Hostname: page-component-7bb8b95d7b-s9k8s Total loading time: 0 Render date: 2024-10-05T14:14:43.321Z Has data issue: false hasContentIssue false

Helicobacter pylori virulence factors: subversion of host immune system and development of various clinical outcomes

Published online by Cambridge University Press:  13 June 2023

Roghayeh Mohammadzadeh
Affiliation:
Antimicrobial Resistance Research Center, Mashhad University of Medical Sciences, Mashhad, Iran Department of Microbiology and Virology, School of Medicine, Mashhad University of Medical Sciences, Mashhad, Iran
Shaho Menbari
Affiliation:
Antimicrobial Resistance Research Center, Mashhad University of Medical Sciences, Mashhad, Iran Department of Microbiology and Virology, School of Medicine, Mashhad University of Medical Sciences, Mashhad, Iran Department of Medical Laboratory Sciences, Faculty of Paramedical, Kurdistan University of Medical Sciences, Sanandaj, Iran
Abbas Pishdadian
Affiliation:
Department of Immunology, School of Medicine, Zabol University of Medical Sciences, Zabol, Iran
Hadi Farsiani*
Affiliation:
Antimicrobial Resistance Research Center, Mashhad University of Medical Sciences, Mashhad, Iran Department of Microbiology and Virology, School of Medicine, Mashhad University of Medical Sciences, Mashhad, Iran
*
Corresponding author: Dr Hadi Farsiani, Email: farsianih@mums.ac.ir
Rights & Permissions [Opens in a new window]

Abstract

Helicobacter pylori (H. pylori) is a worldwide spread bacterium, co-evolving with humans for at least 100 000 years. Despite the uncertainty about the mode of H. pylori transmission, the development of intra-gastric and extra-gastric diseases is attributed to this bacterium. The morphological transformation and production of heterogenic virulence factors enable H. pylori to overcome the harsh stomach environment. Using numerous potent disease-associated virulence factors makes H. pylori a prominent pathogenic bacterium. These bacterial determinants are adhesins (e.g., blood group antigen-binding adhesin (BabA)/sialic acid-binding adhesin (SabA)), enzymes (e.g., urease), toxins (e.g., vacuolating cytotoxin A (VacA)), and effector proteins (e.g., cytotoxin-associated gene A (CagA)) involved in colonisation, immune evasion, and disease induction. H. pylori not only cleverly evades the immune system but also robustly induces immune responses. This insidious bacterium employs various strategies to evade human innate and adaptive immune responses, leading to a life-long infection. Owing to the alteration of surface molecules, innate immune receptors couldn't recognise this bacterium; moreover, modulation of effector T cells subverts adaptive immune response. Most of the infected humans are asymptomatic and only a few of them present severe clinical outcomes. Therefore, the identification of virulence factors will pave the way for the prediction of infection severity and the development of an effective vaccine. H. pylori virulence factors are hereby comprehensively reviewed and the bacterium evasion from the immune response is properly discussed.

Type
Review
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press

Introduction

The spiral-shaped, Gram-negative microaerophilic bacterium, Helicobacter pylori (H. pylori) colonises the mucosal layer of the stomach epithelium of more than 50% of mankind (Reference Hooi1). H. pylori-human co-existence is ~100 000 years old indicating a great adaptation of H. pylori to its niche in the human gastric mucosa (Reference Moodley2, Reference Camilo, Sugiyama and Touati3). Recently, Hooi et al. and Zamani et al. conducted two separate systematic reviews and meta-analyses on H. pylori's global prevalence (Reference Hooi1, Reference Zamani4). A high prevalence of H. pylori was reported by the first meta-analysis (60.3%); the lowest prevalence was found in Northern America (37.1%) and Oceania (24.4%) and the highest prevalence in Africa (79.1%), Latin America, and the Caribbean (63.4%), and Asia (54.7%) (Reference Hooi1). A lower prevalence was suggested by the second meta-analysis (44.3%), ranging from 34.7% in developed countries to 50.8% in developing countries (Reference Zamani4). The presence of the organism in the oral cavity, water and food introduces them as potential routes of H. pylori transmission (Reference Burucoa and Axon5, Reference Stefano6). Considering that the oral cavity provides an optimal pH, temperature, and microaerophilic environment, oral–oral (particularly mother-to-child) is the prominent mode of H. pylori transmission in both developing and developed countries (Reference Kim and Kim7, Reference Yokota8). H. pylori dwelling in the oral cavity is assumed as a risk factor for the recrudescence of gastric H. pylori infection (Reference Sun and Zhang9). Recrudescence or re-infection leads to recurrence (Reference Sun and Zhang9). Therefore, the recurrence of H. pylori is considered a critical issue (Reference Zhao10). The pathogenicity of H. pylori is attributed to several mechanisms including (i) manipulation of the host signalling pathways, (ii) induction of indirect inflammatory responses within the gastric mucosa, and (iii) induction of direct epigenetic changes on gastric epithelial cells (Reference Machlowska11). Despite inducing an inflammatory response, the human immune system could not eradicate bacterial colonisation, resulting in a life-long infection (Reference Abadi12). H. pylori is a class-I carcinogen and it is known as a main risk factor for intra-gastric diseases including chronic gastritis (CG), peptic ulcer diseases (PUD), gastric cancer (GC), MALT lymphoma (MALToma) and biliary tract cancer (Reference Aviles-Jimenez13-Reference Ota17) as well as extra-gastric diseases, including cardiovascular, metabolic, and neurologic disorders (Reference Franceschi, Covino and Roubaud Baudron18). The complex interplay among the three major factors including: (i) bacterial virulence factors (e.g., cytotoxin-associated gene A (CagA), vacuolating cytotoxin A (VacA) and blood group antigen-binding adhesion (BabA), (ii) host genetic traits (e.g., tumour necrosis factor-alpha (TNF-α), interleukin 10 (IL-10) and IL-1β), and (iii) environmental factors (e.g., high salt, diet, meat consumption and smoke) would result in the diverse clinical outcomes from asymptomatic infection to GC (Reference Chang, Yeh and Sheu19-Reference Toh and Wilson21). Many factors, including geographic regions, living environment, socioeconomic status, educational level, and age are involved in the prevalence of H. pylori (Reference Wang22). Gut microbiome, obesity, male gender, consumption of unpasteurised dairy products, and high-risk occupations (e.g., healthcare and sheepherding) predispose individuals to H. pylori infection (Reference Assaad23-Reference Wang29). The development of various adaptations enables H. pylori to effectively colonise in harsh stomach conditions (Reference Camilo, Sugiyama and Touati3). Using various virulence factors, H. pylori colonises the stomach mucosa, escapes the immune response, and induces the disease (Table 1) (Reference Baj30). H. pylori passes through four steps to colonise, persist and cause serious disease in the human host: (i) overcoming the harsh acidic stomach condition by bacterial urease enzyme, (ii) moving toward epithelium cells by polar flagella and penetrating gastric mucosal barrier, (iii) the interaction of bacterial adhesins (e.g., CagL, CagY, OipA, HopQ, HopZ, BabA/B and SabA/B) with glycan structures (Gly-Rs) exposed on the external surface of gastric epithelial cells and the mucus layer, and (iv) releasing the toxins (e.g., VacA and CagA) causing tissue damages (Reference Kao, Sheu and Wu31). Besides, the virulence factors such as urease, flagellin, arginase and TlpB (a transmembrane chemoreceptor) are survival proteins and the virulence factors including neutrophil-activating protein A (NapA), γ-glutamyl transpeptidase (GGT), peptidoglycan (PG) and ADP-Heptose are immunoresponsive elements (Reference Knorr32). Furthermore, morphological transformation (from spiral to coccoid) is an interesting issue that has been recently considered owing to its vital role in the survival of H. pylori in the host gastric microenvironment (Reference Reshetnyak, Burmistrov and Maev33, Reference Reshetnyak and Reshetnyak34). Although H. pylori numerous virulence factors are responsible for the induction of local damage in gastric mucosa, the host's innate and adaptive immunity including the pattern recognition receptors (PRRs), pro-inflammatory cytokines, chemokines, chemotactic factors, monocytes, macrophages, neutrophils, natural killer cells (NK cells), dendritic cells (DCs), B cells, and T cells facilitate the development of subclinical systemic inflammation (Reference Săsăran, Meliț and Dobru35-Reference Guclu and Agan37). Therefore, the presence of H. pylori in the host gastric mucosa motivates the host's immune system to trigger systemic damage (Reference Mărginean, Mărginean and Meliț38). Despite this strong inflammatory response, H. pylori's ability to evasion, subversion and manipulation of the host immune responses guarantees the development of persistent infection in the stomach mucosa. Alteration in the surface molecules is an efficient strategy that preserves the bacterium from being recognised by the innate immune system (Reference Peek, Fiske and Wilson39). Besides, modulation of the function of the T cells by H. pylori disrupts the host's adaptive immunity against this bacterium (Reference Wen and Moss40). H. pylori prevention and eradication are among the challenging issues in the post-antibiotic era; therefore, providing alternative drugs or vaccine targets is an urgent need (Reference Mohammadzadeh41). In this regard, highlighting the importance of bacterial virulence factors in H. pylori pathogenesis is a promising strategy. This review provides comprehensive data on the H. pylori virulence factors and discusses this bacterium's evasion strategies from the immune response.

Table 1. Virulence factors involving in different aspects of H. pylori pathogenesis (colonization, immune escape, and disease induction), their functions, and associated diseases.

cagPAI, cag pathogenicity island; CagT4SS, Cag type IV secretion system; CagA, cytotoxin-associated gene A; VacA, vacuolating cytotoxin A; HtrA, high temperature requirement A; DupA, duodenal ulcer promoting gene A; IceA, induced by contact with epithelium gene A; OipA, outer inflammatory protein A; GGT, γ-glutamyl transpeptidase; BabA/B, blood group antigen-binding adhesin A/B; SabA/B, sialic acid-binding adhesin A/B; AlpA/B, adherence-associated proteins A/B; HopZ/Q, H. pylori outer membrane protein Z/Q; CagL/Y, cytotoxin associated gene L/Y; NapA, neutrophil-activating protein A; LabA, LacdiNAc-binding adhesin; LPS, lipopolysaccharide; HSP60, heat shock protein 60. ROS, reactive oxygen species; NO, nitric oxide; CG, chronic gastritis; GC, gastric cancer; MALToma, MALT lymphoma; PUD, peptic ulcer diseases; DU, duodenal ulcer.

The genome and virulence factors of H. pylori

The genome of H. pylori was among the first bacterial species which was completely sequenced.

According to the obtained data, H. pylori, strain 26 695, possesses a circular genome of 1 667 867 bp and 1590 predicted coding sequences (Reference Alm42, Reference Tomb JF, Kerlavage, Clayton, Sutton and Fleischmann43). The whole-genome sequencing of H. pylori revealed extraordinary genetic flexibility and a high frequency of gene recombination; this unique nature qualifies the bacteria to survive in harsh and dynamic habitats (Reference Cao44, Reference Noto45). Recognition of virulence factors might pave the way to the illustration of the H. pylori pathogenesis and prediction of the risk for inducing intra- and extra-gastric diseases (Reference Kabamba, Yamaoka and Shiotani46).

Cytotoxin-associated gene pathogenicity island

Cytotoxin-associated gene pathogenicity island (cagPAI) is existent in nearly 70% of all H. pylori strains isolated globally, compared with 60% of western isolates and 95% of East Asian isolates (Reference Olbermann47, Reference Yamaoka48). The cagPAI has been integrated into the H. pylori DNA via horizontal gene transfer (HGT), although its origin is unknown and its genes are not essential for H. pylori (Reference Censini49). cagPAI is a ~40 kb genomic region containing 32 open reading frames (ORFs), namely cag1-26, cagA-Z or cagα-ζ, or by locus name of the HP 26 695 or HP J99 strain genomes (Reference Blomstergren50). cagPAI encode effector protein CagA and type IV secretion system (CagT4SS), a syringe-like structure to inject CagA into gastric epithelial cells (Reference Backert, Tegtmeyer and Fischer51, Reference Chung52). Various bacterial molecules including CagA, DNA and PG metabolites are translocated into host cells by CagT4SS (Reference Backert53-Reference Viala55). Backert et al. suggested that intact CagT4SS consists of a core complex (CagT, CagX, CagM, Cagδ, and CagY), with associated factors (CagH, CagN, CagU, CagV, and CagW); pilus components (CagC, CagH, CagI, CagL, and CagY); and energetic components (CagE, Cagα, and Cagβ). Besides, translocation-associated factors (CagF, CagZ, and Cagβ), and a lytic trans-glycosylase (Cagγ) exist in CagT4SS (Reference Backert, Tegtmeyer and Fischer51). Although the complete composition of cagPAI guarantees encoding of intact CagT4SS, cagPAI is absent in nearly 30% of H. pylori strains, and it is incomplete in some strains (Reference Censini49, Reference Akopyants56). The severity of clinical outcomes induced by H. pylori is dependent on the integrity of cagPAI, therefore partial deletions within cagPAI reduce pathogenic features (Reference Patra57, Reference Nilsson58).

Cytotoxin-associated gene A

According to the presence or the absence of the cagA gene in H. pylori strains two main subpopulations are considered: cagA-positive and cagA-negative strains (Reference Censini49). The co-existence of cytotoxin-associated gene A (CagA)-positive and negative strains in a unique host is not far from expected (Reference Canzian59). Although CagA-positive strains cause more severe clinical outcomes which may progress to malignancy (Reference Baj30, Reference Sharndama and Mba60); the results of a meta-analysis study showed that these strains are more curable (Reference Wang61). CagA oncoprotein is the main virulence factor of H. pylori. Following the adherence of H. pylori to the host gastric epithelial cells, the cagA gene is expressed (Reference Chowdhury62). It was shown that CagA expression is regulated in response to environmental conditions including salt concentration, iron limitation, and pH (Reference Noto63-Reference Loh, Torres and Cover65). Via interaction with different cellular factors, CagA (128–145 kDa) interferes with numerous cellular signal transduction cascades and consequently induces gastric carcinogenesis (Reference Hatakeyama15, Reference Knorr32) (Fig. 1). Molecular anatomy shows that CagA includes a conserved N-terminal region, the variable repeats of the EPIYA (Glu-Pro-Ile-Tyr-Ala) motif, a tyrosine phosphorylation motif, and the C-terminal tail (Reference Hayashi66). According to distinct flanking amino acids around the EPIYA motifs, four different peptide segments, EPIYA-A, B, C and D have been determined (Reference Hatakeyama67). Almost all CagA isolates contain the EPIYA-A and B segments. EPIYA-C segment exists in CagA species that distribute in Europe, North America, and Australia; therefore, it is termed ‘Western CagA’. Also, the EPIYA-D segment exists in CagA species that circulate in East Asia; therefore, it is termed ‘East Asian CagA’ (Reference Hatakeyama68). Upon delivery into the host cell, CagA may attach to the inner surface of the cell membrane and its EPIYA motifs are tyrosine phosphorylated by Src and Abl family kinases, respectively (Reference Backert and Selbach69, Reference Nagase70). Subsequently, it binds Src homology 2 (SH2) domain-containing proteins (e.g., SHP2 tyrosine phosphatase and C-terminal Src kinase (Csk)) and adaptor protein Crk to disturb the adhesion, spreading, and migration of the host cell (Reference Yamazaki71-Reference Higashi74). The induction of NF-κB signalling and IL-8 secretion by CagA and CagT4SS leads to enhanced gastric inflammation, which is considered a risk factor for genetic instability and carcinogenesis (Reference Gorrell75). EPIYA-D motif's more binding tendency to SHP2 compared with the EPIYA-C motif suggests a more intense ability of East Asian CagA for induction of cellular transformations (Reference Hayashi76). Via induction of hypermethylation in the DNA promoters or histones, CagA also mediates epigenetic changes, leading to downregulation of the tumour suppressor genes (e.g., MGMT) or microRNAs (e.g., let-7) (Reference Hayashi77, Reference Sepulveda78). CagA destroys the apoptosis-stimulating protein of the p53 (ASPP2) tumour suppressor pathway and also interacts with the gastric tumour suppressor RUNX3 (RUNX family transcription factor 3), which leads to RUNX3 degradation by the proteasome (Reference Tsang79, Reference Buti80). Distinct CagA species have a different number of CagA-multimerisation (CM) motifs, also called CRPIA (conserved repeats responsible for phosphorylation-independent activity) sites, containing a 16-amino-acid sequence located in the CagA C-terminal region (Reference Ren81). Inside the gastric epithelial cells, the CM motif attaches to and downregulates the polarity-regulating kinase, partitioning-defective 1b (PAR1b), which is also termed the microtubule affinity-regulating kinase 2 (MARK2). The inhibition of PAR1b leads to failure of junction and polarity, which predisposes cells to oncogenesis (Reference Nishikawa82). Carcinoembryonic antigen-related cell adhesion molecules (CEACAMs) are a kind of intercellular adhesion molecules that act as functional receptors for H. pylori outer membrane adhesion HopQ (Reference Javaheri83, Reference Königer84). Recent genetic evidence showed that the interaction between HopQ and CEACAM receptors (CEACAM 1, 5, and 6) is critical for H. pylori CagA translocation while CagT4SS interaction with integrins (neither β1 integrin heterodimers (α1β1, α2β1 or α5β1), nor any other αβ integrin heterodimers) is not essential for this process (Reference Königer84-Reference Zhao86).

Figure 1. Signalling pathways induced by H. pylori CagA. CagA oncoprotein is directly injected into host cells by CagT4SS. This protein interacts with myriad signalling factors to manipulate different signal transduction cascades. CagA, cytotoxin-associated gene A; CagT4SS, Cag type IV secretion system; NFAT, nuclear factor of activated T cells; Bcl-2, B-cell lymphoma 2; c-Met, c-mesenchymal epithelial transition factor; Grb2, growth factor receptor-bound protein 2; GSK3β Wnt, glycogen synthase kinase 3β Wnt; SHP2, Src homology-2 domain-containing protein tyrosine phosphatase-2; TRAF, TNF receptor-associated factor; CD44, cluster of differentiation 44; Akt, Ak strain transforming; BIM, BCL-2-interacting mediator of cell death; MCL1, myeloid cell leukaemia-1.

Vacuolating cytotoxin A

All the H. pylori strains have a single chromosomal copy of the vacA gene. This gene encodes vacuolating cytotoxin A (VacA) protein that forms intracellular vacuoles in the eukaryotic host cells (Reference Foegeding87). vacA gene possesses allelic polymorphism owing to the existence of the signal peptide (s1a, s1b, s1c, and s2 variants), the intermediate (i1, i2, and i3 variants), and the middle regions (m1a, m1b, m1c, and m2 variants) which cause various clinical outcomes and toxic activity in different combinations (Reference Atherton88-Reference van Doorn92). Different combinations of s- and m-regions form four different vacA genotypes including s1m1, s1m2, s2m1, and s2m2; these variants have different abilities to induce vacuolisation in infected cells. The s1m1 variants are the most powerful strains in toxin production and vacuolisation in host cells. Cell vacuolation by s1m2 strains depends on the host cell line, s2m2 strains rarely produce cytotoxin, and s2m1 strains rarely induce cell vacuolation (Reference Pagliaccia91, Reference Rhead93-Reference Atherton96). All s1m1i1 strains are vacuolating and all s2m2i2 are non-vacuolating. S1m2i1 strains induce cell vacuolation, whereas s1m2i2 strains could not do this. Therefore, s1m1i1 and s1m2i1 strains are more virulent and more likely associated with serious clinical outcomes such as GC compared with the s2m2i2 and s1m2i2 strains (Reference Rhead93, Reference Ferreira97, Reference Ferreira, Machado and Figueiredo98). There are extra polymorphic regions including the deletion (d)-region (d1 and d2 variants), located between the i- and the m-region, and the c-region (c1 and c2 variants) located at the 3′-end region of vacA gene (Reference Ogiwara99, Reference Bakhti100). The d1- and c1- genotypes are considered biomarkers of a high risk of GC (Reference Ogiwara99, Reference Bakhti100). It seems that vacA expression is regulated in response to host–microbe interactions, salt and iron concentrations, and low pH (Reference Amilon101-Reference van Amsterdam104). The intact VacA toxin (~88 kDa) consists of an N-terminal domain (33 kDa) and a C-terminal domain (55 kDa) that are involved in cytotoxicity and binding to cell surface receptors, respectively (Reference Torres, McClain and Cover105, Reference Yahiro106). This toxin consists of a signal peptide and it is secreted by the type V secretion system (T5SS, autotransporter) to the intracellular milieu (Reference Chauhan107, Reference Fischer108). Epithelial cell's surface receptors including (i) low-density lipoprotein receptor-related protein-1 (LRP-1), (ii) receptor-like protein tyrosine phosphatase alpha and beta (RPTP-α, -β), and (iii) sphingomyelin acts as receptors for VacA protein (Reference McClain, Beckett and Cover109, Reference Yahiro110). Furthermore, it binds to β2 integrin (CD18) receptors on T cells (Reference McClain, Beckett and Cover109). VacA influences the host cells via induction of apoptosis, autophagy, membrane depolarisation, activation of mitogen-activated protein (MAP) kinases, inhibition of T-cell function and mitochondrial dysfunction contributing to H. pylori life-long colonisation and pathogenesis (Reference Foegeding87, Reference Terebiznik111-Reference Jain, Luo and Blanke116). For triggering apoptosis in the gastric epithelial cells, VacA inserts into mitochondrial membranes, and subsequently, cytochrome C is released (Reference Domańska117). VacA molecule's N-terminally encoded hydrophobic amino acids could form hexameric pores in the lysosomal, endosomal, and mitochondrial membranes of epithelial cells and phagocytes facilitating VacA's pro-apoptotic and vacuolating activity (Reference McClain118). Besides, VacA plays a critical role in H. pylori colonisation via inhibition of the proliferation and activation of B cells and T cells (Reference Torres119). During acute disease, VacA induces the autophagy pathways in host cells, whereas, during chronic disease, it induces the disruption of phagosomes and promotes cell vacuolation that facilitates the survival of H. pylori in the host epithelial cells (Reference Greenfield and Jones120, Reference Raju121).

CagA–VacA interactions

As H. pylori major virulence factors, VacA is expressed by all H. pylori strains, whereas CagA is exclusively expressed by specific strains (Reference Baj30). Deep research in bacterial life suggests a theory that bacteria may release various virulence factors able to act either in a synergistic manner or in an antagonistic manner to achieve greater compatibility with the host (Reference Ricci122). Using a functional relationship between CagA and VacA, H. pylori could fine-tune its interaction with the human host stomach (Reference Ricci122). Most of the time, VacA and CagA cellular activities are antagonistic; CagA downregulates the VacA-induced cellular vacuolation and VacA downregulates the CagA-induced cell alterations (Reference Tegtmeyer123). One of the most important functions of VacA is the induction of autophagy which degrades the CagA protein and shortens its half-life in the host cells. VacA-induced autophagy is downregulated in gastric cells and CagA escapes degradation, such as in the cancer stem cells expressing a cell-surface marker called CD44 variant 9 (CD44v9) (Reference Tsugawa124). Interestingly, a recent report provides contradictory evidence showing that in the absence of VacA, CagA undergoes both proteasomal and autophagic degradation (Reference Abdullah125). VacA causes the accumulation of dysfunctional autophagosomes via disrupting a late step of the autophagic pathway, which results in accumulation (rather than degradation) of CagA in the gastric epithelial cells. It seems that VacA-induced CagA accumulation in dysfunctional autophagosomes results in the vigorous limitation of CagA downstream signalling (Reference Abdullah125). CagA-induced suppression of autophagy via the c-Met-PI3 K/AKT-mTOR signalling pathway enables it to establish pro-inflammatory and carcinogenic action (Reference Li126). CagA obstructs the internalization of VacA to the host cells and leads to the blocking of VacA-induced apoptosis (Reference Akada127). CagA activates calcineurin and subsequently activates the transcription factor nuclear factor of activated T cells (NFAT), which exerts pleiotropic actions on cell proliferation and differentiation. Whereas VacA downregulates the NFAT by decreasing the calcium influx blocking calcineurin activation (Reference Yokoyama128). However, the co-participation of VacA and CagA in the iron acquisition by H. pylori is a rare example of their synergistic effect (Reference Tan129).

Urease

H. pylori urease is involved in the survival and colonisation of the bacterium and induces infection in the stomach (Reference Ansari and Yamaoka130, Reference Voland131). A gene cluster containing seven genes is regulated by two promoters that encode the H. pylori urease enzyme. The ureA and ureB genes (encoding catalytic units) are regulated by the first promoter. Downstream genes, including ureI (encoding acid-gated urea channel) and ure E–H (encoding accessory assembly proteins), are under the control of the second promoter (Reference Akada132, Reference Marcus, Sachs and Scott133). In H. pylori species, urease activity is regulated by the availability of the cofactor nickel and the pH of the stomach microenvironment (Reference Weeks134, Reference Belzer135). It was shown that the existence of nickel in the bacterial culture medium dramatically increases urease activity (Reference Belzer135). Besides, the acid-gated urea channels are closed at pH 7.0 and are firmly open at pH 5.0 (Reference Kao, Sheu and Wu31, Reference Weeks134). Urease is a vital virulence factor to the survival of H. pylori in the human host stomach; recently, Madison et al. demonstrated that in response to nitric oxide (NO), as a product of the host innate immunity, CrdRS two-component system (TCS) regulates the expression of ureA gene (Reference Allen136). Urease is a polymeric enzyme (5–10% of the total protein content) that is composed of two subunits, i.e., UreA (29.5 kDa) and UreB (66 kDa), of which UreB is considered the subunit responsible for the enzyme activity (Reference Strugatsky137). It was shown that a cluster of the 12 active sites containing 24 nickel ions on the urease supramolecular structure guarantees enzymatic action and H. pylori survival at low pH (Reference Ha138). The urease enzyme catalyses the urea conversion to ammonia and carbon dioxide, raising the acidic stomach pH to neutral to protect H. pylori from acidity via the formation of a cloud of ammonia which neutralises acidic pH (Reference Dunn and Phadnis139). Following the production of ammonia, a soluble form of occludin, a 65-kDa tetraspan integral membrane protein, is accumulated which leads to the disruption of tight epithelial junctions (Reference Lytton140). Besides, high levels of ammonia cause cytotoxic effects on the gastric epithelial barrier and ruin these cells’ mitochondrial oxygenation (Reference Schoep141). Natural pH generated by the urease enzyme reduces mucin viscoelasticity to convert the gastric mucin structure from gel to sol, leading to bacterial free movement through the mucus (Reference Celli142). Based on localisation, there are internal and external types of urease enzymes. The internal urease is produced by live bacterial cells and is active at pH 2.5–6.5, whereas the external urease is produced during cell lysis and is active at pH 5.0–8.5 (Reference Scott143). Increased urease activity may lead to a higher risk of induction of histopathological alterations within the gastric mucosa and greater gastric carcinogenesis (Reference Ghalehnoei144).

Helicobacter outer membrane porins

H. pylori adherence to the gastric epithelium is necessary for the delivery of toxins (e.g., CagA and VacA) or other virulence factors into the host cells, which results in inflammatory or immune response-mediated direct or indirect damages (Reference Sharndama and Mba60). The interaction between the receptors expressed on the gastric epithelial cells and Helicobacter outer membrane porins (Hop) family, adhesion factors encoded by hop genes, starts H. pylori infection (Reference Oleastro and Ménard145). Hop family porins including HopS, HopP, HopH, HopQ, and HopZ augment H. pylori adherence to the host cell and boost inflammation by promoting the expression of virulence factors and the secretion of inflammatory cytokines (Reference Xu146). Widely studied among the Hop family are blood group antigen-binding adhesin (BabA) and sialic acid-binding adhesin (SabA) (Reference Doohan147). A wide variety of receptors for BabA and SabA binding activity have been found in the human host stomach and saliva (Reference Xu146, Reference Doohan147). BabA (HopS) binds to H-type 1 and ABO/Lewisb (Leb) blood group antigens exposed on the gastric epithelium and mucus layer (Reference Ansari and Yamaoka148). According to X-ray analysis, it was shown that BabA contains three structural domains including one conserved loop (CL2) and two diversity loops (DL1 and DL2) for interaction with Leb (Reference Moonens149). BabA protein has two domains including the extracellular N-terminal domain linked to Leb antigens and an outer membrane C-terminal domain that anchors into the outer membrane (Reference Hage150). BabA not only contributes to bacterial adherence and colonisation but also augments a nonspecific immune including granulocyte infiltration or secretion of IL-8 enhancing gastric inflammation (Reference Rad151). BabA adherence to fucosylated Leb antigen facilitates CagT4SS activity which leads to the release of high amounts of the pro-inflammatory factors inducing carcinogenesis (Reference Ansari and Yamaoka148, Reference Ishijima152). Low BabA production and decreased binding tendency to the Leb leads to detachment of H. pylori from the gastric mucus and induces ulceration within the duodenum and consequently increases the risk of PUD (Reference Saberi153). The existence of BabA in Western countries is correlated to the high prevalence of PUD and GC, but there is no such correlation in Asians (Reference Chen154). SabA (HopP) is an adhesion molecule, which is involved in H. pylori binding and colonisation via the interaction with sialyl-Lex, sialyl-Lea , and Lex, but not with other Lewis's antigens, such as Lea, Leb or Ley (Reference Mahdavi155, Reference Pang156). SabA-expressing strains could promote gastric diseases, redundant neutrophil infiltration and gastric atrophy during infection; furthermore, these strains could vastly colonise (Reference Sheu157, Reference Yanai158). Recently, it was shown that the tropism of BabA along with SabA for spasmolytic polypeptide-expressing metaplasia (SPEM) glands, enables H. pylori to induce metaplastic alterations and trigger the onset of carcinogenesis (Reference Sáenz, Vargas and Mills159). H. pylori's successful colonisation and persistent infection are largely dependent on two major adhesins: including BabA and SabA. Therefore, these adhesins can be considered potential vaccine candidates against H. pylori (Reference Doohan147, Reference Keikha160-Reference Urrutia-Baca162).

γ-Glutamyl-transpeptidase

γ-Glutamyl-transpeptidase (GGT) is a 61 kDa protein that catalyses the conversion of glutamine into glutamate and ammonia as well as glutathione into glutamate and cysteinyl glycine. This enzyme enables H. pylori to take up and incorporate glutamate into the tricarboxylic acid (TCA) cycle (Reference Ricci163). Initially, it was thought that GGT involves in H. pylori colonisation but more studies confirmed that GGT knockout mutant strains can colonise animals (Reference McGovern164, Reference Oertli165). The GGT expression and activity may promote the development of peptic ulcer disease (PUD) (Reference Gong166). The apoptotic effect on the host cells has been recognised for H. pylori and H. suis GGT (Reference Flahou167). However, it seems that GGT is not the main factor in the induction of H. pylori-mediated apoptosis in T cells because a GGT-negative mutant still induces T-cell apoptosis (Reference Wessler168). H. pylori GGT induces cell cycle arrest at the G1/S phase transition in T cells and gastric epithelial cells, and subsequently inhibits their proliferation (Reference Schmees169, Reference Kim170). It was demonstrated that GGT and VacA could inhibit T cells which emphasises the role of secreted virulence factors in the H. pylori pathogenesis (Reference Wessler168). GGT and VacA promote immune tolerance and gastric persistence that facilitate the prevention of asthma in vivo (Reference Oertli165, Reference Engler171).

Tumour necrosis factor-alpha (TNF-α)-inducing protein

Various H. pylori strains have a conserved sequence of tipα gene encoding tumour necrosis factor-alpha (TNF-α)-inducing protein (Tipα) as a small and secretory protein (Reference Suganuma172). It was shown that Tipα has a weak homology to Gram-positive bacterial penicillin-binding proteins. This led us to conceive that the tipα gene has been derived from Gram-positive bacteria and transferred horizontally to H. pylori (Reference Kuzuhara173, Reference Suganuma174). The Crystal Structure of the Tipα indicates that the functional Tipα protein (approximately 37 kDa) is a homodimer one (Reference Gao175, Reference Jang176). According to the well-established data, a functional T4SS induces epithelial cells to produce cytokines against H. pylori infection (Reference Backert, Tegtmeyer and Selbach177). In addition to the mentioned pathway, it was found that Tipα is a strong inducer of pro-inflammatory cytokine and chemokine gene expressions (Reference Kuzuhara178). Through the activation of nuclear factor kappa B (NF-κB), Tipα induces the overexpression of TNF-α in the Bhas 42 (BALB/3T3 cells transfected with v- H-ras gene) and MGT-40 cells (mouse gastric epithelial cell line) (Reference Suganuma174, Reference Suganuma179). TNF-α is well characterised as a master tumour promotor. It is a key regulator of inflammation that plays a critical role in the cytokine network between cancer and inflammation (Reference Bauer180). Accordingly, Tipα is a potent carcinogenic factor that could induce inflammation and/or CG, hyperplasia, and GC in individuals infected with cagPAI-negative H. pylori strains (Reference Suganuma174, Reference Suganuma179, Reference Morningstar-Wright181). Tipα plays a critical role in the colonisation of mouse gastric mucosa (Reference Godlewska182), however, its secretion is independent of the CagT4SS (Reference Suganuma179). Tipα can bind a membrane receptor called nucleolin and subsequently, it can internalise into the gastric epithelial cell's cytoplasm (Reference Watanabe183). Interestingly, the external ligands of nucleolin including hepatocyte growth factor (HGF), K-ras and Tipα are carcinogenic (Reference Fujiki, Watanabe and Suganuma184). Tipα interacts with single- and double-strand forms of DNA in the GC cells suggesting that DNA binding may have a possible role in the molecular mechanisms of carcinogenesis (Reference Suganuma185). Turning to a monomer form leads to the loss of Tipα's functional performances in tumour promotion, TNF-α induction and NF-kB activation (Reference Suganuma174, Reference Suganuma179). Besides, the monomer form of Tipα has poor DNA-binding ability (Reference Kuzuhara186).

Morphological transformation

H. pylori is typically known as spirally twisted rods, whereas its highly heterogenic nature can form various cell shapes, including coccoid forms, elongated (filamentous) forms, and straight or curved rods (Reference Krzyżek and Gościniak187). During the morphological transition to coccoid forms, H. pylori loses its culturability which could be a sign of bacterial death. Using highly advanced genetic and microbiological techniques, it was shown that these cells are alive and they have modified their physiology (Reference Azevedo188). Decreased cell size and dramatic limitations in metabolic activity lead to H. pylori morphological transformation into coccoid form, which translates into a transition to a viable but non-culturable (VNC) phenotype (Reference Azevedo188). Spiral viable culturable form (SVCF) and coccoid viable but non-culturable form (CVNCF) are two distinct forms distinguished via molecular techniques and electron microscopy (Reference Elhariri189). However, several studies have proposed that coccoid forms colonise the mucus layers, produce virulence factors, escape the immune responses, promote carcinogenesis, and play a critical role in therapeutic failures (Reference Reshetnyak, Burmistrov and Maev33, Reference Sisto190-Reference Chaput195). Therefore, blocking the process of morphological transformation of H. pylori might be a promising strategy for the successful eradication of this pathogen (Reference Krzyżek and Gościniak187, Reference Krzyżek and Grande193). Myricetin (MYR; 3,5,7,3′,4′,5′ hexahydroxyflavone, a natural anti-virulence compound) interferes with the morphological transformation of H. pylori from spiral/rod-shaped forms to coccoid forms and increases the activity of antibiotics against this pathogen (Reference Krzyżek196). It was shown that following exposure to MYR, genes related to the H. pylori morphogenesis (e.g., csd3, sd6, csd4, and amiA) were downregulated. These suppressed genes are mostly involved in the shortening of muropeptide monomers, suggesting their major role in the spiral-to-coccoid transition (Reference Krzyżek196). Cultivation under mild sub-optimal growth conditions such as acidic and alkaline pH, high temperature, aerobiosis, extended incubation, and exposure with a proton pump inhibitor and/or antibiotics stimulates H. pylori morphological conversion from spiral to coccoid (Reference Azevedo188, Reference Shahamat197-Reference Bode, Mauch and Malfertheiner202). It is postulated that commensal bacteria undergo an increased cell filamentation, whereas pathogenic bacteria undergo a reduction of elongation (Reference Rossetti203). In agreement with this hypothesis, it was demonstrated that highly virulent strains had shorter cells than the lower virulent strains; indicating that adaptational changes in cell morphology drastically associate with the virulence profile of H. pylori strains (Reference Krzyżek, Biernat and Gościniak204).

Interaction between H. pylori and the human host immune system: a life-long challenge

Although H. pylori was originally considered an extracellular germ, it was shown that a small portion of this pathogenic bacteria can penetrate the intracellular compartments of various cell types (Reference Ricci, Romano and Boquet205, Reference Dubois and Borén206). It seems that the intracellular lifestyle protects H. pylori from the antibiotics and host immune responses which facilitate persistent infection (Reference Lina207). Despite the induction of intense pro-inflammatory response by various Gram-negative bacteria, H. pylori stimulates only a relatively feeble immune response which results in asymptomatic and persistent infection. Various unique immune evasion strategies mediated by numerous virulence factors are raised H. pylori-derived life-long infection (Reference Sijmons208, Reference Neuper209). These strategies include (i) survival in a harsh habitat, (ii) changing macrophages and DCs functions, (iii) induction of the secretion of the anti-inflammatory cytokines, and (iv) the stimulation of Treg cell development (Reference Kaebisch210-Reference Altobelli213). It was shown that after exposure to live H. pylori, human monocytes initially secret a mixture of pro- and anti-inflammatory cytokines, whereas the prolonged-term innate memory promotes the secretion of anti-inflammatory cytokines (Reference Frauenlob212).

Innate immunity evasion strategies

Escape from innate immune recognition

The highly conserved pathogen-associated molecular patterns (PAMPs) expressed by H. pylori are recognised by PRRs that are expressed by the innate immune cells and gastric epithelial cells (Reference Kawai and Akira214). The interactions between PAMPs and PRRs trigger innate immune responses which are followed by adaptive immune responses (Reference Castaño-Rodríguez, Kaakoush and Mitchell215, Reference Mogensen216). PRR families are classified into transmembrane receptors (Toll-like receptors (TLRs) and C-type lectin receptors (CLRs)) and intracellular receptors (nucleotide-binding oligomerisation domain- (NOD-) like receptors (NLRs) and retinoic acid-inducible gene- (RIG-) I-like receptors (RLRs)) (Reference Akira, Uematsu and Takeuchi217, Reference Takeuchi O218). PAMPs sensation by PRRs activates intracellular signal transduction pathways triggering stormy reactions such as antimicrobial activity and inflammatory response to eliminate the pathogenic microorganism (Reference Akira, Uematsu and Takeuchi217). The TLRs are major classes of PRRs involved in the recognition of PAMPs, such as lipopolysaccharide (LPS), lipoteichoic acid (LTA), hypo-methylated CpG-rich regions of DNA, lipoprotein (LP), flagellin and PG (Reference Takeda and Akira219). Numerous scientific reports explain the role of TLRs during H. pylori infection (Reference Neuper220-Reference Koch222). Since several PAMPs of H. pylori trigger anti- but not pro-inflammatory responses, H. pylori could evade innate and adaptive immune responses to survive and persist in life-long infection (Fig. 2) (Reference Müller and Hartung223). Escherichia coli hexa-acylated LPS with a negative charge and potent immunostimulatory properties activates inflammatory signalling via the sensation of TLR4, whereas H. pylori tetra-acylated LPS with less negative charge could not be recognised by TLR4 and resists antimicrobial peptides (Reference Cullen224). Recently, Schmidinger et al. found that lipid A part of H. pylori LPS interacts with human annexins and dramatically suppresses LPS-mediated TLR4 signal transduction which subsequently hinders the innate immune system (Reference Schmidinger225). The N-terminal position of flagellin in Salmonella enterica could be detected by TLR5, but this critical position (TLR5 binding site) has mutated in H. pylori (Reference Andersen-Nissen226). Additional H. pylori PAMPs-PRRs interactions include the detection of H. pylori hypo-methylated CpG DNA by endosomal TLR9, and sensation of H. pylori LPS by TLR2 (Reference Akira, Uematsu and Takeuchi217, Reference Smith227, Reference Hemmi H, Kawai, Kaisho, Sato, Sanjo, Matsumoto, Hoshino, Wagner, Takeda and Akira228). These two TLRs induce anti-inflammatory and tolerogenic responses causing persistent infection (Reference Otani229-Reference Sun231). H. pylori 5ʹ triphosphorylated RNA is recognised by endosomal TLR8; this triggers a downstream signalling pathway that results in the transactivation of type I interferon (IFN-I (IFN-α/β)) in human monocytes (Reference Akira, Uematsu and Takeuchi217, Reference Lee232). H. pylori 5ʹ triphosphorylated RNA is also detected by one kind of RLRs known as RIG-I. The stimulation of the RIG-I receptor leads to the activation of the transcription factors IRF3 and IRF7 and subsequent expression of IFN-I (Reference Rad233). According to in vitro and ex vivo experiments, recently it was shown that through the downregulation of IRF3 activation, H. pylori actively inhibits cyclic GMP-AMP (cGAMP) synthase (cGAS)-stimulator of interferon genes (STING) and RIG-I signalling (Reference Dooyema234). DCs-specific intercellular adhesion molecule-3 grabbing non-integrin (DC-SIGN) is a member of the CLR family which plays a critical role in H. pylori recognition and pathogenesis. DC-SIGN ligands are mannosylated in most of the pathogens causing activation of the pro-inflammatory pathways. Although DC-SIGN ligands of H. pylori are fucosylated leading to the suppression of the signalling pathways downstream of DC-SIGN and also causing activation of the anti-inflammatory genes (Reference Gringhuis235). The macrophage-inducible C-type lectin (MINCLE) is another CLR, it interacts with the Lewis antigens of H. pylori LPS inducing macrophages to secrete low levels of TNF-α and high levels of IL-10 promoting H. pylori persistence (Reference Devi, Rajakumara and Ahmed236). NOD1 (CARD4) and NOD2 (CARD15) are two members of NLRs that can recognise various motifs in PG of various Gram-negative pathogenic bacteria (Reference Girardin237, Reference Girardin238). PG delivered by the H. pylori CagT4SS is sensed by NOD1 in the cytoplasm of epithelial cells and causes the activation of NF-κB signalling and upregulation of pro-inflammatory immune responses (Reference Viala55). It was shown that the immunomodulatory glycoprotein olfactomedin 4 (OLFM4) targets NOD1 and NOD2 in H. pylori-infected cells (Reference Liu239). OLFM4 is a target gene of the NF-κB pathway and associates directly with both NOD1 and NOD2 proteins thereby having a negative feedback effect on NF-κB activation induced by H. pylori infection. It was shown that the knockout of the OLFM4 gene in mice reduces H. pylori loads and increases gastric immune cell infiltration. Via a negative regulatory effect on H. pylori-specific NOD-mediated immune responses, OLFM4 plays an impressive role in the persistence of H. pylori colonisation (Reference Liu239). The bacterial ligand such as PG can access intracellular NOD-1 by CagT4SS and outer membrane vesicles (OMVs) (Reference Minaga240). PG deacetylation allows H. pylori to evade host innate immunity via induction of an NOD-1-dependent negative feedback loop. It seems that NOD-1 alters macrophage polarisation leading to H. pylori persistence (Reference Suarez241). Bacterial sugars including ADP-β-D-manno-heptose (ADP-Heptose) and D-glycero-β-D-manno-heptose-1,7-bisphosphate (βHBP) are key intermediate metabolites of LPS inner heptose core. The biosynthesis of ADP-Heptose and βHBP is a novel potent and cagPAI-mediated cell activation pathway triggering NF-κB augmentation and IL-8 release in human epithelial cells (Reference Pfannkuch242, Reference Zimmermann243). Upon CagT4SS-mediated internalization of βHBP into the gastric epithelial cells, it is sensed by α-kinase 1 (ALPK1) and TNF receptor-associated factors (TRAF)-interacting protein with forkhead-associated domain (TIFA). The activation of the ALPK1-TIFA signalling axis leads to the activation of NF-kB signalling (Reference Pfannkuch242-Reference Zhou244). Recently, Maubach et al. showed that the TIFA has dual functions in H. pylori-induced NF-κB pathways; induction of NF-κB classical pathway via the association of TIFA with TRAF6, and induction of NF-κB alternative pathway via the association of TIFA with TRAF2/TRAF3 (Reference Maubach245). The TIFA is an intrinsic anti-parallel dimer, containing a threonine residue (Thr9), a central forkhead domain (FHA), and a C-terminal TRAF6 binding site (T6BP). Following βHBP sensation, Thr9 is phosphorylated and then intermolecular pThr9–FHA interactions lead to head-to-tail oligomerisation. The TIFAsome (including TIFA oligomers, TRAF2, and additional host factors) recruits and activates TRAF6, triggering NF-κB activation and subsequent inflammatory signalling (Reference Gaudet and Gray-Owen246). Following TIFA activation, PG delivered through the CagT4SS activates NOD1 which leads to NF-κB-mediated pro-inflammatory responses. Within hours of infection and before NOD1 activation, the ALPK1-TIFA signalling pathway is activated and triggers strong NF-ҡB-dependent inflammation. Finally, the interaction of CagA (delivered through the CagT4SS) and host transforming growth factor (TGF)-β-activated kinase 1 (TAK1) triggers NF-κB-mediated pro-inflammatory responses (Reference Gall247). Fig. 3 illustrates the model of H. pylori CagT4SS-mediated, NF-κB-driven innate immune response in gastric epithelial cells. The successive activation of the ALPK1-TIFA signalling pathway and NOD1, and CagA delivery trigger the initial inflammatory response in gastric epithelial cells, motivate the subsequent recruitment of immune cells and lead to CG (Reference Gall247); however, the contribution rates of these pathways to natural infection are not clear. It was reported that mutation in the genes required for the synthesis of βHBP (e.g., rfaE, gene HP0858 in strain 26 695) dramatically reduces IL-8 induction (>95%) and ruins CagT4SS-dependent cellular signalling (Reference Tomb JF, Kerlavage, Clayton, Sutton and Fleischmann43, Reference Stein248). It was shown that ADP-Heptose, a derivative of βHBP, was present in H. pylori at more concentration (10 times) than its origin. Simultaneously, ADP-Heptose was dramatically more potent and cells distinctly recognise the existence of the β-form. The aforementioned findings revealed that ADP-Heptose is not only a new potent NF-kB-activating PAMP in H. pylori but also it is a general Gram-negative bacteria-derived PAMP (Reference Pfannkuch242). Furthermore, ALPK1 is a cytosolic innate immune receptor for bacterial ADP-Heptose (Reference Zhou244). The activation of the ALPK1-TIFA-NF-kB axis and CagA translocation are two distinct functions of CagT4SS because, despite the mutation in bacterial rfaE or the host ALPK1, CagA internalization continues (Reference Zimmermann243). It was demonstrated that cagPAI proteins enhance the colonisation rate via the suppression of antimicrobial peptides (Reference Patel249). The inverse correlation between gastric β-defensin 1 level and colonisation in H. pylori carriers is attributed to the reduced expression of human β-defensin 1 in a CagT4SS-dependent manner (Reference Patel249). β-defensin 3 is another human antimicrobial peptide, with strong anti-H. pylori activity. At the beginning of in vitro infection, β-defensin 3 is induced in an MAP kinase- and epidermal growth factor receptor (EGFR)-dependent manner. The induction of β-defensin 3 is followed by the stable shutdown through CagA-mediated activation of the Src homology domain, containing protein tyrosine phosphatase 2 (SHP2) and downmodulation of the EGFR signalling pathway (Reference Bauer250).

Figure 2. H. pylori evasion of innate immune recognition. Structurally modified pathogen-associated molecular patterns (PAMPs) enable H. pylori to evade the detection by pro-inflammatory Toll-like receptors (TLRs). H. pylori tetra-acylated LPS is less biologically active and is not sensed by TLR4. The TLR5 cannot detect the mutated TLR5 binding site of H. pylori flagellin. TLRs 9 detects H. pylori DNA (CpG DNA) and TLR2 detects H. pylori LPS; these TLRs predominantly activate anti-inflammatory signalling pathways and IL-10 expression. The cytosolic receptor RIG-I and endosomal receptor TLR8 sense 5ʹ triphosphorylated RNA which elicit IFN-α/β response. H. pylori fucosylated DC-SIGN ligands are another activator of anti-inflammatory genes. Besides, theses ligands block IFN-α/β response. Notice that the different cell types express various pattern recognition receptors (PRRs). TLR2/4/5/8/9, Toll-like receptor 2/4/5/8/9; LPS, lipopolysaccharide; DC-SIGN, dendritic cell-specific intercellular adhesion molecule-3 grabbing non-integrin; SRC, steroid receptor coactivator; IFN-α/β, α/β interferon; IRF3/7, interferon regulatory factor3/7; AP-1, activator protein 1; NF-κB, nuclear factor-κB; P50/P65, NF-κB p50/p65 heterodimer; TIR, Toll/interleukin-1 receptor domain; MYD88, myeloid differentiation primary response gene 88; DD, death domain; CARD, caspase activation and recruitment domain; RIG-I, retinoic acid-inducible gene I, CpG DNA, 5'—C—phosphate—G—3'.

Figure 3. CagT4SS-injected virulence factors (e.g., ADP-Heptose, CagA and PG) trigger NF-κB-mediated innate immune response in gastric epithelial cells. Upon binding of ADP-Heptose, an intermediate metabolite produced during the biosynthesis of LPS, to ALPK1, the ALPK1-TIFA signalling pathway is triggered, which results in an NF-kB-dependent pro-inflammatory response. Following TIFA activation, PG activates NOD1, which leads to NF-κB-mediated pro-inflammatory responses. Finally, the interaction of CagA and host TAK1 triggers NF-κB-mediated pro-inflammatory responses. Besides, CagT4SS-injected CpG DNA is detected by TLR9 which result in activation of anti-inflammatory signalling pathways. LPS, lipopolysaccharide; CagT4SS, Cag type IV secretion system; ADP-Heptose, ADP-β-D-manno-heptose; PG, peptidoglycan; CagA, cytotoxin-associated gene A; ALPK1, alpha kinase 1; Tak1, TGF-β activated kinase 1; TRAF6, tumour necrosis factor receptor (TNFR)-associated factor 6; TIFA, TRAF interacting forkhead-associated protein A; NOD1, nucleotide-binding oligomerisation domain 1; RIPK2, receptor-interacting-serine/threonine-protein kinase 2; NF-κB, nuclear factor-κB; P50/P65, NF-κB p50/p65 heterodimer; CpG DNA, 5'—C—phosphate—G—3'; TLR9, Toll-like receptor 9; TIR, Toll/interleukin-1 receptor domain; DD, death domain.

Manipulation of innate immune cells

The innate immune cells in the lamina propria and the mucus layer on the surface of gastric epithelial cells are the main defensive barrier against H. pylori (Reference Chmiela251). H. pylori CagA in macrophages interacts with SHP-1 and targets TRAF6 for K63-linked ubiquitination, thereby this multi-functional protein can down-regulate the expression of pro-inflammatory cytokines and subsequent immune response (Reference He252). Gang Liu et al. showed that cagA-positive H. pylori strains abrogate Cathepsin C (CtsC) to ruin neutrophil activation; this process prevents bacterial clearance and guarantees persistent infection (Reference Gang Liu253). It was shown that the phosphorylated form of CagA over-expresses the gene encoding haem oxygenase-1 which leads to macrophage anti-inflammatory responses and causes persistent infection (Reference Gobert254). It was demonstrated that CEACAM1 expressed mainly by activated immune cells, including NK cells and T cells acts as a suppressive receptor (Reference Gray-Owen and Blumberg255). Recently, Gur et al. showed that the secretion of IFN-γ by CD4+ T cells is suppressed by HopQ; also, NK cell and T-cell functions are inhibited by HopQ-mediated activation of CEACAM1 which may result in immune cells’ harness (Reference Gur256). HopQ–CEACAM interaction modulates the expression and secretion of chemokines in immune cells which leads to H. pylori's survival within neutrophils in a HopQ-dependent manner (Reference Behrens85). Via the suppressive effects on DCs and inductive effect on macrophages, VacA suppresses IL-23 expression and induces IL-10 and TGF-β secretion. The VacA immunomodulatory activity built a tolerogenic environment for H. pylori and leads to a chronic infection (Reference Altobelli213). The reduction of proliferation of immune cells, including neutrophils, macrophages, eosinophils, DCs, B cells, and T cells is also attributed to the VacA protein (Reference Foegeding87, Reference Djekic and Müller257, Reference Utsch and Haas258). VacA downregulates class II major histocompatibility complex (MHC II)-dependent pathways and via the formation of vesicular compartments inside macrophages assists H. pylori's intracellular survival (Reference Zheng and Jones259). NapA manipulates both innate and adaptive immune responses by: (i) upregulation of MHC II, (ii) promotion of T helper1 (Th1) cell differentiation, and (iii) stimulation of IL-12 and IL-23 release from neutrophils and monocytes (Reference Amedei260). Another potent immunomodulator is the urease enzyme; it (i) modifies opsonisation, (ii) augments the chemotaxis of neutrophils and monocytes, (iii) binds to the MHC II receptors, and subsequently facilitates apoptosis, and (iv) increases the secretion of the pro-inflammatory cytokines (Reference Schmalstig261). GGT downregulates T-cell proliferation and DCs differentiation (Reference Oertli165, Reference Schmees169, Reference Gerhard262, Reference Beigier-Bompadre263). To achieve life-long survival in the host, H. pylori downregulates macrophage phagocytosis and inhibit NO production by the induction of macrophage arginase II (ARG2) (Reference Ramarao264, Reference Lewis265). Inhibition of cell division-associated genes by H. pylori attenuates the proliferation of macrophages (Reference Tan266). Besides, H. pylori promotes mitochondrial membrane depolarisation and hydrogen peroxide secretion which induce apoptotic program in macrophages (Reference Chaturvedi267, Reference Asim268). Persisting a life-long H. pylori infection in the host depends on the inhibition of macrophage-mediated functions including phagocytosis, human leukocyte antigen-II (HLA-II) expression and IFN-γ production which results in T-cell suppression (Reference Cheok269). Via the glucosylation of cholesterol, H. pylori escapes from macrophage phagocytosis and could survive (Reference Yang and Hu270). Morey et al. showed that H. pylori depletes cholesterol in gastric glands to inhibit IFN-γ signal transduction and evade inflammatory response (Reference Morey271). Antigen presentation to T cells is one of the most important functions of DCs; abrogation of this function by H. pylori blocks Th1 cell differentiation (Reference Mitchell272).

Adaptive immunity evasion strategies

Despite the high prevalence of H. pylori infection among the human race, the majority of the infected individuals are asymptomatic for life, and only a minority of them develop H. pylori infection-related disease (Reference Müller and Hartung223, Reference Alexander273). Studies on human carriers and mouse models show that the polarisation and severity of the H. pylori-specific Th cell responses are major predictors and drivers of the disease (Reference Arnold274). Compared with the asymptomatic carriers, the PUD patients have a threefold higher anti-H. pylori Th1 cell response, twofold lower Treg response, sixfold higher Th2 response, and dramatically reduced levels of TGF-β and IL-10 in the gastric mucosa (Reference Robinson275). This imbalance suggests an association between inadequate Treg response and the development of H. pylori-derived disease (Reference Robinson275). It was shown that the severity of gastritis has an inverse correlation with the number of gastric Tregs and the cytokines secretion by Tregs (Reference Harris276). These findings are in line with previous work showing that Tregs gather in the gastric mucosa and quench exclusive memory T-cell responses in the infected but not uninfected patients (Reference Lundgren277, Reference Lundgren278). Treg responses to the infection are predominantly launched in the asymptomatic (healthy) carriers resulting in effectively damping immunopathologic reactions and promotion of the persistent infection. Although T-effector-dominated responses are predominantly expressed in symptomatic carriers that promote the disease (Reference Müller and Hartung223). Using two critical cytokines including IL-10 and TGF-β, Treg cells can inhibit T-effector cell-driven immunopathology (Reference Arnold274). Thus, the expression level of these cytokines in the gastric mucosa could effectively predict H. pylori-induced clinical outcomes (Reference Müller and Hartung223). It was shown that the downregulation or neutralisation of IL-10 signalling could effectively trigger strong T-cell-dependent immunopathology and clear H. pylori (Reference Sayi279). Comparison of T-cell responses of the symptomatic versus asymptomatic carriers as well as children with mild gastritis versus adults with severe gastritis suggests that Treg/T-effector cell ratios are associated with the clinical outcome (Reference Robinson275, Reference Harris276). Since vaccination is the only effective strategy to achieve protective immunity, understanding the importance of T-effector versus Treg responses to establish H. pylori clearance or immunopathology is a major issue in H. pylori vaccinology (Reference Hitzler280). For successful colonisation, H. pylori must suppress effector T cells (Th1 and Th17 subsets) activity, proliferation and clonal expansion. GGT and VacA are two critical virulence factors devastating T-cell-mediated immunity (Reference Müller and Hartung223) (Fig. 4). β2 integrin subunit of the heterodimeric transmembrane receptor lymphocyte function-associated antigen-1(LFA-1) acts as a receptor for hexameric VacA (Reference Sewald281). Subsequent ligand–receptor binding, VacA is entered upon protein kinase C-mediated serine/threonine phosphorylation of the β2 integrin cytoplasmic tail (Reference Sewald, Jiménez-Soto and Haas282). Cytoplasmic VacA blocks NFAT dephosphorylation by the Ca2+/calmodulin-dependent phosphatase calcineurin. Then the nuclear transfer of NFAT was prevented resulting in the downregulation of IL-2 production as well as subsequent T-cell activation and proliferation (Reference Müller and Hartung223). GGT could arrest T-cell proliferation in the G1 phase of the cell cycle through the disruption of the Ras signalling pathway (Reference Lina207). Using a cAMP-dependent pathway, VacA and GGT trigger the secretion of miR-155 and Foxp3 in the human lymphocytes (Reference Fehri283). The GGT enzymatic activity is linked to its immunomodulatory effects, whereas VacA vacuolating cytotoxicity is independent of its immunomodulatory effects, as both the non-toxigenic (s2/m2) or toxigenic (s1/m1) types of VacA are equally tolerogenic in vitro (Reference Oertli165). Selective recruitment or activation of Tregs in the preferred niche facilitates the promotion of chronicity by persistent pathogens, such as Mycobacterium tuberculosis and certain helminths (Reference McBride, Konowich and Salgame284). The same is true for H. pylori, which establishes Treg-mediated immunosuppression promoting chronic infection (Reference Arnold274, Reference Robinson275). In vitro and in vivo findings imply that H. pylori not only triggers DCs-derived tolerogenic (i.e., Treg-inducing) responses but also quenches their immunogenic functions (Reference Müller and Hartung223). Mature and immature DCs diversely affect the Th cells differentiation which results in immunity or immunosuppression (Fig. 5). DCs maturation is promoted via the cytosolic or membrane-bound PRRs-mediated recognition of PAMPs. Mature DCs can express high levels of MHC II, maturation markers, co-stimulatory markers (e.g., CD80, CD86, and CD40), Th cell-activating cytokines, and Th cell-differentiating cytokines as well as other pro-inflammatory cytokines (e.g., IL-12, IL-23, TNF-α, and IL-6) (Reference Kaebisch210). Differentiation of naive T cells into Th1 or Th17 cells depends on antigen recognition via the T-cell receptor, soluble cytokine signals, co-stimulatory signals, and high-level of IL-12 and IL-23 (Reference Müller and Hartung223). This scenario will result in immunity and H. pylori control. H. pylori-exposed DCs remain immature and present tolerogenic activity. Despite the high expression of MHC II by the semi-mature DCs, these cells fail to express co-stimulatory markers and Th1/Th17 cells differentiating cytokines, and instead, they produce high levels of the anti-inflammatory cytokine IL-10 (Reference Oertli285). These semi-mature DCs efficiently induce Treg differentiation and immunosuppression in vitro and in vivo (Reference Oertli285); however, LPS treatment can break the tolerance in vivo (Reference Oertli285). The H. pylori virulence and persistence factors including VacA and GGT act as the inhibitors of murine DCs maturation and tolerogenic reprogramming (Reference Oertli285); CagT4SS is their counterpart, which acts on human DCs (Reference Kaebisch210). Mutation of either GGT or VacA (i) reduces the colonisation of mutant strains in mice, (ii) suppresses the prevention of LPS-induced DCs maturation, and (iii) inhibits the DCs tolerisation (Reference Oertli165). Via a partial or total inhibition of activation of T cells in the lamina propria, VacA proteins can subvert the immune response (Reference Reshetnyak, Burmistrov and Maev33, Reference Raju121). All in vivo findings suggest that under H. pylori forces, DCs present tolerogenic properties leading to the anti-inflammatory cytokine secretion and Treg differentiation, as well as downregulation of T-effector cell function, and establishment of persistent infection (Reference Müller and Hartung223). Tregs facilitate the persistence of infection and protect the infected host cells against gastric inflammation. Besides, Tregs encourage bacterial colonisation which likely promotes gastric tumour progression (Reference Laur286).

Figure 4. H. pylori subverts T-cell-mediated immunity using the secreted virulence factors VacA and GGT. Following the internalization, Cytoplasmic VacA prevents nuclear translocation of NFAT by inhibiting its dephosphorylation by the Ca2+/calmodulin-dependent phosphatase calcineurin and thereby blocks IL-2 production and subsequent T-cell activation and proliferation. The GGT prevents the proliferation of T cells via interfering in the G1 phase of the cell cycle. TCR, T-cell receptor; MHCII, major histocompatibility complex class II; VacA, vacuolating cytotoxin A; GGT, γ-glutamyl transpeptidase; NFAT, nuclear factor of activated T cells; IL-2, Interleukin-2; CnA/B, calcineurin A/B subunits; CaM, calmodulin; LFA-1, lymphocyte function-associated antigen-1; PKCζ/ η, protein kinase Cζ/η.

Figure 5. The immunological response elicited by H. pylori infection. H. pylori is genetically highly variable and expresses various virulence factors, including adhesins (BabA/SabA), enzymes (urease), toxins (VacA) and effector proteins (CagA) which are involved in bacterial pathogenesis. The immunity and immunosuppression are the consequences of the maturation status of dendritic cells (DCs), which direct T helper cell differentiation. CagT4SS, Cag type IV secretion system; CagA, cytotoxin-associated gene A; VacA, vacuolating cytotoxin A; BabA, blood group antigen-binding adhesin A; SabA, sialic acid-binding adhesin A; DC, dendritic cell; Mφ, macrophage cell; PMN, polymorphonuclear leukocyte; Treg, regulatory T cell; Th1/17, T helper cell1/17; GM-CSF, granulocyte–macrophage colony-stimulating factor; TNF-α, tumour necrosis factor-α; IFN-γ, interferon-γ; TGF-β, transforming growth factor-beta, CXCL1/2, C-X-C motif chemokine ligand 1/2; IL-1/2/6/8/10/12/17/23, interleukin 1/2/6/8/10/12/17/23.

Vaccine development against H. pylori

Owing to high antibiotic resistance among H. pylori strains, World Health Organization (WHO) put this bacterium on its high-priority pathogen list to prioritise research and development on the deadly threats to global health (Reference Tacconelli287). Owing to increasing antibiotic resistance, high cost, poor patient compliance, and re-infection the current therapeutic strategies remain sub-optimal (Reference Megraud288). Besides, GC is the most drastic clinical outcome of H. pylori infection, and it is ranking fifth for incidence and fourth for mortality globally (Reference Sung289). Therefore, priority for H. pylori eradication comes with vaccination (Reference Hooi1, Reference Mohammadzadeh41, Reference Yoon290). Until now various kinds of vaccines including (i) inactivated whole-cell vaccines, (ii) subunit and recombinant protein vaccines, (iii) epitope-based vaccines, (iv) DNA vaccines, and (v) vector vaccines have been introduced for vaccination against H. pylori (Reference Soudi291-Reference Ansari, Tahmasebi-Birgani and Bijanzadeh295) (Fig. 6). Reviewing the recent progress on H. pylori vaccines shows that these efforts have not been very fruitful and most vaccine candidates are at a very early stage (phase I or even preclinical) (Reference Mohammadzadeh41, Reference Zhang296, Reference Dos Santos Viana297). H. pylori evasion from the host adaptive immunity and its persistent infection are attributed to extreme adaptation to the gastric environment and result in current vaccines’ partial or limited protectivity (Reference Camilo, Sugiyama and Touati3, Reference Robinson, Kaneko and Andersen298). There is no agreement on the role of humoral immunity in H. pylori vaccine-induced protective immunity (Reference Guo299, Reference Akhiani300). Nevertheless, Sun et al. showed that Th1 and Th17 cell responses induced by immunodominant antigens protect mice against H. pylori infection which confirms the vital role of cellular immunity for the clearance of H. pylori infection (Reference Sun301). Acquisition of H. pylori infection occurs in early childhood, so an effective vaccine should protect individuals for 10–15 years and more (Reference Sutton302). H. pylori's successful colonisation, persistence and induction of severe clinical outcomes in the human host is dependent on passing through several steps (Reference Kao, Sheu and Wu31). Designing and developing multivalent and multistage vaccines containing various immunodominant antigens involved in different aspects of H. pylori colonisation and pathogenesis could be a hopeful strategy for the eradication of this insidious bacterium (Reference Mohammadzadeh41).

Figure 6. Vaccine development against H. pylori. CagA, cytotoxin-associated gene A; VacA, vacuolating cytotoxin A; BabA, blood group antigen-binding adhesin A; HspA, heat shock protein A; FliD, Flagellar hook-associated protein 2.

Conclusions

H. pylori is an uninvited guest in the stomach mucosa of more than 50% of mankind. Variable morphology and expression of numerous heterogenic virulence factors are two intrinsic traits that make H. pylori a successful pathogen in the hostile stomach environment. Using various mechanisms, H. pylori manages to escape from the innate and adaptive immune responses. In this condition, the immune response cannot eradicate the bacterium and facilitates bacterial colonisation and survival in the gastric mucosa (Reference Abadi12, Reference Lina207, Reference Mejías-Luque, Gerhard and Tegtmeyer303). Therefore, immune evasion makes vaccination a challenging issue. In the majority of infected individuals, H. pylori and its host are in balance leading to chronic infection without any serious clinical outcomes. However, this balance is not universal, and individuals develop various clinical outcomes from asymptomatic infection to GC. The interaction between the environmental factors (e.g., diet and smoke), host genetic polymorphism (e.g., IL-1β, IL-10, and TNF-α), and the H. pylori virulence factors (e.g., CagA and VacA) will determine the clinical outcomes in individuals. So far, a wide range of virulence factors have been attributed to H. pylori classified as adhesins (BabA), enzymes (urease), toxins (VacA), and effector proteins (CagA) to influence the host-pathogen interactions. The induction of chronic persistent infection despite a strong host immune response is a unique feature of H. pylori pathogenesis. The identification of the host inflammatory pathways that H. pylori activates and subverts might help to understand the complicated pathogenesis of H. pylori. Therefore, the profound understanding of the exclusive role of H. pylori virulence factors in (i) induction of various clinical outcomes and (ii) H. pylori–host interaction to subvert the immune response will pave the way for alternative therapies and vaccine development.

Author contributions

All authors contributed to the writing and review of the manuscript.

Disclosure statement

The authors report no conflicts of interest.

Funding

This work was supported by the National Institute for Medical Research Development (NIMAD) (Grant no. 989320).

References

Hooi, JK, et al. (2017) Global prevalence of Helicobacter pylori infection: systematic review and meta-analysis. Gastroenterology 153(2), 420429.CrossRefGoogle ScholarPubMed
Moodley, Y, et al. (2012) Age of the association between Helicobacter pylori and man. PLoS Pathogen 8(5), e1002693.CrossRefGoogle ScholarPubMed
Camilo, V, Sugiyama, T and Touati, E (2017) Pathogenesis of Helicobacter pylori infection. Helicobacter 22, e12405.CrossRefGoogle ScholarPubMed
Zamani, M, et al. (2018) Systematic review with meta-analysis: the worldwide prevalence of Helicobacter pylori infection. Alimentary pharmacology & therapeutics 47(7), 868876.CrossRefGoogle ScholarPubMed
Burucoa, C and Axon, A (2017) Epidemiology of Helicobacter pylori infection. Helicobacter 22, e12403.CrossRefGoogle ScholarPubMed
Stefano, K, et al. (2018) Helicobacter pylori, transmission routes and recurrence of infection: state of the art. Acta Bio Medica: Atenei Parmensis 89(Suppl. 8), 72.Google Scholar
Kim, N (2016) Prevalence and transmission routes of Helicobacter pylori. In Kim, N (ed.), Helicobacter pylori. Springer, pp. 319.CrossRefGoogle ScholarPubMed
Yokota, SI, et al. (2015) Intrafamilial, preferentially mother-to-child and intraspousal, Helicobacter pylori infection in Japan determined by mutilocus sequence typing and random amplified polymorphic DNA fingerprinting. Helicobacter 20(5), 334342.CrossRefGoogle ScholarPubMed
Sun, Y and Zhang, J (2019) Helicobacter pylori recrudescence and its influencing factors. Journal of Cellular and Molecular Medicine 23(12), 79197925.CrossRefGoogle ScholarPubMed
Zhao, H, et al. (2021) The recurrence rate of Helicobacter pylori in recent 10 years: A systematic review and meta-analysis. Helicobacter 26(6), e12852.Google Scholar
Machlowska, J, et al. (2020) Gastric cancer: epidemiology, risk factors, classification, genomic characteristics and treatment strategies. International Journal of Molecular Sciences 21(11), 4012.CrossRefGoogle ScholarPubMed
Abadi, ATB (2017) Strategies used by Helicobacter pylori to establish persistent infection. World Journal of Gastroenterology 23(16), 2870.CrossRefGoogle ScholarPubMed
Aviles-Jimenez, F, et al. (2016) Microbiota studies in the bile duct strongly suggest a role for Helicobacter pylori in extrahepatic cholangiocarcinoma. Clinical Microbiology and Infection 22(2), 178, e11-78. e22.CrossRefGoogle ScholarPubMed
Graham, DY, Lu, H and Shiotani, A (2021) Vonoprazan-containing Helicobacter pylori triple therapies contribution to global antimicrobial resistance. Journal of Gastroenterology and Hepatology 36(5), 11591163.CrossRefGoogle ScholarPubMed
Hatakeyama, M (2017) Structure and function of Helicobacter pylori CagA, the first-identified bacterial protein involved in human cancer. Proceedings of the Japan Academy. Series B 93(4), 196219.CrossRefGoogle ScholarPubMed
Marshall, BJ and Windsor, HM (2005) The relation of Helicobacter pylori to gastric adenocarcinoma and lymphoma: pathophysiology, epidemiology, screening, clinical presentation, treatment, and prevention. Medical Clinics of North America 89(2), 313344.CrossRefGoogle ScholarPubMed
Ota, H, et al. (2009) Crucial roles of Helicobacter pylori infection in the pathogenesis of gastric cancer and gastric mucosa-associated lymphoid tissue (MALT) lymphoma. Rinsho byori. The Japanese Journal of Clinical Pathology 57(9), 861.Google Scholar
Franceschi, F, Covino, M and Roubaud Baudron, C (2019) Helicobacter pylori and extragastric diseases. Helicobacter 24, e12636.CrossRefGoogle ScholarPubMed
Chang, W-L, Yeh, Y-C and Sheu, B-S (2018) The impacts of Helicobacter pylori virulence factors on the development of gastroduodenal diseases. Journal of Biomedical Science 25(1), 19.CrossRefGoogle ScholarPubMed
Ofori, EG, et al. (2019) Helicobacter pylori infection, virulence genes’ distribution and accompanying clinical outcomes: The West Africa situation. BioMed Research International 2019, 7312908.CrossRefGoogle ScholarPubMed
Toh, JW and Wilson, RB (2020) Pathways of gastric carcinogenesis, Helicobacter pylori virulence and interactions with antioxidant systems, vitamin C and phytochemicals. International Journal of Molecular Sciences 21(17), 6451.CrossRefGoogle ScholarPubMed
Wang, F, et al. (2014) Helicobacter pylori-induced gastric inflammation and gastric cancer. Cancer Letters 345(2), 196202.CrossRefGoogle ScholarPubMed
Assaad, S, et al. (2018) Dietary habits and Helicobacter pylori infection: a cross sectional study at a Lebanese hospital. BMC Gastroenterology 18, 113.CrossRefGoogle Scholar
Baradaran, A, et al. (2021) The association between Helicobacter pylori and obesity: a systematic review and meta-analysis of case–control studies. Clinical Diabetes and Endocrinology 7(1), 111.CrossRefGoogle ScholarPubMed
De Martel, C and Parsonnet, J (2006) Helicobacter pylori infection and gender: a meta-analysis of population-based prevalence surveys. Digestive Diseases and Sciences 51, 22922301.CrossRefGoogle Scholar
Dore, MP, et al. (1999) High prevalence of Helicobacter pylori infection in shepherds. Digestive Diseases and Sciences 44, 11611164.CrossRefGoogle ScholarPubMed
Mastromarino, P, et al. (2005) Does hospital work constitute a risk factor for Helicobacter pylori infection? Journal of Hospital Infection 60(3), 261268.CrossRefGoogle ScholarPubMed
Shatila, M and Thomas, AS (2022) Current and future perspectives in the diagnosis and management of Helicobacter pylori infection. Journal of Clinical Medicine 11(17), 5086.CrossRefGoogle ScholarPubMed
Wang, D, et al. (2019) Alterations in the human gut microbiome associated with Helicobacter pylori infection. FEBS open bio 9(9), 15521560.CrossRefGoogle ScholarPubMed
Baj, J, et al. (2021) Helicobacter pylori virulence factors—mechanisms of bacterial pathogenicity in the gastric microenvironment. Cells 10(1), 27.CrossRefGoogle Scholar
Kao, C-Y, Sheu, B-S and Wu, J-J (2016) Helicobacter pylori infection: An overview of bacterial virulence factors and pathogenesis. Biomedical Journal 39(1), 1423.CrossRefGoogle ScholarPubMed
Knorr, J, et al. (2019) Classification of Helicobacter pylori virulence factors: Is CagA a toxin or not? Trends in Microbiology 27(9), 731738.CrossRefGoogle ScholarPubMed
Reshetnyak, VI, Burmistrov, AI and Maev, IV (2021) Helicobacter pylori: Commensal, symbiont or pathogen? World Journal of Gastroenterology 27(7), 545.CrossRefGoogle ScholarPubMed
Reshetnyak, VI and Reshetnyak, TM (2017) Significance of dormant forms of Helicobacter pylori in ulcerogenesis. World Journal of Gastroenterology 23(27), 4867.CrossRefGoogle ScholarPubMed
Săsăran, MO, Meliț, LE and Dobru, ED (2021) MicroRNA modulation of host immune response and inflammation triggered by Helicobacter pylori. International Journal of Molecular Sciences 22(3), 1406.CrossRefGoogle ScholarPubMed
Meliț, LE, et al. (2021) Innate immunity–the hallmark of Helicobacter pylori infection in pediatric chronic gastritis. World Journal of Clinical Cases 9(23), 6686.CrossRefGoogle ScholarPubMed
Guclu, M and Agan, AF (2017) Association of severity of Helicobacter pylori infection with peripheral blood neutrophil to lymphocyte ratio and mean platelet volume. Euroasian Journal of Hepato-Gastroenterology 7(1), 11.CrossRefGoogle ScholarPubMed
Mărginean, CD, Mărginean, CO and Meliț, LE (2022) Helicobacter pylori-related extraintestinal manifestations—myth or reality. Children 9(9), 1352.CrossRefGoogle ScholarPubMed
Peek, RM Jr, Fiske, C and Wilson, KT (2010) Role of innate immunity in Helicobacter pylori-induced gastric malignancy. Physiological Reviews 90(3), 831858.CrossRefGoogle ScholarPubMed
Wen, S and Moss, SF (2009) Helicobacter pylori virulence factors in gastric carcinogenesis. Cancer Letters 282(1), 18.CrossRefGoogle ScholarPubMed
Mohammadzadeh, R, et al. (2022) Designing and development of epitope-based vaccines against Helicobacter pylori. Critical Reviews in Microbiology 48(4), 489512.CrossRefGoogle ScholarPubMed
Alm, RA, et al. (1999) Genomic-sequence comparison of two unrelated isolates of the human gastric pathogen Helicobacter pylori. Nature 397(6715), 176180.CrossRefGoogle ScholarPubMed
Tomb JF, WO, Kerlavage, AR, Clayton, RA, Sutton, GG, Fleischmann, RD, et al. (1997) The complete genome sequence of the gastric pathogen Helicobacter pylori. Nature 389, 539547.CrossRefGoogle Scholar
Cao, D-M, et al. (2016) Comparative genomics of Helicobacter pylori and non-pylori Helicobacter species to identify new regions associated with its pathogenicity and adaptability. BioMed Research International, 6106029.Google ScholarPubMed
Noto, JM, et al. (2018) Pan-genomic analyses identify key Helicobacter pylori pathogenic loci modified by carcinogenic host microenvironments. Gut 67(10), 17931804.CrossRefGoogle ScholarPubMed
Kabamba, ET and Yamaoka, Y (2019) Helicobacter pylori and related virulence factors for gastrointestinal diseases. In Shiotani, A (ed.), Gastric Cancer. Springer, pp. 3150.CrossRefGoogle Scholar
Olbermann, P, et al. (2010) A global overview of the genetic and functional diversity in the Helicobacter pylori cag pathogenicity island. PLoS Genetics 6(8), e1001069.CrossRefGoogle ScholarPubMed
Yamaoka, Y (2010) Mechanisms of disease: Helicobacter pylori virulence factors. Nature Reviews Gastroenterology & hepatology 7(11), 629.CrossRefGoogle ScholarPubMed
Censini, S, et al. (1996) cag, a pathogenicity island of Helicobacter pylori, encodes type I-specific and disease-associated virulence factors. Proceedings of the National Academy of Sciences 93(25), 1464814653.CrossRefGoogle ScholarPubMed
Blomstergren, A, et al. (2004) Comparative analysis of the complete cag pathogenicity island sequence in four Helicobacter pylori isolates. Gene 328, 8593.CrossRefGoogle ScholarPubMed
Backert, S, Tegtmeyer, N and Fischer, W (2015) Composition, structure and function of the Helicobacter pylori cag pathogenicity island encoded type IV secretion system. Future microbiology 10(6), 955965.CrossRefGoogle ScholarPubMed
Chung, JM, et al. (2020) Structure of the Helicobacter pylori Cag Type IV Secretion System. Biophysical Journal 118(3), 295a.CrossRefGoogle Scholar
Backert, S, et al. (2000) Translocation of the Helicobacter pylori CagA protein in gastric epithelial cells by a type IV secretion apparatus. Cellular microbiology 2(2), 155164.CrossRefGoogle ScholarPubMed
Varga, MG, et al. (2016) Pathogenic Helicobacter pylori strains translocate DNA and activate TLR9 via the cancer-associated cag type IV secretion system. Oncogene 35(48), 62626269.CrossRefGoogle ScholarPubMed
Viala, J, et al. (2004) Nod1 responds to peptidoglycan delivered by the Helicobacter pylori cag pathogenicity island. Nature Immunology 5(11), 11661174.CrossRefGoogle ScholarPubMed
Akopyants, NS, et al. (1998) Analyses of the cag pathogenicity island of Helicobacter pylori. Molecular Microbiology 28(1), 3753.CrossRefGoogle ScholarPubMed
Patra, R, et al. (2011) Intact cag pathogenicity island of Helicobacter pylori without disease association in Kolkata, India. International Journal of Medical Microbiology 301(4), 293302.CrossRefGoogle ScholarPubMed
Nilsson, C, et al. (2003) Correlation between cag pathogenicity island composition and Helicobacter pylori-associated gastroduodenal disease. Infection and immunity 71(11), 65736581.CrossRefGoogle ScholarPubMed
Canzian, F, et al. (2020) Genetic polymorphisms in the cag pathogenicity island of Helicobacter pylori and risk of stomach cancer and high-grade premalignant gastric lesions. International Journal of Cancer 147(9), 24372445.CrossRefGoogle ScholarPubMed
Sharndama, HC and Mba, IE (2022) Helicobacter pylori: An up-to-date overview on the virulence and pathogenesis mechanisms. Brazilian Journal of Microbiology, 118.Google ScholarPubMed
Wang, D, et al. (2017) The association between vacA or cagA status and eradication outcome of Helicobacter pylori infection: A meta-analysis. PloS one 12(5), e0177455.CrossRefGoogle ScholarPubMed
Chowdhury, R (2014) Host cell contact induces fur-dependent expression of virulence factors CagA and VacA in Helicobacter pylori. Helicobacter 19(1), 1725.Google Scholar
Noto, JM, et al. (2012) Iron deficiency accelerates Helicobacter pylori–induced carcinogenesis in rodents and humans. The Journal of Clinical Investigation 123(1), 479492.CrossRefGoogle ScholarPubMed
Merrell, DS, et al. (2003) pH-regulated gene expression of the gastric pathogen Helicobacter pylori. Infection and immunity 71(6), 35293539.CrossRefGoogle ScholarPubMed
Loh, JT, Torres, VJ and Cover, TL (2007) Regulation of Helicobacter pylori cagA expression in response to salt. Cancer Research 67(10), 47094715.CrossRefGoogle ScholarPubMed
Hayashi, T, et al. (2012) Tertiary structure-function analysis reveals the pathogenic signaling potentiation mechanism of Helicobacter pylori oncogenic effector CagA. Cell Host & Microbe 12(1), 2033.CrossRefGoogle ScholarPubMed
Hatakeyama, M (2004) Oncogenic mechanisms of the Helicobacter pylori CagA protein. Nature Reviews Cancer 4(9), 688694.CrossRefGoogle ScholarPubMed
Hatakeyama, M (2011) Anthropological and clinical implications for the structural diversity of the Helicobacter pylori CagA oncoprotein. Cancer Science 102(1), 3643.CrossRefGoogle ScholarPubMed
Backert, S and Selbach, M (2005) Tyrosine-phosphorylated bacterial effector proteins: the enemies within. Trends in Microbiology 13(10), 476484.CrossRefGoogle Scholar
Nagase, L, et al. (2015) Dramatic increase in SHP2 binding activity of Helicobacter pylori Western CagA by EPIYA-C duplication: its implications in gastric carcinogenesis. Scientific Reports 5(1), 15749.CrossRefGoogle ScholarPubMed
Yamazaki, S, et al. (2003) The CagA protein of Helicobacter pylori is translocated into epithelial cells and binds to SHP-2 in human gastric mucosa. The Journal of Infectious Diseases 187(2), 334337.CrossRefGoogle ScholarPubMed
Tsutsumi, R, et al. (2003) Attenuation of Helicobacter pylori CagA⋅ SHP-2 signaling by interaction between CagA and C-terminal Src kinase. Journal of Biological Chemistry 278(6), 36643670.CrossRefGoogle ScholarPubMed
Suzuki, M, et al. (2005) Interaction of CagA with Crk plays an important role in Helicobacter pylori–induced loss of gastric epithelial cell adhesion. The Journal of Experimental Medicine 202(9), 12351247.CrossRefGoogle ScholarPubMed
Higashi, H, et al. (2002) SHP-2 tyrosine phosphatase as an intracellular target of Helicobacter pylori CagA protein. Science 295(5555), 683686.CrossRefGoogle ScholarPubMed
Gorrell, RJ, et al. (2013) A novel NOD1-and CagA-independent pathway of interleukin-8 induction mediated by the Helicobacter pylori type IV secretion system. Cellular Microbiology 15(4), 554570.CrossRefGoogle Scholar
Hayashi, T, et al. (2017) Differential mechanisms for SHP2 binding and activation are exploited by geographically distinct Helicobacter pylori CagA oncoproteins. Cell Reports 20(12), 28762890.CrossRefGoogle ScholarPubMed
Hayashi, Y, et al. (2013) CagA mediates epigenetic regulation to attenuate let-7 expression in Helicobacter pylori-related carcinogenesis. Gut 62(11), 15361546.CrossRefGoogle ScholarPubMed
Sepulveda, AR, et al. (2010) CpG methylation and reduced expression of O6-methylguanine DNA methyltransferase is associated with Helicobacter pylori infection. Gastroenterology 138(5), 18361844, e4.CrossRefGoogle ScholarPubMed
Tsang, Y, et al. (2010) Helicobacter pylori CagA targets gastric tumor suppressor RUNX3 for proteasome-mediated degradation. Oncogene 29(41), 56435650.CrossRefGoogle ScholarPubMed
Buti, L, et al. (2011) Helicobacter pylori cytotoxin-associated gene A (CagA) subverts the apoptosis-stimulating protein of p53 (ASPP2) tumor suppressor pathway of the host. Proceedings of the National Academy of Sciences 108(22), 92389243.CrossRefGoogle ScholarPubMed
Ren, S, et al. (2006) Structural basis and functional consequence of Helicobacter pylori CagA multimerization in cells. Journal of Biological Chemistry 281(43), 3234432352.CrossRefGoogle ScholarPubMed
Nishikawa, H, et al. (2016) Impact of structural polymorphism for the Helicobacter pylori CagA oncoprotein on binding to polarity-regulating kinase PAR1b. Scientific Reports 6(1), 113.CrossRefGoogle ScholarPubMed
Javaheri, A, et al. (2016) Helicobacter pylori adhesin HopQ engages in a virulence-enhancing interaction with human CEACAMs. Nature Microbiology 2(1), 113.Google Scholar
Königer, V, et al. (2016) Helicobacter pylori exploits human CEACAMs via HopQ for adherence and translocation of CagA. Nature Microbiology 2(1), 112.Google ScholarPubMed
Behrens, I-K, et al. (2020) The HopQ-CEACAM interaction controls CagA translocation, phosphorylation, and phagocytosis of Helicobacter pylori in neutrophils. MBio 11(1), e03256–19.CrossRefGoogle ScholarPubMed
Zhao, Q, et al. (2018) Integrin but not CEACAM receptors are dispensable for Helicobacter pylori CagA translocation. PLoS Pathogens 14(10), e1007359.CrossRefGoogle ScholarPubMed
Foegeding, NJ, et al. (2016) An overview of Helicobacter pylori VacA toxin biology. Toxins 8(6), 173.CrossRefGoogle ScholarPubMed
Atherton, JC, et al. (1995) Mosaicism in Vacuolating Cytotoxin Alleles of Helicobacter pylori: ASSOCIATION OF SPECIFIC vacA TYPES WITH CYTOTOXIN PRODUCTION AND PEPTIC ULCERATION (∗). Journal of Biological Chemistry 270(30), 1777117777.CrossRefGoogle ScholarPubMed
Atherton, JC, et al. (1999) Vacuolating cytotoxin (vacA) alleles of Helicobacter pylori comprise two geographically widespread types, m1 and m2, and have evolved through limited recombination. Current Microbiology 39, 211218.CrossRefGoogle ScholarPubMed
Chung, C, et al. (2010) Diversity of VacA intermediate region among Helicobacter pylori strains from several regions of the world. Journal of Clinical Microbiology 48(3), 690696.CrossRefGoogle ScholarPubMed
Pagliaccia, C, et al. (1998) The m2 form of the Helicobacter pylori cytotoxin has cell type-specific vacuolating activity. Proceedings of the National Academy of Sciences 95(17), 1021210217.CrossRefGoogle ScholarPubMed
van Doorn, L-J, et al. (1998) Expanding allelic diversity of Helicobacter pylori vacA. Journal of Clinical Microbiology 36(9), 25972603.CrossRefGoogle ScholarPubMed
Rhead, JL, et al. (2007) A new Helicobacter pylori vacuolating cytotoxin determinant, the intermediate region, is associated with gastric cancer. Gastroenterology 133(3), 926936.CrossRefGoogle ScholarPubMed
Letley, DP, et al. (2003) Determinants of non-toxicity in the gastric pathogen Helicobacter pylori. Journal of Biological Chemistry 278(29), 2673426741.CrossRefGoogle ScholarPubMed
Forsyth, M, et al. (1998) Heterogeneity in levels of vacuolating cytotoxin gene (vacA) transcription among Helicobacter pylori strains. Infection and immunity 66(7), 30883094.CrossRefGoogle ScholarPubMed
Atherton, J, et al. (1997) Clinical and pathological importance of heterogeneity in vacA, the vacuolating cytotoxin gene of Helicobacter pylori. Gastroenterology 112(1), 9299.CrossRefGoogle ScholarPubMed
Ferreira, RM, et al. (2012) A novel method for genotyping the Helicobacter pylori vacA intermediate region directly in gastric biopsy specimens. Journal of Clinical Microbiology 50(12), 39833989.CrossRefGoogle ScholarPubMed
Ferreira, RM, Machado, JC and Figueiredo, C (2014) Clinical relevance of Helicobacter pylori vacA and cagA genotypes in gastric carcinoma. Best Practice & Research Clinical Gastroenterology 28(6), 10031015.CrossRefGoogle ScholarPubMed
Ogiwara, H, et al. (2009) Role of deletion located between the intermediate and middle regions of the Helicobacter pylori vacA gene in cases of gastroduodenal diseases. Journal of Clinical Microbiology 47(11), 34933500.CrossRefGoogle ScholarPubMed
Bakhti, SZ, et al. (2016) Relevance of Helicobacter pylori vacA 3ʹ-end region polymorphism to gastric cancer. Helicobacter 21(4), 305316.CrossRefGoogle ScholarPubMed
Amilon, KR, et al. (2015) Expression of the Helicobacter pylori virulence factor vacuolating cytotoxin A (vac A) is influenced by a potential stem-loop structure in the 5′ untranslated region of the transcript. Molecular Microbiology 98(5), 831846.CrossRefGoogle Scholar
Gancz, H, Jones, KR and Merrell, DS (2008) Sodium chloride affects Helicobacter pylori growth and gene expression. Journal of Bacteriology 190(11), 41004105.CrossRefGoogle ScholarPubMed
Merrell, DS, et al. (2003) Growth phase-dependent response of Helicobacter pylori to iron starvation. Infection and Immunity 71(11), 65106525.CrossRefGoogle ScholarPubMed
van Amsterdam, K, et al. (2003) Induced Helicobacter pylori vacuolating cytotoxin VacA expression after initial colonisation of human gastric epithelial cells. FEMS Immunology & Medical Microbiology 39(3), 251256.CrossRefGoogle ScholarPubMed
Torres, VJ, McClain, MS and Cover, TL (2004) Interactions between p-33 and p-55 domains of the Helicobacter pylori vacuolating cytotoxin (VacA). Journal of Biological Chemistry 279(3), 23242331.CrossRefGoogle ScholarPubMed
Yahiro, K, et al. (2015) Helicobacter pylori VacA toxin causes cell death by inducing accumulation of cytoplasmic connexin 43. Nature Publishing Group.CrossRefGoogle Scholar
Chauhan, N, et al. (2019) Helicobacter pylori VacA, a distinct toxin exerts diverse functionalities in numerous cells: An overview. Helicobacter 24(1), e12544.CrossRefGoogle Scholar
Fischer, W, et al. (2001) Outer membrane targeting of passenger proteins by the vacuolating cytotoxin autotransporter of Helicobacter pylori. Infection and immunity 69(11), 67696775.CrossRefGoogle ScholarPubMed
McClain, MS, Beckett, AC and Cover, TL (2017) Helicobacter pylori vacuolating toxin and gastric cancer. Toxins 9(10), 316.CrossRefGoogle ScholarPubMed
Yahiro, K, et al. (2016) New insights into VacA intoxication mediated through its cell surface receptors. Toxins 8(5), 152.CrossRefGoogle ScholarPubMed
Terebiznik, MR, et al. (2009) Effect of Helicobacter pylori's vacuolating cytotoxin on the autophagy pathway in gastric epithelial cells. Autophagy 5(3), 370379.CrossRefGoogle ScholarPubMed
Sundrud, MS, et al. (2004) Inhibition of primary human T cell proliferation by Helicobacter pylori vacuolating toxin (VacA) is independent of VacA effects on IL-2 secretion. Proceedings of the National Academy of Sciences 101(20), 77277732.CrossRefGoogle Scholar
Radin, JN, et al. (2011) Helicobacter pylori VacA induces programmed necrosis in gastric epithelial cells. Infection and Immunity 79(7), 25352543.CrossRefGoogle ScholarPubMed
McClain, MS, et al. (2003) Essential role of a GXXXG motif for membrane channel formation by Helicobacter pylori vacuolating toxin. Journal of Biological Chemistry 278(14), 1210112108.CrossRefGoogle ScholarPubMed
Hisatsune, J, Masaaki Nakayama, EY, Shirasaka, D, Kurazono, H, Katagata, Y, Inoue, H, Han, J, Sap, J, Yahiro, K, Moss, J and Hirayama, T (2007) Helicobacter pylori VacA enhances prostaglandin E2 production through induction of cyclooxygenase 2 expression via a p38 mitogen-activated protein kinase/activating transcription factor 2 cascade in AZ-521 cells. Infection and Immunity 75(9).CrossRefGoogle Scholar
Jain, P, Luo, Z-Q and Blanke, SR (2011) Helicobacter pylori vacuolating cytotoxin A (VacA) engages the mitochondrial fission machinery to induce host cell death. Proceedings of the National Academy of Sciences 108(38), 1603216037.CrossRefGoogle ScholarPubMed
Domańska, G, et al. (2010) Helicobacter pylori VacA toxin/subunit p34: targeting of an anion channel to the inner mitochondrial membrane. PLoS Pathogens 6(4), e1000878.CrossRefGoogle ScholarPubMed
McClain, MS, et al. (2001) A 12-amino-acid segment, present in type s2 but not type s1 Helicobacter pylori VacA proteins, abolishes cytotoxin activity and alters membrane channel formation. Journal of Bacteriology 183(22), 64996508.CrossRefGoogle Scholar
Torres, VJ, et al. (2007) Helicobacter pylori vacuolating cytotoxin inhibits activation-induced proliferation of human T and B lymphocyte subsets. The Journal of Immunology 179(8), 54335440.CrossRefGoogle Scholar
Greenfield, LK and Jones, NL (2013) Modulation of autophagy by Helicobacter pylori and its role in gastric carcinogenesis. Trends in Microbiology 21(11), 602612.CrossRefGoogle ScholarPubMed
Raju, D, et al. (2012) Vacuolating cytotoxin and variants in Atg16L1 that disrupt autophagy promote Helicobacter pylori infection in humans. Gastroenterology 142(5), 11601171.CrossRefGoogle ScholarPubMed
Ricci, V (2016) Relationship between VacA toxin and host cell autophagy in Helicobacter pylori infection of the human stomach: a few answers, many questions. Toxins 8(7), 203.CrossRefGoogle ScholarPubMed
Tegtmeyer, N, et al. (2009) Importance of EGF receptor, HER2/Neu and Erk1/2 kinase signalling for host cell elongation and scattering induced by the Helicobacter pylori CagA protein: antagonistic effects of the vacuolating cytotoxin VacA. Cellular Microbiology 11(3), 488505.CrossRefGoogle ScholarPubMed
Tsugawa, H, et al. (2012) Reactive oxygen species-induced autophagic degradation of Helicobacter pylori CagA is specifically suppressed in cancer stem-like cells. Cell Host & Microbe 12(6), 764777.CrossRefGoogle ScholarPubMed
Abdullah, M, et al. (2019) VacA promotes CagA accumulation in gastric epithelial cells during Helicobacter pylori infection. Scientific Reports 9(1), 19.CrossRefGoogle ScholarPubMed
Li, N, et al. (2017) Helicobacter pylori CagA protein negatively regulates autophagy and promotes inflammatory response via c-Met-PI3 K/Akt-mTOR signaling pathway. Frontiers in Cellular and Infection Microbiology 7, 417.CrossRefGoogle Scholar
Akada, JK, et al. (2010) Helicobacter pylori CagA inhibits endocytosis of cytotoxin VacA in host cells. Disease Models & Mechanisms 3(9-10), 605617.CrossRefGoogle ScholarPubMed
Yokoyama, K, et al. (2005) Functional antagonism between Helicobacter pylori CagA and vacuolating toxin VacA in control of the NFAT signaling pathway in gastric epithelial cells. Proceedings of the National Academy of Sciences 102(27), 96619666.CrossRefGoogle ScholarPubMed
Tan, S, et al. (2011) Helicobacter pylori perturbs iron trafficking in the epithelium to grow on the cell surface. PLoS Pathogens 7(5), e1002050.CrossRefGoogle ScholarPubMed
Ansari, S and Yamaoka, Y (2020) Helicobacter pylori virulence factor cytotoxin-associated Gene A (CagA)-mediated gastric pathogenicity. International Journal of Molecular Sciences 21(19), 7430.CrossRefGoogle ScholarPubMed
Voland, P, et al. (2006) Human immune response towards recombinant Helicobacter pylori urease and cellular fractions. Vaccine 24(18), 38323839.CrossRefGoogle ScholarPubMed
Akada, JK, et al. (2000) Identification of the urease operon in Helicobacter pylori and its control by mRNA decay in response to pH. Molecular Microbiology 36(5), 10711084.CrossRefGoogle ScholarPubMed
Marcus, EA, Sachs, G and Scott, DR (2018) Acid-regulated gene expression of Helicobacter pylori: insight into acid protection and gastric colonization. Helicobacter 23(3), e12490.CrossRefGoogle ScholarPubMed
Weeks, DL, et al. (2000) A H+-gated urea channel: the link between Helicobacter pylori urease and gastric colonization. Science 287(5452), 482485.CrossRefGoogle ScholarPubMed
Belzer, C, et al. (2005) Differential regulation of urease activity in Helicobacter hepaticus and Helicobacter pylori. Microbiology 151(12), 39893995.CrossRefGoogle ScholarPubMed
Allen, MG, et al. (2023) Regulation of Helicobacter pylori urease and acetone carboxylase genes by nitric oxide and the CrdRS two-component system. Microbiology Spectrum, e04633–22.Google ScholarPubMed
Strugatsky, D, et al. (2013) Structure of the proton-gated urea channel from the gastric pathogen Helicobacter pylori. Nature 493(7431), 255258.CrossRefGoogle ScholarPubMed
Ha, N-C, et al. (2001) Supramolecular assembly and acid resistance of Helicobacter pylori urease. Nature Structural Biology 8(6), 505509.CrossRefGoogle ScholarPubMed
Dunn, BE and Phadnis, SH (1998) Structure, function and localization of Helicobacter pylori urease. The Yale Journal of Biology and Medicine 71(2), 63.Google ScholarPubMed
Lytton, SD, et al. (2005) Production of ammonium by Helicobacter pylori mediates occludin processing and disruption of tight junctions in Caco-2 cells. Microbiology 151(10), 32673276.CrossRefGoogle ScholarPubMed
Schoep, TD, et al. (2010) Surface properties of Helicobacter pylori urease complex are essential for persistence. PloS one 5(11), e15042.CrossRefGoogle ScholarPubMed
Celli, JP, et al. (2009) Helicobacter pylori moves through mucus by reducing mucin viscoelasticity. Proceedings of the National Academy of Sciences 106(34), 1432114326.CrossRefGoogle ScholarPubMed
Scott, DR, et al. (1998) The role of internal urease in acid resistance of Helicobacter pylori. Gastroenterology 114(1), 5870.CrossRefGoogle ScholarPubMed
Ghalehnoei, H, et al. (2016) Relationship between ureB sequence diversity, urease activity and genotypic variations of different Helicobacter pylori strains in patients with gastric disorders. Polish journal of microbiology 65(2).CrossRefGoogle ScholarPubMed
Oleastro, M and Ménard, A (2013) The role of Helicobacter pylori outer membrane proteins in adherence and pathogenesis. Biology 2(3), 11101134.CrossRefGoogle ScholarPubMed
Xu, C, et al. (2020) Virulence of Helicobacter pylori outer membrane proteins: an updated review. European Journal of Clinical Microbiology & Infectious Diseases 39, 18211830.CrossRefGoogle ScholarPubMed
Doohan, D, et al. (2021) Helicobacter pylori BabA–SabA key roles in the adherence phase: The synergic mechanism for successful colonization and disease development. Toxins 13(7), 485.CrossRefGoogle ScholarPubMed
Ansari, S and Yamaoka, Y (2017) Helicobacter pylori BabA in adaptation for gastric colonization. World Journal of Gastroenterology 23(23), 4158.CrossRefGoogle ScholarPubMed
Moonens, K, et al. (2016) Structural insights into polymorphic ABO glycan binding by Helicobacter pylori. Cell Host & Microbe 19(1), 5566.CrossRefGoogle ScholarPubMed
Hage, N, et al. (2015) Structural basis of Lewisb antigen binding by the Helicobacter pylori adhesin BabA. Science Advances 1(7), e1500315.CrossRefGoogle ScholarPubMed
Rad, R, et al. (2002) The Helicobacter pylori blood group antigen-binding adhesin facilitates bacterial colonization and augments a nonspecific immune response. The Journal of Immunology 168(6), 30333041.CrossRefGoogle ScholarPubMed
Ishijima, N, et al. (2011) BabA-mediated adherence is a potentiator of the Helicobacter pylori type IV secretion system activity. Journal of Biological Chemistry 286(28), 2525625264.CrossRefGoogle ScholarPubMed
Saberi, S, et al. (2016) Helicobacter pylori strains from duodenal ulcer patients exhibit mixed babA/B genotypes with low levels of BabA adhesin and Lewis b binding. Digestive Diseases and Sciences 61(10), 28682877.CrossRefGoogle ScholarPubMed
Chen, M-Y, et al. (2013) Association of Helicobacter pylori babA2 with peptic ulcer disease and gastric cancer. World Journal of Gastroenterology 19(26), 4242.CrossRefGoogle ScholarPubMed
Mahdavi, J, et al. (2002) Helicobacter pylori SabA adhesin in persistent infection and chronic inflammation. Science 297(5581), 573578.CrossRefGoogle ScholarPubMed
Pang, SS, et al. (2014) The three-dimensional structure of the extracellular adhesion domain of the sialic acid-binding adhesin SabA from Helicobacter pylori. Journal of Biological Chemistry 289(10), 63326340.CrossRefGoogle ScholarPubMed
Sheu, B-S, et al. (2006) Interaction between host gastric Sialyl-Lewis X and Helicobacter pylori SabA enhances H. pylori density in patients lacking gastric Lewis B antigen. Official journal of the American College of Gastroenterology 101(1), 3644.CrossRefGoogle ScholarPubMed
Yanai, A, et al. (2007) Clinical relevance of Helicobacter pylori sabA genotype in Japanese clinical isolates. Journal of gastroenterology and hepatology 22(12), 22282232.CrossRefGoogle ScholarPubMed
Sáenz, JB, Vargas, N and Mills, JC (2019) Tropism for spasmolytic polypeptide-expressing metaplasia allows Helicobacter pylori to expand its intragastric niche. Gastroenterology 156(1), 160174, e7.CrossRefGoogle ScholarPubMed
Keikha, M, et al. (2019) Potential antigen candidates for subunit vaccine development against Helicobacter pylori infection. Journal of Cellular Physiology 234(12), 2146021470.CrossRefGoogle ScholarPubMed
Naz, A, et al. (2015) Identification of putative vaccine candidates against Helicobacter pylori exploiting exoproteome and secretome: a reverse vaccinology based approach. Infection, Genetics and Evolution 32, 280291.CrossRefGoogle ScholarPubMed
Urrutia-Baca, VH, et al. (2019) Immunoinformatics approach to design a novel epitope-based oral vaccine against Helicobacter pylori. Journal of Computational Biology 26(10), 11771190.CrossRefGoogle ScholarPubMed
Ricci, V, et al. (2014) Helicobacter pylori gamma-glutamyl transpeptidase and its pathogenic role. World Journal of Gastroenterology 20(3), 630.CrossRefGoogle ScholarPubMed
McGovern, K, et al. (2001) γ-Glutamyltransferase is a Helicobacter pylori virulence factor but is not essential for colonization. Infection and Immunity 69(6), 41684173.CrossRefGoogle Scholar
Oertli, M, et al. (2013) Helicobacter pylori γ-glutamyl transpeptidase and vacuolating cytotoxin promote gastric persistence and immune tolerance. Proceedings of the National Academy of Sciences 110(8), 30473052.CrossRefGoogle ScholarPubMed
Gong, M, et al. (2010) Helicobacter pylori γ-glutamyl transpeptidase is a pathogenic factor in the development of peptic ulcer disease. Gastroenterology 139(2), 564573.CrossRefGoogle ScholarPubMed
Flahou, B, et al. (2011) Gastric epithelial cell death caused by Helicobacter suis and Helicobacter pylori γ-glutamyl transpeptidase is mainly glutathione degradation-dependent. Cellular Microbiology 13(12), 19331955.CrossRefGoogle ScholarPubMed
Wessler, S (2016) Emerging novel virulence factors of Helicobacter pylori. In Helicobacter pylori Research. Springer, pp. 165188.CrossRefGoogle Scholar
Schmees, C, et al. (2007) Inhibition of T-cell proliferation by Helicobacter pylori γ-glutamyl transpeptidase. Gastroenterology 132(5), 18201833.CrossRefGoogle ScholarPubMed
Kim, K-M, et al. (2010) Helicobacter pylori γ-glutamyltranspeptidase induces cell cycle arrest at the G1-S phase transition. The Journal of Microbiology 48(3), 372377.CrossRefGoogle ScholarPubMed
Engler, DB, et al. (2014) Effective treatment of allergic airway inflammation with Helicobacter pylori immunomodulators requires BATF3-dependent dendritic cells and IL-10. Proceedings of the National Academy of Sciences 111(32), 1181011815.CrossRefGoogle ScholarPubMed
Suganuma, M, et al. (2008) TNF-α-inducing protein, a carcinogenic factor secreted from Helicobacter pylori, enters gastric cancer cells. International Journal of Cancer 123(1), 117122.CrossRefGoogle ScholarPubMed
Kuzuhara, T, et al. (2005) Presence of a motif conserved between Helicobacter pylori TNF-α inducing protein (Tipα) and penicillin-binding proteins. Biological and Pharmaceutical Bulletin 28(11), 21332137.CrossRefGoogle ScholarPubMed
Suganuma, M, et al. (2006) Carcinogenic role of tumor necrosis factor-alpha inducing protein of Helicobacter pylori in human stomach. Journal of Biochemistry and Molecular Biology 39(1), 1.Google ScholarPubMed
Gao, M, et al. (2012) Crystal structure of TNF-α-inducing protein from Helicobacter pylori in active form reveals the intrinsic molecular flexibility for unique DNA-binding. PLoS ONE 7(7), e41871.CrossRefGoogle ScholarPubMed
Jang, JY, et al. (2009) Crystal structure of the TNF-α-inducing protein (Tipα) from Helicobacter pylori: insights into Its DNA-binding activity. Journal of Molecular Biology 392(1), 191197.CrossRefGoogle ScholarPubMed
Backert, S, Tegtmeyer, N and Selbach, M (2010) The versatility of Helicobacter pylori CagA effector protein functions: The master key hypothesis. Helicobacter 15(3), 163176.CrossRefGoogle ScholarPubMed
Kuzuhara, T, et al. (2007a) Helicobacter pylori-secreting protein Tipα is a potent inducer of chemokine gene expressions in stomach cancer cells. Journal of Cancer Research and Clinical Oncology 133(5), 287296.CrossRefGoogle ScholarPubMed
Suganuma, M, et al. (2005) New tumor necrosis factor-α-inducing protein released from Helicobacter pylori for gastric cancer progression. Journal of Cancer Research and Clinical Oncology 131(5), 305313.CrossRefGoogle ScholarPubMed
Bauer, J, et al. (2012) Lymphotoxin, NF-ĸB, and cancer: the dark side of cytokines. Digestive Diseases 30(5), 453468.CrossRefGoogle ScholarPubMed
Morningstar-Wright, L, et al. (2022) The TNF-alpha inducing protein is associated with gastric inflammation and hyperplasia in a murine model of Helicobacter pylori infection. Frontiers in Pharmacology 13, 241.CrossRefGoogle Scholar
Godlewska, R, et al. (2008) Tip-α (hp0596 gene product) is a highly immunogenic Helicobacter pylori protein involved in colonization of mouse gastric mucosa. Current Microbiology 56(3), 279286.CrossRefGoogle ScholarPubMed
Watanabe, T, et al. (2010) Nucleolin as cell surface receptor for tumor necrosis factor-α inducing protein: a carcinogenic factor of Helicobacter pylori. Journal of Cancer Research and Clinical Oncology 136(6), 911921.CrossRefGoogle ScholarPubMed
Fujiki, H, Watanabe, T and Suganuma, M (2014) Cell-surface nucleolin acts as a central mediator for carcinogenic, anti-carcinogenic, and disease-related ligands. Journal of Cancer Research and Clinical Oncology 140(5), 689699.CrossRefGoogle ScholarPubMed
Suganuma, M, et al. (2007) The unique carcinogenic factor Tipα in cancer microenvironment of Helicobacter pylori infection. Proc. AACR Annual Meeting 48(2007), 1337.Google Scholar
Kuzuhara, T, et al. (2007b) DNA-binding activity of TNF-α inducing protein from Helicobacter pylori. Biochemical and Biophysical Research Communications 362(4), 805810.CrossRefGoogle ScholarPubMed
Krzyżek, P and Gościniak, G (2018) Morphology of Helicobacter pylori as a result of peptidoglycan and cytoskeleton rearrangements. Przeglad Gastroenterologiczny 13(3), 182.Google ScholarPubMed
Azevedo, N, et al. (2007) Coccoid form of Helicobacter pylori as a morphological manifestation of cell adaptation to the environment. Applied and Environmental Microbiology 73(10), 34233427.CrossRefGoogle ScholarPubMed
Elhariri, M, et al. (2018) Occurrence of cagA+ vacA s1a m1 i1 Helicobacter pylori in farm animals in Egypt and ability to survive in experimentally contaminated UHT milk. Scientific Reports 8(1), 14260.CrossRefGoogle ScholarPubMed
Sisto, F, et al. (2000) Helicobacter pylori: ureA, cagA and vacA expression during conversion to the coccoid form. International Journal of Antimicrobial Agents 15(4), 277282.CrossRefGoogle Scholar
Saxena, A, Mukhopadhyay, AK and Nandi, SP (2020) Helicobacter pylori: Perturbation and restoration of gut microbiome. Journal of Biosciences 45, 115.CrossRefGoogle ScholarPubMed
Loke, MF, et al. (2016) Understanding the dimorphic lifestyles of human gastric pathogen Helicobacter pylori using the SWATH-based proteomics approach. Scientific Reports 6(1), 18.CrossRefGoogle ScholarPubMed
Krzyżek, P and Grande, R (2020) Transformation of Helicobacter pylori into coccoid forms as a challenge for research determining activity of antimicrobial substances. Pathogens 9(3), 184.CrossRefGoogle ScholarPubMed
Kadkhodaei, S, Siavoshi, F and Akbari Noghabi, K (2020) Mucoid and coccoid Helicobacter pylori with fast growth and antibiotic resistance. Helicobacter 25(2), e12678.CrossRefGoogle ScholarPubMed
Chaput, C, et al. (2006) Role of AmiA in the morphological transition of Helicobacter pylori and in immune escape. PLoS pathogens 2(9), e97.CrossRefGoogle ScholarPubMed
Krzyżek, P, et al. (2021) Myricetin as an antivirulence compound interfering with a morphological transformation into coccoid forms and potentiating activity of antibiotics against Helicobacter pylori. International Journal of Molecular Sciences 22(5), 2695.CrossRefGoogle ScholarPubMed
Shahamat, M, et al. (1993) Use of autoradiography to assess viability of Helicobacter pylori in water. Applied and Environmental Microbiology 59(4), 12311235.CrossRefGoogle ScholarPubMed
Ng, BL, et al. (2003) Immune responses to differentiated forms of Helicobacter pylori in children with epigastric pain. Clinical and Vaccine Immunology 10(5), 866869.CrossRefGoogle ScholarPubMed
Mizoguchi, H, et al. (1998) Diversity in protein synthesis and viability of Helicobacter pylori coccoid forms in response to various stimuli. Infection and Immunity 66(11), 55555560.CrossRefGoogle ScholarPubMed
Cellini, L, et al. (1994) Helicobacter pylori a fickle germ. Microbiology and Immunology 38(1), 2530.CrossRefGoogle ScholarPubMed
Catrenich, C and Makin, K (1991) Characterization of the morphologic conversion of Helicobacter pylori from bacillary to coccoid forms. Scandinavian Journal of Gastroenterology 26(Supp.181), 5864.CrossRefGoogle Scholar
Bode, G, Mauch, F and Malfertheiner, P (1993) The coccoid forms of Helicobacter pylori. Criteria for their viability. Epidemiology & Infection 111(3), 483490.CrossRefGoogle ScholarPubMed
Rossetti, V, et al. (2013) Phenotypic diversity of multicellular filamentation in oral streptococci. PLoS ONE 8(9), e76221.CrossRefGoogle ScholarPubMed
Krzyżek, P, Biernat, MM and Gościniak, G (2019) Intensive formation of coccoid forms as a feature strongly associated with highly pathogenic Helicobacter pylori strains. Folia Microbiologica 64(3), 273281.CrossRefGoogle ScholarPubMed
Ricci, V, Romano, M and Boquet, P (2011) Molecular cross-talk between Helicobacter pylori and human gastric mucosa. World Journal of Gastroenterology 17(11), 1383.CrossRefGoogle ScholarPubMed
Dubois, A and Borén, T (2007) Helicobacter pylori is invasive and it may be a facultative intracellular organism. Cellular Microbiology 9(5), 11081116.CrossRefGoogle ScholarPubMed
Lina, TT, et al. (2014) Immune evasion strategies used by Helicobacter pylori. World Journal of Gastroenterology 20(36), 12753.CrossRefGoogle ScholarPubMed
Sijmons, D, et al. (2022) Helicobacter pylori and the role of lipopolysaccharide variation in innate immune evasion. Frontiers in Immunology 13(13), 868225.CrossRefGoogle ScholarPubMed
Neuper, T, et al. (2022) Beyond the gastric epithelium—the paradox of Helicobacter pylori-induced immune responses. Current Opinion in Immunology 76, 102208.CrossRefGoogle ScholarPubMed
Kaebisch, R, et al. (2014) Helicobacter pylori cytotoxin-associated gene A impairs human dendritic cell maturation and function through IL-10–mediated activation of STAT3. The Journal of Immunology 192(1), 316323.CrossRefGoogle ScholarPubMed
Käbisch, R, et al. (2016) Helicobacter pylori γ-glutamyltranspeptidase induces tolerogenic human dendritic cells by activation of glutamate receptors. The Journal of Immunology 196(10), 42464252.CrossRefGoogle ScholarPubMed
Frauenlob, T, et al. (2022) Helicobacter pylori infection of primary human monocytes boosts subsequent immune responses to LPS. Frontiers in Immunology 13, 847958.CrossRefGoogle ScholarPubMed
Altobelli, A, et al. (2019) Helicobacter pylori VacA targets myeloid cells in the gastric lamina propria to promote peripherally induced regulatory T-cell differentiation and persistent infection. MBio 10(2), e00261–19.CrossRefGoogle ScholarPubMed
Kawai, T and Akira, S (2011) Toll-like receptors and their crosstalk with other innate receptors in infection and immunity. Immunity 34(5), 637650.CrossRefGoogle ScholarPubMed
Castaño-Rodríguez, N, Kaakoush, NO and Mitchell, HM (2014) Pattern-recognition receptors and gastric cancer. Frontiers in Immunology 5, 336.Google ScholarPubMed
Mogensen, TH (2009) Pathogen recognition and inflammatory signaling in innate immune defenses. Clinical Microbiology Reviews 22(2), 240273.CrossRefGoogle ScholarPubMed
Akira, S, Uematsu, S and Takeuchi, O (2006) Pathogen recognition and innate immunity. Cell 124(4), 783801.CrossRefGoogle ScholarPubMed
Takeuchi O, AS (2010) Pattern recognition receptors and inflammation. Cell 140(6), 805820.CrossRefGoogle ScholarPubMed
Takeda, K and Akira, S (2004) TLR signaling pathways. Seminars in Immunology 16, 39.CrossRefGoogle ScholarPubMed
Neuper, T, et al. (2020) TLR2, TLR4 and TLR10 shape the cytokine and chemokine release of Helicobacter pylori -infected human DCs. International Journal of Molecular Sciences 21(11), 3897.CrossRefGoogle ScholarPubMed
Kumar Pachathundikandi, S, et al. (2011) Induction of TLR-2 and TLR-5 expression by Helicobacter pylori switches cag PAI-dependent signalling leading to the secretion of IL-8 and TNF-α. PLoS ONE 6(5), e19614.CrossRefGoogle Scholar
Koch, KN, et al. (2015) Helicobacter urease–induced activation of the TLR2/NLRP3/IL-18 axis protects against asthma. The Journal of Clinical Investigation 125(8), 32973302.CrossRefGoogle ScholarPubMed
Müller, A and Hartung, ML (2016) Helicobacter pylori and the Host Immune Response. In Helicobacter pylori Research. Springer, pp. 299323.CrossRefGoogle Scholar
Cullen, TW, et al. (2011) Helicobacter pylori versus the host: remodeling of the bacterial outer membrane is required for survival in the gastric mucosa. PLoS Pathogens 7(12), e1002454.CrossRefGoogle ScholarPubMed
Schmidinger, B, et al. (2022) Helicobacter pylori binds human Annexins via Lipopolysaccharide to interfere with Toll-like receptor 4 signaling. PLoS Pathogens 18(2), e1010326.CrossRefGoogle ScholarPubMed
Andersen-Nissen, E, et al. (2005) Evasion of Toll-like receptor 5 by flagellated bacteria. Proceedings of the National Academy of Sciences 102(26), 92479252.CrossRefGoogle ScholarPubMed
Smith, MF, et al. (2003) Toll-like receptor (TLR) 2 and TLR5, but not TLR4, are required for Helicobacter pylori-induced NF-κB activation and chemokine expression by epithelial cells. Journal of Biological Chemistry 278(35), 3255232560.CrossRefGoogle Scholar
Hemmi H, TO, Kawai, T, Kaisho, T, Sato, S, Sanjo, H, Matsumoto, M, Hoshino, K, Wagner, H, Takeda, K and Akira, S (2000) A Toll-like receptor recognizes bacterial DNA. Nature 408(6813), 740745.CrossRefGoogle ScholarPubMed
Otani, K, et al. (2012) Toll-like receptor 9 signaling has anti-inflammatory effects on the early phase of Helicobacter pylori-induced gastritis. Biochemical and Biophysical Research Communications 426(3), 342349.CrossRefGoogle ScholarPubMed
Varga, MG, et al. (2016) TLR9 activation suppresses inflammation in response to Helicobacter pylori infection. American Journal of Physiology-Gastrointestinal and Liver Physiology 311(5), G852G858.CrossRefGoogle ScholarPubMed
Sun, X, et al. (2013) TLR2 mediates Helicobacter pylori-induced tolerogenic immune response in mice. PLoS ONE 8(9), e74595.CrossRefGoogle ScholarPubMed
Lee, CYQ, et al. (2022) Helicobacter pylori infection elicits Type I interferon response in human monocytes via toll-like receptor 8 signaling. Journal of Immunology Research, 3861518.Google ScholarPubMed
Rad, R, et al. (2009) Extracellular and intracellular pattern recognition receptors cooperate in the recognition of Helicobacter pylori. Gastroenterology 136(7), 22472257.CrossRefGoogle ScholarPubMed
Dooyema, SD, et al. (2022) Helicobacter pylori actively suppresses innate immune nucleic acid receptors. Gut Microbes 14(1), 2105102.CrossRefGoogle ScholarPubMed
Gringhuis, SI, et al. (2009) Carbohydrate-specific signaling through the DC-SIGN signalosome tailors immunity to Mycobacterium tuberculosis, HIV-1 and Helicobacter pylori. Nature Immunology 10(10), 10811088.CrossRefGoogle ScholarPubMed
Devi, S, Rajakumara, E and Ahmed, N (2015) Induction of Mincle by Helicobacter pylori and consequent anti-inflammatory signaling denote a bacterial survival strategy. Scientific Reports 5, 15049.CrossRefGoogle ScholarPubMed
Girardin, SE, et al. (2003a) Nod1 detects a unique muropeptide from gram-negative bacterial peptidoglycan. Science 300(5625), 15841587.CrossRefGoogle ScholarPubMed
Girardin, SE, et al. (2003b) Nod2 is a general sensor of peptidoglycan through muramyl dipeptide (MDP) detection. Journal of Biological Chemistry 278(11), 88698872.CrossRefGoogle ScholarPubMed
Liu, W, et al. (2010) Olfactomedin 4 down-regulates innate immunity against Helicobacter pylori infection. Proceedings of the National Academy of Sciences 107(24), 1105611061.CrossRefGoogle ScholarPubMed
Minaga, K, et al. (2018) Nucleotide-binding oligomerization domain 1 and Helicobacter pylori infection: A review. World Journal of Gastroenterology 24(16), 1725.CrossRefGoogle ScholarPubMed
Suarez, G, et al. (2019) Nod1 imprints inflammatory and carcinogenic responses toward the gastric pathogen Helicobacter pylori. Cancer Research 79(7), 16001611.CrossRefGoogle ScholarPubMed
Pfannkuch, L, et al. (2019) ADP heptose, a novel pathogen-associated molecular pattern identified in Helicobacter pylori. The FASEB Journal 33(8), 90879099.CrossRefGoogle ScholarPubMed
Zimmermann, S, et al. (2017) ALPK1-and TIFA-dependent innate immune response triggered by the Helicobacter pylori type IV secretion system. Cell Reports 20(10), 23842395.CrossRefGoogle ScholarPubMed
Zhou, P, et al. (2018) Alpha-kinase 1 is a cytosolic innate immune receptor for bacterial ADP-heptose. Nature 561(7721), 122126.CrossRefGoogle ScholarPubMed
Maubach, G, et al. (2021) TIFA has dual functions in Helicobacter pylori-induced classical and alternative NF-κB pathways. EMBO Reports 22(9), e52878.CrossRefGoogle ScholarPubMed
Gaudet, RG and Gray-Owen, SD (2016) Heptose sounds the alarm: innate sensing of a bacterial sugar stimulates immunity. PLoS Pathogens 12(9), e1005807.CrossRefGoogle ScholarPubMed
Gall, A, et al. (2017) TIFA signaling in gastric epithelial cells initiates the cag type 4 secretion system-dependent innate immune response to Helicobacter pylori infection. MBio 8(4), e01168–17.CrossRefGoogle ScholarPubMed
Stein, SC, et al. (2017) Helicobacter pylori modulates host cell responses by CagT4SS-dependent translocation of an intermediate metabolite of LPS inner core heptose biosynthesis. PLoS Pathogens 13(7), e1006514.CrossRefGoogle ScholarPubMed
Patel, S, et al. (2013) Helicobacter pylori downregulates expression of human β-defensin 1 in the gastric mucosa in a type IV secretion-dependent fashion. Cellular Microbiology 15(12), 20802092.CrossRefGoogle Scholar
Bauer, B, et al. (2012) The Helicobacter pylori virulence effector CagA abrogates human β-defensin 3 expression via inactivation of EGFR signaling. Cell Host & Microbe 11(6), 576586.CrossRefGoogle ScholarPubMed
Chmiela, M, et al. (2017) Host pathogen interactions in Helicobacter pylori related gastric cancer. World Journal of Gastroenterology 23(9), 1521.CrossRefGoogle ScholarPubMed
He, H, et al. (2021) Helicobacter pylori CagA interacts with SHP-1 to suppress the immune response by targeting TRAF6 for K63-linked ubiquitination. The Journal of Immunology 206(6), 11611170.CrossRefGoogle ScholarPubMed
Gang Liu, Y, et al. (2019) Abrogation of cathepsin C by Helicobacter pylori impairs neutrophil activation to promote gastric infection. The FASEB Journal 33(4), 50185033.CrossRefGoogle Scholar
Gobert, AP, et al. (2014) Heme oxygenase-1 dysregulates macrophage polarization and the immune response to Helicobacter pylori. The Journal of Immunology 193(6), 30133022.CrossRefGoogle ScholarPubMed
Gray-Owen, SD and Blumberg, RS (2006) CEACAM1: contact-dependent control of immunity. Nature Reviews Immunology 6(6), 433446.CrossRefGoogle ScholarPubMed
Gur, C, et al. (2019) The Helicobacter pylori HopQ outermembrane protein inhibits immune cell activities. Oncoimmunology 8(4), e1553487.CrossRefGoogle ScholarPubMed
Djekic, A and Müller, A (2016) The immunomodulator VacA promotes immune tolerance and persistent Helicobacter pylori infection through its activities on T-cells and antigen-presenting cells. Toxins 8(6), 187.CrossRefGoogle ScholarPubMed
Utsch, C and Haas, R (2016) VacA's induction of VacA-containing vacuoles (VCVs) and their immunomodulatory activities on human T cells. Toxins 8(6), 190.CrossRefGoogle ScholarPubMed
Zheng, PY and Jones, NL (2003) Helicobacter pylori strains expressing the vacuolating cytotoxin interrupt phagosome maturation in macrophages by recruiting and retaining TACO (coronin 1) protein. Cellular Microbiology 5(1), 2540.CrossRefGoogle ScholarPubMed
Amedei, A, et al. (2006) The neutrophil-activating protein of Helicobacter pylori promotes Th1 immune responses. The Journal of Clinical Investigation 116(4), 10921101.CrossRefGoogle ScholarPubMed
Schmalstig, AA, et al. (2018) Noncatalytic antioxidant role for Helicobacter pylori urease. Journal of Bacteriology 200(17), e00124–18.CrossRefGoogle ScholarPubMed
Gerhard, M, et al. (2005) A secreted low-molecular-weight protein from Helicobacter pylori induces cell-cycle arrest of T cells. Gastroenterology 128(5), 13271339.CrossRefGoogle ScholarPubMed
Beigier-Bompadre, M, et al. (2011) Modulation of the CD4+ T-cell response by Helicobacter pylori depends on known virulence factors and bacterial cholesterol and cholesterol α-glucoside content. Journal of Infectious Diseases 204(9), 13391348.CrossRefGoogle ScholarPubMed
Ramarao, N, et al. (2000) Helicobacter pylori inhibits phagocytosis by professional phagocytes involving type IV secretion components. Molecular Microbiology 37(6), 13891404.CrossRefGoogle ScholarPubMed
Lewis, ND, et al. (2011) Immune evasion by Helicobacter pylori is mediated by induction of macrophage arginase II. The Journal of Immunology 186(6), 36323641.CrossRefGoogle ScholarPubMed
Tan, GMY, et al. (2015) Suppression of cell division-associated genes by Helicobacter pylori attenuates proliferation of RAW264. 7 monocytic macrophage cells. Scientific Reports 5(1), 116.Google ScholarPubMed
Chaturvedi, R, et al. (2004) Induction of polyamine oxidase 1 by Helicobacter pylori causes macrophage apoptosis by hydrogen peroxide release and mitochondrial membrane depolarization. Journal of Biological Chemistry 279(38), 4016140173.CrossRefGoogle ScholarPubMed
Asim, M, et al. (2010) Helicobacter pylori induces ERK-dependent formation of a phospho-c-Fos⋅ c-Jun activator protein-1 complex that causes apoptosis in macrophages. Journal of Biological Chemistry 285(26), 2034320357.CrossRefGoogle ScholarPubMed
Cheok, YY, et al. (2022) Innate immunity crosstalk with Helicobacter pylori: pattern recognition receptors and cellular responses. International Journal of Molecular Sciences 23(14), 7561.CrossRefGoogle ScholarPubMed
Yang, H and Hu, B (2022) Immunological perspective: Helicobacter pylori infection and gastritis. Mediators of Inflammation 2022, 156189.CrossRefGoogle ScholarPubMed
Morey, P, et al. (2018) Helicobacter pylori depletes cholesterol in gastric glands to prevent interferon gamma signaling and escape the inflammatory response. Gastroenterology 154(5), 13911404. e9.CrossRefGoogle ScholarPubMed
Mitchell, P, et al. (2007) Chronic exposure to Helicobacter pylori impairs dendritic cell function and inhibits Th1 development. Infection and Immunity 75(2), 810819.CrossRefGoogle ScholarPubMed
Alexander, SM, et al. (2021) Helicobacter pylori in human stomach: the inconsistencies in clinical outcomes and the probable causes. Frontiers in Microbiology 12, 713955.CrossRefGoogle ScholarPubMed
Arnold, IC, et al. (2011) Tolerance rather than immunity protects from Helicobacter pylori-induced gastric preneoplasia. Gastroenterology 140(1), 199209. e8.CrossRefGoogle ScholarPubMed
Robinson, K, et al. (2008) Helicobacter pylori-induced peptic ulcer disease is associated with inadequate regulatory T cell responses. Gut 57(10), 13751385.CrossRefGoogle ScholarPubMed
Harris, PR, et al. (2008) Helicobacter pylori gastritis in children is associated with a regulatory T-cell response. Gastroenterology 134(2), 491499.CrossRefGoogle ScholarPubMed
Lundgren, A, et al. (2005a) Mucosal FOXP3-expressing CD4+ CD25high regulatory T cells in Helicobacter pylori-infected patients. Infection and Immunity 73(1), 523531.CrossRefGoogle ScholarPubMed
Lundgren, A, et al. (2005b) Helicobacter pylori-specific CD4+ T cells home to and accumulate in the human Helicobacter pylori-infected gastric mucosa. Infection and Immunity 73(9), 56125619.CrossRefGoogle ScholarPubMed
Sayi, A, et al. (2011) TLR-2–Activated B cells suppress helicobacter-induced preneoplastic gastric immunopathology by inducing T regulatory-1 cells. The Journal of Immunology 186(2), 878890.CrossRefGoogle ScholarPubMed
Hitzler, I, et al. (2011) Dendritic cells prevent rather than promote immunity conferred by a Helicobacter vaccine using a mycobacterial adjuvant. Gastroenterology 141(1), 186196, e1.CrossRefGoogle ScholarPubMed
Sewald, X, et al. (2008) Integrin subunit CD18 Is the T-lymphocyte receptor for the Helicobacter pylori vacuolating cytotoxin. Cell Host & Microbe 3(1), 2029.CrossRefGoogle ScholarPubMed
Sewald, X, Jiménez-Soto, L and Haas, R (2011) PKC-dependent endocytosis of the Helicobacter pylori vacuolating cytotoxin in primary T lymphocytes. Cellular Microbiology 13(3), 482496.CrossRefGoogle ScholarPubMed
Fehri, LF, et al. (2010) Helicobacter pylori induces miR-155 in T cells in a cAMP-Foxp3-dependent manner. PLoS ONE 5(3), e9500.CrossRefGoogle Scholar
McBride, A, Konowich, J and Salgame, P (2013) Host defense and recruitment of Foxp3+ T regulatory cells to the lungs in chronic Mycobacterium tuberculosis infection requires toll-like receptor 2. PLoS Pathogens 9(6), e1003397.CrossRefGoogle Scholar
Oertli, M, et al. (2012) DC-derived IL-18 drives Treg differentiation, murine Helicobacter pylori-specific immune tolerance, and asthma protection. The Journal of Clinical Investigation 122(3), 10821096.CrossRefGoogle ScholarPubMed
Laur, AM, et al. (2016) Regulatory T cells may participate in Helicobacter pylori persistence in gastric MALT lymphoma: lessons from an animal model. Oncotarget 7(3), 3394.CrossRefGoogle ScholarPubMed
Tacconelli, E, et al. (2018) Discovery, research, and development of new antibiotics: the WHO priority list of antibiotic-resistant bacteria and tuberculosis. The Lancet Infectious Diseases 18(3), 318327.CrossRefGoogle ScholarPubMed
Megraud, F, et al. (2013) Helicobacter pylori resistance to antibiotics in Europe and its relationship to antibiotic consumption. Gut 62(1), 3442.CrossRefGoogle ScholarPubMed
Sung, H, et al. (2021) Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA: a Cancer Journal for Clinicians 71(3), 209249.Google ScholarPubMed
Yoon, H, et al. (2013) Meta-analysis: Is sequential therapy superior to standard triple therapy for Helicobacter pylori infection in Asian adults? Journal of Gastroenterology and Hepatology 28(12), 18011809.CrossRefGoogle ScholarPubMed
Soudi, H, et al. (2021) Evaluation of Helicobacter pylori OipA protein as a vaccine candidate and propolis as an adjuvant in C57BL/6 mice. Iranian Journal of Basic Medical Sciences 24(9), 1220.Google ScholarPubMed
Holmgren, J, et al. (2018) Preclinical immunogenicity and protective efficacy of an oral Helicobacter pylori inactivated whole cell vaccine and multiple mutant cholera toxin: a novel and non-toxic mucosal adjuvant. Vaccine 36(41), 62236230.CrossRefGoogle ScholarPubMed
Guo, L, et al. (2019) Therapeutic protection against Helicobacter pylori infection in Mongolian gerbils by oral immunization with a tetravalent epitope-based vaccine with polysaccharide adjuvant. Frontiers in Immunology 10, 1185.CrossRefGoogle ScholarPubMed
Cen, Q, et al. (2021) Immune evaluation of a Saccharomyces cerevisiae-based oral vaccine against Helicobacter pylori in mice. Helicobacter 26(1), e12772.CrossRefGoogle ScholarPubMed
Ansari, H, Tahmasebi-Birgani, M and Bijanzadeh, M (2021) DNA vaccine containing Flagellin A gene induces significant immune responses against Helicobacter pylori infection: An in vivo study. Iranian Journal of Basic Medical Sciences 24(6), 796.Google ScholarPubMed
Zhang, Y, et al. (2022) Perspectives from recent advances of Helicobacter pylori vaccines research. Helicobacter 27(6), e12926.CrossRefGoogle ScholarPubMed
Dos Santos Viana, I, et al. (2021) Vaccine development against Helicobacter pylori: From ideal antigens to the current landscape. Expert Review of Vaccines 20(8), 989999.CrossRefGoogle Scholar
Robinson, K, Kaneko, K and Andersen, LP (2017) Helicobacter: Inflammation, immunology and vaccines. Helicobacter 22, e12406.CrossRefGoogle ScholarPubMed
Guo, L, et al. (2013) Immunological features and efficacy of the reconstructed epitope vaccine CtUBE against Helicobacter pylori infection in BALB/c mice model. Applied Microbiology and Biotechnology 97(6), 23672378.CrossRefGoogle ScholarPubMed
Akhiani, AA, et al. (2005) IgA antibodies impair resistance against Helicobacter pylori infection: studies on immune evasion in IL-10-deficient mice. The Journal of Immunology 174(12), 81448153.CrossRefGoogle ScholarPubMed
Sun, H, et al. (2018) Immunodominant antigens that induce Th1 and Th17 responses protect mice against Helicobacter pylori infection. Oncotarget 9(15), 12050.CrossRefGoogle ScholarPubMed
Sutton, P (2015) At last, vaccine-induced protection against Helicobacter pylori. Lancet (London. England 386(10002), 1424.Google Scholar
Mejías-Luque, R and Gerhard, M (2017) Immune evasion strategies and persistence of Helicobacter pylori. In Tegtmeyer, N (ed.), Molecular Pathogenesis and Signal Transduction by Helicobacter pylori. Springer, pp. 5371.CrossRefGoogle ScholarPubMed
Figure 0

Table 1. Virulence factors involving in different aspects of H. pylori pathogenesis (colonization, immune escape, and disease induction), their functions, and associated diseases.

Figure 1

Figure 1. Signalling pathways induced by H. pylori CagA. CagA oncoprotein is directly injected into host cells by CagT4SS. This protein interacts with myriad signalling factors to manipulate different signal transduction cascades. CagA, cytotoxin-associated gene A; CagT4SS, Cag type IV secretion system; NFAT, nuclear factor of activated T cells; Bcl-2, B-cell lymphoma 2; c-Met, c-mesenchymal epithelial transition factor; Grb2, growth factor receptor-bound protein 2; GSK3β Wnt, glycogen synthase kinase 3β Wnt; SHP2, Src homology-2 domain-containing protein tyrosine phosphatase-2; TRAF, TNF receptor-associated factor; CD44, cluster of differentiation 44; Akt, Ak strain transforming; BIM, BCL-2-interacting mediator of cell death; MCL1, myeloid cell leukaemia-1.

Figure 2

Figure 2. H. pylori evasion of innate immune recognition. Structurally modified pathogen-associated molecular patterns (PAMPs) enable H. pylori to evade the detection by pro-inflammatory Toll-like receptors (TLRs). H. pylori tetra-acylated LPS is less biologically active and is not sensed by TLR4. The TLR5 cannot detect the mutated TLR5 binding site of H. pylori flagellin. TLRs 9 detects H. pylori DNA (CpG DNA) and TLR2 detects H. pylori LPS; these TLRs predominantly activate anti-inflammatory signalling pathways and IL-10 expression. The cytosolic receptor RIG-I and endosomal receptor TLR8 sense 5ʹ triphosphorylated RNA which elicit IFN-α/β response. H. pylori fucosylated DC-SIGN ligands are another activator of anti-inflammatory genes. Besides, theses ligands block IFN-α/β response. Notice that the different cell types express various pattern recognition receptors (PRRs). TLR2/4/5/8/9, Toll-like receptor 2/4/5/8/9; LPS, lipopolysaccharide; DC-SIGN, dendritic cell-specific intercellular adhesion molecule-3 grabbing non-integrin; SRC, steroid receptor coactivator; IFN-α/β, α/β interferon; IRF3/7, interferon regulatory factor3/7; AP-1, activator protein 1; NF-κB, nuclear factor-κB; P50/P65, NF-κB p50/p65 heterodimer; TIR, Toll/interleukin-1 receptor domain; MYD88, myeloid differentiation primary response gene 88; DD, death domain; CARD, caspase activation and recruitment domain; RIG-I, retinoic acid-inducible gene I, CpG DNA, 5'—C—phosphate—G—3'.

Figure 3

Figure 3. CagT4SS-injected virulence factors (e.g., ADP-Heptose, CagA and PG) trigger NF-κB-mediated innate immune response in gastric epithelial cells. Upon binding of ADP-Heptose, an intermediate metabolite produced during the biosynthesis of LPS, to ALPK1, the ALPK1-TIFA signalling pathway is triggered, which results in an NF-kB-dependent pro-inflammatory response. Following TIFA activation, PG activates NOD1, which leads to NF-κB-mediated pro-inflammatory responses. Finally, the interaction of CagA and host TAK1 triggers NF-κB-mediated pro-inflammatory responses. Besides, CagT4SS-injected CpG DNA is detected by TLR9 which result in activation of anti-inflammatory signalling pathways. LPS, lipopolysaccharide; CagT4SS, Cag type IV secretion system; ADP-Heptose, ADP-β-D-manno-heptose; PG, peptidoglycan; CagA, cytotoxin-associated gene A; ALPK1, alpha kinase 1; Tak1, TGF-β activated kinase 1; TRAF6, tumour necrosis factor receptor (TNFR)-associated factor 6; TIFA, TRAF interacting forkhead-associated protein A; NOD1, nucleotide-binding oligomerisation domain 1; RIPK2, receptor-interacting-serine/threonine-protein kinase 2; NF-κB, nuclear factor-κB; P50/P65, NF-κB p50/p65 heterodimer; CpG DNA, 5'—C—phosphate—G—3'; TLR9, Toll-like receptor 9; TIR, Toll/interleukin-1 receptor domain; DD, death domain.

Figure 4

Figure 4. H. pylori subverts T-cell-mediated immunity using the secreted virulence factors VacA and GGT. Following the internalization, Cytoplasmic VacA prevents nuclear translocation of NFAT by inhibiting its dephosphorylation by the Ca2+/calmodulin-dependent phosphatase calcineurin and thereby blocks IL-2 production and subsequent T-cell activation and proliferation. The GGT prevents the proliferation of T cells via interfering in the G1 phase of the cell cycle. TCR, T-cell receptor; MHCII, major histocompatibility complex class II; VacA, vacuolating cytotoxin A; GGT, γ-glutamyl transpeptidase; NFAT, nuclear factor of activated T cells; IL-2, Interleukin-2; CnA/B, calcineurin A/B subunits; CaM, calmodulin; LFA-1, lymphocyte function-associated antigen-1; PKCζ/ η, protein kinase Cζ/η.

Figure 5

Figure 5. The immunological response elicited by H. pylori infection. H. pylori is genetically highly variable and expresses various virulence factors, including adhesins (BabA/SabA), enzymes (urease), toxins (VacA) and effector proteins (CagA) which are involved in bacterial pathogenesis. The immunity and immunosuppression are the consequences of the maturation status of dendritic cells (DCs), which direct T helper cell differentiation. CagT4SS, Cag type IV secretion system; CagA, cytotoxin-associated gene A; VacA, vacuolating cytotoxin A; BabA, blood group antigen-binding adhesin A; SabA, sialic acid-binding adhesin A; DC, dendritic cell; Mφ, macrophage cell; PMN, polymorphonuclear leukocyte; Treg, regulatory T cell; Th1/17, T helper cell1/17; GM-CSF, granulocyte–macrophage colony-stimulating factor; TNF-α, tumour necrosis factor-α; IFN-γ, interferon-γ; TGF-β, transforming growth factor-beta, CXCL1/2, C-X-C motif chemokine ligand 1/2; IL-1/2/6/8/10/12/17/23, interleukin 1/2/6/8/10/12/17/23.

Figure 6

Figure 6. Vaccine development against H. pylori. CagA, cytotoxin-associated gene A; VacA, vacuolating cytotoxin A; BabA, blood group antigen-binding adhesin A; HspA, heat shock protein A; FliD, Flagellar hook-associated protein 2.