Next Article in Journal
Differential Susceptibility of Fetal Retinal Pigment Epithelial Cells, hiPSC- Retinal Stem Cells, and Retinal Organoids to Zika Virus Infection
Previous Article in Journal
Fast and Sensitive Detection of Soil-Borne Cereal Mosaic Virus in Leaf Crude Extract of Durum Wheat
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mutation Profile of HPV16 L1 and L2 Genes in Different Geographic Areas

by
Dimitris Tsakogiannis
1,*,
Marios Nikolaidis
2,
Flora Zagouri
3,
Eleni Zografos
3,
Christine Kottaridi
4,
Zaharoula Kyriakopoulou
5,
Lamprini Tzioga
1,
Panayotis Markoulatos
1,
Grigoris D. Amoutzias
2 and
Garyfalia Bletsa
1
1
Research Center, Hellenic Anticancer Institute, 10680 Athens, Greece
2
Bioinformatics Laboratory, Department of Biochemistry and Biotechnology, University of Thessaly, 41500 Larissa, Greece
3
Department of Clinical Therapeutics, Alexandra Hospital, National and Kapodistrian University of Athens School of Medicine, 11528 Athens, Greece
4
Department of Genetics, Development and Molecular Biology, School of Biology, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
5
Department of Environment, School of Technology, University of Thessaly, Gaiopolis, 41500 Larissa, Greece
*
Author to whom correspondence should be addressed.
Viruses 2023, 15(1), 141; https://doi.org/10.3390/v15010141
Submission received: 18 October 2022 / Revised: 16 December 2022 / Accepted: 19 December 2022 / Published: 31 December 2022
(This article belongs to the Section Human Virology and Viral Diseases)

Abstract

:
The causal relationship between HPV and cervical cancer in association with the high prevalence of high risk HPV genotypes led to the design of HPV vaccines based on the major capsid L1 protein. In recent years, capsid protein L2 has also become a focal point in the field of vaccine research. The present review focuses on the variability of HPV16 L1 and L2 genes, emphasizing the distribution of specific amino acid changes in the epitopes of capsid proteins. Moreover, a substantial bioinformatics analysis was conducted to describe the worldwide distribution of amino acid substitutions throughout HPV16 L1, L2 proteins. Five amino acid changes (T176N, N181T; EF loop), (T266A; FG loop), (T353P, T389S; HI loop) are frequently observed in the L1 hypervariable surface loops, while two amino acid substitutions (D43E, S122P) are adjacent to L2 specific epitopes. These changes have a high prevalence in certain geographic regions. The present review suggests that the extensive analysis of the amino acid substitutions in the HPV16 L1 immunodominant loops may provide insights concerning the ability of the virus in evading host immune response in certain populations. The genetic variability of the HPV16 L1 and L2 epitopes should be extensively analyzed in a given population.

1. Introduction

Human Papillomaviruses (HPVs) are non-enveloped, double–stranded circular DNA viruses of 8 kb in size that infect basal keratinocytes of mucosal and cutaneous epithelia [1,2]. Based on the sequence similarity of the L1 gene, HPVs are grouped into five genera including Alphapapillomaviruses (alpha), Betapapillomaviruses (beta), Gammapapillomaviruses (gamma), Mupapillomaviruses (mu), and Nupapillomaviruses (nu) [3,4]. HPV genomes are further classified into intratypic variants according to the diversity of their genome sequence. Members of the same lineage display a nucleotide divergence of 1% to 10%, whereas members of the same sub-lineage display a nucleotide divergence of 0.5% to 1% [5,6,7,8]. Considering their carcinogenic potential, mucosal alpha-PVs are designated as high-risk (HR-HPV) and low-risk (LR-HPV) [3,4]. Currently, 15 HPV genotypes have been described as high-risk (HPV16, 18, 31, 33, 35, 39, 45,51, 52, 56, 58, 59, 68, 73 and 82), and 12HPV genotypes have been characterized as low-risk (HPV6, 11, 40, 42, 43, 44, 54, 61, 70, 72, 81, and CP6108) [9]. Long-term HR-HPV infection is the leading cause of cervical cancer development with HPV16 and HPV18 being identified in 70% of cervical cancer incidences worldwide [10]. It is considerable to highlight that HPV infection is also responsible for the high rates of vulvar, vaginal, anal, penile, and oropharyngeal cancers [11].
HPVs infect the basal layer of the epithelium via micro-wounds [12,13]. To initiate viral infection, the L1 capsid protein binds to herparin sulfate proteoglycans (HSPGs) located on the epithelial cell surface or on the basement membrane or it can interact with laminin-332 on the extracellular matrix [14,15]. After binding, the viral capsid undergoes cyclophilin (CyP) B-mediated conformational modifications that eventually expose the minor capsid protein L2 [14,15]. The binding of HPV to a secondary receptor is necessary for the effective viral insertion [14,16]. After viral entry the replication of HPV DNA is associated with the epithelial differentiation program [17]. First an initial replication phase preserves a stable number of viral episomes (50–100 per cell). Second, during the latent infection, the copy number of viral DNA remains stable and the viral genome is replicated along with the host cell DNA. Third, during the productive infection that occurs in the middle layers of the epitheliun the viral copy number increases guided by the different functions of E6, E7, E1 and E2 early proteins [18,19]. Over the late stage of viral infection that takes place in the upper layers of the epithelium, the viral capsid L1, L2 proteins form the new viral capsids, while the E4 protein triggers the total collapse of the epithelial cell intermediate filament, facilitating the release of newly generated viruses [20,21].
HPVs have established molecular mechanisms in order to prevent cell-cycle arrest and apoptosis. The viral E6 and E7 oncoproteins are involved in the uncontrolled cell proliferation as they inactivate tumor suppressor proteins p53 and pRB, respectively [22,23,24]. The E6 oncoprotein generates a complex with the E3 ubiquitin ligase E6-associated protein (E6AP) that subsequently triggers the polyubiquitination of p53 via the ubiquitin-mediated degradation pathway [22]. Accordingly, the E7 oncorotein interacts with pRb and releases the E2F transcription factor from the pRB/E2F suppressor complex. The disruption of the pRB/E2F complex enables the stimulation of the mitotic phase of cell cycle and consequently triggers cell proliferation [25,26]. An additional mechanism that is strongly implicated in viral carcinogenicity is the integration of HPV DNA into the host genome. The integration of viral DNA can occur through disruption of the E1, E2, L2 or L1 genes, whereas the E6 and E7 genes always remain intact and integrated into the host chromosome [27,28,29]. The viral integration causes additional chromosomal damages and genome destabilization in the infected cells, whereas the E6 and E7 oncogenes are extensively expressed [28,30,31,32,33].
The genome of HPV16 is organized in three major regions; the long control region (LCR), the early region that encodes for the early genes E1, E2, E4, E5, E6 and E7, and the late region that encodes for the late genes L1 and L2 [27,28,34]. A large-scale evolutionary analysis enabled the classification of HPV16 into four major variant intratypic lineages (A–D) and 16 sub-lineages (A1–4, B1–4, C1–4 and D1–4) [6,7,35]. The prevalence of the different HPV16 lineages is associated with the geographic origin, ethnicity and tumorigenic capacity of viral DNA [6]. Specific nucleotide variations and amino acid substitutions have been extensively associated with specific HPV16 lineages, whereas particular mutations of HPV16 DNA have been extensively investigated and related with a higher risk of severe dysplasia and cervical cancer development, including E6: Q14H, H78Y, L83V, Ε7: N29S, S63F, E2: H35Q, P219S, T310K and E5: I65V [36]. Notably, a recent bioinformatics tool (HPV16-Genotyper) was designed to facilitate the rapid identification of these cancer-related mutations simplifying the entire computational process of HPV16 lineage genotyping and the detection of potential intratypic recombination events or genomic mis-assemblies due to co-infections [37].
Cervical cancer is the fourth most common type of cancer and the second leading cause of cancer-related death in women in developing countries [38]. Cytological testing (the Papanicolaou and the Pap test) and HPV screening have substantially reduced the incidence of cervical cancer; however, it is still a major global public health issue [28,39]. In developed countries, HPV vaccination programs have significantly reduced the incidences of HPV infection [40,41]. The prophylactic HPV vaccines are produced based on the ability of HPV L1 major capsid proteins to be self-organized into empty capsid like structures, called virus-like particles (VLPs) [42,43]. HPV vaccines confer high immunogenicity due to structural characteristics of L1 VLPs which are responsible for the production of long-lasting antigen-specific antibody-generating cells [44,45,46]. HPV vaccines trigger the formation of neutralizing antibodies against the specific HPV vaccine genotypes [45,46]. Moreover, previous studies have proposed that intratypic HPV variants may exhibit novel epitopes specific to geographic locations and/or ethnic groups [47,48]. The accumulation of non-synonymous mutations in the L1 protein can modify protein structure and function and subsequently provide the virus with a selective advantage to escape immune response [47,48].
In the present study, we review the non-synonymous nucleotide variations that have been reported in the L1 and L2 capsid genes of HPV16 DNA with a special focus on those which are located in antigenic epitopes. Moreover, the impact of the L2 minor protein on prevention of viral transmission is also discussed. Finally, we performed a large-scale analysis of publicly available HPV16 L1 and L2 gene sequences in order to identify the global distribution of particular amino acid changes which may affect the aggression of the virus.

2. Types of HPV Vaccines and Mechanism of Antibody Protection

Currently, the US Food and Drug Administration (FDA) has approved three prophylactic HPV vaccines. In 2006, the FDA first approved Gardasil (Merck), which provides protection against four HPV genotypes (HPV 6, 11, 16, 18). Next, Cervarix (GlaxoSmithKline) was approved by the FDA in 2009, and it provides immunity against HPV16 and HPV18. Finally, in 2014 the FDA approved a supplemental use of Gardasil 9 (HPV nine—valent vaccine) expanding its administration in women and men of 27 to 45 years of age. Gardasil 9 protects against HPV 6, 11, 16, 18, 31, 33, 45, 52 and 58 [49]. The design of the currently available vaccines is based on virus-like particles (VLPs) that are generated by the major capsid protein L1 and mimics the structure of virions [50,51,52,53]. VLPs do not contain viral genome and they are considered safer than attenuated or inactivated viruses that could be turned into infectious ones [54,55].
>The HPV VLPs have a highly ordered structure that exposes the L1 epitopes, promoting T-cell activation. In particular, the major histocompatibility complex class II molecules (MHC-II) that are expressed in the antigen-presenting cells (APCs) enable the presentation of the L1 epitopes to CD4+T cells. CD4+T cells in turn secrete cytokines, thus leading to the activation of B-cells, T-cells and macrophages [56,57,58,59] (Figure 1). HPV vaccination generates high-quality and prolonged antibody titers which are mostly immunoglobulin (Ig) G against the respective HPV L1 proteins [60,61]. Notably, HPV vaccination produces 10- to 100-fold higher titers of L1 specific neutralizing antibodies compared to natural infection, thus preventing viral infection and subsequently the development of premalignant dysplasias [61]. Intramuscular vaccination produces antibodies which are transferred to the site of the viral infection (cervix, vulva) by two mechanisms. First, IgG can traverse the epithelial layer into mucosal secretions through the neonatal Fc receptor that is expressed at the cervix [62] (Figure 1). Nevertheless, this mechanism may play a secondary role in vaccine protection, since L1 specific neutralizing antibodies in the cervical mucus of vaccinated women are 10 to 100 fold lower compared to serum concentrations [63]. The second and the most important mechanism includes the direct transfer of interstitial antibodies at the site of injury in the basal epithelial cell layer, thus blocking viral infection.
Although the development of HPV vaccines has been successful, the design of a vaccine against a wide range of HPV genotypes is difficult, since the L1 gene demonstrates significant variability among different HPV genotypes. On the other hand, the L2 gene is more conserved among HPV genotypes, thus it may also be considered for the design of future vaccines [64,65,66,67,68]. Antibodies against the L2 protein could provide a broad protection against various HPV genotypes, though the L2 protein alone cannot form VLPs and the antibody titers against it are considerably low [64,65,66,67,68]. Recent analyses have focused on the insertion of specific L2 peptides to bacteriophage coat proteins (MS2, PP7 or AP205) (45–46). The generation of chimeric bacteriophage coat proteins and their assembly into VLPs is a promising approach that provides strong protection against numerous HPV genotypes [69,70]. In particular, previous studies involved immunized mice with a MS2-L2 VLP that exposes a concatemer peptide comprising amino acids 17–31 from HPV16 L2 and a peptide containing amino acids 20–31 from HPV31 L2 in combination with a MS2-L2 VLP that exposes a consensus L2 peptide containing amino acids 69–86 [69,70]. This combination offered wide protection against various HPV genotypes including HPV 11, 16, 18, 31, 33, 35, 39, 45, 52, 53, 56, and 58 [69,70].
A special focus has been given to the HPV16 L2 region between amino acids 17 and 36, since it has been proposed that the respective region acts as a consensus sequence that could trigger the generation of antibodies against a broad range of HPV genotypes including HPV 6, 11, 16, 18, 26, 31, 33, 34, 35, 39, 43, 44, 45, 51, 52, 53, 56, 58, 59, 66, 68, and 73 [64]. In particular, the region between amino acids 17–36 from the HPV16 L2 gene is 90% similar to a consensus sequence of amino acid 17–36 from this region [64]. The immunization with the HPV16 L2 peptide (aa 17–36) seems to simulate the immunization with a consensus peptide, thus making the respective region a promising candidate for future vaccines [64]. As a result, the high conservation of the minor capsid protein L2 among different HPV genotypes constitutes for the respective protein an appealing target for the development of next generation vaccines that could provide protection against a wider range of HPV genotypes.

3. Genetic Variability of the HPV16 L1 Gene

The HPV16 L1 gene is located between nucleotides 5559 and 7154 and it encodes the major capsid protein L1, which is 531 amino acids in size. The L2 gene is located between nucleotides 4235 and 5656 and it encodes the minor capsid protein L2 of 473 amino acids in size. The HPV capsid upon progeny virion formation is assembled in a T = 7 icosahedral symmetry around viral DNA via L1–L2 protein interaction [71]. Cryo-electron microscopy reconstructions and antibody binding analyses showed that the major capsid protein forms 72 pentamers that comprise a total of 360 copies of the L1 molecule, while the minor L2 capsid protein is included into the L1 pentamers [72,73]. The L1–L2 protein interaction is hydrophobic. Specifically, the L2 protein is inserted into the central lumen of the L1 pentamer, whereas an N-terminal “external loop” of approximately 60 amino acids of L2 proteins seems to be disclosed on the capsid surface [74,75]
Intratypic HPV16 variants share more than 98% L1 sequence similarity with the reference sequence of the viral genome (reference HPV16 genome NC_001526) [72,76]. Considering the HPV16 L1 gene, specific nucleotide sequence variations have been associated with particular HPV16 lineages. More specifically, the A1–A3 sub-lineages harbor the T6861C variation; the A4 sub-lineage harbors the G7059A variation; the B lineage harbors 12 variations (G5697A, C5863T, T5910C, C6164A, T6246C, A6315G, C6558T, T6567A, G6720A, C6853T, C6969T, G6993A, G7059T); the C lineage harbors 15 variations (G5697A, C5863T, T5910C, C6164A, T6246C, A6315G, T6481C, C6558T, A6694C, G6720A, C6853T, C6864T, C6969T, G6993A, G7059T); the D lineage harbors 15 variations (G5697A, C5863T, T5910C, C6164A, T6246C, A6315G, C6558T, A6694C, G6720A, A6802T, C6853T, C6864T, C6969T, G6993A, and G7059T) [77].
The analysis of the L1 gene revealed that non-synonymous mutations were situated adjacent to the neutralizing epitopes, thus offering capsid L1 protein a considerable advantage to avoid immune response [78]. In particular, five L1 hypervariable surface loops of 10–30 amino acids in size [DC loop (aa 50–69), DE loop (aa 110–153), EF loop (aa 160–189), FG loop (aa 262–291) and HI loop (aa 348–396)] have been characterized through crystallography, and it has been proposed that they are recognized by human antibodies during immune response [79,80]. Although these hypervariable loops are exposed to circulating human antibodies, they have a crucial impact on capsid formation, as well. In particular, the DE and FG loops are responsible for the interaction of the L1 with the proline rich region of the L2, while the EF loop is essential for the generation of L1-L1 di-sulphidic bonds [81,82]. As a result, the amino-acid changes accumulated in the respective regions can influence not only the immunological host response but they can also modify the L1 protein structure and subsequently the efficient assembly of L1, L2 proteins into viral particles [83,84].
Research data collected from different studies revealed that numerous amino acid changes are found throughout the L1 ORF, while a considerable variability is detected near or within the region of the L1 gene that encodes for the respective immunodominant loops. In particular, a total of twenty-seven non synonymous mutations in L1 gene were recorded in patients from the Netherlands, while twenty one of these variations are located in the region that encodes for the immunodominant FG loop [85]. Nevertheless, the impact of these changes in immunological response is yet to be clarified. Interestingly, a previous analysis that was performed in India revealed seven non-synonymous nucleotide changes, namely C6163A (T202N), G6171A (A205T), C6240G (H228D), A6432G (T292A), G6693A (T379P) C6863T (P435L), and G7058T (L500F). The role of these sequence changes in the L1 epitopes were predicted in silico and their impact on immunogenicity was evaluated in vivo. According to these findings, it was proposed that L500F exhibits a 2.7-fold (p < 0.002) increase in antibody titer, while T379P presents a 0.4 fold decrease in antibody titer [47]. In addition, a previous analysis that was conducted in Morocco revealed a total of three non-synonymous nucleotide changes, A6694C (T389P; HI loop), G6800A (M424I;H2 helix) and G6818A (M430A; H2 helix), while 3D prediction models revealed that these changes do not affect the overall structure of the viral protein [48]. Moreover, it was found that the amino acid substitution T389P (HI loop) may influence the susceptibility of the respective L1 mutant to the vaccine induced immunity [48]. Finally, deletions and insertions have also been observed in the L1 sequence. In particular, an ATC insertion at position 6901 (H4 helix) and a GAT deletion at position 6950 (BC loop) are prevalent in all the examined cases from India, Morocco and Southwest China [47,48,86]. Nevertheless, 3D prediction models showed that these sequence changes do not modify the L1 protein structure [48].
A previous analysis in the Indian population revealed a total of 20 non synonymous mutations, nine of which are located in the immunodominant loops including N56T (BC loop), T176N, V178G, A179T, N181T/I (EF loop), T266A, N285T (FG loop) and T353P (HI loop).Three-dimensional prediction models suggest that the amino acid mutations L158F, A179T, K236T, T266A, S296R, T353P, S396P, and K454T might influence the stability of the protein structure, while the amino acid mutations N56T, H76Y, N92T, T176N, V178G, N181T/I, N285T, T389S, K443Q, and L474F might prevent the structure of the L1 monomer [87].
A more recent analysis that examined all the available complete and partial sequences of the HPV16 L1 gene demonstrated that the vast majority of nucleotide polymorphisms are accumulated in the DE (27.38%) and FG (31%) loops, while four mutations are frequently detected in other loops such as T176N (EF loop), N181T (EF loop), A266T (FG loop), and T353P/I/N (HI loop) [88]. Notably, the vast majority of L1 mutations located at the FG loop are found mainly in Europe, while mutations in the EF and HI loops are widely distributed in Asia. In contrast, sequence changes that were recorded in the DC and DE loops exhibit no significant association with the ethnicity of cases [88]. Considering research data from clinical and bioinformatics analyses, we postulated that different nucleotide changes in the L1 gene are prevalent in diverse geographic populations, thus affecting the sufficient immune response of hosts and consequently the development of cervical cancer. As a consequence, an extensive record of the most frequent amino acid substitutions of the HPV16 L1 protein could provide targeted improvements to vaccines specific to each geographic population.

4. Genetic Variability of the HPV16 L2 Gene

Although the L1 gene has been the focus of many studies, little is known concerning the variability of the HPV16 L2 gene. The role of the minor L2 capsid protein in viral pathogenicity is regarded as critical, since it is implicated not only in viral particle assembly and viral entry, but also ensures the efficient transport of viral DNA in the nucleus of infected cells and is involved in host immune response as well [76,89,90]. As was mentioned above, the N-terminal “external loop” of the L2 protein is exposed on the viral capsid surface, which is highly conserved, while it contains cross-neutralizing epitopes that are recognized by either neutralizing or cross-neutralizing antibodies [74,91,92]. Specific epitopes have been characterized, including the amino acid residues aa 17–36, aa 56–81, aa 65–81, and aa 108–120 [67,93,94,95,96,97]. However, the capsid L2 protein is regarded as immunologically subdominant to L1 [98,99].
Certain sequence variations within the HPV16 L2 gene are related to the different lineages. In particular, the A1–A3 sub-lineages harbor two variations (C4724T, G5235A); the A4 sub-lineage harbors four variations (G5044A, A5225C, C5368T, A5517C); the B lineage harbors 18 variations (T4280C, G4307A, G4427A, G4460A, T4599C, T4640C, T4643A, A4910T, G5141A, G5235A, A5289G, T5309C, G5378A, G5388A, T5402C, T5494C, G5505A, C5563G); the C lineage contains 22 variations (T4280C, G4427T, A4517G, T4544G, T4556G, T4599C, T4643A, C4853T, A4886G, G5141A, G5235A, A5258G, A5289C, T5309C, C5368T, G5378A, G5388A, T5402C, C5486T, T5494C, G5505A, C5563G); and the D lineage contains 28 variations (T4280C, G4427T, T4451C, G4460A, A4598C, T4599C, T4643A, A4886G, A4943G, A4949G, A4968G, A5033T, A5151C, G5235A, T5285A, A5294C, T5309A, T5367G, C5368T, G5378A, T5385G, G5388A, T5402C, T5474A, C5486T, T5494C, G5505A, C5563G) [77].
The analyses of the HPV16 L2 gene in clinical samples are rather limited. However, it has been observed that the L2 gene is polymorphic, whereas the region that encodes for the N-terminal domain of the L2 protein was highly conserved. In particular, a previous analysis that was carried out in patients from Southwest China revealed nine non-synonymous nucleotide variations, while no sequence changes in the residues 65–78 and 108–120 were detected [100]. In addition, in the Indian population, a total of 38 non-synonymous variations were observed in HPV16 positive malignant cases [101]. Interestingly, a more recent study in the Indian population revealed 16 novel amino acid substitutions in the HPV16 L2 protein with T245A, L266F, S378V and S384A found to be significantly associated with severe dysplasia [102]. Finally, a study that was performed in Thai women revealed a total of seven non-synonymous sequence variations, while no mutations were recorded in the N-terminal domain [103]. The L2 protein plays a pivotal role in viral function and structure. In addition, its N-terminal domain is highly conserved among different HPV16 strains.

5. The Frequency of non-Synonymous Changes within the HPV16 L1 and L2 Proteins across Different Populations

In order to perform a sequence analysis of publicly available sequence data, we searched the NCBI taxonomy database for Human papillomavirus type 16 (txid;333760) and retrieved 11,018 (10 June 2022) nucleotide records. These records were filtered based on the availability of country of isolation, thus resulting in 9658 sequences. These nucleotide sequences were subsequently scanned by the BLASTn algorithm [104] using the L1 and L2 CDS of the HPV16 type strain (NC_001526) as queries (e-value cutoff: 1e-10) resulting in 5013 and 4172 sequences, respectively. The sequences for each gene were translated and only those that had known amino acids in at least 90% of the reference protein length were kept, resulting in 3697 and 2059 sequences for the L1 and L2 genes, respectively. The final protein sequences were aligned using the MUSCLE multiple alignment algorithm [105] embedded in the Seaview software [106]. The consensus sequences were calculated for each of the final protein alignments with Jalview [107]. The number of sequences retrieved from the different geographic regions is presented in Table 1.
A global amino acid mutation analysis of HPV16 L1 revealed a total of seven high frequency amino acid substitutions (H76Y, T176N, N181T, T266A, T353P, T389S, L474F) (63). The location of the respective mutations in the L1 protein as well as their distribution among different geographic locations is summarized in Table 2 and Figure 2. Interestingly, the mutations H76Y (N-terminal domain), T176N (EF loop), N181T (EF loop), T353P (HI loop), and L474F (C-terminal domain) are most frequently detected in South America followed by Asia, North America and Europe (Table 2, Figure 2). Moreover, the T266A mutation (FG loop) was most frequent in North America (34%), followed by Europe (22%), South America (9.1%) and Asia (5.3%) (Figure 2), while amino acid substitution T389S (HI loop) is most frequently found in South America (22%), followed by North America (6%), Asia (2.8%), and Europe (0.9%) (Table 2, Figure 2). The DC loop and the DE loop seems to be more conserved, as the frequency of amino acid substitutions at the corresponding positions was less than 0.1% in all geographic locations (see supplementary file).
The large-scale sequence analysis of the L2 gene revealed that the specific L2 epitopes (aa 17–36, aa 56–81, aa 65–81, aa 108–120) appear to be conserved among the different HPV16 isolates globally (Figure 2). However, a total of 17 amino acid mutations (D43E, S122P, V243I, T245A, L266F/V, S269P, L330F, D334N, T351P/S, T352P/A, S378V/F, S384A, V385I, I420T, A424T, I428L, A443G) were found with high frequency in the L2 protein (see supplementary file). The location of the respective mutations in the L2 protein and their distribution in various geographic areas is summarized in Figure 2. Although the N-terminal domain of the L2 protein was found to be highly conserved, two amino acid substitutions, D43E and S122P, exhibit high distribution among HPV16 isolates. These changes were found to be adjacent to L2 epitopes (Table 2, Figure 2). Considering their geographic distribution, it was revealed that D43E is observed exclusively in Asia (Table 2, Figure 2). In addition, the amino acid substitution S122P is most common in South America (46% of isolates), but is also observed in North America (17%), Asia (11%), and Europe (5.1%) (Table 2, Figure 2).
The large-scale sequence analysis of the L2 gene revealed that the specific L2 epitopes (aa 17–36, aa 56–81, aa 65–81, aa 108–120) appear to be conserved among the different HPV16 isolates globally (Figure 2). However, a total of 17 amino acid mutations (D43E, S122P, V243I, T245A, L266F/V, S269P, L330F, D334N, T351P/S, T352P/A, S378V/F, S384A, V385I, I420T, A424T, I428L, A443G) were found with high frequency in the L2 protein (see supplementary file). The location of the respective mutations in the L2 protein and their distribution in various geographic areas is summarized in Figure 2. Although the N-terminal domain of the L2 protein was found to be highly conserved, two amino acid substitutions, D43E and S122P, exhibit high distribution among HPV16 isolates. These changes were found to be adjacent to L2 epitopes (Table 2, Figure 2). Considering their geographic distribution, it was revealed that D43E is observed exclusively in Asia (Table 2, Figure 2). In addition, the amino acid substitution S122P is most common in South America (46% of isolates), but is also observed in North America (17%), Asia (11%), and Europe (5.1%) (Table 2, Figure 2).
The residual mutations of the HPV16 L2 protein are detected in the central region and the C-terminal domain (Figure 2). Notably, the amino acid substitutions T351P and I428L were characterized as population-specific. Specifically, the amino acid change T351P was predominant in South America (11.4%), but was rarely detected elsewhere. The amino acid mutation I428L was present almost uniquely in Asia, accounting for 26% of Asian isolates (Table 2, Figure 2), and the frequency of S269P and L330F was higher than that of the reference amino acid. In particular, at position 269, proline is detected more frequently than the reference amino acid serine. This phenomenon was only observed in Asian isolates (see supplementary file). Moreover, at position 330 phenylalanine is more common than the reference amino acid leucine in Europe, Asia, and North America (see supplementary file).

6. Discussion

Τhe US Food and Drug Administration (FDA) has approved a total of three prophylactic HPV vaccines based on the L1 protein. Until June 2020, 107 out of 194 WHO member states have started national HPV vaccination programs, whereas it has been estimated that in 2019 approximately 15% of girls and 4% of boys have been fully vaccinated [108]. The WHO Cervical Cancer Elimination Strategy aims to expand HPV vaccination to 90% of all adult women, double lifetime cervical screening to 70% and treatment of cervical intraepithelial lesions and cervical cancer to 90% by 2030 [109]. Currently available vaccines contain virus-like particles (VLPs) that are formed by the major capsid protein L1, imitating the structure of virions [50,51,52,53]. In recent years, a special focus has been given to the capsid protein L2 in the field of HPV vaccines. In the present review, we describe the variability in the HPV16 late genes (L1, L2) according to publicly available sequence data collected from different populations. Moreover, a large-scale sequence analysis was conducted in order to report the worldwide distribution of amino acid substitutions throughout HPV16 L1 and L2 proteins as well as to investigate whether these changes are found in specific immunodominant regions of viral capsid proteins.
Various amino acid substitutions have been recorded within or adjacent to the HPV16 L1 immunodominant loops [47,48,85,86,88]. However, the impact of these changes on immunological response has not been clarified thus fur. Although it has been proven that the amino acid changes L500F and T379P (HI loop) influence antibody titers in the Indian population, there are no available data concerning the role of these changes in different geographic populations [47]. Considering our large-scale analysis, we found seven amino acid substitutions (H76Y, T176N, N181T, T266A, T353P, T389S, L474F) that exhibit high frequency in the L1 protein globally. Five of these amino acid changes (T176N, N181T; EF loop), (T266A; FG loop), (T353P, T389S; HI loop) are of particular interest since they are located within the L1 hypervariable surface loops and they are preferentially detected in specific geographic areas [88]. In particular, the T176N, N181T, and T353P are predominant in America and Asia, the T266A was found in 22–33% of isolates from Europe and America, while the T389S is specific to the Americas (Table 2). It is worthy of note that recent 3D prediction models suggested that the amino acid substitutions T176N, N181T and T389S might be involved in stabilization of the L1 monomer structure, whereas the amino acid changes T266A and T353P seem to affect the stability of the L1 pentamer [87]. In particular, the T266A mutation leads to the loss of two hydrogen bonds with the amino acids at positions 360 and 361 of the adjacent chain, while the amino acid substitution T353P results in the loss of one hydrogen bond with the amino acid at position 26 of the adjacent pentamer [87]. The impact of these changes on the host-immune response requires further investigation, while more analyses are necessary to better assess whether these mutations enable an escape from L1 vaccines.
On the other hand, the L2 gene is considered as a polymorphic region, although its N-terminal domain is highly conserved [100,101,103]. Our large-scale analysis confirmed these observations. Moreover, we detected a total of 17 L2 amino acid substitutions that are frequently found in various geographic locations. Interestingly, two of these changes (D43E, S122P) were found adjacent to L2 epitopes and they had high frequency in specific geographic regions. In particular, the D43E is highly specific to Asians, while the S122P is predominant in specific geographic regions, including South America (46%), North America (17%), Asia (11%), and Europe (5.1%) (Table 2). More recent findings suggested that the S122P is common to the epitope region recognized by MHC-I and MHC-II, according to B and T cell epitope prediction models [102]. However, the role of these changes in the tertiary structure of viral proteins as well as their impact on immunological response requires further investigation.
It is of note that the amino acid changes that were found next to or within the immunodominant regions of L1 and L2 proteins are highly distributed in specific geographic areas. This variability might influence the potential of certain HPV16 strains to avoid immunological response in some populations, providing HPV16 strains with a selective advantage to establish persistent infection, thus leading to more severe dysplasia. However, further analyses of these amino acid changes need to be performed in order to confirm this hypothesis.
Highly conserved regions of the HPV16 L1, L2 proteins have been extensively characterized, including the DC loop, DE loop of the L1 protein, the L2 specific epitopes and the N-terminal domain of the L2 protein [48,88,100,101,103]. The high conservation of these regions might aid the virus to maintain the integrity of the viral capsid. Moreover, it is concluded that these highly conserved regions might be suitable targets for future vaccines. It is important to underline that HPV16 L2 amino acid region 17–36 is highly conserved not only among the different HPV16 isolates, but also among different HPV genotypes [64,110]. Therefore, L2 should be considered as an appealing target for future broad-range vaccines [64,110]. A limitation of the present analysis was the small number of the available HPV16 L1 and L2 sequences in certain geographic regions. Hence, more studies in the future with more available data would be essential to provide a thorough overview of L1 and L2 gene diversity in these geographic areas, underlying mutations that could change the L1 and L2 protein structure and consequently influence the specific immunodominant regions.

7. Conclusions

To the best of our knowledge, this is the first extensive literature review and large-scale sequence analysis concerning the global variability of both HPV16 L1 and L2 genes. We presented important data concerning the variability of the late region of HPV16 DNA. Our observations suggest that the high prevalence of certain amino acid substitutions within or adjacent to specific immunodominant regions might influence the stability of the viral capsid as well as modify the ability of HPV16 in evading host—immune response in specific populations. The extensive mapping of amino acid substitutions in the HPV16 L1 and L2 epitopes should be cautiously considered in a given population.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/v15010141/s1, Supplementary file: Distribution of amino acid substitutions in the complete amino acid sequence of HPV16 L1 and L2 proteins in different geographic locations.

Author Contributions

Conceptualization D.T., G.D.A. and P.M.; Methodology, D.T., G.D.A., M.N., L.T., F.Z., E.Z., Z.K. and C.K.; Validation, D.T. and G.D.A.; formal analysis D.T., G.D.A. and M.N.; writing—original draft preparation D.T. and G.B.; writing—review and editing D.T., G.B., G.D.A. and P.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is included in the manuscript.

Acknowledgments

The authors thank the president of the Hellenic Anticancer Institute, M. Vrontakis, and the members of the Board for their continuous support.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Zur Hausen, H. Papillomavirus infections—A major cause of human cancers. Biochim. Biophys. Acta 1996, 1288, F55–F78. [Google Scholar] [CrossRef]
  2. Van Doorslaer, K.; Chen, Z.; Bernard, H.-U.; Chan, P.K.S.; DeSalle, R.; Dillner, J.; Forslund, O.; Haga, T.; McBride, A.A.; Villa, L.L.; et al. ICTV Virus Taxonomy Profile: Papillomaviridae. J. Gen. Virol. 2018, 99, 989–990. [Google Scholar] [CrossRef]
  3. Bernard, H.-U.; Burk, R.D.; Chen, Z.; Van Doorslaer, K.; zur Hausen, H.; de Villiers, E.-M. Classification of papillomaviruses (PVs) based on 189 PV types and proposal of taxonomic amendments. Virology 2010, 401, 70–79. [Google Scholar] [CrossRef] [Green Version]
  4. Bernard, H.-U.; Calleja-Macias, I.E.; Dunn, S.T. Genome variation of human papillomavirus types: Phylogenetic and medical implications. Int. J. Cancer 2006, 118, 1071–1076. [Google Scholar] [CrossRef]
  5. Chen, Z.; Schiffman, M.; Herrero, R.; DeSalle, R.; Anastos, K.; Segondy, M.; Sahasrabuddhe, V.V.; Gravitt, P.E.; Hsing, A.W.; Burk, R.D. Evolution and Taxonomic Classification of Human Papillomavirus 16 (HPV16)-Related Variant Genomes: HPV31, HPV33, HPV35, HPV52, HPV58 and HPV67. PLoS ONE 2011, 6, e20183. [Google Scholar] [CrossRef]
  6. Burk, R.D.; Harari, A.; Chen, Z. Human papillomavirus genome variants. Virology 2013, 445, 232–243. [Google Scholar] [CrossRef] [Green Version]
  7. Mirabello, L.; Clarke, M.A.; Nelson, C.W.; Dean, M.; Wentzensen, N.; Yeager, M.; Cullen, M.; Boland, J.F.; NCI HPV Workshop; Schiffman, M.; et al. The Intersection of HPV Epidemiology, Genomics and Mechanistic Studies of HPV-Mediated Carcinogenesis. Viruses 2018, 10, 80. [Google Scholar] [CrossRef] [Green Version]
  8. Mirabello, L.; Yeager, M.; Yu, K.; Clifford, G.M.; Xiao, Y.; Zhu, B.; Cullen, M.; Boland, J.F.; Wentzensen, N.; Nelson, C.W.; et al. HPV16 E7 Genetic Conservation Is Critical to Carcinogenesis. Cell 2017, 170, 1164–1174.e6. [Google Scholar] [CrossRef] [Green Version]
  9. Muñoz, N.; Bosch, F.X.; De Sanjosé, S.; Herrero, R.; Castellsagué, X.; Shah, K.V.; Snijders, P.J.F.; Meijer, C.J.L.M. Epidemiologic Classification of Human Papillomavirus Types Associated with Cervical Cancer. N. Engl. J. Med. 2003, 348, 518–527. [Google Scholar] [CrossRef] [Green Version]
  10. Li, Y.; Xu, C. Human Papillomavirus-Related Cancers. Adv. Exp. Med. Biol. 2017, 1018, 23–34. [Google Scholar] [CrossRef]
  11. Alhamlan, F.S.; Alfageeh, M.B.; Al Mushait, M.A.; Al-Badawi, I.A.; Al-Ahdal, M.N. Human Papillomavirus-Associated Cancers. Adv. Exp. Med. Biol. 2021, 1313, 1–14. [Google Scholar] [CrossRef]
  12. Kines, R.C.; Thompson, C.D.; Lowy, D.R.; Schiller, J.T.; Day, P.M. The initial steps leading to papillomavirus infection occur on the basement membrane prior to cell surface binding. Proc. Natl. Acad. Sci. USA 2009, 106, 20458–20463. [Google Scholar] [CrossRef] [Green Version]
  13. Schiller, J.T.; Day, P.M.; Kines, R.C. Current understanding of the mechanism of HPV infection. Gynecol. Oncol. 2010, 118, S12–S17. [Google Scholar] [CrossRef] [Green Version]
  14. Raff, A.B.; Woodham, A.W.; Raff, L.M.; Skeate, J.G.; Yan, L.; Da Silva, D.M.; Schelhaas, M.; Kast, W.M. The Evolving Field of Human Papillomavirus Receptor Research: A Review of Binding and Entry. J. Virol. 2013, 87, 6062–6072. [Google Scholar] [CrossRef] [Green Version]
  15. Gheit, T. Mucosal and Cutaneous Human Papillomavirus Infections and Cancer Biology. Front. Oncol. 2019, 9, 355. [Google Scholar] [CrossRef] [Green Version]
  16. Horvath, C.A.; Boulet, G.A.; Renoux, V.M.; Delvenne, P.O.; Bogers, J.-P.J. Mechanisms of cell entry by human papillomaviruses: An overview. Virol. J. 2010, 7, 11. [Google Scholar] [CrossRef] [Green Version]
  17. White, E.A. Manipulation of Epithelial Differentiation by HPV Oncoproteins. Viruses 2019, 11, 369. [Google Scholar] [CrossRef] [Green Version]
  18. Gyöngyösi, E.; Szalmás, A.; Ferenczi, A.; Póliska, S.; Kónya, J.; Veress, G. Transcriptional regulation of genes involved in keratinocyte differentiation by human papillomavirus 16 oncoproteins. Arch. Virol. 2015, 160, 389–398. [Google Scholar] [CrossRef]
  19. McBride, A.A. Oncogenic human papillomaviruses. Philos. Trans. R. Soc. B Biol. Sci. 2017, 372, 20160273. [Google Scholar] [CrossRef] [Green Version]
  20. Tsakogiannis, D.; Ruether, I.G.A.; Kyriakopoulou, Z.; Pliaka, V.; Skordas, V.; Gartzonika, C.; Levidiotou-Stefanou, S.; Markoulatos, P. Molecular and phylogenetic analysis of the HPV 16 E4 gene in cervical lesions from women in Greece. Arch. Virol. 2012, 157, 1729–1739. [Google Scholar] [CrossRef]
  21. Doorbar, J. The E4 protein; structure, function and patterns of expression. Virology 2013, 445, 80–98. [Google Scholar] [CrossRef] [Green Version]
  22. Martinez-Zapien, D.; Ruiz, F.X.; Poirson, J.; Mitschler, A.; Ramirez, J.; Forster, A.; Cousido-Siah, A.; Masson, M.; Vande Pol, S.; Podjarny, A.; et al. Structure of the E6/E6AP/p53 complex required for HPV-mediated degradation of p53. Nature 2016, 529, 541–545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Olmedo-Nieva, L.; Muñoz-Bello, J.O.; Contreras-Paredes, A.; Lizano, M. The Role of E6 Spliced Isoforms (E6*) in Human Papillomavirus-Induced Carcinogenesis. Viruses 2018, 10, 45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Aarthy, M.; Kumar, D.; Giri, R.; Singh, S.K. E7 oncoprotein of human papillomavirus: Structural dynamics and inhibitor screening study. Gene 2018, 658, 159–177. [Google Scholar] [CrossRef] [PubMed]
  25. Dick, F.A.; Goodrich, D.W.; Sage, J.; Dyson, N.J. Non-canonical functions of the RB protein in cancer. Nat. Rev. Cancer 2018, 18, 442–451. [Google Scholar] [CrossRef]
  26. Tsakogiannis, D.; Moschonas, G.D.; Daskou, M.; Stylianidou, Z.; Kyriakopoulou, Z.; Kottaridi, C.; Dimitriou, T.; Gartzonika, C.; Markoulatos, P. Polymorphic variability in the exon 19 of the RB1 gene and its flanking intronic sequences in HPV16-associated precancerous lesions in the Greek population. J. Med. Microbiol. 2018, 67, 1638–1644. [Google Scholar] [CrossRef]
  27. Tsakogiannis, D.; Kyriakopoulou, Z.; Ruether, I.G.A.; Amoutzias, G.; Dimitriou, T.; Diamantidou, V.; Kotsovassilis, C.; Markoulatos, P. Determination of human papillomavirus 16 physical status through E1/E6 and E2/E6 ratio analysis. J. Med. Microbiol. 2014, 63, 1716–1723. [Google Scholar] [CrossRef] [PubMed]
  28. Tsakogiannis, D.; Gartzonika, C.; Levidiotou-Stefanou, S.; Markoulatos, P. Molecular approaches for HPV genotyping and HPV-DNA physical status. Expert Rev. Mol. Med. 2017, 19, e1. [Google Scholar] [CrossRef]
  29. Tsakogiannis, D.; Gortsilas, P.; Kyriakopoulou, Z.; Ruether, I.G.A.; Dimitriou, T.G.; Orfanoudakis, G.; Markoulatos, P. Sites of disruption within E1 and E2 genes of HPV16 and association with cervical dysplasia. J. Med. Virol. 2015, 87, 1973–1980. [Google Scholar] [CrossRef]
  30. Akagi, K.; Li, J.; Broutian, T.R.; Padilla-Nash, H.; Xiao, W.; Jiang, B.; Rocco, J.W.; Teknos, T.N.; Kumar, B.; Wangsa, D.; et al. Genome-wide analysis of HPV integration in human cancers reveals recurrent, focal genomic instability. Genome Res. 2014, 24, 185–199. [Google Scholar] [CrossRef]
  31. Tsakogiannis, D.; Bletsa, M.; Kyriakopoulou, Z.; Dimitriou, T.; Kotsovassilis, C.; Panotopoulou, E.; Markoulatos, P. Identification of rearranged sequences of HPV16 DNA in precancerous and cervical cancer cases. Mol. Cell. Probes 2016, 30, 6–12. [Google Scholar] [CrossRef] [PubMed]
  32. Kamal, M.; Lameiras, S.; Deloger, M.; Morel, A.; Vacher, S.; Lecerf, C.; Dupain, C.; Jeannot, E.; Girard, E.; Baulande, S.; et al. Human papilloma virus (HPV) integration signature in Cervical Cancer: Identification of MACROD2 gene as HPV hot spot integration site. Br. J. Cancer 2021, 124, 777–785. [Google Scholar] [CrossRef] [PubMed]
  33. Shen-Gunther, J.; Cai, H.; Wang, Y. HPV Integration Site Mapping: A Rapid Method of Viral Integration Site (VIS) Analysis and Visualization Using Automated Workflows in CLC Microbial Genomics. Int. J. Mol. Sci. 2022, 23, 8132. [Google Scholar] [CrossRef] [PubMed]
  34. Tsakogiannis, D.; Darmis, F.; Gortsilas, P.; Ruether, I.G.A.; Kyriakopoulou, Z.; Dimitriou, T.G.; Amoutzias, G.; Markoulatos, P. Nucleotide polymorphisms of the human papillomavirus 16 E1 gene. Arch. Virol. 2014, 159, 51–63. [Google Scholar] [CrossRef]
  35. Schiffman, M.; Rodriguez, A.C.; Chen, Z.; Wacholder, S.; Herrero, R.; Hildesheim, A.; Desalle, R.; Befano, B.; Yu, K.; Safaeian, M.; et al. A Population-Based Prospective Study of Carcinogenic Human Papillomavirus Variant Lineages, Viral Persistence, and Cervical Neoplasia. Cancer Res. 2010, 70, 3159–3169. [Google Scholar] [CrossRef] [Green Version]
  36. Bletsa, G.; Zagouri, F.; Amoutzias, G.D.; Nikolaidis, M.; Zografos, E.; Markoulatos, P.; Tsakogiannis, D. Genetic variability of the HPV16 early genes and LCR. Present and future perspectives. Expert Rev. Mol. Med. 2021, 23, e19. [Google Scholar] [CrossRef]
  37. Nikolaidis, M.; Tsakogiannis, D.; Bletsa, G.; Mossialos, D.; Kottaridi, C.; Iliopoulos, I.; Markoulatos, P.; Amoutzias, G.D. HPV16-Genotyper: A Computational Tool for Risk-Assessment, Lineage Genotyping and Recombination Detection in HPV16 Sequences, Based on a Large-Scale Evolutionary Analysis. Diversity 2021, 13, 497. [Google Scholar] [CrossRef]
  38. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  39. Berman, T.A.; Schiller, J.T. Human papillomavirus in cervical cancer and oropharyngeal cancer: One cause, two diseases. Cancer 2017, 123, 2219–2229. [Google Scholar] [CrossRef]
  40. Bruni, L.; Diaz, M.; Barrionuevo-Rosas, L.; Herrero, R.; Bray, F.; Bosch, F.X.; de Sanjosé, S.; Castellsagué, X. Global estimates of human papillomavirus vaccination coverage by region and income level: A pooled analysis. Lancet Glob. Health 2016, 4, e453–e463. [Google Scholar] [CrossRef]
  41. Torre, L.A.; Bray, F.; Siegel, R.L.; Ferlay, J.; Lortet-Tieulent, J.; Jemal, A. Global cancer statistics, 2012. CA Cancer J. Clin. 2015, 65, 87–108. [Google Scholar] [CrossRef] [Green Version]
  42. Safaeian, M.; Porras, C.; Pan, Y.; Kreimer, A.; Schiller, J.T.; Gonzalez, P.; Lowy, D.R.; Wacholder, S.; Schiffman, M.; Rodriguez, A.C.; et al. Durable Antibody Responses Following One Dose of the Bivalent Human Papillomavirus L1 Virus-Like Particle Vaccine in the Costa Rica Vaccine Trial. Cancer Prev. Res. 2013, 6, 1242–1250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Safaeian, M.; Sampson, J.N.; Pan, Y.; Porras, C.; Kemp, T.J.; Herrero, R.; Quint, W.; van Doorn, L.J.; Schussler, J.; Lowy, D.R.; et al. Durability of Protection Afforded by Fewer Doses of the HPV16/18 Vaccine: The CVT Trial. J. Natl. Cancer. Inst. 2018, 110, 205–212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Amanna, I.J.; Slifka, M.K. Mechanisms that determine plasma cell lifespan and the duration of humoral immunity. Immunol. Rev. 2010, 236, 125–138. [Google Scholar] [CrossRef] [PubMed]
  45. Schiller, J.; Lowy, D. Explanations for the high potency of HPV prophylactic vaccines. Vaccine 2018, 36, 4768–4773. [Google Scholar] [CrossRef]
  46. Schiller, J.T.; Castellsagué, X.; Garland, S.M. A Review of Clinical Trials of Human Papillomavirus Prophylactic Vaccines. Vaccine 2012, 30 (Suppl. 5), F123–F138. [Google Scholar] [CrossRef] [Green Version]
  47. Kumar, A.; Hussain, S.; Sharma, G.; Mehrotra, R.; Gissmann, L.; Das, B.C.; Bharadwaj, M. Identification and validation of immunogenic potential of India specific HPV-16 variant constructs: In-silico & in-vivo insight to vaccine development. Sci. Rep. 2015, 5, 15751. [Google Scholar] [CrossRef] [Green Version]
  48. El-Aliani, A.; El Alaoui, M.A.; Chaoui, I.; Ennaji, M.M.; Attaleb, M.; El Mzibri, M. Naturally occurring capsid protein variants L1 of human papillomavirus genotype 16 in Morocco. Bioinformation 2017, 13, 241–248. [Google Scholar] [CrossRef] [Green Version]
  49. Hirth, J. Disparities in HPV vaccination rates and HPV prevalence in the United States: A review of the literature. Hum. Vaccines Immunother. 2019, 15, 146–155. [Google Scholar] [CrossRef]
  50. Day, P.M.; Gambhira, R.; Roden, R.B.S.; Lowy, D.R.; Schiller, J.T. Mechanisms of Human Papillomavirus Type 16 Neutralization by L2 Cross-Neutralizing and L1 Type-Specific Antibodies. J. Virol. 2008, 82, 4638–4646. [Google Scholar] [CrossRef]
  51. The GlaxoSmithKline Vaccine HPV-007 Study Group; Romanowski, B.; de Borba, P.C.; Naud, P.S.; Roteli-Martins, C.M.; De Carvalho, N.S.; Teixeira, J.C.; Aoki, F.; Ramjattan, B.; Shier, R.M.; et al. Sustained efficacy and immunogenicity of the human papillomavirus (HPV)-16/18 AS04-adjuvanted vaccine: Analysis of a randomised placebo-controlled trial up to 6.4 years. Lancet 2009, 374, 1975–1985. [Google Scholar] [CrossRef] [PubMed]
  52. Harper, D.M.; Franco, E.L.; Wheeler, C.M.; Moscicki, A.-B.; Romanowski, B.; Roteli-Martins, C.M.; Jenkins, D.; Schuind, A.; Costa Clemens, S.A.; Dubin, G.; et al. Sustained efficacy up to 4.5 years of a bivalent L1 virus-like particle vaccine against human papillomavirus types 16 and 18: Follow-up from a randomised control trial. Lancet 2006, 367, 1247–1255. [Google Scholar] [CrossRef] [Green Version]
  53. Mao, C.; Koutsky, L.A.; Ault, K.A.; Wheeler, C.M.; Brown, D.R.; Wiley, D.J.; Alvarez, F.B.; Bautista, O.M.; Jansen, K.U.; Barr, E. Efficacy of Human Papillomavirus-16 Vaccine to Prevent Cervical Intraepithelial Neoplasia: A randomized controlled trial. Obstet. Gynecol. 2006, 107, 18–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Roldao, A.; Mellado, M.C.M.; Castilho, L.R.; Carrondo, M.J.; Alves, P.M. Virus-like particles in vaccine development. Expert Rev. Vaccines 2010, 9, 1149–1176. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, J.W.; Roden, R.B. Virus-like particles for the prevention of human papillomavirus-associated malignancies. Expert Rev. Vaccines 2013, 12, 129–141. [Google Scholar] [CrossRef] [Green Version]
  56. Chackerian, B.; Lenz, P.; Lowy, D.R.; Schiller, J.T. Determinants of Autoantibody Induction by Conjugated Papillomavirus Virus-Like Particles. J. Immunol. 2002, 169, 6120–6126. [Google Scholar] [CrossRef] [Green Version]
  57. Lenz, P.; Day, P.M.; Pang, Y.-Y.S.; Frye, S.A.; Jensen, P.N.; Lowy, D.R.; Schiller, J.T. Papillomavirus-Like Particles Induce Acute Activation of Dendritic Cells. J. Immunol. 2001, 166, 5346–5355. [Google Scholar] [CrossRef] [Green Version]
  58. Lenz, P.; Thompson, C.D.; Day, P.M.; Bacot, S.M.; Lowy, D.R.; Schiller, J.T. Interaction of papillomavirus virus-like particles with human myeloid antigen-presenting cells. Clin. Immunol. 2003, 106, 231–237. [Google Scholar] [CrossRef]
  59. Couture, A.; Garnier, A.; Docagne, F.; Boyer, O.; Vivien, D.; Le-Mauff, B.; Latouche, J.-B.; Toutirais, O. HLA-Class II Artificial Antigen Presenting Cells in CD4+T Cell-Based Immunotherapy. Front. Immunol. 2019, 10, 1081. [Google Scholar] [CrossRef]
  60. Paavonen, J.; Jenkins, D.; Bosch, F.X.; Naud, P.; Salmerón, J.; Wheeler, C.M.; Chow, S.-N.; Apter, D.L.; Kitchener, H.C.; Castellsague, X.; et al. Efficacy of a prophylactic adjuvanted bivalent L1 virus-like-particle vaccine against infection with human papillomavirus types 16 and 18 in young women: An interim analysis of a phase III double-blind, randomised controlled trial. Lancet 2007, 369, 2161–2170. [Google Scholar] [CrossRef]
  61. Stanley, M. HPV-immune response to infection and vaccination. Infect. Agents Cancer 2010, 5, 19. [Google Scholar] [CrossRef] [Green Version]
  62. Li, Z.; Palaniyandi, S.; Zeng, R.; Tuo, W.; Roopenian, D.C.; Zhu, X. Transfer of IgG in the female genital tract by MHC class I-related neonatal Fc receptor (FcRn) confers protective immunity to vaginal infection. Proc. Natl. Acad. Sci. USA 2011, 108, 4388–4393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Nardelli-Haefliger, D.; Wirthner, D.; Schiller, J.T.; Lowy, D.R.; Hildesheim, A.; Ponci, F.; De Grandi, P. Specific Antibody Levels at the Cervix During the Menstrual Cycle of Women Vaccinated With Human Papillomavirus 16 Virus-like Particles. J. Natl. Cancer Inst. 2003, 95, 1128–1137. [Google Scholar] [CrossRef] [PubMed]
  64. Yadav, R.; Zhai, L.; Tumban, E. Virus-like Particle-Based L2 Vaccines against HPVs: Where Are We Today? Viruses 2019, 12, 18. [Google Scholar] [CrossRef] [Green Version]
  65. Huber, B.; Wang, J.; Roden, R.; Kirnbauer, R. RG1-VLP and Other L2-Based, Broad-Spectrum HPV Vaccine Candidates. J. Clin. Med. 2021, 10, 1044. [Google Scholar] [CrossRef] [PubMed]
  66. Schellenbacher, C.; Huber, B.; Skoll, M.; Shafti-Keramat, S.; Roden, R.; Kirnbauer, R. Incorporation of RG1 epitope into HPV16L1-VLP does not compromise L1-specific immunity. Vaccine 2019, 37, 3529–3534. [Google Scholar] [CrossRef] [PubMed]
  67. Schellenbacher, C.; Kwak, K.; Fink, D.; Shafti-Keramat, S.; Huber, B.; Jindra, C.; Faust, H.; Dillner, J.; Roden, R.B.; Kirnbauer, R. Efficacy of RG1-VLP Vaccination against Infections with Genital and Cutaneous Human Papillomaviruses. J. Investig. Dermatol. 2013, 133, 2706–2713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Schellenbacher, C.; Roden, R.; Kirnbauer, R. Chimeric L1-L2 Virus-like Particles as Potential Broad-Spectrum Human Papillomavirus Vaccines. J. Virol. 2009, 83, 10085–10095. [Google Scholar] [CrossRef] [Green Version]
  69. Ong, H.K.; Tan, W.S.; Ho, K.L. Virus like particles as a platform for cancer vaccine development. Peerj 2017, 5, e4053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Zhai, L.; Yadav, R.; Kunda, N.K.; Anderson, D.; Bruckner, E.; Miller, E.K.; Basu, R.; Muttil, P.; Tumban, E. Oral immunization with bacteriophage MS2-L2 VLPs protects against oral and genital infection with multiple HPV types associated with head & neck cancers and cervical cancer. Antivir. Res. 2019, 166, 56–65. [Google Scholar] [CrossRef]
  71. Conway, M.J.; Meyers, C. Replication and Assembly of Human Papillomaviruses. J. Dent. Res. 2009, 88, 307–317. [Google Scholar] [CrossRef] [PubMed]
  72. Buck, C.B.; Cheng, N.; Thompson, C.D.; Lowy, D.R.; Steven, A.C.; Schiller, J.T.; Trus, B.L. Arrangement of L2 within the Papillomavirus Capsid. J. Virol. 2008, 82, 5190–5197. [Google Scholar] [CrossRef] [Green Version]
  73. Cardone, G.; Moyer, A.L.; Cheng, N.; Thompson, C.D.; Dvoretzky, I.; Lowy, D.R.; Schiller, J.T.; Steven, A.C.; Buck, C.B.; Trus, B.L. Maturation of the Human Papillomavirus 16 Capsid. mBio 2014, 5, e01104-14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Kondo, K.; Ishii, Y.; Ochi, H.; Matsumoto, T.; Yoshikawa, H.; Kanda, T. Neutralization of HPV16, 18, 31, and 58 pseudovirions with antisera induced by immunizing rabbits with synthetic peptides representing segments of the HPV16 minor capsid protein L2 surface region. Virology 2007, 358, 266–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Kawana, Y.; Kawana, K.; Yoshikawa, H.; Taketani, Y.; Yoshiike, K.; Kanda, T. Human Papillomavirus Type 16 Minor Capsid Protein L2 N-Terminal Region Containing a Common Neutralization Epitope Binds to the Cell Surface and Enters the Cytoplasm. J. Virol. 2001, 75, 2331–2336. [Google Scholar] [CrossRef] [Green Version]
  76. Broniarczyk, J.; Massimi, P.; Pim, D.; Bergant Marusic, M.; Myers, M.P.; Garcea, R.L.; Banks, L. Phosphorylation of Human Papillomavirus Type 16 L2 Contributes to Efficient Virus Infectious Entry. J. Virol. 2019, 93. [Google Scholar] [CrossRef] [Green Version]
  77. Yamada, T.; Wheeler, C.M.; Halpern, A.L.; Stewart, A.C.; Hildesheim, A.; Jenison, S.A. Human papillomavirus type 16 variant lineages in United States populations characterized by nucleotide sequence analysis of the E6, L2, and L1 coding segments. J. Virol. 1995, 69, 7743–7753. [Google Scholar] [CrossRef] [Green Version]
  78. Da Silva, D.M.; Pastrana, D.V.; Schiller, J.T.; Kast, W. Effect of Preexisting Neutralizing Antibodies on the Anti-tumor Immune Response Induced by Chimeric Human Papillomavirus Virus-like Particle Vaccines. Virology 2001, 290, 350–360. [Google Scholar] [CrossRef] [Green Version]
  79. Bishop, B.; Dasgupta, J.; Klein, M.; Garcea, R.L.; Christensen, N.D.; Zhao, R.; Chen, X.S. Crystal Structures of Four Types of Human Papillomavirus L1 Capsid Proteins: Understanding the specificity of neutralizing monoclonal antibodies. J. Biol. Chem. 2007, 282, 31803–31811. [Google Scholar] [CrossRef] [Green Version]
  80. Olcese, V.A.; Chen, Y.; Schlegel, R.; Yuan, H. Characterization of HPV16 L1 loop domains in the formation of a type-specific, conformational epitope. BMC Microbiol. 2004, 4, 11–29. [Google Scholar] [CrossRef]
  81. Carter, J.J.; Wipf, G.C.; Benki, S.F.; Christensen, N.D.; Galloway, D.A. Identification of a Human Papillomavirus Type 16-Specific Epitope on the C-Terminal Arm of the Major Capsid Protein L1. J. Virol. 2003, 77, 11625–11632. [Google Scholar] [CrossRef] [Green Version]
  82. Bissett, S.L.; 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, 39730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Kirnbauer, R.; Taub, J.; Greenstone, H.; Roden, R.; Dürst, M.; Gissmann, L.; Lowy, D.R.; Schiller, J.T. Efficient self-assembly of human papillomavirus type 16 L1 and L1-L2 into virus-like particles. J. Virol. 1993, 67, 6929–6936. [Google Scholar] [CrossRef] [Green Version]
  84. Ellis, J.; Keating, P.; Baird, J.; Hounsell, E.F.; Renouf, D.V.; Rowe, M.; Hopkins, D.; Duggan-Keen, M.; Bartholomew, J.; Young, L.; et al. The association of an HPV16 oncogene variant with HLA-B7 has implications for vaccine design in cervical cancer. Nat. Med. 1995, 1, 464–470. [Google Scholar] [CrossRef] [PubMed]
  85. King, A.J.; Sonsma, J.A.; Vriend, H.J.; van der Sande, M.A.B.; Feltkamp, M.C.; Boot, H.J.; Koopmans, M.P.G. on behalf of the Medical Microbiological Laboratories and Municipal Health Services. Genetic Diversity in the Major Capsid L1 Protein of HPV-16 and HPV-18 in the Netherlands. PLoS ONE 2016, 11, e0152782. [Google Scholar] [CrossRef] [Green Version]
  86. Cao, M.; Chenzhang, Y.; Ding, X.; Zhang, Y.; Jing, Y.; Chen, Z. Genetic variability and lineage phylogeny of human papillomavirus type-16 and -53 based on the E6, E7, and L1 genes in Southwest China. Gene 2016, 592, 49–59. [Google Scholar] [CrossRef]
  87. Mane, A.; Patil, L.; Limaye, S.; Nirmalkar, A.; Kulkarni-Kale, U. Characterization of major capsid protein (L1) variants of Human papillomavirus type 16 by cervical neoplastic status in Indian women: Phylogenetic and functional analysis. J. Med. Virol. 2020, 92, 1303–1308. [Google Scholar] [CrossRef] [PubMed]
  88. El Aliani, A.; El-Abid, H.; Kassal, Y.; Khyatti, M.; Attaleb, M.; Ennaji, M.M.; El Mzibri, M. HPV16 L1 diversity and its potential impact on the vaccination-induced immunity. Gene 2020, 747, 144682. [Google Scholar] [CrossRef]
  89. Day, P.M.; Baker, C.C.; Lowy, D.R.; Schiller, J.T. Establishment of papillomavirus infection is enhanced by promyelocytic leukemia protein (PML) expression. Proc. Natl. Acad. Sci. USA 2004, 101, 14252–14257. [Google Scholar] [CrossRef] [Green Version]
  90. Day, P.M.; Roden, R.B.S.; Lowy, D.R.; Schiller, J.T. The Papillomavirus Minor Capsid Protein, L2, Induces Localization of the Major Capsid Protein, L1, and the Viral Transcription/Replication Protein, E2, to PML Oncogenic Domains. J. Virol. 1998, 72, 142–150. [Google Scholar] [CrossRef]
  91. Bossis, I.; Roden, R.B.S.; Gambhira, R.; Yang, R.; Tagaya, M.; Howley, P.; Meneses, P.I. Interaction of tSNARE Syntaxin 18 with the Papillomavirus Minor Capsid Protein Mediates Infection. J. Virol. 2005, 79, 6723–6731. [Google Scholar] [CrossRef] [Green Version]
  92. Pastrana, D.V.; Gambhira, R.; Buck, C.B.; Pang, Y.-Y.S.; Thompson, C.D.; Culp, T.D.; Christensen, N.D.; Lowy, D.R.; Schiller, J.T.; Roden, R.B. Cross-neutralization of cutaneous and mucosal Papillomavirus types with anti-sera to the amino terminus of L2. Virology 2005, 337, 365–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Lowe, J.; Panda, D.; Rose, S.; Jensen, T.; Hughes, W.A.; Tso, F.Y.; Angeletti, P.C. Evolutionary and structural analyses of alpha-papillomavirus capsid proteins yields novel insights into L2 structure and interaction with L1. Virol. J. 2008, 5, 150. [Google Scholar] [CrossRef] [Green Version]
  94. Gambhira, R.; Karanam, B.; Jagu, S.; Roberts, J.N.; Buck, C.B.; Bossis, I.; Alphs, H.; Culp, T.; Christensen, N.D.; Roden, R.B.S. A Protective and Broadly Cross-Neutralizing Epitope of Human Papillomavirus L2. J. Virol. 2007, 81, 13927–13931. [Google Scholar] [CrossRef] [Green Version]
  95. Slupetzky, K.; Gambhira, R.; Culp, T.D.; Shafti-Keramat, S.; Schellenbacher, C.; Christensen, N.D.; Roden, R.B.; Kirnbauer, R. A papillomavirus-like particle (VLP) vaccine displaying HPV16 L2 epitopes induces cross-neutralizing antibodies to HPV11. Vaccine 2007, 25, 2001–2010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Jagu, S.; Kwak, K.; Karanam, B.; Huh, W.K.; Damotharan, V.; Chivukula, S.V.; Roden, R.B.S. Optimization of Multimeric Human Papillomavirus L2 Vaccines. PLoS ONE 2013, 8, e55538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kawana, K.; Matsumoto, K.; Yoshikawa, H.; Taketani, Y.; Kawana, T.; Yoshiike, K.; Kanda, T. A Surface Immunodeterminant of Human Papillomavirus Type 16 Minor Capsid Protein L2. Virology 1998, 245, 353–359. [Google Scholar] [CrossRef] [Green Version]
  98. Karanam, B.; Jagu, S.; Huh, W.K.; Roden, R.B.S. Developing vaccines against minor capsid antigen L2 to prevent papillomavirus infection. Immunol. Cell Biol. 2009, 87, 287–299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Chabeda, A.; van Zyl, A.R.; Rybicki, E.P.; Hitzeroth, I.I. Substitution of Human Papillomavirus Type 16 L2 Neutralizing Epitopes Into L1 Surface Loops: The Effect on Virus-Like Particle Assembly and Immunogenicity. Front. Plant Sci. 2019, 10, 779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Yue, Y.; Yang, H.; Wu, K.; Yang, L.; Chen, J.; Huang, X.; Pan, Y.; Ruan, Y.; Zhao, Y.; Shi, X.; et al. Genetic Variability in L1 and L2 Genes of HPV-16 and HPV-58 in Southwest China. PLoS ONE 2013, 8, e55204. [Google Scholar] [CrossRef]
  101. Bhattacharjee, B.; Mandal, N.R.; Roy, S.; Sengupta, S. Characterization of sequence variations within HPV16 isolates among Indian women: Prediction of causal role of rare non-synonymous variations within intact isolates in cervical cancer pathogenesis. Virology 2008, 377, 143–150. [Google Scholar] [CrossRef] [PubMed]
  102. Mane, A.; Limaye, S.; Patil, L.; Kulkarni-Kale, U. Genetic variability in minor capsid protein (L2 gene) of human papillomavirus type 16 among Indian women. Med. Microbiol. Immunol. 2022, 211, 153–160. [Google Scholar] [CrossRef]
  103. Lurchachaiwong, W.; Junyangdikul, P.; Payungporn, S.; Chansaenroj, J.; Sampathanukul, P.; Tresukosol, D.; Termrungruanglert, W.; Theamboonlers, A.; Poovorawan, Y. Entire genome characterization of human papillomavirus type 16 from infected Thai women with different cytological findings. Virus Genes 2009, 39, 30–38. [Google Scholar] [CrossRef]
  104. Camacho, C.; Coulouris, G.; Avagyan, V.; Ma, N.; Papadopoulos, J.; Bealer, K.; Madden, T.L. BLAST+: Architecture and applications. BMC Bioinform. 2009, 10, 421. [Google Scholar] [CrossRef] [Green Version]
  105. Edgar, R.C. MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004, 32, 1792–1797. [Google Scholar] [CrossRef] [Green Version]
  106. Gouy, M.; Guindon, S.; Gascuel, O. SeaView Version 4: A Multiplatform Graphical User Interface for Sequence Alignment and Phylogenetic Tree Building. Mol. Biol. Evol. 2010, 27, 221–224. [Google Scholar] [CrossRef] [Green Version]
  107. Waterhouse, A.M.; Procter, J.B.; Martin, D.M.A.; Clamp, M.; Barton, G.J. Jalview Version 2—A multiple sequence alignment editor and analysis workbench. Bioinformatics 2009, 25, 1189–1191. [Google Scholar] [CrossRef] [Green Version]
  108. Bruni, L.; Saura-Lázaro, A.; Montoliu, A.; Brotons, M.; Alemany, L.; Diallo, M.S.; Afsar, O.Z.; LaMontagne, D.S.; Mosina, L.; Contreras, M.; et al. HPV vaccination introduction worldwide and WHO and UNICEF estimates of national HPV immunization coverage 2010–2019. Prev. Med. 2021, 144, 106399. [Google Scholar] [CrossRef]
  109. Canfell, K.; Kim, J.J.; Brisson, M.; Keane, A.; Simms, K.T.; Caruana, M.; Burger, E.A.; Martin, D.; Nguyen, D.T.N.; Bénard, É.; et al. Mortality impact of achieving WHO cervical cancer elimination targets: A comparative modelling analysis in 78 low-income and lower-middle-income countries. Lancet 2020, 395, 591–603. [Google Scholar] [CrossRef] [Green Version]
  110. Olczak, P.; Roden, R.B. Progress in L2-Based Prophylactic Vaccine Development for Protection against Diverse Human Papillomavirus Genotypes and Associated Diseases. Vaccines 2020, 8, 568. [Google Scholar] [CrossRef]
Figure 1. (A)After HPV vaccination, the antigen-presenting cells (APCs) expose the L1 epitopes to CD4+T cells through the major histocompatibility complex class II molecules (MHC-II). CD4+T cells in turn secrete cytokines that activate B-cells, T-cells (TH1; T-Helper 1 cells) and macrophages. HPV vaccination generates prolonged IgG antibody titers. (B) The produced antibodies (IgG) traverse the epithelial layer through the neonatal Fc receptor that is expressed at the epithelial cells of the cervix. In addition, IgG can also be transferred directly at the site of injury in the basal epithelial cell layer in order to block viral infection.
Figure 1. (A)After HPV vaccination, the antigen-presenting cells (APCs) expose the L1 epitopes to CD4+T cells through the major histocompatibility complex class II molecules (MHC-II). CD4+T cells in turn secrete cytokines that activate B-cells, T-cells (TH1; T-Helper 1 cells) and macrophages. HPV vaccination generates prolonged IgG antibody titers. (B) The produced antibodies (IgG) traverse the epithelial layer through the neonatal Fc receptor that is expressed at the epithelial cells of the cervix. In addition, IgG can also be transferred directly at the site of injury in the basal epithelial cell layer in order to block viral infection.
Viruses 15 00141 g001
Figure 2. Mutation profile of HPV16 L1 and L2 proteins. The immunogenic regions of viral proteins are highlighted. (A) Seven amino acid mutations were detected in the L1 protein with high frequency in certain geographic areas. Five of these mutations (T176N, N181T, T266A, T353P, T389S) are located within the L1 hypervariable surface loops. (B) Seventeen amino acid mutations were found in the L2 protein with high frequency in various geographic populations. Two of these amino acid substitutions (D43E, S122P) are adjacent to the L2 specific epitopes.
Figure 2. Mutation profile of HPV16 L1 and L2 proteins. The immunogenic regions of viral proteins are highlighted. (A) Seven amino acid mutations were detected in the L1 protein with high frequency in certain geographic areas. Five of these mutations (T176N, N181T, T266A, T353P, T389S) are located within the L1 hypervariable surface loops. (B) Seventeen amino acid mutations were found in the L2 protein with high frequency in various geographic populations. Two of these amino acid substitutions (D43E, S122P) are adjacent to the L2 specific epitopes.
Viruses 15 00141 g002
Table 1. Sequences retrieved from the different geographic regions.
Table 1. Sequences retrieved from the different geographic regions.
GeneGeographic LocationSequences (n)
L1Africa1
L1Asia356
L1Europe456
L1America2884
L1South America143
L1North America2741
L2Africa0
L2Asia233
L2Europe175
L2America1651
L2South America105
L2North America1546
Table 2. Distribution of amino acid substitutions in the L1 and L2 proteins in various geographic regions.
Table 2. Distribution of amino acid substitutions in the L1 and L2 proteins in various geographic regions.
Total DataEuropeAsiaTotal AmericaNorth AmericaSouth America
PosRef PosRef aaMajor aaaa %Subs aaaa %Major aaaa %Subs aaaa %Major aaaa %Subs aaaa %Major aaaa %Subs aaaa %Major aaaa %Subs aaaa %Major aaaa %Subs aaaa %
L2w4343DD98.6E1.4D100-0D88E12D100-0D100-0D100-0
122122SS83.5P16S94.9P5.1S88P11.2S81.6P18S83P16.6S54.3P45.7
L1 176176TT85.8N14T93.2N6.8T82.3N17.7T85.1N15T86N13.6T60.1N39.9
EF loop181181NN90.8T9.2N96.7T3.3N83.7T16N90.7T9.3N91T8.65N78.3T21.7
FG loop266266TA71T29A77.9T22.1A94.7T5.3A67.1T33A66T34.2A90.9T9.1
353353TT88P12T94.1P5.9T83.1P16.9T87.6P12T89P10.8T60.6P39.4
HI loop389389TT94.3S5.7T99.1S0.9T96.9S2.8T93.2S6.8T94S5.98T78.2S21.8
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

Tsakogiannis, D.; Nikolaidis, M.; Zagouri, F.; Zografos, E.; Kottaridi, C.; Kyriakopoulou, Z.; Tzioga, L.; Markoulatos, P.; Amoutzias, G.D.; Bletsa, G. Mutation Profile of HPV16 L1 and L2 Genes in Different Geographic Areas. Viruses 2023, 15, 141. https://doi.org/10.3390/v15010141

AMA Style

Tsakogiannis D, Nikolaidis M, Zagouri F, Zografos E, Kottaridi C, Kyriakopoulou Z, Tzioga L, Markoulatos P, Amoutzias GD, Bletsa G. Mutation Profile of HPV16 L1 and L2 Genes in Different Geographic Areas. Viruses. 2023; 15(1):141. https://doi.org/10.3390/v15010141

Chicago/Turabian Style

Tsakogiannis, Dimitris, Marios Nikolaidis, Flora Zagouri, Eleni Zografos, Christine Kottaridi, Zaharoula Kyriakopoulou, Lamprini Tzioga, Panayotis Markoulatos, Grigoris D. Amoutzias, and Garyfalia Bletsa. 2023. "Mutation Profile of HPV16 L1 and L2 Genes in Different Geographic Areas" Viruses 15, no. 1: 141. https://doi.org/10.3390/v15010141

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