Skip to main content

Genetic diversity and bioinformatic analysis in the L1 gene of HPV genotypes 31, 33, and 58 circulating in women with normal cervical cytology

Abstract

Background

HPV-31, -33, and -58, along with HPV-45 and -52, account for almost 11% of HPV-associated cancers. Our previous studies showed that after HPV-16 and -51, HPV-58 was common and HPV-31 was as frequent as HPV-18 among Iranian women with normal cytology. Hence, in this study, we aimed to investigate the intra-type variations in L1 genes of HPV-58, -31, and -33 to find the predominant lineages circulating in women with normal cytology.

Methods

Complete coding sequencing of the L1 gene was amplified and nucleotide and amino acid sequences were compared to those of the references. The selective pressure on L1 protein and whether the variations of the L1 genes embed in L1 loops, or N-glycosylated sites were also investigated.

Results

B1, A, and A1 (sub)lineages were common in the HPV-58, -33, and -31 samples, respectively. Ninety nucleotide mutations were observed. Twenty nine nucleotide changes corresponded to nonsynonymous substitutions in which seventeen mutations were located in L1 loops. Only one codon position in HPV-58 sequences was found as the positive selection. No difference was observed in N-glycosylation sites between reference and understudied amino acid sequences.

Conclusion

In the current study, we reported, for the first time, the (sub) lineages, amino acid, and genetic diversity in the L1 gene of circulating HPV-58, -33, and -31, in women with normal cytology, in Iran. Such studies can not only have epidemiological values, but also aid to set vaccination programs.

Background

Human papillomaviruses (HPVs) are small, non-enveloped, double stranded DNA viruses with a genome consisting of approximately 8000 bp, which contains of early region (E6, E7, E1, E2, E4, and E5) and late region (L1, L2) [1]. HPVs cause benign lesions (warts) and cancers in humans. Up to now, more than 220 HPVs have been identified and classified into five genera of alpha, beta, gamma, mu, and nu [2], among which 14 HPV genotypes are known as high-risk (HR) HPVs (16, 18, 31, 33, 35, 39, 45, 51, 52, 56, 58, 59, 66, and 68) [3]. HR HPVs are involved in 10% of humans’ cancer. Cervical cancer, which is the third most prevalent cancer among women, results from the persistent infection of HPVs. HPVs are also the main reason for anal cancer as well as other carcinomas in vulva, vagina, and penis [4]. HPV genotypes 31, 33, and 58, along with HPV-45 and 52, account for almost 15% of cervical cancers and 11% of HPV-associated cancers which can be prevented by 9-valent vaccine [3]. This vaccine is not available in developing countries which raises the importance of four HPV genotypes other than HPV-16 and -18, which can be controlled by 2- and 4-valent vaccines.

Considering the fact that HPV types are able to diversify their genetic content from their whole genome, they can also be separated into lineages and sub-lineages, diverging from 1 to 10% and 0.5 to 1%, respectively [5]. There are three lineages in HPV-31: A, B and C; three lineages in HPV-33: A, B and C and two sub-lineages (A1 and A2); four lineages in HPV-58 (A, B, C, and D) with seven sub-lineages (A1, A2, A3, B1, B2, D1, and D2) [6, 7]. With the diversification of lineages in HPV genotypes, the prognosis of HPV infections may change significantly as lineages have different abilities in terms of developing cervical cancer, persistence and escaping the immune system [5]. Indeed, one of the effective immune evasion mechanisms developed by HPVs are mutations occurring, especially, in L1 gene, which hamper the induction of B-cell and T-cell associated immune responses [8]. With the detection of cross reactivity among HPV vaccine recipients in response to non-HPV vaccine genotypes, HPV-16 L1 loops are used as a model to investigate the immunogenic domains in other HR HPVs [9]. The predominant lineage of HPV-45 which was B as reported in our previous study, with the same criterion for sampling [10]. Also, we found B variants as a frequent lineage in HPV-52 samples (unpublished data). Thus, in the current study, we decided to report the prevalent variants of HPV-31, -33, and -58 and investigate the effect of amino acid changes on selective pressure, N-glycosylation sites, and whether these mutations were located in the five external loos; BC, DE, EF, FG, and HI.

Methods

Ethics statement

The current study was approved by the ethics committee of Iran University of medical sciences, Tehran, Iran with ID IR.IUMS.REC 1396.32308. According to Helsinki declaration, all patients were informed by a written consent prior to sampling and their information was attentively protected.

Sample collection

In this cross-sectional study, 32, 21, and 11 Thin Prep pap samples with normal cytology were included that were positive for the single infection of HPV-58, -31, and -33, respectively. Indeed, the samples were collected from women who were referred to hospitals affiliated to Iran University of medical sciences in Tehran province for routine screening tests from October 2019 to September 2021.

HPV DNA extraction and the L1 gene amplification

HPV DNA was isolated by using QIAamp DNA Mini kit (Qiagen GmbH, Germany). The positive samples were approved by High + Low PapillomaStrip kit (Opegen Kit, Spain) for HPV genotyping, according to the manufacturer’s instructions. The L1 gene in samples were amplified, using 50 µl of total PCR reaction, including 33.9 µl DNase-free water, 5 µl 10 × reaction buffer, 4 µl dNTP mix (10 mM for each), 0.8 µl of each forward and reverse specific primers (10 pmol/µl), 0.5 µl mi-Pfu DNA polymerase (metabion, Germany), and 5 µl DNA template. Also, β-globin primers were used as control and to determine the quality of DNA. The primers and thermocycler programs that used in this study are shown in Table 1.

Table 1 The primers and programs of the thermocycler with temperatures cycle

Phylogenetic trees

Multiple sequence alignment (MSA) of sequenced samples was done and phylogenetic trees were drawn by using maximum likelihood method, Kimura 2-parameter method in the bootstrap test, 1000 replicates, by Mega 10 software [11]. The sequences with Genbank accession numbers were used for each sub-lineage as references.

N-glycosylation analysis

The understudied L1 gene nucleotides were translated to amino acid sequences through an online tool of the European Bioinformatics Institute (EMBL-EBI, EMBOSS Transeq Tool) (https://www.ebi.ac.uk/Tools/st/emboss_transeq). Complete amino acid sequences of HPV-31, -33, and -58 isolates were compared with the references UniProt accession numbers P17388, P06416, and P26535, respectively, to determine the changes. Consequently, NetNGlyc 1.0 server was applied to predict N-glycosylated sites through artificial neural networks and threshold 0.5 defined by the conserved Asparagine-Xaa-Serine/Threonine sequons [14].

Selective pressure analysis and homology modeling

Datamonkey was used to detect amino acid sites under selection. This free online server detects selective pressure based on statistical methods, such as the Fixed Effects Likelihood (FEL), in which the rate of synonymous (alpha) and nonsynonymous (beta) substitution is used to estimate the negative and positive selection at each codon based on the neutral model (alpha = beta) and p value ≤ 0.1 for a given coding alignment and corresponding phylogeny [12, 13]. The homology models of the HPV genotypes L1 protein were constructed by Swiss-model [14] in comparison with the HPV-16 L1 protein.

Results

The average age of infected females with HPV-58, -31, and -33 were 33.6, 29.8, and 33, respectively. Two samples from HPV-58 and one sample from HPV-33 were excluded due to the undesirable sequencing, thus we continued the study with 30 and 10 samples for HPV-58 and HPV-33, respectively.

Phylogenetic trees

The phylogenetic trees were constructed by maximum likelihood method with Mega 10 software (Fig. 1). Four lineages were distinguished in HPV-58, A (13.3%, n = 4), B (63.3%, n = 19), C (6.7%, n = 2), and D (16.7%, n = 5); however, there were six sub-lineages among HPV-58 isolates, including A1 (10.0%, n = 3), A3 (3.3%, n = 1), B1 (50.0%, n = 15), B2 (13.3%. n = 4), D1 (3.3%, n = 1), and D2 (13.4%, n = 4). HPV-31 was divided into three lineages, A (61.9%, n = 13), B (14.3%, n = 3), and C (23.8%, n = 5). Each lineage fell into two sub-lineages, including A1 (57.1%, n = 12), A2 (4.8%, n = 1), B1 (4.8%, n = 1), B2 (9.5%, n = 2), C1 (9.5%, n = 2), and C3 (14.3%, n = 3). All HPV-33 isolates (100.0%, n = 10) accounted for lineage A (50.0% A1 and 50.0% A2).

Fig. 1
figure 1figure 1

Phylogenetic trees of three HPV genotypes 58 (A), 31 (B) and 33 (C) based on alignments of the L1 genes. The trees were constructed in Mega 10 by using maximum likelihood/Kimura 2-parameter method and 1000 bootstrap replicates and the values greater than 70% are shown above the branches. The isolates from this study are shown with small black circles and remaining accession numbers are available HPV genotypes 31, 33, and 58 sequences in the Genbank as references. All HPV genotypes of this study are available in the NCBI database and Genbank accession numbers OQ412837-82, MT267729, MZ221065-73, and MZ221053-57

Amino acid and genetic variability

HPV-31

Nucleotide sequences analysis on the L1 gene of HPV-31 showed 34 changes (Additional file 1: Table S1). The most frequent mutation, which was seen in all isolates, was C1266A, followed by C186T, in 61.9% (n = 13) of isolates. The third most frequent mutations were T370C, C821A, and G1245A, accounting for 38.1% (n = 8) of samples. Eight mutations were found in all lineage C isolates, T30A, A468G, G777A, A799G, T1035G, G1245A, C534T, and C821A; the two formers of which led to T267A and T274N missense mutations, respectively (Table 2).

Table 2 Amino acid mutations in all HPV-58, -31, and -33 isolates

HPV-33

Compared to the reference gene, the L1 sequence mutations of HPV-33 isolates showed nine variations (Additional file 1: Table S2). Among these, C797A, A1071G, and three changes C167A, G397A, and T885G occurred in 80.0% (n = 8), 70% (n = 7), and 50.0% (n = 5) of isolates, respectively. Table 2 represents all mutations in the amino acid sequences in comparison to the reference. Three replacements T266K, T56N, and G133S accounted for the most frequent amino acid changes.

HPV-58

Fourty seven nucleotides and 19 nonsynonymous mutations were observed in the L1 gene of HPV-58 isolates with 86.7% (n = 26) of each G430A, G1258A, C1263A, and A1264G, as well as G840A 83.3% (n = 25) as major substitutions (Additional file 1: Table S3). Among these variations, G430A, G1258A, and A1264G led to missense variations V144I, D420N, and N422D, respectively as predominant amino acid changes. Other variations are represented in Table 2.

N-glycosylation analysis

N- and O-linked glycosylation are two main glycosylation in which glycans attach to side chains of Asparagine and mainly Serin/Threonine, respectively. N-glycosylation is a post translational modification that is involved in myriad biological process, such as protein folding, stability, and host cell membrane-ligand interactions [15]. Compared with the references, none of mutations in studied sequences led to any changes in N-glycosylation sites (Table 3). Although no changes were found between the N-glycosylated sites in references and HPV-58 and -31 isolates, at position 54 of HPV-33 reference a proline locates just after asparagine, which is predicted as a N-glycosylation site (Table 3). Since, in most cases, a proline situated after asparagine may inhibit the N-glycosylation by rendering the asparagine inaccessible, more experimental evidence is needed to confirm this N-link glycosylation site. In HPV-33 isolates 99–1217, 99–1216, 99–1215, 99–1211, 99–1210, and 98–274, threonine was replaced by asparagine at position 56 which was the only difference between these isolates and the N-glycosylation site in the reference.

Table 3 Prediction of N-glycosylation sites

Selective analysis

During natural selection some variants can be adapted or deleted in a population. In positive selection, the variants that confer a fitness advantage are fixed in the population, while in negative selection the variants with a deleterious effect on the fitness are gradually removed [16]. Datamonkey server estimates alpha and beta substitution rates codon-by-codon under selection. A nonsynonymous replacement with a negligible effect on its fitness is classifies as neutral but a variation which leads to an increase in fitness, is considered a positive selection. Also, a purifying substitution is interpreted as a negative selection that means the mutation may gradually be removed from the genome. Tables 4, 5 and 6 indicates the codon-by-codon results from the FEL analysis in case of HPV genotypes 31, 33, and 58. Among 85 nucleotide variations which were under different selective pressures, 43.5% (n = 37) were under negative selection and 55.3% (n = 47) accounted for neutral selection. 17, 17, and 3 codons with negative (purifying) selection were observed in the HPV-58, -31, and -33 sequences, respectively. Only one positive (diversifying) selection at position 150 in the HPV-58 sequences was found. We reported all neutral variations in the case of these HPV genotypes in Additional file 1: Table S4a–c.

Table 4 The codon-by-codon results from the FEL analysis for HPV-58 isolates
Table 5 The codon-by-codon results from the FEL analysis for HPV-33 isolates
Table 6 The codon-by-codon results from the FEL analysis for HPV-31 isolates

Homology analysis on L1 loops

According to the HPV-16 major capsid protein, the pentameric L1 protein consists of five external surface loops, which are designated as BC, DE, EF, FG, and HI loop [17]. The structure of the HPV-16 L1 protein was compared with each HPV-31, HPV-33, and HPV-58 references by using homology model to predict the changes in the L1 loops (Fig. 2). Among missense variations in HPV-31 samples, T267A and T274N were in FG loop and S67L in BC loop. Out of four nonsynonymous replacements in the HPV-33 isolates, three changes T56N, G133S, and T266K were located in BC, DE, and FG loop, respectively. Among nineteen changes in the HPV-58 amino acid sequences, eleven replacements were located in DE, FG, and HI loops, including V144I, L150F, S159G, and P163T in the DE loop; K292T, A296P, D299N, and V311A/G in the FG loop; T375N, G378D, and D383N in the HI loop. Remaining mutations occurred out of the loops.

Fig. 2
figure 2

The L1 gene in reference genomes of HPV-58, -33, and -31 compared with that of HPV-16. Red = BC Loop, pink = DE Loop, blue = EF Loop, green = FG Loop, orange = HI Loop. The mutations in the understudied isolates are presented by brown boxes

Discussion

As HPV-16 and -18 comprise 70% of worldwide cervical cancers [3], in Iran like other parts of the world, most research has been conducted of these two genotypes, focusing on their prevalence, genetic variability, carcinogenicity, integration status, and lineages [18,19,20,21]. According to our previous study, HPV-52 and -58 followed by HPV-31 were the second more common types of the virus after HPV-16 [22], which indicates the importance of other HR HPVs, including HPV-31, -33, and -58. Furthermore, the L1 gene can be used for (sub)lineage classification and is critical for the induction of neutralizing antibodies [23]. Therefore, we decided to investigate the amino acid changes and genetic diversity of the L1 gene and attempted to fill the gap as the information about HPV-31, -33, and -58 in Iran is extensively restricted.

Our phylogenetic analysis illustrated that 63.3% of HPV-58 isolates fell into the lineage B, 16.7% in D, 13.3% in A, and 6.7% in C lineages. The lineage A accounted for 100.0% of HPV-33 samples. Among HPV-31 samples, 61.9%, 23.8%, and 14.3% were in lineages A, C, and B, respectively. Our findings about HPV-31 are in consistent with Hosseini’s findings that variants A and C are common in Iran [24] but to the best of our knowledge, this study is the first in Iran. Some variants of HPV-16 are more aggressive than others and even lineage-specific nucleotide changes were reported [25, 26]. Our samples were collected from females with normal pap-test; however, as shown by previous studies the (sub)lineage of other HR HPVs, like HPV-16, can be involved in the development of neoplasia. On the other hand, few studies were undertaken on the genetic variability in the L1 gene of HPV genotypes 31, 33, and 58 to identify the nucleotide changes associated with cervical cancer and more studies focused on E6 and E7 regions. For example, the association of HPV-31, -33, and -58 variants with the risk of high grade of abnormal lesions was reported. Women with A and B variants were at greater risk of HPV-31-related severe lesions [27]. Sub-lineage A1 in HPV-33 has been mostly relevant to cervical cancer [5]; however, due to the low number of understudied samples more studies are required to confirm this conclusion. Also, the A variants of HPV-58 was found to be associated with perseverance and severity of lesions [28, 29]. In addition to oncogenicity, nucleotide changes in L1 may have critical role in protein folding and self-assembly process [30], and even sensitivity to neutralizing antibodies (HPV-33).

Negative selection or purifying selection focuses on deleterious mutations [16]. In contrast, variations with positive pressure contribute to productive infection. Host factors, including immune cells, and viral proteins also influence positive selection [31] and can even be a main factor in HPV persistency. We only found one codon position in HPV-58 sequences that is under positive selection, at position 150, and no negative selection site was observed. N-glycosylation highly affects the protein stability and its third structure, especially in interaction with another protein and receptor signaling, which this modification has been mainly conserved in evolution [32]. As it was expected, no difference was observed between our sequences and those of the references.

The majority of differences among HPV-16 and other HPV genotypes are observed in the five surface-exposed loops and C-termini, against which anti bodies are produced. A comparison of the L1 amino acid sequences with the HPV-16 L1 reveals the main homology in internal regions of the L1 capsid. 41.4% (12/29) of nonsynonymous mutations of our samples occurred out of the loops and among those mutations occurring in loops (58.6%, 17/29) they were mostly located in DE and FG loops (70.6%). The FG and DE loops of L1 proteins contribute to the epitopes for cross-neutralizing antibodies. Deletion at positions 281 and 282, which are located in the late region of the FG loop, leads to the loss of cross-neutralizing antibodies, which are produced through vaccines containing HPV-16 virus-like particles (VLPs) [9]. In this study, although 70.6% (n = 12) missense mutations were located in the FG and DE loops, none of them occurred at positions 281–282. Furthermore, sub-lineages can influence the intensity of the immune responses. Godi and colleagues illustrated that HPV-33 A1 variants were more sensitive than other HPV-33 sub-lineages against the cross-neutralizing antibodies produced by 4-valent vaccine. They pointed that the occurrence of amino acid changes in the FG and DE loops of A2, B, and C (sub)lineages was the main cause for the difference in antibody responses [33]. More studies are needed to find the impact of these mutations in both FG and DE loops on the neutralizing antibodies.

This study had some limitations which should be addressed; (1) According to the report of Ministry of Health of Iran, cervical cancer did not rank in ten common cancers among Iranian women. As it was highly unlike to collect adequate positive cancerous samples for HPV-31, -33, and -58, we designed our study on samples from women with normal cytology to estimate the predominant circulating genotypes of interest in the population. However, our sample size of positive specimens for the three HPV genotypes was small. (2) Our study contained only samples with normal cytology. In contrast, some specific-lineage variations in coding sequence regions or the long control region (LCR) of HPV-16 are associated with severe lesions [25]. Thus, investigations on other regions and samples with high grade neoplasia are needed. (3) The region/gene that we selected may have an effect on sub-lineage classification. The accuracy of L1 in terms of lineage and sub-lineage classification is (97.75 and 97.75) and (100 and 48.72) for HPV-58 and -16, respectively [34], which proves L1 to be an appropriate gene for HPV-58 classification, while E2 was reported the best region for sub-lineage HPV-16 classification. Although variations in one gene is compatible with changes in HPV complete genome, different regions have different accuracy in lineage and sub-lineage classification.; therefore, more studies on different genes/regions, especially for HPV-31 and -33 is suggested to find the best region for sub-lineage classification.

Conclusion

In the current study, phylogenetic analysis of HPV-58, -33, and -31 showed B1, A, and A1 as the frequent (sub)lineages circulating among women with normal cytology. We also reported, for the first time, the genetic diversity in the L1 loops of these HPV genotypes and the effect of amino acid changes on selective pressure and N-glycosylation sites. Additional studies with larger sample size, containing cancerous samples, in different geographical regions in Iran are required to provide a reliable insight into the distribution of these HR HPV genotypes.

Availability of data and materials

The data supporting the findings of this study are available in GenBank, both the current study, and the cited references.

References

  1. Brianti P, De Flammineis E, Mercuri SR. Review of HPV-related diseases and cancers. New Microbiol. 2017;40(2):80–5.

    CAS  PubMed  Google Scholar 

  2. Center IHPHR. Human Reference clones 2020. Available from: https://www.hpvcenter.se/human_reference_clones/.

  3. Elissa Meites JG, Elizabeth Unger, and Lauri Markowitz. Human Papillomavirus November 2, 2020 [Updated October 2020]. Available from: https://www.cdc.gov/vaccines/pubs/pinkbook/hpv.html.

  4. Xu J, Tan L, Wang T, Cui F, Ding X, Wan Q, et al. Genetic variability of human papillomavirus type 51 E6, E7, L1 and L2 genes in Southwest China. Gene. 2019;690:99–112.

    Article  CAS  PubMed  Google Scholar 

  5. Siqueira JD, Alves BM, Prellwitz IM, Furtado C, Meyrelles ÂR, Machado ES, et al. Identification of novel human papillomavirus lineages and sublineages in HIV/HPV-coinfected pregnant women by next-generation sequencing. Virology. 2016;493:202–8.

    Article  CAS  PubMed  Google Scholar 

  6. Chen Z, Schiffman M, Herrero R, DeSalle R, Anastos K, Segondy M, et al. Evolution and taxonomic classification of human papillomavirus 16 (HPV16)-related variant genomes: HPV31, HPV33, HPV35, HPV52, HPV58 and HPV67. PLoS ONE. 2011;6(5): e20183.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Chen AA, Heideman DA, Boon D, Chen Z, Burk RD, De Vuyst H, et al. Human papillomavirus 33 worldwide genetic variation and associated risk of cervical cancer. Virology. 2014;448:356–62.

    Article  CAS  PubMed  Google Scholar 

  8. Yang R, Wheeler CM, Chen X, Uematsu S, Takeda K, Akira S, et al. Papillomavirus capsid mutation to escape dendritic cell-dependent innate immunity in cervical cancer. J Virol. 2005;79(11):6741–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Bissett SL, Godi A, Beddows S. The DE and FG loops of the HPV major capsid protein contribute to the epitopes of vaccine-induced cross-neutralising antibodies. Sci Rep. 2016;6(1):39730.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Kesheh MM, Shavandi S, Jahromi ZK, Esghaei M, Keyvani H. Genetic diversity and bioinformatic analysis of the HPV-45 L1 gene in Iranian women with normal cytology. Human Gene. 2022;34: 201092.

    Article  Google Scholar 

  11. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35(6):1547–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Pond SLK, Frost SDW. Datamonkey: rapid detection of selective pressure on individual sites of codon alignments. Bioinformatics. 2005;21(10):2531–3.

    Article  CAS  PubMed  Google Scholar 

  13. Kosakovsky Pond SL, Frost SDW. Not so different after all: a comparison of methods for detecting amino acid sites under selection. Mol Biol Evol. 2005;22(5):1208–22.

    Article  PubMed  Google Scholar 

  14. Schwede T, Kopp J, Guex N, Peitsch MC. SWISS-MODEL: an automated protein homology-modeling server. Nucleic Acids Res. 2003;31(13):3381–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Lee HS, Qi Y, Im W. Effects of N-glycosylation on protein conformation and dynamics: protein data bank analysis and molecular dynamics simulation study. Sci Rep. 2015;5(1):8926.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Loewe L. Negative selection. Nat Educ. 2008;1(1):59.

    Google Scholar 

  17. Chen XS, Garcea RL, Goldberg I, Casini G, Harrison SC. Structure of small virus-like particles assembled from the L1 protein of human papillomavirus 16. Mol Cell. 2000;5(3):557–67.

    Article  CAS  PubMed  Google Scholar 

  18. Niya MHK, Kesheh MM, Keshtmand G, Basi A, Rezvani H, Imanzade F, et al. Integration rates of human papilloma virus genome in a molecular survey on cervical specimens among Iranian patients. Eur J Cancer Prev. 2019;28(6):537–43.

    Article  Google Scholar 

  19. Salehi-Vaziri M, Sadeghi F, Hashemi FS, Haeri H, Bokharaei-Salim F, Monavari SH, et al. Distribution of human papillomavirus genotypes in Iranian women according to the severity of the cervical lesion. Iran Red Crescent Med J. 2016;18(4):e24458.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Hossein R, Behzad S, Tahar M, Azadeh NA. Prevalence of human papillomavirus genotypes associated with cervical and breast cancers in Iran. Monoclon Antib Immunodiagn Immunother. 2013;32(6):399–403.

    Article  CAS  PubMed  Google Scholar 

  21. Mobini Kesheh M, Barazandeh M, Kaffashi A, Shahkarami MK, Nadji SA. Genetic diversity of HPV 16 and HPV 18 based on partial long control region in Iranian women. Can J Infect Diseases Med Microbiol. 2022;2022:4759871.

    Article  Google Scholar 

  22. Mobini Kesheh M, Keyvani H. The prevalence of HPV genotypes in Iranian population: an update. Iran J Pathol. 2019;14(3):197–205.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Zhang T, Chen X, Liao G, Hu M, Xu J, Xu X. Induction of cross-neutralizing antibodies by sequential immunization with heterologous papillomavirus L1VLPs and its implications for HPV prophylactic vaccines. J Med Virol. 2020;92(12):3750–8.

    Article  CAS  Google Scholar 

  24. Hosseini N, Shoja Z, Younesi S, Shafiei-Jandaghi NZ, Jalilvand S. Lineage analysis of human papillomavirus types 31 and 45 in cervical samples of Iranian women. J Med Virol. 2021;93(6):3857–64.

    Article  CAS  PubMed  Google Scholar 

  25. Cornet I, Gheit T, Iannacone MR, Vignat J, Sylla BS, Del Mistro A, et al. HPV16 genetic variation and the development of cervical cancer worldwide. Br J Cancer. 2013;108(1):240–4.

    Article  CAS  PubMed  Google Scholar 

  26. Clifford GM, Tenet V, Georges D, Alemany L, Pavon MA, Chen Z, et al. Human papillomavirus 16 sub-lineage dispersal and cervical cancer risk worldwide: Whole viral genome sequences from 7116 HPV16-positive women. Papillomavirus Res. 2019;7:67–74.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Xi LF, Schiffman M, Koutsky LA, Hulbert A, Lee SK, DeFilippis V, et al. Association of human papillomavirus type 31 variants with risk of cervical intraepithelial neoplasia grades 2–3. Int J Cancer. 2012;131(10):2300–7.

    Article  CAS  PubMed  Google Scholar 

  28. Godínez J, Heideman DA, Gheit T, Alemany L, Snijders PJ, Tommasino M, et al. Differential presence of Papillomavirus variants in cervical cancer: an analysis for HPV33, HPV45 and HPV58. Infect Genet Evol. 2013;13:96–104.

    Article  PubMed  Google Scholar 

  29. Schiffman M, Rodriguez AC, Chen Z, Wacholder S, Herrero R, Hildesheim A, et al. A population-based prospective study of carcinogenic human papillomavirus variant lineages, viral persistence, and cervical neoplasia. Can Res. 2010;70(8):3159–69.

    Article  CAS  Google Scholar 

  30. Gurgel APAD, Chagas BS, do Amaral CM, Nascimento KCG, Leal LRS, Silva Neto JDC, et al. Prevalence of human papillomavirus variants and genetic diversity in the L1 gene and long control region of HPV16, HPV31, and HPV58 found in North-East Brazil. BioMed Res Int. 2015;2015.

  31. Nijmeijer BM, Geijtenbeek TBH. Negative and positive selection pressure during sexual transmission of transmitted founder HIV-1. Front Immunol. 2019;10:1599.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Arey BJ. The role of glycosylation in receptor signaling. Glycosylation. 2012;26(10):50262.

    Google Scholar 

  33. Godi A, Bissett SL, Miller E, Beddows S. Impact of naturally occurring variation in the human papillomavirus (HPV) 33 capsid proteins on recognition by vaccine-induced cross-neutralizing antibodies. J Gen Virol. 2017;98(7):1755–61.

    Article  CAS  PubMed  Google Scholar 

  34. Ou Z, Chen Z, Zhao Y, Lu H, Liu W, Li W, et al. Genetic signatures for lineage/sublineage classification of HPV16, 18, 52 and 58 variants. Virology. 2021;553:62–9.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors wish to special thanks to Dr. Davod Javanmard and Dr. Sareh Arjmand for helping with us on the study.

Funding

This research was supported by Iran University of Medical Sciences (Grant No. 96-03-30-32308) as a part of Ph. D thesis.

Author information

Authors and Affiliations

Authors

Contributions

The study was designed by MMK. MMK and SSh collected the samples. MMK, SSh, JA, and ME analyzed the bioinformatics parts. MMK and SSh wrote the first draft. HK supervised the study and provided material supports. SSh and JA edited the final version. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Hossein Keyvani.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1.

Single nucleotide mutations and neutral variations in the understudied HPV genotypes 31, 33, and 58 isolates.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mobini Kesheh, M., Shavandi, S., Azami, J. et al. Genetic diversity and bioinformatic analysis in the L1 gene of HPV genotypes 31, 33, and 58 circulating in women with normal cervical cytology. Infect Agents Cancer 18, 19 (2023). https://doi.org/10.1186/s13027-023-00499-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13027-023-00499-7

Keywords