Next Article in Journal
Understanding Knowledge and Attitude of Farmers towards Antibiotic Use and Antimicrobial Resistance in Jhunjhunu District, Rajasthan India
Next Article in Special Issue
Exploring Structure–Activity Relationships of Niclosamide-Based Colistin Potentiators in Colistin-Resistant Gram-Negative Bacteria
Previous Article in Journal
Lipid-Centric Approaches in Combating Infectious Diseases: Antibacterials, Antifungals and Antivirals with Lipid-Associated Mechanisms of Action
Previous Article in Special Issue
Comparison between Colistin and Polymyxin B in the Treatment of Bloodstream Infections Caused by Carbapenem-Resistant Pseudomonas aeruginosa and Acinetobacter baumannii-calcoaceticus Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Meropenem Pharmacodynamics and Emergence of Resistance against Carbapenem-Susceptible Non-Carbapenemase-Producing and Carbapenemase-Producing Enterobacterales: A Pharmacodynamic Study in a Hollow-Fiber Infection Model

by
Maria V. Golikova
1,*,
Kamilla N. Alieva
1,
Elena N. Strukova
1,
Daria A. Kondratieva
1,
Nika F. Petrova
2,
Mayya A. Petrova
2 and
Stephen H. Zinner
3
1
Department of Pharmacokinetics & Pharmacodynamics, Gause Institute of New Antibiotics, 11 Bolshaya Pirogovskaya Street, 119021 Moscow, Russia
2
National Research Centre “Kurchatov Institute”, 123182 Moscow, Russia
3
Harvard Medical School, Department of Medicine, Mount Auburn Hospital, 330 Mount Auburn Street., Cambridge, MA 02138, USA
*
Author to whom correspondence should be addressed.
Antibiotics 2023, 12(12), 1717; https://doi.org/10.3390/antibiotics12121717
Submission received: 10 November 2023 / Revised: 7 December 2023 / Accepted: 8 December 2023 / Published: 12 December 2023
(This article belongs to the Special Issue Antibiotics Use and Therapy in Gram-Negative Bacterial Infection)

Abstract

:
Resistance to carbapenems has become a problem due to Klebsiella pneumoniae (K. pneumoniae), harboring carbapenemases. Among them, there are isolates that are recognized as carbapenem-susceptible; however, these carbapenemase-producing strains with low meropenem minimal inhibitory concentrations (MICs) may pose a threat to public health. We aimed to investigate the impact of the ability to produce carbapenemases by a bacterial isolate on the effectiveness of meropenem in the hollow-fiber infection model. K. pneumoniae and Escherichia coli (E. coli) strains with equal meropenem MICs but differing in their ability to produce carbapenemases were used in pharmacodynamic simulations with meropenem. In addition to standard MIC determination, we assessed the MICs against tested strains at high inoculum density to test if the inoculum effect occurs. According to pharmacodynamic data, the carbapenemase-producing strains were characterized with a relatively decreased meropenem effectiveness compared to non-producers. Meanwhile, the effect of meropenem perfectly correlated with the meropenem exposure expressed as the DOSE/MIC ratio when high-inoculum (HI) MICs but not standard-inoculum (SI) MICs were used for regression analysis. It could be concluded that meropenem-susceptible carbapenemase-producing strains may not respond to meropenem therapy; the antibiotic inoculum effect (IE) may have a prognostic value to reveal the meropenem-susceptible Enterobacterales that harbor carbapenemase genes.

1. Introduction

Beta-lactam antibiotics are among the most commonly prescribed antimicrobials for a wide variety of clinical indications. Among them, carbapenems are effective against many organisms and are typically used to treat complicated infections. Unfortunately, over the last decade, resistance to carbapenems among Gram-negative bacteria has become a major problem, especially carbapenemase-producing K. pneumoniae [1,2,3]. These organisms can spread easily among hospitalized patients and can transmit carbapenemase genes located on different mobile genetic elements (plasmids, transposons, and insertion sequences) to other Gram-negative bacteria by horizontal gene transfer [4,5,6]. The most common carbapenemase genes are blaKPC-2 and blaOXA-48 [7,8]. Among carbapenemase producers circulating in hospitals, isolates characterized by carbapenem MICs up to the 2 mg/L susceptibility breakpoint are recognized as carbapenem-susceptible by European Committee on Antimicrobial Susceptibility Testing [9] and Clinical & Laboratory Standards Institute [10] guidelines. EUCAST guidelines do not consider whether or not the key pathogen is able to produce carbapenemases; if a given isolate is categorized as susceptible but produces carbapenemases, “it still should be reported as tested”. It is not clear if carbapenemase-producing strains with low meropenem MICs actually pose a threat to public health. As far as we know, the studies aimed at these issues are lacking. Previously, our group conducted a pharmacodynamic study in a hollow-fiber infection model (HFIM) with meropenem and carbapenemase-producing K. pneumoniae strains with a wide range of meropenem susceptibility, including strains with MICs below the 2 mg/L breakpoint [11]. As a result, these susceptible strains displayed intensive regrowth and decreased meropenem susceptibility during meropenem exposure, despite simulating the maximal dose of 2 g over a 3 h infusion.
In the current study, we aimed to investigate the impact of carbapenemase production by bacterial isolates with meropenem susceptibility or intermediate susceptibility, on the effectiveness of meropenem in the HFIM. The HFIM is an effective tool to study antimicrobial efficacy and has many applications, including studying the emergence of bacterial resistance and estimating the clinical efficacy of antimicrobial drugs [12,13]. To address this goal, well-characterized parental carbapenemase non-producing K. pneumoniae ATCC 700603 and E. coli C600 strains with extremely low meropenem MICs were used as recipients in mating experiments. The clinical isolates containing plasmids carrying the blaOXA-48 and blaKPC-2 genes of carbapenemases from D and A classes were used as donors. The resulting transconjugants were able to produce carbapenemases but retained meropenem MICs below the resistance breakpoint. In addition, we used the same parental K. pneumoniae ATCC 700603 and E. coli C600 strains to obtain less meropenem-susceptible variants with non-carbapenemase-producing resistance mechanisms through serial passage on meropenem-containing media. Thus, we obtained derivative strains with equal meropenem MICs but different resistance mechanisms that were used as model strains in pharmacodynamic simulations with meropenem. This allowed us to compare the influence of carbapenemase production by bacterial strains with comparable meropenem susceptibility on meropenem pharmacodynamics and the emergence of resistance. In addition to standard MIC determinations, we assessed meropenem activity against strains at high inoculum density to test if the inoculum effect is present [14]. It is well known that beta-lactam antibiotics are frequently prone to the IE if a bacterial strain produces beta-lactamase. In this study, peritoneal fluid (PF) meropenem pharmacokinetics [15] was simulated following the administration of 0.5 g every 8 h in a 0.5 h infusion during 5-day treatments in an HFIM.
The investigation of issues, referring to the role of carbapenemase production by the meropenem-susceptible strains in their ability to resist meropenem and worse treatment outcome, determines the clinical significance of the current study.

2. Results

2.1. Transconjugant and Mutant K. pneumoniae and E. coli Strains, Meropenem Susceptibility at Standard and High Inocula

The flowchart depicted in Figure 1 displays all manipulations followed to obtain model bacterial strains for the pharmacodynamic study. In the first step, we obtained transconjugants of the K. pneumoniae ATCC 700603 and E. coli C600 strains carrying plasmids containing two different carbapenemase genes. After mating of the recipient strain E. coli C600 with the donor strains K. pneumoniae 485 or 565, the respective E. coli tcC600/485 and tcC600/565 transconjugants were obtained. Meropenem MICs for these strains were 0.5 and 1 mg/L, respectively. After mating of the recipient strain K. pneumoniae ATCC 700603 and the donor strains K. pneumoniae 485 or 565, respective K. pneumoniae tc700603/485 and tc700603/565 transconjugants were obtained. Meropenem MICs for these strains were 0.5 and 4 mg/L, respectively (Figure 1).
In the second step, through passaging on meropenem-containing media, we selected mutant K. pneumoniae ATCC 700603 and E. coli C600 variants (from the parent E. coli C600 and K. pneumoniae ATCC 70060) with the same meropenem MICs as the transconjugants: (1) E. coli mC600/0.5 with meropenem MIC of 0.5 mg/L (the same MIC as E. coli tcC600/485); (2) E. coli mC600/1 with meropenem MIC of 1 mg/L (the same MIC as E. coli tcC600/565); (3) K. pneumoniae m700603/0.5 with meropenem MIC of 0.5 mg/L (the same MIC as K. pneumoniae tc700603/485); and (4) K. pneumoniae m700603/4 with meropenem MIC of 4 mg/L (the same MIC as K. pneumoniae tc700603/565).
As a result, we obtained eight bacterial strains with different meropenem MICs and the presence or absence of carbapenemase genes that were used in further pharmacodynamic simulations. Along with the meropenem MICs determined at SI, to determine if the tested bacterial strains are prone to the inoculum effect (IE), we also estimated MICs at HI. The resultant meropenem MICs for all tested strains at SI and HI are summarized in Table 1. As seen from Table 1, the susceptibility to meropenem did not depend on the inoculum density for the E. coli C600 and its carbapenemase-free mutants. The IE was weakly pronounced for the K. pneumoniae ATCC 700603 strain and its carbapenemase-free mutants. However, in the presence of carbapenemase genes, the meropenem susceptibility at HI was even more decreased. The blaOXA-48 gene affected meropenem HI MICs 2–4 times less than blaKPC-2.

2.2. Meropenem Pharmacodynamics with E. coli and K. pneumoniae Total and Resistant Subpopulations

At first, the pharmacodynamic experiments with all parental strains exposed to meropenem were conducted. Figure 2 shows the results of pharmacodynamic simulations with meropenem against the highly susceptible E. coli C600 and K. pneumoniae ATCC 700603 strains, as well as meropenem-resistant K. pneumoniae 485 and 565 strains. Not surprisingly, with the two carbapenemase-non-producing strains, the rapid and complete elimination of bacteria was observed. Conversely, carbapenemase-producers were characterized by pronounced growth and the concomitant selection of meropenem-resistant cells even though meropenem was infused into the system.
The results of simulations with transconjugant and mutant strains of E. coli and K. pneumoniae are shown in Figure 3. With both E. coli mC600/0.5 and tcC600/485 that have equivalent meropenem MICs of 0.5 mg/L, the antibiotic demonstrated a considerable effect, which was more pronounced with the mutant strain E. coli mC600/0.5; at the end of the simulations, both bacterial strains were eliminated. With E. coli mC600/1 and tcC600/565 (MICs of 1 mg/L), meropenem demonstrated different effects. The carbapenemase-non-producing mC600/1 was fully suppressed, while the numbers of the KPC-producing tcC600/565 increased rapidly after an initial 6 h decline; bacterial re-growth was accompanied by the intensive selection of meropenem-resistant cells (Figure 3).
Although K. pneumoniae m700603/0.5 and tc700603/485 had identical meropenem MICs of 0.5 mg/L, they behaved differently under antibiotic exposure. Numbers of the mutant carbapenemase-non-producing strain m700603/0.5 decreased gradually throughout the observation period, but the transconjugant OXA-48-producing strain tc700603/485 showed a modest meropenem effect and regrowth, and the selection of meropenem-resistant cells was observed after 24 h. Against K. pneumoniae m700603/4 and tc700603/565 with MICs of 4 mg/L, meropenem had a minimal effect, and both strains behaved similarly with rapid regrowth after an initial decline and concomitant selection of meropenem-resistant cells.

2.3. DOSE/MIC Relationships with the Meropenem Effect and Emergence of Resistance

To quantitatively compare the meropenem effect against bacteria with equal antibiotic MICs but different carbapenemase production, the integral parameter of the effect, ABBC, was calculated (Figure 4).
As seen in the figure, ABBCs that correspond to the mutant carbapenemase-non-producing E. coli and K. pneumoniae strains are higher than for transconjugant carbapenemase-producing strains. This difference in meropenem effect is obvious with the mC600/1–tcC600/565 pair. Acquisition by the recipient C600 strain of a plasmid carrying the KPC-carbapenemase genes dramatically affects the strain’s ability to resist meropenem.
To analyze if differences in the meropenem effect (expressed as ABBC) against strains with equal MICs but different antibiotic resistance mechanisms can be predicted using MICs determined at HI, relationships of “DOSE/MIC—ABBC” were constructed for SI (for comparison) and HI (Figure 5).
The “DOSE/MIC—ABBC” relationship is described by a sigmoid function and was stronger when HI meropenem MICs were used for the DOSE/MIC calculation compared with SI MICs: r2 0.98 versus 0.72, respectively. Using the DOSE/MICSI index, the correlation was reduced due to the stratification of data points for carbapenemase-producing and non-producing strains. When MICs at HI were used for the analysis, this stratification was not observed.
The relationships of DOSE/MIC index with resistance were also constructed for cells at all resistance levels (from 2× to 16 × MIC) using both SI and HI MICs and were described with the Gaussian function. As an example, data for cells resistant to 4 × MIC meropenem are shown in Figure 6. The relationships for E. coli and K. pneumoniae cells resistant to 2×, 8×, and 16×MIC had similar patterns. The use of HI MICs allowed the demonstration of a stronger relationship of meropenem resistance with exposure compared with SI MICs: r2 0.96 versus 0.54, respectively.

3. Discussion

The present study was designed to investigate if the effectiveness of meropenem against meropenem-susceptible Enterobacterales can be influenced by the presence or absence of bacterial carbapenemase genes in a certain strain. We used well-characterized parental highly meropenem-susceptible carbapenemase-free E. coli and K. pneumoniae strains to obtain pairs of model isogenic strains with equal meropenem MICs but differing in the presence of carbapenemases. To obtain carbapenemase-producing strains, horizontal gene transfer of the conjugative plasmids with carbapenemase genes was performed in mating experiments with donor clinical K. pneumoniae strains. As a result, we obtained four transconjugant strains, two each of E. coli and K. pneumoniae, differing by levels of meropenem susceptibility (Figure 1). Although donor K. pneumoniae strains were highly resistant to meropenem (MICs of 32 and 64 mg/L), after plasmid transfer to the recipient strains, the resultant plasmid-related meropenem resistance was low (meropenem MICs from 0.5 to 4 mg/L). The plasmid containing the blaOXA-48 carbapenemase gene provided meropenem MICs of 0.5 mg/L in both recipient strains, whereas the plasmid containing blaKPC-2 provided meropenem MICs of 1 and 4 mg/L in E. coli and K. pneumoniae recipient strains, respectively. The low meropenem MICs against strains with the transferred blaOXA-48 gene can be explained by the fact that the enzyme encoded by the gene has weak carbapenemase activity [15]. The KPC-carbapenemases are more powerful carbapenem-hydrolyzing enzymes [16] and provided higher meropenem MICs than the OXA-48-carbapenemase. Apparently, the high meropenem MICs of donor K. pneumoniae strains are due to carbapenemase production and alternative heterogeneous resistance mechanisms (i.e., reduced outer membrane permeability [17], efflux pump activation [18]).
Assuming the given meropenem MICs in transconjugants, the respective mutant carbapenemase-free strains of E. coli and K. pneumoniae were selected on meropenem-containing media. Then, all these strains were used in the comparative pharmacodynamic study in HFIM. In the HFIM, we simulated the pharmacokinetic profile of meropenem in peritoneal fluid to mimic an intraabdominal infection [15]. We chose this scenario because this infection provides optimal conditions for horizontal gene transfer during this intra-abdominal infection usually caused by Enterobacterales [19]. A series of kinetic time-kill curves displaying the course of both total and resistant bacterial populations were obtained for each strain. The meropenem antibacterial effect was assessed with the ABBC integral parameter; this allowed quantitative comparisons of meropenem effectiveness among related strains with equal meropenem MICs but differing by the ability of carbapenemase production. As seen in Figure 3, the ability of a strain to produce carbapenemases determines its survival capability during antibiotic exposure. This can be most clearly seen in the example of the E. coli transconjugant strain tcC600/565 with a meropenem MIC of 1 mg/L, which exhibited extremely intensive growth under meropenem exposure, unlike its carbapenemase-free pair strain mC600/1 that was eliminated. Therefore, it can be concluded that the specific resistance mechanism related to carbapenemase production plays a key role in the effectiveness of the antibiotic even against susceptible strains; this is not obvious with standard MIC testing. However, HI MIC testing revealed dramatic differences between strains that had equal SI MICs. All carbapenemase producers had high HI MICs and demonstrated the IE. It is worth noting that the K. pneumoniae ATCC 700603 strain was also characterized with the IE despite it being a non-carbapenemase producer. However, this is rather an exception as, in general, carbapenemase non-producers are rarely associated with IE [14,20]. The prognostic value of HI MICs in terms of meropenem effectiveness was confirmed with the “exposure—effect” and “exposure—resistance” relationships, which were much stronger with HI than SI MICs (Figure 5 and Figure 6).
This study provides evidence in support of the utility of detecting carbapenemase genes even in meropenem-susceptible Enterobacterales, especially in intra-abdominal infections. In addition, determining carbapenem MICs at high bacterial inocula to study the IE is a useful, inexpensive, and easy-to-perform option to detect organisms that harbor carbapenemases. This is especially important given that the IE is clearly linked to bacterial beta-lactamase production.
Numerous clinical studies have investigated carbapenem efficacy in infections caused by carbapenemase-producing and non-producing Enterobacterales [21,22,23,24,25]. However, these studies did not highlight strains with meropenem MICs below the resistance breakpoint (2–4 mg/L) that might also harbor carbapenemase genes. Attempts to distinguish carbapenemase-producers and non-producers using meropenem MICs have been made and the MIC breakpoint of 2 mg/L was proposed (i.e., strains with MICs above or below the 2 mg/L are accounted as producers or not, respectively) [26]. Unfortunately, isolates with meropenem MICs below 2 mg/L that carry carbapenemase genes are frequently found. Moreover, there is a hidden threat that initially highly susceptible non-carbapenemase-producing organisms might acquire plasmids that carry carbapenemase genes. As we have shown in this study, strains having no resistance mechanisms other than carbapenemase production may actually have meropenem MICs below the susceptibility breakpoint. Obviously, these transconjugant cells may not respond to meropenem therapy since they can produce carbapenemases.
Our study is a pilot and therefore did not include a large number of E. coli and K. pneumoniae strains with a wide range of MICs or other carbapenemase classes, for example, metallo-beta-lactamases. Additional pharmacodynamic studies with E. coli and K. pneumoniae and other clinically significant Gram-negative bacteria (e.g., Pseudomonas aeruginosa, Acinetobacter baumannii) are necessary to fully evaluate the role of carbapenemase-production on the antimicrobial effect of beta-lactam. Subsequent studies with a wider range of beta-lactam antibiotics would enhance the generalizability of our results.

4. Materials and Methods

4.1. Antimicrobial Agent, Bacterial Isolates, and Susceptibility Testing

Meropenem powder was purchased from Sigma-Aldrich (St. Louis, MO, USA). Two clinical isolates of K. pneumoniae, 485 (meropenem MIC = 32 mg/L; isolated in Saint-Petersburg in 2012 year) and 565 (meropenem MIC = 64 mg/L; isolated in Saint-Petersburg in 2011 year), carrying plasmids with blaOXA-48 (POXAAPSS2, IncL, KU159086.1) and blaKPC-2 (PKPCAPSS plasmid, IncFII, KP008371, [27]) carbapenemase genes, respectively, were used as plasmid donors in mating experiments. Strains of E. coli K-12 C600 and K. pneumoniae ATCC 700603 were used as recipients in mating experiments and as parental strains in meropenem resistance selection. Before and after each testing or experiment, carbapenemase production was verified for each bacterial strain using a modified carbapenem-inactivation method [28].
Susceptibility testing was carried out using broth microdilution techniques with a standard inoculum of approximately 5 × 105 CFU/mL (SI) and a high inoculum of approximately 5 × 107 CFU/mL (HI). Meropenem MICs at SI were determined according to standard recommendations using cation-supplemented Mueller–Hinton broth (CSMHB) (Becton Dickinson, Franklin Lakes, NJ, USA) [29]. Before reading, plates were incubated at 37 °C for 18–20 h. In case of determination of MICs at HI, bacterial growth in the trays was quantified by optical density at 600 nm (OD) before and after 18 h of incubation at 37 °C. MIC was taken as the minimal dilution at which the 18 h OD was equal to or less than that at time 0. Inoculum effect was defined as an eightfold or greater increase in MIC when tested with HI compared to at SI. MIC values in each case were obtained at least in triplicate, and modal MICs were estimated.

4.2. Mating Experiments

For mating experiments to distinguish between donor and transconjugant cells, the recipient strains of K. pneumoniae ATCC 700603 and E. coli C600 were selected on the media with rifampicin to produce rifampicin-resistant variants (rifampicin MICs of 256 mg/L). For all experiments, these rifampicin-resistant K. pneumoniae ATCC 700603 and E. coli C600 strains were used.
Bacteria were grown in Luria Broth (LB, Becton Dickinson, Franklin Lakes, NJ, USA) and Luria Agar (LA, Becton Dickinson, Franklin Lakes, NJ, USA) Media [30] at 37 °C. When required, LB Agar was supplemented with antimicrobial agents at the following final concentrations (mg/mL): meropenem (0.5–2) and rifampicin (150).
Matings were performed overnight on the LA plates. Briefly, the 1:1 mixture of the donor and recipient in the late logarithmic growth phase was plated on the LA surface and incubated at 37 °C for 18–20 h. The mixed growth was then scraped from the plate surface and then resuspended in 1 mL of saline, and to quantify the numbers of donor, recipient, and transconjugant cells, the cell mixture was diluted as appropriate and 100 µL samples were spread on appropriate selective plates with respective antibiotics. Parent strains were plated in parallel with the matings and then processed similarly to the matings as controls. Isolated colonies from matings presumed as transconjugant and parental strains from controls were used to identify recombinants and parental forms. The plasmid acquisition by recipient strains was confirmed by PCR with primers specific to genes encoding plasmid replication proteins and relaxases (positions 1039–2159 and 22713–23548 in KU159086.1 and 77542–78366 and 83825–84614 in KP008371.1, respectively). PCR was performed according to a standard protocol for amplification of fragments with a size of 1 kb [31].

4.3. Resistance Selection Studies under Static Conditions

To select mutants of E. coli and K. pneumoniae with lowered meropenem susceptibility (equal to the respective transconjugant MICs), a previously described serial passages technique on antibiotic-containing media was performed [32]. Briefly, parental strains (E. coli C600 and K. pneumoniae ATCC 700603) were passaged in CSMHB, containing consecutively increasing concentrations of meropenem (from 0.016 to 4 mg/L). For each subsequent passage, an inoculum was taken from the tube with the maximal meropenem concentration that showed visual growth with a turbidity equivalent to or exceeding that of a 0.3 McFarland standard. A sample was then plated on Mueller–Hinton agar (MHA, Becton Dickinson, Franklin Lakes, NJ, USA) containing the same meropenem concentration and the cycle was repeated. The incubation period of each step was 24 h. The passages in meropenem containing CSMHB were repeated 7 to 23 times, depending on the strain.

4.4. In Vitro Dynamic Model and Operational Procedure Used in the Pharmacodynamic Experiments

The HFIM was used to evaluate meropenem pharmacodynamics and to conduct growth control experiments. The studies were performed using a hollow-fiber bioreactor (Fresenius dialyzer, model AV400S, Fresenius Medical Care AG, Bad Homburg Germany) that represents the infection site.
The operational procedure is described in detail elsewhere [33]. Briefly, antibiotic dosing and sampling were processed automatically, using computer-assisted controls. The system was filled with sterile CSMHB and placed in an incubator at 37 °C. The inoculum of an 18 h culture of E. coli or K. pneumoniae was injected into the hollow-fiber bioreactor to produce a bacterial concentration of 108 CFU/mL. After a 2 h incubation, samples were obtained to determine the starting bacterial concentration; then, the infusion of CSMHB with antibiotic was initiated. The duration of each experiment was 120 h. To verify the reliability of pharmacokinetic simulations, throughout each experiment, the bioreactor was sampled immediately after the end of infusion (0.5 h) and at the 5th hour of the dosing interval.

4.5. Antibiotic Dosing Regimens and Simulated Pharmacokinetic Profiles

Meropenem treatment mimicked the therapeutic dosing regimen: 0.5 g administered every 8 h, as a 0.5 h intravenous infusion. A mono-exponential profile in peritoneal fluid (PF) after thrice-daily dosing of meropenem with a half-life (t1/2) of 1.4 h was simulated [15] for 5 consecutive days with a total of 15 infusions. The pharmacokinetic parameter values were as follows: CMAX = 24 mg/L, 24 h area under the concentration–time curve (AUC) = 160 (mg × h)/L. Before all pharmacodynamic simulations, the system was calibrated and preliminary in vitro pharmacokinetic experiments in CSMHB without bacteria were conducted. For “dose—response” relationships, the total daily dose of meropenem (in mg) that was infused into the system to simulate the desired meropenem pharmacokinetic profile was calculated.

4.6. Quantitation of the Antimicrobial Effect

Bacteria-containing samples from the hollow-fiber bioreactor were obtained to determine bacterial titer and antibiotic concentrations throughout the observation period. Total bacterial enumeration was performed by plating 100 µL samples onto tryptic soy agar (Becton Dickinson, Franklin Lakes, NJ, USA) plates. If necessary, the samples were serially diluted and were then plated onto tryptic soy agar. All plates were incubated at 37 °C for 24 h to quantify the growth. The lower limit of accurate detection was 2 log CFU/mL (equivalent to 10 colonies per plate).
To determine the time course of meropenem-resistant bacterial numbers, samples from the hollow-fiber bioreactor were plated on MHA plates with meropenem concentrations equal to 2×, 4×, 8×, and 16× MIC of the tested strain, incubated for 24–48 h at 37 °C, and then screened visually for growth. The lower limit of detection was 1 log CFU/mL (equivalent to at least one colony per plate).
To quantify the meropenem total antimicrobial effect, an integral parameter, determined as the area between the control growth and time-kill curves (ABBC) [11], was calculated. The interrelation of ABBC with the antibacterial effect is direct: the greater the effect, the higher the ABBC. Time courses of bacterial counts resistant to meropenem (for each meropenem resistance level, from 2× to 16× MIC) in pharmacodynamic experiments were characterized by the area under the bacterial mutant concentration–time curve (AUBCM) [11] determined from the beginning of treatment to 120 h. The interrelation of AUBCM with the emergence of resistance is direct: the greater the resistant cells growth, the higher the AUBCM.

4.7. Statistical Analysis

The reported SI and HI MIC data were obtained by calculation of the respective modal values. In pharmacodynamic and growth control experiments, bacterial count data were calculated as arithmetic mean ± standard deviations for three replicate experiments. Based on these data, kinetic growth and time-kill curves were constructed. Assuming that the coefficient of variation for log CFU/mL data was less than 10%, to facilitate figure viewing, we did not place data point error bars, in order to not interfere with the kinetic curves.
The ABBC versus log(DOSE/MIC) curve was fitted by the sigmoid function:
Y = Y0 + a/{1 + exp[−(xx0)/b]}
where Y is ABBC; x is log(DOSE/MIC); a are maximal values of the antimicrobial effect; x0 is x corresponding to a/2; b is a parameter reflecting sigmoidicity.
A Gaussian function was used to fit the AUBCM versus DOSE/MIC data sets:
Y = a/{1 + exp[−(xx0)/b]}
where Y is AUBCM; x is log(DOSE/MIC); x0 is log DOSE/MIC that corresponds to the maximal value of Y; a and b are parameters.
All calculations were performed using SigmaPlot 12 software (Systat Software Inc., headquartered in San Jose, CA, USA).

5. Conclusions

The present study revealed that (1) unlike carbapenemase-non-producing E. coli and K. pneumoniae, meropenem-susceptible carbapenemase-producing strains may not respond to meropenem therapy; (2) the antibiotic IE may have a prognostic value to reveal meropenem-susceptible Enterobacterales strains between the isolates that harbor carbapenemase genes; and (3) there is a hidden threat that bacteria devoid of antibiotic resistance mechanisms may contain carbapenemase genes even if their meropenem MICs fall below the susceptibility breakpoint.

Author Contributions

Conceptualization, M.V.G., M.A.P. and S.H.Z.; methodology, M.V.G., M.A.P. and S.H.Z.; validation, E.N.S., K.N.A., D.A.K. and N.F.P.; formal analysis, E.N.S., K.N.A., D.A.K. and N.F.P.; investigation, M.V.G., E.N.S., K.N.A. and D.A.K.; resources, M.V.G.; funding acquisition, M.V.G.; data curation, E.N.S. and K.N.A.; writing—original draft preparation, M.V.G. and S.H.Z.; writing—review and editing, M.A.P. and E.N.S.; visualization, M.V.G. and K.N.A.; supervision, M.V.G. and M.A.P.; project administration, M.V.G. and E.N.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation (grant number 23-25-00525).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

We wish to acknowledge Sidorenko S.V. (Pediatric Research and Clinical Center for Infectious Diseases, Saint Petersburg, Russia) and Ageevets A.V. (Pediatric Research and Clinical Center for Infectious Diseases, Saint Petersburg, Russia) for providing the K. pneumoniae clinical isolates.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lai, C.-C.; Yu, W.-L. Klebsiella pneumoniae Harboring Carbapenemase Genes in Taiwan: Its Evolution over 20 Years, 1998–2019. Int. J. Antimicrob. Agents 2021, 58, 106354. [Google Scholar] [CrossRef] [PubMed]
  2. Ragheb, S.M.; Tawfick, M.M.; El-Kholy, A.A.; Abdulall, A.K. Phenotypic and Genotypic Features of Klebsiella pneumoniae Harboring Carbapenemases in Egypt: OXA-48-Like Carbapenemases as an Investigated Model. Antibiotics 2020, 9, 852. [Google Scholar] [CrossRef]
  3. Raro, O.H.F.; da Silva, R.M.C.; Filho, E.M.R.; Sukiennik, T.C.T.; Stadnik, C.; Dias, C.A.G.; Iglesias, J.O.; Pérez-Vázquez, M. Carbapenemase-Producing Klebsiella pneumoniae From Transplanted Patients in Brazil: Phylogeny, Resistome, Virulome and Mobile Genetic Elements Harboring blaKPC–2 or blaNDM–1. Front. Microbiol. 2020, 11, 1563. [Google Scholar] [CrossRef] [PubMed]
  4. Michaelis, C.; Grohmann, E. Horizontal Gene Transfer of Antibiotic Resistance Genes in Biofilms. Antibiotics 2023, 12, 328. [Google Scholar] [CrossRef] [PubMed]
  5. Dell’annunziata, F.; Folliero, V.; Giugliano, R.; De Filippis, A.; Santarcangelo, C.; Izzo, V.; Daglia, M.; Galdiero, M.; Arciola, C.R.; Franci, G. Gene Transfer Potential of Outer Membrane Vesicles of Gram-Negative Bacteria. Int. J. Mol. Sci. 2021, 22, 5985. [Google Scholar] [CrossRef]
  6. Emamalipour, M.; Seidi, K.; Vahed, S.Z.; Jahanban-Esfahlan, A.; Jaymand, M.; Majdi, H.; Amoozgar, Z.; Chitkushev, L.T.; Javaheri, T.; Jahanban-Esfahlan, R.; et al. Horizontal Gene Transfer: From Evolutionary Flexibility to Disease Progression. Front. Cell Dev. Biol. 2020, 8, 229. [Google Scholar] [CrossRef] [PubMed]
  7. Forero-Hurtado, D.; Corredor-Rozo, Z.L.; Ruiz-Castellanos, J.S.; Márquez-Ortiz, R.A.; Abril, D.; Vanegas, N.; Lafaurie, G.I.; Chambrone, L.; Escobar-Pérez, J. Worldwide Dissemination of blaKPC Gene by Novel Mobilization Platforms in Pseudomonas aeruginosa: A Systematic Review. Antibiotics 2023, 12, 658. [Google Scholar] [CrossRef]
  8. Meng, L.; Liu, Z.; Liu, C.; Li, C.; Shen, H.; Cao, X. The distribution characteristics of global blaOXA-carrying Klebsiella pneumoniae. BMC Infect. Dis. 2023, 23, 1–9. [Google Scholar] [CrossRef]
  9. The European Committee on Antimicrobial Susceptibility Testing. Breakpoint Tables for Interpretation of MICs and Zone Diameters. Version 13.1. 2023. Available online: http://www.eucast.org (accessed on 1 December 2023).
  10. CLSI document M100; Clinical and Laboratory Standards Institute (CLSI). Performance Standards for Antimicrobial Susceptibility Testing, 33rd ed. Clinical and Laboratory Standards Institute (CLSI): Wayne, PA, USA, 2023.
  11. Alieva, K.N.; Golikova, M.V.; Dovzhenko, S.A.; Kobrin, M.B.; Strukova, E.N.; Ageevets, V.A.; Avdeeva, A.A.; Sulian, O.S.; Sidorenko, S.V.; Zinner, S.H. Testing the mutant selection window hypothesis with meropenem: In vitro model study with OXA-48-producing Klebsiella pneumoniae. PLoS ONE 2023, 18, e0288660. [Google Scholar] [CrossRef]
  12. Blaser, J.; Zinner, S.H. In vitro models for the study of antibiotic activities. Prog Drug Res. 1987, 31, 349–381. [Google Scholar] [CrossRef]
  13. Sadouki, Z.; McHugh, T.D.; Aarnoutse, R.; Canseco, J.O.; Darlow, C.; Hope, W.; van Ingen, J.; Longshaw, C.; Manissero, D.; Mead, A.; et al. Application of the hollow fibre infection model (HFIM) in antimicrobial development: A systematic review and recommendations of reporting. J. Antimicrob. Chemother. 2021, 76, 2252–2259. [Google Scholar] [CrossRef] [PubMed]
  14. Lenhard, J.R.; Bulman, Z.P. Inoculum effect of β-lactam antibiotics. J. Antimicrob. Chemother. 2019, 74, 2825–2843. [Google Scholar] [CrossRef] [PubMed]
  15. Veeraraghavan, B.; Bakthavatchalam, Y.D.; Anandan, S. Laboratory detection and clinical implication of oxacillinase-48 like carbapenemase: The hidden threat. J. Glob. Infect. Dis. 2016, 8, 41–50. [Google Scholar] [CrossRef] [PubMed]
  16. Hleba, L.; Hlebová, M.; Kováčik, A.; Čuboň, J.; Medo, J. Carbapenemase Producing Klebsiella pneumoniae (KPC): What Is the Best MALDI-TOF MS Detection Method. Antibiotics 2021, 10, 1549. [Google Scholar] [CrossRef] [PubMed]
  17. Zhou, G.; Wang, Q.; Wang, Y.; Wen, X.; Peng, H.; Peng, R.; Shi, Q.; Xie, X.; Li, L. Outer Membrane Porins Contribute to Antimicrobial Resistance in Gram-Negative Bacteria. Microorganisms 2023, 11, 1690. [Google Scholar] [CrossRef] [PubMed]
  18. Padilla, E.; Llobet, E.; Doménech-Sánchez, A.; Martínez-Martínez, L.; Bengoechea, J.A.; Albertí, S. Klebsiella pneumoniae AcrAB Efflux Pump Contributes to Antimicrobial Resistance and Virulence. Antimicrob. Agents Chemother. 2010, 54, 177–183. [Google Scholar] [CrossRef]
  19. Bodmann, K.-F.; Der Infektliga, U.D.E. Komplizierte intraabdominelle Infektionen: Erreger, Resistenzen. Der Chir. 2009, 81, 38–49. [Google Scholar] [CrossRef] [PubMed]
  20. Golikova, M.V.; Strukova, E.N.; Alieva, K.N.; Ageevets, V.A.; Avdeeva, A.A.; Sulian, O.S.; Zinner, S.H. Meropenem MICs at Standard and High Inocula and Mutant Prevention Concentration Inter-Relations: Comparative Study with Non-Carbapenemase-Producing and OXA-48-, KPC- and NDM-Producing Klebsiella pneumoniae. Antibiotics 2023, 12, 872. [Google Scholar] [CrossRef]
  21. Su, C.-F.; Chuang, C.; Lin, Y.-T.; Chan, Y.-J.; Lin, J.-C.; Lu, P.-L.; Wang, J.-T.; Siu, L.K.; Fung, C.-P. Treatment outcome of non-carbapenemase-producing carbapenem-resistant Klebsiella pneumoniae infections: A multicenter study in Taiwan. Eur. J. Clin. Microbiol. Infect. Dis. 2017, 37, 651–659. [Google Scholar] [CrossRef]
  22. Lim, F.K.; Liew, Y.X.; Cai, Y.; Lee, W.; Teo, J.Q.M.; Lay, W.Q.; Chung, J.; Kwa, A.L.H. Treatment and Outcomes of Infections Caused by Diverse Carbapenemase-Producing Carbapenem-Resistant Enterobacterales. Front. Cell. Infect. Microbiol. 2020, 10, 579462. [Google Scholar] [CrossRef]
  23. Debnath, A.; Pillinger, K.E.; Martin, A.J.; Dobrzynski, D.; Cameron, A.; Shulder, S. Clinical Outcomes and Treatment Strategies in Patients with Non-Carbapenemase-producing Carbapenem-Resistant Versus Carbapenem-Susceptible Enterobacterales Infections. Ann. Pharmacother. 2022, 57, 803–812. [Google Scholar] [CrossRef] [PubMed]
  24. Oka, K.; Matsumoto, A.; Tetsuka, N.; Morioka, H.; Iguchi, M.; Ishiguro, N.; Nagamori, T.; Takahashi, S.; Saito, N.; Tokuda, K.; et al. Clinical characteristics and treatment outcomes of carbapenem-resistant Enterobacterales infections in Japan. J. Glob. Antimicrob. Resist. 2022, 29, 247–252. [Google Scholar] [CrossRef] [PubMed]
  25. Lee, N.-Y.; Tsai, C.-S.; Syue, L.-S.; Chen, P.-L.; Li, C.-W.; Li, M.-C.; Ko, W.-C. Treatment Outcome of Bacteremia Due to Non–Carbapenemase-producing Carbapenem-Resistant Klebsiella pneumoniae Bacteremia: Role of Carbapenem Combination Therapy. Clin. Ther. 2020, 42, e33–e44. [Google Scholar] [CrossRef] [PubMed]
  26. Tamma, P.D.; Huang, Y.; Opene, B.N.A.; Simner, P.J. Determining the Optimal Carbapenem MIC That Distinguishes Carbapenemase-Producing and Non-Carbapenemase-Producing Carbapenem-Resistant Enterobacteriaceae. Antimicrob. Agents Chemother. 2016, 60, 6425–6429. [Google Scholar] [CrossRef] [PubMed]
  27. Soga, Y.; Ohge, H.; Ikawa, K.; Morikawa, N.; Ikeda, K.; Sueda, T. Peritoneal Pharmacokinetics and Pharmacodynamic Target Attainment of Meropenem in Patients Undergoing Abdominal Surgery. J. Chemother. 2010, 22, 98–102. [Google Scholar] [CrossRef] [PubMed]
  28. Ageevets, V.; Sopova, J.; Lazareva, I.; Malakhova, M.; Ilina, E.; Kostryukova, E.; Babenko, V.; Carattoli, A.; Lobzin, Y.; Uskov, A.; et al. Genetic Environment of the blaKPC-2 Gene in a Klebsiella pneumoniae Isolate That May Have Been Imported to Russia from Southeast Asia. Antimicrob. Agents Chemother. 2017, 61, e01856-16. [Google Scholar] [CrossRef]
  29. Van Der Zwaluw, K.; De Haan, A.; Pluister, G.N.; Bootsma, H.J.; De Neeling, A.J.; Schouls, L.M. The Carbapenem Inactivation Method (CIM), a Simple and Low-Cost Alternative for the Carba NP Test to Assess Phenotypic Carbapenemase Activity in Gram-Negative Rods. PLoS ONE 2015, 10, e0123690. [Google Scholar] [CrossRef]
  30. ISO 20776-1:2019; Susceptibility Testing of Infectious Agents and Evaluation of Performance of Antimicrobial Susceptibility Test Devices—Part 1: Broth Micro-Dilution Reference Method for Testing the in vitro Activity of Antimicrobial Agents against Rapidly Growing Aerobic Bacteria Involved in Infectious Diseases. ISO: Geneva, Switzerland, 2019.
  31. Canene-Adams, K. General PCR. Lab. Methods Enzymol. DNA 2013, 529, 291–298. [Google Scholar] [CrossRef]
  32. Sambrook, J.; Russell, D.W. Molecular Cloning: A Laboratory Manual, Cold Spring Harbor Laboratory, 3rd ed.; Cold Spring Harbor: New York, NY, USA, 2001. [Google Scholar]
  33. Firsov, A.A.; Alieva, K.N.; Strukova, E.N.; Golikova, M.V.; Portnoy, Y.A.; Dovzhenko, S.A.; Kobrin, M.B.; Romanov, A.V.; Edelstein, M.V.; Zinner, S.H. Testing the mutant selection window hypothesis with Staphylococcus aureus exposed to linezolid in an in vitro dynamic model. J. Antimicrob. Chemother. 2017, 72, 3100–3107. [Google Scholar] [CrossRef]
Figure 1. Flowchart followed to obtain bacterial strains for pharmacodynamic study.
Figure 1. Flowchart followed to obtain bacterial strains for pharmacodynamic study.
Antibiotics 12 01717 g001
Figure 2. Time courses of the total bacterial population (0 × MIC) and meropenem-resistant (2×, 4×, 8× and 16 × MIC) sub-populations of parental carbapenemase-non-producing (left panel) and carbapenemase-producing (right panel) strains of E. coli and K. pneumoniae in pharmacodynamic experiments.
Figure 2. Time courses of the total bacterial population (0 × MIC) and meropenem-resistant (2×, 4×, 8× and 16 × MIC) sub-populations of parental carbapenemase-non-producing (left panel) and carbapenemase-producing (right panel) strains of E. coli and K. pneumoniae in pharmacodynamic experiments.
Antibiotics 12 01717 g002
Figure 3. Time courses of the total bacterial population (0 × MIC) and meropenem-resistant (2×, 4×, 8× and 16 × MIC) sub-populations of mutant carbapenemase-non-producing (left panel) and transconjugant carbapenemase-producing (right panel) strains of E. coli and K. pneumoniae in pharmacodynamic experiments. Arrows indicate the start of meropenem infusion.
Figure 3. Time courses of the total bacterial population (0 × MIC) and meropenem-resistant (2×, 4×, 8× and 16 × MIC) sub-populations of mutant carbapenemase-non-producing (left panel) and transconjugant carbapenemase-producing (right panel) strains of E. coli and K. pneumoniae in pharmacodynamic experiments. Arrows indicate the start of meropenem infusion.
Antibiotics 12 01717 g003
Figure 4. The ABBC values for mutant and transconjugant E. coli and K. pneumoniae strains grouped by meropenem susceptibility. “CARB+” indicates that the strain is a carbapenemase-producer, “CARB−” indicates that the strain is a non-carbapenemase-producer.
Figure 4. The ABBC values for mutant and transconjugant E. coli and K. pneumoniae strains grouped by meropenem susceptibility. “CARB+” indicates that the strain is a carbapenemase-producer, “CARB−” indicates that the strain is a non-carbapenemase-producer.
Antibiotics 12 01717 g004
Figure 5. Relationship between the meropenem effectiveness and its total daily dose related to the K. pneumoniae MIC at SI (a) and at HI (b). Red circles—carbapenemase-non-producers; blue circles—carbapenemase-producers. DOSE/MIC relationship with ABBC fitted by Equation (1): (SI) x0 = 0.798, a = 920, b = 0.267; (HI) x0 = −0.391, a = 920, b = 0.204.
Figure 5. Relationship between the meropenem effectiveness and its total daily dose related to the K. pneumoniae MIC at SI (a) and at HI (b). Red circles—carbapenemase-non-producers; blue circles—carbapenemase-producers. DOSE/MIC relationship with ABBC fitted by Equation (1): (SI) x0 = 0.798, a = 920, b = 0.267; (HI) x0 = −0.391, a = 920, b = 0.204.
Antibiotics 12 01717 g005
Figure 6. Relationship between the development of meropenem resistance (expressed as parameter AUBCM) of tested E. coli and K. pneumoniae strains under antibiotic exposure and meropenem total daily dose related to MIC at SI (a) and at HI (b) against respective bacterial isolate. Red circles—carbapenemase-non-producers; blue circles—carbapenemase-producers. DOSE/MIC relationship with AUBCM for mutants resistant to 4 × MIC of meropenem; data fitted by Equation (2): (a) x0 = 0.044, a = 550, b = 0.605; (b) x0 = −1.004, a = 550, b = 0.474.
Figure 6. Relationship between the development of meropenem resistance (expressed as parameter AUBCM) of tested E. coli and K. pneumoniae strains under antibiotic exposure and meropenem total daily dose related to MIC at SI (a) and at HI (b) against respective bacterial isolate. Red circles—carbapenemase-non-producers; blue circles—carbapenemase-producers. DOSE/MIC relationship with AUBCM for mutants resistant to 4 × MIC of meropenem; data fitted by Equation (2): (a) x0 = 0.044, a = 550, b = 0.605; (b) x0 = −1.004, a = 550, b = 0.474.
Antibiotics 12 01717 g006
Table 1. Modal MICs (mg/L) of meropenem at SI and HI against E. coli and K. pneumoniae strains.
Table 1. Modal MICs (mg/L) of meropenem at SI and HI against E. coli and K. pneumoniae strains.
Bacterial StrainCarbapenemase ProducerMeropenem
MIC at SI, mg/L
Meropenem
MIC at HI, mg/L
Fold MIC IncreaseIE
C600No0.030.031
ATCC 700603No0.030.516 *+
485Yes322568 *+
565Yes645128 *+
mC600/0.5No0.524
mC600/1No144
tcC600/485Yes0.5816 *+
tcC600/565Yes16464 *+
m700603/0.5No0.548 *+
m700603/4No4328 *+
tc700603/485Yes0.5816 *+
tc700603/565Yes412832 *+
* MIC changes associated with IE. “−“-indicate the IE was not detected, “+”-indicate the IE was detected.
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

Golikova, M.V.; Alieva, K.N.; Strukova, E.N.; Kondratieva, D.A.; Petrova, N.F.; Petrova, M.A.; Zinner, S.H. Comparative Meropenem Pharmacodynamics and Emergence of Resistance against Carbapenem-Susceptible Non-Carbapenemase-Producing and Carbapenemase-Producing Enterobacterales: A Pharmacodynamic Study in a Hollow-Fiber Infection Model. Antibiotics 2023, 12, 1717. https://doi.org/10.3390/antibiotics12121717

AMA Style

Golikova MV, Alieva KN, Strukova EN, Kondratieva DA, Petrova NF, Petrova MA, Zinner SH. Comparative Meropenem Pharmacodynamics and Emergence of Resistance against Carbapenem-Susceptible Non-Carbapenemase-Producing and Carbapenemase-Producing Enterobacterales: A Pharmacodynamic Study in a Hollow-Fiber Infection Model. Antibiotics. 2023; 12(12):1717. https://doi.org/10.3390/antibiotics12121717

Chicago/Turabian Style

Golikova, Maria V., Kamilla N. Alieva, Elena N. Strukova, Daria A. Kondratieva, Nika F. Petrova, Mayya A. Petrova, and Stephen H. Zinner. 2023. "Comparative Meropenem Pharmacodynamics and Emergence of Resistance against Carbapenem-Susceptible Non-Carbapenemase-Producing and Carbapenemase-Producing Enterobacterales: A Pharmacodynamic Study in a Hollow-Fiber Infection Model" Antibiotics 12, no. 12: 1717. https://doi.org/10.3390/antibiotics12121717

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