Holt, H.; Perdek, N.; Maltby, S.; Talaei, M.; Greenig, M.; Vivaldi, G.; Tydeman, F.; Symons, J.; et al. Determinants of Antibody Responses to SARS-CoV-2 Vaccines: Population-Based Longitudinal Study (COVIDENCE UK). Vaccines 2022, 10, 1601. [Google Scholar] [CrossRef]
  18. Dominici, R.; Cattaneo, M.; Malferrari, G.; Archi, D.; Mariani, C.; Grimaldi, L.; Biunno, I. Cloning and functional analysis of the allelic polymorphism in the transcription regulatory region of interleukin-1α. Immunogenetics 2002, 54, 82–86. [Google Scholar] [CrossRef]
  19. Jiménez-Sousa, M.; Medrano, L.M.; Liu, P.; Almansa, R.; Fernández-Rodríguez, A.; Gómez-Sánchez, E.; Rico, L.; Heredia-Rodríguez, M.; Gómez-Pesquera, E.; Tamayo, E.; et al. IL-1Brs16944 polymorphism is related to septic shock and death. Eur. J. Clin. Investig. 2017, 47, 53–62. [Google Scholar] [CrossRef]
  20. Smith, A.J.; Humphries, S.E. Cytokine and cytokine receptor gene polymorphisms and their functionality. Cytokine Growth Factor Rev. 2009, 20, 43–59. [Google Scholar] [CrossRef]
  21. Meyer, N.J.; Ferguson, J.F.; Feng, R.; Wang, F.; Patel, P.N.; Li, M.; Xue, C.; Qu, L.; Liu, Y.; Boyd, J.H.; et al. A Functional Synonymous Coding Variant in the IL1RN Gene Is Associated with Survival in Septic Shock. Am. J. Respir. Crit. Care Med. 2014, 190, 656–664. [Google Scholar] [CrossRef]
  22. Vasilyev, F.; Silkov, A.; Sennikov, S. Relationship between interleukin-1 type 1 and 2 receptor gene polymorphisms and the expression level of membrane-bound receptors. Cell. Mol. Immunol. 2015, 12, 222–230. [Google Scholar] [CrossRef]
  23. Arimitsu, J.; Hirano, T.; Higa, S.; Kawai, M.; Naka, T.; Ogata, A.; Shima, Y.; Fujimoto, M.; Yamadori, T.; Hagiwara, K.; et al. IL-18 gene polymorphisms affect IL-18 production capability by monocytes. Biochem. Biophys. Res. Commun. 2006, 342, 1413–1416. [Google Scholar] [CrossRef]
  24. Yoo, J.K.; Kwon, H.; Khil, L.-Y.; Zhang, L.; Jun, H.-S.; Yoon, J.-W. IL-18 Induces Monocyte Chemotactic Protein-1 Production in Macrophages through the Phosphatidylinositol 3-Kinase/Akt and MEK/ERK1/2 Pathways. J. Immunol. 2005, 175, 8280–8286. [Google Scholar] [CrossRef] [PubMed]
  25. Fishman, D.; Faulds, G.; Jeffery, R.; Mohamed-Ali, V.; Yudkin, J.S.; Humphries, S.; Woo, P. The effect of novel polymorphisms in the interleukin-6 (IL-6) gene on IL-6 transcription and plasma IL-6 levels, and an association with systemic-onset juvenile chronic arthritis. J. Clin. Investig. 1998, 102, 1369–1376. [Google Scholar] [CrossRef] [PubMed]
  26. Read, R.C.; Teare, D.M.; Pridmore, A.C.; Naylor, S.C.; Timms, J.M.; Kaczmarski, E.B.; Borrow, R.; Wilson, A.G. The tumor necrosis factor polymorphism TNF (−308) is associated with susceptibility to meningococcal sepsis, but not with lethality*. Crit. Care Med. 2009, 37, 1237–1243. [Google Scholar] [CrossRef] [PubMed]
  27. Hsia, T.-C.; Chang, W.-S.; Wang, S.; Shen, T.-C.; Hsiao, W.-Y.; Liu, C.-J.; Liang, S.-J.; Chen, W.-C.; Tu, C.-Y.; Tsai, C.-W.; et al. The Contribution of Interleukin-10 Promoter Genotypes to Susceptibility to Asthma in Adults. Vivo 2015, 29, 695–699. [Google Scholar]
  28. Imran, M.; Laddha, N.; Dwivedi, M.; Mansuri, M.; Singh, J.; Rani, R.; Gokhale, R.; Sharma, V.; Marfatia, Y.; Begum, R. Interleukin-4 genetic variants correlate with its transcript and protein levels in patients with vitiligo. Br. J. Dermatol. 2012, 167, 314–323. [Google Scholar] [CrossRef]
  29. Meng, J.-F.; Rosenwasser, L.J. Unraveling the Genetic Basis of Asthma and Allergic Diseases. Allergy Asthma Immunol. Res. 2010, 2, 215–227. [Google Scholar] [CrossRef]
  30. Forte, G.I.; Scola, L.; Misiano, G.; Milano, S.; Mansueto, P.; Vitale, G.; Bellanca, F.; Sanacore, M.; Vaccarino, L.; Rini, G.B.; et al. Relevance of Gamma Interferon, Tumor Necrosis Factor Alpha, and Interleukin-10 Gene Polymorphisms to Susceptibility to Mediterranean Spotted Fever. Clin. Vaccine Immunol. 2009, 16, 811–815. [Google Scholar] [CrossRef]
  31. Hijikata, M.; Shojima, J.; Matsushita, I.; Tokunaga, K.; Ohashi, J.; Hang, N.T.L.; Horie, T.; Sakurada, S.; Hoang, N.P.; Thuong, P.H.; et al. Association of IFNGR2 gene polymorphisms with pulmonary tuberculosis among the Vietnamese. Hum. Genet. 2012, 131, 675–682. [Google Scholar] [CrossRef]
  32. Bonura, F.; Genovese, D.; Amodio, E.; Calamusa, G.; Sanfilippo, G.L.; Cacioppo, F.; Giammanco, G.M.; De Grazia, S.; Ferraro, D. Neutralizing Antibodies Response against SARS-CoV-2 Variants of Concern Elicited by Prior Infection or mRNA BNT162b2 Vaccination. Vaccines 2022, 10, 874. [Google Scholar] [CrossRef] [PubMed]
  33. Bonelli, F.; Sarasini, A.; Zierold, C.; Calleri, M.; Bonetti, A.; Vismara, C.; Blocki, F.A.; Pallavicini, L.; Chinali, A.; Campisi, D.; et al. Clinical and Analytical Performance of an Automated Serological Test That Identifies S1/S2-Neutralizing IgG in COVID-19 Patients Semiquantitatively. J. Clin. Microbiol. 2020, 58, e01224-20. [Google Scholar] [CrossRef] [PubMed]
  34. Scola, L.; Giarratana, R.; Marinello, V.; Cancila, V.; Pisano, C.; Ruvolo, G.; Frati, G.; Lio, D.; Balistreri, C. Polymorphisms of Pro-Inflammatory IL-6 and IL-1β Cytokines in Ascending Aortic Aneurysms as Genetic Modifiers and Predictive and Prognostic Biomarkers. Biomolecules 2021, 11, 943. [Google Scholar] [CrossRef]
  35. Liu, Y.; Mao, B.; Liang, S.; Yang, J.-W.; Lu, H.-W.; Chai, Y.-H.; Wang, L.; Zhang, L.; Li, Q.-H.; Zhao, L.; et al. Association between age and clinical characteristics and outcomes of COVID-19. Eur. Respir. J. 2020, 55, 2001112. [Google Scholar] [CrossRef] [PubMed]
  36. Müller, L.; Andrée, M.; Moskorz, W.; Drexler, I.; Walotka, L.; Grothmann, R.; Ptok, J.; Hillebrandt, J.; Ritchie, A.; Rabl, D.; et al. Age-dependent Immune Response to the Biontech/Pfizer BNT162b2 Coronavirus Disease 2019 Vaccination. Clin. Infect. Dis. 2021, 73, 2065–2072. [Google Scholar] [CrossRef]
  37. Collier, D.A.; Ferreira, I.A.T.M.; Kotagiri, P.; Datir, R.; Lim, E.; Touizer, E.; Meng, B.; Abdullahi, A.; CITIID-NIHR BioResource COVID-19 Collaboration; Elmer, A.; et al. Age-related immune response heterogeneity to SARS-CoV-2 vaccine BNT162b2. Nature 2021, 596, 417–422. [Google Scholar] [CrossRef]
  38. Lio, D.; Scola, L.; Giarratana, R.M.; Candore, G.; Colonna-Romano, G.; Caruso, C.; Balistreri, C.R. SARS CoV2 infection _The longevity study perspectives. Ageing Res. Rev. 2021, 67, 101299. [Google Scholar] [CrossRef]
  39. Collatuzzo, G.; Visci, G.; Violante, F.S.; Porru, S.; Spiteri, G.; Monaco, M.G.L.; Fillon, F.L.; Negro, C.; Janke, C.; Castelletti, N.; et al. Determinants of anti-S immune response at 6 months after COVID-19 vaccination in a multicentric European cohort of healthcare workers—ORCHESTRA project. Front. Immunol. 2022, 13, 986085. [Google Scholar] [CrossRef]
  40. Bánki, Z.; Seekircher, L.; Falkensammer, B.; Bante, D.; Schäfer, H.; Harthaller, T.; Kimpel, J.; Willeit, P.; von Laer, D.; Borena, W. Six-Month Follow-Up of Immune Responses after a Rapid Mass Vaccination against SARS-CoV-2 with BNT162b2 in the District of Schwaz/Austria. Viruses 2022, 14, 1642. [Google Scholar] [CrossRef]
  41. Nelli, F.; Fabbri, A.; Panichi, V.; Giannarelli, D.; Topini, G.; Berrios, J.R.G.; Virtuoso, A.; Marrucci, E.; Mazzotta, M.; Schirripa, M.; et al. Peripheral lymphocyte subset counts predict antibody response after SARS-CoV-2 mRNA-BNT162b2 vaccine in cancer patients: Results from the Vax-On-Profile study. Int. Immunopharmacol. 2022, 108, 108774. [Google Scholar] [CrossRef] [PubMed]
  42. Decru, B.; Van Elslande, J.; Steels, S.; Van Pottelbergh, G.; Godderis, L.; Van Holm, B.; Bossuyt, X.; Van Weyenbergh, J.; Maes, P.; Vermeersch, P. IgG Anti-Spike Antibodies and Surrogate Neutralizing Antibody Levels Decline Faster 3 to 10 Months After BNT162b2 Vaccination Than After SARS-CoV-2 Infection in Healthcare Workers. Front. Immunol. 2022, 13, 909910. [Google Scholar] [CrossRef] [PubMed]
  43. Mangia, A.; Serra, N.; Cocomazzi, G.; Giambra, V.; Antinucci, S.; Maiorana, A.; Giuliani, F.; Montomoli, E.; Cantaloni, P.; Manenti, A.; et al. Cellular and Humoral Immune Responses and Breakthrough Infections After Two Doses of BNT162b Vaccine in Healthcare Workers (HW) 180 Days After the Second Vaccine Dose. Front. Public Health 2022, 10, 847384. [Google Scholar] [CrossRef] [PubMed]
  44. Tahtinen, S.; Tong, A.-J.; Himmels, P.; Oh, J.; Paler-Martinez, A.; Kim, L.; Wichner, S.; Oei, Y.; McCarron, M.J.; Freund, E.C.; et al. IL-1 and IL-1ra are key regulators of the inflammatory response to RNA vaccines. Nat. Immunol. 2022, 23, 532–542. [Google Scholar] [CrossRef] [PubMed]
  45. Heo, J.Y.; Bin Seo, Y.; Kim, E.J.; Lee, J.; Kim, Y.R.; Yoon, J.G.; Noh, J.Y.; Cheong, H.J.; Kim, W.J.; Yoon, S.-Y.; et al. COVID-19 vaccine type-dependent differences in immunogenicity and inflammatory response: BNT162b2 and ChAdOx1 nCoV-19. Front. Immunol. 2022, 13, 975363. [Google Scholar] [CrossRef]
  46. Gordon, S.; Akopyan, G.; Garban, H.; Bonavida, B. Transcription factor YY1: Structure, function, and therapeutic implications in cancer biology. Oncogene 2006, 25, 1125–1142. [Google Scholar] [CrossRef]
  47. Yu, K.L.; Jung, Y.M.; Park, S.H.; Lee, S.D.; You, J.C. Human transcription factor YY1 could upregulate the HIV-1 gene expression. BMB Rep. 2020, 53, 248–253. [Google Scholar] [CrossRef]
  48. Warowicka, A.; Broniarczyk, J.; Węglewska, M.; Kwaśniewski, W.; Goździcka-Józefiak, A. Dual Role of YY1 in HPV Life Cycle and Cervical Cancer Development. Int. J. Mol. Sci. 2022, 23, 3453. [Google Scholar] [CrossRef]
  49. Hess, J.; Angel, P.; Schorpp-Kistner, M. AP-1 subunits: Quarrel and harmony among siblings. J. Cell Sci. 2004, 117, 5965–5973. [Google Scholar] [CrossRef]
  50. Wan, P.; Zhang, S.; Ruan, Z.; Liu, X.; Yang, G.; Jia, Y.; Li, Y.; Pan, P.; Wang, W.; Li, G.; et al. AP-1 signaling pathway promotes pro-IL-1β transcription to facilitate NLRP3 inflammasome activation upon influenza A virus infection. Virulence 2022, 13, 502–513. [Google Scholar] [CrossRef]
  51. Park, J.-Y.; Chung, T.-W.; Jeong, Y.-J.; Kwak, C.-H.; Ha, S.-H.; Kwon, K.-M.; Abekura, F.; Cho, S.-H.; Lee, Y.-C.; Ha, K.-T.; et al. Ascofuranone inhibits lipopolysaccharide–induced inflammatory response via NF-kappaB and AP-1, p-ERK, TNF-α, IL-6 and IL-1β in RAW 264.7 macrophages. PLoS ONE 2017, 12, e0171322. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, R.; King, A.; Tarlinton, D.; Heierhorst, J. The ASCIZ-DYNLL1 Axis Is Essential for TLR4-Mediated Antibody Responses and NF-κB Pathway Activation. Mol. Cell. Biol. 2021, 41, e0025121. [Google Scholar] [CrossRef]
  53. Yeung, Y.T.; Aziz, F.; Guerrero-Castilla, A.; Argüelles, S. Signaling Pathways in Inflammation and Anti-inflammatory Therapies. Curr. Pharm. Des. 2018, 24, 1449–1484. [Google Scholar] [CrossRef] [PubMed]
  54. Kunimoto, D.; Nordan, R.P.; Strober, W. IL-6 is a potent cofactor of IL-1 in IgM synthesis and of IL-5 in IgA synthesis. J. Immunol. 1989, 143, 2230–2235. [Google Scholar] [CrossRef] [PubMed]
  55. Kishimoto, T. Factors affecting B-cell growth and differentiation. Annu. Rev. Immunol. 1985, 3, 133–157. [Google Scholar] [CrossRef]
  56. Ma, C.S.; Deenick, E.K.; Batten, M.; Tangye, S.G. The origins, function, and regulation of T follicular helper cells. J. Exp. Med. 2012, 209, 1241–1253. [Google Scholar] [CrossRef]
  57. Wang, C.-Y.; Liang, C.-Y.; Feng, S.-C.; Lin, K.-H.; Lee, H.-N.; Shen, Y.-C.; Wei, L.-C.; Chang, C.-J.; Hsu, M.-Y.; Yang, Y.-Y.; et al. Analysis of the Interleukin-6 (-174) Locus Polymorphism and Serum IL-6 Levels with the Severity of Normal Tension Glaucoma. Ophthalmic Res. 2017, 57, 224–229. [Google Scholar] [CrossRef] [PubMed]
  58. Pedersen, S.F.; Ho, Y.-C. SARS-CoV-2: A storm is raging. J. Clin. Investig. 2020, 130, 2202–2205. [Google Scholar] [CrossRef]
  59. Wu, D.; Yang, X.O. TH17 responses in cytokine storm of COVID-19: An emerging target of JAK2 inhibitor Fedratinib. J. Microbiol. Immunol. Infect. 2020, 53, 368–370. [Google Scholar] [CrossRef]
  60. Chen, T.; Lin, Y.-X.; Zha, Y.; Sun, Y.; Tian, J.; Yang, Z.; Lin, S.-W.; Yu, F.; Chen, Z.-S.; Kuang, B.-H.; et al. A Low-Producing Haplotype of Interleukin-6 Disrupting CTCF Binding Is Protective against Severe COVID-19. Mbio 2021, 12, e0137221. [Google Scholar] [CrossRef]
  61. Kirtipal, N.; Bharadwaj, S. Interleukin 6 polymorphisms as an indicator of COVID-19 severity in humans. J. Biomol. Struct. Dyn. 2021, 39, 4563–4565. [Google Scholar] [CrossRef] [PubMed]
  62. Kaltoum, A.B.O. Mutations and polymorphisms in genes involved in the infections by covid 19: A review. Gene Rep. 2021, 23, 101062. [Google Scholar] [CrossRef] [PubMed]
  63. Linnik, J.; Egli, A. Impact of host genetic polymorphisms on vaccine induced antibody response. Hum. Vaccines Immunother. 2016, 12, 907–915. [Google Scholar] [CrossRef]
  64. Poland, G.A.; Ovsyannikova, I.G.; Jacobson, R. Immunogenetics of seasonal influenza vaccine response. Vaccine 2008, 26, D35–D40. [Google Scholar] [CrossRef] [PubMed]
  65. Kalish, R.B.; Vardhana, S.; Gupta, M.; Perni, S.C.; Witkin, S.S. Interleukin-4 and -10 gene polymorphisms and spontaneous preterm birth in multifetal gestations. Am. J. Obstet. Gynecol. 2004, 190, 702–706. [Google Scholar] [CrossRef]
  66. Sun, Y.H.; Wei, S.T.; Zong, S.H. Correlation between IL-4 gene polymorphism as well as its mRNA expression and rheumatoid arthritis. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 3879–3885. [Google Scholar]
  67. Akbari, H.; Tabrizi, R.; Lankarani, K.B.; Aria, H.; Vakili, S.; Asadian, F.; Noroozi, S.; Keshavarz, P.; Faramarz, S. The role of cytokine profile and lymphocyte subsets in the severity of coronavirus disease 2019 (COVID-19): A systematic review and meta-analysis. Life Sci. 2020, 258, 118167. [Google Scholar] [CrossRef]
  68. Mountz, J.D.; Gao, M.; Ponder, D.M.; Liu, S.; Sun, C.-W.; Alduraibi, F.; Sullivan, K.; Pat, B.; Dell’Italia, L.J.; Hsu, H.-C. IL-4 receptor blockade is a global repressor of naïve B cell development and responses in a dupilumab-treated patient. Clin. Immunol. 2022, 244, 109130. [Google Scholar] [CrossRef]
  69. Turner, M.D.; Nedjai, B.; Hurst, T.; Pennington, D.J. Cytokines and chemokines: At the crossroads of cell signalling and inflammatory disease. Biochim. et Biophys. Acta Mol. Cell Res. 2014, 1843, 2563–2582. [Google Scholar] [CrossRef]
  70. de Paula, C.B.V.; de Azevedo, M.L.V.; Nagashima, S.; Martins, A.P.C.; Malaquias, M.A.S.; Miggiolaro, A.F.R.D.S.; Júnior, J.D.S.M.; Avelino, G.; Carmo, L.A.P.D.; Carstens, L.B.; et al. IL-4/IL-13 remodeling pathway of COVID-19 lung injury. Sci. Rep. 2020, 10, 18689. [Google Scholar] [CrossRef]
  71. Hasanvand, A. COVID-19 and the role of cytokines in this disease. Inflammopharmacology 2022, 30, 789–798. [Google Scholar] [CrossRef] [PubMed]
  72. Lotfi, R.; Kalmarzi, R.N.; Roghani, S.A. A review on the immune responses against novel emerging coronavirus (SARS-CoV-2). Immunol. Res. 2021, 69, 213–224. [Google Scholar] [CrossRef] [PubMed]
  73. Rokni, M.; Sarhadi, M.; Nia, M.H.; Khosroshahi, L.M.; Asghari, S.; Sargazi, S.; Mirinejad, S.; Saravani, R. Single nucleotide polymorphisms located in TNFA, IL1RN, IL6R and IL6 genes are associated with COVID-19 risk and severity in an Iranian population. Cell Biol. Int. 2022, 46, 1109–1127. [Google Scholar] [CrossRef] [PubMed]
  74. Patel, J.A.; Nair, S.; Ochoa, E.E.; Huda, R.; Roberts, N.J.; Chonmaitree, T. Interleukin-6−174 and Tumor Necrosis Factor α−308 Polymorphisms Enhance Cytokine Production by Human Macrophages Exposed to Respiratory Viruses. J. Interf. Cytokine Res. 2010, 30, 917–921. [Google Scholar] [CrossRef] [PubMed]
  75. Stieber, F.; Allen, N.; Carpenter, K.; Hu, P.; Alagna, R.; Rao, S.; Manissero, D.; Howard, J.; Nikolayevskyy, V. Durability of COVID-19 vaccine induced T-cell mediated immune responses measured using the QuantiFERON SARS-CoV-2 assay. Pulmonology, 2022; published online ahead of print. [Google Scholar] [CrossRef] [PubMed]
Table 1. SNPs typed.
Table 1. SNPs typed.
GeneSNPPosition Minor AlleleBiological EffectReferences
IL-1Ars18005872:112785383TThe minor allele is associated with a greater production of the cytokine[18]
IL-1Brs11436342:112832813TThe minor allele is associated with a greater production of the cytokine[19]
rs169442:112837290AThe minor allele is associated with a reduced production of the cytokine[20]
IL-1RNrs3159522:113132727CMinor allele is associated with an increased efficiency in inflammation control [21]
IL-1R1rs22346502:102141867TAlleles create two alternative putative binding site for two different transcription factors[22]
IL-18rs18723811:112164265GThe minor allele is associated with a greater production of the cytokine[23]
rs194651811:112164735TThe minor allele is associated with a reduced production of the cytokine[24]
IL-6rs18007957:22727026CThe minor allele is associated with a reduced production of the cytokine[25]
TNFArs18006296:31575254AThe minor allele is associated with a greater production of the cytokine[26]
IL-10rs18008961:206773552GThe minor allele is associated with a greater production of the cytokine[27]
rs18008721:206773062AThe minor allele is associated with a reduced production of the cytokine
rs30210971:206773289TThe minor allele is associated with a reduced production of the cytokine
IL-4rs22432505:132673462TThe minor allele is associated with a greater production of the cytokine[28]
IL-13rs18009255:132657117TThe minor allele is associated with a greater production of the cytokine[29]
IFNGrs243056112:68158742TThe minor allele is associated with a greater production of the cytokine[30]
IFNGR2rs283421321:33420603GThe minor allele is associated with a greater production of the cytokine[31]
Table 2. Anti-SARS-CoV-2 antibody titers (AU/mL) mean, standard deviation (SD), standard error (SE), median, and range (Max-Min) modifications at 31 and 105 days after the administration of the second dose of anti-SARS-CoV-2 vaccine.
Table 2. Anti-SARS-CoV-2 antibody titers (AU/mL) mean, standard deviation (SD), standard error (SE), median, and range (Max-Min) modifications at 31 and 105 days after the administration of the second dose of anti-SARS-CoV-2 vaccine.
AU/mL 31 Days after 2nd AdministrationAU/mL 105 Days after 2nd AdministrationPercentage of Titer Decrease
Nr. of subjects12283
Mean259.95162.3946.37
SD156.0094.8217.37
SE14.1210.411.91
Median259.50135.0046.42
Max458.00400.0077.73
Min1.301.263.07
Table 3. Analysis of the significant differences in age distribution between different groups and subgroups of subjects.
Table 3. Analysis of the significant differences in age distribution between different groups and subgroups of subjects.
GroupsNrAge ± SDp
Study population12249.451 ± 13.419----
Ab titer ≥ median at day 316246.806 ± 13.3170.0263
Ab titer < median at day 316052.183 ± 13.077
Uninfected with Ab titer ≥ median at day 1054144.341 ± 13.6780.0226
Uninfected with Ab titer < median at day 1054250.786 ± 11.514
Uninfected at day 105 8347.602 ± 12.9630.0258
Infected at day 105 3953.385 ± 13.689
Table 4. Anti-SARS CoV-2 antibody titer mean (AU/mL) ± standard deviation (SD) at day 31 in vaccinated subjects who then developed SARS-CoV-2 infection compared with levels detected in subjects still uninfected at 105 days post-administration of BNT162b2 second doses.
Table 4. Anti-SARS CoV-2 antibody titer mean (AU/mL) ± standard deviation (SD) at day 31 in vaccinated subjects who then developed SARS-CoV-2 infection compared with levels detected in subjects still uninfected at 105 days post-administration of BNT162b2 second doses.
N.AU/mL ± SDp
Infected by day 10539167.53 ± 149.91<0.0001
Uninfected by day 10583303.38 ± 139.73
Table 5. Analysis of the significant differences in gender distribution between different groups and subgroups.
Table 5. Analysis of the significant differences in gender distribution between different groups and subgroups.
GroupsMen%Women%OR (95% CI)p
Study population5645.906654.10-------
Ab titer ≥ median at day 312419.673831.150.55 (0.27–1.14)0.146
Ab titer < median at day 313226.232822.95
Uuninfected with Ab titer ≥ median at day 1051922.892226.511.27 (0.53–3.03)0.661
Uninfected with Ab titer < median at day 1051720.482530.12
Uninfected at day 1053629.514738.531.37 (0.64–3.00)0.441
Infected at day 1052016.391915.57
Table 6. Cytokine gene polymorphisms significantly affecting anti-SARS-CoV-2 S1/S2 antibody titers measured 31 days after administration of the second dose of anti-SARS-CoV-2 mRNA vaccine.
Table 6. Cytokine gene polymorphisms significantly affecting anti-SARS-CoV-2 S1/S2 antibody titers measured 31 days after administration of the second dose of anti-SARS-CoV-2 mRNA vaccine.
Genes and AllelesDay 31OR (95% CI)p *
Titer < MedianTiter ≥ Median
NFreq.NFreq.
IL-1R1rs2234650C730.61610.491.60 (0.97–2.67)0.07
T470.39630.51
C/C160.2750.080.24 (0.08–0.71)0.0081
C/T410.68510.822.00 (0.84–4.77)0.11
T/T30.0560.12.47 (0.56–10.88)0.22
IL-6rs1800795G750.62990.80.42 (0.24–0.75)0.003
C450.38250.2
G/G270.45430.692.77 (1.31–5.81)0.011
C/G210.35130.210.49 (0.21–1.12)0.086
C/C120.260.10.53 (0.18–1.57)0.24
* adjusted by age and gender. IL-1R1 rs2234650: dominant model C/T-T/T Vs C/C, OR 4.14 (1.41–12.19) p 0.0071. IL-6 rs1800795: dominant model G/C-C/C vs G/G, OR 0.36 (0.17–0.76) p 0.017.
Table 7. Significant associations of genotypes of IL-1R1 rs2234650 and IL-4 rs2243250 with anti-SARS-CoV-2 S1/S2 antibody titers measured after 105 days from the administration of the second dose of recombinant mRNA vaccine (83 uninfected subjects).
Table 7. Significant associations of genotypes of IL-1R1 rs2234650 and IL-4 rs2243250 with anti-SARS-CoV-2 S1/S2 antibody titers measured after 105 days from the administration of the second dose of recombinant mRNA vaccine (83 uninfected subjects).
Genes and AllelesDay 105 OR (95% CI)p *
Titer < MedianTiter ≥ Median
NFreq.NFreq.
IL-1R1rs2234650C500.6400.491.54 (0.83–2.85)0.212
T340.4420.51
C/C120.2940.100.27 (0.08–0.92)0.0495
C/T260.62320.782.01 (0.74–5.47)0.175
T/T40.150.121.60 (0.37–6.86)0.52
IL-4rs2243250C680.81600.730.64 (0.31–1.33)0.270
T160.19220.27
C/C300.71220.540.46 (0.19–1.14)0.115
C/T80.19160.393.06 (1.17–8.00)0.035
T/T40.130.070.81 (0.16–4.03)0.799
* adjusted by age and gender. IL-1R1 RS2234650 C/T dominant model C/T-T/T vs C/C, OR 3.67 (1.04–12.99) p 0.033.
Table 8. Cytokine gene polymorphisms significantly associated with the detection of SARS-CoV-2 infection before day 105 post-administration of the second dose of recombinant mRNA vaccine.
Table 8. Cytokine gene polymorphisms significantly associated with the detection of SARS-CoV-2 infection before day 105 post-administration of the second dose of recombinant mRNA vaccine.
Genes and AllelesUninfected SubjectsInfected SubjectsOR (95% CI)p *
NFreq.NFreq.
IL-1R1rs2234650C900.54440.560.92 (0.53–1.57)0.784
T760.46340.44
C/C160.1950.131.62 (0.55–4.81)0.449
C/T580.70340.873.33 (1.14–9.74)0.018
T/T90.1100------
IL-6rs1800795G1280.77460.592.34 1.31–4.18)0.006
C380.23320.41
G/G530.64170.440.44 (0.20–0.95) 0.049
C/G220.27120.311.21 (0.52–2.83)0.651
C/C80.1100.262.89 (1.01–8.26)0.048
TNFArs1800629G1420.86600.771.77 (0.91–3.51)0.104
A240.14180.23
G/G600.72250.641.46 (0.65–3.290.402
G/A220.27100.260.86 (0.35–2.09)0.732
A/A10.0140.18.72 (0.92–82.84)0.031
IL-4rs2243250C1280.77490.631.99 (1.11–3.58)0.022
T380.23290.37
C/C520.63150.380.37 (0.17–0.82)0.019
C/T240.29190.492.37 (1.07–5.28)0.033
T/T70.0850.131.52 (0.44–5.25)0.511
* adjusted by age and gender. IL-6 rs1800795: dominant model G/C-C/C vs G/G, OR 2.29 (1.05–4.96), p 0.045. IL-4 rs2243250: dominant model C/T-T/T vs C/C, OR 2.67 (1.21–5.88), p 0.014.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Scola, L.; Ferraro, D.; Sanfilippo, G.L.; De Grazia, S.; Lio, D.; Giammanco, G.M. Age and Cytokine Gene Variants Modulate the Immunogenicity and Protective Effect of SARS-CoV-2 mRNA-Based Vaccination. Vaccines 2023, 11, 413. https://doi.org/10.3390/vaccines11020413

AMA Style

Scola L, Ferraro D, Sanfilippo GL, De Grazia S, Lio D, Giammanco GM. Age and Cytokine Gene Variants Modulate the Immunogenicity and Protective Effect of SARS-CoV-2 mRNA-Based Vaccination. Vaccines. 2023; 11(2):413. https://doi.org/10.3390/vaccines11020413

Chicago/Turabian Style

Scola, Letizia, Donatella Ferraro, Giuseppa Luisa Sanfilippo, Simona De Grazia, Domenico Lio, and Giovanni Maurizio Giammanco. 2023. "Age and Cytokine Gene Variants Modulate the Immunogenicity and Protective Effect of SARS-CoV-2 mRNA-Based Vaccination" Vaccines 11, no. 2: 413. https://doi.org/10.3390/vaccines11020413

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop