Skip to main content
Erschienen in: Gut Pathogens 1/2023

Open Access 01.12.2023 | Research

Polymorphism of virulence genes and biofilm associated with in vitro induced resistance to clarithromycin in Helicobacter pylori

verfasst von: Naim Asyraf Rosli, Anis Rageh Al-Maleki, Mun Fai Loke, Eng Guan Chua, Mohammed Abdelfatah Alhoot, Jamuna Vadivelu

Erschienen in: Gut Pathogens | Ausgabe 1/2023

Abstract

Background

Clarithromycin-containing triple therapy is commonly used to treat Helicobacter pylori infections. Clarithromycin resistance is the leading cause of H. pylori treatment failure. Understanding the specific mutations that occur in H. pylori strains that have evolved antibiotic resistance can help create a more effective and individualised antibiotic treatment plan. However, little is understood about the genetic reprogramming linked to clarithromycin exposure and the emergence of antibiotic resistance in H. pylori. Therefore, this study aims to identify compensatory mutations and biofilm formation associated with the development of clarithromycin resistance in H. pylori. Clarithromycin-sensitive H. pylori clinical isolates were induced to develop clarithromycin resistance through in vitro exposure to incrementally increasing concentration of the antibiotic. The genomes of the origin sensitive isolates (S), isogenic breakpoint (B), and resistant isolates (R) were sequenced. Single nucleotide variations (SNVs), and insertions or deletions (InDels) associated with the development of clarithromycin resistance were identified. Growth and biofilm production were also assessed.

Results

The S isolates with A2143G mutation in the 23S rRNA gene were successfully induced to be resistant. According to the data, antibiotic exposure may alter the expression of certain genes, including those that code for the Cag4/Cag protein, the vacuolating cytotoxin domain-containing protein, the sel1 repeat family protein, and the rsmh gene, which may increase the risk of developing and enhances virulence in H. pylori. Enhanced biofilm formation was detected among R isolates compared to B and S isolates. Furthermore, high polymorphism was also detected among the genes associated with biofilm production.

Conclusions

Therefore, this study suggests that H. pylori may acquire virulence factors while also developing antibiotic resistance due to clarithromycin exposure.
Hinweise

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1186/​s13099-023-00579-4.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
BHI
Brain Heart Infusion
CagA
Cytotoxin-associated gene A
CA
Chocolate agar
CFU
Colony forming unit
EUCAST
European Committee on Antimicrobial Susceptibility Testing
H. pylori
Helicobacter pylori
InDels
Insertions or deletions
MIC
Minimum Inhibitory Concentration
OD
Optical density
PBS
Phosphate buffer saline
PPI
Proton pump inhibitor
QC
Quality control
RAPD
Random amplification of polymorphic DNA
SNVs
Single nucleotide variations
T4SS
Type IV secretion system
S
Parental sensitive isolate
B
Breakpoint isogenic isolate
R
Induced resistant isogenic isolate

Introduction

Helicobacter pylori is a spiral-shaped Gram-negative bacteria that thrives in the mucus and epithelial stomach mucosa of more than half of the world’s adult population [1]. It causes a variety of gastrointestinal disorders, such as gastritis, gastric ulcer, and duodenal ulcer, and is also associated with gastric cancer [2]. The standard triple therapy for H. pylori infection includes a proton pump inhibitor (PPI) and two antibiotics (amoxicillin with clarithromycin, or metronidazole) [3]. However, misuse and overuse of antimicrobials is causing the steady growth in antibiotic resistance and is severely hindering the eradication of H. pylori. Similarly, the success rates of clarithromycin as a component of the first-line therapy for H. pylori infections will continue to fall because of overuse of the antibiotic for the treatment of H. pylori [4, 5].
Clarithromycin is a bacteriostatic antibiotic that targets the peptidyl transferase loop of the V domain of the 23S ribosomal RNA (23S rRNA) molecule [6]. H. pylori clarithromycin resistance has been reported to be closely associated with point mutations in two neighbouring 23S rRNA nucleotides, 2142 and 2143 [4], which reduces ribosome affinity for the macrolide, leading to enhanced resistance [7, 8]. However, other investigations revealed only 40–80% of clarithromycin resistant H. pylori had these 23S rRNA point mutations [6, 912]. Consistently, the A2143G point mutation has been observed in both clarithromycin-sensitive and -resistant H. pylori [6, 13, 14]. Therefore, alternative mechanisms might play a role in clarithromycin resistance in H. pylori [10, 15, 16]. Biofilm formation, which is also a virulence mechanism, is also a potential resistance mechanism used by bacteria [17]. Bacteria protect themselves from host defence, disinfectants, and antibiotics by forming a biofilm. The bacteria in a biofilm are more resistant to antimicrobial agents and can exhibit a 10 to 1000-fold increase in antibiotic resistance compared to the same bacteria existing in a planktonic form [17, 18].
The prevalence of resistant patients with no prior history of clarithromycin-containing eradication treatment was 13.3% while resistance increased to 51.4% in previously treated patients as secondary (acquired) resistance [7]. Some H. pylori strains developed clarithromycin resistance in response to exposure to the antibiotic while others do not. We hypothesized that H. pylori strains that developed resistance to the antibiotic undergo specific genetic reprogramming and understanding these specific mutations can aid in determining a more effective and personalized antibiotic therapeutic regime. However, little is known about the genetic reprogramming associated with exposure to clarithromycin and antibiotic resistance development in H. pylori. Therefore, in this study, comparative genomic analysis was performed on clarithromycin-sensitive B isogenic isolates of H. pylori in comparison to their parental clarithromycin-sensitive clinical isolates and in vitro induced R isogenic isolates. Induced isolates collected one passage immediately prior to becoming clarithromycin-resistant were taken as the B isogenic isolates. Genetic alterations found in breakpoint isolates may not be directly associated to clarithromycin resistance, but they may serve to condition the organism to develop clarithromycin resistance. In addition, biofilm formed by these H. pylori isolates were compared to investigate for possible correlation between biofilm formation and exposure to clarithromycin or development of antibiotic resistance.

Methodology

Bacterial growth and culture conditions

Helicobacter pylori from the glycerol stocks of the clinical bacterial archival collection of the Helicobacter Research Laboratory (UM Marshall Centre) at the Universiti Malaya were used in this study. The H. pylori isolates were cultured on non-selective chocolate agar (CA) plates (Oxoid Ltd., UK) supplemented with 5% defibrinated horse blood and incubated at 37 °C in a 10% CO2 incubator for 3 days. To minimize the chance of mixed cultures, all the stock archival cultures (sweep cultures) were grown on selective CA plates supplemented with vancomycin (10 μg/mL) (Amresco Inc., Ohio), amphotericin B (5 μg/mL) (Bio-world Inc., USA), trimethoprim (5 μg/mL) (Santa Cruz Biotechnology Inc., USA), and nalidixic acid (20 μg/mL) (Bio-world Inc., USA) to obtain well-isolated colonies. Each of these well-isolated colonies are treated as individual clonal isolates and sub-cultured on fresh CA plates to get sufficient material. H. pylori was confirmed by rapid urease test, catalase test, oxidase test, and 16S rRNA PCR using forward primer 5′-CTG GAG AGA CTA AGC CCT CC-3′ and reverse primer 5′-ATT ACT GAC GCT GAT TGT GC-3′ [19].

Minimum inhibitory concentration (MIC)

The MICs of clarithromycin against H. pylori was determined on non-selective CA plate using MIC Test Strip (Calbiochem, Germany) according to the manufacturer’s instructions. Briefly, viable H. pylori colonies from non-selective CA plates grown for 3 days (72 h) were harvested and inoculated into the Brain Heart Infusion (BHI) broth. The turbidity of the suspension was adjusted by visual comparison to the McFarland standard no. 3, which is approximately ∼ 9.0 × 108 CFU/mL and the suspension was spread onto a fresh non-selective CA plate with a clarithromycin-impregnated strip. The CA plates were incubated at 37 °C in a 10% CO2 incubator for 3 days. A drop-shaped inhibition zone intersects the graded test strip at the MIC of the antibiotic. The experiments were performed in triplicate with 3 biological replicates. Based on European Committee on Antimicrobial Susceptibility Testing (EUCAST) standards (version 13.0), the MIC breakpoint for clarithromycin is > 0.25 μg/mL. Clarithromycin sensitive isolates from the collection were selected for induction experiment.

In vitro clarithromycin resistance induction

Clarithromycin-sensitive H. pylori isolates were induced by the method according to Yan et al. [20] with modifications. Briefly, H. pylori isolates were exposed to incrementally doubling concentrations of clarithromycin from 0.0156 to 32 µg/mL incorporated into CA plate. To ensure that the induced strains were stable, MICs of the R, B and S isolates were confirmed using MIC Strip Test as previously described after ten passages of the R isolates on non-selective CA plate to determine the stability of resistance and after storage frozen at − 80 °C for 3–5 months (Additional file 2 and Additional file 3: Fig. S1).

Random amplification of polymorphic DNA-polymerase chain reaction (RAPD-PCR)

The identity between resistant strains and their corresponding parental sensitive strains before induction were verified by RAPD-PCR typing using primers as previously described [21]. The primers were 1254 5′-CCG CAG CCA A-3′, 1281 5′-AAC GCG CAA C-3′ and 1283 5′-GCG ATC CCC A-3′. The conditions for PCR amplification were denaturation at 95 °C for 3 min, followed by 45 cycles of denaturation at 95 °C for 1 min, annealing at 36 °C for 1 min, and extension at 72 °C for 1 min: and then a final extension at 72 °C for 5 min.

Sanger sequencing

The bacteria were grown for 3 days in 10 mL of BHI broth supplemented with 1% ß-cyclodextrin and 0.4% yeast extract and incubated at 37 °C in a 10% CO2 incubator. To collect the pellet, the bacterial broth was centrifuged for 10 min at 8000 rpm. The DNA of S, B and R isolates were extracted using the MasterPure™ Complete DNA and RNA Purification Kit (Lucigen, USA) and used for Sanger sequencing as well as whole genome sequencing. Two primers corresponding to bases 1820–1839 [Hp23-1: 5′-CCACAGCGATGTGGTCTCAG-3′] and from positions 2244–2225 [Hp23-2: 5′-CTCCatAAGAGCCAAAGCCC-3′] flanking a region of 425 bp within bacterial 23S rRNA peptidyl transferase as described by Ho et al. [22]. The PCR amplified products were sequenced on a ABI PRISM 3730xl Genetic Analyzer (Applied Biosystems, USA) by 1st Base (Singapore). Multiple sequence alignment was done by Bioedit version 7.2.5 and CodonCode Aligner version 10.0. The sequences were compared with the 23S rRNA of the reference genome (H. pylori UM 032 and H. pylori 26695).

Library preparation, and sequencing

The extracted DNA was used in library preparation. Following the manufacturer’s instructions, preparations for next-generation sequencing libraries were constructed using VAHTS Universal Pro DNA Library Prep Kit for Illumina V1. For each sample, 200 μg genomic DNA was randomly fragmented by Covaris ultrasonication system to an average size of 300–350 bp. The fragments underwent treatment with End Prep Enzyme Mix for end repairing, 5′ Phosphorylation and 3′ adenylated, to add adaptors to both ends. Next, DNA Cleanup beads selected the size of the adaptor-ligated DNA. Then, using P5 and P7 primers, each sample was then amplified by PCR for 8 cycles, with both primers carrying sequences which can anneal with flowcell to perform bridge PCR and P7 primer carrying a six-base index allowing for multiplexing. An Agilent 2100 Bioanalyzer was used to clean up and validate the PCR products. The qualified libraries were sequenced pair end PE150 (V1) on the Illumina Novaseq 6000.

Single-nucleotide variations (SNVs) and insertion and deletions (InDels) identification

The sequences of adaptors, PCR primers, content of N bases greater than 10%, and bases of poorer quality than 20 were removed using Cutadapt (V1.9.1). Using BWA (V0.7.17), clean data were mapped to the reference genome (H. pylori UM032, CP005490.3). Mapping results were processed by Picard (V2.25.7) to remove duplication. The HaplotypeCaller calls SNV/InDel with GATK (V3.8.1) software. Annotation for SNV/InDel was performed by Annovar (V21 Apr 2018).

Probability of mutation occurrence

The mutation rate of a gene was calculated by dividing the number of mutations (identified SNVs and InDels) within the gene with the number of base pair of the gene. Meanwhile, each strain’s threshold was calculated by dividing the total number of mutations within the strain with the total number of base pairs of the genome. The probability of occurrence of these mutations was computed by dividing the rate of mutation of the gene with the threshold of the corresponding strain.

Growth curve

The growth pattern of the H. pylori strains was performed according to Al-Maleki et al. [23]. Briefly, 12 isolates of H. pylori of S, B and R were cultured in BHI broth (Oxoid, UK) supplemented with 1% ß-cyclodextrin and 0.4% yeast extract in a 24-well plate and incubated at 37 °C in a 10% CO2 incubator. The optical density (OD600 nm) of the bacterial broth suspension was standardized to optical density (OD600 nm) of 0.02 using spectrophotometer (Thermo Fisher Scientific, USA) at t = 0 h. Bacterial broth suspension was collected and the OD600 nm is measured every 24 h over 7 days. Then, a serial tenfold dilutions in 1× sterile phosphate buffer saline (PBS) was performed and subsequently plated on CA plate followed by incubation at 37 °C in a 10% CO2 incubator for 3 days. The viable count was then performed to calculate the CFU which represent the number of living cells in the broth at every time point. Growth curve was performed as three independent replicates.

Biofilm formation

The inhibition of biofilm formation was assessed using methods that were described previously [24]. Briefly, H. pylori cultured on CA plate for 3 days in a 10% CO2 incubator were harvested and inoculated in BHI broth supplemented with 1% β-cyclodextrin and 0.4% yeast extract in a 24-well plate (Corning, USA) for another 3 days. The bacterial suspension was adjusted to 1–2 × 108 CFU/mL. The development of the biofilm was visually inspected at days 3, 5, and 7. The amount of biofilm produced was measured after day 7 using 0.1% crystal violet staining. After 30 min of gentle agitation at 100 rpm, the unbound crystal violet was removed. The biofilm was destained with a 19:1 ethanol-acetic acid solution after the crystal violet-treated wells were washed with distilled water. The solution collected was measured at OD595 nm on a spectrophotometer. The amount of crystal violet absorbed by the biofilm was determined by taking the mean absorbance value. The experiment was performed in triplicate.

Statistical analysis

IBM SPSS statistics version 22 software was used to perform the statistical analyses for growth curve and biofilm formation assays. One-way Analysis of Variance (ANOVA) and two-sample t-test were used to compare the means between variants. A p-value < 0.05 was considered statistically significant.

Results

Clarithromycin-resistant isogenic isolates

In total, 86 H. pylori clinical isolates collected from patients presenting for endoscopy at the Universiti Malaya Medical Centre between November 2011 and January 2015 were screened for clarithromycin resistance. Based on EUCAST resistant breakpoint of > 0.25 μg/mL, 60 isolates (69.77%) were susceptible to clarithromycin and 26 isolates (30.23%) were resistant to the antibiotic. Twenty randomly selected clarithromycin-sensitive H. pylori clinical isolates and one standard strain (NCTC 11637) were inducted by exposure to incrementally doubling concentrations of clarithromycin (Table 1 and Additional file 2). After 10 to 12 passages on clarithromycin CA plates, four clarithromycin-sensitive H. pylori isolates were successfully induced in vitro to become resistant with a success rate of 19.0% (Table 2). Clarithromycin CA plates were used to continue the induction until the isolates showed resistance to > 64 µg/mL on the MIC Test Strip. Notably, all the four successfully induced isolates originally harbour the A2143G variant despite been phenotypically susceptible to clarithromycin. On the other hand, the remaining 17 clarithromycin-sensitive isolates did not harbour the A2143G variant and were not successfully induced by exposure to the antibiotic.
Table 1
Twenty randomly selected clarithromycin-sensitive H. pylori clinical isolates and one standard strain were inducted by exposure to clarithromycin
No
Isolate
CLR sensitivity/resistance
CLR MIC (µg/mL)
23S rRNA (A2142G/C)
23S rRNA (A2143G)
Induction success
1
NCTC 11637
S
0.032
No
No
No
2
UM025B
S
0.032
No
No
No
3
UM067B SW
S
0.125
No
No
No
4
UM113B
S
0.125
No
No
No
5
UM113B C5
S
0.032
No
No
No
6
UM114
S
0.032
No
No
No
7
UM171
S
0.064
No
Yes
Yes
8
UM276 R (71)
S
0.032
No
No
No
9
UM276 S (53)
S
0.032
No
No
No
10
UM303A C6
S
0.032
No
No
No
11
UM408 SW
S
0.032
No
No
No
12
UM411 SW
S
0.032
No
No
No
13
UM443A (33)
S
0.032
No
No
No
14
UM443A SW
S
0.032
No
No
No
15
UM622A1a
S
0.032
No
No
No
16
UM626A1
S
0.064
No
Yes
Yes
17
UM650B
S
< 0.016
No
Yes
Yes
18
UM663B2
S
0.064
No
No
No
19
UM678A
S
0.064
No
Yes
Yes
20
UM692B2
S
0.032
No
No
No
21
UM113A SW
S
0.032
No
No
No
22
H. pylori 26695
S
0.016
No
No
Not Done
23
UM032
S
0.016
No
No
Not Done
24
UM202
R
No
No
Not Done
25
UM233
S
No
No
Not Done
26
UM370
R
No
Yes
Not Done
aThe isolate number consist of: UM stands for Unversiti Malaya, followed by isolate number, letter designations (A: Antrum, B: Body, C: Cardia, SW: Sweep), and the number of clonal isolates
Table 2
MICs of H. pylori isolates before, and after in vitro induction with clarithromycin
No
Isolate
MIC in µg/mL (passage number)
S isolates before induction
B isolates at breakpoint
R isolates after breakpointa
R isolates after resistance inductionb
1
UM171
0.064 (P0)
0.125 (P4)
1 (P5)
> 64 (P10)
2
UM626A1
0.064 (P0)
0.125 (P5)
2 (P6)
> 64 (P11)
3
UM650B
< 0.016 (P0)
0.125 (P7)
1 (P8)
> 64 (P12)
4
UM678A
0.064 (P0)
0.125 (P5)
1 (P6)
> 64 (P10)
aPassage number when isolate developed resistance for the first time
bPassage number when there is no inhibition zone using MIC Test Strip

Stability of resistance

All the four R isogenic isolates maintained their MICs of > 64 µg/mL against clarithromycin after 10 successful rounds of growth on non-selective CA plate, and after storage frozen at − 80 °C for 3–5 months in a BHI broth with 20% glycerol. The MICs were confirmed using MIC Test Strip (Fig. 1). These induced isolates probably underwent stable genetic reprogramming that contributed to the persistence of antibiotic resistance even in the absence of selective pressure.

RAPD genotypes

RAPD PCR was performed to verify the identity of S, B, R isolates. Based on Fig. 2, the RAPD analysis of the four parental S isolates were not related, each isolate showed a distinctive pattern of bands. In addition, the B and R isogenic isolates were identical in genotype to their respective parental S isolates.

23SrRNA genotypes and resistance to clarithromycin

The S, B, and R isolates were tested for clarithromycin susceptibility, and the results showed that all S and B isolates were susceptible while all the induced R isolates were resistant to clarithromycin. Interestingly, despite been susceptible to clarithromycin, all four S parental isolates had A2143G variation of 23S rRNA (Fig. 3). There were no variations in positions 2142 and 2143 of the gene between the parental isolates and the induced isogenic isolates. The 23S rRNA Sanger sequencing datasets supporting the conclusions of this article are available in the NCBI’s GenBank repository under the accession OR357686-OR357715.

Quality of H. pylori genomes

After trimming the low-quality reads, there were 19 to 36 million cleaned reads. Cleaned reads from S, B, and R samples were directly mapped to the UM032 reference H. pylori genome. H. pylori UM032 was chosen as the reference genome for mapping because UM032 genome was fully sequenced, extensively studied, and was derived from H. pylori isolated in the same human population. Sequencing coverage and average depth ranged from 95.08 to 95.47% and from 1152.47 to 2151.57, respectively. Therefore, the efficient reads were sufficient for SNV/InDel analysis. SNV/InDel analysis of the B and R isolates was carried out with reference to the corresponding S isolates. The number of identified SNVs ranged from 5300 to 35,742 and InDels ranged from 646 to 2679. The WGS sequencing datasets supporting the conclusions of this article are available in the NCBI’s Sequence Read Archive (SRA) database repository under the accession ID PRJNA999133, unique persistent identifier and hyperlink to datasets in https://​www.​ncbi.​nlm.​nih.​gov/​sra/​PRJNA999133.

Clarithromycin resistance-associated variants

SNV and InDel

The SNV and InDel analysis of the corresponding S isolates was used to compare the SNV and InDel of the B and R isolates. A total of 67,688 SNV mutations were detected among the B isolates and 65,476 SNV mutations were detected among the R isolates. Moreover, 5442 InDel mutations detected among the B isolates, and 5244 InDel mutations detected among the R isolates (Table 3).
Table 3
Classification of SNVs and InDels
Isolate
Group
Isolate ID
SNV
InDel
Non-synonymousa
Synonymousa
Stop gain
Stop loss
Non-frameshift deletion
Frameshift deletion
Non-frameshift insertion
Frameshift insertion
UM171
B
UM171-B
7006
15,900
71
18
133
420
115
409
R
UM171-R
6764
15,358
70
19
136
411
112
395
UM626A1
B
UM626A1-B
9101
21,425
83
27
169
510
135
485
R
UM626A1-R
8635
20,422
78
27
148
465
139
443
UM650B
B
UM650B-B
1432
2978
20
3
52
136
38
139
R
UM650B-R
1504
3108
18
3
53
136
40
143
The corresponding S isolates were used as reference
aThose numbers were detected after mapping against reference strains (UM32) for the purpose of annotation
Among of all identified variations in B and R, 119,432 were in the coding region, 1888 were located in the intergenic region, 4995 were located in the upstream, 7827 were located in the downstream, 9702 were located in the upstream; downstream, and 6 were located in the UTR5 (Table 4).
Table 4
Distribution of SNV/InDel on genome
Isolate
Group
Isolate ID
Categories
Exonic
Intergenic
Upstream
Downstream
Upstream; downstream
UTR5a
UM171
B
UM171-B
24,072
320
953
1536
1881
1
R
UM171-R
23,265
327
938
1479
1898
0
UM626A1
B
UM626A1-B
31,935
559
1352
2032
2541
2
R
UM626A1-R
30,357
533
1375
1998
2454
2
UM650B
B
UM650B-B
4798
61
202
440
445
0
R
UM650B-R
5005
88
175
342
483
1
The corresponding S isolates were used as reference
aUTR5: untranslated reg

Genes associated with virulence and antibiotic resistance

Genes with the highest rate of mutation in response to clarithromycin may have a higher likelihood of been associated directly (causation) or indirectly (compensation) with clarithromycin resistance. An additional file shows this in more detail of the genes with high number of mutations (see Additional file 1). Interestingly, mutations above threshold of > 1 were detected in genes that play a role in virulence and survival (cag4, rsmH, gene encoding sel1 repeat family protein, and gene encoding vacuolating cytotoxin domain-containing protein) in the induced isolates (B and R) against their corresponding S isolates (Table 5). Additionally, the specific mutations of these genes were also noted which may associate with the development of antibiotic resistance in H. pylori in response to clarithromycin (Table 6).
Table 5
Rate of mutations of genes associated with clarithromycin resistance development in the UM171, UM626A1, and UM650B
Locus tag
Protein ID
Protein name
Isolate
Rate of mutation
Threshold
Probability of mutation occurrence
K747_RS06340
WP_015644956.1
Sel1 repeat family protein
UM171-B
0.1054
0.0180
5.8387
UM171-R
0.0948
0.0175
5.4160
UM626A1-B
0.0948
0.0241
3.9339
UM626A1-R
0.0948
0.0230
4.1162
UM650B-B
0.0141
0.0037
3.7658
UM650B-R
0.0070
0.0038
1.8372
K747_RS06360
WP_015644959.1
Transglycosylase SLT domain-containing protein (Cag4)
UM171-B
0.0180
0.0180
6.4219
UM171-R
0.0175
0.0175
5.3848
UM626A1-B
0.0241
0.0241
3.8298
UM626A1-R
0.0230
0.0230
4.0073
UM650B-B
0.0037
0.0037
16.3223
UM650B-R
0.0038
0.0038
15.9259
K747_RS06805
WP_015645027.1
Vacuolating cytotoxin domain-containing protein
UM171-B
0.0568
0.0180
3.1475
UM171-R
0.0509
0.0175
2.9048
UM626A1-B
0.0586
0.0241
2.4298
UM626A1-R
0.0591
0.0230
2.5651
UM650B-B
0.0182
0.0037
4.8890
UM650B-R
0.0215
0.0038
5.6153
K747_RS07265
WP_015645103.1
16S rRNA (cytosine(1402)-N(4))-methyltransferase (RsmH)
UM171-B
0.0335
0.0180
1.8547
UM171-R
0.0432
0.0175
2.4666
UM626A1-B
0.0443
0.0241
1.8364
UM626A1-R
0.0292
0.0230
1.2654
UM650B-B
0.0076
0.0037
2.0259
UM650B-R
0.0108
0.0038
2.8239
Table 6
Specific mutations associated with clarithromycin resistance development in UM171, UM626A1, and UM650B
Type
Locus tag (gene)/protein
Position
Type of variation
Bases change (amino acid change)
UM171
UM626A1
UM650B
S
B
R
S
B
R
S
B
R
SNV
K747_RS06360 (cag4)/transglycosylase SLT domain-containing protein
NC_021215.3:1292240
Non-synonymous, exonic
G (R)
G>A (R47K)
G>A (R47K)
G (R)
G>A (R47K)
G>A (R47K)
G (R)
G>A (R47K)
G>A (R47K)
NC_021215.3:1292491
Non-synonymous, exonic
G (V)
G>A (V131M)
G>A (V131M)
G (V)
G>A (V131M)
G>A (V131M)
G (V)
G>A (V131M)
G>A (V131M)
NC_021215.3:1292500
Non-synonymous, exonic
A (I)
A>T (I134L)
A>T (I134L)
A (I)
A>T (I134L)
A>T (I134L)
A (I)
A>T (I134L)
A>T (I134L)
K747_RS07265(rsmH)/16S rRNA (cytosine(1402)-N(4))-methyltransferase
NC_021215.3:1499504
Non-synonymous, exonic
T (S)
T (S)
T>G (S244A)
T (S)
T>G (S244A)
T>G (S244A)
T (S)
T (S)
T>G (S244A)
NC_021215.3:1499056
Non-synonymous, exonic
T (N)
T (N)
T>A (N94K)
T (N)
T>A (N94K)
T>A (N94K)
T (N)
T>A (N94K)
T>A (N94K)
NC_021215.3:1499175
Non-synonymous, exonic
A (D)
A (D)
A>G (D134G)
A (D)
A>G (D134G)
A>G (D134G)
A (D)
A>G (D134G)
A>G (D134G)
NC_021215.3:1499345
Non-synonymous, exonic
C (L)
C (L)
C>T (L191F)
C (L)
C>T (L191F)
C>T (L191F)
C (L)
C>T (L191F)
C>T (L191F)
K747_RS06340/Sel1 repeat family protein
NC_021215.3:1289818
Non-synonymous, exonic
C (D)
C>T (D39N)
C>T (D39N)
C (D)
C>T (D39N)
C>T (D39N)
C (D)
C>T (D39N)
C>T (D39N)
InDel
K747_RS06360 (cag4)/transglycosylase SLT domain-containing protein
NC_021215.3:1292229_1292232
Frameshift deletion, exonic
AGTG (E)
– (E43fs)
AGTG (E)
AGTG (E)
– (E43fs)
– (E43fs)
AGTG (E)
– (E43fs)
– (E43fs)
NC_021215.3:1292233_1292234
Frameshift insertion, exonic
– (V)
ATAA (V45fs)
– (V)
– (V)
ATAA (V45fs)
ATAA (V45fs)
– (V)
ATAA (V45fs)
ATAA (V45fs)
K747_RS06805/vacuolating cytotoxin domain-containing protein
NC_021215.3:1393215_1393219
Frameshift deletion, exonic
ACTAG (L)
– (L2430fs)
ACTAG (L)
ACTAG (L)
– (L2430fs)
– (L2430fs)
ACTAG (L)
– (L2430fs)
– (L2430fs)
NC_021215.3:1393222_1393223
Frameshift insertion, exonic
– (S)
CAAAC (S2433fs)
– (S)
– (S)
CAAAC (S2433fs)
CAAAC (S2433fs)
– (S)
CAAAC (S2433fs)
CAAAC (S2433fs)

Growth curves

The growth rate of the H. pylori S, B, and R isolates were comparable within the initial 1 day of growth. However, S and B isolates showed an increase in growth compared to R isolate from 2 to 3 days of growth (Fig. 4). The OD600 nm of S, B, and R were found to be associated well with the bacterial viable count.

Biofilm assessment

The multicellular survival tactic of biofilm formation, which occurs at the population level, indirectly improves the fitness of bacteria for overall survival. In this study, the R isolates produced significantly (p-value < 0.05, > twofold changes) more biofilm compared to S. Meanwhile, B isolates produced more biofilm compared to S isolates (Table 7). The average biofilm development of H. pylori isolates on day 7 was then divided against their corresponding growth level of day 7. The results showed that the R isolates produced significantly (p-value < 0.05) more biofilm compared to S. Meanwhile, B isolates produced more biofilm compared to S isolates but was only statistically significant (p-value < 0.05) for UM171 (Fig. 5).
Table 7
Average H. pylori biofilm formed on day 7
Isolate
Group
Sample ID
Average OD595 nm reading
Fold difference
B vs S (p-value)
R vs S (p-value)
R vs B (p-value)
UM171
S
UM171-S
0.14
2.07 (< 0.05)
2.36 (< 0.05)
1.14 (≥ 0.05)
B
UM171-B
0.29
R
UM171-R
0.33
UM626A1
S
UM626A1-S
0.22
1.39 (< 0.05)
2.03 (< 0.05)
1.45 (< 0.05)
B
UM626A1-B
0.36
R
UM626A1-R
0.49
UM650B
S
UM650B-S
0.21
1.86 (< 0.05)
2.19 (< 0.05)
1.18 (< 0.05)
B
UM650B-B
0.40
R
UM650B-R
0.47
UM678A
S
UM678A-S
0.19
1.95 (< 0.05)
2.63 (< 0.05)
1.35 (< 0.05)
B
UM678A-B
0.38
R
UM678A-R
0.52

Identification of genes associated with biofilm formation

The presence of annotated genes in the H. pylori genomic sequences of B and R mutants was shown to be substantially linked with the capacity to build biofilm (Table 8). It is interesting to note that bacteria in biofilms frequently display mutator phenotypes and phenotypic variety, indicating that genetic instability and mutation are key components of biofilm formation. Interestingly, mutations were found in the B and R mutants. These included the genes hypE, hypF, and gene encoding cag pathogenicity island (Table 8).
Table 8
Mutations associated with biofilm formation in UM171, UM626A1, and UM650B
Type
Locus tag (gene)/protein
Position
Type of variation
Bases change (amino acid change)
UM171
UM626A1
UM650B
S
B
R
S
B
R
S
B
R
SNV
K747_RS04025(hypE)/hydrogenase expression/formation protein HypE
NC_021215.3:820809
Non-synonymous, exonic
C (G)
C>G (G145A)
C>G (G145A)
C (G)
C>G (G145A)
C>G (G145A)
C (G)
C>G (G145A)
C>G (G145A)
NC_021215.3:821112
Non-synonymous, exonic
T (H)
T>C (H44R)
T>C (H44R)
T (H)
T>C (H44R)
T>C (H44R)
T (H)
T>C (H44R)
T>C (H44R)
NC_021215.3:821113
Non-synonymous, exonic
G (H)
G>T (H44N)
G>T (H44N)
G (H)
G>T (H44N)
G>T (H44N)
G (H)
G>T (H44N)
G>T (H44N)
K747_RS04030(hypF)/carbamoyltransferase HypF
NC_021215.3:822687
Non-synonymous, exonic
C (A)
C>T (A271T)
C>T (A271T)
C (A)
C T (A271T)
C>T (A271T)
C (A)
C>T (A271T)
C>T (A271T)
NC_021215.3:822837
Non-synonymous, exonic
C (D)
C>T (D221N)
C>T (D221N)
C (D)
C>T (D221N)
C>T (D221N)
C (D)
C>T (D221N)
C>T (D221N)
K747_RS06380/Cag pathogenicity island protein
NC_021215.3:1300640
Non-synonymous, exonic
T (K)
T>G (K608Q)
T>G (K608Q)
T (K)
T>G (K608Q)
T>G (K608Q)
T (K)
T>G (K608Q)
T>G (K608Q)
NC_021215.3:1300687
Non-synonymous, exonic
C (S)
C>T (S592N)
C>T (S592N)
C (S)
C>T (S592N)
C>T (S592N)
C (S)
C>T (S592N)
C>T (S592N)

Discussions

The global resistance rate of clarithromycin increased significantly from 24.28% in 2010–2017 to 32.14% in 2018–2021 with Switzerland, Portugal, and Israel having the highest resistance rate [25]. Similarly, clarithromycin resistance in Malaysia is also increasing from 6.8% between July 2011 and August 2012 [26] to 35.6% between April 2014 and August 2015 [27]. Thus, constantly monitoring clarithromycin-resistant rates of H. pylori is crucial for making informed decision of the most appropriate eradication therapies with good clinical outcomes. In this study, the rate of clarithromycin resistance was estimated to be 29.9%. However, this must be interpreted with caution as different resistance breakpoint and testing methods were used by different researchers in this field. Notably, Hanafiah et al. [27] and our data did not distinguish primary and secondary resistance cases while the earlier study [26] had excluded all known cases of treatment failure.
Helicobacter pylori clarithromycin resistance is mostly caused by point mutations (A2142G/C and A2143G) in the 23S rRNA gene’s peptidyl transferase loop region [28]. This mutation in 23S rRNA gene (A2143G) has also been observed in clarithromycin-sensitive and clarithromycin-resistant H. pylori strains by other researchers. Zhang et al. [13] has detected the A2143G mutation in 45.5% (5/11) clarithromycin-sensitive strains. Moreover, A2143G mutation was also found in two out of six clarithromycin-sensitive H. pylori strains in China [6]. Similarly, sensitive strains with A2143G mutation were also reported previously among H. pylori stains in Mexico [14]. In this study, sequencing of clarithromycin-sensitive H. pylori detected mutation A2143G mutations of the 23S rRNA in four out of 26 isolates. Interestingly, these four isolates, which were phenotypically clarithromycin-sensitive but harbour the A2143G, were successfully induced to clarithromycin resistant by exposure to the antibiotic in vitro. On the other hand, none of the other clarithromycin-sensitive isolates could be induced. Additionally, sensitive strains could carry silent antimicrobial resistance genes, often known as cryptic genes. Bacteria may carry these silent genes on their chromosomal DNA or plasmids but do not show the appropriate phenotypic antibiotic resistance [29, 30]. The majority of strains with silent genes are clinical strains [31]. Several Gram-negative bacteria have been reported to carry cryptic genes [3237]. In each of these cases, the genes’ promoter and resistance gene sequences were intact, indicating that the process of silencing is not well understood. This implies that under some circumstances, it is possible for genes to spread silently throughout bacterial populations. Such silent genes “off status” may express phenotypic resistance “on status” when they are subjected to selective pressure, such as pressure from antibiotics. It was previously observed by Stasiak et al. [36] that antibiotic pressure can cause the activation of silent antimicrobial genes. Moreover, H. pylori possesses regulatory genes that regulate the expression of various antibiotic resistance genes [38]. Therefore, investigating the gene expression linked to antibiotic resistance can reveal insights into the mechanisms H. pylori employ to survive antibiotic treatment. Therefore, we hypothesized that these clarithromycin sensitive strains with A2143G mutation in the present study were potentially resistant to clarithromycin and resistance were “switched on” when exposed to the antibiotic.
On the other hand, the result of WGS showed changes in genes associated with virulence and antibiotic resistance and may influence in the development of clarithromycin resistance among H. pylori strains. Although four pairs of induced resistant isolates were sequenced, based on the quality control (QC) data, UM678A was excluded from genomic analysis. In the B and R strains, mutations in the cag4 and gene encoding vacuolating cytotoxin domain-containing protein were detected, which have been shown to contribute to virulence in H. pylori and promote bacterial survival [39, 40]. Cag4, which is also known as Cagγ (hp0523), refers to one of the proteins encoded by the cag pathogenicity island (cagPAI) [41]. The cagPAI is a genomic region found in H. pylori that has been linked to enhanced virulence. The cagPAI gene encodes a type IV secretion system (T4SS), a complex molecular system that allows the bacterium to directly inject bacterial proteins into host cells [42]. CagA (Cytotoxin-associated gene A) protein, Cag4/Cagγ, and other proteins expressed by numerous genes within this region are among the proteins encoded by cagPAI [43]. The putative peptidoglycan hydrolase Cag4/Cag protein is part of the T4SS and is essential for CagA protein secretion and delivery into host gastric epithelial cells [44]. Once inside the host cells, the CagA protein can disrupt a variety of biological processes, altering host cell signalling and causing inflammation which contributes to the development of gastritis, peptic ulcers, and potentially gastric cancer [45]. The vacuolating cytotoxin domain-containing protein, which is also known as FaaA protein from a representative H. pylori strain (J99), has been found to improve H. pylori colonisation capacity in animal models, and transcription of each gene is elevated in the gastric environment relative to the level of transcription during bacterial growth in vitro [46]. The VacA-like proteins of H. pylori are found on the bacterial surface, while the FaaA protein is found on the flagella [47]. A study found that the faaA mutant mislocalized the flagella and reduced bacterial mobility [48]. Additionally, SNV mutations was also noted in gene encoding Sel1 repeat family protein in the B and R isolates which is involved in signal transduction pathways between eukaryotes and bacteria [49]. The interactions between bacterial and eukaryotic host cells are thought to be mediated by bacterial Sel1-like repeat (SLR) [50]. Five of the nine secreted proteins from H. pylori (HcpAD, HcpA, HcpE, HcpB, and HcpC) folds into a stable three-dimensional structure composed of six disulfides bridged SLRs [51]. These proteins are known to trigger an immune response, causing inflammation [52]. Likewise, Newton et al. [53] has noted that Sel1 repeat protein as the virulence determinant of Legionella pneumophila which influences vacuolar trafficking. Furthermore, SNV mutations were found in rsmH genes among the B and R isolates, which have been linked to antibiotic resistance. Interestingly, mutations in the 16S RNA methyltransferase family (which includes the rsmH gene) have been shown to confer aminoglycoside resistance in aerobic Gram-negative bacteria [54, 55]. Helicobacter heilmannii isolates with high MIC against neomycin have been shown to have a SNV in the ribosomal RNA small subunit methyltransferase H (RsmH) gene [56].
The association between enhanced virulence and resistance development in bacteria is a complex and multifaceted topic. While they are distinct characteristics, there are scenarios where enhanced virulence and resistance development may be interconnected or even co-selected under certain circumstances [57]. Acquiring antibiotic resistance in bacteria may be advantageous for their survival and enhance their virulence [58]. Therefore, H. pylori may simultaneously enhance its virulence through exposure to clarithromycin [59]. To combat the emergence and spread of both virulence and resistance in bacteria, it is critical to promote responsible antibiotic use, implement infection prevention measures, monitor resistance patterns, and conduct additional research to understand the underlying mechanisms and interactions between these two traits [60]. Therefore, mutations that occur in R isolates compared to S isolates suggests that mutations are probably involved in antibiotic resistance. However, if mutations that occur in B isolate compared to S isolate, it may not be directly linked to resistance, but it may condition the organism to develop antibiotic resistance.
The development of antibiotic resistance is closely associated with the formation of biofilms in bacterial populations. The biofilm matrix provides protection and shelter to the bacteria within, making them highly resistant to the effects of antibiotics [61]. The continuous presence of increasing concentrations of antibiotics within biofilms can lead to adaptive resistance [62]. The biofilm mass of H. pylori may be seen after 3 days of in vitro incubation [63, 64] and can last for up to 7 days under different culture conditions [24, 6567]. However, some of our samples took longer time to form biofilm and we were unable to see any visible biofilm within 3 days; as a result, we left them for 7 days. The results of this study showed that H. pylori produced more biofilm as they developed resistance against clarithromycin. Moreover, bacteria in the biofilm may undergo genetic changes to become more resistant to the specific antibiotics present [68]. It is interesting to note that both B and R isolates have SNV mutations in several genes (hypE, hypF, and cag pathogenicity island) associated to the development of biofilms. Hydrogenase activity in H. pylori is mediated by hypE and hypF, both of which have been shown to contribute to biofilm formation [63, 69]. A cag pathogenicity island protein is one of the proteins that have been identified as being frequently present with strains that form good biofilms. It has been determined that the CagA protein, which is encoded by the Cag pathogenicity islands, is induced in H. pylori biofilms [61]. The Cag pathogenicity island may have a substantial impact on the establishment of the H. pylori biofilm. CagA and the cag pathogenicity island may be implicated in the production of H. pylori biofilms through their influence on bacteria-bacteria interactions, in addition to their function in bacteria-host interactions [61, 63, 70]. It is important to note that bacterial infections generated by biofilms are frequently more difficult to treat than infections caused by planktonic (free-floating) bacteria [71]. Therefore, researchers are looking at several strategies to combat biofilm-associated antibiotic resistance, including the development of new antimicrobial agents, the use of combination therapy, and the development of biofilm-disrupting techniques. For better treatment outcomes and to solve the problem of worldwide antibiotic resistance, it is essential to comprehend the characteristics of biofilms and how they contribute to antibiotic resistance [72].

Conclusions

In conclusion, the clarithromycin-sensitive H. pylori isolates with the A2143G mutation were successfully induced to be resistant and numerous genes were subjected to genetic reprogramming in response to increasing concentration of clarithromycin. Furthermore, antibiotic exposure may reprogram certain genes such as genes encoding Cag4/Cagγ protein, vacuolating cytotoxin domain-containing protein, sel1 repeat family protein, and rsmh gene which could possibly increase the likelihood of antibiotic resistance development and enhances virulence factor in H. pylori. Therefore, further studies are required to elucidate these genes mechanisms in antibiotic resistance in H. pylori which will help in improving H. pylori eradication and develop a new treatment for H. pylori infection.

Acknowledgements

The authors would like to express their utmost gratitude and appreciation to the late Emeritus Professor Dato' Dr. Khean Lee Goh for his help and support in setting up the Helicobacter Research Laboratory (UM Marshall Centre) at the Universiti Malaya and establishing the centre’s H. pylori culture collection.

Declarations

Not applicable.
Not applicable.

Competing interests

Not applicable.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
2.
Zurück zum Zitat Graham DY. Helicobacter pylori update: gastric cancer, reliable therapy, and possible benefits. Gastroenterology. 2015;148:719-31.e3.PubMedCrossRef Graham DY. Helicobacter pylori update: gastric cancer, reliable therapy, and possible benefits. Gastroenterology. 2015;148:719-31.e3.PubMedCrossRef
4.
Zurück zum Zitat Thung I, Aramin H, Vavinskaya V, Gupta S, Park JY, Crowe SE, et al. Review Article: the global emergence of Helicobacter pylori antibiotic resistance. AP&T. 2016;43:514–33. Thung I, Aramin H, Vavinskaya V, Gupta S, Park JY, Crowe SE, et al. Review Article: the global emergence of Helicobacter pylori antibiotic resistance. AP&T. 2016;43:514–33.
5.
Zurück zum Zitat Puah SM, Goh KL, Ng HK, Chua KH. Current status of Helicobacter pylori resistance to clarithromycin and levofloxacin in Malaysia—findings from a molecular based study. PeerJ. 2021;9: e11518.PubMedPubMedCentralCrossRef Puah SM, Goh KL, Ng HK, Chua KH. Current status of Helicobacter pylori resistance to clarithromycin and levofloxacin in Malaysia—findings from a molecular based study. PeerJ. 2021;9: e11518.PubMedPubMedCentralCrossRef
6.
Zurück zum Zitat Chen J, Ye L, Jin L, Xu X, Xu P, Wang X, et al. Application of next-generation sequencing to characterize novel mutations on clarithromycin-susceptible Helicobacter pylori strains with A2143G Of 23s rRNA gene. Ann Clin Microbiol. 2018;17:10. Chen J, Ye L, Jin L, Xu X, Xu P, Wang X, et al. Application of next-generation sequencing to characterize novel mutations on clarithromycin-susceptible Helicobacter pylori strains with A2143G Of 23s rRNA gene. Ann Clin Microbiol. 2018;17:10.
7.
Zurück zum Zitat Kocsmár É, Buzás GM, Szirtes I, Kocsmár I, Kramer Z, Szijártó A, et al. Primary and secondary clarithromycin resistance in Helicobacter pylori and mathematical modeling of the role of macrolides. Nat Commun. 2021;12:2255.PubMedPubMedCentralCrossRef Kocsmár É, Buzás GM, Szirtes I, Kocsmár I, Kramer Z, Szijártó A, et al. Primary and secondary clarithromycin resistance in Helicobacter pylori and mathematical modeling of the role of macrolides. Nat Commun. 2021;12:2255.PubMedPubMedCentralCrossRef
8.
Zurück zum Zitat Redondo JJ, Keller PM, Zbinden R, Wagner K. A novel RT-PCR for the detection of Helicobacter pylori and identification of clarithromycin resistance mediated by mutations in the 23s rRNA gene. Diagn Microbiol Infect Dis. 2018;90:1–6.PubMedCrossRef Redondo JJ, Keller PM, Zbinden R, Wagner K. A novel RT-PCR for the detection of Helicobacter pylori and identification of clarithromycin resistance mediated by mutations in the 23s rRNA gene. Diagn Microbiol Infect Dis. 2018;90:1–6.PubMedCrossRef
9.
Zurück zum Zitat Albasha AM, Elnosh MM, Osman EH, Zeinalabdin DM, Fadl AAM, Ali MA, et al. Helicobacter pylori 23S rRNA gene A2142G, A2143G, T2182C, and C2195T mutations associated with clarithromycin resistance detected in Sudanese patients. BMC Microbiol. 2021;21:38.PubMedPubMedCentralCrossRef Albasha AM, Elnosh MM, Osman EH, Zeinalabdin DM, Fadl AAM, Ali MA, et al. Helicobacter pylori 23S rRNA gene A2142G, A2143G, T2182C, and C2195T mutations associated with clarithromycin resistance detected in Sudanese patients. BMC Microbiol. 2021;21:38.PubMedPubMedCentralCrossRef
10.
Zurück zum Zitat Bińkowska A, Biernat MM, Łaczmański Ł, Gościniak G. Molecular patterns of resistance among Helicobacter pylori strains in South-Western Poland. Front Microbiol. 2018;9:3154.PubMedPubMedCentralCrossRef Bińkowska A, Biernat MM, Łaczmański Ł, Gościniak G. Molecular patterns of resistance among Helicobacter pylori strains in South-Western Poland. Front Microbiol. 2018;9:3154.PubMedPubMedCentralCrossRef
11.
Zurück zum Zitat Çağan-Appak Y, Gazi H, Ayhan S, Cengiz-Özyurt B, Kurutepe S, Kasırga E. Clarithromycin resistance and 23s rRNA gene point mutations of Helicobacter pylori infection in children. Turk J Pediatr. 2016;58:371–6.PubMedCrossRef Çağan-Appak Y, Gazi H, Ayhan S, Cengiz-Özyurt B, Kurutepe S, Kasırga E. Clarithromycin resistance and 23s rRNA gene point mutations of Helicobacter pylori infection in children. Turk J Pediatr. 2016;58:371–6.PubMedCrossRef
12.
Zurück zum Zitat Matta AJ, Zambrano DC, Pazos AJ. Punctual mutations in 23S rRNA gene of clarithromycin-resistant Helicobacter pylori in Colombian populations. World J Gastroenterol. 2018;24:1531–9.PubMedPubMedCentralCrossRef Matta AJ, Zambrano DC, Pazos AJ. Punctual mutations in 23S rRNA gene of clarithromycin-resistant Helicobacter pylori in Colombian populations. World J Gastroenterol. 2018;24:1531–9.PubMedPubMedCentralCrossRef
13.
Zurück zum Zitat Zhang Y, Wen Y, Xiao Q, Zheng W, Long G, Chen B, et al. Mutations in the antibiotic target genes related to clarithromycin, metronidazole and levofloxacin resistance in Helicobacter pylori strains from children in China. Infect Drug Resist. 2020;13:311–22.PubMedPubMedCentralCrossRef Zhang Y, Wen Y, Xiao Q, Zheng W, Long G, Chen B, et al. Mutations in the antibiotic target genes related to clarithromycin, metronidazole and levofloxacin resistance in Helicobacter pylori strains from children in China. Infect Drug Resist. 2020;13:311–22.PubMedPubMedCentralCrossRef
14.
Zurück zum Zitat Camorlinga-Ponce M, Gómez-Delgado A, Aguilar-Zamora E, Torres RC, Giono-Cerezo S, Escobar-Ogaz A, et al. Phenotypic and genotypic antibiotic resistance patterns in Helicobacter pylori strains from ethnically diverse population in México. Front Cell Infect Microbiol. 2021;10: 539115.PubMedPubMedCentralCrossRef Camorlinga-Ponce M, Gómez-Delgado A, Aguilar-Zamora E, Torres RC, Giono-Cerezo S, Escobar-Ogaz A, et al. Phenotypic and genotypic antibiotic resistance patterns in Helicobacter pylori strains from ethnically diverse population in México. Front Cell Infect Microbiol. 2021;10: 539115.PubMedPubMedCentralCrossRef
15.
Zurück zum Zitat Hashemi SJ, Sheikh AF, Goodarzi H, Yadyad MJ, Seyedian SS, Aslani S, et al. Genetic basis for metronidazole and clarithromycin resistance in Helicobacter pylori strains isolated from patients with gastroduodenal disorders. Infect Drug Resist. 2019;12:535–43.PubMedPubMedCentralCrossRef Hashemi SJ, Sheikh AF, Goodarzi H, Yadyad MJ, Seyedian SS, Aslani S, et al. Genetic basis for metronidazole and clarithromycin resistance in Helicobacter pylori strains isolated from patients with gastroduodenal disorders. Infect Drug Resist. 2019;12:535–43.PubMedPubMedCentralCrossRef
16.
Zurück zum Zitat Li XH, Huang YY, Lu LM, Zhao LJ, Luo XK, Li RJ, et al. Early genetic diagnosis of clarithromycin resistance in Helicobacter pylori. World J Gastroenterol. 2021;27:3595–608.PubMedPubMedCentralCrossRef Li XH, Huang YY, Lu LM, Zhao LJ, Luo XK, Li RJ, et al. Early genetic diagnosis of clarithromycin resistance in Helicobacter pylori. World J Gastroenterol. 2021;27:3595–608.PubMedPubMedCentralCrossRef
17.
Zurück zum Zitat Sharma D, Misba L, Khan AU. Antibiotics versus biofilm: an emerging battleground in microbial communities. Antimicrob Resist Infect Control. 2019;8:76.PubMedPubMedCentralCrossRef Sharma D, Misba L, Khan AU. Antibiotics versus biofilm: an emerging battleground in microbial communities. Antimicrob Resist Infect Control. 2019;8:76.PubMedPubMedCentralCrossRef
18.
Zurück zum Zitat Abebe GM. The role of bacterial biofilm in antibiotic resistance and food contamination. Int J Med Microbiol. 2020;2020:1705814. Abebe GM. The role of bacterial biofilm in antibiotic resistance and food contamination. Int J Med Microbiol. 2020;2020:1705814.
19.
Zurück zum Zitat Chong SK, Lou Q, Fitzgerald JF, Lee CH. Evaluation of 16S rRNA gene PCR with primers Hp1 and Hp2 for detection of Helicobacter pylori. J Clin Microbiol. 1996;34:2728–30.PubMedPubMedCentralCrossRef Chong SK, Lou Q, Fitzgerald JF, Lee CH. Evaluation of 16S rRNA gene PCR with primers Hp1 and Hp2 for detection of Helicobacter pylori. J Clin Microbiol. 1996;34:2728–30.PubMedPubMedCentralCrossRef
20.
Zurück zum Zitat Yan WH, Chen J, Hu HJ, Yu JD, Huang XL, Li ZY. Preliminary study on in-vitro induction of antibiotic resistance in Helicobacter pylori strains isolated from children. Zhonghua Er Ke Za Zhi. 2007;45:708–11.PubMed Yan WH, Chen J, Hu HJ, Yu JD, Huang XL, Li ZY. Preliminary study on in-vitro induction of antibiotic resistance in Helicobacter pylori strains isolated from children. Zhonghua Er Ke Za Zhi. 2007;45:708–11.PubMed
21.
Zurück zum Zitat Akopyanz N, Bukanov NO, Westblom TU, Kresovich S, Berg DE. DNA diversity among clinical isolates of Helicobacter pylori detected by PCR-based RAPD fingerprinting. Nucleic Acids Res. 1992;20:5137–42.PubMedPubMedCentralCrossRef Akopyanz N, Bukanov NO, Westblom TU, Kresovich S, Berg DE. DNA diversity among clinical isolates of Helicobacter pylori detected by PCR-based RAPD fingerprinting. Nucleic Acids Res. 1992;20:5137–42.PubMedPubMedCentralCrossRef
22.
Zurück zum Zitat Ho S-L, Tan EL, Sam CK, Goh K-L. Clarithromycin resistance and point mutations in the 23S rRNA gene in Helicobacter pylori isolates from Malaysia. J Dig Dis. 2010;11:101–5.PubMedCrossRef Ho S-L, Tan EL, Sam CK, Goh K-L. Clarithromycin resistance and point mutations in the 23S rRNA gene in Helicobacter pylori isolates from Malaysia. J Dig Dis. 2010;11:101–5.PubMedCrossRef
23.
Zurück zum Zitat Al-Maleki AR, Loke MF, Lui SY, Ramli NSK, Khosravi Y, Ng CG, et al. Helicobacter pylori outer inflammatory protein A (OipA) suppresses apoptosis of AGS gastric cells in vitro. Cell Microbiol. 2017;19: e12771.CrossRef Al-Maleki AR, Loke MF, Lui SY, Ramli NSK, Khosravi Y, Ng CG, et al. Helicobacter pylori outer inflammatory protein A (OipA) suppresses apoptosis of AGS gastric cells in vitro. Cell Microbiol. 2017;19: e12771.CrossRef
24.
Zurück zum Zitat Wong EH, Ng CG, Chua EG, Tay AC, Peters F, Marshall BJ, et al. Comparative genomics revealed multiple Helicobacter pylori genes associated with biofilm formation in vitro. PLoS ONE. 2016;11: e0166835.PubMedPubMedCentralCrossRef Wong EH, Ng CG, Chua EG, Tay AC, Peters F, Marshall BJ, et al. Comparative genomics revealed multiple Helicobacter pylori genes associated with biofilm formation in vitro. PLoS ONE. 2016;11: e0166835.PubMedPubMedCentralCrossRef
25.
Zurück zum Zitat Sholeh M, Khoshnood S, Azimi T, Mohamadi J, Kaviar VH, Hashemian M, et al. The prevalence of clarithromycin-resistant Helicobacter pylori isolates: a systematic review and meta-analysis. PeerJ. 2023;11: e15121.PubMedPubMedCentralCrossRef Sholeh M, Khoshnood S, Azimi T, Mohamadi J, Kaviar VH, Hashemian M, et al. The prevalence of clarithromycin-resistant Helicobacter pylori isolates: a systematic review and meta-analysis. PeerJ. 2023;11: e15121.PubMedPubMedCentralCrossRef
26.
Zurück zum Zitat Teh X, Khosravi Y, Lee WC, Leow AHR, Loke MF, Vadivelu J, et al. Functional and molecular surveillance of Helicobacter pylori antibiotic resistance in Kuala Lumpur. PLoS ONE. 2014;9: e101481.PubMedPubMedCentralCrossRef Teh X, Khosravi Y, Lee WC, Leow AHR, Loke MF, Vadivelu J, et al. Functional and molecular surveillance of Helicobacter pylori antibiotic resistance in Kuala Lumpur. PLoS ONE. 2014;9: e101481.PubMedPubMedCentralCrossRef
27.
Zurück zum Zitat Hanafiah A, Binmaeil H, Raja Ali RA, Mohamed Rose I, Lopes BS. Molecular characterization and prevalence of antibiotic resistance in Helicobacter pylori isolates in Kuala Lumpur, Malaysia. Infect Drug Resist. 2019;12:3051–61.PubMedPubMedCentralCrossRef Hanafiah A, Binmaeil H, Raja Ali RA, Mohamed Rose I, Lopes BS. Molecular characterization and prevalence of antibiotic resistance in Helicobacter pylori isolates in Kuala Lumpur, Malaysia. Infect Drug Resist. 2019;12:3051–61.PubMedPubMedCentralCrossRef
28.
Zurück zum Zitat Goldman RC, Zakula D, Flamm R, Beyer J, Capobianco J. Tight binding of clarithromycin, its 14-(R)-hydroxy metabolite, and erythromycin to Helicobacter pylori ribosomes. Antimicrob Agents Chemother. 1994;38:1496–500.PubMedPubMedCentralCrossRef Goldman RC, Zakula D, Flamm R, Beyer J, Capobianco J. Tight binding of clarithromycin, its 14-(R)-hydroxy metabolite, and erythromycin to Helicobacter pylori ribosomes. Antimicrob Agents Chemother. 1994;38:1496–500.PubMedPubMedCentralCrossRef
30.
Zurück zum Zitat Tamburini E, Mastromei G. Do bacterial cryptic genes really exist? Res Microbiol. 2000;151:179–82.PubMedCrossRef Tamburini E, Mastromei G. Do bacterial cryptic genes really exist? Res Microbiol. 2000;151:179–82.PubMedCrossRef
31.
Zurück zum Zitat Zhao S, White DG, Ge B, Ayers S, Friedman S, English L, et al. Identification and characterization of integron-mediated antibiotic resistance among shiga toxin-producing Escherichia coli isolates. Appl Environ Microbiol. 2001;67:1558–64.PubMedPubMedCentralCrossRef Zhao S, White DG, Ge B, Ayers S, Friedman S, English L, et al. Identification and characterization of integron-mediated antibiotic resistance among shiga toxin-producing Escherichia coli isolates. Appl Environ Microbiol. 2001;67:1558–64.PubMedPubMedCentralCrossRef
32.
Zurück zum Zitat Carvalho KR, Carvalho-Assef AP, Santos LG, Pereira MJ, Asensi MD. Occurrence of blaOXA-23 gene in imipenem-susceptible Acinetobacter baumannii. Mem Inst Oswaldo Cruz. 2011;106:505–6.PubMedCrossRef Carvalho KR, Carvalho-Assef AP, Santos LG, Pereira MJ, Asensi MD. Occurrence of blaOXA-23 gene in imipenem-susceptible Acinetobacter baumannii. Mem Inst Oswaldo Cruz. 2011;106:505–6.PubMedCrossRef
33.
Zurück zum Zitat Fernandes MR, Moura Q, Sartori L, Silva KC, Cunha MP, Esposito F, et al. Silent dissemination of colistin-resistant Escherichia coli in South America could contribute to the global spread of the mcr-1 gene. Euro Surveill. 2016;21:30214.CrossRef Fernandes MR, Moura Q, Sartori L, Silva KC, Cunha MP, Esposito F, et al. Silent dissemination of colistin-resistant Escherichia coli in South America could contribute to the global spread of the mcr-1 gene. Euro Surveill. 2016;21:30214.CrossRef
34.
Zurück zum Zitat Lanz R, Kuhnert P, Boerlin P. Antimicrobial resistance and resistance gene determinants in clinical Escherichia coli from different animal species in Switzerland. Vet Microbiol. 2003;91:73–84.PubMedCrossRef Lanz R, Kuhnert P, Boerlin P. Antimicrobial resistance and resistance gene determinants in clinical Escherichia coli from different animal species in Switzerland. Vet Microbiol. 2003;91:73–84.PubMedCrossRef
35.
Zurück zum Zitat Ma M, Wang H, Yu Y, Zhang D, Liu S. Detection of antimicrobial resistance genes of pathogenic Salmonella from swine with DNA microarray. J Vet Diagn Invest. 2007;19:161–7.PubMedCrossRef Ma M, Wang H, Yu Y, Zhang D, Liu S. Detection of antimicrobial resistance genes of pathogenic Salmonella from swine with DNA microarray. J Vet Diagn Invest. 2007;19:161–7.PubMedCrossRef
36.
Zurück zum Zitat Stasiak M, Maćkiw E, Kowalska J, Kucharek K, Postupolski J. Silent genes: antimicrobial resistance and antibiotic production. Pol J Microbiol. 2021;70:421–9.PubMedPubMedCentralCrossRef Stasiak M, Maćkiw E, Kowalska J, Kucharek K, Postupolski J. Silent genes: antimicrobial resistance and antibiotic production. Pol J Microbiol. 2021;70:421–9.PubMedPubMedCentralCrossRef
37.
Zurück zum Zitat Xu F, Nazari B, Moon K, Bushin LB, Seyedsayamdost MR. Discovery of a cryptic antifungal compound from Streptomyces albus J1074 using high-throughput elicitor screens. J Am Chem Soc. 2017;139:9203–12.PubMedPubMedCentralCrossRef Xu F, Nazari B, Moon K, Bushin LB, Seyedsayamdost MR. Discovery of a cryptic antifungal compound from Streptomyces albus J1074 using high-throughput elicitor screens. J Am Chem Soc. 2017;139:9203–12.PubMedPubMedCentralCrossRef
38.
Zurück zum Zitat Kinoshita-Daitoku R, Kiga K, Miyakoshi M, Otsubo R, Ogura Y, Sanada T, et al. A bacterial small RNA regulates the adaptation of Helicobacter pylori to the host environment. Nat Commun. 2021;12:2085.PubMedPubMedCentralCrossRef Kinoshita-Daitoku R, Kiga K, Miyakoshi M, Otsubo R, Ogura Y, Sanada T, et al. A bacterial small RNA regulates the adaptation of Helicobacter pylori to the host environment. Nat Commun. 2021;12:2085.PubMedPubMedCentralCrossRef
39.
Zurück zum Zitat Backert S, Haas R, Gerhard M, Naumann M. The Helicobacter pylori type IV secretion system encoded by the cag pathogenicity island: architecture, function, and signaling. Curr Top Microbiol Immunol. 2017;413:187–220.PubMed Backert S, Haas R, Gerhard M, Naumann M. The Helicobacter pylori type IV secretion system encoded by the cag pathogenicity island: architecture, function, and signaling. Curr Top Microbiol Immunol. 2017;413:187–220.PubMed
41.
Zurück zum Zitat Zhong Q, Shao S, Mu R, Wang H, Huang S, Han J, et al. Characterization of peptidoglycan hydrolase in cag pathogenicity island of Helicobacter pylori. Mol Biol Rep. 2011;38:503–9.PubMedCrossRef Zhong Q, Shao S, Mu R, Wang H, Huang S, Han J, et al. Characterization of peptidoglycan hydrolase in cag pathogenicity island of Helicobacter pylori. Mol Biol Rep. 2011;38:503–9.PubMedCrossRef
42.
Zurück zum Zitat Hammond CE, Beeson C, Suarez G, Peek RM Jr, Backert S, Smolka AJ. Helicobacter pylori virulence factors affecting gastric proton pump expression and acid secretion. Am J Physiol Gastrointest. 2015;309:G193–201.CrossRef Hammond CE, Beeson C, Suarez G, Peek RM Jr, Backert S, Smolka AJ. Helicobacter pylori virulence factors affecting gastric proton pump expression and acid secretion. Am J Physiol Gastrointest. 2015;309:G193–201.CrossRef
43.
Zurück zum Zitat Ansari S, Yamaoka Y. Helicobacter pylori virulence factor cytotoxin-associated gene A (CagA)-mediated gastric pathogenicity. Int J Mol Sci. 2020;21:7430.PubMedPubMedCentralCrossRef Ansari S, Yamaoka Y. Helicobacter pylori virulence factor cytotoxin-associated gene A (CagA)-mediated gastric pathogenicity. Int J Mol Sci. 2020;21:7430.PubMedPubMedCentralCrossRef
44.
Zurück zum Zitat Backert S, Tegtmeyer N, Fischer W. Composition, structure and function of the Helicobacter pylori cag pathogenicity island encoded type IV secretion system. Future Microbiol. 2015;10:955–65.PubMedCrossRef Backert S, Tegtmeyer N, Fischer W. Composition, structure and function of the Helicobacter pylori cag pathogenicity island encoded type IV secretion system. Future Microbiol. 2015;10:955–65.PubMedCrossRef
45.
Zurück zum Zitat Takahashi-Kanemitsu A, Knight CT, Hatakeyama M. Molecular anatomy and pathogenic actions of Helicobacter pylori CagA that underpin gastric carcinogenesis. Cell Mol Immunol. 2020;17:50–63.PubMedCrossRef Takahashi-Kanemitsu A, Knight CT, Hatakeyama M. Molecular anatomy and pathogenic actions of Helicobacter pylori CagA that underpin gastric carcinogenesis. Cell Mol Immunol. 2020;17:50–63.PubMedCrossRef
47.
Zurück zum Zitat Voss BJ, Gaddy JA, McDonald WH, Cover TL. Analysis of surface-exposed outer membrane proteins in Helicobacter pylori. J Bacteriol. 2014;196:2455–71.PubMedPubMedCentralCrossRef Voss BJ, Gaddy JA, McDonald WH, Cover TL. Analysis of surface-exposed outer membrane proteins in Helicobacter pylori. J Bacteriol. 2014;196:2455–71.PubMedPubMedCentralCrossRef
49.
Zurück zum Zitat Mittl PRE, Schneider-Brachert W. Sel1-like repeat proteins in signal transduction. Cell Signalling. 2007;19:20–31.PubMedCrossRef Mittl PRE, Schneider-Brachert W. Sel1-like repeat proteins in signal transduction. Cell Signalling. 2007;19:20–31.PubMedCrossRef
50.
Zurück zum Zitat Li M-S, Langford PR, Kroll JS. Inactivation of NMB0419, encoding a sel1-like repeat (SLR) protein, in Neisseria meningitidis is associated with differential expression of genes belonging to the fur regulon and reduced intraepithelial replication. Infect Immun. 2017;85:10–1128. https://doi.org/10.1128/iai.00574-16.CrossRef Li M-S, Langford PR, Kroll JS. Inactivation of NMB0419, encoding a sel1-like repeat (SLR) protein, in Neisseria meningitidis is associated with differential expression of genes belonging to the fur regulon and reduced intraepithelial replication. Infect Immun. 2017;85:10–1128. https://​doi.​org/​10.​1128/​iai.​00574-16.CrossRef
51.
Zurück zum Zitat Dumrese C, Slomianka L, Ziegler U, Choi SS, Kalia A, Fulurija A, et al. The secreted Helicobacter cysteine-rich protein A causes adherence of human monocytes and differentiation into a macrophage-like phenotype. FEBS Lett. 2009;583:1637–43.PubMedPubMedCentralCrossRef Dumrese C, Slomianka L, Ziegler U, Choi SS, Kalia A, Fulurija A, et al. The secreted Helicobacter cysteine-rich protein A causes adherence of human monocytes and differentiation into a macrophage-like phenotype. FEBS Lett. 2009;583:1637–43.PubMedPubMedCentralCrossRef
52.
53.
Zurück zum Zitat Newton HJ, Sansom FM, Dao J, McAlister AD, Sloan J, Cianciotto NP, et al. Sel1 repeat protein LpnE is a Legionella pneumophila virulence determinant that influences vacuolar trafficking. Infect Immun. 2007;75:5575–85.PubMedPubMedCentralCrossRef Newton HJ, Sansom FM, Dao J, McAlister AD, Sloan J, Cianciotto NP, et al. Sel1 repeat protein LpnE is a Legionella pneumophila virulence determinant that influences vacuolar trafficking. Infect Immun. 2007;75:5575–85.PubMedPubMedCentralCrossRef
54.
Zurück zum Zitat Lee KH, Park SY, Jeong SJ, Jung DH, Kim J-H, Jeong SH, et al. Can aminoglycosides be used as a new treatment for Helicobacter pylori? In vitro activity of recently isolated Helicobacter pylori. Infect Chemother. 2019;51:10–20.PubMedPubMedCentralCrossRef Lee KH, Park SY, Jeong SJ, Jung DH, Kim J-H, Jeong SH, et al. Can aminoglycosides be used as a new treatment for Helicobacter pylori? In vitro activity of recently isolated Helicobacter pylori. Infect Chemother. 2019;51:10–20.PubMedPubMedCentralCrossRef
55.
Zurück zum Zitat Lioy VS, Goussard S, Guerineau V, Yoon EJ, Courvalin P, Galimand M, et al. Aminoglycoside resistance 16S rRNA methyltransferases block endogenous methylation, affect translation efficiency and fitness of the host. RNA. 2014;20:382–91.PubMedPubMedCentralCrossRef Lioy VS, Goussard S, Guerineau V, Yoon EJ, Courvalin P, Galimand M, et al. Aminoglycoside resistance 16S rRNA methyltransferases block endogenous methylation, affect translation efficiency and fitness of the host. RNA. 2014;20:382–91.PubMedPubMedCentralCrossRef
56.
Zurück zum Zitat Matos R, De Witte C, Smet A, Berlamont H, De Bruyckere S, Amorim I, et al. Antimicrobial susceptibility pattern of Helicobacter heilmannii and Helicobacter ailurogastricus isolates. Microorganisms. 2020;8:957.PubMedPubMedCentralCrossRef Matos R, De Witte C, Smet A, Berlamont H, De Bruyckere S, Amorim I, et al. Antimicrobial susceptibility pattern of Helicobacter heilmannii and Helicobacter ailurogastricus isolates. Microorganisms. 2020;8:957.PubMedPubMedCentralCrossRef
59.
Zurück zum Zitat Brink A. Antibiotic resistance and virulence. Int J Infect Dis. 2014;21:64.CrossRef Brink A. Antibiotic resistance and virulence. Int J Infect Dis. 2014;21:64.CrossRef
60.
Zurück zum Zitat Uddin TM, Chakraborty AJ, Khusro A, Zidan BMRM, Mitra S, Emran TB, et al. Antibiotic resistance in microbes: history, mechanisms, therapeutic strategies and future prospects. J Infect Public Health. 2021;14:1750–66.PubMedCrossRef Uddin TM, Chakraborty AJ, Khusro A, Zidan BMRM, Mitra S, Emran TB, et al. Antibiotic resistance in microbes: history, mechanisms, therapeutic strategies and future prospects. J Infect Public Health. 2021;14:1750–66.PubMedCrossRef
62.
Zurück zum Zitat Penesyan A, Paulsen IT, Gillings MR, Kjelleberg S, Manefield MJ. Secondary effects of antibiotics on microbial biofilms. Front Microbiol. 2020;11:2109.PubMedPubMedCentralCrossRef Penesyan A, Paulsen IT, Gillings MR, Kjelleberg S, Manefield MJ. Secondary effects of antibiotics on microbial biofilms. Front Microbiol. 2020;11:2109.PubMedPubMedCentralCrossRef
63.
Zurück zum Zitat Hathroubi S, Hu S, Ottemann KM. Genetic requirements and transcriptomics of Helicobacter pylori biofilm formation on abiotic and biotic surfaces. NPJ Biofilms Microbiomes. 2020;6:56.PubMedPubMedCentralCrossRef Hathroubi S, Hu S, Ottemann KM. Genetic requirements and transcriptomics of Helicobacter pylori biofilm formation on abiotic and biotic surfaces. NPJ Biofilms Microbiomes. 2020;6:56.PubMedPubMedCentralCrossRef
64.
Zurück zum Zitat Luo P, Huang Y, Hang X, Tong Q, Zeng L, Jia J, et al. Dihydrotanshinone I is effective against drug-resistant Helicobacter pylori in vitro and in vivo. Antimicrob Agents Chemother. 2021;65:10–1128.CrossRef Luo P, Huang Y, Hang X, Tong Q, Zeng L, Jia J, et al. Dihydrotanshinone I is effective against drug-resistant Helicobacter pylori in vitro and in vivo. Antimicrob Agents Chemother. 2021;65:10–1128.CrossRef
65.
Zurück zum Zitat Windham IH, Servetas SL, Whitmire JM, Pletzer D, Hancock REW, Merrell DS. Helicobacter pylori biofilm formation is differentially affected by common culture conditions, and proteins play a central role in the biofilm matrix. Appl Environ Microbiol. 2018;84:e00391-18.PubMedPubMedCentralCrossRef Windham IH, Servetas SL, Whitmire JM, Pletzer D, Hancock REW, Merrell DS. Helicobacter pylori biofilm formation is differentially affected by common culture conditions, and proteins play a central role in the biofilm matrix. Appl Environ Microbiol. 2018;84:e00391-18.PubMedPubMedCentralCrossRef
66.
Zurück zum Zitat Bugli F, Palmieri V, Torelli R, Papi M, De Spirito M, Cacaci M, et al. In vitro effect of clarithromycin and alginate lyase against Helicobacter pylori biofilm. Biotechnol Prog. 2016;32:1584–91.PubMedCrossRef Bugli F, Palmieri V, Torelli R, Papi M, De Spirito M, Cacaci M, et al. In vitro effect of clarithromycin and alginate lyase against Helicobacter pylori biofilm. Biotechnol Prog. 2016;32:1584–91.PubMedCrossRef
67.
Zurück zum Zitat Ratthawongjirakul P, Thongkerd V, Chaicumpa W. The impacts of a fliD mutation on the biofilm formation of Helicobacter pylori. Asian Pac J Trop Biomed. 2016;6:1008–14.CrossRef Ratthawongjirakul P, Thongkerd V, Chaicumpa W. The impacts of a fliD mutation on the biofilm formation of Helicobacter pylori. Asian Pac J Trop Biomed. 2016;6:1008–14.CrossRef
68.
Zurück zum Zitat Ji J, Yang H. In vitro effects of Lactobacillus plantarum LN66 and antibiotics used alone or in combination on Helicobacter pylori mature biofilm. Microorganisms. 2021;9:424.PubMedPubMedCentralCrossRef Ji J, Yang H. In vitro effects of Lactobacillus plantarum LN66 and antibiotics used alone or in combination on Helicobacter pylori mature biofilm. Microorganisms. 2021;9:424.PubMedPubMedCentralCrossRef
69.
Zurück zum Zitat Benoit S, Mehta N, Wang G, Gatlin M, Maier RJ. Requirement of hydD, hydE, hypC and hypE genes for hydrogenase activity in Helicobacter pylori. Microb Pathog. 2004;36:153–7.PubMedCrossRef Benoit S, Mehta N, Wang G, Gatlin M, Maier RJ. Requirement of hydD, hydE, hypC and hypE genes for hydrogenase activity in Helicobacter pylori. Microb Pathog. 2004;36:153–7.PubMedCrossRef
70.
Zurück zum Zitat Wilkinson D, Alsharaf L, Thompson S, Paulin A, Takor R, Zaitoun A, et al. Characterization of a Helicobacter pylori strain with high biofilm-forming ability. J Med Microbiol. 2023;72:001710.CrossRef Wilkinson D, Alsharaf L, Thompson S, Paulin A, Takor R, Zaitoun A, et al. Characterization of a Helicobacter pylori strain with high biofilm-forming ability. J Med Microbiol. 2023;72:001710.CrossRef
71.
72.
Zurück zum Zitat Khan J, Tarar SM, Gul I, Nawaz U, Arshad M. Challenges of antibiotic resistance biofilms and potential combating strategies: a review. 3 Biotech. 2021;11:169.PubMedPubMedCentralCrossRef Khan J, Tarar SM, Gul I, Nawaz U, Arshad M. Challenges of antibiotic resistance biofilms and potential combating strategies: a review. 3 Biotech. 2021;11:169.PubMedPubMedCentralCrossRef
Metadaten
Titel
Polymorphism of virulence genes and biofilm associated with in vitro induced resistance to clarithromycin in Helicobacter pylori
verfasst von
Naim Asyraf Rosli
Anis Rageh Al-Maleki
Mun Fai Loke
Eng Guan Chua
Mohammed Abdelfatah Alhoot
Jamuna Vadivelu
Publikationsdatum
01.12.2023
Verlag
BioMed Central
Erschienen in
Gut Pathogens / Ausgabe 1/2023
Elektronische ISSN: 1757-4749
DOI
https://doi.org/10.1186/s13099-023-00579-4

Weitere Artikel der Ausgabe 1/2023

Gut Pathogens 1/2023 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Die „Zehn Gebote“ des Endokarditis-Managements

30.04.2024 Endokarditis Leitlinie kompakt

Worauf kommt es beim Management von Personen mit infektiöser Endokarditis an? Eine Kardiologin und ein Kardiologe fassen die zehn wichtigsten Punkte der neuen ESC-Leitlinie zusammen.

Strenge Blutdruckeinstellung lohnt auch im Alter noch

30.04.2024 Arterielle Hypertonie Nachrichten

Ältere Frauen, die von chronischen Erkrankungen weitgehend verschont sind, haben offenbar die besten Chancen, ihren 90. Geburtstag zu erleben, wenn ihr systolischer Blutdruck < 130 mmHg liegt. Das scheint selbst für 80-Jährige noch zu gelten.

Frauen bekommen seltener eine intensive Statintherapie

30.04.2024 Statine Nachrichten

Frauen in den Niederlanden erhalten bei vergleichbarem kardiovaskulärem Risiko seltener eine intensive Statintherapie als Männer. Ihre LDL-Zielwerte erreichen sie aber fast ähnlich oft.

Reizdarmsyndrom: Diäten wirksamer als Medikamente

29.04.2024 Reizdarmsyndrom Nachrichten

Bei Reizdarmsyndrom scheinen Diäten, wie etwa die FODMAP-arme oder die kohlenhydratreduzierte Ernährung, effektiver als eine medikamentöse Therapie zu sein. Das hat eine Studie aus Schweden ergeben, die die drei Therapieoptionen im direkten Vergleich analysierte.

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.