Next Article in Journal
Rectovaginal Colonization with Serotypes of Group B Streptococci with Reduced Penicillin Susceptibility among Pregnant Women in León, Nicaragua
Previous Article in Journal
The Complexity of Swine Caliciviruses. A Mini Review on Genomic Diversity, Infection Diagnostics, World Prevalence and Pathogenicity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

H. pylori Infection and Virulence Factors cagA and vacA (s and m Regions) in Gastric Adenocarcinoma from Pará State, Brazil

by
Igor Brasil-Costa
1,*,
Cintya de Oliveira Souza
2,
Leni Célia Reis Monteiro
2,
Maria Elisabete Silva Santos
3,
Edivaldo Herculano Correa De Oliveira
4 and
Rommel Mario Rodriguez Burbano
5
1
Laboratório de Imunologia, Seção de Virologia, Instituto Evandro Chagas, Ananindeua 67030-000, PA, Brazil
2
Laboratório de Enteroinfecções Bacterianas, Seção de Bacteriologia e Micologia, Instituto Evandro Chagas, Ananindeua 67030-000, PA, Brazil
3
Instituto de Ciências Biológicas, Universidade Federal do Pará, Belém 66075-110, PA, Brazil
4
Laboratório de Células e Tecidos, Seção de Meio Ambiente, Instituto Evandro Chagas, Ananindeua 67030-000, PA, Brazil
5
Laboratório de Citogenética Humana, Universidade Federal do Pará, Belém 66075-110, PA, Brazil
*
Author to whom correspondence should be addressed.
Pathogens 2022, 11(4), 414; https://doi.org/10.3390/pathogens11040414
Submission received: 20 January 2022 / Revised: 17 March 2022 / Accepted: 21 March 2022 / Published: 29 March 2022
(This article belongs to the Section Bacterial Pathogens)

Abstract

:
H. pylori shows a great variability in genes associated with virulence, which may influence properties related to gastric adenocarcinoma initiation and progression. Among them, cagA and vacA show a strong positive association with the disease. Therefore, a cross-sectional study was carried out with 281 samples of gastric adenocarcinoma, collected at a cancer reference center in the Brazilian Amazon. Detection of H. pylori was proceeded by PCR of the ureA and 16S genes. Positive samples were subjected to the cagA detection and vacA typing. The bacteria were observed in 32.03% of the samples. Positivity for H. pylori was associated with advanced age (p = 0.0093) and metastases (p = 0.0073). Among the positive cases, 80% (72/90) had the cagA gene. For the “s” position of the vacA gene, 98.8% (83/84) of the bacteria had genotype s1 and 1.2% (1/84) were genotyped as s2. For the “m” position, the results were: 63.6% (56/88) with m1 genotype, 2.3% (2/88) genotyped as m2 and 34.1% (30/88) m1/m2. Virulence factors did not impact an increase in the association with age or metastases. In conclusion, H. pylori infection is associated with malignant phenotype cases of gastric adenocarcinoma, involving metastases. The virulence factors related to the cagA and vacA genes showed a high prevalence in the Brazilian Amazon.

1. Introduction

Since 1994, the World Health Organization (WHO) and the International Agency for Research on Cancer (IARC) determined that there was sufficient evidence to support carcinogenicity of H. pylori infection in humans, then classified the bacterium as a class I carcinogen [1]. The infection can cause chronic gastritis, peptic ulcer, gastric adenocarcinoma and lymphoma of the lymphoid tissue associated with the gastric mucosa (MALT) [2,3]. H. pylori infection is the main risk factor for stomach cancer [4]. It is a Gram-negative, spiral or slightly curved bacillus, growing in a microaerophilic environment that colonizes the stomach [5]. There is great genomic variability among H. pylori strains, which may be associated with genesis or progression of gastric adenocarcinoma, mainly due to the presence of virulence factors associated with the bacterium genome [6,7]. Half of the differentiated genes in each strain are grouped in a hypervariable region, known as the “plasticity zone”. In addition to this “plasticity zone”, the cag pathogenicity island (CPI), a region of approximately 40 Kb that confers virulence, also shows high genetic variability among strains [8,9,10].
Within the CPI, the most studied gene is cagA, which differs between strains of H. pylori and confers virulence characteristics [9,11,12,13]. This gene is present in about 60% of the strains described in western countries. Positive cagA is one of the most virulent strains of H. pylori, being associated with gastric ulcer, duodenal ulcer and gastric cancer [14,15].
In some regions of the world, including Asia and most of Africa, almost all strains of H. pylori show intact CPI, while 30% of strains of the bacterium found in Europe and North America do not have this island, being considered cagA negative. However, it is possible that cagA positive and negative strains co-exist in the same stomach [8].
After entering the cytoplasm of gastric cells, CagA is phosphorylated at EPIYA sites (glutamate/proline/isoleucine/tyrosine/alanine) by the family of kinases Abl and Src [16]. Once phosphorylated, the protein can bind to SHP-2 and deregulate the normal signaling process in cells [17]. This action may involve important processes such as the host’s cellular metabolism, including cell proliferation [18], apoptosis and maintenance of normal cytoskeleton structure, prerequisite for neoplastic transformation [16,17,19]. Additionally, the presence of CagA induces an increased production of interleukin 8 by epithelial cells [20,21]. This interleukin produces an intense local inflammatory response and stimulates the production of free radicals such as reactive oxygen species (ROS) that can lead to damage to the DNA of adjacent cells [22]. The accumulation of DNA damage causes genetic changes in gastric epithelial cells, which in turn can culminate in carcinogenesis [15,23].
In addition to the CPI, H. pylori has several other gene regions related to pathogenicity and/or virulence. Among them, we can highlight the genes vacA, iceA and babA [24]. The vacA gene is present in the bacterial chromosome and encodes a high molecular weight multimeric protein (VacA) characterized as a multifunctional toxin that induces cell vacuolization, formation of membrane channels, dysregulation of endosomal/lysosomal function, apoptosis and immunomodulation [11,25,26,27,28]. Unlike cagA, the vacA gene is present in all strains of H. pylori but exhibits a high level of genetic diversity. The gene contains two polymorphic regions: the “s” region and the “m” region. The “s” region encodes the signal peptide in the 5′ gene region and can have the s1 or s2 genotypes. In turn, the “m” region is located in the middle region of the gene and encodes the host cell binding site and exists in the m1 or m2 genotypes (Figure 1A). Clinical studies have shown that the s1m1 genotype is more virulent because it has greater vacuolating activity in several cell lines [6,9,12,29,30]. This genotype is associated with an increased production of the vacuolating toxin, culminating in high levels of inflammation and the presence of epithelial damage in the gastric mucosa, which increases the risk of gastric adenocarcinoma [6]. There is also a strong association between the vacA s1 genotype and the presence of the CPI, which enhances the virulence of the strain [29,31]. Thus, CagA and VacA are the most important virulence factors during H. pylori infection [12,30,32].
The present study aimed to determine the prevalence of H. pylori infection in gastric adenocarcinomas, as well as the cagA gene and s1, s2, m1 and m2 genotypes of the vacA gene, associating these factors to gender, age, histological types, primary location tumor and gastric adenocarcinoma staging indices.

2. Results

H. pylori was detected in 32.03% (90/281) of the samples. The Chi square test demonstrated a borderline statistical association (p < 0.05) between the H. pylori and clinical-epidemiological information in females, which presented a relatively higher number of positive cases, compared to the males. A significant difference was observed in the analysis of the patients’ age (0.0093). Kruskal–Wallis test showed that the association with age was reinforced (p = 0.0005), because positive H. pylori patients were older (average of 65.54) compared to negative patients (average of 58.66) (Table 1).
A robust association was also obtained between positivity for H. pylori and the presence of metastases (p = 0.0073). Among H. pylori positive samples, 59.6% (53/89) had metastases, while only 41.5% (78/188) of the negative ones evolved to this complication (Table 1). The simple logistic regression test reinforced the association of the bacterium with the appearance of metastases (p = 0.0053; OR = 2.0762; 95% CI = 1.24 to 3.47).
Among the H. pylori positive samples, 80% (72/90) were positive for the cagA gene. For the “s” position of the vacA gene, 98.8% (83/84) of the bacteria had genotype s1 and 1.2% (1/84) were genotyped as s2, with 6 not obtaining results for the region. For the “m” position, the results were: 63.6% (56/88) with m1 genotype; 2.3% (2/88) genotyped as m2 and 34.1% (30/88) m1/m2, with 2 cases without information from the “m” region (Figure 1B).
For the bacteria combined vacA genotype, 63.9% (53/83) were defined as s1m1, 1.2% (1/83) s1m2, 33.7% (28/83) s1m1/m2 and 1.2% (1/83) s2m2, and in seven samples a combined vacA genotype were not defined.
The combined cagA and vacA genotype showed the following result: 53% (44/83) with combined cagA+ s1m1 genotype, 30.1% (25/83) with cagA+ s1m1/2, 10.9% (9/83) with cagA- s1m1, 3.6% (3/83) with cagA- s1m1/2, 1.2% (1/83) cagA- s1m2 cases and 1.2% (1/83) cagA- s2m2.
The comparison of the H. pylori genotypes for cagA, “s” or “m” vacA gene with clinical-epidemiological information demonstrated the association only between patients age and the presence of cagA gene (p = 0, 0402, by the Chi-square; p = 0.0001, by the Kruskal–Wallis test). However, it did not follow the same association tendency with the presence of H. pylori, as older individuals were found in the negative cagA group (mean of 75.11) when compared with patients diagnosed with the presence of the gene (mean of 61.90).

3. Discussion

Although chronic H. pylori infection is estimated to contribute to more than 80% of gastric cancer cases [33,34], its prevalence in gastric adenocarcinoma samples varies greatly. For instance, in Asia, where previous studies have consistently reported superior gastric cancer outcomes [35], the prevalence of H. pylori in gastric adenocarcinoma samples varies from 40% in China [36] to 89.9% in South Korea [37]. In Brazil, the presence of H. pylori showed high prevalence in different regions—85.7% in São Paulo [38,39], 93% in Ceará [9] and 88% in Pará [40]. Despite this, only 32.03% of our samples were positive for H. pylori presence, suggesting that other factors may have favored the emergence of the neoplasia [41,42,43,44] in this population, such as genetic constitution, eating habits and other possible exposures. Although H. pylori infection is an important promoter step of gastric carcinogenesis, treatment benefits and eradication on gastric cancer prevention is controversial due to intestinal metaplasia and dysplasia be considered a “point of no return” in the precancerous cascade [45,46]. In this sense, cancer occurrence is observed, however, the bacteria detection is negative. In addition, other microbiological and molecular aspect in detection may have influenced the low prevalence found in this study, such as the low-threshold bacterial activity or low-copy H. pylori load occurs in the patients with gastric atrophy, intestinal metaplasia, gastric mucosa-associated lymphoid tissue (MALT) lymphoma and upper gastrointestinal bleeding [47].
The presence of H. pylori was slightly higher in females (p = 0.0437); however, we judge this result as a statistical artifact due to the lack of biological basis for the phenomenon and because the epidemiological studies point to an equal proportion between men and women infected with H. pylori or increased risk of infection among the male population [12,42,43]. As gastric cancer belongs to a set of complex diseases, it has a multifactorial character in terms of risk factors, leading to some erroneous statistical findings [44], since hardly all factors that contribute to the outcome will be controlled.
The bacterium is normally associated with the intestinal type of gastric adenocarcinoma, with distal or non-cardiac location and with the presence of distant metastases [40,48,49]. However, a study also associated H. pylori with the cardiac location [50]. The same frequency is observed in different degrees of metastasis [51], however, in the present study, the association of the presence of bacteria in metastatic cancer strongly associated only with the presence of metastases (p = 0.0073, by X2 and p = 0.0053 by RLS). Kong et al. demonstrated that the bacteria can positively regulate the expression of CACUL1, a protein associated with the cell cycle that promotes cell proliferation by activating CDK2 in the transition from G1 to S phase [52]. CACUL1 is also capable to stimulate the expression of the MMP-9 and SLUG genes, important in the process of cell invasion and migration, which confers a great risk of metastases. Infection by the bacterium can produce a more invasive phenotype of malignant cells by releasing gamma-glutamyltransferase, which contributes significantly to the establishment of metastases [26]. A higher expression of MMP-7 was also found in patients infected with H. pylori [53,54]. Bagheri et al. observed an increased expression of MMP9 in non-neoplastic diseases of the stomach in patients infected with bacteria, suggesting that changes occur in premalignant stages [53].
As for the bacterial virulence factors, the cagA gene was found in 80% of the strains. A study carried out in the same geographic region, addressing the same neoplasia, observed the presence of the virulence gene in 88.3% of cases [41]. Another study, developed in the northeast region of Brazil, found 64.9% of the strains showing the gene [9]. The association found between the presence of cagA and younger patients suggests parallel risks for the development of gastric adenocarcinoma, where these patients develop the neoplasia primarily in the presence of the virulence factor, while in the absence of the factor, the bacterium is favored by senility. Additionally, the presence of other pathogenicity island genes may confer greater or lesser effectiveness on the carcinogenic action of CagA [55].
It is important to highlight the high number of cases with mixed infection by more than one strain: 33.7% of the vacA genotypes were determined as m1/m2. This result reflects an increased exposure to the transmission of the bacterium or an accumulation of strains due to the failure to resolve the first infection [56]. Another study, in the state of Pará, Brazil, found only 2.4% of infections by more than one strain [41]. It is suggested that there is a greater susceptibility to the persistence of H. pylori in the studied population; however, further studies are needed to better elucidate this finding. A study carried out in Colombia demonstrated that there is a greater number of infections with multiple strains and greater bacterial genetic variability in regions of greater risk for the development of gastric cancer, when compared to regions of lower risk [57].
H. pylori genotype associated as a risk factor (cagA+ s1m1) for gastric adenocarcinoma was found in 83.1% of cases, a percentage similar to that observed by Vinagre et al. at 86.9%, also in the state of Pará, Brazil. This result demonstrates that the strains of bacteria we have found have great oncogenic potential [41], which could explain the high prevalence of gastric adenocarcinoma in the studied population.
The associations found between the presence of the bacterium and the clinical-epidemiological variables had a loss of statistical significance when comparing only the genotypes considered more virulent (cagA+, s1, m1, s1m1 and cagA+ s1m1). Some studies observed an association with the development of gastric cancer and the presence of metastases for the cagA+ strains, unlike our results [40,58,59]. This may be due to the multifactorial nature of the disease. However, Dabiri et al. found no association between the presence of cagA and the neoplasm, demonstrating that the presence of this virulence factor is still controversial for the increased risk of developing gastric cancer and more accelerated disease progression [58]. A hypothesis to justify the divergent results may be the genetic variability that is not normally observed in most studies, such as the number of copies of cagA (ranging from 1 to 4) and the ignored variability in regions of low coverage by sequencing [60,61]. On the other hand, the involvement of other virulent factors cannot be ignored [24,29,62]. Gastli et al. suggest the use of H. pylori quantification associated with cagA genotyping in routine workflow for a sensitive and reliable diagnosis, to identify patients at high risk and to manage eradication therapies [62].
Regardless of the strain, the importance of H. pylori in the development of gastric cancer appears to be indisputable. The eradication of the bacteria by itself causes a significant decline in cases of the disease [63,64,65,66,67,68,69]. In addition to the eradication of the bacteria with antibiotics, new approaches to gastric cancer prevention have emerged, such as piperine treatment, which suppresses H. pylori toxin entry into gastric epithelium and minimizes β-catenin mediated oncogenesis and IL-8 secretion in vitro [70].

4. Materials and Methods

A cross-sectional study was carried out with 281 DNA samples extracted from fresh tumor biopsies from cases of gastric adenocarcinoma collected at the Ophir Loyola Hospital, State of Pará, Brazil, a reference in diagnosis and treatment for this tumor type. Histological types were classified as intestinal or diffuse. Epidemiological and histological information was obtained from the Human Cytogenetics Laboratory gastric tumor database, Federal University of Pará (Brazil).
DNA samples were extracted using the QIAamp tissue kit (Qiagen, Hilden, Germany), according to the manufacturer’s protocol, and subjected to quantification with the Qubit® dsDNA BR Assay kit on the Qubit® equipment (Life Technologies, Carlsbad, CA, USA). All samples with concentrations higher or lower than 100 ng/µL were adjusted for this value.
Molecular detection of H. pylori was performed by PCR, using two target regions: the urease A (ureA) genes and the 16SrRNA (16S) gene. Positive samples for at least one of these genes were directed to PCR [32] for the cagA virulence gene (cytotoxin-associated gene A) and genotyping through the allelic search for the vacA gene (vacuolating cytotoxin gene A): signal region (s1/s2) and median region of the gene (m1/m2). The PCR primers to search for these genes and the sizes of the amplified fragments are shown in Table 2.
The PCR reactions were performed with a final volume of 25 µL, using the components and their respective concentrations and volumes according to the specific target as observed in Table 3.
To perform the cycling, the thermal cycler Veriti Thermal Cycler (Applied Biosystems, Waltham, MA, USA) was used. Cycling programs were performed at a denaturation temperature of 95 °C for 5 min, followed by 40 cycles of denaturation, annealing and extension for one minute each and finally an extension step of 7 min at 72 °C. For the diagnostic targets 16SrRNA and ureA, annealing temperatures of 50 °C and 60 °C, respectively, were used. For the virulence factor targets cagA and vacA, the annealing temperature was 55 °C.
For positive control of the reactions for H. pylori, ATCC 43504 strain and DNA extracted from biopsies with positive urease test were used, while DNA extracted from biopsies with negative urease test were used as negative control.
The PCR products were subjected to electrophoresis on 2.0% agarose gel containing 6 µL of Sybr Safe (Invitrogen, Waltham, MA, USA) for visualization and determination of the sizes of the amplified DNA fragments. Together with the samples, the positive controls and the molecular weight marker (Ready-load 1 kb DNA Ladder, Invitrogen) were applied to the gels for comparison. Electrophoretic migrations were performed under constant voltage of 100 V in TE buffer (Tris-EDTA Buffer) for 45 min. After the migration, the gels were visualized on a transilluminator (ultraviolet light) for visual analysis of the amplified DNA fragments, and the photographic record was performed by the UPV Biolmaging Systems (EpiChemi Darkroom, Singapore) photo-documentation system.
Bioestat v5.3 (https://www.mamiraua.org.br/downloads/programas/, accessed on 30 October 2021) was used to perform the statistical analyses. Simple and multiple logistic regression tests, chi-square, and Kruskal–Wallis were conducted to verify associations between the epidemiological and histological variables with the presence or factors of H. pylori virulence.

5. Conclusions

Our study demonstrated that the prevalence of H. pylori infection in gastric adenocarcinoma in the State of Pará, Brazilian Amazon, was considerably lower when compared to other similar studies, however, the majority of positive cases had the usual more virulent strains. The presence of the bacteria was associated with advanced age and the presence of metastases, reinforcing the importance of infection for the progression of the neoplasia. Our results suggest that the bacteria may not be present throughout the course of gastric adenocarcinoma, but it has the potential virulence to accelerate the development of the disease in the Brazilian Amazon.

Author Contributions

Conceptualization, formal analysis, investigation, funding acquisition and writing—original draft, I.B.-C.; methodology, validation, funding acquisition and writing—review and editing, C.d.O.S.; methodology, L.C.R.M.; visualization and writing—review and editing, M.E.S.S. and E.H.C.D.O.; project administration and supervision, R.M.R.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Health Surveillance Secretariat of the Brazilian Ministry of Health (SVS/MS) (Brasília, Distrito Federal, Brazil); there is no grant number.

Institutional Review Board Statement

The study was submitted and approved (Process number 04133612.4.0000.0019; approval date: 27 September 2012) by the Research Ethics Committee of the Evandro Chagas Institute in accordance with the regulatory standards for research involving human beings.

Informed Consent Statement

It was proposed and accepted (Process number 04133612.4.0000.0019; approval date: 27 September 2012) by the Ethics Committee in Research with Human at Instituto Evandro Chagas to waive the consent form, as the participants in the original project (responsible for the collection of samples) had already died. The original project normally applied the informed consent.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Takahashi-Kanemitsu, A.; Knight, C.T.; Hatakeyama, M. Molecular anatomy and pathogenic actions of Helicobacter pylori CagA that underpin gastric carcinogenesis. Cell. Mol. Immunol. 2020, 17, 50–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. González, I.; Araya, P.; Rojas, A. Helicobacter pylori Infection and Lung Cancer: New Insights and Future Challenges. Chin. J. Lung Cancer 2018, 21, 658–662. [Google Scholar] [CrossRef]
  3. Joo, M. Rare Gastric Lesions Associated with Helicobacter pylori Infection: A Histopathological Review. J. Pathol. Transl. Med. 2017, 51, 341–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Poorolajal, J.; Moradi, L.; Mohammadi, Y.; Cheraghi, Z.; Gohari-Ensaf, F. Risk factors for stomach cancer: A systematic review and meta-analysis. Epidemiol. Health 2020, 42, e2020004. [Google Scholar] [CrossRef]
  5. Palamides, P.; Jolaiya, T.; Idowu, A.; Loell, E.; Onyekwere, C.; Ugiagbe, R.; Agbo, I.; Lesi, O.; Ndububa, D.; Adekanle, O.; et al. Helicobacter pylori patient isolates from South Africa and Nigeria differ in virulence factor pathogenicity profile and associated gastric disease outcome. Sci. Rep. 2020, 10, 11409. [Google Scholar] [CrossRef]
  6. Mucito-Varela, E.; Castillo-Rojas, G.; Calva, J.J.; López-Vidal, Y. Integrative and Conjugative Elements of Helicobacter pylori Are Hypothetical Virulence Factors Associated with Gastric Cancer. Front. Cell. Infect. Microbiol. 2020, 10, 525335. [Google Scholar] [CrossRef]
  7. Sharndama, H.C.; Mba, I.E. Helicobacter pylori: An up-to-date overview on the virulence and pathogenesis mechanisms. Braz. J. Microbiol. 2022, 53, 33–50. [Google Scholar] [CrossRef]
  8. Canzian, F.; Rizzato, C.; Obazee, O.; Stein, A.; Flores-Luna, L.; Camorlinga-Ponce, M.; Mendez-Tenorio, A.; Vivas, J.; Trujillo, E.; Jang, H.; et al. Genetic polymorphisms in the cag pathogenicity island of Helicobacter pylori and risk of stomach cancer and high-grade premalignant gastric lesions. Int. J. Cancer 2020, 147, 2437–2445. [Google Scholar] [CrossRef]
  9. Lima, V.P.; Silva-Fernandes, I.J.d.L.; Alves, M.K.S.; Rabenhorst, S.H.B. Prevalence of Helicobacter pylori genotypes (vacA, cagA, cagE and virB11) in gastric cancer in Brazilian’s patients: An association with histopathological parameters. Cancer Epidemiol. 2011, 35, e32–e37. [Google Scholar] [CrossRef] [Green Version]
  10. Silva, B.; Nunes, A.; Vale, F.F.; Rocha, R.; Gomes, J.P.; Dias, R.; Oleastro, M. The expression of Helicobacter pylori tfs plasticity zone cluster is regulated by pH and adherence, and its composition is associated with differential gastric IL-8 secretion. Helicobacter 2017, 22, e12390. [Google Scholar] [CrossRef]
  11. Foegeding, N.J.; Caston, R.R.; McClain, M.S.; Ohi, M.D.; Cover, T.L. An Overview of Helicobacter pylori VacA Toxin Biology. Toxins 2016, 8, 173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Shetty, V.; Lingadakai, R.; Pai, G.C.; Ballal, M. Profile of Helicobacter pylori cagA & vacA genotypes and its association with the spectrum of gastroduodenal disease. Indian J. Med. Microbiol. 2021, 39, 495–499. [Google Scholar] [CrossRef] [PubMed]
  13. Watanabe, Y.; Oikawa, R.; Kodaka, Y.; Sato, Y.; Ono, S.; Kenmochi, T.; Suzuki, H.; Futagami, S.; Kato, M.; Yamamoto, H.; et al. Cancer-related genetic variants of Helicobacter pylori strains determined using gastric wash-based whole-genome analysis with single-molecule real-time technology. Int. J. Cancer 2021, 148, 178–192. [Google Scholar] [CrossRef] [PubMed]
  14. Boonyanugomol, W.; Kongkasame, W.; Palittapongarnpim, P.; Baik, S.-C.; Jung, M.-H.; Shin, M.-K.; Kang, H.-L.; Lee, W.-K. Genetic variation in the cag pathogenicity island of Helicobacter pylori strains detected from gastroduodenal patients in Thailand. Braz. J. Microbiol. 2020, 51, 1093–1101. [Google Scholar] [CrossRef]
  15. Tegtmeyer, N.; Neddermann, M.; Asche, C.I.; Backert, S. Subversion of host kinases: A key network in cellular signaling hijacked by Helicobacter pylori CagA. Mol. Microbiol. 2017, 105, 358–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Nagase, L.; Hayashi, T.; Senda, T.; Hatakeyama, M. Dramatic increase in SHP2 binding activity of Helicobacter pylori Western CagA by EPIYA-C duplication: Its implications in gastric carcinogenesis. Sci. Rep. 2015, 5, 15749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Tagoe, E.A.; Awandare, G.A.; Quaye, O.; Asmah, R.H.; Archampong, T.N.; Osman, M.A.; Brown, C.A. Helicobacter pylori Variants with ABC-Type Tyrosine Phosphorylation Motif in Gastric Biopsies of Ghanaian Patients. BioMed Res. Int. 2021, 2021, 6616059. [Google Scholar] [CrossRef]
  18. Liu, B.; Li, X.; Sun, F.; Tong, X.; Bai, Y.; Jin, K.; Liu, L.; Dai, F.; Li, N. HP-CagA+ Regulates the Expression of CDK4/CyclinD1 via reg3 to Change Cell Cycle and Promote Cell Proliferation. Int. J. Mol. Sci. 2019, 21, 224. [Google Scholar] [CrossRef] [Green Version]
  19. Hayashi, T.; Senda, M.; Suzuki, N.; Nishikawa, H.; Ben, C.; Tang, C.; Nagase, L.; Inoue, K.; Senda, T.; Hatakeyama, M. Differential Mechanisms for SHP2 Binding and Activation Are Exploited by Geographically Distinct Helicobacter pylori CagA Oncoproteins. Cell Rep. 2017, 20, 2876–2890. [Google Scholar] [CrossRef] [Green Version]
  20. Fazeli, Z.; Alebouyeh, M.; Tavirani, M.R.; Azimirad, M.; Yadegar, A. Helicobacter pylori CagA induced interleukin-8 secretion in gastric epithelial cells. Gastroenterol. Hepatol. Bed Bench 2016, 9, S42–S46. [Google Scholar]
  21. Ferreira, R.; Pinto-Ribeiro, I.; Wen, X.; Marcos-Pinto, R.; Dinis-Ribeiro, M.; Carneiro, F.; Figueiredo, C. Helicobacter pylori cagA Promoter Region Sequences Influence CagA Expression and Interleukin 8 Secretion. J. Infect. Dis. 2016, 213, 669–673. [Google Scholar] [CrossRef] [PubMed]
  22. Kim, S.H.; Lim, J.W.; Kim, H. Astaxanthin Inhibits Mitochondrial Dysfunction and Interleukin-8 Expression in Helicobacter pylori-Infected Gastric Epithelial Cells. Nutrients 2018, 10, 1320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Bae, S.; Lim, J.; Kim, H. β-Carotene Inhibits Expression of Matrix Metalloproteinase-10 and Invasion in Helicobacter pylori-Infected Gastric Epithelial Cells. Molecules 2021, 26, 1567. [Google Scholar] [CrossRef]
  24. Dadashzadeh, K.; Peppelenbosch, M.; Adamu, A.I. Helicobacter pylori Pathogenicity Factors Related to Gastric Cancer. Can. J. Gastroenterol. Hepatol. 2017, 2017, e7942489. [Google Scholar] [CrossRef] [Green Version]
  25. Pyburn, T.M.; Foegeding, N.J.; González-Rivera, C.; McDonald, N.A.; Gould, K.L.; Cover, T.L.; Ohi, M.D. Structural organization of membrane-inserted hexamers formed by Helicobacter pylori VacA toxin. Mol. Microbiol. 2016, 102, 22–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Ricci, V. Relationship between VacA Toxin and Host Cell Autophagy in Helicobacter pylori Infection of the Human Stomach: A Few Answers, Many Questions. Toxins 2016, 8, 203. [Google Scholar] [CrossRef] [Green Version]
  27. McClain, M.S.; Czajkowsky, D.M.; Torres, V.J.; Szabo, G.; Shao, Z.; Cover, T.L. Random Mutagenesis of Helicobacter pylori vacA to Identify Amino Acids Essential for Vacuolating Cytotoxic Activity. Infect. Immun. 2006, 74, 6188–6195. [Google Scholar] [CrossRef] [Green Version]
  28. McClain, M.S.; Iwamoto, H.; Cao, P.; Vinion-Dubiel, A.D.; Li, Y.; Szabo, G.; Shao, Z.; Cover, T. Essential Role of a GXXXG Motif for Membrane Channel Formation by Helicobacter pylori Vacuolating Toxin. J. Biol. Chem. 2003, 278, 12101–12108. [Google Scholar] [CrossRef] [Green Version]
  29. Bakhti, S.Z.; Latifi-Navid, S.; Zahri, S. Unique constellations of five polymorphic sites of Helicobacter pylori vacA and cagA status associated with risk of gastric cancer. Infect. Genet. Evol. 2020, 79, 104167. [Google Scholar] [CrossRef]
  30. Pormohammad, A.; Ghotaslou, R.; Leylabadlo, H.E.; Nasiri, M.J.; Dabiri, H.; Hashemi, A. Risk of gastric cancer in association with Helicobacter pylori different virulence factors: A systematic review and meta-analysis. Microb. Pathog. 2018, 118, 214–219. [Google Scholar] [CrossRef]
  31. Lee, D.-H.; Ha, J.-H.; Shin, J.-I.; Kim, K.-M.; Choi, J.-G.; Park, S.; Park, J.-S.; Seo, J.-H.; Park, J.-S.; Shin, M.-K.; et al. Increased Risk of Severe Gastric Symptoms by Virulence Factors vacAs1c, alpA, babA2, and hopZ in Helicobacter pylori Infection. J. Microbiol. Biotechnol. 2021, 31, 368–379. [Google Scholar] [CrossRef]
  32. Nejati, S.; Karkhah, A.; Darvish, H.; Validi, M.; Ebrahimpour, S.; Nouri, H.R. Influence of Helicobacter pylori virulence factors CagA and VacA on pathogenesis of gastrointestinal disorders. Microb. Pathog. 2018, 117, 43–48. [Google Scholar] [CrossRef] [PubMed]
  33. de Martel, C.; Georges, D.; Bray, F.; Ferlay, J.; Clifford, G.M. Global burden of cancer attributable to infections in 2018: A worldwide incidence analysis. Lancet Glob. Health 2020, 8, e180–e190. [Google Scholar] [CrossRef] [Green Version]
  34. O’Brien, V.P.; Koehne, A.L.; Dubrulle, J.; Rodriguez, A.E.; Leverich, C.K.; Kong, V.P.; Campbell, J.S.; Pierce, R.H.; Goldenring, J.R.; Choi, E.; et al. Sustained Helicobacter pylori infection accelerates gastric dysplasia in a mouse model. Life Sci. Alliance 2021, 4, e202000967. [Google Scholar] [CrossRef] [PubMed]
  35. Rahman, R.; Asombang, A.W.; Ibdah, J.A. Characteristics of gastric cancer in Asia. World J. Gastroenterol. 2014, 20, 4483–4490. [Google Scholar] [CrossRef]
  36. Wang, R.; Chen, X.-Z. High mortality from hepatic, gastric and esophageal cancers in mainland China: 40 years of experience and development. Clin. Res. Hepatol. Gastroenterol. 2014, 38, 751–756. [Google Scholar] [CrossRef]
  37. Kim, J.Y.; Lee, H.S.; Kim, N.; Shin, C.M.; Lee, S.H.; Park, Y.S.; Hwang, J.-H.; Kim, J.-W.; Jeong, S.-H.; Lee, D.H.; et al. Prevalence and Clinicopathologic Characteristics of Gastric Cardia Cancer in South Korea. Helicobacter 2012, 17, 358–368. [Google Scholar] [CrossRef]
  38. Zaterka, S.; Eisig, J.N.; Chinzon, D.; Rothstein, W. Factors Related to Helicobacter pylori Prevalence in an Adult Population in Brazil. Helicobacter 2007, 12, 82–88. [Google Scholar] [CrossRef]
  39. Coelho, L.G.V.; Marinho, J.R.; Genta, R.; Ribeiro, L.T.; Passos, M.D.C.F.; Zaterka, S.; Assumpção, P.P.; Barbosa, A.A.J.; Barbuti, R.; Braga, L.L.; et al. Ivth Brazilian Consensus Conference on Helicobacter pylori Infection. Arq. Gastroenterol. 2018, 55, 97–121. [Google Scholar] [CrossRef]
  40. De Souza, C.R.T.; De Oliveira, K.S.; Ferraz, J.J.S.; Leal, M.F.; Calcagno, D.Q.; Seabra, A.D.; Khayat, A.S.; Montenegro, R.C.; Alves, A.P.N.N.; Assumpção, P.P.; et al. Occurrence of Helicobacter pylori and Epstein-Barr virus infection in endoscopic and gastric cancer patients from Northern Brazil. BMC Gastroenterol. 2014, 14, 179. [Google Scholar] [CrossRef] [Green Version]
  41. Vinagre, I.D.F.; De Queiroz, A.L.; Silva Júnior, M.R.D.; Vinagre, R.M.D.F.; Martins, L.C. Helicobacter pylori Infection in Patients with Different Gastrointestinal Diseases from Northern Brazil. Arq. Gastroenterol. 2015, 52, 266–271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. De Larrea-Baz, N.F.; Michel, A.; Romero, B.; Pérez-Gómez, B.; Moreno, V.; Martín, V.; Dierssen-Sotos, T.; Jimenez-Moleon, J.J.; Castilla, J.; Tardon, A.; et al. Helicobacter pylori Antibody Reactivities and Colorectal Cancer Risk in a Case-control Study in Spain. Front. Microbiol. 2017, 8, 888. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, L.; Chen, Z.; Xia, X.; Chi, J.; Li, H.; Liu, X.; Li, R.; Li, Y.; Liu, D.; Tian, D.; et al. Helicobacter pylori infection selectively increases the risk for carotid atherosclerosis in young males. Atherosclerosis 2019, 291, 71–77. [Google Scholar] [CrossRef]
  44. Machlowska, J.; Baj, J.; Sitarz, M.; Maciejewski, R.; Sitarz, R. Gastric Cancer: Epidemiology, Risk Factors, Classification, Genomic Characteristics and Treatment Strategies. Int. J. Mol. Sci. 2020, 21, 4012. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, K.S.-H.; Wong, I.O.-L.; Leung, W.K. Helicobacter pylori associated gastric intestinal metaplasia: Treatment and surveillance. World J. Gastroenterol. 2016, 22, 1311–1320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Chen, H.-N.; Wang, Z.; Li, X.; Zhou, Z.-G. Helicobacter pylori eradication cannot reduce the risk of gastric cancer in patients with intestinal metaplasia and dysplasia: Evidence from a meta-analysis. Gastric Cancer 2015, 19, 166–175. [Google Scholar] [CrossRef] [PubMed]
  47. Gong, L.; El-Omar, E.M. Application of molecular techniques in Helicobacter pylori detection: Limitations and improvements. Helicobacter 2021, 26, e12841. [Google Scholar] [CrossRef]
  48. Jiang, H.; Zhou, Y.; Liao, Q.; Ouyang, H. Helicobacter pylori infection promotes the invasion and metastasis of gastric cancer through increasing the expression of matrix metalloproteinase-1 and matrix metalloproteinase-10. Exp. Ther. Med. 2014, 8, 769–774. [Google Scholar] [CrossRef] [Green Version]
  49. Liu, L.P.; Sheng, X.P.; Shuai, T.K.; Zhao, Y.X.; Li, B.; Li, Y.M. Helicobacter pylori promotes invasion and metastasis of gastric cancer by enhancing heparanase expression. World J. Gastroenterol. 2018, 24, 4565–4577. [Google Scholar] [CrossRef]
  50. Xie, S.; Wang, S.; Xue, L.; Middleton, D.R.S.; Guan, C.; Hao, C.; Wang, J.; Li, B.; Chen, R.; Li, X.; et al. Helicobacter pylori Is Associated with Precancerous and Cancerous Lesions of the Gastric Cardia Mucosa: Results of a Large Population-Based Study in China. Front. Oncol. 2020, 10, 205. [Google Scholar] [CrossRef] [Green Version]
  51. Tsai, K.-F.; Liou, J.-M.; Chen, M.-J.; Chen, C.-C.; Kuo, S.-H.; Lai, I.-R.; Yeh, K.-H.; Lin, M.-T.; Wang, H.-P.; Cheng, A.-L.; et al. Distinct Clinicopathological Features and Prognosis of Helicobacter pylori Negative Gastric Cancer. PLoS ONE 2017, 12, e0170942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Kong, Y.; Ma, L.-Q.; Bai, P.-S.; Da, R.; Sun, H.; Qi, X.-G.; Ma, J.-Q.; Zhao, R.-M.; Chen, N.-Z.; Nan, K.-J. Helicobacter pylori promotes invasion and metastasis of gastric cancer cells through activation of AP-1 and up-regulation of CACUL1. Int. J. Biochem. Cell Biol. 2013, 45, 2666–2678. [Google Scholar] [CrossRef] [PubMed]
  53. Bagheri, N.; Sadeghiani, M.; Rahimian, G.; Mahsa, M.; Shafigh, M.; Rafieian-Kopaei, M.; Shirzad, H. Correlation between expression of MMP-9 and MMP-3 in Helicobacter pylori infected patients with different gastroduodenal diseases. Arab J. Gastroenterol. 2018, 19, 148–154. [Google Scholar] [CrossRef] [PubMed]
  54. Sadeghiani, M.; Bagheri, N.; Shahi, H.; Reiisi, S.; Rahimian, G.; Rashidi, R.; Mahsa, M.; Shafigh, M.; Salimi, E.; Rafieian-Kopaei, M.; et al. cag Pathogenicity island-dependent upregulation of matrix metalloproteinase-7 in infected patients with Helicobacter pylori. J. Immunoass. Immunochem. 2017, 38, 595–607. [Google Scholar] [CrossRef] [PubMed]
  55. Lin, A.S.; Dooyema, S.D.R.; Frick-Cheng, A.; Harvey, M.L.; Suarez, G.; Loh, J.T.; McDonald, W.H.; McClain, M.S.; Peek, R.M.; Cover, T.L. Bacterial Energetic Requirements for Helicobacter pylori Cag Type IV Secretion System-Dependent Alterations in Gastric Epithelial Cells. Infect. Immun. 2020, 88, e00790-19. [Google Scholar] [CrossRef] [PubMed]
  56. Mansour, K.B.; Fendri, C.; Battikh, H.; Garnier, M.; Zribi, M.; Jlizi, A.; Burucoa, C. Multiple and mixed Helicobacter pylori infections: Comparison of two epidemiological situations in Tunisia and France. Infect. Genet. Evol. 2016, 37, 43–48. [Google Scholar] [CrossRef]
  57. Matta, A.J.; Pazos, A.J.; Bustamante-Rengifo, J.A.; Bravo, L.E. Genomic Varia-bility of Helicobacter pylori Isolates of Gastric Regions from Two Colombian Populations. World J. Gastroenterol. 2017, 23, 800–809. [Google Scholar] [CrossRef]
  58. Dabiri, H.; Jafari, F.; Baghaei, K.; Shokrzadeh, L.; Abdi, S.; Pourhoseingholi, M.A.; Mohammadzadeh, A. Prevalence of Helicobacter pylori vacA, cagA, cagE, oipA, iceA, babA2 and babB genotypes in Iranian dyspeptic patients. Microb. Pathog. 2017, 105, 226–230. [Google Scholar] [CrossRef]
  59. Safaralizadeh, R.; Dastmalchi, N.; Hosseinpourfeizi, M.; Latifi-Navid, S. Helicobacter pylori virulence factors in relation to gastrointestinal diseases in Iran. Microb. Pathog. 2017, 105, 211–217. [Google Scholar] [CrossRef]
  60. Draper, J.L.; Hansen, L.M.; Bernick, D.L.; Abedrabbo, S.; Underwood, J.G.; Kong, N.; Huang, B.C.; Weis, A.M.; Weimer, B.C.; van Vliet, A.H.M.; et al. Fallacy of the Unique Genome: Sequence Diversity within Single Helicobacter pylori Strains. mBio 2017, 8, e02321-16. [Google Scholar] [CrossRef] [Green Version]
  61. Keikha, M.; Karbalaei, M. EPIYA motifs of Helicobacter pylori cagA genotypes and gastrointestinal diseases in the Iranian population: A systematic review and meta-analysis. New Microbes New Infect. 2021, 41, 100865. [Google Scholar] [CrossRef] [PubMed]
  62. Gastli, N.; Allain, M.; Lamarque, D.; Abitbol, V.; Billoët, A.; Collobert, G.; Coriat, R.; Terris, B.; Kalach, N.; Raymond, J. Diagnosis of Helicobacter pylori Infection in a Routine Testing Workflow: Effect of Bacterial Load and Virulence Factors. J. Clin. Med. 2021, 10, 2755. [Google Scholar] [CrossRef] [PubMed]
  63. Choi, I.J.; Kim, C.G.; Lee, J.Y.; Kim, Y.-I.; Kook, M.-C.; Park, B.; Joo, J. Family History of Gastric Cancer and Helicobacter pylori Treatment. N. Engl. J. Med. 2020, 382, 427–436. [Google Scholar] [CrossRef] [PubMed]
  64. Kim, J.; Wang, T.C. Helicobacter pylori and Gastric Cancer. Gastrointest. Endosc. Clin. N. Am. 2021, 31, 451–465. [Google Scholar] [CrossRef]
  65. Tsuda, M.; Asaka, M.; Kato, M.; Matsushima, R.; Fujimori, K.; Akino, K.; Kikuchi, S.; Lin, Y.; Sakamoto, N. Effect on Helicobacter pylori eradication therapy against gastric cancer in Japan. Helicobacter 2017, 22, e12415. [Google Scholar] [CrossRef] [Green Version]
  66. Rota, C.A.; Pereira-Lima, J.C.; Blaya, C.; Nardi, N.B. Consensus and Variable Region PCR Analysis of Helicobacter pylori 3′ Region of cagA Gene in Isolates from Individuals with or without Peptic Ulcer. J. Clin. Microbiol. 2001, 39, 606–612. [Google Scholar] [CrossRef] [Green Version]
  67. Palau, M.; Piqué, N.; Ramírez-Lázaro, M.; Lario, S.; Calvet, X.; Miñana-Galbis, D. Whole-Genome Sequencing and Comparative Genomics of Three Helicobacter pylori Strains Isolated from the Stomach of a Patient with Adenocarcinoma. Pathogens 2021, 10, 331. [Google Scholar] [CrossRef]
  68. Fan, R.; Han, X.; Gong, Y.; He, L.; Xue, Z.; Yang, Y.; Sun, L.; Fan, D.; You, Y.; Meng, F.; et al. Alterations of Fucosyltransferase Genes and Fucosylated Glycans in Gastric Epithelial Cells Infected with Helicobacter pylori. Pathogens 2021, 10, 168. [Google Scholar] [CrossRef]
  69. Kumar, S.; Patel, G.K.; Ghoshal, U.C. Helicobacter pylori-Induced Inflammation: Possible Factors Modulating the Risk of Gastric Cancer. Pathogens 2021, 10, 1099. [Google Scholar] [CrossRef]
  70. Tharmalingam, N.; Park, M.; Lee, M.H.; Woo, H.J.; Kim, H.W.; Yang, J.Y.; Rhee, K.-J.; Kim, J.-B. Piperine treatment suppresses Helicobacter pylori toxin entry in to gastric epithelium and minimizes β-catenin mediated oncogenesis and IL-8 secretion in vitro. Am. J. Transl. Res. 2016, 8, 885–898. [Google Scholar]
Figure 1. H. pylori vacA gene allelic diversity and molecular detection by PCR assay. (A) Gene architecture of three regions of the VacA gene: the signal region (s1 and s2), the intermediate region (i) and the mid-region (m1 and m2). (B) Amplification of vacA s1 (B1) and vacA m1/m2 (B2) alleles. Subtitle: (B1) (lane 1 to 4: strains vacA s1 [176 pb]; lane 5: positive control (vacA s1); lane 6: 100 to 1000 pb marker); (B2) (lanes 1 and 2: strains vacA m2 [476 pb]; lanes 3 and 4: strains vacA m1/m2 [401 pb/476 pb]; lane 5: 100 to 1000 pb marker; lane 6: strains vacA m1 [401 pb]; lane 7: positive control (vacA m1/m2). Note: Result of vacA s2 not showing in this figure.
Figure 1. H. pylori vacA gene allelic diversity and molecular detection by PCR assay. (A) Gene architecture of three regions of the VacA gene: the signal region (s1 and s2), the intermediate region (i) and the mid-region (m1 and m2). (B) Amplification of vacA s1 (B1) and vacA m1/m2 (B2) alleles. Subtitle: (B1) (lane 1 to 4: strains vacA s1 [176 pb]; lane 5: positive control (vacA s1); lane 6: 100 to 1000 pb marker); (B2) (lanes 1 and 2: strains vacA m2 [476 pb]; lanes 3 and 4: strains vacA m1/m2 [401 pb/476 pb]; lane 5: 100 to 1000 pb marker; lane 6: strains vacA m1 [401 pb]; lane 7: positive control (vacA m1/m2). Note: Result of vacA s2 not showing in this figure.
Pathogens 11 00414 g001
Table 1. Comparison of positivity for H. pylori with clinical-epidemiological variables.
Table 1. Comparison of positivity for H. pylori with clinical-epidemiological variables.
Clinical and Epidemiological VariablePositive H. pyloriUndetectable H. pyloriValue of p
Age≥6065.6% (59/90)48.2% (92/191)0.0093
<6034.4% (31/90)51.8% (99/191)
GenderMan60% (54/90)72.8% (139/191)0.0437
Woman40% (36/90)27.2% (52/191)
LocationProximal37.8% (34/90)42.4% (81/191)0.4614
Distal62.2% (56/90)57.6% (110/191)
Histological typeIntestinal53.3% (48/90)57.1% (109/191)0.5563
Diffuse46.7% (42/90)42.9% (82/191)
Presence of metastasespresence59.6% (53/89)41.5% (78/188)0.0073
Absent40.4% (36/89)58.5% (110/188)
Table 2. Description of PCR primers used for molecular detection and genotyping of H. pylori.
Table 2. Description of PCR primers used for molecular detection and genotyping of H. pylori.
Target Primer PCRSize (pb)Reference
ureA5′-GCCAATGGTAAATTAGTT-3′394[64]
5′-CTCCTTAATTGTTTTTAC-3′
16SrRNA5′-CCCATTTGACTCAATGCGATG-3′132[65]
5′-TGGGATTAGCGAGTATGTCGG-3′
cagA5′-GTGCCTGCTAGTTTGTCAGCG-3′402[66]
5′-TTGGAAACCACCTTTTGTATTAGC-3′
vacA m1/m25′-CACAGCCACTTTCAATAACGA-3′401/476[67]
5′-CGTCAAAATAATTCCAAGGG-3′
vacA s1/s25′-ATGGAAATACAACAAACACAC-3′176/203[67]
5′-CCTGARACCGTTCCTACAGC-3′
Table 3. Components, concentrations and volumes of PCR constituents for molecular detection and genotyping.
Table 3. Components, concentrations and volumes of PCR constituents for molecular detection and genotyping.
Components/ConcentrationsVolumes
16SrRNAureA, cagA, vacA
Ultra-pure water 14.55 μL16.55 μL
Reaction buffer 10X (Invitrogen)2.5 μL2.5 μL
Magnesium chloride-MgCl2 50 mM (Invitrogen)0.5 μL0.5 μL
Deoxynucleotides-dNTP 10 mM (Invitrogen)1.0 µL1.0 µL
Oligonucleotides-10 pmol/μL (Invitrogen)1.0 μL/1.0 μL1.0 μL/1.0 μL
Platinum-Taq DNA Polymerase 5 U/µL (Invitrogen)0.2 μL0.2 μL
DNA extracted from gastric biopsy 4 μL2 μL
Final reaction volume 25 μL25 μL
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Brasil-Costa, I.; Souza, C.d.O.; Monteiro, L.C.R.; Santos, M.E.S.; Oliveira, E.H.C.D.; Burbano, R.M.R. H. pylori Infection and Virulence Factors cagA and vacA (s and m Regions) in Gastric Adenocarcinoma from Pará State, Brazil. Pathogens 2022, 11, 414. https://doi.org/10.3390/pathogens11040414

AMA Style

Brasil-Costa I, Souza CdO, Monteiro LCR, Santos MES, Oliveira EHCD, Burbano RMR. H. pylori Infection and Virulence Factors cagA and vacA (s and m Regions) in Gastric Adenocarcinoma from Pará State, Brazil. Pathogens. 2022; 11(4):414. https://doi.org/10.3390/pathogens11040414

Chicago/Turabian Style

Brasil-Costa, Igor, Cintya de Oliveira Souza, Leni Célia Reis Monteiro, Maria Elisabete Silva Santos, Edivaldo Herculano Correa De Oliveira, and Rommel Mario Rodriguez Burbano. 2022. "H. pylori Infection and Virulence Factors cagA and vacA (s and m Regions) in Gastric Adenocarcinoma from Pará State, Brazil" Pathogens 11, no. 4: 414. https://doi.org/10.3390/pathogens11040414

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