Next Article in Journal
Antiprotozoal Nor-Triterpene Alkaloids from Buxus sempervirens L.
Previous Article in Journal
Super-Cationic Peptide Dendrimers—Synthesis and Evaluation as Antimicrobial Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molecular Characterization of KPC-2-Producing Enterobacter cloacae Complex Isolates from Cali, Colombia

1
Grupo de Investigación en Microbiología, Industria y Ambiente (GIMIA), Facultad de Ciencias Básicas, Universidad Santiago de Cali, Cali 760035, Colombia
2
Clínica Imbanaco, Cali 760042, Colombia
3
Laboratorio de Salud Pública Departamental, Secretaria Departamental de Salud del Valle del Cauca, Gobernación del Valle del Cauca, Cali 760045, Colombia
4
Grupo de Investigación en Química y Biotecnología (QUIBIO), Facultad de Ciencias Básicas, Universidad Santiago de Cali, Cali 760035, Colombia
*
Author to whom correspondence should be addressed.
Antibiotics 2021, 10(6), 694; https://doi.org/10.3390/antibiotics10060694
Submission received: 27 April 2021 / Revised: 31 May 2021 / Accepted: 31 May 2021 / Published: 10 June 2021

Abstract

:
The Enterobacter cloacae complex is an emerging opportunistic pathogen whose increased resistance to carbapenems is considered a public health problem. This is due to the loss of efficacy of beta-lactam antibiotics, which are used as the first treatment option in the management of infections caused by Gram-negative bacteria. The objective of this study was to perform the molecular characterization of 28 isolates of the E. cloacae complex resistant to cephalosporins and carbapenems isolated between 2011 and 2018 from five hospitals located in the municipality of Santiago de Cali, Colombia. Molecular detection of blaKPC, blaVIM, blaNDM and blaOXA-48-like genes was performed on these isolates and the genetic relationship between the isolates was assessed using multilocus sequence typing (MLST). Forty-three percent of the isolates carried the blaKPC-2 gene variant. MLST showed high genetic diversity among isolates, the most frequent being the sequence type ST510 with a frequency of 50%. The identification of the genes involved in carbapenem resistance and dispersing genotypes is an important step toward the development of effective prevention and epidemiological surveillance strategies in Colombian hospitals.

1. Introduction

The Enterobacter cloacae complex includes several species (Enterobacter asburiae, Enterobacter carcinogenus, Enterobacter cloacae, Enterobacter hormaechei, Enterobacter kobei, Enterobacter nimipressuralis, Enterobacter ludwigii and Enterobacter mori) [1,2] that are opportunistic pathogens, which belong to the Family Enterobacteriaceae and are considered one of the causal agents of healthcare-associated infections (HAIs) both internationally [3,4,5,6] and nationally [7,8]. Because carbapenems are used to treat infections caused by extended-spectrum beta-lactamase (ESBL)-producing E. cloacae complex, the increase in carbapenem resistance in this species is a matter of concern. Studies worldwide report that carbapenem resistance in this species is mainly associated with the production of Klebsiella pneumoniae carbapenemases (KPCs) [9,10,11,12,13,14]; however, there are other resistance mechanisms associated with this phenotype such as changes in permeability of cell membrane, efflux pumps and hyper-expression of AmpC [15,16]. The National Institute of Health in Colombia reports that, between 2012 and 2018, 90% of Enterobacteriaceae isolates causing HAIs were resistant to carbapenems [17]. Of these, 17% are E. cloacae complex isolates, being the second most frequently reported. Ninety four percent of these were carbapenems resistant and 67% were KPC-producers, followed by VIM with 3.5% and NDM with 2.2%. This highlights the importance of KPC-producing E. cloacae complex in Colombia [8,17]. Although other studies in Colombia have reported carbapenemase-producing E. cloacae complex isolates in the country [7,18,19], there is no updated data specific to the region. Hence, the molecular characterization of KPC-producing E. cloacae complex isolated from five tertiary care hospitals located in the municipality of Santiago de Cali from 2011 to 2018 has been performed in the present study.

2. Results

2.1. Collection of E. cloacae Complex Clinical Isolates

According to the selection criteria (see Materials and Methods), 28 clinical isolates of E. cloacae complex were selected and showing a phenotypic profile of total or intermediate resistance to third and fourth generation cephalosporins and to at least one carbapenem (Table 1). Twenty-one percent (6/28) were isolated from a public tertiary care hospital located in commune 19 (Figure 1, Table 1). The remaining 79% (22/28) were isolated from four private tertiary care hospitals, distributed as follows: 54% (15/28) from clinic 1 (commune 19), 14% (4/28) from clinic 2 (commune 17), 7% (2/28) from clinic 3 (commune 2) and 4% (1/28) from clinic 4 (commune 1) (Figure 1, Table 1). All the communes are in the municipality of Santiago de Cali, Department of Valle del Cauca, in southwest Colombia (Figure 1).
Additionally, Table 1 presents the information related to the clinical isolates (year of isolation and hospital of origin) and the demographic data associated with the samples (age and sex of the patient and site of infection).

2.2. Detection of Genes Encoding Beta-Lactamase Enzymes

The blaKPC-2 gene was detected in 43% (12/28) of the isolates (Table 2). None of them carried blaVIM, blaNDM, or blaOXA-48-like genes that code for other carbapenemases.

2.3. Clinical and Demographic Characteristics of the E. cloacae Complex Isolates Carrying the blaKPC-2 Gene

We analyzed the twelve E. cloacae complex isolates found to carry the blaKPC-2 gene, which were isolated from five tertiary care hospitals. Twenty-five percent (3/12) of the isolates were obtained from a public hospital between 2017 and 2018. The remaining 75% (9/12) were obtained from private entities and were distributed as follows: 42% (5/12) from clinic 1 between 2013 and 2017; 17% (2/12) from clinic 2 in 2017 and 8% (1/12) from clinics 3 and 4 in 2017 and 2018, respectively.
The demographic characteristics are summarized in Table 3. Eighty-three percent of infected patients were male and 32% were elderly. The most common site of infection was blood, affecting 42% of patients (Table 3).

2.4. Molecular Genotyping of KPC-Producing E. cloacae Complex Isolates

The genotyping analysis of the 12 evaluated KPC-producing E. cloacae complex isolates yielded five STs, namely ST510 (50%, 6/12), ST45 (17%, 2/12), ST456 (17%, 2/12), ST1483 (8%, 1/12) and ST513 (8%, 1/12) (Figure 2, Table 2).
Two STs were found in two different hospitals. The first one was ST510, which was identified in four isolates from clinic 1 and two isolates from the public entity. Both were in commune 19 (Figure 1), which indicates that ST510 circulated from 2013 to 2018 in the southwest of the municipality of Santiago de Cali. The second ST was ST456, which was found in one isolate from clinic 3 (commune 2) and one isolate from clinic 4 (commune 1) (Figure 1). Therefore, ST456 circulated in the northwest of the city between 2017 and 2018.
The other three STs were found in a single hospital, with one isolate each. Thus, ST1483 was found in clinic 1 (commune 19) in 2014, ST513 in the public entity (commune 19) in 2017 and ST45 in clinic 2 (commune 17) in 2017.
The results obtained using the goeBURST algorithm indicated high genetic diversity because there were no single, double, or triple locus variants. However, ST510 and ST45 are satellites, because they share three of the seven genes with each other (dnaA, fusA and rplB) (Figure 3, Table 3), whereas ST510 and ST45 share the rplB allele with ST1483 (Figure 3, Table 3). Additionally, there were two STs, ST513 and ST456, that remained unclustered, i.e., as singletons (Figure 3, Table 3).

3. Discussion

According to our results, 43% (12/28) of the carbapenem-resistant E. cloacae complex isolates evaluated in this study carried the variant of the blaKPC-2 gene. This percentage is lower than that reported by De la Cadena et al. (2017), who also performed a molecular characterization of 28 strains of carbapenem-resistant E. cloacae complex isolated between 2009 and 2013 from eight cities in Colombia. These authors reported that 100% (12/12) of the isolates from Cali carried the blaKPC-2 gene [7]. In another report by Rada et al. (2020), in which the dynamics of blaKPC-2 gene transmission through plasmids was studied in strains of Enterobacteriaceae in the city of Medellín, four isolates of the E. cloacae complex with ST456 were found to be circulating in three hospitals between 2013 and 2015 [18]. Additionally, Vanegas et al. (2016) studied the molecular epidemiology of Gram-negative bacilli resistant to carbapenems, isolated from a pediatric population in five hospitals in Medellín between 2012 and 2014. Out of the 24 E. cloacae complex isolates evaluated, one was found to carry the blaKPC-3 gene, although its ST was not reported [19]. Specifically, Colombia has been described as an endemic country for the gene that codes for KPC, with the KPC-2 and KPC-3 variants—those that circulate in the country—being more frequently detected in K. pneumoniae [11,20,21].
Regarding the STs detected in this study, 50% of the KPC-producing E. cloacae complex isolates belonged to ST510 and circulated in Cali from 2013 to 2018. These results agree with those of De la Cadena et al. (2017), who also reported ST510 as the most frequent ST found in the cities of Cali, Pereira and Medellín between 2009 and 2013. In particular, 80% of the E. cloacae complex isolates found in Cali carried the blaKPC-2 gene [7]. According to the MLST database of E. cloacae, apart from Colombia, ST510 has only been reported in Japan in 2016.
ST456 was found in 16.7% of the E. cloacae complex isolates carrying the blaKPC-2 gene, which circulated in Cali in 2017 and 2018. These results coincide with those reported by Rada et al. (2020), who found four E. cloacae complex isolates with ST456 carrying the blaKPC-2 gene in three hospitals in Medellín from 2013 to 2015 [18]. ST456 has been previously reported in a E. cloacae complex isolate carrying the blaKPC-2 gene in Norway in 2017 [22] and, according to the MLST database of E. cloacae, in other countries such as India (2013), Ghana (2015), Togo (2016) and the United States (2016 and 2018).
This is the first study reporting the presence of ST45, ST513 and ST1483 in Colombia. ST45 was found in 16.7% of the isolates circulating in Cali in 2017. This ST has been previously reported in Spain, in an E. cloacae isolate carrying the blaOXA-48-like gene [23], and in China, in three CTXM-9- and SHV-12-producing isolates [24]. According to the MLST database of E. cloacae, ST45 has also been reported in Japan in 2013.
ST513 was found in 8.3% of the E. cloacae complex isolates carrying the blaKPC-2 gene and circulated in Cali in 2017. ST513 has been previously reported in Vietnam, in a colistin-resistant E. cloacae isolate that carried the blaNDM-1 gene and was isolated from a male patient in 2010 [25]. According to the MLST database of E. cloacae, ST513 has also been reported in Japan in 2016. Finally, ST1483 was found in 8.3% of the E. cloacae complex isolates carrying the blaKPC-2 gene, which circulated in Cali in 2014 and, according to the MLST database of E. cloacae, has also been reported by the Pasteur Institute of Guadeloupe in 2020.
The analysis of demographic data indicated that the KPC-producing E. cloacae complex mostly affected males (83%) and the elderly (32%), causing bloodstream infections (BSIs) (42%). In particular, BSIs are frequently acquired in hospital facilities and are considered a serious clinical condition that can worsen the prognosis of sepsis, leading to extended hospital stay, higher care costs, as well as increased morbidity and mortality [26]. According to the World Health Organization, 8.7% of nosocomial infections correspond to bacteremia and have gained importance in Europe, North America and Latin America, because they are usually caused by Gram-negative bacteria resistant to the available antibiotics [27]. For this reason, different antibiotics have been used to treat these microorganisms, including ceftazidime/avibactam which has shown activity against Enterobacteriaceae isolates producing carbapenemases from classes A and D, turning into a highly useful tool for the management of infections by multidrug-resistant Enterobacteriaceae [28]. However, resistance to this antibiotic has already been reported in class C β-lactamases [29] and KPC-producing Enterobacteriaceae [30]. Due to this, the therapeutic option will depend on the sensitivity profile of the isolate. This was the case of a 52-year-old female with infected arthritis of the right shoulder, whose joint aspirate culture showed a cefazolin-resistant E. cloacae. She was treated with levofloxacin, and she stopped experiencing shoulder swelling and severe pain [31]. Precisely, because sometimes there are few therapeutic options available to treat infections, in vitro studies using essential oils from plants have shown to have activity against multidrug-resistant Enterobacteriaceae isolates [32,33], which would be a basis for the development of new and effective antibacterial treatments.
In Colombia, bacterial isolates carrying the genes coding for metallo-beta-lactamases blaVIM [34,35,36,37] and blaNDM [38,39,40,41] have been reported. However, none of these genes have been previously found in E. cloacae isolates, which is in line with the results obtained in this study. This is also true for the blaOXA-48-like gene, which was found for the first and only time in 2016 in Medellín, in an elderly patient infected with OXA-48-producing K. oxytoca [42], but not in the E. cloacae complex.
It is worth noting that none of the evaluated genes was detected in 57% of the isolates included in this study. Nonetheless, they were resistant to cephalosporins and at least to one carbapenem, which indicates that there is some other mechanism causing the resistance phenotype. Other carbapenemase not tested in this study, such as IMP or alteration or loss of nonspecific porins and hyper production of intrinsic, chromosomally encoded AmpC-type beta-lactamases, could cause such resistance [20]. Additional studies are required to elucidate these mechanisms.
The results show that, in addition to K. pneumoniae and E. coli, there are other Enterobacteriaceae such as KPC-2-producing E. cloacae complex, which are emerging as opportunistic pathogens resistant to carbapenems in Cali, Colombia. Therefore, it provides valuable information to further reinforce the epidemiological surveillance in the municipality of Santiago de Cali. Additionally, it was observed that some lineages are maintained over time not only in the region, but also in the country. This may be due to the referral of patients, which in turn makes the spread of these resistance mechanisms possible. It is crucial to pursue the molecular characterization of this bacterial species in different hospitals in the Department of Valle del Cauca and the rest of Colombia to design effective prevention and epidemiological surveillance strategies at the local, regional, and national levels. This will allow the establishment and implementation of institutional policies for the rational use of antibiotics in the hospitals of Colombia to prevent the spread of KPC-producing bacteria to all health institutions in the country.

4. Materials and Methods

4.1. Selection Criteria for E. cloacae Complex Isolates and Antibiotic Sensitivity Tests

In this retrospective and cross-sectional study, twenty-eight E. cloacae complex clinical isolates were sampled from 2011 to 2018. All isolates showing a phenotypic profile of total or intermediate resistance to third and fourth generation cephalosporins and to at least one carbapenem were selected according to the Clinical and Laboratory Standards Institute 2018 criteria [43]. Each of the clinical laboratories at the hospitals that participated in the study performed the identification and antibiotic sensitivity test using the automated system VITEK® (BioMérieux, Marcy l’E’toile, France).

4.2. Detection of Genes Encoding Beta-Lactamase Enzymes

E. cloacae complex isolates were grown on MacConkey agar and incubated at 37 °C overnight. A single colony was resuspended in 100 µL of distilled water and bacteria were lysed in a water bath at 100 °C for 10 min. Cellular debris was removed by centrifugation at 13,000 rpm for 10 min and the supernatant was used as DNA template to perform polymerase chain reaction (PCR) [44]. To determine the quality of the cell lysate, we amplified a partial sequence of the 16S rRNA genes using universal primers U1 (5′-CCAGCAGCCGCGGTAATACG-3′) and U2 (5′-ATCGG(C/T)TACCTTGTTACGACTTC-3′), according with Lu et al., 2000 [44]. Then, the genes blaKPC, blaNDM, blaVIM and blaOXA-48-like were amplified by PCR using the primers shown in Table 4.
PCR was performed in an Eppendorf 950,000,040 Mastercycler thermal cycler using specific programs based on the hybridization temperature of the primers, as well as the size of the expected amplicons (Table 4). For this, 2X PCR 100 mixture (OPTBM-00006, CorpoGen, Bogotá, Colombia) was used following the manufacturer’s instructions. The reaction mix was prepared as follows: 1X master mix, 0.2 µM each of forward and reverse primers, <250 ng DNA and sterile water to make up the final volume to 25 µL. The following strains were used as positive controls for PCR: K. pneumoniae carrying the blaKPC gene, K. pneumoniae carrying the blaOXA-48-like gene, P. aeruginosa carrying the blaVIM gene and P. aeruginosa carrying the blaNDM gene. Sterile water was used to replace the volume corresponding to the DNA in negative controls.
Following PCR, the reaction products were purified using the commercial Qiaquick PCR Spin columns kit (Qiagen, Hilden, Germany) and were sequenced with the Sanger method in Macrogen, Korea. The sequences obtained were compared with the data from the National Center for Biotechnology Information database (https://blast.ncbi.nlm.nih.gov/Blast.cgi, accessed on 10 January 2021) and Beta-Lactamase Data Resources (https://www.ncbi.nlm.nih.gov/pathogens/beta-lactamase-data-resources/ accessed on 11 January 2021).

4.3. Molecular Genotyping of the KPC-Producing E. cloacae Complex Isolates

The genetic relationships between the KPC-producing E. cloacae complex isolates were determined using the PCR amplification technique multilocus sequence typing (MLST), according to Miyoshi-Akiyama et al. [49]. Following PCR, the amplified fragments were sequenced by Macrogen, Korea. Alleles and sequence types (STs) were assigned using the MLST website of E. cloacae (https://pubmlst.org/organisms/enterobacter-cloacae, accessed on 15 January 2021).
To establish the relationships between STs, the goeBURST algorithm [50] of the PHYLOViZ software (http://www.phyloviz.net/, accessed on 25 April 2021) was used.

5. Conclusions

This study identified the resistance mechanism and circulating genotypes of E. cloacae complex strains resistant to carbapenems isolated from five tertiary care hospitals in the municipality of Santiago de Cali. In 43% of the cases, the resistance to these antibiotics was due to the production of KPCs. Nonetheless, there are other mechanisms not detected in this study that confer resistance to beta-lactams. With regard to the circulating genotypes in the city, it was found that they are diverse and that ST510 and ST456 showed the highest proportion between 2011 and 2018. Moreover, these STs have been reported in other cities of the country with the greatest degree of spread. The other STs (ST45, ST513 and ST1483) are reported in this study for the first time in Colombia. The results obtained in this study demonstrate that, in addition to K. pneumoniae and E. coli, there are other Enterobacteriaceae, such as the E. cloacae complex carrying the variant of the blaKPC-2 gene, which could be emerging as carbapenem-resistant opportunistic pathogens in Cali, Colombia.

Author Contributions

Conceptualization, A.F. and A.C.; methodology, A.F. and A.C.; software, A.F.; validation, S.R. and A.C.; formal analysis, A.F., A.C. and C.A.; investigation, A.F., D.G., I.G., A.C., S.R., M.B.O. and C.A.; resources, A.F., A.C. and S.R.; data curation, A.F.; writing—original draft preparation, A.F.; writing—review and editing, A.F., A.C., S.R. and C.A.; visualization, A.F.; supervision, A.F.; project administration, A.F.; funding acquisition, A.F., A.C. and S.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was funded by the Dirección General de Investigaciones of the Universidad Santiago de Cali, grant number 934-621118-9, and under call No. 01-2021.

Institutional Review Board Statement

The study was conducted according to the guidelines of the Declaration of Helsinki and approved by Ethics Committee of Centro Médico Imbanaco (Act number 226 of 1 August 2019).

Informed Consent Statement

Not applicable.

Data Availability Statement

The data used to support the findings of this study are included within the article and are available from the corresponding author upon request.

Acknowledgments

We thank Elsa de La Cadena for her contributions in reviewing this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mustafa, A.; Ibrahim, M.; Rasheed, M.A.; Kanwal, S.; Hussain, A.; Sami, A.; Ahmed, R.; Bo, Z. Genome-wide Analysis of Four Enterobacter cloacae complex type strains: Insights into Virulence and Niche Adaptation. Sci. Rep. 2020, 10, 8150. [Google Scholar] [CrossRef]
  2. Davin-Regli, A.; Lavigne, J.P.; Pagès, J.M. Enterobacter spp.: Update on taxonomy, clinical aspects, and emerging antimicrobial resistance. Clin. Microbiol. Rev. 2019, 32, 1–32. [Google Scholar] [CrossRef]
  3. Yang, X.; Guo, R.; Xie, B.; Lai, Q.; Xu, J.; Hu, N.; Wan, L.; Dai, M.; Zhang, B. Drug resistance of pathogens causing nosocomial infection in orthopedics from 2012 to 2017: A 6-year retrospective study. J. Orthop. Surg. Res. 2021, 16, 100. [Google Scholar] [CrossRef] [PubMed]
  4. Cruz-López, F.; Villarreal-Treviño, L.; Morfin-Otero, R.; Martinez-Melendez, A.; Camacho-Ortiz, A.; Rodríguez-Noriega, E.; Garza-González, E. Dynamics of colonization in patients with health care-associated infections at step-down care units from a tertiary care hospital in Mexico. Am. J. Infect. Control. 2020, 48, 1329–1335. [Google Scholar] [CrossRef]
  5. Annavajhala, M.K.; Gomez-Simmonds, A.; Uhlemann, A.-C. Multidrug-Resistant Enterobacter cloacae Complex Emerging as a Global, Diversifying Threat. Front. Microbiol. 2019, 10, 44. [Google Scholar] [CrossRef] [Green Version]
  6. Sievert, D.M.; Ricks, P.; Edwards, J.R.; Schneider, A.; Patel, J.; Srinivasan, A.; Kallen, A.; Limbago, B.; Fridkin, S.; National Healthcare Safety Network; et al. Antimicrobial-resistant pathogens associated with healthcare-associated infections: Summary of data reported to the National Healthcare Safety Network at the Centers for Disease Control and Prevention, 2009–2010. Infect. Control. Hosp. Epidemiol. 2013, 34, 1–14. [Google Scholar] [CrossRef] [PubMed]
  7. De la Cadena, E.; Adriana, C.; Sebastian, M.J.; Rojas, L.J.; Cristhian, H.-G.; Pallares, C.; Perez, F.; Bonomo, R.A.; Villegas, M.V.; Pallares, C. Molecular characterisation of carbapenem-resistant Enterobacter cloacae complex in Colombia: Bla KPC and the “changing landscape.”. J. Glob. Antimicrob. Resist. 2018, 13, 184–189. [Google Scholar] [CrossRef]
  8. Salud I nacional de. Resultados Del Programa de Vigilancia Por Laboratorio De Resistencia Antimicrobiana En Infecciones Asociadas a La Atención En Salud (IAAS) 2016. Epidemiología de las Infecciones Asociadas a la Atención en Salud. 2017; 1–43. Available online: https://www.ins.gov.co/BibliotecaDigital/informe-vigilancia-por-laboratorio-resistencia-antimicrobiana-y-whonet-iaas-2016.pdf#search=Resultados%20Del%20Programa%20de%20Vigilancia%20Por%20Laboratorio%20De%20Resistencia%20Antimicrobiana%20En%20Infecciones%20Asociadas%20a%20La%20Atenci%C3%B3n%20En%20Salud%20%28IAAS%29%202016 (accessed on 8 March 2021).
  9. Rosa, J.F.; Rizek, C.; Marchi, A.P.; Guimaraes, T.; Miranda, L.; Carrilho, C.; Levin, A.S.; Costa, S.F. Clonality, outer-membrane proteins profile and efflux pump in KPC- producing Enterobacter sp. in Brazil. BMC Microbiol. 2017, 17, 1–9. [Google Scholar] [CrossRef] [Green Version]
  10. Sacsaquispe-Contreras, R.; Bailón-Calderón, H. Original Breve Identificación De Genes De Resistencia A Carbapenémicos En Enterobacterias De Hospitales Identification of Carbapenem-Resistant Genes IN. Rev. Peru Med. Exp. Salud Publica 2018, 35, 2013–2017. [Google Scholar]
  11. Logan, L.K.; Weinstein, R.A. The Epidemiology of Carbapenem-Resistant Enterobacteriaceae: The Impact and Evolution of a Global Menace. J. Infect. Dis. 2017, 215, S28–S36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Chavda, K.D.; Chen, L.; Fouts, D.E.; Sutton, G.; Brinkac, L.; Jenkins, S.G.; Bonomo, R.A.; Adams, M.D.; Kreiswirth, B.N. Comprehensive Genome Analysis of Carbapenemase-Producing Enterobacter spp.: New Insights into Phylogeny, Population Structure, and Resistance Mechanisms. mBio 2016, 7, e02093-16. [Google Scholar] [CrossRef] [Green Version]
  13. Girlich, D.; Poirel, L.; Nordmann, P. Clonal distribution of multidrug-resistant Enterobacter cloacae. Diagn. Microbiol. Infect. Dis. 2015, 81, 264–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kiedrowski, L.M.; Guerrero, D.M.; Perez, F.; Viau, R.A.; Rojas, L.J.; Mojica, M.F.; Rudin, S.D.; Hujer, A.M.; Marshall, S.H.; Bonomo, R.A. Carbapenem-ResistantEnterobacter cloacaeIsolates Producing KPC-3, North Dakota, USA. Emerg. Infect. Dis. 2014, 20, 1583–1585. [Google Scholar] [CrossRef] [Green Version]
  15. Doumith, M.; Ellington, M.J.; Livermore, D.M.; Woodford, N. Molecular mechanisms disrupting porin expression in ertapenem-resistant Klebsiella and Enterobacter spp. clinical isolates from the UK. J. Antimicrob. Chemother. 2009, 63, 659–667. [Google Scholar] [CrossRef] [PubMed]
  16. Bradford, P.A.; Urban, C.; Mariano, N.; Projan, S.J.; Rahal, J.J.; Bush, K. Imipenem resistance in Klebsiella pneumoniae is associated with the combination of ACT-1, a plasmid-mediated AmpC beta-lactamase, and the foss of an outer membrane protein. Antimicrob. Agents Chemother. 1997, 41, 563–569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Salud I nacional de. Informe De Resultados De La Vigilancia Por Laboratorio De Resistencia Antimicrobiana En Infecciones Asociadas a La Atención En Salud (IAAS) 2017. Epidemiol las Infecc Asoc a la Atención en Salud. 2018; 1–31. Available online: https://www.ins.gov.co/buscador-eventos/Informacin%20de%20laboratorio/Informe%20Vigilancia%20por%20Laboratorio%20Resistencia%20Antimicrobiana%20y%20Whonet%20IAAS%202017.pdf#search=Informe%20de%20Resultados%20de%20la%20Vigilancia%20por%20Laboratorio%20de%20Resistencia%20antimicrobiana%20en%20Infecciones%20Asociadas%20a%20la%20Atenci%C3%B3n%20en%20Salud%20%28IAAS%29%202017 (accessed on 8 March 2021).
  18. Rada, A.M.; De La Cadena, E.; Agudelo, C.; Capataz, C.; Orozco, N.; Pallares, C.; Dinh, A.Q.; Panesso, D.; Ríos, R.; Diaz, L.; et al. Dynamics of blaKPC-2 Dissemination from Non-CG258 Klebsiella pneumoniae to Other Enterobacterales via IncN Plasmids in an Area of High Endemicity. Antimicrob. Agents Chemother. 2020, 64, 64. [Google Scholar] [CrossRef]
  19. Vanegas, J.M.; Parra, O.L.; Jiménez, J.N. Molecular epidemiology of carbapenem resistant gram-negative bacilli from infected pediatric population in tertiary—care hospitals in Medellín, Colombia: An increasing problem. BMC Infect. Dis. 2016, 16, 1–10. [Google Scholar] [CrossRef] [Green Version]
  20. Escandón-Vargas, K.; Reyes, S.; Gutiérrez, S.; Villegas, M.V. The epidemiology of carbapenemases in Latin America and the Caribbean. Expert Rev. Anti. Infect. Ther. 2016, 15, 277–297. [Google Scholar] [CrossRef]
  21. Pacheco, R.; Osorio, L.; Correa, A.M.; Villegas, M.V. Prevalencia de bacterias Gram negativas portadoras del gen blaKPC en hospitales de Colombia. Biomédica 2013, 34, 81. [Google Scholar] [CrossRef] [Green Version]
  22. Samuelsen, Ø.; Overballe-Petersen, S.; Bjørnholt, J.V.; Brisse, S.; Doumith, M.; Woodford, N.; Hopkins, K.; Aasnæs, B.; Haldorsen, B.; Sundsfjord, A. Molecular and epidemiological characterization of carbapenemase-producing Enterobacteriaceae in Norway, 2007 to 2014. PLoS ONE 2017, 12, e0187832. [Google Scholar] [CrossRef] [Green Version]
  23. Fernández, J.F.; Montero, I.; Martinez, O.F.; Fleites, A.; Poirel, L.; Nordmann, P.; Rodicio, M.D.R.R. Dissemination of multiresistant Enterobacter cloacae isolates producing OXA-48 and CTX-M-15 in a Spanish hospital. Int. J. Antimicrob. Agents 2015, 46, 469–474. [Google Scholar] [CrossRef] [Green Version]
  24. Cao, X.-L.; Cheng, L.; Zhang, Z.-F.; Ning, M.-Z.; Zhou, W.-Q.; Zhang, K.; Shen, H. Survey of Clinical Extended-Spectrum Beta-Lactamase-Producing Enterobacter cloacae Isolates in a Chinese Tertiary Hospital, 2012–2014. Microb. Drug Resist. 2017, 23, 83–89. [Google Scholar] [CrossRef]
  25. Le-Ha, T.D.; Le, L.; Le-Vo, H.N.; Anda, M.; Motooka, D.; Nakamura, S.; Tran, L.K.; Tran, P.T.-H.; Lida, T.; Cao, V. Characterization of a carbapenem- and colistin-resistant Enterobacter cloacae carrying Tn6901 in blaNDM-1 genomic context. Infect. Drug Resist. 2019, 12, 733–739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. De La Rosa, G.; León, A.L.; Jaimes, F. Epidemiología y pronóstico de pacientes con infección del torrente sanguíneo en 10 hospitales de Colombia. Rev. Chil Infectol. 2016, 33, 141–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Cortés, J.A.; Leal, A.L.; Montañez, A.M.; Buitrago, G.; Castillo, J.S.; Guzman, L. Frequency of microorganisms isolated in patients with bacteremia in intensive care units in Colombia and their resistance profiles. Braz. J. Infect. Dis. 2013, 17, 346–352. [Google Scholar] [CrossRef] [Green Version]
  28. Martínez-Miranda, R.; Gastélum-Acosta, M.; Guerrero-Estrada, P.; Ayala-Figueroa, R.I.; Osuna-Álvarez, L.E. Actividad antimicrobiana de ceftolozano-tazobactam y ceftazidima-avibactam contra bacilos gramnegativos clínicamente relevantes aislados en México. Gac Med. Mex 2020, 156, 592–597. [Google Scholar] [CrossRef]
  29. Shields, R.K.; Nguyen, M.H.; Chen, L.; Press, E.G.; Kreiswirth, B.N.; Clancy, C.J. Pneumonia and Renal Replacement Therapy Are Risk Factors for Ceftazidime-Avibactam Treatment Failures and Resistance among Patients with Carbapenem-Resistant Enterobacteriaceae Infections. Antimicrob. Agents Chemother. 2018, 62, e02497-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Shields, R.K.; Iovleva, A.; Kline, E.G.; Kawai, A.; McElheny, C.L.; Doi, Y. Clinical evolution of AmpC-mediated ceftazidime-avibactam and cefiderocol resistance in enterobacter cloacae complex following exposure to cefepime. Clin. Infect. Dis. 2020, 71, 2713–2716. [Google Scholar] [CrossRef]
  31. Huang, J.; Xu, Q.; Liu, F.; Xiong, H.; Yang, J. Enterobacter cloacae infection of the shoulder in a 52-year-old woman without apparent predisposing risk factor: A case report and literature review. BMC Infect. Dis. 2021, 21, 1–9. [Google Scholar] [CrossRef] [PubMed]
  32. Le, N.T.; Donadu, M.G.; Ho, D.V.; Doan, T.Q.; Le, A.T.; Raal, A.; Usai, D.; Sanna, G.; Marchetti, M.; Usai, M.; et al. Biological activities of essential oil extracted from leaves of Atalantia sessiflora Guillauminin Vietnam. J. Infect. Dev. Ctries 2020, 14, 1054–1064. [Google Scholar] [CrossRef]
  33. Benameur, Q.; Gervasi, T.; Pellizzeri, V.; Pľuchtová, M.; Tali-Maama, H.; Assaous, F.; Guettou, B.; Rahal, K.; Gruľová, D.; Dugo, G.; et al. Antibacterial activity of Thymus vulgaris essential oil alone and in combination with cefotaxime against blaESBL producing multidrug resistant Enterobacteriaceae isolates. Nat. Prod. Res. 2018, 33, 2647–2654. [Google Scholar] [CrossRef]
  34. Crespo, M.P.; Woodford, N.; Sinclair, A.; Kaufmann, M.E.; Turton, J.; Glover, J.; Velez, J.D.; Castaneda, C.R.; Recalde, M.; Livermore, D.M. Outbreak of Carbapenem-Resistant Pseudomonas aeruginosa Producing VIM-8, a Novel Metallo- -Lactamase, in a Tertiary Care Center in Cali, Colombia. J. Clin. Microbiol. 2004, 42, 5094–5101. [Google Scholar] [CrossRef] [Green Version]
  35. Villegas, M.V.; Lolans, K.; Olivera, M.D.R.; Suarez, C.J.; Correa, A.; Queenan, A.M.; Quinn, J.P. First Detection of Metallo-β-Lactamase VIM-2 in Pseudomonas aeruginosa Isolates from Colombia. Antimicrob. Agents Chemother. 2006, 50, 226–229. [Google Scholar] [CrossRef] [Green Version]
  36. Montealegre, M.C.; Correa, A.; Briceño, D.F.; Rosas, N.C.; De La Cadena, E.; Ruiz, S.J.; Mojica, M.F.; Camargo, R.D.; Zuluaga, I.; Marin, A.; et al. Novel VIM Metallo-β-Lactamase Variant, VIM-24, from a Klebsiella pneumoniae Isolate from Colombia. Antimicrob. Agents Chemother. 2011, 55, 2428–2430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Pacheco, T.; Bustos-Cruz, R.H.; Abril, D.; Arias, S.; Uribe, L.; Rincón, J.; García, J.-C.; Escobar-Perez, J. Pseudomonas aeruginosa Coharboring BlaKPC-2 and BlaVIM-2 Carbapenemase Genes. Antibiotics 2019, 8, 98. [Google Scholar] [CrossRef] [Green Version]
  38. Pérez, J.A.E.; Escobar, N.M.O.; Castro-Cardozo, B.; Márquez, I.A.V.; Aguilar, M.I.G.; De La Barrera, L.M.; Barreto, E.R.B.; Marquez-Ortiz, R.A.; Guayazán, M.V.M.; Gómez, N.V. Outbreak of NDM-1-Producing Klebsiella pneumoniae in a Neonatal Unit in Colombia. Antimicrob. Agents Chemother. 2013, 57, 1957–1960. [Google Scholar] [CrossRef] [Green Version]
  39. Ministerio de Salud de Colombia. Informe Quincenal Epidemiológico Nacional. Inst Nac Salud Colomb 2017, 22, 65–77. Available online: https://www.ins.gov.co/buscador-eventos/IQEN/IQEN (accessed on 16 March 2021).
  40. Saavedra-Rojas, S.-Y.; Duarte-Valderrama, C.; González-De-Arias, M.-N.; Ovalle-Guerro, M.V. Emergencia de Providencia rettgeri NDM-1 en dos departamentos de Colombia, 2012–2013. Enferm. Infecc. y Microbiol. Clínica 2017, 35, 354–358. [Google Scholar] [CrossRef] [PubMed]
  41. Rojas, L.J.; Wright, M.S.; De La Cadena, E.; Motoa, G.; Hujer, K.M.; Villegas, M.V.; Adams, M.D.; Bonomo, R.A. Initial assessment of the molecular epidemiology of blaNDM-1 in Colombia. Antimicrob. Agents Chemother. 2016, 60, 4346–4350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Vanegas, J.M.; Ospina, W.P.; Higuita-Gutiérrez, L.F.; Jiménez, J.N.; Higuita, L.F. First reported case of an OXA-48-producing isolate from a Colombian patient. J. Glob. Antimicrob. Resist. 2016, 6, 67–68. [Google Scholar] [CrossRef]
  43. Clinical and Laboratory Standards Institute. Performance Standards for Antimicrobial Susceptibility Testing: 28th Edition Informational Supplement M100-S27; CLSI: Wayne, PA, USA, 2018. [Google Scholar]
  44. Lu, J.J.; Perng, C.L.; Lee, S.Y.; Wan, C.C. Use of PCR with universal primers and restriction endonuclease digestions for detection and identification of common bacterial pathogens in cerebrospinal fluid. J. Clin. Microbiol. 2000, 38, 2076–2080. [Google Scholar] [CrossRef]
  45. Yigit, H.; Queenan, A.M.; Anderson, G.J.; Doménech-Sánchez, A.; Biddle, J.W.; Steward, C.D.; Alberti, S.; Bush, K.; Tenover, F.C. Novel Carbapenem-Hydrolyzing β-Lactamase, KPC-1, from a Carbapenem-Resistant Strain of Klebsiella pneumoniae. Antimicrob. Agents Chemother. 2001, 45, 1151–1161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Liu, Z.; Li, W.; Wang, J.; Pan, J.; Sun, S.; Yu, Y.; Zhao, B.; Ma, Y.; Zhang, T.; Qi, J.; et al. Identification and Characterization of the First Escherichia coli Strain Carrying NDM-1 Gene in China. PLoS ONE 2013, 8, e66666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Toleman, M.A.; Simm, A.M.; Murphy, T.A.; Gales, A.C.; Biedenbach, D.J.; Jones, R.N.; Walsh, T.R. Molecular characterization of SPM-1, a novel metallo-β-lactamase isolated in Latin America: Report from the SENTRY antimicrobial surveillance programme. J. Antimicrob Chemother. 2002, 50, 673–679. [Google Scholar] [CrossRef]
  48. Poirel, L.; Castanheira, M.; Carrër, A.; Rodriguez, C.P.; Jones, R.N.; Smayevsky, J.; Nordmann, P. OXA-163, an OXA-48-Related Class D β-Lactamase with Extended Activity Toward Expanded-Spectrum Cephalosporins. Antimicrob. Agents Chemother. 2011, 55, 2546–2551. [Google Scholar] [CrossRef] [Green Version]
  49. Miyoshi-Akiyama, T.; Hayakawa, K.; Ohmagari, N.; Shimojima, M.; Kirikae, T. Multilocus Sequence Typing (MLST) for Characterization of Enterobacter cloacae. PLoS ONE 2013, 8, e66358. [Google Scholar] [CrossRef] [Green Version]
  50. Francisco, A.P.; Bugalho, M.; Ramirez, M.; Carriço, J.A. Global optimal eBURST analysis of multilocus typing data using a graphic matroid approach. BMC Bioinform. 2009, 10, 152. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Location of the hospitals from which E. cloacae complex isolates were obtained. Four of them are private (in yellow) and one is public (in red); all are in four communes of the municipality of Santiago de Cali, Department of Valle del Cauca, Colombia (Made through https://www.visme.co/, accessed on 29 March 2021).
Figure 1. Location of the hospitals from which E. cloacae complex isolates were obtained. Four of them are private (in yellow) and one is public (in red); all are in four communes of the municipality of Santiago de Cali, Department of Valle del Cauca, Colombia (Made through https://www.visme.co/, accessed on 29 March 2021).
Antibiotics 10 00694 g001
Figure 2. Percentage of sequence types (STs) of the E. cloacae complex isolates carrying the blaKPC-2 gene circulating in the municipality of Santiago de Cali from 2013 to 2018.
Figure 2. Percentage of sequence types (STs) of the E. cloacae complex isolates carrying the blaKPC-2 gene circulating in the municipality of Santiago de Cali from 2013 to 2018.
Antibiotics 10 00694 g002
Figure 3. Cluster of the E. cloacae complex isolates carrying the blaKPC-2 gene generated with the goeBURST algorithm.
Figure 3. Cluster of the E. cloacae complex isolates carrying the blaKPC-2 gene generated with the goeBURST algorithm.
Antibiotics 10 00694 g003
Table 1. Antibiotic susceptibility profile and demographic data of the E. cloacae complex isolates included in this study.
Table 1. Antibiotic susceptibility profile and demographic data of the E. cloacae complex isolates included in this study.
KeyYear IsolateCeftazidimeCefepimeCeftriaxoneErtapenemMeropenemOriginInfectionAge (Years)Gender
14162011>644>644<1Clinic 1Urine81F
16662012168>64>81Clinic 1Bloodstream9M
20722012162>644>16Clinic 1Urine41M
13642013>64>64>64>8>16Clinic 1Bloodstream49M
22052013>642>6442Clinic 1Urine6M
22492013162421Clinic 1Urine84F
13592014>64>64>64>8>16Clinic 1Skin93F
429020143232>64>8>16Clinic 1Bloodstream23M
43472014>64>64>64>8>16Clinic 1Urine67M
47302015>64>64>64>8>16Clinic 1Ulcer20F
5047201616>64>64>8>16Clinic 1Urine17M
52272017>64>64>64>8>16Clinic 1Wound58M
3822017>642>644<1PublicBloodstream76F
474201732163242PublicPeritoneal fluid22M
38120171616>3241PublicBloodstream74M
7132017164>4>18PublicPeritoneal fluid54M
3322017162>6422Clinic 2Bloodstream21 daysM
3312017>64>64>642<1Clinic 2Bloodstream30M
3332017>64>64>64>8>16Clinic 2Bloodstream47 daysM
3302017>64>64>64>8>16Clinic 2Bloodstream12 daysM
3052017>64>64>64>8>16Clinic 3Sore30M
312017>642>6442Clinic 3Wound20F
22018>64>64>64>8>16Clinic 4Urine culture70M
1,510,0062018>64>64>64>8>16PublicCatheter62M
62018>64>64>64>8>16PublicBloodstream62M
53852018>64>64>64>8>16Clinic 1Sore55M
54382018162≥6441Clinic 1Urine18M
5521-122018162≥6441Clinic 1Urine1F
Table 2. Genetic relationship of E. cloacae complex isolates carrying the blaKPC variant.
Table 2. Genetic relationship of E. cloacae complex isolates carrying the blaKPC variant.
KeyAlleleSTdnaAfusAgyrBleuSpyrGrplBrpoB
1364blaKPC-25104442091714115
1359blaKPC-214833762194445433
4290blaKPC-25104442091714115
4730blaKPC-25104442091714115
5227blaKPC-25104442091714115
474blaKPC-25131711190168122113
333blaKPC-245441463946
330blaKPC-245441463946
305blaKPC-2456149446118015211
2blaKPC-2456149446118015211
1,510,006blaKPC-25104442091714115
6blaKPC-25104442091714115
Abbreviations: Sequence type (ST).
Table 3. Demographic characteristics corresponding to the E. cloacae complex isolates carrying the blaKPC gene.
Table 3. Demographic characteristics corresponding to the E. cloacae complex isolates carrying the blaKPC gene.
Characteristic% of Isolates
Gender
Female17% (2/12)
Male83% (10/12)
Age
Newborn (0–30 days)8% (1/12)
Infant (1–12 months)8% (1/12)
Teenagers (12–20 years)8% (1/12)
Young adult (21–40 years)16% (2/12)
Middle adult (41–60 years)24% (3/12)
Elderly (>60 years)32% (4/12)
Infection site
Bloodstream42% (5/12)
Skin8.3% (1/12)
Ulcer8.3% (1/12)
Wound8.3% (1/12)
Peritoneal fluid8.3% (1/12)
Sore8.3% (1/12)
Urine culture8.3% (1/12)
Catheter8.3% (1/12)
Table 4. Primers used for polymerase chain reaction of the carbapenem-resistant E. cloacae complex isolates.
Table 4. Primers used for polymerase chain reaction of the carbapenem-resistant E. cloacae complex isolates.
Primer NamePrimer SequenceProduct Size (bp)Annealing Temperature (°C)References
KPC F5-TGTCACTGTATCGCCGTC-389453Yigit y col., 2001 [45]
KPC R5-CTCAGTGCTCTACAGAAAACC-3
NDM F5-ATGGAATTGCCCAATATTATGC-381352Liu y col., 2013 [46]
NDM R5-TCAGCGCAGCTTGTCGGCCAT-3
VIM F5-GTCTATTTGACCGCGTC-377552 Toleman y col., 2002 [47]
VIM R5-CTACTCAACGACTGAGCG-3
OXA-48 F5-TATATTGCATTAAGCAAGGG-384856Poirel y col., 2011 [48]
OXA-48 R5-CACACAAATACGCGCTAACC-3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Falco, A.; Guerrero, D.; García, I.; Correa, A.; Rivera, S.; Olaya, M.B.; Aranaga, C. Molecular Characterization of KPC-2-Producing Enterobacter cloacae Complex Isolates from Cali, Colombia. Antibiotics 2021, 10, 694. https://doi.org/10.3390/antibiotics10060694

AMA Style

Falco A, Guerrero D, García I, Correa A, Rivera S, Olaya MB, Aranaga C. Molecular Characterization of KPC-2-Producing Enterobacter cloacae Complex Isolates from Cali, Colombia. Antibiotics. 2021; 10(6):694. https://doi.org/10.3390/antibiotics10060694

Chicago/Turabian Style

Falco, Aura, Daniela Guerrero, Isabella García, Adriana Correa, Sandra Rivera, María Beatriz Olaya, and Carlos Aranaga. 2021. "Molecular Characterization of KPC-2-Producing Enterobacter cloacae Complex Isolates from Cali, Colombia" Antibiotics 10, no. 6: 694. https://doi.org/10.3390/antibiotics10060694

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop