Next Article in Journal
Can MRI Accurately Diagnose and Stage Endometrial Adenocarcinoma?
Previous Article in Journal
Differences in Pathophysiology and Treatment Efficacy Based on Heterogeneous Out-of-Hospital Cardiac Arrest
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Intra-Hospital Spread of Colistin-Resistant K. pneumoniae Isolates—Epidemiological, Clinical, and Genetic Analysis

by
Georgi Popivanov
1,*,
Rumyana Markovska
2,
Ivanka Gergova
3,
Marina Konaktchieva
4,
Roberto Cirocchi
5,
Kirien Kjossev
1 and
Ventsislav Mutafchiyski
1
1
Department of Surgery, Military Medical Academy, 1606 Sofia, Bulgaria
2
Department of Medical Microbiology, Medical Faculty, Medical University, 1431 Sofia, Bulgaria
3
Department of Microbiology and Virology, Military Medical Academy, 1606 Sofia, Bulgaria
4
Department of Gastroenterology and Hepatology, Military Medical Academy, 1606 Sofia, Bulgaria
5
Department of Surgical Science, University of Perugia, 06100 Perugia, Italy
*
Author to whom correspondence should be addressed.
Medicina 2024, 60(3), 511; https://doi.org/10.3390/medicina60030511
Submission received: 29 January 2024 / Revised: 18 March 2024 / Accepted: 19 March 2024 / Published: 21 March 2024
(This article belongs to the Section Infectious Disease)

Abstract

:
Background and Objective: Klebsiella pneumoniae appears to be a significant problem due to its ability to accumulate antibiotic-resistance genes. After 2013, alarming colistin resistance rates among carbapenem-resistant K. pneumoniae have been reported in the Balkans. The study aims to perform an epidemiological, clinical, and genetic analysis of a local outbreak of COLr CR-Kp. Material and Methods: All carbapenem-resistant and colistin-resistant K. pneumoniae isolates observed among patients in the ICU unit of Military Medical Academy, Sofia, from 1 January to 31 October 2023, were included. The results were analyzed according to the EUCAST criteria. All isolates were screened for blaVIM, blaIMP, blaKPC, blaNDM, and blaOXA-48. Genetic similarity was determined using the Dice coefficient as a similarity measure and the unweighted pair group method with arithmetic mean (UPGMA). mgrB genes and plasmid-mediated colistin resistance determinants (mcr-1, mcr-2, mcr-3, mcr-4, and mcr-5) were investigated. Results: There was a total of 379 multidrug-resistant K. pneumoniae isolates, 88% of which were carbapenem-resistant. Of these, there were nine (2.7%) colistin-resistant isolates in six patients. A time and space cluster for five patients was found. Epidemiology typing showed that two isolates belonged to clone A (pts. 1, 5) and the rest to clone B (pts. 2–4) with 69% similarity. Clone A isolates were coproducers of blaNDM-like and blaOXA-48-like and had mgrB-mediated colistin resistance (40%). Clone B isolates had only blaOXA-48-like and intact mgrB genes. All isolates were negative for mcr-1, -2, -3, -4, and -5 genes. Conclusions: The study describes a within-hospital spread of two clones of COLr CR-Kp with a 60% mortality rate. Clone A isolates were coproducers of NDM-like and OXA-48-like enzymes and had mgrB-mediated colistin resistance. Clone B isolates had only OXA-48-like enzymes and intact mgrB genes. No plasmid-mediated resistance was found. The extremely high mortality rate and limited treatment options warrant strict measures to prevent outbreaks.

1. Introduction

HAIs (hospital-acquired infections) account for 5–15% of all admissions worldwide (9 million), but the rate is probably higher because of significant underreporting [1]. The total annual cost for the five significant HAIs in the USA is USD 9.8 billion [2]. Almost 100 years after the discovery of antibiotics, we are faced with unprecedented antibiotic resistance, leading to a catastrophic crisis worldwide. Multidrug resistance (MDR) is defined as resistance to one or more antimicrobials from at least three different antimicrobial classes; extensive drug resistance (XDR) is non-susceptibility to at least one agent in all but two or fewer antimicrobial categories (i.e., bacterial isolates remain susceptible to only one or two categories); and pan-drug resistance (PDR) is resistance to all agents in all antimicrobial categories [3]. MDR pathogens account for 670,000 infections and 33,000 deaths in the European Union with healthcare costs of USD 1.1 billion [4,5]. Antimicrobial Resistance Collaborators estimated that 1.27 million deaths were attributable to antibiotic resistance during 2019. Six of twenty-three analyzed pathogens (Escherichia coli, Staphylococcus aureus, Klebsiella pneumoniae, Streptococcus pneumoniae, Acinetobacter baumannii, and Pseudomonas aeruginosa) were responsible for 929,000 of these deaths [6]. A UK report warned that by 2050 approximately 10 million deaths would occur if no action was taken [7]. These figures, however, were questioned by others, mainly due to a lack of reliable estimates of the antibiotic resistance burden [8]. Nevertheless, WHO declared priority status to the most frequently reported MDR bacteria called “ESKAPE” (Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter species) due to their ability to escape the antibiotics [9,10]. The MDR Enterobacteriaceae are essential because they are part of the human microbiota. Among these, K. pneumoniae appeared to be a significant problem due to its ability to accumulate antibiotic resistance genes (ARGs) by de novo mutations or via horizontal gene transfer by plasmid transfer (conjugation) or by bacteriophage (transduction); thus, Navon-Venezia et al. called it a “source and shuttle for antibiotic resistance” [11]. Via the above-mentioned mechanisms, ARGs could be easily transferred from harmless commensals to pathogenic bacteria [12]. According to a recent survey, K. pneumoniae is among the five deadliest bacteria, with more than 500,000 deaths yearly [13]. The carbapenem-resistant K. pneumoniae was first reported in 2001 in the USA [14]. The main reason for the appearance of carbapenem-resistant isolates is the production of carbapenemases. They are classified into two main groups: the first is the serine active group (class A Klebsiella-producing carbapenemase (KPC) and class D oxacillinases (OXA; OXA-48 and OXA-181), and their variants, are the main representatives) and the second comprises class B metallo-carbapenemases (New Delhi Metallo beta-lactamase (NDM), Verona Integron metallo-carbapenemase (VIM), and Imipenemase (IMP) enzymes) [15]. The European survey of carbapenemase-producing Enterobacteriaceae (EuSCAPE) demonstrated that the epidemic of carbapenem-resistant Klebsiella pneumoniae in Europe is driven by “within-hospital transmission and interhospital spread rather than between countries” [16]. According to ECDC (2021), the MDR strains account for 21%, with an increasing trend for carbapenemase-producing strains (11.7%) [5].
The Balkans are considered a reservoir of MDR and XDR K. pneumoniae. NDM-1-producing strains appear to spread within a similar time frame (2013–2016) [4,17,18,19]. The first polyclonal outbreak caused by NDM-1-producing K. pneumoniae in Bulgaria was described in 2015 at our and other institutions [17,18]. In 2019, Markovska et al. demonstrated the rapid interregional spread of NDM-1-producing ST11 strains with plasmid-mediated carbapenem resistance [20]. Unfortunately, the last link of the chain was the emergence of colistin-resistant strains. Polymyxin was discovered in 1947 and has been available on the market since 1959 for the treatment of Gram-negative infections [21]. Due to its nephrotoxicity, in the 1970s it was replaced by other antibiotics. Due to the emerging MDR crisis, colistin was “re-discovered” in 2000 and began to play a strategic role in the treatment of MDR Gram-negative infections [22,23]. At the same time, the rapid increase in its use rapidly led to an increase in resistance [21]. The studies published in the period 2008–2011 demonstrated a rate of colistin-resistant K. pneumoniae (COLr CR-Kp) between 1.5% and 28% [21]. According to Binsker et al., citing the ATLAS database, the global colistin resistance rate for 2014–2019 varied between 2.6% and 4.6%, and between 2.4 and 3.4% for Europe [24]. The first cases in Bulgaria were reported by Markovska et al. in 2015, and in our institution and other hospitals in Sofia and other cities in 2016 [25,26,27,28]. After 2013, alarming colistin resistance among carbapenem-resistant K. pneumoniae was reported in the Balkans–Bulgaria (37%), Greece (40%), Romania (27.5%), Serbia (10.6%), Türkiye (25.5%), and Italy (27%) [24,26,27,28,29,30,31,32]. More recent work from Poland for the 2019/2021 period reported a resistance rate of 14.5% [33]. Unfortunately, there are not enough data for the COVID-19 and post-COVID-19 periods for our region. The imminent disaster highlights the need for emergent measures to prevent its spread and to find new treatments because of the high mortality rate (41–70%) [34,35].
This study aims to perform an epidemiological, clinical, and genetic analysis of a local outbreak of COLr CR-Kp.

2. Material and Methods

2.1. Bacterial Isolates and Patients

The colistin-resistant and carbapenem-resistant K. pneumoniae isolates observed among patients in the ICU unit of Military Medical Academy, Sofia from 1 January to 31 October 2023 were included in the study. The hospital has 800 beds and more than 40,000 admissions per year.
The identifications of the microbial isolates were performed by MALDI-TOF mass spectrometry (MALDI-TOF MS, Bruker Corp. Billerica, MA, USA), following the manufacturer’s instructions.

2.2. Antimicrobial Susceptibility Testing

The antimicrobial susceptibility testing (AST) was determined by Vitek 2 (bioMerieux, Marcy-l’Étoile, France). The reference method broth microdilution (ComASP Colistin, Liofilchem, Roseto degli Abruzzi (Te) Italy) was used to perform AST of colistin. The results were analyzed according to the criteria of the European Committee on Antimicrobial Susceptibility Testing (EUCAST) [36].

2.3. Molecular-Genetic Investigations

All isolates were PCR screened for the presence of blaVIM, blaIMP, blaKPC, blaNDM, and blaOXA-48, as previously described [37]. The mgrB gene was amplified and sequenced with primers reported previously by Kanateli. Nucleotide and deduced amino acid sequences were analyzed and multiple alignments were performed using Chromas Lite 2.01 (Technelysium Pty Ltd., Brisbane, Australia) and DNAMAN version 8.0 Software (Lynnon BioSoft, Vaudreuil-Dorion, QC, Canada).
Total bacterial DNA was prepared using the boiling method. ERIC (Enterobacterial Repetitive Intergenic Consensus) PCR with ERIC1R and ERIC2 primer sets was performed as previously described [20]. Genetic similarity of ERIC fingerprints was determined using the simple clustering method, UPGMA (unweighted pair group method with arithmetic mean), and the Dice coefficient as a similarity measure (http://genomes.urv.cat/UPGMA/, accessed on 20 January 2024). A clone was defined if the isolate showed a coefficient of similarity above 0.8.
Plasmid-mediated colistin resistance determinants (mcr-1, mcr-2, mcr-3, mcr-4, and mcr-5) were investigated with multiplex PCR, as suggested by Lescat et al. [38]. The chromosomal mgrB gene was amplified with primer sets as described previously [39].

3. Results

3.1. Bacterial Isolates and Patient’s Characteristics

A total of 379 MDR-K. pneumoniae isolates were isolated in the ICU unit, of which 333 (87.9%) were carbapenem resistant. Of these, six patients had nine (2.7%) colistin-resistant isolates. A time and space cluster was observed for five patients with eight colistin-resistant K. pneumoniae isolates. They were treated in the same room in the ICU within a two-week overlapping interval. All COLr CR-Kp were isolated within 13 days (31.08–13.09). The characteristics of the patients and isolates are shown in Table 1 and Table 2. All patients, except the fifth, had previous endoscopic interventions and initially colistin-susceptible K. pneumoniae; however, in the table they are given as only COLr CR-Kp.
The putative index patient in the series was a 69-year-old woman who underwent a left nephrectomy due to acute bleeding after retrograde intrarenal surgery, lithotripsy, and JJ stent. The urine culture revealed colistin-susceptible, carbapenem-resistant K. pneumoniae (22.08, data not shown), which was treated with a suboptimal dose regimen of colistin (2 × 1 MM U) due to kidney failure. The first COLr CR-Kp was isolated from the wound tissue (31.08). The wound was treated locally, but after seven days a reoperation was performed due to organ space infection and severe necrotizing fasciitis. The patient developed sepsis with a positive blood culture (8.09) caused by the same strain and died on the 42nd postoperative day (POD).
The second and third patients underwent a Traverso–Longmire procedure for pancreatic cancer; both had bile duct stenting one month before the operation. The first isolates from the bile during the index operation in both patients were colistin-susceptible, carbapenem-resistant K. pneumoniae. In the second patient, COLr CR-Kp was recovered from the wound (3.09) and ascites (12.09). In the third patient, it was also isolated from the wound (12.09). Both patients were treated in one surgical clinic and both underwent reoperation on the same day (12.09).
In the fourth patient, colistin-susceptible, carbapenem-resistant K. pneumoniae was found in the urine (31.09), which was also treated with a suboptimal dose regimen of colistin (2 × 1 MM U). COLr CR-Kp was isolated from the tracheobronchial tree (4.09), on the seventh POD followed by positive blood culture (13.09). In the fifth patient, COLr CR-Kp was isolated from the urine on the sixth POD (11.09). Three of the five patients died (60%).

3.2. Antimicrobial Susceptibility Testing

All investigated K pneumoniae isolates were PDR and showed identical results and were resistant to amoxicillin, ampicillin, ampicillin/sulbactam, amoxicillin/clavulanic acid, piperacillin/tazobactam, cefalexine, cefuroxime, cefixime, ceftazidime, ceftriaxone, cefepime, imipenem, meropenem, colistin, gentamicin, amikacin, levofloxacin, fosfomycin, ciprofloxacin, moxifloxacin, and trimethoprim/sulfamethoxazole.

3.3. Molecular-Genetic Investigations

PCR reactions confirmed the production of carbapenemases (NDM or/and OXA-48) in all eight isolates. Two clones were confirmed by ERIC PCR and UPGMA analysis showing a 0.69 coefficient of similarity. Clone A was a coproducer of blaNDM-like and blaOXA-48-like enzymes (isolates K pneumoniae 1, 2, and 8 (patients 1 and 5)), whereas clone B (isolates 3, 4, 5, 6, 7 (pts. 2–4)) harbored only blaOXA-48-like enzymes. The entire mgrB gene was amplified by PCR. In the isolates from clone A, the mgrB genes could not be amplified, showing truncated genes. In clone B, there were amplicons in all three isolates (intact mgrB genes), (Table 1). All isolates were negative for mcr-1, -2, -3, -4, and -5 genes.
A broad hospital infection-control campaign was initiated encompassing the other patients, staff, and working environment, but no other COLr CR-Kp strains were isolated. The campaign included tightened control on the patients with XDR infections, with isolation in boxes or single rooms with access only for dedicated personnel. The patients were reoperated on in a dedicated “septic” operative room without interference with the “clean” patients. Regular microbiological specimens were taken from all the personnel, clinics, and operating theatres at random. For acutely ill patients referred from other hospitals, a microbiological survey was performed at admission. The Antimicrobial Stewardship Program was continued with a more restrictive antibiotic policy. The antimicrobial treatment of the complex cases was discussed by a multidisciplinary team including microbiologists.

4. Discussion

The present analysis revealed two clones of COLr CR-Kp that coincided with the time and space in the ICU unit. The most important was clone A, co-producer of OXA-48-like and NDM-like enzymes, and harbored disrupted mgrB genes responsible for the colistin resistance. The isolates of this clone were resistant to all tested antimicrobials, leaving no therapeutic alternative. Clone B harbored only OXA-48-like and lacked mutations in mgrB genes, so the colistin resistance might be explained by other chromosomal mutations [38,39]. The finding suggests that our series represents a within-hospital spread of COLr CR-Kp with two foci that coincided in the time and space (ICU). We speculate that the second focus originates from the second and third patients, who were operated on and managed in one clinic by the same team with subsequent stays in the ICU close to the fourth patient.
Of note, all patients in the presented series, except the fifth, underwent endoscopic intervention before the index operation (two endourology procedures and two bile duct stentings). All of them had an initial culture of colistin-susceptible K. pneumoniae, which also demonstrates the within-hospital spread.
Our finding is similar to the EuSCAPE survey, which demonstrated the central role of the inter-hospital, but more pronounced, transmission at the hospital level, similar to the results of Markovska et al. [16,20]. Other authors, however, reported that “carbapenem resistance reveals remarkable diversity and unexplained mechanisms” and that not all outbreaks could be linked to transmissions [40].
The first and fourth patients were treated with suboptimal doses of colistin (2 × 1 MM U daily for ten days). In the first patient, COLr CR-KP was detected on the 9th day of the treatment with colistin, whereas it was detected in the fourth patient on the 5th day. However, we can only speculate that this selective antibiotic pressure might explain the transition from colistin susceptibility to COLr CR-Kp in our series [41].
Although it was declared as strategic by WHO, colistin has been increasingly used in clinical practice, with a steep increase after 2005 [20]. Moreover, in 2017, the overall consumption of polymyxins in food-producing animals in 28 EU countries was 340 times higher than that in human medicine [25]. An increasing trend was also found in our institution. The consumption of colistin increased from 0.5 definitive daily doses per bed-day in 2017 to 1.98 in 2022 for the whole hospital, and from 11.7 to 25.5 in the ICU. A logical consequence of this worrisome trend is the rapid emergence of colistin-resistant strains. A recent meta-analysis demonstrated a significant increase in bloodstream COLr CR-Kp during the last decade, from 3% in 2015 to 13% in 2020 and after [42]. In the Balkans, the first COLr CR-Kp strains were isolated in 2012, with rapid expansion leading to multiple outbreaks in Greece, Bulgaria, and probably other countries (Table 3) [28,29,30,31,32,43].
A recent study showed significant heterogeneity in the molecular mechanisms behind colistin resistance [44]. Colistin resistance in K. pneumoniae is driven by the change (decrease) in the negatively charged lipopolysaccharides on the membrane surface, which precludes the binding with the cationic colistin. Several genes, mainly encoding the proteins of two-component regulatory systems PhoPQ, PmrAB, and CrrAB, are responsible for the decrease in this negative charge and resistance [44,45]. PhoPQ and PmrAB are negatively regulated by the mgrB protein on the inner surface of the membrane. The recently described CrrAB two-component system also regulates the PmrAB system. Kim et al. demonstrated higher survival rates in crrAB-positive isolates with early exposure to high colistin concentrations [46].
In a recent meta-analysis, Yusof et al. reported a pooled prevalence of mutated colistin resistance in K. pneumoniae of about 75% [47]. The most common genetic mechanism of resistance includes mutations in the genes mgrB (88%), pmrA/pmrB (54%), phoQ (44%), and phoP (36%). Plasmid-mediated resistance via mcr-1 was noted in 14%, while other genetic mechanisms were noted in 40%. In Bulgaria, a recent study by Markovska et al. found a lack, disruption, or mutation of the mgrB gene in 9/37 cases (24%), whereas, in the rest, the mechanism of resistance was not elucidated [28]. No plasmid-mediated resistance was found. These data are lower than our results (lack of mgrB in three of eight colistin-resistant isolates (37.5%)). These genes, however, could not explain all cases with colistin resistance. Macesic et al. demonstrated multiple genetic events in 71% of clusters with more than two patients [46]. According to the authors, several other genes constitute the so-called secondary resistome. Moreover, they state that polymixin exposure with de novo mutations rather than transmissions lies behind the colistin resistance.
The effect of mgrB gene inactivation, however, goes beyond the colistin resistance. In a murine model, Bray et al. demonstrated that the inactivation of mgrB leads to increased environmental survival of K. pneumoniae and facilitates host-to-host transmission [48]. Given its ability to survive on different surfaces, these consequences make the mutations in mgrB of particular importance for the spreading of K. pneumoniae even after strict infection control measures in hospital settings.
A very important but poorly estimated phenomenon is the mgrB-induced heteroresistance to colistin. A recent work of Alousi et al. underscores that the colistin-resistant subpopulations in the background of selective colistin pressure may become dominant, with worrisome consequences. In their study, they found heteroresistance in 21.9% of the CR-K. pneumoniae isolates [49]. A recent series from Bulgaria reported 8% heteroresistance [28]. The heteroresistance may hamper the interpretation of the antibiogram and, if unrecognized, it may lead to selective colistin pressure. Therefore, the early detection of heteroresistance is crucial to avoid this phenomenon.
The mortality rate of the present series is 60% and 100% in cases with bloodstream infection, which is in unison with the literature [34,36,50]. As of today, a few treatment options exist, such as ceftazidime/avibactam with or without aztreonam, plazomicin (not approved by EMA), cefiderocol, and fosfomycin. If the MIC of imipenem is below 8 mg/L, it could also be included in combination schemes. An excellent review by Petrosillo et al. demonstrated the characteristics of these antimicrobials and highlighted the need for an analysis of the meropenem MIC value, and OXA-48-like and NDM status, to guide the treatment [51].
The extremely high mortality rate and limited treatment options warrant strict measures to prevent outbreaks. Given the overtaking bacterial resistance and the difficult control of the chaotic and frequently defensive use of antibiotics, it appears more prudent to improve the prevention. Despite the high risk of bias (99%), the published literature suggests a sustained potential for reduction in HAI rates of between 35% and 55% using multifaceted interventions irrespective of a country’s income level [52]. Keeping the ten golden rules for optimal antibiotic use is of paramount importance [53].
The main limitation of the present study is the small sample size, which might be overcome with a multicenter national survey. Future research should be focused on a better understanding of the role of the number of the genes involved in colistin resistance, the role of CrrAB, phoPQ, and PmrAB, the early detection of heteroresistance, the influence of the dosage of colistin, and the duration of the treatment with this resistance [44,46]. Last, but not least, solving the puzzle also requires better elucidation of the biological cost of the gene mutations (increased environmental survival and facilitated host-to-host transmission) [48].

5. Conclusions

The present study describes a within-hospital spread of two clones of COLr CR-Kp with a 60% mortality rate. Clone A comprised co-producers of NDM-like and OXA-48-like enzymes and had mgrB-mediated colistin resistance. Clone B isolates had only OXA-48-like enzymes and intact mgrB genes. No plasmid-mediated resistance was found. The study also confirms the central role of the transmission at the hospital level, not only for COLr CR-Kp, but also for colistin-susceptible K. pneumoniae. The extremely high mortality rate and limited treatment options warrant strict measures to prevent the outbreaks. The decisive first step is to increase awareness about this threat and to implement in practice well-known principles, such as a hospital surveillance system, prevention, closer collaboration with the microbiology laboratory, an Antimicrobial Stewardship Program with a restrictive antibiotic policy, and creation of a multidisciplinary team discussing the antimicrobial treatment of the complex cases.

Author Contributions

Conceptualization, G.P., K.K. and V.M.; methodology, G.P., R.M. and I.G.; formal analysis, G.P., M.K., R.C. and V.M.; investigation, R.M.; data curation, R.M., I.G. and M.K.; writing—original draft preparation, G.P., R.M., M.K., R.C. and K.K.; writing—review and editing, G.P., R.M., I.G., M.K., R.C., K.K. and V.M.; supervision, K.K. and V.M.; project administration, V.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Available online: https://apps.who.int/gb/MSPI/pdf_files/2022/03/Item1_07-03.pdf (accessed on 7 March 2022).
  2. Zimlichman, E.; Henderson, D.; Tamir, O.; Franz, C.; Song, P.; Yamin, C.K.; Keohane, C.; Denham, C.R.; Bates, D.W. Health Care-associated infections. A meta-analysis of costs and financial impact on the US Health Care System. JAMA Intern. Med. 2013, 173, 2039–2046. [Google Scholar] [CrossRef]
  3. Magiorakos, A.; Srinivasan, A.; Carey, R.; Carmeli, Y.; Falagas, M.; Giske, C.; Harbarth, S.; Hindler, J.F.; Kahlmeter, G.; Olsson-Liljequist, B.; et al. Multidrug-resistant, extensively drug-resistant and pandrug-resistant bacteria: An international expert proposal for interim standard definitions for acquired resistance. Clin. Microbiol. Infect. 2012, 18, 268–281. [Google Scholar] [CrossRef]
  4. Cassini, A.; Högberg, L.; Plachouras, D.; Quattrocchi, A.; Hoxha, A.; Simonsen, G.; Colomb-Cotinat, M.; Kretzschmar, M.E.; Devleesschauwer, B.; Cecchini, M.; et al. Attributable deaths and disability-adjusted life-years caused by infections with antibiotic-resistant bacteria in the EU and the European Economic Area in 2015: A population-level modelling analysis. Lancet Infect. Dis. 2019, 19, 56–66. [Google Scholar] [CrossRef]
  5. Antimicrobial Resistance in the EU/EEA (EARS-Net)—Annual Epidemiological Report for 2021. Available online: https://www.ecdc.europa.eu/en/publications-data/surveillance-antimicrobial-resistance-europe-2021 (accessed on 17 November 2023).
  6. Antimicrobial Resistance Collaborators. Global burden of bacterial antimicrobial resistance in 2019: A systematic analysis. Lancet 2022, 399, 629–655. [Google Scholar] [CrossRef]
  7. O’Neill, J. Tackling Drug-Resistant Infections Globally: Final Report and Recommendations. The Review on Antimicrobial Resistance. 2016. Available online: https://amr-review.org/sites/default/files/160525_Final%20paper_with%20cover.pdf (accessed on 1 May 2016).
  8. de Kraker, M.E.A.; Stewardson, A.J.; Harbarth, S. Will 10 million people die a year due to antimicrobial resistance by 2050? PLoS Med. 2016, 13, e1002184. [Google Scholar] [CrossRef]
  9. Boucher, H.; Talbot, G.; Bradley, J.; Edwards, J.; Gilbert, D.; Rice, L.B.; Scheld, M.; Spellberg, B.; Bartlett, J. Bad bugs, no drugs: No ESKAPE! An update from the Infectious Diseases Society of America. Clin. Infect. Dis. 2009, 48, 1–12. [Google Scholar] [CrossRef]
  10. World Health Organization. Global Priority List of Antibiotic-Resistant Bacteria to Guide Research, Discovery, and Development of New Antibiotics. 2017. Available online: http://www.who.int/medicines/publications/WHO-PPL-Short_Summary_25Feb-ET_NM_WHO.pdf?ua_1 (accessed on 27 February 2017).
  11. Navon-Venezia, S.; Kondratyeva, K.; Carattoli, A. Klebsiella pneumoniae: A major worldwide source and shuttle for antibiotic resistance. FEMS Microbiol. Rev. 2017, 41, 252–275. [Google Scholar] [CrossRef]
  12. MacLean, R.; San Milla, A. The evolution of antibiotic resistance. Science 2019, 365, 1082–1083. [Google Scholar] [CrossRef] [PubMed]
  13. GBD 2019 Antimicrobial Resistance Collaborators. Global mortality associated with 33 bacterial pathogens in 2019: A systematic analysis for the Global Burden of Disease Study 2019. Lancet 2022, 400, 2221–2248. [Google Scholar] [CrossRef] [PubMed]
  14. Yigit, H.; Queenan, A.; Anderson, G.; Domenech-Sanchez, A.; Biddle, J.; Steward, C.; Alberti, S.; Bush, K.; Tenover, F.C. Novel carbapenem-hydrolyzing beta-lactamase, KPC-1, from a carbapenem-resistant strain of Klebsiella pneumoniae. Antimicrob. Agents Chemother. 2001, 45, 1151–1161. [Google Scholar] [CrossRef] [PubMed]
  15. Nordmann, P.; Poirel, L. Epidemiology and diagnostics of carbapenem resistance in Gram-negative bacteria. Clin. Infect. Dis. 2019, 69 (Suppl. 7), 521–528. [Google Scholar] [CrossRef] [PubMed]
  16. David, S.; Reuter, S.; Harris, S.R.; Glasner, C.; Feltwell, T.; Argimon, S.; EuSCAPE Working Group; ESGEM Study Group; Feil, E.J.; Rossolini, G.M.; et al. Epidemic of carbapenem-resistant Klebsiella pneumoniae in Europe is driven by nosocomial spread. Nat. Microbiol. 2019, 4, 1919–1929. [Google Scholar] [CrossRef] [PubMed]
  17. Savov, E.; Politi, L.; Spanakis, N.; Trifonova, A.; Kioseva, E.; Tsakris, A. NDM-1 Hazard in the Balkan States: Evidence of the First Outbreak of NDM-1-Producing Klebsiella pneumoniae in Bulgaria. Microb. Drug Resist. 2018, 24, 253–259. [Google Scholar] [CrossRef] [PubMed]
  18. Todorova, B.; Sabtcheva, S.; Ivanov, I.; Lesseva, M.; Chalashkanov, T.; Ioneva, M.; Bachvarova, A.; Dobreva, E.; Kantardjiev, T. First clinical cases of NDM-1-producing Klebsiella pneumoniae from two hospitals in Bulgaria. J. Infect. Chemother. 2016, 22, 837–840. [Google Scholar] [CrossRef] [PubMed]
  19. Politi, L.; Gartzonika, K.; Spanakis, N.; Zarkotou, O.; Poulou, A.; Skoura, L.; Vrioni, G.; Tsakris, A. Emergence of NDM-1-producing Klebsiella pneumoniae in Greece: Evidence of a widespread clonal outbreak. J. Antimicrob. Chemother. 2019, 74, 2197–2202. [Google Scholar] [CrossRef]
  20. Markovska, R.; Stoeva, T.; Boyanova, L.; Stankova, P.; Schneider, I.; Keuleyan, E.; Mihova, K.; Murdjeva, M.; Sredkova, M.; Lesseva, M.; et al. Multicentre investigation of carbapenemase-producing Klebsiella pneumoniae and Escherichia coli in Bulgarian hospitals—Interregional spread of ST11 NDM-1-producing K. pneumoniae. Infect. Genet. Evol. 2019, 69, 61–67. [Google Scholar] [CrossRef]
  21. Biswas, S.; Brunel, J.-M.; Dubus, J.-C.; Reynaud-Gaubert, M.; Rolain, J.-M. Colistin: An update on the antibiotic of the 21st century. Expert Rev. Anti-Infect. Ther. 2012, 10, 917–934. [Google Scholar] [CrossRef]
  22. Evans, M.E.; Feola, D.J.; Rapp, R.P. Polymyxin B sulfate and colistin: Old antibiotics for emerging multiresistant Gram-negative bacteria. Ann. Pharmacother. 1999, 33, 960–967. [Google Scholar] [CrossRef]
  23. Landman, D.; Georgescu, C.; Martin, D.A.; Quale, J. Polymyxins revisited. Clin. Microbiol. Rev. 2008, 21, 449–465. [Google Scholar] [CrossRef]
  24. Binsker, U.; Käsbohrer, A.; Hammerl, J.A. Global colistin use: A review of the emergence of resistant Enterobacterales and the impact on their genetic basis. FEMS Microbiol. Rev. 2022, 46, fuab049. [Google Scholar] [CrossRef]
  25. Savov, E.; Todorova, I.; Politi, L.; Trifonova, A.; Borisova, M.; Kioseva, E.; Tsakris, A. Colistin Resistance in KPC-2- and SHV-5-Producing Klebsiella pneumoniae Clinical Isolates in Bulgaria. Chemotherapy 2017, 62, 339–342. [Google Scholar] [CrossRef] [PubMed]
  26. Markovska, R.; Stoeva, T.; Schneider, I.; Boyanova, L.; Popova, V.; Dacheva, D.; Kaneva, R.; Bauernfeind, A.; Mitev, V.; Mitov, I. Clonal dissemination of multilocus sequence type ST15 KPC-2-producing Klebsiella pneumoniae in Bulgaria. APMIS 2015, 123, 887–894. [Google Scholar] [CrossRef] [PubMed]
  27. Marteva-Proevska, Y.; Velinov, T.; Markovska, R.; Dobrikova, D.; Pavlov, I.; Boyanova, L.; Mitov, I. Antibiotic combinations with colistin against carbapenem-resistant Klebsiella pneumoniae—In vitro assessment. J. IMAB 2018, 24, 2258–2266. [Google Scholar] [CrossRef]
  28. Markovska, R.; Marteva-Proevska, Y.; Velinov, T.; Pavlov, I.; Kaneva, R.; Boyanova, L. Detection of different colistin resistance mechanisms among multidrug-resistant Klebsiella pneumoniae isolates in Bulgaria. Acta Microbiol. Immunol. Hung. 2022, 69, 220–227. [Google Scholar] [CrossRef] [PubMed]
  29. Popescu, G.A.; Serban, R.; Niculcea, A.; Leustean, M.; Pistol, A. Consumul de antibiotice, Rezistența microbiană și Infecții Asociate Asistenței Medicale în România—2018. Available online: https://www.cnscbt.ro/index.php/analiza-date-supraveghere/infectii-nosocomiale-1/2025-consumul-de-antibiotice-rezistenta-microbiana-si-infectiile-asociate-asistentei-medicale-romania-2018/file (accessed on 18 March 2024).
  30. Galani, I.; Karaiskos, I.; Karantani, I.; Papoutsaki, V.; Maraki, S.; Papaioannou, V.; Kazila, P.; Tsorlini, H.; Charalampaki, N.; Toutouza, M.; et al. Epidemiology and resistance phenotypes of carbapenemase-producing Klebsiella pneumoniae in Greece, 2014 to 2016. Eurosurveillance 2018, 23, 2–13. [Google Scholar] [CrossRef]
  31. Cizmeci, Z.; Aktas, E.; Otlu, B.; Acikgoz, O.; Ordekci, S. Molecular characterization of carbapenem-resistant Enterobacteriaceae yields increasing rates of NDM-1 carbapenemases and colistin resistance in an OXA-48-endemic area. J. Chemother. 2017, 29, 344–350. [Google Scholar] [CrossRef]
  32. Palmieri, M.; D’Andrea, M.; Pelegrin, A.; Mirande, C.; Brkic, S.; Cirkovic, I.; Goossens, H.; Rossolini, G.M.; van Belkum, A. Genomic Epidemiology of Carbapenem- and Colistin-Resistant Klebsiella pneumoniae Isolates From Serbia: Predominance of ST101 Strains Carrying a Novel OXA-48 Plasmid. Front. Microbiol. 2020, 11, 294. [Google Scholar] [CrossRef]
  33. Pruss, A.; Kwiatkowski, P.; Masiuk, H.; Jursa-Kulesza, J.; Bilska, I.; Lubecka, A.; Cettler, M.; Roszkowska, P.; Dołęgowska, B. Analysis of the prevalence of colistin resistance among clinical strains of Klebsiella pneumoniae. Ann. Agric. Environ. Med. 2022, 29, 518–522. [Google Scholar] [CrossRef]
  34. Kaur, A.; Gandran, S.; Gupta, P.; Mehta, Y.; Laxminarayan, R.; Sengupta, S. Clinical outcome of dual colistin- and carbapenem-resistant Klebsiella pneumoniae bloodstream infections: A single-center retrospective study of 75 cases in India. Am. J. Infect. Control 2017, 45, 1289–1291. [Google Scholar] [CrossRef]
  35. Aydın, M.; Ergönül, Ö.; Azap, A.; Bilgin, H.; Aydın, G.; Çavuş, S.; Demiroğlu, Y.Z.; Alışkan, H.E.; Memikoğlu, O.; Menekşe, Ş.; et al. Rapid emergence of colistin resistance and its impact on fatality among healthcare-associated infections. J. Hosp. Infect. 2018, 98, 260–263. [Google Scholar] [CrossRef]
  36. The European Committee on Antimicrobial Susceptibility (EUCAST). Testing Breakpoint Tables for Interpretation of MICs and Zone Diameters. Version 13.1. 2023. Available online: www.eucast.org/clinical_breakpoints (accessed on 25 November 2023).
  37. Poirel, L.; Walsh, T.R.; Cuvillier, V.; Nordmann, P. Multiplex PCR for detection of acquired carbapenemase genes. Diagn. Microbiol. Infect. Dis. 2011, 70, 119–123. [Google Scholar] [CrossRef]
  38. Lescat, M.; Poirel, L.; Nordmann, P. Rapid multiplex PCR for detection of mcr-1 to -5 genes. Diagn. Microbiol. Infect. Dis. 2018, 92, 267–269. [Google Scholar] [CrossRef] [PubMed]
  39. Cannatelli, A.; Giani, T.; D’Andrea, M.M.; Di Pilato, V.; Arena, F.; Conte, V.; Tryfinopoulou, K.; Vatopoulos, A.; Rossolini, G.M. MgrB Inactivation Is a Common Mechanism of Colistin Resistance in KPC-Producing Klebsiella pneumoniae of Clinical Origin. Antimicrob. Agents Chemother. 2014, 58, 5696–5703. [Google Scholar] [CrossRef] [PubMed]
  40. Cerqueira, G.C.; Earl, A.M.; Ernst, C.M.; Grad, Y.H.; Dekker, J.P.; Feldgarden, M.; Chapman, S.B.; Reis-Cunha, J.L.; Shea, T.P.; Young, S.; et al. Multi-institute analysis of carbapenem resistance reveals remarkable diversity, unexplained mechanisms, and limited clonal outbreaks. Proc. Natl. Acad. Sci. USA 2017, 114, 1135–1140. [Google Scholar] [CrossRef] [PubMed]
  41. Cannatelli, A.; Di Pilato, V.; Giani, T.; Arena, F.; Ambretti, S.; Gaibani, P.; D’Andrea, M.M.; Rossolini, G.M. In vivo evolution to colistin resistance by PmrB sensor kinase mutation in KPC-producing Klebsiella pneumoniae is associated with low-dosage colistin treatment. Antimicrob. Agents Chemother. 2014, 58, 4399–4403. [Google Scholar] [CrossRef] [PubMed]
  42. Uzairue, L.I.; Rabaan, A.A.; Adewumi, F.A.; Okolie, O.J.; Folorunso, J.B.; Bakhrebah, M.A.; Garout, M.; Alfouzan, W.A.; Halwani, M.A.; Alamri, A.A.; et al. Global Prevalence of Colistin Resistance in Klebsiella pneumoniae from Bloodstream Infection: A Systematic Review and Meta-Analysis. Pathogens 2022, 11, 1092. [Google Scholar] [CrossRef] [PubMed]
  43. Labarca, J.; Poirel, L.; Ozdamar, M.; Turkoglu, S.; Hakko, E.; Nordmann, P. KPC-producing Klebsiella pneumoniae, finally targeting Turkey. New Microbes New Infect. 2014, 2, 50–51. [Google Scholar] [CrossRef] [PubMed]
  44. Macesic, N.; Nelson, B.; Mcconville, T.H.; Giddins, M.J.; Green, D.A.; Stump, S.; Gomez-Simmonds, A.; Annavajhala, M.K.; Uhlemann, A.-C. Emergence of Polymyxin Resistance in Clinical Klebsiella pneumoniae Through Diverse Genetic Adaptations: A Genomic, Retrospective Cohort Study. Clin. Infect. Dis. 2020, 70, 2084–2091. [Google Scholar] [CrossRef] [PubMed]
  45. Poirel, L.; Jayol, A.; Nordmann, P. Polymyxins: Antibacterial activity, susceptibility testing, 424 and resistance mechanisms encoded by plasmids or chromosomes. Clin. Microbiol. Rev. 2017, 30, 557–596. [Google Scholar] [CrossRef]
  46. Kim, S.J.; Shin, J.H.; Kim, H.; Ko, K.S. Roles of crrAB two-component regulatory system in Klebsiella pneumoniae: Growth yield, survival in initial colistin treatment stage, and virulence. Int. J. Antimicrob. Agents 2024, 63, 107011. [Google Scholar] [CrossRef]
  47. Yusof, N.; Norazzman, N.; Hakim, S.; Azlan, M.; Anthony, A.; Mustafa, F.; Ahmed, N.; Rabaan, A.A.; Almuthree, S.A.; Alawfi, A.; et al. Prevalence of Mutated Colistin-Resistant Klebsiella pneumoniae: A Systematic Review and Meta-Analysis. Trop. Med. Infect. Dis. 2022, 7, 414. [Google Scholar] [CrossRef] [PubMed]
  48. Bray, A.S.; Smith, R.D.; Hudson, A.W.; Hernandez, G.E.; Young, T.M.; George, H.E.; Ernst, R.K.; Zafar, M.A. MgrB-Dependent Colistin Resistance in Klebsiella pneumoniae Is Associated with an Increase in Host-to-Host Transmission. mBio 2022, 13, 0359521. [Google Scholar] [CrossRef] [PubMed]
  49. Alousi, S.; Saad, J.; Panossian, B.; Makkhlouf, R.; Al Khoury, C.; Rahy, K.; Thoumi, S.; Araj, G.F.; Khnayzer, R.; Tokajian, S. Genetic and structural basis of Colistin resistance in Klebsiella pneumoniae: Unraveling the molecular mechanisms. bioRxiv 2023. [Google Scholar] [CrossRef]
  50. Guducuoglu, H.; Gursoy, N.; Yakupogullari, Y.; Parlak, M.; Karasin, G.; Sunnetcioglu, M.; Otlu, B. Hospital Outbreak of a Colistin-Resistant, NDM-1- and OXA-48-Producing Klebsiella pneumoniae: High Mortality from Pandrug Resistance. Microb. Drug. Resist. 2017, 24, 966–972. [Google Scholar] [CrossRef] [PubMed]
  51. Petrosillo, N.; Taglietti, F.; Granata, G. Treatment Options for Colistin Resistant Klebsiella pneumoniae: Present and Future. J. Clin. Med. 2019, 8, 934. [Google Scholar] [CrossRef]
  52. Schreiber, P.; Sax, H.; Wolfensberger, A.; Clack, L.; Kuster, S. The preventable proportion of healthcare-associated infections 2005–2016: Systematic review and meta-analysis. Infect. Control Hosp. Epidemiol. 2018, 39, 1277–1295. [Google Scholar] [CrossRef]
  53. Worldwide Antimicrobial Resistance National/International Network Group (WARNING) Collaborators. Ten golden rules for optimal antibiotic use in hospital settings: The WARNING call to action. World J. Emerg. Surg. 2023, 18, 50. [Google Scholar] [CrossRef]
Table 1. Characteristics, treatments, and outcomes of the included patients.
Table 1. Characteristics, treatments, and outcomes of the included patients.
PatientsGenderAgeDiagnosisInterventionPrevious InterventionICU StayOutcomeTreatment
1f69Acute kidney bleedingLeft nephrectomy* RIRS + LT + JJ stent
cystoscopy
18.08–3.10diedamp/sulb, ceftriaxon, meropenem, colistin
2m71Pancreatic cancerTraverso-Longmirebile duct stent1–5.09
12.09–2.10
diedpiperacilin/tazobactam, cfp/sulb, linezolid, colistin
3m46Pancreatic cancerTraverso-Longmirebile duct stent18–19.08
12–14.09
dischargedpiperacilin/tazobactam, ciprofloxacin, colistin
4m82Urine bladder cancerCystectomy-28.08–14.09diedamp/sulb, levofloxacin, doxycycline, colisitn
5m56Perforated duodenal ulcerSuture-5–15.09dischargedmeropenem
Abbreviations: * RIRS—retrograde intrarenal surgery; LT—lithotripsy; amp/sulb—ampicillin/sulbactam; cfp/sulb—cefoperazone/sulbactam.
Table 2. Source and characteristics of the isolates.
Table 2. Source and characteristics of the isolates.
PatientIsolateSampleDate of IsolationCarbapenemaseERIC *mcr 1–5 mgrB
11 wound31.08OXA-48-like and NDM-likeANEGNEG
2 blood culture8.09OXA-48-like and NDM-like ANEGNEG
23 wound3.09OXA-48-likeBNEGPOS
4 ascites12.09OXA-48-likeBNEGPOS
35wound12.09OXA-48-likeBNEGPOS
46 tracheo-bronchial4.09OXA-48-likeBNEGPOS
7 blood 13.09OXA-48-like BNEGPOS
58 urine11.09OXA-48-like and NDM-likeANEGNEG
Abbreviations: * ERIC—Enterobacterial Repetitive Intergenic Consensus, NEG—negative, POS—positive.
Table 3. The rate of colistin resistance in carbapenem-resistant K. pneumoniae in Balkan countries.
Table 3. The rate of colistin resistance in carbapenem-resistant K. pneumoniae in Balkan countries.
AuthorCountry, Study Period% of MDR
Markovska, et al. [28]Bulgaria, 2017–201837
Galani, et al. [30]Greece, 2014–201640.4
epi-net.eu/records/12313/12313/[29]Romania, 201827.5
Palmieri, et al. [32]Serbia, 2013–201710.6
Cizmeci, et al. [31]Türkiye, 201627.5
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Popivanov, G.; Markovska, R.; Gergova, I.; Konaktchieva, M.; Cirocchi, R.; Kjossev, K.; Mutafchiyski, V. An Intra-Hospital Spread of Colistin-Resistant K. pneumoniae Isolates—Epidemiological, Clinical, and Genetic Analysis. Medicina 2024, 60, 511. https://doi.org/10.3390/medicina60030511

AMA Style

Popivanov G, Markovska R, Gergova I, Konaktchieva M, Cirocchi R, Kjossev K, Mutafchiyski V. An Intra-Hospital Spread of Colistin-Resistant K. pneumoniae Isolates—Epidemiological, Clinical, and Genetic Analysis. Medicina. 2024; 60(3):511. https://doi.org/10.3390/medicina60030511

Chicago/Turabian Style

Popivanov, Georgi, Rumyana Markovska, Ivanka Gergova, Marina Konaktchieva, Roberto Cirocchi, Kirien Kjossev, and Ventsislav Mutafchiyski. 2024. "An Intra-Hospital Spread of Colistin-Resistant K. pneumoniae Isolates—Epidemiological, Clinical, and Genetic Analysis" Medicina 60, no. 3: 511. https://doi.org/10.3390/medicina60030511

Article Metrics

Back to TopTop