Abstract
Antimicrobial resistance is a major global health challenge. As few new efficacious antibiotics will become available in the near future, peptide antibiotics continue to be major therapeutic options for treating infections caused by multidrug-resistant pathogens. Rational use of antibiotics requires optimisation of the pharmacokinetics and pharmacodynamics for the treatment of different types of infections. Toxicodynamics must also be considered to improve the safety of antibiotic use and, where appropriate, to guide therapeutic drug monitoring. This review focuses on the pharmacokinetics/pharmacodynamics/toxicodynamics of peptide antibiotics against multidrug-resistant Gram-negative and Gram-positive pathogens. Optimising antibiotic exposure at the infection site is essential for improving their efficacy and minimising emergence of resistance.
Keywords: Antimicrobial resistance, antimicrobial peptide, antibiotics, pharmacokinetics, pharmacodynamics, safety, therapeutic drug monitoring
1. Introduction
1.1. Antimicrobial resistance: A major global health challenge
In 2020, the World Health Organization (WHO) declared antimicrobial resistance one of the top 10 global public human health threats [1]. Over the last several decades, multidrug-resistant (MDR) bacterial pathogens that are resistant to three or more antibiotic categories [2] have been increasingly reported worldwide [3, 4]. Infections caused by these highly-resistant microorganisms result in high morbidity and mortality [5]. Currently, the annual global death toll from antibiotic-resistant organisms is at least 700,000, including more than 23,000 deaths in the USA and 175,000 in Europe [6]. This figure is expected to rise to approximately 10 million deaths annually by 2050 if efforts to curtail resistance and develop new antibiotics are not implemented [7]. Were this to occur, the associated global cost is estimated to be US$100 trillion by 2050 [8]. Without effective antimicrobials, the successes of modern medicine that depend on the availability of effective antimicrobial therapy such as cancer chemotherapy, organ transplantation and other major surgeries, is threatened [9, 10].
In response to this major global medical crisis, the WHO highlighted a list of priority pathogens urgently requiring the discovery and development of novel antibiotic treatments [11]. The pathogens given the highest priority (Priority 1: Critical) were carbapenem-resistant Acinetobacter baumannii and Pseudomonas aeruginosa, and carbapenem-resistant or 3rd generation cephalosporin-resistant Enterobacteriaceae including Klebsiella pneumoniae and Escherichia coli [12]. Several other microorganisms including Gram-positive species (e.g., vancomycin-intermediate and -resistant S. aureus) are included in the Priority 2 (High) and Priority 3 (Medium) categories. Unfortunately, in parallel with the increasing resistance to existing antimicrobials, the development pipeline for new antimicrobials has dried up [13, 14]. In 2019, the WHO identified only 32 antibiotics in clinical development that target organisms contained in the WHO list of priority pathogens, and only six of these were classified as innovative [1]. As this review will demonstrate, antimicrobial peptides remain active against many pathogens on the WHO Priority Pathogens list, although worrying levels of resistance are beginning to emerge [15, 16]. Given the slow pace of antimicrobial development, optimising the use of existing peptide antibiotics is critical to combat the ever-increasing antimicrobial resistance.
1.2. Antimicrobial peptide antibiotics
Antimicrobial peptides (AMPs) are the first line of defense in many organisms ranging from bacteria, fungi, plants, invertebrates, non-mammalian vertebrates, to mammals; and they are a key source of new anti-infective agents [17, 18]. More than 3,000 AMPs have been reported and characterized, although relatively few have had their therapeutic potential assessed and fewer still have progressed to clinical trials [19]. To date, the only AMPs approved by FDA are the polymyxins (polymyxin B and colistin [polymyxin E]) for Gram-negative bacteria, the glycopeptides (vancomycin, oritavancin, dalbavancin, and telavancin) and daptomycin for Gram-positive bacteria, and gramicidin (not discussed in this review as it is only administered topically) for both Gram-positive and Gram-negative bacteria [20, 21]. Although not approved for use in humans by the FDA, the glycopeptide teicoplanin has been approved in Asia and Europe for the treatment of serious infections caused by Gram-positive bacteria [22].
Bacterial killing by the abovementioned AMPs primarily involves membrane disorganization and/or inhibition of cell wall synthesis. The polymyxins primarily cause disruption and permeabilization of the outer membrane of Gram-negative bacteria, leading to cell death [23]. Daptomycin disorganizes the cell membrane of Gram-positive bacteria and causes arrest of DNA, RNA and protein synthesis [24, 25]. The glycopeptides vancomycin, dalbavancin and teicoplanin bind to cell wall precursors leading to interference of penicillin-binding protein (PBP) enzymes, thus inhibiting cell wall synthesis [19]. Oritavancin and telavancin have a dual antimicrobial mechanism that targets both the cell membrane and cell wall [19]. These modes of action will be described in detail in the relevant sections below. As the AMPs target the cell membrane or cell wall, they act rapidly and often have a relatively broad spectrum of activity [26]. Consequently, AMPs may be used in combination with other classes of antibiotics to improve efficacy and minimise the emergence of resistance [17]. This review focuses on the pharmacokinetics/pharmacodynamics (PK/PD) of existing AMP antibiotics in order to promote their optimal use in clinical practice.
2. Polymyxins: The last-line therapy against MDR Gram-negative bacteria
2.1. Chemistry and mechanisms of activity and resistance
The polymyxins are cyclic lipopeptides naturally produced by the Gram-positive Paenibacillus polymyxa (previously known as Bacillus polymyxa) [27]. While various polymyxins have been described, only polymyxin B and E (the latter known as colistin) are available for clinical use, having first entered the clinic over 60 years ago [28, 29]. The basic structure of the polymyxins consists of a cyclic heptapeptide with a tripeptide side chain linked to an N-terminal fatty acyl tail (Figure 1) [27]. The five α,γ-diaminobutyric acid (Dab) residues at positions 1, 3, 5, 8 and 9 are positively-charged at physiological pH. Polymyxin B and colistin differ only at position 6 where polymyxin B contains D-phenylalanine and colistin contains D-leucine (Figure 1) [30]. Both polymyxin B and colistin are mixtures of multiple components. The major constituents of polymyxin B are polymyxin B1, B2, B3 and polymyxin B1-Ile (isoleucine), which mainly differ in the fatty acyl moiety [31]. For colistin, the major components are colistin A and B [32]. As will be discussed in Section 2.2, polymyxin B is administered parenterally in its active form as the sulphate salt [33]. In contrast, colistin is administered as the inactive prodrug sodium colistin methanesulphonate (CMS) which subsequently converts to the active moiety, colistin [34], except in China where a product of colistin sulfate is also available for intravenous administration [35]. The different entities administered to patients create major differences in the pharmacokinetics.
Figure 1.

Structures of polymyxin B and colistin [36]. Permission obtained from Elsevier Press.
The exact mechanism(s) by which polymyxins kill Gram-negative bacteria are unknown. However, their antimicrobial activity is initiated by an electrostatic attraction between the electropositive Dab residues of the polymyxin and the electronegative lipid A of lipopolysaccharide (LPS) in the bacterial outer membrane (OM) [37]. This interaction is a requirement for the initial encounter and causes displacement of divalent cations (Ca2+ and Mg2+) that bridge adjacent LPS molecules, enabling the hydrophobic moieties of polymyxins (i.e., the N-terminal fatty acyl tail and D-Phe6-L-Leu7 [polymyxin B]/D-Leu6-L-Leu7 [colistin]; Figure 1) to insert into the outer membrane [38, 39]. These interactions cause disruption of lipid A fatty acyl chains resulting in OM leaflet expansion, disruption of the membrane integrity, and increased membrane permeability [38, 40]. The traditional view of polymyxin activity sees the disruption of OM structures as the primary mechanism by which polymyxins ultimately cause bacterial cell death. However, it has been suggested that polymyxins in the periplasmic space induce vesicle-contacts between the periplasmic leaflets of the outer and inner membrane, promoting anionic phospholipid exchange that leads to an osmotic imbalance that contributes to cell death [41, 42]. While disruption of OM integrity is still considered the primary mode of action of polymyxins, secondary modes of action have also been proposed including inhibition of bacterial respiration via the inhibition of essential respiratory enzymes (e.g., type II NADH-quinone oxidoreductases [NDH-2]) [43], and induction of a hydroxyl radical death pathway [44].
While the polymyxins still retain excellent activity against relevant Gram-negative pathogens (discussed in Section 2.2), polymyxin resistance has been increasingly reported worldwide [45]. As described above, the initial interaction of polymyxins with lipid A is necessary for bacterial killing. It is not surprising that the primary mechanism of resistance to polymyxins involves modifications of lipid A that reduce the initial electrostatic interactions with polymyxins. This includes covalent modification of phosphate groups of lipid A with positively charged motifs, such as 4-amino-4-deoxy-L-arabinose (L-Ara4N), phosphoethanolamine (pEtN) and galactosamine [46–48]. These modifications are mainly mediated by the arn (pmr) operon which is regulated by different two-component regulatory systems (e.g., PhoPQ, PmrAB and ParRS) [49, 50]. Complete loss of LPS (including the lipid A moiety anchoring LPS to the OM) due to mutations in key LPS synthesis genes has also been reported in A. baumannii [51, 52].
In addition to these primary resistance mechanisms, other mechanisms of polymyxin resistance including binding to anionic capsular polysaccharide (in K. pneumoniae) [53], exclusion from cells via efflux pumps [54, 55], and OM proteins (mechanism unclear) [56, 57] have also been reported. While a putative serine protease designated colistinase has been reported in the Gram-positive bacterium P. polymyxa [58], polymyxin resistance in other microorganisms has never been associated with polymyxin degradation.
Suboptimal antibiotic exposure can lead to resistance, including to AMPs [59]. The evolutionary mechanism of resistance to polymyxins was investigated in P. aeruginosa, especially in highly resistant strains [60]. Strong intergenic epistasis explains varying levels of resistance. For example, mutated two-component regulatory systems PhoPQ and PmrAB usually constitute critical first steps in the evolution of polymyxin resistance. Mutations in opr86 (encoding an outer membrane protein) or lpxC (lipid A biosynthesis) often lead to a higher level of resistance than with mutations merely in pmrB or phoQ in P. aeruginosa [60]. Mathematical models describing the relationship between single or multiple mutations and treatment failure revealed that if a mutation causes high-level resistance, the risk of bacteria surviving the treatment is very high [61]. However, if more than two mutations are required to achieve high-level resistance, the risk of treatment failure drops dramatically [61]. Therefore, optimising the PK/PD of antibiotics at the infection site to achieve maximum killing is critical to minimising the emergence of resistance.
2.2. Pharmacodynamics and pharmacokinetics
a). Pharmacodynamics
Antimicrobial pharmacodynamics (PD) examines the relationship between concentration and antimicrobial effect, with the minimum inhibitory concentration (MIC) a major parameter used to quantify the activity [62]. Polymyxin B and colistin share similar MICs and have a moderate spectrum of activity that includes many of the Gram-negative pathogens commonly responsible for MDR nosocomial infections.
Polymyxins exert rapid, concentration- and inoculum-dependent killing against Gram-negative bacteria including P. aeruginosa, A. baumannii and Enterobacterales. For example, against polymyxin-susceptible P. aeruginosa at an initial inoculum of 106 CFU/mL, bactericidal activity (i.e., ≥3 log10 CFU/mL reduction) was achieved within 1 h with concentrations ranging from 1–4× MIC, whereas no viable bacteria were detected within 30 min with concentrations ≥16× MIC [63]. At a 108 CFU/mL inoculum, bactericidal activity was 23-fold slower compared to that at 106 CFU/mL. Importantly, even with clinically unachievable concentrations of colistin (e.g., >16 mg/L) at the lower inoculum (106 CFU/mL), regrowth of polymyxin-resistant bacteria was observed [63]. Similar patterns of inoculum-dependent bacterial killing followed by regrowth of resistant bacteria (as early as ≤2 h) have been observed for polymyxin B against P. aeruginosa [64, 65] and for both polymyxins against other important Gram-negative pathogens [66–69]. Regrowth following initial extensive bacterial killing has primarily been attributed to the phenomenon of polymyxin heteroresistance, namely, isolates that are considered polymyxin-susceptible based upon their MICs but which contain resistant subpopulations [70]. The pattern of bacterial killing often followed by regrowth described above has similarly been observed in vivo and suggests caution with polymyxin monotherapy [71–73].
At high concentrations, colistin possesses at best a moderate post-antibiotic effect (PAE) against P. aeruginosa (2–3 h 16× MIC) and K. pneumoniae (1.6 h at ≥64× MIC) [67, 74]. The reported colistin PAE against A. baumannii is more variable which may be due to varying exposure times [66, 75, 76]. No studies have examined the PAE of polymyxin B and the clinical utility of the polymyxin PAE is unclear.
b). Pharmacokinetics in animals
In the clinic, polymyxin B is administered as its sulfate salt via intravenous infusion or inhalation, with patients therefore receiving the antibacterial entity directly [77, 78]; a colistin sulfate product for intravenous administration is only available in China [35]. In contrast, in most of the world colistin is administered intravenously, intrathecally and via inhalation as colistin methanesulphonate (CMS), an inactive prodrug that is converted in vivo to the active entity, colistin [34]. Over the last two decades, the development of specific chromatographic assays has resulted in substantial improvements in our understanding of the pharmacokinetics of CMS/colistin [79–82].
This review focuses on the pharmacokinetics of CMS and formed colistin after intravenous administration. In aqueous media including plasma and urine, CMS is unstable and can convert to colistin [83]. In rats, approximately 7–10% of intravenously administered CMS was converted systemically into colistin [83, 84], showing only a small fraction of the active entity is generated from the prodrug. CMS is predominantly cleared by renal excretion [83], while the formed colistin undergoes extensive renal tubular reabsorption via a carrier-mediated process and is cleared primarily by non-renal pathways [85, 86]. Such extensive renal tubular reabsorption of colistin (and polymyxin B) leads to their accumulation in kidney tissue and renal toxicity [87]. Despite this, urinary concentrations of colistin after administration of CMS can reach a high level due to conversion from CMS (which is extensively renally excreted) within the urinary tract [88]. The half-lives (t1/2) of CMS and colistin in rats ranged from 20.6–25.5 min and 32.4–45.2 min across the dosage range of 5–120 mg/kg, respectively, whereas the clearance of CMS ranged from 9.5 to 14.9 mL/min/kg; the colistin clearance was estimated to be half that of CMS. Also in rats, the pharmacokinetic parameters of the main components of polymyxin B (polymyxin B1, B2, B3 and polymyxin B1-Ile; the percentage of each component was not stated) were all similar, with a t1/2 of ~1.5 h and a total body clearance of ~1.6 mL/min [89]. Less than 1% of the intravenously administered dose of 4 mg/kg polymyxin B was recovered unchanged in urine over 48 h [89].
Plasma protein binding of polymyxin B and colistin is species dependent. In mice, the plasma protein binding of colistin was 91.4 ± 1.65% and constant over a wide concentration range of ~0.9–37 mg/L [73]; while in dogs [90] and calves [91], binding was ~50%. Similarly, determined by ultracentrifugation methods, 92.9 ± 3.3% of polymyxin B was bound to mouse plasma proteins over a concentration range of ~2–50 mg/L [71]. Ultracentrifugation is the most reliable method to determine the plasma protein binding of polymyxins by minimising the potential for non-specific binding. The different protein binding values of polymyxin B and colistin in different species should be considered when the unbound PK/PD indices are determined for optimising their clinical use.
c). PK/PD of polymyxins in vitro and in infected mice
Among the three key PK/PD indices, namely ƒCmax/MIC, ƒAUC/MIC and %ƒT>MIC, ƒAUC/MIC is most closely correlated with optimal microbiological outcomes [71–73, 92]. In an early hollow-fibre infection model study investigating the pharmacodynamics of polymyxin B against two strains of P. aeruginosa, similar bacterial killing was observed when the same total daily dose was administered 8-, 12-, or 24-hourly, suggesting that the pharmacodynamics of polymyxin B was most closely linked to AUC/MIC [64]. More recently, mouse thigh and lung infection studies revealed that polymyxin B antibacterial activity against three strains of K. pneumoniae was well correlated with fAUC/MIC (R2=0.89; Figure 2) [71]. While bacterial killing of 2 log10 CFU/thigh was not achieved for any strain of K. pneumoniae, target values of fAUC/MIC were 1.22–13.5 and 3.72–28.0 for stasis and 1 log10 kill, respectively [71]. In another study utilizing mouse thigh and lung infection models, bacterial killing of P. aeruginosa and A. baumannii in the thigh by colistin was also well correlated with fAUC/MIC (R2=0.82–0.94 for P. aeruginosa and R2=0.84–0.95 for A. baumannii) [73]. Target values for 2 log10 kill were 7.4–13.7 for P. aeruginosa and 7.4–17.6 for A. baumannii (Figure 3). Notably, in both the aforementioned studies bacterial killing was greatly reduced in the lung infection model, even at the highest doses tolerated by mice [71, 73]. Therefore, treating pneumonia with parenteral administration of polymyxins is very challenging due to the poor PK/PD in the infected lung.
Figure 2.

Relationship between fAUC/MIC of polymyxin B and log10 CFU/thigh at 24 h against K. pneumoniae FADDI-KP032 [71]. Log10 CFU/thigh represents the bacterial burden (counted as colony forming unit [CFU]) in a single mouse thigh. The dotted line represents the average bacterial burden in the thighs at the start of polymyxin B treatment. Permission obtained from Oxford Press.
Figure 3.

Relationships between fAUC/MIC of colistin and log10 CFU/thigh at 24 h against P. aeruginosa ATCC 27853 (a) and A. baumannii ATCC 19606 (b) [73]. Log10 CFU/thigh represents the bacterial burden (counted as colony forming unit [CFU]) in a single mouse thigh. The dotted line represents the average bacterial burden in the thighs at the start of colistin treatment. Permission obtained from Oxford Press.
d). Pharmacokinetics of CMS/colistin in healthy volunteers and critically ill patients
There are two different conventions used in different parts of the world to label the contents of parenteral CMS products, namely colistin base activity (CBA, primarily in North America, South America and Asia), and international units (IU, primarily in Europe) [93]. It is essential to understand both labelling conventions and the conversion (~33.3 mg of CBA = ~80 mg of the chemical CMS = 1 million IU [MIU]) for correct interpretation of polymyxin pharmacokinetics reported in the literature [93].
Three pharmacokinetic studies of CMS/colistin have been conducted in healthy subjects (Table 1) [88, 94, 95]. In a French study involving 12 healthy male volunteers administered a single dose of 1 MIU CMS intravenously over 1 h, the Cmax of CMS was 4.8 mg/L and the mean Cmax of formed colistin (achieved at ~2 to 3 h after commencement of the CMS infusion) was 0.83 mg/L [88]. The terminal t1/2 of colistin (~3 h) was longer than that of CMS (~2 h), indicating that the disposition of formed colistin was rate-limited by its own elimination and not conversion from CMS. The renal clearance of colistin was only 1.9 mL/min and it was estimated that only ~30% of CMS underwent conversion to colistin prior to renal excretion. However, the ~30% estimate assumes that the only clearance pathways of CMS are renal excretion and conversion to colistin. Given the percentage systemic conversion of CMS in animals is lower (~2.5–12%) [83, 84, 96], the ~30% estimate likely overestimates this conversion. The volumes of distribution of CMS (14.0 L) and colistin (12.4 L) were consistent with a distribution restricted to the extracellular space, likely due to their large size and positive charge at physiological pH (see Section 2.1) restricting their crossing of membranes and physiological barriers. Another two studies conducted in Japanese (15 males) [94] and Chinese (9 males and 9 females) [95] volunteers utilised a CMS dose of ~2.5 mg/kg CBA infused over 0.5 h (Japanese study) or 1.5 h (Chinese study). The Chinese study also examined the pharmacokinetics of two different brands of CMS (CTTQ from CHIA TAI TIAN-QING Pharmaceutical Group Co., Ltd. vs Coly-Mycin@ M from Parkedale Pharmaceuticals Inc.) in healthy subjects. The Cmax of formed colistin at steady state was 4.38 ± 1.56 mg/L in Japanese subjects, compared to 1.30 ± 0.19 mg/L (CTTQ) and 1.06 ± 0.16 mg/L (Parkedale) in Chinese subjects [94, 95]. The terminal t1/2 of CMS and formed colistin ranged from 0.73–2.76 h and 3–5.13 h, respectively, across all three studies, with the renal clearance of CMS (65.4–103 mL/min) being significantly greater than that of formed colistin (1.5–10.5 mL/min) [88, 94, 95].
Table 1.
Pharmacokinetics of CMS and formed colistin in healthy subjects.
| Study | Country | Volunteers (N) | Dosage regimen | PK parameters of CMS | PK parameters of colistin |
|---|---|---|---|---|---|
| Couet et al [88] | French | 12 males | 1 MIU CMS infused over 1 h | Cmax 4.8 mg/L, t1/2 2 h, CLr 103 mL/min, Vd 14.0 L |
Cmax 0.83 mg/L t1/2 3 h, CLr 1.9 mL/min, Vd 12.4 L |
| Mizuya-chi et al [95] | Japanese | 15 males | 2.5 mg/kg CBA infused over 0.5 h | Cmax 17.21 ± 2.50 mg/L, t1/2 0.47 ± 0.17 h, CLr 0.088 ± 0.024 L/h/kg, Vd 0.249 ± 0.044 L/kg | Cmax 4.38 ± 1.56 mg/L, t1/2 4.98 ± 0.99 h, CLr 0.0073 ± 0.0030 L/h/kg, Vd 1.032 ± 0.237 L/kg |
| Fan et al [96] | Chinese | 9 males and 9 females | 2.5 mg/kg CBA infused over 1.5 h | Cmax 22.6 ± 2.0 mg/L, t1/2 2.76 ± 1.46 h, CLr 0.06 ± 0.01 L/h/kg, Vd 0.27 ± 0.14 L/kg (CTTQ) | Cmax 1.30 ± 0.19 mg/L t1/2 5.02 ± 0.46 h, CLr 0.0014 ± 0.0003 L/h/kg, Vd 1.28 ± 0.18 L/kg (CTTQ) |
| Cmax 19.7 ± 1.8 mg/L, t1/2 2.36 ± 1.06 h, CLr 0.06 ± 0.01 L/h/kg, Vd 0.23 ± 0.10 L/kg (Parkedale) | Cmax 1.06 ± 0.16 mg/L t1/2 5.13 ± 0.48 h, CLr 0.0015 ± 0.0005 L/h/kg, Vd 1.39 ± 0.24 L/kg (Parkedale) |
For critically ill patients, the conversion ratio of CMS to colistin varies substantially according to the renal function [97]. For this reason, renal function significantly impacts the conversion of CMS to colistin (and thus the concentration of colistin achieved) in vivo following CMS administration. In 14 critically ill patients (creatinine clearance, 46–200 mL/min) receiving ~3 MIU of CMS every 8 or 12 h, the t1/2 of formed colistin ranged from 3.8 to 9.5 h (mean, 7.4 ± 1.7 h); the maximum and minimum plasma colistin concentrations ranged from 1.15–5.14 mg/L (mean, 2.93 ± 1.24 mg/L) and 0.35–1.70 mg/L (mean, 1.03 ± 0.44 mg/L), respectively [98]. The largest population pharmacokinetic analysis of CMS and formed colistin thus far involved 215 critically ill patients with a wide range of renal function (0–236 mL/min), including 29 patients receiving various forms of renal replacement therapy [99]. The apparent clearance of formed colistin was related to the renal function of the patients [99], which confirmed the interim analysis of the same study involving 105 critically ill patients [100]. Specifically, in patients with a creatinine clearance greater than ~80 mL/min/1.73 m2, CMS administration at or close to the maximum recommended daily dose of 9 MIU was unable to reliably achieve a plasma colistin Css,avg of 2 mg/L [100]. In critically ill patients, the plasma protein binding of colistin is approximately 50% [99, 101]. Based upon the current preclinical PK/PD and clinical pharmacokinetic findings, a plasma colistin Css,avg of 2 mg/L has been proposed as a rational target for patients receiving the European Medicines Agency (EMA) recommended CMS dosage regimens based on creatinine clearance (e.g., 9 MIU daily for a creatinine clearance of ≥50 mL/min) [102]. As this target cannot be reliably achieved in most patients with a creatine clearance greater than ~80 mL/min/1.73 m2 when receiving the maximum recommended dose (~9 MIU CMS per day), various loading and daily doses of CMS have been proposed for critically ill patients with different renal functions and for those on different renal replacement therapies in order to achieve a colistin plasma Css,avg of 2 mg/L [99, 103]. It should be noted that the current PK/PD findings do not support intravenous CMS for the treatment of pneumonia due to the poor drug exposure in the lungs.
e). Pharmacokinetics of polymyxin B in healthy volunteers and patients
Polymyxin B sulfate is available for intravenous administration except in Australia, New Zealand and Europe [104]. Polymyxin B sulfate products are labeled with international units (1 mg of polymyxin B equals 10,000 IU) [93]. The recommended dose of polymyxin B sulfate is 1.5–2.5 mg/kg/day or 50–100 mg/day administered in two divided doses [105, 106]. Similar to colistin, polymyxin B undergoes extensive tubular reabsorption [107] and its urinary recovery is low [107–109]. However, very few studies to date have examined the pharmacokinetics of polymyxin B. A recent phase I study of polymyxin B in 20 healthy Chinese volunteers is the only one to investigate the safety and pharmacokinetics of polymyxin B in healthy subjects [108]. Following intravenous administration at 0.75 mg/kg (1-h infusion; n = 10) and 1.5 mg/kg (1.5-h infusion; n = 9), the Cmax achieved were 4.66 ± 0.46 and 8.53 ± 1.07 mg/L, respectively. In the 0.75 mg/kg group, the total clearance, volume of distribution and t1/2 of polymyxin B were 0.028 ± 0.002 L/h/kg, 0.219 ± 0.023 L/kg and 5.44 ± 0.741 h, respectively; similar results were obtained in the 1.5 mg/kg group. The low urinary recovery of 3.7 ± 1.1% (0.75 mg/kg dosing) and 8.1 ± 1.3% (1.5 mg/kg dosing) is in agreement with previous studies undertaken in critically ill patients (<1% [107] and ~4% [109]) and suggests that extensive renal tubular reabsorption occurred.
The first modern pharmacokinetic study of polymyxin B examined eight critically ill patients using dosage regimens ranging from 0.5 mg/kg 48-hourly to 1.25 mg/kg 12-hourly (administered via a 1-h infusion) [107]. The maximum concentrations at steady state (Css,max) ranged from 2.38 to 13.9 mg/L. There was little inter-individual variability in total body clearance (range, 0.27–0.81 mL/min/kg) and volume of distribution (range, 71–194 mL/kg), and the urinary recovery of unchanged drug was <1% (0.04%–0.86%). The lack of variability in total body clearance in this study is a most important finding given the diverse renal functions in the patients (creatinine clearance <10 to 246 mL/min), implying that renal clearance is only a very small proportion of total body clearance. A subsequent population pharmacokinetic study involving 24 critically ill patients reported a similarly low urinary recovery of polymyxin B (4.04%, range 0.98–17.4%) and a median unbound fraction of polymyxin B of 0.42 (range, 0.26 – 0.64) in patient plasma [109]. Notably, creatinine clearance was not identified as a covariate of total body clearance [109]. However, several recent studies that have compared the population pharmacokinetic properties of polymyxin B in patients with different renal functions have identified creatinine clearance as a significant covariate in their patient populations [110–113]. The clinical significance of any such relationship in critically ill patients requires further examination, and current guidelines do not suggest the dose of intravenous polymyxin B be adjusted for renal impairment or dialysis [114, 115].
f). Pharmacokinetics/Pharmacodynamics based breakpoints of polymyxins
Prior to 2020, the Clinical and Laboratory Standards Institute (CLSI) provided susceptible and resistant breakpoints for colistin and polymyxin B (e.g., M100-S29). However, in 2020 the susceptible interpretive category (previously ≤2 mg/L in all cases before 2020) was removed (M100-S30). In contrast, the ‘susceptible’ category has been maintained by the European Committee on Antimicrobial Susceptibility Testing (EUCAST; for colistin; polymyxin B breakpoints not reported) [116] and the United States Committee on Antimicrobial Susceptibility Testing (USCAST; for both colistin and polymyxin B) [117]. The determination of breakpoints should be species- and infection-site specific based upon the antibiotic PK/PD [118]. The current breakpoints for polymyxins (polymyxin B and colistin) are shown in Table 2.
Table 2.
Breakpoints of polymyxin B and colistin.
| Organism | Breakpoint (mg/L)* | ||
|---|---|---|---|
| S | I | R | |
| CLSI | |||
| P. aeruginosa | ≤2 | ≥4 | |
| Acinetobacter spp. | ≤2 | ≥4 | |
| Enterobacterales | ≤2 | ≥4 | |
| EUCAST † | |||
| Pseudomonas spp. | ≤2 | >2 | |
| Acinetobacter spp. | ≤2 | >2 | |
| Enterobacterales | ≤2 | >2 | |
| USCAST # | |||
| Pseudomonas spp. | ≤2 | ≥4 | |
| Acinetobacter spp. | ≤2 | ≥4 | |
| Enterobacterales | ≤2 | ≥4 | |
S, susceptible; I, intermediate; R, resistant
EUCAST: EUCAST only provides the breakpoints for colistin.
USCAST recommends no polymyxin breakpoints for lower respiratory tract infections, and no breakpoint for polymyxin B for lower urinary tract infections.
g). Optimisation of intravenous polymyxins in patients
The different pharmacokinetics of CMS/colistin and polymyxin B reviewed above indicate that each polymyxin has its own advantages for different types of infections. Given that polymyxin B is administered as the active entity (i.e., polymyxin B sulfate), it rapidly attains therapeutic concentrations and has a relatively lower potential to cause nephrotoxicity (Section 2.3 below). Therefore, polymyxin B is the preferred agent for bloodstream and wound infections [114]. In a single-center cohort study involving 43 septic patients infected with 58 carbapenem-resistant Gram-negative strains (57/58 [98.3%] strains susceptible to polymyxin B), high doses of polymyxin B were administered for >72 h (median daily dose and duration of ~32,000 IU/kg/day IV and 14 days, respectively) [119]. Of the 24 patients with bloodstream infections, favorable microbiologic eradication was achieved in 23 patients (95.8%) [119]. A retrospective study examining the relationship between polymyxin B dose and clinical outcomes included patients with bloodstream infections due to carbapenem-resistant Gram-negative rods and who received >48 h of intravenous polymyxin B [120]. The overall 30-day mortality was 37.8% (54/151) and a polymyxin B dose of <1.3 mg/kg/day was significantly associated with higher 30-day mortality (46.5% versus 26.3%; P = 0.02). However, doses of ≥250 mg/day were independently associated with acute kidney injury (AKI) [120].
As discussed in Section 2.2e, CMS/colistin and polymyxin B are handled very differently by the kidneys [107, 109]. The urinary recovery of polymyxin B and colistin (which are extensively reabsorbed by the kidneys) is often very low (<1% of unchanged drug recovered in urine [107]). In contrast, >60% of the intravenous dose of CMS (which is renally excreted) is recovered in urine where subsequent conversion to colistin achieves very high urinary colistin concentrations [88]. Therefore, CMS has significant PK/PD advantages for treating urinary tract infections (UTIs). Sorlí et al. [121] examined the relationship between the AUC/MIC of formed colistin and clinical outcomes in 19 patients who received CMS monotherapy (range <3 MIU/day to ≥9 MIU/day) for UTIs (16 lower UTI and 3 pyelonephritis) caused by extremely drug resistant (XDR) P. aeruginosa. Clinical cure was achieved in 17 of 19 (89.4%) patients with only 5 (29.4%) attaining an ‘optimal’ plasma colistin AUC/MIC of 60 mg·h/L and 1 (5.9%) achieving a plasma colistin concentration of 2.5 mg/L at steady state [121]. Considering the high concentrations of colistin in urine, the currently recommended dosing regimens of CMS are efficacious for treating UTIs caused by XDR Gram-negative pathogens in non-critically ill patients. Despite the low urinary recovery of polymyxin B, several studies have suggested that it is also efficacious for UTIs [122–124] and further well-designed clinical studies are warranted.
To date, neither polymyxin B nor CMS has shown favorable efficacy for the treatment of pneumonia when administered solely intravenously. In a retrospective observational study, critically ill patients with pneumonia caused by MDR A. baumannii were treated with CMS (2.5–5 mg/kg/day CBA; n = 84) or tigecycline (50 mg/12h; n = 84) [125]. Eleven patients in the CMS group and 14 patients in the tigecycline group were treated with combination therapy, primarily involving a carbapenem. The tigecycline group has greater mortality (60.7% vs. 44%, 95% confidence interval (CI) 0.9%–32.4%, P = 0.04), although both CMS and tigecycline had limited efficacy [125]. In a retrospective study with 98 patients diagnosed with ventilator-associated pneumonia (VAP) caused by carbapenem-resistant A. baumannii, 66 were treated with intravenous CMS (2 MIU CBA 8-hourly) and 32 with intravenous ampicillin–sulbactam (3 g 6-hourly) [126]. CMS treatment resulted in a higher 7-day microbiological failure rate [16/33 (48%) vs. 3/17 (18%); P = 0.03] and an increased 30-day mortality (adjusted-odds ratio, 6.5; 95% CI, 1.348–31.342; P = 0.02) [126]. In a prospective study, 45 patients with VAP or ventilator-associated tracheobronchitis (VAT) were administered intravenous polymyxin B at the currently recommended dosage (2.5 mg/kg/day 12-hourly), while 22 patients received other antibacterial therapy (e.g., cephalosporin, levofloxacin, carbapenem or piperacillin-tazobactam) [127]. The polymyxin B group showed a significantly higher crude 30-day mortality than the comparator group (53% versus 27%) [127]. In another prospective study evaluating intravenous polymyxin B (dosing regimens not stated) in patients with VAP caused by carbapenem-resistant P. aeruginosa (n = 29), the infection-related mortality within 14 days was 51.7% [128]. Numerous other clinical studies involving the intravenous polymyxin B or CMS for the treatment of lung infections have shown similarly poor results [129–131].
The poor efficacy of intravenous polymyxins for the treatment of pulmonary infections is probably not surprising given their poor penetration into lung tissue and relatively low concentrations achieved in epithelial lining fluid (ELF) [71, 73, 132, 133]. Significant binding of polymyxins to lung surfactant also contributes to the poor efficacy against pneumonia. The addition of bovine pulmonary surfactant to growth media substantially inhibited bacterial killing by several polymyxin lipopeptides, including polymyxin B and colistin [134], and both polymyxin B and colistin bound strongly to porcine mucin, substantially increasing the MICs (>100-fold) of various Gram-negative bacteria [135, 136]. Thus, the poor efficacy of intravenously administered polymyxins against lung infections mostly likely results from low achieved free drug concentrations in the ELF.
To overcome the poor pharmacokinetics in the lungs after intravenous administration, inhaled polymyxins have been investigated for the treatment of respiratory tract infections such as VAP or those occurring in people with cystic fibrosis [133, 137–139]. Inhalation provides a substantial targeting advantage [140] while simultaneously minimizing systemic exposure and nephrotoxicity [140]. Currently, the majority of polymyxin inhalation studies undertaken in humans have utilized CMS/colistin, and similar studies examining the pharmacokinetics and efficacy of polymyxin B following inhalation are urgently required. In a study involving 20 critically ill patients with VAP, inhaled CMS (1 million IU 8-hourly) resulted in median ELF concentrations of formed colistin (25–75% interquartile range) of 6.7 mg/L (4.8–10.1 mg/L), 3.9 mg/L (2.5–6.0 mg/L) and 2.0 mg/L (1.0–3.8 mg/L) at 1, 4 and 8 h, respectively; serum concentrations were ~5-fold lower than in ELF [139]. The median AUC following a single dose was 29.8 mg·h/L (25–75% interquartile range: 21.9–54.5 mg·h/L). Boisson et al. examined two dosing regimens of inhaled CMS in 12 critically ill patients: 2 MIU of nebulized CMS administered over 30 min, followed by 2-MIU doses of CMS administered intravenously over 60 min every 8 h until the end of the treatment [141]. Colistin ELF concentrations following a single nebulized CMS dose ranged from 9.53 to 1,137 mg/L, compared to 0.15–0.73 mg/L following intravenous administration. PK/PD analysis indicated substantially higher antimicrobial efficacy following nebulization than after intravenous administration, with a time to ‘total-kill’ of ~12 h following inhalation of the first dose of CMS (assuming a starting inoculum of 106 CFU/mL); simulations for intravenous CMS alone were unable to achieve this target [141].
Given the different pharmacokinetics between inhaled and intravenously administered CMS/colistin described above, it is not surprising that a systematic review and meta-analysis concluded that the overall clinical response (improvement and cure) in patients with pneumonia caused by MDR Gram-negative pathogens was significantly higher with intravenous plus aerosolized polymyxin therapy than with intravenous polymyxin therapy alone [OR = 1.81, 95% CI 1.30–2.53; P = 0.0005] [142]. Significantly higher rates of pathogen eradication (OR = 1.66, 95% CI 1.11–2.49; P = 0.01) and lower all-cause mortality were also observed in patients treated with inhalation plus intravenous administration compared with intravenous administration alone (OR = 0.69, 95% CI 0.50–0.95; P = 0.02) [142]. However, a single retrospective clinical study examining polymyxin B for the treatment MDR Gram-negative bacterial respiratory tract infections did not detect any differences in mortality or favorable clinical outcome between intravenous plus aerosolized administration versus intravenous administration alone, although the number of patients (n = 25) was small [143]. Well-designed large clinical studies are required to evaluate the efficacy of inhaled polymyxins in patients with respiratory tract infections.
h). Pharmacokinetics/Pharmacodynamics of intraventricular and intrathecal polymyxins for central nervous system infections
Numerous case studies have examined the efficacy of CMS after intravenous, intrathecal, or intraventricular administration for the treatment of CNS infections [144–148]. However, very few have measured the concentrations of CMS and formed colistin achieved in the cerebrospinal fluid (CSF) using specific and accurate chromatographic assays. Penetration of CMS and formed colistin into the CSF following intravenous administration of CMS is minimal [144, 145]. In a prospective study involving five critically ill patients without CNS infection administered 3 MIU of CMS intravenously 8-hourly, the CSF concentrations of formed colistin ranged from 72 to 152 ng/mL and the maximum CSF/serum concentration ratio of colistin was 0.074 ± 0.002, indicating extremely low penetration of CMS and colistin into CSF (Table 3) [144]. In the same study, three patients with external ventricular drain-associated ventriculitis caused by A. baumannii and K. pneumoniae (EVDViv group) received the same intravenous dose of CMS, while four similar patients additionally received 10 mg (125,000 IU) of intraventricularly administered CMS once daily (EVDVcomb group) [144]. The range of CSF concentrations and the maximum CSF/serum ratio were, respectively, 119–259 ng/mL and 0.118 in the EVDViv group and 611–1,449 ng/mL and 0.456 in the EVDVcomb group. CSF colistin concentrations above 0.5 mg/L were achieved only in patients receiving both intraventricular and intravenous administration [144]. Imberti et al. examined CSF pharmacokinetics of colistin following intraventricular CMS administration in 9 adult patients with CNS infection caused by pandrug-resistant K. pneumoniae (6 patients), A. baumannii (2 patients), and P. aeruginosa (1 patient) [146]. The doses of CMS administered were 2.61 mg 12-hourly (3 patients), 2.61 mg 24-hourly (1 patient), 5.22 mg 12-hourly (3 patients), and 5.22 mg 24-hourly (2 patients). Doses of CMS >5.22 mg/day produced colistin CSF concentrations continuously above 2 mg/L throughout the dosing interval of 12 or 24 h, with trough concentrations ranging between 2.0–9.7 mg/L [146]. Microbiological cure was achieved in 8 of 9 patients. Given that external CSF efflux is variable and can influence the clearance of colistin in CSF, the authors suggested that the daily dose of 10 mg CMS recommended by the Infectious Diseases Society of America (IDSA) may be more prudent [146]. A subsequent case report and review examined intrathecal/intraventricular CMS in external ventricular device (EVD)-related infections caused by MDR Gram-negative bacteria [147]. The authors suggested that the intrathecal/intraventricular dosage of CMS should be at least 10 mg/day and possibly increased to 20 mg/day in patients with previous exposure to colistin (due to the possible occurrence of heteroresistance), with recurrent CNS infection, or with CSF drainage through the EVD. When comparing the efficacy of IV CMS and intrathecal/intraventricular CMS against extensively drug-resistant A. baumannii meningitis, the CSF sterilization rate of intravenous CMS alone was significantly lower than intrathecal/intraventricular CMS (33.3% vs 100%, respectively; P = 0.009) [148]. A recent retrospective clinical study suggested early intraventricular polymyxin B supplemented by continuous EVD treatment is an effective and safe strategy for intracranial infections caused by MDR/XDR Gram-negative bacteria [149]. Clearly, our current knowledge of the pharmacokinetics of CMS/colistin and polymyxin B in CSF following intrathecal/intraventricular administration is limited and the corresponding fAUC/MIC targets for polymyxin B and colistin to achieve significant bacterial killing remain unknown. Large prospective clinical PK/PD studies are required to optimise the dosage regimens of intrathecal/intraventricular polymyxins [150].
Table 3.
Colistin concentrations in CSF with different dosage regimens and disease conditions in patients [144].
| Patients | Dosage regimen | CSF concentration of colistin | Maximum CSF/serum concentration ratio of colistin |
|---|---|---|---|
| 5 critically ill patients without CNS infection | 3 MIU CMS iv q8h | 72 – 152 ng/mL | 0.074 ± 0.002 |
| 3 patients with external ventricular drain-associated ventriculitis | 3 MIU CMS iv q8h | 119 – 259 ng/mL | 0.118 |
| 4 patients with external ventricular drain-associated ventriculitis | 3 MIU CMS iv q8h + 10 mg (125,000 IU) intraventricularly | 611 – 1,449 ng/mL | 0.456 |
2.3. Toxicity and therapeutic drug monitoring
The most common adverse effect of systemically administered CMS and polymyxin B are nephrotoxicity and neurotoxicity [150–153]. While early studies indicated that up to half of all patients receiving either drug experienced nephrotoxicity [154, 155], inconsistent definitions of nephrotoxicity have made estimating the true incidence of AKI difficult [152, 153]. More recently, several clinical observational studies used the RIFLE criteria (i.e., Risk, Injury, Failure, Loss of kidney function, and End-stage kidney disease) to compare the rates of nephrotoxicity between patients receiving intravenous CMS and polymyxin B [156–160]. The incidence of AKI in patients treated with intravenous CMS and polymyxin B ranged from ~26–74% and ~20–46%, respectively, with the incidence of AKI with CMS being substantially greater than with polymyxin B in all but one study. A meta-analysis of 95 observational cohort studies reported no significant difference in nephrotoxicity between either polymyxin, with the prevalence of nephrotoxicity being 26.7% (95% CI 22.8–30.9) for CMS and 29.8% (95% CI 23.8–36.7) for polymyxin B [161]. Notably, no randomized controlled trials were included. A recent systematic review [152] estimated the rates of nephrotoxicity and neurotoxicity in patients treated with polymyxin B and CMS and reported an overall nephrotoxicity rate of 28.1% (95% CI 25.9–30.7), similar to that reported by Oliota et al. [161]. However, consistent with previous meta-analyses [162, 163], pairwise analysis of comparative studies did show the rate of nephrotoxicity was significantly lower with intravenous polymyxin B (OR 1.65 [95% CI 1.16–2.35]) and higher doses were associated with a higher risk of nephrotoxicity (OR 1.89 [95% CI 1.40–2.55]) [152]. Nevertheless, nephrotoxicity remains the major dose-limiting adverse effect for both polymyxins and severely impacts their optimal use.
Knowledge of the mechanisms underpinning polymyxin-induced nephrotoxicity has substantially increased in recent years, with apoptosis in renal proximal tubular cells now known to play a key role [158, 164, 165]. In a single-cell synchrotron-based X-ray fluorescence microscopy study, significant accumulation of polymyxins in human kidney tubular cells was observed with concentrations up to ~4,760-fold higher than the extracellular concentrations [166]. The extraordinarily high intracellular polymyxin concentrations cause oxidative stress, mitochondrial dysfunction, cytochrome c release, activation of multiple apoptosis pathways and eventually cell death [158, 164, 165]. With whole-genome CRISPR screening, we recently reported that potassium channels Kir4.2 and Kir5.1 and mitochondrial respiratory complex I and IV are critical for polymyxin-induced nephrotoxicity in human kidney HK-2 cells [167]. Neurotoxicity involving both the central and peripheral nervous systems [168, 169] is another major adverse effect of polymyxins which may lead to respiratory failure and apnea [152, 153, 170–174]. Other adverse effects associated with polymyxin use include skin hyperpigmentation [175–177], pulmonary toxicity [178–181] [182] and immunotoxicity [183].
Given the narrow therapeutic window of polymyxins, therapeutic drug monitoring (TDM) and adaptive feedback control (AFC) have been recommended to individualize dosing of polymyxin B and CMS and reduce their toxicity in patients [114, 184, 185]. A meta-analysis showed that the rate of nephrotoxicity was significantly associated with polymyxin B exposure [185]. Using multiple linear regression models, Kim et al. proposed that a blood sample at 2-h post-dose (for monitoring efficacy) and a trough blood sample (for monitoring renal toxicity) are most appropriate for colistin TDM in critically ill patients [186]. Similarly for polymyxin B, a limited sampling strategy undertaken in Chinese patients showed that a two-point model (i.e., immediately before a dose [i.e., trough] and 2-h post-dose) could predict polymyxin B exposure (r2 > 0.98), while a four-point model (C1h, C1.5h, C4h, and C8h) performed best in predicting the AUC of polymyxin B (r2 > 0.99) [187]. In another recent study in Chinese patients, a C6 scheme (i.e., sampling at 6 h after the end of 1-h infusion) was proposed for accurate prediction of AUCss and the TDM of polymyxin B [188]. A target of 1.9–4.2 mg/L at C6 could achieve polymyxin B AUCss of 50–100 mg·h/L [188] which has been demonstrated efficacious in a mouse thigh infection model [71]. Monte Carlo simulations showed that without adaptive feedback control, the current polymyxin B dosage regimens can only achieve the AUCss,0–24 of 50–100 mg·h/L in 71% of simulated subjects [185]. However, using a single pharmacokinetic sample collected at 24 h and an adaptive feedback control algorithm, >95% of simulated subjects with personalized dosing regimens can achieve AUCss,0–24 values within the target AUCss,0–24 window [185].
In a TDM study with 102 patients receiving intravenous CMS, Sorli et al. reported that the trough plasma concentration ‘breakpoints’ of formed colistin to predict AKI (defined by the RIFLE criteria) were 3.33 mg/L and 2.42 mg/L on day 7 and at the end of therapy, respectively [189]. In a large clinical PK/PD study, the relationship of AKI and colistin exposure in 153 critically ill patients revealed that the rate of AKI was substantially higher when the Css,avg of formed colistin was >2 mg/L [190]. Recent international dosing guidelines recommend that at steady state (AUCss,0–24) a colistin plasma AUCss,0–24 of ~50 mg·h/L (i.e., Css,avg of ~2 mg/L) should be targeted for maximum efficacy; concentrations higher than this increase both the incidence and severity of AKI [114]. Further clinical investigations examining the clinical PK/PD/TD of polymyxin B and colistin and how their therapeutic benefit can be maximised in various patient populations are warranted.
2.4. New-generation polymyxins under development
The clinical utility of polymyxin B and colistin is mainly limited by nephrotoxicity, acute toxicity, and poor pulmonary pharmacokinetics following intravenous administration. New-generation polymyxin lipopeptides MRX-8, SPR206 and QPX9003 show enhanced bacterial killing in lung infection models [191–198] and have entered Phase-I clinical trials. MRX-8 is a novel polymyxin analog containing a fatty acyl tail attached via an ester bond [199]. It is active against P. aeruginosa, K. pneumoniae, E. coli and A. baumannii in the mouse thigh and lung models; and is currently in Phase-1 clinical trial [199]. SPR206 has a β-branched aminobutyrate N-terminus with an aryl substituent that showed enhanced in vitro activity against P. aeruginosa and A. baumannii [200, 201]. SPR206 and polymyxin B have similar in vivo efficacy against Acinetobacter spp., E. coli, Klebsiella spp. and P. aeruginosa [201]. SPR206 was active against P. aeruginosa and A. baumannii in the mouse lung infection model following subcutaneous administration. The GLP 14-day toxicology assessment of SPR206 indicated that renal toxicity was monitorable and reversible in monkeys [201]. Phase I clinical studies showed that SPR206 was well-tolerated in healthy subjects who received doses up to 100 mg/8h (i.e., 300 mg/day) for 14 consecutive days [202, 203]. MRX-8 is a novel polymyxin analog containing a fatty acyl tail attached via an ester bond [199]. It is active against P. aeruginosa, K. pneumoniae, E. coli and A. baumannii in the mouse thigh and lung models; and is currently in Phase-1 clinical trial [199]. QPX9003 is another promising new-generation polymyxin with much-enhanced safety and activity over the current polymyxins [197]. The development of QPX9003 was underpinned by a total chemical synthesis strategy and extensive pharmacological findings on the structure-activity and structure-toxicity relationships of polymyxins. In vitro susceptibility testing of QPX9003 showed improved activity compared to colistin and polymyxin B against MDR Gram-negative pathogens; and notably, animal studies showed much-reduced acute toxicity and nephrotoxicity and improved bacterial killing, especially for lung infections, compared to polymyxin B and colistin [198]. Phase-1 clinical trials of QPX9003 are currently underway in the USA. The development of new-generation polymyxins is very challenging due to the narrow chemical space and the interplay between the chemical structure and antibacterial activity, acute toxicity, nephrotoxicity and pharmacokinetics, all of which must be optimised to enhance the therapeutic index.
3. Antimicrobial glycopeptides against Gram-positive bacteria
3.1. Chemistry and mechanisms of activity and resistance
The glycopeptide antibiotics vancomycin, teicoplanin and semisynthetic telavancin, dalbavancin and oritavancin have a broad antibacterial spectrum against Gram-positive bacteria, especially against Staphylococcus, Enterococcus and Clostridium difficile [204, 205]. These glycopeptide antibiotics consist of a group of glycosylated cyclic or polycyclic heptapeptides connected to a glycosyl side chain (Figure 4). They inhibit cell wall synthesis of Gram-positive bacteria by binding to cell wall precursors rather than acting directly on an enzyme active site [206, 207]. This section mainly focuses on the two major glycopeptide antibiotics vancomycin and teicoplanin.
Figure 4.

Structures of vancomycin, teicoplanin, telavancin, oritavancin and dalbavancin.
Vancomycin (from the root word “vanquish”) was first isolated from Amycolatopsis orientalis (previously known as Streptomyces orientalis) and has been used clinically against penicillin-resistant Staphylococci since 1955 [208]. The vancomycin cyclic heptapeptide core contains both aromatic (5 of 7) and aliphatic (2 of 7) residues with extensive cross-linking between the aromatic amino acid residues; the aliphatic residues form a unique binding pocket for the D-alanine (D-Ala) dipeptide D-Ala-D-Ala located at the C-terminus of the peptidoglycan precursor (discussed below) [209, 210]. The closely related teicoplanin was isolated from Actinoplanes teichomyceticus ~30 years after vancomycin [211] and was first marketed in Europe in the 1990s [212]. Similar to vancomycin, teicoplanin consists of two sugar moieties attached to a core cyclic heptapeptide [213]. However, it also contains a hydrophobic substituent that leads to better cellular and tissue penetration and substantially differentiates its properties from vancomycin [210, 214]. Telavancin is a semisynthetic derivative of vancomycin with a lipophilic decylaminoethyl chain and negatively charged hydrophilic group [215]. The decylaminoethyl chain enhances the binding affinity to D-Ala-D-Ala containing peptidoglycan intermediates and provides activity against MRSA and non-vanA enterococci [216, 217]. The negatively charged hydrophilic group of dalbavancin improves the tissue distribution, metabolism and clearance [217, 218]. Dalbavancin is a semisynthetic derivative of the teicoplanin-like antibiotic A-40926 produced by Nonomuria spp. [217]. It differs from teicoplanin by the presence of a chlorine atom and a terminal methylamino group, the addition of a dimethylaminopropanol-amine group via amidation of the C-terminal carboxyl moiety, and lack of the acetylglucosamine group [217]. Dalbavancin has an acylaminoglucuronate instead of an acylglucosamine on ring 4 [217]. The lipophilic side chain dramatically extends the terminal t1/2 of dalbavancin to 6–11 days, allowing for once-weekly administration, as well as enhancing antimicrobial activity by increasing the affinity for the terminal D-Ala-D-Ala [217, 219]. The amidated carboxyl side group enhances anti-staphylococcal activity of dalbavancin [220]. Oritavancin is a synthetic derivative of the naturally occurring glycopeptide chloroeremomycin (A82846B) produced by Amycolatopsis orientalis [221]. It has a hydrophobic N-alkyl-p-chlorophenylbenzyl substituent attached to the vancosamine, with an additional aminosugar (4-epi-vancosamine) on the phenylserine hydroxyl group [222]. The hydrophobic side chain and extensive protein binding of oritavancin produces a prolonged terminal t1/2 similar to that of dalbavancin, and probably accounts for the strong antimicrobial effect by anchoring the molecule in the cell membrane and stabilizing the formation of oritavancin dimers [223]. The N-alkyl-p-chlorophenylbenzyl substituent in oritavancin enhances the activity against both vancomycin-susceptible enterococci (VSE) and VRE, including vanA enterococci [224, 225]. Oritavancin has a similar antibacterial spectrum to vancomycin, although it is more active against streptococci [226].
Peptidoglycan synthesis for the bacterial cell wall begins in the cytoplasm where a racemase converts L-Ala to D-Ala followed by formation of a D-Ala-D-Ala dipeptide by a ligase [227, 228]. This dipeptide is then added to uracil diphosphate-N-acetylmuramyl-tripeptide to form uracil diphosphate-N-acetylmuramyl-pentapeptide followed by the addition of N-acetylglucosamine. Translocation to the outer surface of the cytoplasmic membrane occurs via the undecaprenol lipid carrier, whereupon transgylcosylation incorporates the N-acetylmuramyl-pentapeptide into nascent peptidoglycan, with cross bridges subsequently formed via transpeptidation [227, 228]. Glycopeptides do not enter the bacterial cytoplasm but rather exert their antibacterial activity on the peptidoglycan precursors following the translocation step [229]. The heptapeptide nucleus of the glycopeptide binds with high affinity to the D-Ala-D-Ala C-terminus of the pentapeptide precursor via hydrophobic van der Waals bonds and five hydrogen bonds [227, 230]. Once bound, these glycopeptide molecules (~1,450 Da for vancomycin and ~1,880 Da for teicoplanin) creates steric hindrance that blocks the formation of β−1,4 glycosidic bonds by transglycosylation and subsequent cross-linking by transpeptidation [207, 227]. The formation of vancomycin dimers via self-association is important in this interaction, which increases the binding affinity by orders of magnitude [231, 232]. Teicoplanin interacts with its target as a monomer and its hydrophobic moiety interacts with the lipid bilayer of bacterial membrane, localising the antibiotic in close proximity to its substrate [232, 233]. In addition to substrate sequestration, vancomycin also antagonizes peptidoglycan remodelling (“autolysis”) [234, 235] and selectively inhibits ribonucleic acid synthesis in S. aureus [236]. Gram-negative bacteria are intrinsically resistant to the glycopeptides as the outer membrane prevents these antibiotics from entering the cell, rendering them unable to reach their peptidoglycan precursor targets.
Telavancin, dalbavancin and oritavancin have lipid side chains that anchor the molecule to the cell membrane and increase potency compared to vancomycin and teicoplanin [217, 237, 238]. Such anchoring in the membrane may also partially destabilize the membrane, resulting in loss of membrane potential [216, 217]. These effects may increase the activity of all three semisynthetic glycopeptides and minimise the emergence of resistance [238]. Telavancin strongly inhibits peptidoglycan biosynthesis, with a 10-fold greater activity than vancomycin in terms of inhibition of peptidoglycan biosynthesis in MRSA bacterial cells [239, 240]. Telavancin also causes rapid concentration-dependent effects on bacterial cell membrane potential and structure, which explains its rapid bactericidal activity when compared to vancomycin [216, 240]. Similar effects have been observed with oritavancin [238, 241, 242]. Both dalbavancin and oritavancin form dimers which increase binding affinity for the D-Ala-D-Ala target site [223, 242]. Notably, oritavancin dimers bind to both peptidoglycan precursors D-Ala-D-Ala and D-Ala-D-Lac, which supports its good antibacterial activity against vanA-carrying vancomycin-resistant bacteria [243–245]. In addition, oritavancin can also inhibit bacterial RNA synthesis [238].
Emergence of vancomycin resistance in the clinic was relatively slow with the first reports of resistance in Enterococcus species in Europe in 1988 [246, 247]. The VRE spread rapidly throughout US hospitals in the 1990s, European hospitals in the 2000s, and ultimately worldwide [248]. Horizontal transfer of the vanA gene from vancomycin-resistant enterococci led to the first report of VRSA in 2002 [249]. Resistance to teicoplanin is also now well documented [250, 251]. The main mechanism of resistance to glycopeptides involves target modification, such as the production of either of the two pentapeptide precursors with a low binding affinity for the glycopeptide. Replacement of D-Ala-D-Ala by D-Ala-D-lactate (D-Lac) removes an essential hydrogen bond, reducing vancomycin affinity 1,000-fold compared to the normal dipeptide and causing high-level vancomycin and teicoplanin resistance (i.e., MIC >64 mg/L vancomycin) [252]. The vanA operon is the most important cause of this type of vancomycin resistance and is the most frequently encountered mechanism of resistance in enterococci [253]. The vanB, vanD, vanF and vanM operons also cause vancomycin resistance via a similar mechanism [254, 255]. The development of heterogeneous vancomycin-intermediate S. aureus (hVISA) is believed to be an epigenetic process rather than due to gene mutations [256]. VISA develops from hVISA in individuals treated with glycopeptide antibiotics over extended time periods due to the accumulation of mutations with various effects on the cell wall [257]. In VRSA, both the vanA gene and the accumulation of gene mutations related to cell wall synthesis (that result in the development of thicker cell walls) may contribute to vancomycin resistance [257–259]. Target modification also occurs when D-Ala-D-Ala is replaced by D-Ala-D-Serine (D-Ser), creating a steric hindrance that results in a 6-fold reduction in affinity and a low level of vancomycin resistance (i.e., MIC 4 to 32 mg/L) [260]. The vanC, vanG, vanE, vanL and vanN resistance cassettes are involved in this modification. Compared to vancomycin and teicoplanin, the rate of resistance to the three semisynthetic glycopeptides is relatively low [228]. Telavancin and dalbavancin have a low frequency of spontaneous resistance [261, 262]. However, in Enterococcus the vanA genotype is moderately resistant to telavancin and dalbavancin (MIC90 < 32 mg/L), but is generally susceptible to oritavancin (MIC90 < 2 mg/L) [217]. Oritavancin resistance (MIC ≤ 16 mg/L) has been reported in the laboratory in Enterococcus isolates carrying vanA, vanB and vanZ [263]. VanZ reduces the binding of these three semisynthetic glycopeptides to S. aureus and St. pneumoniae cells [264].
Vancomycin-resistant Enterococcus (VRE), vancomycin-intermediate S. aureus (VISA) and vancomycin-resistant S. aureus (VRSA) are increasingly reported in patients [265]; therefore, optimising glycopeptide PK/PD is critical for maximising the efficacy and minimizing resistance.
3.2. Pharmacodynamics and pharmacokinetics
Vancomycin and teicoplanin are mainly active against Gram-positive bacteria including Staphylococcus, Streptococcus, Enterococcus and Clostridium [266–268]. S. aureus with an MIC of ≤2 mg/L (vancomycin-sensitive S. aureus, VSSA) are considered susceptible to vancomycin, whereas intermediacy and resistance are defined by MICs of 4–8 mg/L (VISA) and ≥16 mg/L (VRSA), respectively (CLSI, M100-S31). For non-aureus Staphylococcus spp. and Enterococcus spp., the equivalent breakpoints are ≤4, 8–16 and ≥32 mg/L, respectively (CLSI, M100-S31). For teicoplanin, the corresponding breakpoints for all Staphylococci and Enterococcus spp. are ≤8, 16 and ≥32 mg/L (CLSI, M100-S31). The MIC90 of teicoplanin against Streptococcus pneumoniae is approximately one quarter that of vancomycin [269–271] and its activity against Enterococcus is similar to or slightly better than vancomycin [272]. Compared with vancomycin, telavancin, dalbavancin and oritavancin have better activity against MDR Gram-positive bacteria, longer half-lives, and are suitable for daily or weekly administration [217, 223]. In vitro studies have shown that telavancin and dalbavancin are effective for VISA [273], whereas oritavancin is effective against both VISA and VRSA and has potent activity against vanA-mediated VRE (MIC90 = 0.06–0.5 mg/L). Telavancin showed a low frequency of spontaneous resistance in staphylococci and enterococci [218]. Resistance to telavancin and dalbavancin is associated with vanA [223], although MIC increases are low after serial passaging in media containing these two drugs [261, 262]. Oritavancin-resistant clinical isolates are rarely reported to date.
In vitro static and dynamic time-kill studies of vancomycin and teicoplanin showed time-dependent bactericidal killing against common Gram-positive bacteria, including S. aureus and S. epidermis [274–278]. In a one-compartment PK/PD model (vancomycin t1/2, 6 h) against a single strain of methicillin-resistant S. aureus, no difference was observed in bacterial killing with initial concentrations ranging from 5–40 mg/L [275]. A similar lack of concentration-dependent killing was observed against S. aureus and S. epidermis with vancomycin concentrations ranging from 2–64× MIC [274]. Killing by vancomycin is subject to a moderate to high inoculum effect [279–281]. Telavancin, dalbavancin and oritavancin all exhibit concentration-dependent killing against Gram-positive bacteria in vitro [223, 282–284]. In an one-compartment dynamic model simulating a typical human dosage regimen of vancomycin (1 g 12-hourly; t1/2, 6 h; AUC0–24 ~600 mg·h/L) against endocardial vegetations of methicillin-susceptible S. aureus (MSSA), the maximum bacterial killing at low (5.5 log10 CFU/mL) and high (9.5 log10 CFU/mL) inocula was 3.28 and 2.38 log10 CFU/mL, respectively [281]. The inoculum effect of teicoplanin against staphylococci and streptococci appears greater than with vancomycin [285, 286], while no inoculum effect was observed for S. epidermis [287]. The PAE of vancomycin against MSSA, MRSA, S. epidermis and E. faecalis is ~1–2 h and concentration-dependent, although longer PAEs (up to ~5 h) were reported against some strains of S. epidermis [274, 288, 289]. The PAE of teicoplanin against most isolates of MSSA is ~1–2 h, with longer PAEs for MRSA strains [287, 288]; a concentration-dependent PAE (range, ~0.9–10 h) was also reported against E. faecalis [287, 288].
Oral glycopeptides are used for the treatment of intestinal infections due to poor absorption [217, 290]. For the treatment of systemic infections, both vancomycin and teicoplanin are administered intravenously and subsequently eliminated primarily via the kidneys, with >80% eliminated unchanged in urine [213, 291–293]. Only ~5–9% of the administered dose of vancomycin is cleared non-renally, possibly by hepatic conjugation [294, 295]. Therefore, it is not surprising that in both adults and children the pharmacokinetics of vancomycin and teicoplanin are significantly affected by renal function, with renally-impaired patients usually having increased drug exposure in blood and increased risk of toxicity [296–299]. The elimination t1/2 of vancomycin in adult patients with normal renal function is 6–12 h [300, 301]. The t1/2 is longer in children (median 21.1 h, mean 20.5 h), especially in low-birth-weight infants and neonates [302, 303] and patients with renal impairment, which necessitates dose adjustment (increased as creatinine clearance decreased) [304]. The clearance (CL) of vancomycin derived from population pharmacokinetic analyses in adults varies, with a median of 3.22 L/h (equivalent to ~0.0458 L/h/kg) and a range of 0.334–8.75 L/h (0.0054–0.1279 L/h/kg), decreasing with the decline in renal function [305]. Higher CL values (greater than the third interquartile value of 4.9 L/h) were observed in certain patient groups including Chinese and obese patients, patients with post-craniotomy meningitis, and patients with a creatine clearance < 80 mL/min, whereas lower clearance values (lower than the first interquartile value of 2.32 L/h) were reported in other patient groups including Thai patients, patients with post-sternotomy mediastinitis, and mechanically-ventilated ICU patients [305]. Remarkably wide variation was also observed for the total volume of distribution of vancomycin (7.12–501.8 L) [305]. A recent study in 182 Chinese neonates reported a mean volume of distribution of 0.86 L, with physiological maturation, renal function, and concomitant use of vasoactive agents significantly affecting vancomycin pharmacokinetics [306]. Teicoplanin displays a much longer elimination t1/2 of ~30–170 h which substantially increases the time to achieve steady state [307–312]. Therefore, a loading dose is recommended to rapidly achieve target concentrations [311, 313, 314]. The total body clearance of teicoplanin is ~0.00036–0.019 L/h/kg and the volume of distribution at steady state is 0.28–1.1 L/kg in aduIts with impaired renal function [315].
Telavancin exhibited linear pharmacokinetics in healthy subjects over the dose range of 7.5–15 mg/kg when infused over 30–120 min [290], with a long t1/2 (7 – 9 h) permitting once-daily dosing [217, 316]. Its penetration into skin blister fluid is good, with an AUCblister of ~40% of AUCplasma following a 1-h infusion of 7.5 mg/kg/day telavancin for a total of three doses, generating sufficient concentrations in both plasma (Cmax, 84.8 ± 5.3 mg/L) and blister fluid (Cmax, 16.0 ± 2.0 mg/L) for pathogens causing skin and soft tissue infections (e.g., S. aureus and Streptococcus spp.) [317]. Good penetration into ELF has also been reported for telavancin in healthy volunteers, with a median AUCELF/fAUCplasma penetration ratio of 0.73 (compared to 0.39 for vancomycin) [318]. In a separate study in healthy subjects administered with 10 mg/kg/day telavancin, good penetration was observed in ELF (3.73 ± 1.28 mg/L at 8 h and 0.89 ± 1.03 mg/L at 24 h, while the mean plasma concentration at 24 h was 7.28 mg/L) and alveolar macrophages (mean alveolar macrophage/plasma ratio of 6.67 at 24 h) [319]; these concentrations well exceed the MIC90 for MRSA. Pulmonary surfactant did not affect the in vitro antibacterial activity of telavancin. In a rabbit model, only a small percentage of unbound telavancin crossed the meninges (1% and 2% in un-inflamed and inflamed meninges, respectively) [320]. Clearance of telavancin is primarily via the renal route, with 65% (for 7.5 mg/kg) and 72% (for 15 mg/kg) recovered in urine as unchanged drug [321]. Therefore, clearance is substantially reduced and the t1/2 extended in patients with renal dysfunction, necessitating dosage adjustments [322].
Dalbavancin and oritavancin also display linear pharmacokinetics in humans [217, 237, 323, 324]. In rats, tissue penetration of dalbavancin is extensive with the highest concentrations occurring in the liver and kidney [325]; for oritavancin, significant amounts of the administered dose were detected in the liver (59–64%) kidneys (2.7%), spleen (1.8%) and lungs (1.7%) [217]. Blister fluid concentrations of dalbavancin are higher than both telavancin and oritavancin, with penetration ratios of ~0.60 (based on the AUC over 7 days after a single intravenous dose of 1,000 mg) [326] and 0.83–1.11 (method of calculation not reported) [327] having been reported. Dalbavancin is eliminated via both renal and non-renal mechanisms, with ~33% of the dose excreted unchanged in urine [327]. Oritavancin is eliminated in urine and faeces, with renal elimination the major route [328]. Over 14 days, <5% of an oritavancin dose was excreted in urine and <1% eliminated in faeces in healthy subjects [328].
Vancomycin binds primarily to IgA and albumin, with no binding to IgM, IgG or α−1 acid glycoprotein [329, 330]. The reported vancomycin protein binding in human plasma in the literature has varied considerably, ranging anywhere from 0 to 98% [330, 331], although 50–55% is most often reported [332, 333]. The variability in binding likely results from the different assay condition used [334]; however, patient characteristics also play a key role with significantly lower plasma binding (median bound fraction) of ~19% in paediatric patients compared to ~40% in three adult patient cohorts (hematology, ICU, and orthopaedics) [335]. In contrast, teicoplanin is ~90–95% bound to serum albumin in human plasma [336, 337]. Although telavancin binds extensively to plasma proteins (~90% bound in human plasma), the binding affinity is weak and its antibacterial activity is not substantially affected in the presence of serum [217]. Dalbavancin and oritavancin also display high protein binding in human plasma (93–98% binding for dalbavancin and 86–90% for oritavancin) [217, 237, 323, 324]. The extensive (but reversible) protein binding of dalbavancin and oritavancin contributes to their long terminal t1/2 in humans (147–258 h for dalbavancin and ~400 h for oritavancin) which permits daily (for oritavancin) or once-weekly (for dalbavancin) dosing [217, 323, 327, 338].
The large size and hydrophilic nature of vancomycin and teicoplanin limit their tissue penetration [315, 339, 340]. Penetration of vancomycin into skin tissue is very low with a median tissue-to-plasma ratio of 0.1 (range, 0.01–0.45 mg/L) and 0.3 in diabetic and non-diabetic patients (range, 0.46–0.94 mg/L), respectively [341]. Penetration of vancomycin into the CSF of patients with uninflamed meninges is also low, with CSF concentrations ranging from 0–3.45 mg/L and corresponding CSF-to-serum ratios of 0–0.18 [342–344]. In a recent study with 21 critical care patients with proven or suspected ventriculitis, the median serum Cmax and Cmin of vancomycin following 4-h infusions targeting serum trough concentrations of 10–15 mg/L were 25.7 mg/L (range, 10.6–50.8) and 9.60 mg/L (range, 4.46–23.6), respectively; the corresponding median concentrations of vancomycin in CSF were 0.65 mg/L (<0.24–3.83 mg/L) and 0.58 mg/L (<0.24–3.95 mg/L) [343]. Clearly, there was high inter-subject pharmacokinetic variability in the CSF penetration of vancomycin. Penetration of vancomycin into CSF was substantially improved when the meninges were inflamed (CSF concentrations of 6.4–11.1 mg/L and CSF-to-serum ratios of 0.36–0.48) [342]. In a recent systematic review, the CSF-to-serum ratio of vancomycin ranged 0.05–0.17 in patients with ventriculitis, 0.06–0.81 in meningitis, 0.00–0.36 in other types of infections, and 0–0.13 in uninfected patients [344]. Nevertheless, despite the low and variable penetration of vancomycin into CSF after intravenous administration, clinical cure was achieved in 100% of patients with ventriculitis and 83% of patients with meningitis. No factors were identified to predict vancomycin penetration and CSF vancomycin concentrations did not predict clinical cure [344].
Although vancomycin is recommended for the treatment of ventilator-associated pneumonia (VAP) caused by MRSA [345], studies examining the penetration of vancomycin into lung tissue in diverse patient populations have shown variable, although generally poor, results [346–349]. An overall ELF penetration of ~50% has been reported in healthy volunteers [349, 350]. However, in 14 critically ill patients receiving 15 mg/kg of intravenous vancomycin for no less than 5 days, the mean ELF vancomycin concentration was 4.5 mg/L (range, 0.4–8.1 mg/L), substantially lower than the corresponding plasma concentration of 24 mg/L (range, 9.0–37.4 mg/L), with an overall blood-to–ELF penetration ratio of 6:1 [346]. In another study involving 10 critically ill patients receiving 30 mg/kg/day 6-hourly via intravenous infusion, vancomycin in the ELF was only detected in 4 patients with a mean ELF concentration in these patients of 2.03 ± 0.49 mg/L (range, 1–2.77 mg/L) and a corresponding mean plasma concentration of 22.2 ± 0.83 mg/L [347]. The generally poor penetration of vancomycin into ELF after intravenous administration is likely due to protein binding and a limited ability to cross the alveolar-capillary membrane [351]. Therefore, intravenous vancomycin for nosocomial pneumonia and VAP may result in poor clinical responses or therapeutic failure [205]. In a phase-II randomized, double-blind, multicenter study vancomycin (1 g/12 h IV) had a clinical cure rate of 52.2% (12 of 23) and a 28-day death rate of 21.7% (5 of 23) for patients with HAP [352]. Larger clinical PK/PD studies are warranted to examine the efficacy of intravenous vancomycin for the treatment of pneumoniae caused by Gram-positive organisms.
Compared to intravenous administration, nebulized vancomycin can produce substantially higher lung concentrations with reduced systemic toxicity [346, 347]. AeroVanc™, a dry powder formulation of vancomycin for inhalation, is currently undergoing phase III clinical trials [353]. In phase I and II studies, single doses (16, 32, or 80 mg) were well tolerated with the average Cmin of vancomycin in sputum (3.05 mg/L and 8.0 mg/L for the 32 mg and 80 mg doses, respectively) remaining above the usual MRSA MIC values (typically 0.5–2.0 mg/L) for up to 24 h; mean systemic absorption was 49% [354, 355]. Based on the limited PK/PD literature, inhaled vancomycin may be superior to intravenous administration for the treatment of respiratory tract infections.
For teicoplanin, high drug concentrations following intravenous administration have been reported in the abdominal cavity, liver, bile, pancreas and mucosal tissues [213], however lower concentrations were reported in cartilage [315] and fat [356], with variable concentrations reported in bone [357]. As penetration into the CSF is poor even with meningeal inflammation, intravenous teicoplanin is rarely used to treat CNS infection. In a case study, none of four patients receiving intravenous teicoplanin (400 mg/day) for neurosurgical shunt infections achieved CSF concentrations above 1 mg/L [358]. Compared to vancomycin, much less is known on the penetration of teicoplanin into the lung. In a rat study, peak concentrations of teicoplanin in lung tissue were 0.1 mg/g following a single intravenous dose of 50 mg/kg, and 3.9 mg/g after a single instantaneous inhalation of the same dose [359]. Peak serum concentrations were substantially higher following intravenous administration than inhalation (128.6 mg/L vs. 51.7 mg/L, respectively). In a study with 13 adult patients receiving 12 mg/kg/24h teicoplanin intravenously for the treatment of nosocomial pneumonia, the steady-state median total and unbound Cmin concentrations of teicoplanin in serum were 15.9 mg/L (range 8.8–29.9 mg/L) and 3.7 mg/L (2.0–5.4 mg/L), respectively, with a median ELF concentration of 4.9 mg/L (2.0–11.8 mg/L) [360]. An inhaled formulation of teicoplanin is currently undergoing phase I trials (NCT04176328) [361].
Vancomycin has been widely used in the treatment of intracranial infection [362] with demonstrated advantages in safety, efficacy and PK/PD compared to intravenous administration when administered intrathecally via lumbar puncture [363]. For example, compared to intravenous administration, intrathecal vancomycin significantly shortened the treatment time and reduced the treatment cost, and had a better safety profile, in patients with severe traumatic brain injury [364].
3.3. Dosing strategies to improve clinical efficacy and minimise toxicity
A study involving 108 patients with lower respiratory tract infections caused by S. aureus first suggested that a vancomycin AUC0–24/MIC of ≥400 was associated with improved clinical efficacy and bacterial clearance [365]. Subsequent clinical studies support this PK/PD target [366–371]. However, when the AUC0–24/MIC ≥400 target was first adopted [372], the AUC0–24/MIC was not routinely available in clinical practice and consequently trough concentration (Cmin) has been used as a surrogate of the AUC0–24/MIC to predict vancomycin efficacy. A Cmin of 15–20 mg/L indicates effective achievement of the AUC0–24/MIC target ≥400 [365, 372], whereas AUC0–24/MICs <400 (often considered as a Cmin <10 mg/L) is often associated with reduced bacterial killing and the emergence of vancomycin resistance [373, 374]. A Cmin between 15–20 mg/L is deemed suitable in patients with normal renal function if the MIC is ≤1 mg/L [372]. However, reliance on vancomycin Cmin as a surrogate for AUC0–24/MIC has recently been questioned and there has been a shift back to recommending AUC-based dosing approaches [375]. Adoption of serum Cmin was purely for convenience given the difficulty of estimating AUC0–24. However, Cmin does not correlate well with AUC0–24 values as it provides only a snapshot at the end of the dosing interval, and trough monitoring is associated with higher nephrotoxicity [375, 376]. Therefore, Cmin is not an ideal PK/PD surrogate for vancomycin. The recommended dosage regimen of vancomycin was recently revisited by several infectious diseases societies and an AUC-based dosing approach was proposed [375]. This new recommendation was based on numerous studies over the last decade which concluded that the dosing of vancomycin based on the AUC0–24 targets is more accurate and maximises bacterial killing while minimizing the risk of nephrotoxicity [377, 378]. It was determined that for serious infections caused by MRSA, an AUC0–24/MIC of 400–600 (with the MIC determined by broth microdilution) in both adult and paediatric patients should be targeted [375]. Post-dose concentrations (e.g., a peak and a trough level) integrated with Bayesian dosing was the preferred method of monitoring the AUC. Studies are currently underway to examine potential urinary biomarkers (e.g., KIM-1, clusterin, and osteopontin) for determining the onset and extent of kidney injury [379]. As such biomarkers are potentially superior to serum creatinine for AKI [379–381], these studies will substantially assist in optimising the use of vancomycin in patients. Prolonged or continuous infusion of vancomycin has been proposed to optimise its PK/PD for critically ill patients [382, 383]. However, in pediatric patients the evidence of improved clinical and microbiologic outcomes for continuous infusion is limited [384].
As for vancomycin, AUC/MIC is the most predictive PK/PD index of telavancin, dalbavancin and oritavancin (Table 4). Compared with vancomycin, these three new glycopeptide antibiotics show improved PK/PD characteristics, a lower incidence of adverse reactions, and lower rates of resistance.
Table 4.
PK/PD targets for telavancin, dalbavancin and oritavancin.
| Telavancin | Dalbavancin | Oritavancin | |
|---|---|---|---|
| Predictive PK/PD index | AUC/MIC [385] | AUC/MIC [283] | Vary with models and isolates |
| In vitro PK/PD models | AUC/MIC: 50–404 for MSSA and MRSA [386]; 43.1 for VSSA [387]; 15.1 for Enterococci [387] |
||
| Animal models | fAUC/MIC for S. aureus: 83.0 in thigh infection model and 40.4 in lung infection model [388] | fAUC/MIC for S. aureus: 25, 50, 100 for stasis, 1-log and 2-log reductions in thigh infection model, respectively [389] | Cmax/MIC or AUC/MIC for S. aureus: (thigh infection model) [390, 391] |
| Clinical studies | %T>MIC for S. aureus: ≥ 22% for 93% microbiological success; < 22% for 76% microbiological success [392] |
Adverse effects of vancomycin and teicoplanin include hypersensitivity reactions (most commonly macular cutaneous rashes and anaphylaxis), Stevens-Johnson syndrome, and toxic epidermal necrolysis [315, 393, 394]. The red man syndrome can occur with vancomycin and is characterized by intense redness over the upper body due to histamine release and painful trunk muscle spasms [395, 396]; however, the red man syndrome is very rarely seen with teicoplanin which can usually be safely administered to patients with a history of vancomycin-induced red man syndrome [397, 398]. To minimise these adverse events, vancomycin should be infused slowly (maximum rate, 10 mg/min) over at least 1 h [393]. Thrombocytopenia is more common with teicoplanin than vancomycin, especially when administered at higher than recommended doses [398, 399]. Vestibular and cochlear damage associated with tinnitus and sensorineural hearing loss have also been reported following vancomycin and teicoplanin administration in patients with Down’s Syndrome and those with serious Gram-positive infections [400, 401]. However, given many of these patients had pre-existing hearing loss or received other ototoxic drugs (e.g., aminoglycosides or loop diuretics) concomitantly with vancomycin, proving direct causation by vancomycin is difficult [402]. Therefore, it is generally recommended that routine audiogram monitoring for ototoxicity is not required for adult patients and that use of vancomycin should be avoided in patients with previously diagnosed hearing loss [372, 403]. Recently, a retrospective study examined the prevalence of negative changes in audiograms in 92 patients receiving intravenous vancomycin for ≥14 days [404]. The prevalence of negative changes in audiograms was low among patients receiving long-term intravenous vancomycin, with 7 patients (8%) experiencing a decline in hearing from the baseline audiogram, two patients (2%) mild sensory neural loss, two patients (2%) mild to moderate loss, and three patients (3%) moderate to severe loss [404]. Therefore, the authors concluded that routine audiogram testing remains questionable except in high-risk patients and larger prospective clinical studies are required. Among 3,377 patients treated with teicoplanin, only 11 cases of ototoxicity were reported [212, 398]. To date, a causal relation between alterations in auditory function and teicoplanin therapy has not been established in controlled clinical studies.
The adverse effect of most concern with intravenous vancomycin is acute kidney injury (AKI) [405]. Although nephrotoxicity has been reported with teicoplanin, the incidence of AKI is significantly less frequent than with vancomycin [398, 406]. Most renal damage caused by vancomycin occurs in the proximal tubule [407–409], where vancomycin causes oxidative stress in rats [410–412] and alters mitochondrial function in porcine tubular LLCPK1 cells [412]. Determining the prevalence of AKI is challenging as multiple definitions are used in the literature [372]. In a systematic review and meta-analysis that examined vancomycin-induced nephrotoxicity associated with dosing regimens that maintained Cmin between 15 and 20 mg/L, the prevalence of nephrotoxicity ranged from 5–43% when assessed by a variety of definitions (including the Acute Kidney Injury Network classification (AKIN), Risk-Injury-Failure-Loss End-stage renal disease criteria (RIFLE), and an increase in serum creatinine of 0.5 mg/dL) [413]. A recent systematic review and meta-analysis examining only randomized controlled trials and cohort studies compared the nephrotoxicity caused by vancomycin to another non-glycopeptide antibiotic in a total of 4,033 patients [414]. The relative risk of AKI with vancomycin was 2.45 (95% CI, 1.69–3.55), with an attributable risk of 59% [414]. Significantly deteriorated renal function with vancomycin was only observed in 2 of the 6 included cohort studies [414]. Other studies comparing the nephrotoxicity caused by vancomycin and other drugs (e.g., penicillin and gentamicin) also showed similar rates of AKI, although only the study by Carreno et al. specified the definition used (the AKIN criteria) [415, 416]. Current clinical data suggest that the risk of vancomycin-induced AKI increases with increasing Cmin, particularly if it is maintained above 15–20 mg/L [413] and when the AUC0–24 exceeds 650–1,300 mg·h/L [417–419]. Vancomycin-associated nephrotoxicity can progress to acute renal failure which requires dialysis [420].
Considering the narrow therapeutic window, TDM is recommended in patients receiving intravenous vancomycin [375]. However, the appropriate target and sampling times for TDM remain uncertain [421]. The American Society of Health-System Pharmacists, IDSA, Pediatric Infectious Diseases Society, and the Society of Infectious Diseases Pharmacists recently revised their guidelines for vancomycin monitoring [375], The revised guidelines recommend that an AUC0–24/MIC of 400–600 in adult and paediatric patients be targeted to optimise vancomycin efficacy and minimise the risk of AKI, rather than targeting a Cmin of 15–20 mg/L as was recommended previously [375]. In practice, the AUC0–24 can now be accurately estimated using limited PK sampling (1 level [Cmin only or random level] or 2 levels [Cmin and Cmax]) and Bayesian software, which allows for dose adjustment during the first 24–48 h of therapy [375]. Nevertheless, simpler and more accurate methods of TDM are still needed to guide vancomycin therapy [422]. The current Chinese guidelines for adults with MRSA infections still recommend monitoring the Cmin with a target range of between 10–20 mg/L (a target acknowledged as a weak recommendation based on low-quality evidence), or the AUC0–24 with a target range of between 400–650 mg·h/L (strong recommendation, moderate-quality evidence) [423]. For paediatric patients, the Chinese guidelines recommend a Cmin of 5–15 mg/L (strong recommendation, low quality evidence); no AUC0–24 target is provided [423].
While the new guidelines emphasise the importance of early treatment and early TDM [372, 375], early AUC monitoring for clinical success and toxicity is challenging [424]. TDM is arguably most needed in acutely ill and critically ill patients; however, such patients often have poor renal function and altered volume of distribution which render early PK determinations (e.g., in the first 24–48 h) of little value to subsequent pharmacokinetic assessments once these parameters normalize [424, 425]. As the increased risk of vancomycin-induced AKI occurs predominately after at least 5 days of therapy and with large doses [426], most patients receiving empiric therapy have already discontinued vancomycin by this time.
Obese patients may be at higher risk of vancomycin-induced nephrotoxicity [375]. Obese patients often have an increased vancomycin volume of distribution which is not always linear with body weight [427], and high-quality PK/PD/TD data to guide vancomycin dosing in obese patients is lacking [428]. Nevertheless, early and frequent monitoring of AUC is recommended for dose adjustment in obese patients. Immunosuppressed patients with hematologic malignancies often have altered vancomycin pharmacokinetics including a lower AUC0–24 which may cause clinical failure [429–431]. The significance of creatinine clearance in neutropenia patients is controversial and provides another challenge [432, 433], as TDM-informed dosage adjustments may help decrease the incidence of nephrotoxicity but not clinical response [434]. The current literature on TDM of vancomycin in neonates and pediatric patients is very limited and many patients failed to achieve the Cmin of 15–20 mg/L [435, 436], especially in those with higher base-line creatinine clearance [436]. Determining the correlations among clinical outcome, Cmin and AUC0–24 is a significant challenge when using vancomycin and conducting TDM in children [437, 438].
Two studies have reported improved clinical outcomes of intravenous vancomycin with a loading dose, although the total number of patients included was small (76 and 124 patients) [439, 440]. Although a loading dose is not in the product information approved by the FDA [375, 441], clinical practice guidelines in the United States [372] and Japan [442] recommend a loading dose based on actual body weight (25–30 mg/kg) for the rapid attainment of therapeutic concentrations in seriously ill patients with suspected MRSA infection. A recent systematic review and meta-analysis concluded that not only do loading doses significantly increase the attainment of therapeutic concentrations, but they also reduce the incidence of nephrotoxicity without increasing other adverse effects [443].
For teicoplanin, there is very limited toxicokinetic data and hence TDM is not routinely undertaken. Both fCmax/MIC and fAUC0–24/MIC are predictive for teicoplanin efficacy against MSSA and MRSA in the murine thigh infection model [444]. Current clinical data indicate that clinical efficacy of teicoplanin is associated with the AUC0–24/MIC [310, 445–448] and Cmin [449]. AUC0–24/MIC targets of ≥125 and ≥345 have been proposed for moderate and severe infections caused by various staphylococci species including S. aureus [448, 450]. Two studies involving 24 and 33 patients reported that an AUC0–24 of ≥750–800 mg·h/L of teicoplanin on day 3 was associated with clinical success in patients infected with MRSA with an MIC of 1 mg/L [310, 447]. In a recent retrospective study with 46 patients infected by MRSA, an AUC0–24/MIC of ≥900 mg·h/L led to significantly better microbiological responses to intravenous teicoplanin than an AUC0–24/MIC of <900 mg·h/L [446]. Further clinical studies are required to confirm the AUC0–24/MIC targets for different types of infections caused by S. aureus and other bacterial species. Cmin is also closely associated with teicoplanin clinical efficacy and a target Cmin of at least 10 and 15 mg/L has been suggested for moderate infections (e.g., respiratory tract infections, urinary tract infections and skin and soft-tissue infections) and severe infections (e.g., sepsis, infective endocarditis, bone and joint infections), respectively [451]. Due to the low rate of Cmin attainment and potential toxicity (Cmin > 60mg/L), TDM in children has been recently recommended for teicoplanin [452–454]. While the optimal dosage regimen of teicoplanin remains unknown, its high plasma protein binding and long elimination t1/2 result in a slow onset of action (~72 h). Therefore, teicoplanin is not recommended as the first-line treatment of severe infections.
4. Daptomycin against Gram-positive bacteria
Daptomycin (Figure 5) is a cyclic lipopeptide antibiotic produced by Streptomyces rososporus [455]. It has a narrow antibacterial spectrum and is mainly used for the treatment of infections caused by Gram-positive bacteria, including MRSA, VRE and St. pneumoniae [456, 457]. It was first discovered in the 1980s but the development was halted after Phase I/II trials as high doses of daptomycin caused skeletal muscle toxicity [458]. Cubist Pharmaceuticals acquired daptomycin and developed it using a PK/PD approach. In 2003 the FDA approved daptomycin for the treatment of complicated skin and skin-structure infections (cSSSIs), albeit at a lower dose than used in the clinical trials [459].
Figure 5.

Structure of daptomycin.
Daptomycin consists of a 13-member amino acid cyclic lipopeptide (hydrophilic core) with a decanoyl side chain (lipophilic tail) [460–462]. It contains several non-proteinogenic amino acids including ornithine, L-3-methyl-glutamic acid, and L-kynurenine [460–462]. At neutral pH, daptomycin is negatively charged [463, 464]. Its unique chemical structure confers a mechanism of action that is different from that of the glycopeptides, acting mainly on the cell membrane of Gram-positive bacteria. Initial binding of daptomycin to the bacterial membrane is not well understood, although it shows increased binding affinity to negatively charged membranes containing phosphatidylglycerols, an important lipid component of bacterial membranes [465]. Insertion of daptomycin into bacterial membrane occurs in a Ca2+-dependent manner, with Ca2+-daptomycin forming a tripartite complex with undecaprenyl-coupled cell envelope precursors and the anionic phospholipid phosphatidylglycerol in the membrane [466–470]. Daptomycin does not lyse bacterial cells and generally does not cause serious inflammatory reactions due to the release of inflammatory mediators (e.g. lipoteichoic acid) from bacteria [471].
Resistance to daptomycin in patients was first reported only two years after it was approved in 2003 [472]. Mutations, thickening of the bacterial cell wall, changes in cell surface charge, and altered lipid metabolism are all associated with daptomycin resistance in Gram-positive pathogens [473] including En. faecalis [474], S. aureus [475], St. mitis/oralis [476] and Bacillus subtilis [465]. Mutations in mprF (multiple peptide resistance factor), rpoB and rpoC, and membrane spanning sensor/histidine kinase yycG are involved in daptomycin resistance in S. aureus [467].
Daptomycin displays rapid, concentration-dependent bactericidal activity against Gram-positive bacteria including VRE, MRSA and streptococci [477]. In the neutropenic mouse thigh model, daptomycin (0.20–400 mg/kg/day in divided doses) displayed concentration-dependent killing and produced in vivo PAEs of ~ 5–11 h against strains of S. aureus, St. pneumoniae and En. faecium [478]. Daptomycin is mainly excreted renally (78%), with ~50% of the unchanged drug recovered in urine within 24 h [477]. Total clearance of daptomycin in patients undergoing continuous renal replacement therapy varied and was 0.22, 0.24, 0.94 and 0.53 L/h in patients undergoing haemodialysis, continuous ambulatory peritoneal dialysis (CAPD), continuous veno-venous haemodialysis (CVVHD) and continuous veno-venous haemodiafiltration (CVVHDF), respectively. For patients with creatinine clearance ≥ 30 mL/min, the clearance was 0.75 L/h [479]. Protein binding of daptomycin in human plasma is ~92%, with most bound to albumin [480, 481]. Following intravenous administration, daptomycin cannot penetrate the blood-brain or placental barriers [459], and hence remains mostly in plasma and interstitial fluid (volume of distribution, ~0.1 L/kg) [477, 482]. The elimination t1/2 of daptomycin is ~8–9 h in patients [483, 484]. The PK of daptomycin is different between adults and children (especially neonates and infants), with the drug being cleared more rapidly in the latter resulting in lower AUC and Cmax levels [485]. Consequently, higher doses of daptomycin are required in younger children to achieve adequate bacterial killing. Higher than recommended doses may also be required in patients with special clinical conditions (e.g., sepsis, obesity, chronic renal disease, renal replacement therapy, and hypoproteinemia) where pathological changes lead to variable drug exposure [482].
Cmax/MIC and AUC/MIC are the most predictive PK/PD indices for daptomycin efficacy in mice (Figure 6) [478, 486]. In the neutropenic mouse thigh infection model, PK/PD targets for bacteriostasis against S. aureus, St. pneumoniae and En. faecium were 59–94, 12–36 and 0.14–0.25 for Cmax/MIC and 388–537, 75–237 and 0.94–1.67 for AUC/MIC, respectively [478]. More recently, clinical studies have shown improved efficacy with an AUC/MIC >666 [487–489] and a Cmax/MIC between 12–94 for bacteriostasis [488] and resistance suppression [490]; a Cmin <~3 mg/L was associated with poor outcomes [484]. However, PK alternations in critically ill patients and other patient groups mean that achieving adequate daptomycin exposure can be difficult [491]. In a clinical study involving 35 patients with severe Gram-positive infections, 11 patients receiving 8 mg/kg/day of daptomycin had significantly higher Cmax and AUC0–24 values (and consequently Cmax/MIC and AUC/MIC ratios) than in the 24 patients receiving a dose of 6 mg/kg/day (Cmax/MIC of 138 ± 35 and 87 ± 31, and AUC/MIC of 903 ± 280 and 692 ± 210 with the 8 and 6 mg/kg/day regimens, respectively) [490]. This suggests a dose of 8 mg/kg/day has a greater likelihood of achieving clinical success.
Figure 6.

Relationships between PK/PD parameters (a) %T>MIC, (b) Cmax/MIC, and (c) AUC0–24/MIC with bacterial load in the thighs of neutropenic mice after 24 h of treatment with multiple dosing regimens of daptomycin. The dotted lines represent the log10 CFU/thigh at the start of therapy [478]. Permission obtained from the American Society of Microbiology.
While the PK/PD targets outlined above have been proposed, there is high variability in plasma concentrations of daptomycin following intravenous administration which is only partially explained by the dose administered and underlying renal function. For example, Galar et al. reported a median Cmin and Cmax of 10.6 mg/L (range, 1.3–44.7 mg/L) and 44.0 mg/L (range, 3.0–93.7 mg/L) following a median daptomycin dose of 7 mg/kg (interquartile range, IQR, 5.0 – 9.0) [484]; whereas the equivalent values reported by Reiber et al. were 16.7 mg/L (range, 2–68 mg/L) and 66.2 mg/L (20–236 mg/L) following a median dose of 6 mg/kg/day (range, 2.7–13.8 mg/kg) [492]. Creatinine clearance, 48-h dose interval, albumin and ICU hospitalisation are all known to contribute to unpredictable plasma concentrations of daptomycin [492]. Given such high variability, TDM has been suggested to provide guidance on dose adjustment to maximise efficacy and reduce the likelihood of serious adverse events, especially in those patients who have substantially altered PK [483, 484, 490–492]. For example, a population PK/PD analysis showed that increasing the dose helped to increase the probability of target attainment (AUC/MIC of ≥666) in patients with a creatinine clearance ≥60 mL/min, in those with a creatinine clearance <60 mL/min but receiving CRRT, or with MICs ≥1 mg/L [493]. Compared with emipirical treatment, targeted treatment has significently improved the clinical outcome with P = 0.033 [484]. Overall, TDM is critical for determining AUC/MIC and adjusting doses accordingly to maximise efficacy [483, 484, 490, 492].
5. Conclusions
As novel antibiotics are unlikely to become available in the foreseeable future, antimicrobial peptide antibiotics are, and will continue to be, important therapeutic options for the treatment of infections caused by MDR bacterial infections. Optimising dosage regimens of peptide antibiotics using PK/PD/TD approaches is essential for maximising bacterial killing and minimising toxicity and the emergence of resistance. For the treatment of pulmonary infections, inhaled peptide antibiotics may provide superior drug exposure in the lungs and well-designed clinical studies are warranted to evaluate their safety and efficacy, including those that examine the simultaneous use of intravenous antibiotics. Due to the relatively narrow therapeutic window of peptide antibiotics, timely TDM and adaptive feedback control approaches can substantially benefit the therapy in patients.
Acknowledgements
Q.T.Z. and J.L. are supported by the National Institute of Allergy and Infectious Diseases of the National Institute of Health under Award Numbers R01AI146160 and R01AI132681. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institute of Allergy and Infectious Diseases or the National Institutes of Health. J.L. is an Australian National Health and Medical Research Council (NHMRC) Principal Research Fellow.
Abbreviations:
- L-Ara4N
4-amino-4-deoxy-L-arabinose
- AKI
acute kidney injury
- AFC
adaptive feedback control
- AMPs
antimicrobial peptides
- AUC
area under the drug concentration-time curve
- fAUC
area under the free drug concentration-time curve
- CSF
cerebrospinal fluid
- CBA
colistin base activity
- cSSSIs
complicated skin and skin-structure infections
- CI
confidence interval
- CAPD
continuous ambulatory peritoneal dialysis
- CVVHDF
continuous veno-venous haemodiafiltration
- CVVHD
continuous veno-venous haemodialysis
- Dab
diaminobutyric acid
- ELF
epithelial lining fluid
- EVD
external ventricular device
- XDR
extremely drug resistant
- hVISA
heterogeneous vancomycin-intermediate S. aureus
- IQR
interquartile range
- Ile
isoleucine
- LPS
lipopolysaccharide
- MSSA
methicillin-susceptible S. aureus
- MIU
million international unit
- MIC
minimum inhibitory concentration
- MDR
multidrug-resistant
- OR
odds ratio
- OM
outer membrane
- PBP
penicillin-binding protein
- PK/PD
pharmacokinetics/pharmacodynamics
- pEtN
phosphoethanolamine
- PAE
post-antibiotic effect
- CMS
sodium colistin methanesulphonate
- TDM
therapeutic drug monitoring
- UTIs
urinary tract infections
- VISA
vancomycin-intermediate S. aureus
- VRE
vancomycin-resistant enterococci
- VRSA
vancomycin-resistant S. aureus
- VSSA
vancomycin-sensitive S. aureus
- VSE
vancomycin-susceptible enterococci
- VAP
ventilator-associated pneumonia
References
- [1].World Health Organization, Antimicrobial resistance (2020), https://www.who.int/news-room/fact-sheets/detail/antimicrobial-resistance.
- [2].Magiorakos AP, Srinivasan A, Carey RB, Carmeli Y, Falagas ME, Giske CG, Harbarth S, Hindler JF, Kahlmeter G, Olsson-Liljequist B, Paterson DL, Rice LB, Stelling J, Struelens MJ, Vatopoulos A, Weber JT, Monnet DL, Multidrug-resistant, extensively drug-resistant and pandrug-resistant bacteria: An international expert proposal for interim standard definitions for acquired resistance, Clin Microbiol Infect, 18 (2012) 268–281. [DOI] [PubMed] [Google Scholar]
- [3].Prestinaci F, Pezzotti P, Pantosti A, Antimicrobial resistance: A global multifaceted phenomenon, Pathog Glob Health, 109 (2015) 309–318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [4].Vivas R, Barbosa AAT, Dolabela SS, Jain S, Multidrug-resistant bacteria and alternative methods to control them: An overview, Microb Drug Resist, 25 (2019) 890–908. [DOI] [PubMed] [Google Scholar]
- [5].Kollef MH, J C, Jy F, B F, Ms N, J R, A T, Jl V, Rg W, Kw G, Rehm C, Global prospective epidemiologic and surveillance study of ventilator-associated pneumonia due to Pseudomonas aeruginosa, Crit Care Med, 42 (2014) 2178–2187. [DOI] [PubMed] [Google Scholar]
- [6].Luong HX, Thanh TT, Tran TH, Antimicrobial peptides - Advances in development of therapeutic applications, Life Sci, 260 (2020) 118407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [7].Tagliabue A, Rappuoli R, Changing priorities in vaccinology: Antibiotic resistance moving to the top, Front Immunol, 9 (2018) 1068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [8].O’Neill J, Review on antimicrobial resistance. antimicrobial resistance: Tackling a crisis for the Health and Wealth of Nations, (2014).
- [9].Nesher L, Rolston KV, The current spectrum of infection in cancer patients with chemotherapy related neutropenia, Infection, 42 (2014) 5–13. [DOI] [PubMed] [Google Scholar]
- [10].Kawecki D, Pacholczyk M, Lagiewska B, Sawicka-Grzelak A, Durlik M, Mlynarczyk G, Chmura A, Bacterial and fungal infections in the early post-transplantation period after liver transplantation: Etiologic agents and their susceptibility, 64 (2014) 2777–2781. [DOI] [PubMed] [Google Scholar]
- [11].World Health Organization, WHO publishes list of bacteria for which new antibiotics are urgently needed, (2017).
- [12].World Health Organization, Global priority list of antibiotic-resistant bacteria to guide research, discovery, and development of new antibiotics, (2017).
- [13].Ventola CL, The antibiotic resistance crisis: Part 1: Causes and threats, P T, 40 (2015) 277–283. [PMC free article] [PubMed] [Google Scholar]
- [14].World Health Organization, antibacterial agents in clinical development: An analysis of the antibacterial clinical development pipeline, (2019).
- [15].Srinivas P, Hunt LN, Pouch SM, Thomas K, Goff DA, Pancholi P, Balada-Llasat JM, Bauer KA, Detection of colistin heteroresistance in Acinetobacter baumannii from blood and respiratory isolates, Diagn Microbiol Infect Dis, 91 (2018) 194–198. [DOI] [PubMed] [Google Scholar]
- [16].Hermes DM, Pormann Pitt C, Lutz L, Teixeira AB, Ribeiro VB, Netto B, Martins AF, Zavascki AP, Barth AL, Evaluation of heteroresistance to polymyxin B among carbapenem-susceptible and -resistant Pseudomonas aeruginosa, J Med Microbiol, 62 (2013) 1184–1189. [DOI] [PubMed] [Google Scholar]
- [17].Lewies A, Du Plessis LH, Wentzel JF, Antimicrobial peptides: The Achilles’ heel of antibiotic resistance, Probiotics Antimicrob Proteins, 11 (2019) 370–381. [DOI] [PubMed] [Google Scholar]
- [18].Radek K, Gallo R, Antimicrobial peptides: Natural effectors of the innate immune system, Semin Immunopathol, 29 (2007) 27–43. [DOI] [PubMed] [Google Scholar]
- [19].Chen CH, Lu TK, Development and challenges of antimicrobial peptides for therapeutic applications, Antibiotics (Basel), 9 (2020) 24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [20].Usmani SS, Bedi G, Samuel JS, Singh S, Kalra S, Kumar P, Ahuja AA, Sharma M, Gautam A, Raghava GPS, THPdb: Database of FDA-approved peptide and protein therapeutics, PLoS One, 12 (2017) e0181748. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [21].Guan Q, Huang S, Jin Y, Campagne R, Alezra VA-O, Wan YA-O, Recent advances in the exploration of therapeutic analogues of gramicidins, an old but still potent antimicrobial peptide, J. Med Chem, 62 (2019) 7603–7617. [DOI] [PubMed] [Google Scholar]
- [22].Parenta DM, Laplante KL, Anti-infective therapy, 145-glycopetides. In Infectious Diseases, 4th ed.; Cohen J, Opal SM, Powderly WG, Eds.; Elsevier Ltd.: Amsterdam, The Netherlands, (2017) 1249–1255. [Google Scholar]
- [23].Nang SC, Azad MAK, Velkov T, Zhou Q, Li J, Rescuing the last-line polymyxins: achievements and challenges, Pharmacol Rev, 73 (2021) 679–728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [24].Taylor SD, Palmer M, The action mechanism of daptomycin, Bioorg Med Chem, 24 (2016) 6253–6268. [DOI] [PubMed] [Google Scholar]
- [25].Steenbergen JN, Alder J, Thorne GM, Tally FP, Daptomycin: A lipopeptide antibiotic for the treatment of serious Gram-positive infections, J Antimicrob Chemother, 55 (2005) 283–288. [DOI] [PubMed] [Google Scholar]
- [26].Huan Y, Kong Q, Mou H, Yi H, Antimicrobial peptides: Classification, design, application and research progress in multiple fields, Front Microbiol, 11 (2020) 582779. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [27].Velkov T, Thompson PE, Azad MAK, Roberts KD, Bergen PJ, History, chemistry and antibacterial spectrum, In: Li J, Nation R, Kaye K (eds) Polymyxin antibiotics: From laboratory bench to bedside, Adv Exp Med Biol, 1145. Springer, Cham: (2019) 15–36. [DOI] [PubMed] [Google Scholar]
- [28].Clifford H, Stewart G, Intraventricular administration of a new derivative of polymyxin B in meningitis due to Ps. pyocyanea, Lancet, 2 (1961) 177–180. [DOI] [PubMed] [Google Scholar]
- [29].Ross S, Puig J, Zaremba E, Colistin: Some preliminary laboratory and clinical observations in specific gastroenteritis in infants and children, Antibiot Annu, 7 (1959) 89–100. [PubMed] [Google Scholar]
- [30].Storm DR, Rosenthal KS, Swanson PE, Polymyxin and related peptide antibiotics, Annu Rev Biochem, 46 (1977) 723–763. [DOI] [PubMed] [Google Scholar]
- [31].He J, Ledesma KR, Lam WY, Figueroa DA, Lim TP, Chow DS, Tam VH, Variability of polymyxin B major components in commercial formulations, Int J Antimicrob Agents, 35 (2010) 308–310. [DOI] [PubMed] [Google Scholar]
- [32].Gregoire N, Aranzana-Climent V, Magreault S, Marchand S, Couet W, Clinical Pharmacokinetics and pharmacodynamics of colistin, Clin Pharmacokinet, 56 (2017) 1441–1460. [DOI] [PubMed] [Google Scholar]
- [33].Zavascki AP, Goldani LZ, Li J, Nation RL, Polymyxin B for the treatment of multidrug-resistant pathogens: A critical review, J Antimicrob Chemother, 60 (2007) 1206–1215. [DOI] [PubMed] [Google Scholar]
- [34].Bergen PJ, Li J, Rayner CR, Nation RL, Colistin methanesulfonate is an inactive prodrug of colistin against Pseudomonas aeruginosa, Antimicrob Agents Chemother, 50 (2006) 1953–1958. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [35].Shanghai New Asia Pharmaceutical Co., Ltd. Colistin sulfate for injection [package insert]. China: (2019). [Google Scholar]
- [36].Bergen PJ, Landersdorfer CB, Zhang J, Zhao M, Lee HJ, Nation RL, Li J, Pharmacokinetics and pharmacodynamics of ‘old’ polymyxins: What is new, Diagn Microbiol Infect Dis, 74 (2012) 213–223. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [37].Li Z, Velkov T, Polymyxins: Mode of Action. In: Li J, Nation R, Kaye K (eds) Polymyxin Antibiotics: From laboratory bench to bedside, Adv Exp Med Biol, 1145. Springer, Cham: (2019) 37–54. [DOI] [PubMed] [Google Scholar]
- [38].Rabanal F, Cajal Y, Recent advances and perspectives in the design and development of polymyxins, Nat Prod Rep, 34 (2017) 886–908. [DOI] [PubMed] [Google Scholar]
- [39].Velkov T, Deris ZZ, Huang JX, Azad MA, Butler M, Sivanesan S, Kaminskas LM, Dong YD, Boyd B, Baker MA, Cooper MA, Nation RL, Li J, Surface changes and polymyxin interactions with a resistant strain of Klebsiella pneumoniae, Innate Immun, 20 (2014) 350–363. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [40].Berglund NA, Piggot TJ, Jefferies D, Sessions RB, Bond PJ, Khalid S, Interaction of the antimicrobial peptide polymyxin B1 with both membranes of E. coli: A molecular dynamics study, PLoS Comput Biol, 11 (2015) e1004180. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [41].Adrià C, Francesc R, Maria G-S, Asunción AM, Yolanda C, Membrane association and contact formation by a synthetic analogue of polymyxin B and its fluorescent derivatives, J Phys Chem B, 110 (2006) 4465–4471. [DOI] [PubMed] [Google Scholar]
- [42].Cajal Y, Berg OG, Jain MK, Direct vesicle-vesicle exchange of phospholipids mediated by polymyxin B, Biochem Biophys Res Commun, 210 (1995) 746–752. [DOI] [PubMed] [Google Scholar]
- [43].Deris ZZ, Akter J, Sivanesan S, Roberts KD, Thompson PE, Nation RL, Li J, Velkov T, A secondary mode of action of polymyxins against Gram-negative bacteria involves the inhibition of NADH-quinone oxidoreductase activity, J Antibiot (Tokyo), 67 (2014) 147–151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [44].Sampson TR, Liu X, Schroeder MR, Kraft CS, Burd EM, Weiss DS, Rapid killing of Acinetobacter baumannii by polymyxins is mediated by a hydroxyl radical death pathway, Antimicrob Agents Chemother, 56 (2012) 5642–5649. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [45].Li Z, Cao Y, Yi L, Liu JH, Yang Q, Emergent Polymyxin Resistance: End of an Era, Open Forum Infect Dis, 6 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [46].Breazeale SD, Ribeiro AA, McClerren AL, Raetz CR, A formyltransferase required for polymyxin resistance in Escherichia coli and the modification of lipid A with 4-Amino-4-deoxy-L-arabinose: Identification and function of UDP-4-deoxy-4-formamido-L-arabinose, J Biol Chem, 280 (2005) 14154–14167. [DOI] [PubMed] [Google Scholar]
- [47].Moffatt J, Harper M, Boyce J, Mechanisms of polymyxin resistance, In: Li J, Nation R, Kaye K (eds) Polymyxin antibiotics: From laboratory bench to bedside, Adv Exp Med Biol, 1145. Springer, Cham: (2019) 55–71. [DOI] [PubMed] [Google Scholar]
- [48].Mark RP, Leila GC, Jace WJ, Mark DA, Daniel VZ, Karsten ROH, Yohei D, Robert KE, Unique structural modifications are present in the lipopolysaccharide from colistin-resistant strains of Acinetobacter baumannii, Antimicrob Agents Chemother, 57 (2013) 4831–4840. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [49].Schurek KN, Sampaio JL, Kiffer CR, Sinto S, Mendes CM, Hancock RE, Involvement of pmrAB and phoPQ in polymyxin B adaptation and inducible resistance in non-cystic fibrosis clinical isolates of Pseudomonas aeruginosa, Antimicrob Agents Chemother, 53 (2009) 4345–4351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [50].Gunn JS, Lim KB, Krueger J, Kim K, Guo L, Hackett M, Miller SI, PmrA-PmrB-regulated genes necessary for 4-aminoarabinose lipid A modification and polymyxin resistance, Mol Microbiol, 27 (1998) 1171–1182. [DOI] [PubMed] [Google Scholar]
- [51].Moffatt JH, Harper M, Harrison P, Hale JD, Vinogradov E, Seemann T, Henry R, Crane B, St Michael F, Cox AD, Adler B, Nation RL, Li J, Boyce JD, Colistin resistance in Acinetobacter baumannii is mediated by complete loss of lipopolysaccharide production, Antimicrob Agents Chemother, 54 (2010) 4971–4977. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [52].Pelletier MR, Casella LG, Jones JW, Adams MD, Zurawski DV, Hazlett KR, Doi Y, Ernst RK, Unique structural modifications are present in the lipopolysaccharide from colistin-resistant strains of Acinetobacter baumannii, Antimicrob Agents Chemother, 57 (2013) 4831–4840. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [53].Campos MA, Vargas MA, Regueiro V, Llompart CM, Albertí S, & Bengoechea JA, Capsule polysaccharide mediates bacterial resistance to antimicrobial peptides, Infection and immunity, 72 (2004) 7107–7114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [54].Srinivasan VB, Singh BB, Priyadarshi N, Chauhan NK, Rajamohan G, Role of novel multidrug efflux pump involved in drug resistance in Klebsiella pneumoniae, PLoS One, 9 (2014) e96288. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [55].Pamp SJ, Gjermansen M, Johansen HK, Tolker-Nielsen T, Tolerance to the antimicrobial peptide colistin in Pseudomonas aeruginosa biofilms is linked to metabolically active cells, and depends on the pmr and mexAB-oprM genes, Mol Microbiol, 68 (2008) 223–240. [DOI] [PubMed] [Google Scholar]
- [56].Mathur J, Davis BM, Waldor MK, Antimicrobial peptides activate the Vibrio cholerae sigmaE regulon through an OmpU-dependent signalling pathway, Mol Microbiol, 63 (2007) 848–858. [DOI] [PubMed] [Google Scholar]
- [57].Mathur J, Waldor MK, The Vibrio cholerae ToxR-regulated porin OmpU confers resistance to antimicrobial peptides, Infect Immun, 72 (2004) 3577–3583. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [58].Ito-Kagawa M, Koyama Y, Selective cleavage of a peptide antibiotic, colistin by colistinase, J Antibiot (Tokyo), 33 (1980) 1551–1555. [DOI] [PubMed] [Google Scholar]
- [59].Spohn R, Daruka L, Lazar V, Martins A, Vidovics F, Grezal G, Mehi O, Kintses B, Szamel M, Jangir PK, Csorgo B, Gyorkei A, Bodi Z, Farago A, Bodai L, Foldesi I, Kata D, Maroti G, Pap B, Wirth R, Papp B, Pal C, Integrated evolutionary analysis reveals antimicrobial peptides with limited resistance, Nat Commun, 10 (2019) 4538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [60].Jochumsen N, Marvig RL, Damkiaer S, Jensen RL, Paulander W, Molin S, Jelsbak L, Folkesson A, The evolution of antimicrobial peptide resistance in Pseudomonas aeruginosa is shaped by strong epistatic interactions, Nat Commun, 7 (2016) 13002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [61].Igler C, Rolff J, Regoes R, Multi-step vs. single-step resistance evolution under different drugs, pharmacokinetics, and treatment regimens, eLife, 10 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [62].Levison ME, Levison JH, Pharmacokinetics and pharmacodynamics of antibacterial agents, Infect Dis Clin North Am, 23 (2009) 791–815. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [63].Bulitta JB, Yang JC, Yohonn L, Ly NS, Brown SV, D’Hondt RE, Jusko WJ, Forrest A, Tsuji BT, Attenuation of colistin bactericidal activity by high inoculum of Pseudomonas aeruginosa characterized by a new mechanism-based population pharmacodynamic model, Antimicrob Agents Chemother, 54 (2010) 2051–2062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [64].Tam VH, Schilling AN, Vo G, Kabbara S, Kwa AL, Wiederhold NP, Lewis RE, Pharmacodynamics of polymyxin B against Pseudomonas aeruginosa, Antimicrob Agents Chemother, 49 (2005) 3624–3630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [65].Ly NS, Bulman ZP, Bulitta JB, Baron C, Rao GG, Holden PN, Li J, Sutton MD, Tsuji BT, Optimization of polymyxin B in combination with doripenem to combat mutator Pseudomonas aeruginosa, Antimicrob Agents Chemother, 60 (2016) 2870–2880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [66].Owen RJ, Li J, Nation RL, Spelman D, In vitro pharmacodynamics of colistin against Acinetobacter baumannii clinical isolates, J Antimicrob Chemother, 59 (2007) 473–477. [DOI] [PubMed] [Google Scholar]
- [67].Poudyal A, Howden BP, Bell JM, Gao W, Owen RJ, Turnidge JD, Nation RL, Li J, In vitro pharmacodynamics of colistin against multidrug-resistant Klebsiella pneumoniae, J Antimicrob Chemother, 62 (2008) 1311–1318. [DOI] [PubMed] [Google Scholar]
- [68].Deris ZZ, Yu HH, Davis K, Soon RL, Jacob J, Ku CK, Poudyal A, Bergen PJ, Tsuji BT, Bulitta JB, Forrest A, Paterson DL, Velkov T, Li J, Nation RL, The combination of colistin and doripenem is synergistic against Klebsiella pneumoniae at multiple inocula and suppresses colistin resistance in an in vitro pharmacokinetic/pharmacodynamic model, Antimicrob Agents Chemother, 56 (2012) 5103–5112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [69].Pournaras S, Vrioni G, Neou E, Dendrinos J, Dimitroulia E, Poulou A, Tsakris A, Activity of tigecycline alone and in combination with colistin and meropenem against Klebsiella pneumoniae carbapenemase (KPC)-producing Enterobacteriaceae strains by time-kill assay, Int J Antimicrob Agents, 37 (2011) 244–247. [DOI] [PubMed] [Google Scholar]
- [70].Li J, Rayner CR, Nation RL, Owen RJ, Spelman D, Tan KE, Liolios L, Heteroresistance to colistin in multidrug-resistant Acinetobacter baumannii, Antimicrob Agents Chemother, 50 (2006) 2946–2950. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [71].Landersdorfer CB, Wang J, Wirth V, Chen K, Kaye KS, Tsuji BT, Jian L, Nation RL, Pharmacokinetics/pharmacodynamics of systemically administered polymyxin B against Klebsiella pneumoniae in mouse thigh and lung infection models, J Antimicrob Chemother, 73 (2018) 462–468. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [72].Lin Y, Zhou Q, Onufrak NJ, Wirth V, Chen K, Wang J, Forrest A, Chan H-K, Li J, Aerosolized polymyxin B for treatment of respiratory tract infections: Determination of pharmacokinetic-pharmacodynamic indices for aerosolized polymyxin B against Pseudomonas aeruginosa in a mouse lung infection model, Antimicrob Agents Chemother, 61 (2017) e00211–00217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [73].Cheah SE, Wang J, Nguyen VT, Turnidge JD, Jian L, Nation RL, New pharmacokinetic/pharmacodynamic studies of systemically administered colistin against Pseudomonas aeruginosa and Acinetobacter baumannii in mouse thigh and lung infection models: Smaller response in lung infection, J Antimicrob Chemother, 70 (2015) 3291–3297. [DOI] [PubMed] [Google Scholar]
- [74].Li J, Turnidge J, Milne R, Nation RL, Coulthard K, In vitro pharmacodynamic properties of colistin and colistin methanesulfonate against Pseudomonas aeruginosa isolates from patients with cystic fibrosis, Antimicrob Agents Chemother, 45 (2001) 781–785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [75].Plachouras D, Giamarellos-Bourboulis N, Kentepozidis F, Baziaka V, Karagianni H, Giamarellou H, In vitro postantibiotic effect of colistin on multidrug-resistant Acinetobacter baumannii, Diagn Microbiol Infect Dis, 57 (2007) 419–422. [DOI] [PubMed] [Google Scholar]
- [76].Ozbek B, Sentürk A, Postantibiotic effects of tigecycline, colistin sulfate, and levofloxacin alone or tigecycline-colistin sulfate and tigecycline-levofloxacin combinations against Acinetobacter baumannii, Chemotherapy, 56 (2010) 466–471. [DOI] [PubMed] [Google Scholar]
- [77].Nation RL, Velkov T, Li J, Colistin and polymyxin B: Peas in a pod, or chalk and cheese?, Clin Infect Dis, 59 (2014) 88–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [78].Zhou LL, Li CT, Weng QY, Wu JT, Luo HT, Xue ZQ, Xiao Y, Chen CR, Clinical study on intravenous combined with aerosol inhalation of polymyxin B for the treatment of pneumonia caused by multidrug-resistant Gram-negative bacteria, Zhonghua Wei Zhong Bing Ji Jiu Yi Xue, 33 (2021) 416–420. [DOI] [PubMed] [Google Scholar]
- [79].Li J, Milne RW, Nation RL, Turnidge JD, Coulthard K, Johnson DW, A simple method for the assay of colistin in human plasma, using pre-column derivatization with 9-fluorenylmethyl chloroformate in solid-phase extraction cartridges and reversed-phase high-performance liquid chromatography, J Chromatogr B Biomed Sci Appl, 761 (2001) 167–175. [DOI] [PubMed] [Google Scholar]
- [80].Gobin P, Lemaitre F, Marchand S, Couet W, Olivier JC, Assay of colistin and colistin methanesulfonate in plasma and urine by liquid chromatography-tandem mass spectrometry, Antimicrob Agents Chemother, 54 (2010) 1941–1948. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [81].He J, Gao S, Hu M, Chow DS, Tam VH, A validated ultra-performance liquid chromatography-tandem mass spectrometry method for the quantification of polymyxin B in mouse serum and epithelial lining fluid: Application to pharmacokinetic studies, J Antimicrob Chemother, 68 (2013) 1104–1110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [82].Milne RW Bioanalysis and Stability of Polymyxins. In: Li J, Nation R, Kaye K (eds) Polymyxin antibiotics: From laboratory bench to bedside, Adv Exp Med Biol, 1145. Springer, Cham: (2019) 73–87. [DOI] [PubMed] [Google Scholar]
- [83].Li J, Milne RW, Nation RL, Turnidge JD, Smeaton TC, Coulthard K, Pharmacokinetics of colistin methanesulphonate and colistin in rats following an intravenous dose of colistin methanesulphonate, J Antimicrob Chemother, 53 (2004) 837–840. [DOI] [PubMed] [Google Scholar]
- [84].Marchand S, Lamarche I, Gobin P, Couet W, Dose-ranging pharmacokinetics of colistin methanesulphonate (CMS) and colistin in rats following single intravenous CMS doses, J Antimicrob Chemother, 65 (2010) 1753–1758. [DOI] [PubMed] [Google Scholar]
- [85].Li J, Milne RW, Nation RL, Turnidge JD, Smeaton TC, Coulthard K, Use of high-performance liquid chromatography to study the pharmacokinetics of colistin sulfate in rats following intravenous administration, Antimicrob Agents Chemother, 47 (2003) 1766–1770. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [86].Ma Z, Wang J, Nation RL, Li J, Turnidge JD, Coulthard K, Milne RW, Renal disposition of colistin in the isolated perfused rat kidney, Antimicrob Agents Chemother, 53 (2009) 2857–2864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [87].Marchand Sandrine, Grégoire Nicolas, Couet William, Pharmacokinetics of polymyxins in animals. In: Li J, Nation R, Kaye K (eds) Polymyxin antibiotics: From laboratory bench to bedside, Adv Exp Med Biol, 1145. Springer, Cham: (2019) 89–103. [DOI] [PubMed] [Google Scholar]
- [88].Couet W, Gregoire N, Gobin P, Saulnier PJ, Frasca D, Marchand S, Mimoz O, Pharmacokinetics of colistin and colistimethate sodium after a single 80-mg intravenous dose of CMS in young healthy volunteers, Clin Pharmacol Ther, 89 (2011) 875–879. [DOI] [PubMed] [Google Scholar]
- [89].Abdelraouf K, He J, Ledesma KR, Hu M, Tam VH, Pharmacokinetics and renal disposition of polymyxin B in an animal model, Antimicrob Agents Chemother, 56 (2012) 5724–5727. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [90].al-Khayyat AL, Aronson AL, Pharmacologic and toxicologic studies with the polymyxins. II. Comparative pharmnacologic studies of the sulfate and methanesulfonate salts of polymyxin B and colistin in dogs, Chemotherapy, 19 (1973) 82–97. [DOI] [PubMed] [Google Scholar]
- [91].Ziv G, Wanner M, Nicolet J, Clinical pharmacology of polymyxin B, colistin, and colistimethate in young dairy calves, J Vet Pharmacol Ther, 3 (1980) 87–94. [Google Scholar]
- [92].Hengzhuang W, Wu H, Ciofu O, Song Z, Hoiby N, In vivo pharmacokinetics/pharmacodynamics of colistin and imipenem in Pseudomonas aeruginosa biofilm infection, Antimicrob Agents Chemother, 56 (2012) 2683–2690. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [93].Li J, Coulthard K, Nation RL Labelling conventions and product package insert of parenteral polymyxins: Factors causing potential medication errors and impeding optimal clinical use. In: Li J, Nation R, Kaye K (eds) Polymyxin antibiotics: From laboratory bench to bedside, Adv Exp Med Biol, 1145. Springer, Cham: (2019) 133–141. [DOI] [PubMed] [Google Scholar]
- [94].Mizuyachi K, Hara K, Wakamatsu A, Nohda S, Hirama T, Safety and pharmacokinetic evaluation of intravenous colistin methanesulfonate sodium in Japanese healthy male subjects, Curr Med Res Opin, 27 (2011) 2261–2270. [DOI] [PubMed] [Google Scholar]
- [95].Fan YX, Chen YC, Li Y, Yu JC, Bian XC, Li X, Li WZ, Guo BN, Wu HL, Liu XF, Wang Y, Xu XY, Hu JL, Wang JJ, Wu XJ, Cao GY, Wu JF, Xue CJ, Feng J, Zhang YY, Zhang J, Effects of different component contents of colistin methanesulfonate on the pharmacokinetics of prodrug and formed colistin in human, Pharm Res, 38 (2021) 79–87. [DOI] [PubMed] [Google Scholar]
- [96].Yapa SWS, Li J, Porter CJ, Nation RL, Patel K, McIntosh MP, Population pharmacokinetics of colistin methanesulfonate in rats: Achieving sustained lung concentrations of colistin for targeting respiratory infections, Antimicrob Agents Chemother, 57 (2013) 5087–5095. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [97].Fiaccadori E, Antonucci E, Morabito S, d’Avolio A, Maggiore U, Regolisti G, Colistin use in patients with reduced kidney function, Am J Kidney Dis, 68 (2016) 296–306. [DOI] [PubMed] [Google Scholar]
- [98].Markou N, Markantonis SL, Dimitrakis E, Panidis D, Boutzouka E, Karatzas S, Rafailidis P, Apostolakos H, Baltopoulos G, Colistin serum concentrations after intravenous administration in critically ill patients with serious multidrug-resistant, Gram-negative bacilli infections: A prospective, open-label, uncontrolled study, Clin Ther, 30 (2008) 143–151. [DOI] [PubMed] [Google Scholar]
- [99].Nation RL, Garonzik SM, Thamlikitkul V, Giamarellos-Bourboulis EJ, Forrest A, Paterson DL, Jian L, Silveira FP, Dosing guidance for intravenous colistin in critically-ill patients, Clin Infect Dis, 64 (2017) 565–571. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [100].Garonzik SM, Li J, Thamlikitkul V, Paterson DL, Shoham S, Jacob J, Silveira FP, Forrest A, Nation RL, Population pharmacokinetics of colistin methanesulfonate and formed colistin in critically ill patients from a multicenter study provide dosing suggestions for various categories of patients, Antimicrob Agents Chemother, 55 (2011) 3284–3294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [101].Mohamed AF, Karaiskos I, Plachouras D, Karvanen M, Pontikis K, Jansson B, Papadomichelakis E, Antoniadou A, Giamarellou H, Armaganidis A, Cars O, Friberg LE, Application of a loading dose of colistin methanesulfonate in critically ill patients: Population pharmacokinetics, protein binding, and prediction of bacterial kill, Antimicrob Agents Chemother, 56 (2012) 4241–4249. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [102].Nation RL, Garonzik SM, Li J, Thamlikitkul V, Giamarellos-Bourboulis EJ, Paterson DL, Turnidge JD, Forrest A, Silveira FP, Updated US and European dose recommendations for intravenous colistin: How do they perform, Clin Infect Dis, 62 (2016) 552–558. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [103].Koomanachai P, Landersdorfer CB, Chen G, Lee HJ, Jitmuang A, Wasuwattakul S, Sritippayawan S, Li J, Nation RL, Thamlikitkul V, Pharmacokinetics of colistin methanesulfonate and formed colistin in end-stage renal disease patients receiving continuous ambulatory peritoneal dialysis, Antimicrob Agents Chemother, 58 (2014) 440–446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [104].Wertheim H, Van Nguyen K, Hara GL, Gelband H, Laxminarayan R, Mouton J, Cars O, Global survey of polymyxin use: A call for international guidelines, J Glob Antimicrob Resist, 1 (2013) 131–134. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [105].Shanghai First Biochemical Pharmaceutical Co., Ltd. Polymyxin B for injection [package Insert]. China: (2017). [Google Scholar]
- [106].X-Gen Pharmaceuticals I. Polymyxin B for Injection [package Insert]. USP. 2017. [Google Scholar]
- [107].Zavascki AP, Goldani LZ, Cao G, Superti SV, Lutz L, Barth AL, Ramos F, Boniatti MM, Nation RL, Li J, Pharmacokinetics of intravenous polymyxin B in critically ill patients, Clin Infect Dis, 47 (2008) 1298–1304. [DOI] [PubMed] [Google Scholar]
- [108].Liu X, Chen Y, Yang H, Li J, Yu J, Yu Z, Cao G, Wu X, Wang Y, Wu H, Fan Y, Wang J, Wu J, Jin Y, Guo B, Hu J, Bian X, Li X, Zhang J, Acute toxicity is a dose-limiting factor for intravenous polymyxin B: A safety and pharmacokinetic study in healthy Chinese subjects, J Infect, 82 (2021) 207–215. [DOI] [PubMed] [Google Scholar]
- [109].Sandri AM, Landersdorfer CB, Jacob J, Boniatti MM, Dalarosa MG, Falci DR, Behle TF, Bordinhao RC, Wang J, Forrest A, Nation RL, Jian L, Zavascki AP, Population pharmacokinetics of intravenous polymyxin B in critically ill patients: Implications for selection of dosage regimens, Clin Infect Dis, 57 (2013) 524–531. [DOI] [PubMed] [Google Scholar]
- [110].Yu XB, Jiao Z, Zhang CH, Dai Y, Zhou ZY, Han L, Wen X, Sheng CC, Lin GY, Pan JY, Population pharmacokinetic and optimization of polymyxin B dosing in adult patients with various renal functions, Br J Clin Pharmacol, 87 (2021) 1869–1877. [DOI] [PubMed] [Google Scholar]
- [111].Wang P, Zhang Q, Zhu Z, Pei H, Feng M, Sun T, Yang J, Zhang X, Comparing population pharmacokinetics and acute kidney injury of polymyxin B in Chinese patients with and without renal insufficiency, Antimicrob Agents Chemother, 65 (2021) e01900–01920. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [112].Avedissian SN, Miglis C, Kubin CJ, Rhodes NJ, Yin MT, Cremers S, Prickett M, Scheetz MH, Polymyxin B pharmacokinetics in adult cystic fibrosis patients, Pharmacotherapy, 38 (2018) 730–738. [DOI] [PubMed] [Google Scholar]
- [113].Manchandani P, Thamlikitkul V, Dubrovskaya Y, Babic JT, Lye DC, Lee LS, Tam VH, Population Pharmacokinetics of Polymyxin B, Clin Pharmacol Ther, 104 (2018) 534–538. [DOI] [PubMed] [Google Scholar]
- [114].Tsuji BT, Pogue JM, Zavascki AP, Paul M, Daikos GL, Forrest A, Giacobbe DR, Viscoli C, Giamarellou H, Karaiskos I, Kaye D, Mouton JW, Tam VH, Thamlikitkul V, Wunderink RG, Jian L, Nation RL, Kaye KS, International consensus guidelines for the optimal use of the polymyxins: Endorsed by the American College of Clinical Pharmacy (ACCP), European Society of Clinical Microbiology and Infectious Diseases (ESCMID), Infectious Diseases Society of America (IDSA), International Society for Anti-infective Pharmacology (ISAP), Society of Critical Care Medicine (SCCM), and Society of Infectious Diseases Pharmacists (SIDP), Pharmacotherapy, 39 (2019) 10–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [115].Avedissian SN, Scheetz MH, Does renal function matter for polymyxin B?, Br J Clin Pharmacol, 87 (2021) 2629–2632. [DOI] [PubMed] [Google Scholar]
- [116].The European Committee on Antimicrobial Susceptibility Testing. Breakpoint tables for interpretation of MICs and zone diameters. Version 12.0, 2021. http://www.eucast.org.
- [117].The United States Committee on Antimicrobial Susceptibility Testing (USCAST). Breakpoint tables for interpretation of MIC and zone diameter results. Version 7.0, 2021. http://www.uscast.org.
- [118].Bian X, Liu X, Hu F, Feng M, Chen Y, Bergen PJ, Li J, Li X, Guo Y, Zhang J, Pharmacokinetic/pharmacodynamic based breakpoints of polymyxin B for bloodstream infections caused by multidrug-resistant gram-negative pathogens, Front Pharmacol, 12 (2022) 785893. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [119].Cai Y, Leck H, Tan RW, Teo JQ, Lim TP, Lee W, Chlebicki MP, Kwa AL, Clinical experience with high-dose polymyxin B against carbapenem-resistant Gram-negative bacterial infections-A cohort study, Antibiotics (Basel), 9 (2020) 451. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [120].Nelson BC, Eiras DP, Gomez-Simmonds A, Loo AS, Satlin MJ, Jenkins SG, Whittier S, Calfee DP, Furuya EY, Kubin CJ, Clinical outcomes associated with polymyxin B dose in patients with bloodstream infections due to carbapenem-resistant Gram-negative rods, Antimicrob Agents Chemother, 59 (2015) 7000–7006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [121].Sorli L, Luque S, Li J, Campillo N, Danes M, Montero M, Segura C, Grau S, Horcajada JP, Colistin for the treatment of urinary tract infections caused by extremely drug-resistant Pseudomonas aeruginosa: Dose is critical, J Infect, 79 (2019) 253–261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [122].A Pastewski A, Caruso P, R Parris A, Dizon R, Kopec R, Sharma S, Mayer S, Ghitan M, Chapnick EK, Infectious diseases: Parenteral polymyxin B use in patients with multidrug-resistant gram-negative bacteremia and urinary tract infections: A retrospective case series, Ann Pharmacother, 42 (2008) 1177–1187. [DOI] [PubMed] [Google Scholar]
- [123].Dubrovskaya Y, Chen TY, Scipione MR, Esaian D, Phillips MS, Papadopoulos J, Mehta SA, Risk factors for treatment failure of polymyxin B monotherapy for carbapenem-resistant Klebsiella pneumoniae infections, Antimicrob Agents Chemother, 57 (2013) 5394–5397. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [124].Mostardeiro MM, Pereira CA, Marra AR, Pestana JO, Camargo LF, Nephrotoxicity and efficacy assessment of polymyxin use in 92 transplant patients, Antimicrob Agents Chemother, 57 (2013) 1442–1446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [125].Chuang YC, Cheng CY, Sheng WH, Sun HY, Wang JT, Chen YC, Chang SC, Effectiveness of tigecycline-based versus colistin- based therapy for treatment of pneumonia caused by multidrug-resistant Acinetobacter baumannii in a critical setting: A matched cohort analysis, BMC Infect Dis, 14 (2014) 102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [126].Zalts Ronen, Neuberger Ami, Hussein Khetam, Raz-Pasteur Ayelet, Geffen Yuval, Mashiach Tanya, Finkelstein R, Treatment of carbapenem-resistant Acinetobacter baumannii ventilator-associated pneumonia: retrospective comparison between intravenous colistin and intravenous ampicillin-sulbactam, Am J Ther, 23 (2016) e78–85. [DOI] [PubMed] [Google Scholar]
- [127].Rigatto MH, Ribeiro VB, Konzen D, Zavascki AP, Comparison of polymyxin B with other antimicrobials in the treatment of ventilator-associated pneumonia and tracheobronchitis caused by Pseudomonas aeruginosa or Acinetobacter baumannii, Infection, 41 (2013) 321–328. [DOI] [PubMed] [Google Scholar]
- [128].Prevalence and clinical outcomes of episodes of ventilator-associated pneumonia caused by SPM-1-producing and non-producing imipenem-resistant Pseudomonas aeruginosa, Rev Soc Bras Med Tro, 44 (2011) 604–606. [DOI] [PubMed] [Google Scholar]
- [129].Aydemir H, Akduman D, Piskin N, Comert F, Horuz E, Terzi A, Kokturk F, Ornek T, Celebi G, Colistin vs. the combination of colistin and rifampicin for the treatment of carbapenem-resistant Acinetobacter baumannii ventilator-associated pneumonia, Epidemiol Infect, 141 (2013) 1214–1222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [130].Garnacho-Montero J, Ortiz-Leyba C, Jiménez-Jiménez FJ, Barrero-Almodóvar AE, García-Garmendia JL, Bernabeu-Wittell M, Gallego-Lara SL, Madrazo-Osuna J, Treatment of multidrug-resistant Acinetobacter baumannii ventilator-associated pneumonia (VAP) with intravenous colistin: A comparison with imipenem-susceptible VAP, Clin Infect Dis, 36 (2003) 1111–1118. [DOI] [PubMed] [Google Scholar]
- [131].Sirijatuphat R, Thamlikitkul V, Preliminary study of colistin versus colistin plus fosfomycin for treatment of carbapenem-resistant Acinetobacter baumannii infections, Antimicrob Agents Chemother, 58 (2014) 5598–5601. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [132].Imberti R, Cusato M, Villani P, Carnevale L, Iotti GA, Langer M, Regazzi M, Steady-state pharmacokinetics and BAL concentration of colistin in critically Ill patients after IV colistin methanesulfonate administration, Chest, 138 (2010) 1333–1339. [DOI] [PubMed] [Google Scholar]
- [133].Yapa SWS, Li J, Patel K, Wilson JW, Dooley MJ, George J, Clark D, Poole S, Williams E, Porter CJ, Nation RL, McIntosh MP, Pulmonary and systemic pharmacokinetics of inhaled and intravenous colistin methanesulfonate in cystic fibrosis patients: Targeting advantage of inhalational administration, Antimicrob Agents Chemother, 58 (2014) 2570–2579. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [134].Schneider-Futschik EK, Paulin OKA, Hoyer D, Roberts KD, Ziogas J, Baker MA, Karas J, Li J, Velkov T, Sputum active polymyxin lipopeptides: Activity against cystic fibrosis Pseudomonas aeruginosa isolates and their interactions with sputum biomolecules, ACS Infect Dis, 4 (2018) 646–655. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [135].Huang JX, Blaskovich MA, Pelingon R, Ramu S, KavanAgh A, Elliott AG, Butler MS, Montgomery AB, Cooper MA, Mucin binding reduces colistin antimicrobial activity, Antimicrob Agents Chemother, 59 (2015) 5925–5931. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [136].Schwameis R, Erdogan-Yildirim Z, Manafi M, Zeitlinger MA, Strommer S, Sauermann R, Effect of pulmonary surfactant on antimicrobial activity in vitro, Antimicrob Agents Chemother, 57 (2013) 5151–5154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [137].Tumbarello M, De Pascale G, Trecarichi EM, De Martino S, Bello G, Maviglia R, Spanu T, Antonelli M, Effect of aerosolized colistin as adjunctive treatment on the outcomes of microbiologically documented ventilator-associated pneumonia caused by colistin-only susceptible Gram-negative bacteria, Chest, 144 (2013) 1768–1775. [DOI] [PubMed] [Google Scholar]
- [138].Kalin G, Alp E, Coskun R, Demiraslan H, Gundogan K, Doganay M, Use of high-dose IV and aerosolized colistin for the treatment of multidrug-resistant Acinetobacter baumannii ventilator-associated pneumonia: Do we really need this treatment, J Infect Chemother, 18 (2012) 872–877. [DOI] [PubMed] [Google Scholar]
- [139].Athanassa ZE, Markantonis SL, Fousteri MZ, Myrianthefs PM, Boutzouka EG, Tsakris A, Baltopoulos GJ, Pharmacokinetics of inhaled colistimethate sodium (CMS) in mechanically ventilated critically ill patients, Intensive Care Med, 38 (2012) 1779–1786. [DOI] [PubMed] [Google Scholar]
- [140].Nation RL, Forrest A Clinical pharmacokinetics, pharmacodynamics and toxicodynamics of polymyxins: implications for therapeutic use. In: Li J, Nation R, Kaye K (eds) Polymyxin antibiotics: From laboratory bench to bedside, Adv Exp Med Biol, 1145. Springer, Cham: (2019) 219–249. [DOI] [PubMed] [Google Scholar]
- [141].Boisson M, Jacobs M, Gregoire N, Gobin P, Marchand S, Couet W, Mimoz O, Comparison of intrapulmonary and systemic pharmacokinetics of colistin methanesulfonate (CMS) and colistin after aerosol delivery and intravenous administration of CMS in critically ill patients, Antimicrob Agents Chemother, 58 (2014) 7331–7339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [142].Liu D, Zhang J, Liu HX, Zhu YG, Qu JM, Intravenous combined with aerosolised polymyxin versus intravenous polymyxin alone in the treatment of pneumonia caused by multidrug-resistant pathogens: A systematic review and meta-analysis, Int J Antimicrob Agents, 46 (2015) 603–609. [DOI] [PubMed] [Google Scholar]
- [143].Sobieszczyk ME, Furuya EY, Hay CM, Pancholi P, Della-Latta P, Hammer SM, Kubin CJ, Combination therapy with polymyxin B for the treatment of multidrug-resistant Gram-negative respiratory tract infections, J Antimicrob Chemother, 54 (2004) 566–569. [DOI] [PubMed] [Google Scholar]
- [144].Ziaka M, Markantonis SL, Fousteri M, Zygoulis P, Panidis D, Karvouniaris M, Makris D, Zakynthinos E, Combined intravenous and intraventricular administration of colistin methanesulfonate in critically ill patients with central nervous system infection, Antimicrob Agents Chemother, 57 (2013) 1938–1940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [145].Imberti R, Iotti GA, Regazzi M, Intraventricular or intrathecal colistin for the treatment of central nervous system infections caused by multidrug-resistant Gram-negative bacteria, Expert Rev Anti Infect Ther, 12 (2014) 471–478. [DOI] [PubMed] [Google Scholar]
- [146].Imberti R, Cusato M, Accetta G, Marino V, Procaccio F, Del Gaudio A, Iotti GA, Regazzi M, Pharmacokinetics of colistin in cerebrospinal fluid after intraventricular administration of colistin methanesulfonate, Antimicrob Agents Chemother, 56 (2012) 4416–4421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [147].Bargiacchi O, Rossati A, Car P, Brustia D, Brondolo R, Rosa F, Garavelli PL, De Rosa FG, Intrathecal/intraventricular colistin in external ventricular device-related infections by multi-drug resistant Gram negative bacteria: Case reports and review, Infection, 42 (2014) 801–809. [DOI] [PubMed] [Google Scholar]
- [148].De Bonis P, Lofrese G, Scoppettuolo G, Spanu T, Cultrera R, Labonia M, Cavallo MA, Mangiola A, Anile C, Pompucci A, Intraventricular versus intravenous colistin for the treatment of extensively drug resistant Acinetobacter baumannii meningitis, Eur J Neurol, 23 (2016) 68–75. [DOI] [PubMed] [Google Scholar]
- [149].Chen H, Guo X, Xie D, Dong X, Niu J, Chen G, A clinical study on the use of intraventricular polymyxin B supplemented by continuous external ventricular drainage in the treatment of drug-resistant Gram-negative bacilli intracranial infection, Infect Drug Resist, 13 (2020) 2963–2970. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [150].Velkov T, Dai C, Ciccotosto GD, Cappai R, Hoyer D, Li J, Polymyxins for CNS infections: Pharmacology and neurotoxicity, Pharmacol Ther, 181 (2018) 85–90. [DOI] [PubMed] [Google Scholar]
- [151].Zavascki A, Nation R, Nephrotoxicity of polymyxins: Is there any difference between colistimethate and polymyxin B, Antimicrob Agents Chemother, 61 (2016) e02319–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [152].Wagenlehner F, Lucenteforte E, Pea F, Soriano A, Tavoschi L, Steele VR, Henriksen AS, Longshaw C, Manissero D, Pecini R, Pogue JM, Systematic review on estimated rates of nephrotoxicity and neurotoxicity in patients treated with polymyxins, Clin Microbiol Infect, (2021) S1198–743X(20)30764–3. [DOI] [PubMed] [Google Scholar]
- [153].Falagas ME, Kasiakou SK, Toxicity of polymyxins: A systematic review of the evidence from old and recent studies, Crit Care, 10 (2006) R27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [154].Brown JM, Dorman DC, Roy LP, Acute renal failure due to overdosage of colistin, Med J Aust, 2 (1970) 923–924. [DOI] [PubMed] [Google Scholar]
- [155].Koch-Weser J, Sidel VW, Federman EB, Kanarek P, Finer DC, E. AE, Adverse effects of sodium colistimethate. Manifestations and specific reaction rates during 317 courses of therapy, Ann Intern Med, 72 (1970) 857–868. [DOI] [PubMed] [Google Scholar]
- [156].Rigatto MH, Oliveira MS, Perdigao-Neto LV, Levin AS, Carrilho CM, Tanita MT, Tuon FF, Cardoso DE, Lopes NT, Falci DR, Zavascki AP, Multicenter prospective cohort study of renal failure in patients treated with colistin versus polymyxin B, Antimicrob Agents Chemother, 60 (2016) 2443–2449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [157].Tuon FF, Rigatto MH, Lopes CK, Kamei LK, Rocha JL, Zavascki AP, Risk factors for acute kidney injury in patients treated with polymyxin B or colistin methanesulfonate sodium, Int J Antimicrob Agents, 43 (2014) 349–352. [DOI] [PubMed] [Google Scholar]
- [158].Azad MAK, Finnin BA, Poudyal A, Davis K, Li J, Hill PA, Nation RL, Velkov T, Li J, Polymyxin B induces apoptosis in kidney proximal tubular cells, Antimicrob Agents Chemother, 57 (2013) 4329–4335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [159].Oliveira MS, Prado GV, Costa SF, Grinbaum RS, Levin AS, Polymyxin B and colistimethate are comparable as to efficacy and renal toxicity, Diagn Microbiol Infect Dis, 65 (2009) 431–434. [DOI] [PubMed] [Google Scholar]
- [160].Phe K, Lee Y, McDaneld PM, Prasad N, Yin T, Figueroa DA, Musick WL, Cottreau JM, Hu M, Tam VH, In vitro assessment and multicenter cohort study of comparative nephrotoxicity rates associated with colistimethate versus polymyxin B therapy, Antimicrob Agents Chemother, 58 (2014) 2740–2746. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [161].Oliota AF, Penteado SuelemT., Tonin FernandaS., Fernandez-Llimos F, Sanches AC, Nephrotoxicity prevalence in patients treated with polymyxins: A systematic review with meta-analysis of observational studies, Diagn Microbiol Infect Dis, 94 (2019) 41–49. [DOI] [PubMed] [Google Scholar]
- [162].Vardakas KZ, Falagas ME, Colistin versus polymyxin B for the treatment of patients with multidrug-resistant Gram-negative infections: A systematic review and meta-analysis, Int J Antimicrob Agents, 49 (2017) 233–238. [DOI] [PubMed] [Google Scholar]
- [163].Chien HT, Lin YC, Sheu CC, Hsieh KP, Chang JS, Is colistin-associated acute kidney injury clinically important in adults? A systematic review and meta-analysis, Int J Antimicrob Agents, 55 (2020) 105889. [DOI] [PubMed] [Google Scholar]
- [164].Dai C, Li J, Tang S, Li J, Xiao X, Colistin-induced nephrotoxicity in mice involves the mitochondrial, death receptor, and endoplasmic reticulum pathways, Antimicrob Agents Chemother, 58 (2014) 4075–4085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [165].Azad MA, Akter J, Rogers KL, Nation RL, Velkov T, Li J, Major pathways of polymyxin-induced apoptosis in rat kidney proximal tubular cells, Antimicrob Agents Chemother, 59 (2015) 2136–2143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [166].Azad MA, Roberts KD, Yu HH, Liu B, Schofield AV, James SA, Howard DL, Nation RL, Rogers K, de Jonge MD, Thompson PE, Fu J, Velkov T, Li J, Significant accumulation of polymyxin in single renal tubular cells: A medicinal chemistry and triple correlative microscopy approach, Anal Chem, 87 (2015) 1590–1595. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [167].Lu J, Mohammad Azad AK, Moreau L. M. Julie., Zhu Y, Jiang XK, Tonta M, Lam R, Wickremasinghe H, Zhao JX, Wang JP, Coleman HA, Formosa LE, Velkov T, Parkington HC, Combes AN, Rosenbluh J, Li J, Polymyxin-induced nephrotoxicity involves potassium channel and energy generation in human kidney proximal tubular cells. Poster presented at: ASM Microbe 2020, Jun 18 – 22, Chicago, Illinois. [Google Scholar]
- [168].Dai C, Ciccotosto GD, Cappai R, Wang Y, Tang S, Xiao X, Velkov T, Minocycline attenuates colistin-induced neurotoxicity via suppression of apoptosis, mitochondrial dysfunction and oxidative stress, J Antimicrob Chemother, 72 (2017) 1635–1645. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [169].Dai C, Ciccotosto GD, Cappai R, Tang S, Li D, Xie S, Xiao X, Velkov T, Curcumin attenuates colistin-induced neurotoxicity in N2a cells via anti-inflammatory activity, suppression of oxidative stress, and apoptosis, Mol Neurobiol, 55 (2018) 421–434. [DOI] [PubMed] [Google Scholar]
- [170].Weinstein L, Doan TL, Smith MA, Neurotoxicity in patients treated with intravenous polymyxin B: Two case reports, Am J Health Syst Pharm, 66 (2009) 345–347. [DOI] [PubMed] [Google Scholar]
- [171].Holloway KP, Rouphael NG, Wells JB, King MD, Blumberg HM, Polymyxin B and doxycycline use in patients with multidrug-resistant Acinetobacter baumannii infections in the intensive care unit, Ann Pharmacother, 40 (2006) 1939–1945. [DOI] [PubMed] [Google Scholar]
- [172].Wunsch H, Moitra VK, Patel M, Dzierba AL, Polymyxin use associated with respiratory arrest, Chest, 141 (2012) 515–517. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [173].Wahby K, Chopra T, Chandrasekar P, Intravenous and inhalational colistin-induced respiratory failure, Clin Infect Dis, 50 (2010) e38–40. [DOI] [PubMed] [Google Scholar]
- [174].Honore PM, Jacobs R, Lochy S, De Waele E, Van Gorp V, De Regt J, Martens G, Joannes-Boyau O, Boer W, Spapen HD, Acute respiratory muscle weakness and apnea in a critically ill patient induced by colistin neurotoxicity: Key potential role of hemoadsorption elimination during continuous venovenous hemofiltration, Int J Nephrol Renovasc Dis, 6 (2013) 107–111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [175].Knueppel Ryan C., Rahimian J, Diffuse cutaneous hyperpigmentation due to tigecycline or polymyxin B, Clin Infect Dis, 45 (2007) 136–138. [DOI] [PubMed] [Google Scholar]
- [176].Zheng G, Cao L, Che Z, Mao E, Chen E, He J, Polymyxin B-induced skin hyperpigmentation: A rare case report and literature review, BMC Pharmacol Toxicol, 19 (2018) 41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [177].Wen X, Luo C, Lyu W, Polymyxin B-induced skin hyperpigmentation, Case Rep Med, (2020) 6461329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [178].Alothman GA, Ho B, Alsaadi MM, Ho SL, O’Drowsky L, Louca E, Coates AL, Bronchial constriction and inhaled colistin in cystic fibrosis, Chest, 127 (2005) 522–529. [DOI] [PubMed] [Google Scholar]
- [179].Maddison J, Dodd M, Webb AK, Nebulized colistin causes chest tightness in adults with cystic fibrosis, Respir Med, 88 (1994) 145–147. [DOI] [PubMed] [Google Scholar]
- [180].Dodd ME, Abbott J, Maddison J, Moorcroft AJ, Webb AK, Effect of tonicity of nebulised colistin on chest tightness and pulmonary function in adults with cystic fibrosis, Thorax, 52 (1997) 656–658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [181].Westerman EM, Le Brun PP, Touw DJ, Frijlink HW, Heijerman HG, Effect of nebulized colistin sulphate and colistin sulphomethate on lung function in patients with cystic fibrosis: A pilot study, J Cyst Fibros, 3 (2004) 23–28. [DOI] [PubMed] [Google Scholar]
- [182].Ahmed MU, Velkov T, Lin YW, Yun B, Nowell CJ, Zhou F, Zhou QT, Chan K, Azad MAK, Li J, Potential toxicity of polymyxins in human lung epithelial cells, Antimicrob Agents Chemother, 61 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [183].Fathalla AM, Chow SH, Naderer T, Zhou QT, Velkov T, Azad MAK, Li J, Polymyxin-induced cell death of human macrophage-likeTHP-1 and neutrophil-like HL-60 cells associated with theactivation of apoptotic pathways, Antimicrob Agents Chemother, 64 (2020) e00013–00020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [184].Avila MP, Pacheco T, Arias S, Bustos RH, Garcia JC, Jaimes D, is there a role for the therapeutic drug monitoring of colistin? An overview, Pharmaceuticals (Basel), 13 (2020) 42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [185].Lakota EA, Landersdorfer CB, Nation RL, Li J, Kaye KS, Rao GG, Forrest A, Personalizing polymyxin B dosing using an adaptive feedback control algorithm, Antimicrob Agents Chemother, 62 (2018) e00483–00418. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [186].Kim EJ, Oh J, Lee K, Yu KS, Chung JY, Hwang JH, Nam EY, Kim HS, Kim M, Park JS, Song KH, Kim ES, Song J, K. HB, Pharmacokinetic characteristics and limited sampling strategies for therapeutic drug monitoring of colistin in patients with multidrug-resistant Gram-negative bacterial infections., Ther Drug Monit, 41 (2019) 102–106. [DOI] [PubMed] [Google Scholar]
- [187].Wang P, Zhang Q, Zhu Z, Feng M, Sun T, Yang J, Zhang X, Population pharmacokinetics and limited sampling strategy for therapeutic drug monitoring of polymyxin B in Chinese patients with multidrug-resistant Gram-negative bacterial infections, Front Pharmacol, 11 (2020) 829. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [188].Chen W, Liu H, Wang Q, Wang X, Kong X, Wang X, Zhang X, Zhan Q, Li P, Estimation of the area under concentration-time curve of polymyxin B based on limited sampling concentrations in Chinese patients with severe pneumonia, Eur J Clin Pharmacol, 77 (2021) 95–105. [DOI] [PubMed] [Google Scholar]
- [189].Sorlí L, Luque S, Grau S, Berenguer N, Segura C, Montero MM, Alvarez-Lerma F, Knobel H, Benito N, Horcajada JP, Trough colistin plasma level is an independent risk factor for nephrotoxicity: A prospective observational cohort study., BMC Infect Dis, 13 (2013) 380. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [190].Forrest A, Garonzik SM, Thamlikitkul V, Giamarellos-Bourboulis EJ, Paterson DL, Li J, Silveira FP, Nation RL, Pharmacokinetic/toxicodynamic analysis of colistin-associated acute kidney injury in critically Ill Patients, Antimicrob Agents Chemother, 61 (2017) e01367–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [191].Vaara M, Polymyxins and their potential next generation as therapeutic antibiotics, Front Microbiol, 10 (2019) 1689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [192].Velkov T, Roberts KD, Discovery of novel polymyxin-like antibiotics, Adv Exp Med Biol, In: Li J, Nation R, Kaye K (eds) Polymyxin antibiotics: From laboratory bench to bedside, Adv Exp Med Biol, 1145. Springer, Cham: (2019) 343–362. [DOI] [PubMed] [Google Scholar]
- [193].Velkov T, Roberts KD, Nation RL, Wang J, Thompson PE, Li J, Teaching ‘old’ polymyxins new tricks: New-generation lipopeptides targeting Gram-negative ‘superbugs’, ACS Chem Biol, 9 (2014) 1172–1177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [194].Grosser L, Heang K, Teague J, Warn P, Corbett D, Dawson MJ, et al. (2018b). “In vivo efficacy of SPR206 in murine lung and thigh infection models caused by multidrug resistant pathogens Pseudomonas aeruginosa and Acinetobacter baumannii,” in Poster at P-139, ASM ESCMID (2018), (Lisbon). [Google Scholar]
- [195].Boakes S, Duperchy E, Brown P, Simonovic M, Abdulle O, Divall N, et al. (2015). “Novel polymyxin derivative CA824: Efficacy in neutropaenic mouse infection and lung infection models,” in Poster at F-735 ICAAC, (San Diego, CA: ICAAC; ). [Google Scholar]
- [196].Gordeev MF, Liu J, Wang Z, and Yuan Z (2016). Antimicrobial polymyxins for treatment of bacterial infections. Patent Application WO2016/100578
- [197].Castanheira Mariana, Lindley Jill, Huynh Holly, Mendes Rodrigo E., Lomovskaya O, Activity of a novel polymyxin analog, QPX9003, Tested against resistant Gram-negative pathogens, including carbapenem-resistant Acinetobacter, Enterobacterales, and Pseudomonas, ID Week (2019). [Google Scholar]
- [198].Sabet M, Tarazi Z, Parkinson J, Roberts KD, Thompson PE, Nation RL, Velkov T, Hecker SJ, Lomovskaya O, Dudley MN, Li J, Griffith DC, QPX9003: Pharmacology of a novel polymyxin in mice and rats, ID week, (2019). [Google Scholar]
- [199].Lepak Alexander J, Wen Wang, Andes DR, Pharmacodynamic evaluation of MRX-8, a novel polymyxin, in the neutropenic mouse thigh and lung infection models against Gram-negative pathogens, Antimicrob Agents Chemother, 64 (2020) e01517–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [200].Arends SJR, Rhomberg PR, Lister T, Cotroneo N, Rubio A, Flamm RK, Mendes R, In vitro Activity Evaluation of a Next-Generation Polymyxin, SPR206, against Non-Fermentative Gram-negative Bacilli Responsible for Human Infections, ASM-ESCMID, (2018). [Google Scholar]
- [201].Lister T, SPR206: Next Generation Polymyxin for the Treatment of Highly Resistant Gram-negative Bacterial Infections, ASM Microbe, (2019). [Google Scholar]
- [202].Pharmacology studies planned to initiate in the second half of 2020, Spero Therapeutics, Inc. [Google Scholar]
- [203].Bruss J, Lister T, Gupta VK, Stone E, Morelli L, Lei Y, Melnick D, Single- and Multiple-ascending dose study of the safety, tolerability, and pharmacokinetics of the polymyxin derivative, SPR206, Antimicrob Agents Chemother, (2021) AAC0073921. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [204].Vazquez-Guillamet C, Kollef MH, Treatment of Gram-positive infections in critically ill patients, BMC Infect Dis, 14 (2014) 92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [205].Ferrara AM, Treatment of hospital-acquired pneumonia caused by methicillin-resistant Staphylococcus aureus, Int J Antimicrob Agents, 30 (2007) 19–24. [DOI] [PubMed] [Google Scholar]
- [206].Kang HK, Park Y, Glycopeptide antibiotics: Structure and mechanisms of action, J Bacteriol Virol, 45 (2015) 67–78. [Google Scholar]
- [207].Barna JC, Williams DH, The structure and mode of action of glycopeptide antibiotics of the vancomycin group, Annu Rev Microbiol, 38 (1984) 339–357. [DOI] [PubMed] [Google Scholar]
- [208].McCormick MH, McGuire JM, Pittenger GE, Pittenger RC, Stark WM, Vancomycin, a new antibiotic. I. Chemical and biologic properties, Antibiot Annu, 3 (1955) 606–611. [PubMed] [Google Scholar]
- [209].James RC, Pierce JG, Okano A, Xie J, Boger DL, Redesign of glycopeptide antibiotics: Back to the future, ACS Chem Biol, 7 (2012) 797–804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [210].Treviño J, Bayón C, Ardá A, Marinelli F, Gandolfi R, Molinari F, Jimenez-Barbero J, Hernáiz MJ, New insights into glycopeptide antibiotic binding to cell wall precursors using SPR and NMR spectroscopy, Chemistry, 20 (2014) 7363–7372. [DOI] [PubMed] [Google Scholar]
- [211].Parenti F, Beretta G, Berti M, Arioli V, Teichomycins, new antibiotics from Actinoplanes teichomyceticus Nov. Sp. I. Description of the producer strain, fermentation studies and biological properties, J Antibiot (Tokyo), 31 (1978) 276–283. [DOI] [PubMed] [Google Scholar]
- [212].Davey PG, Williams AH, A review of the safety profile of teicoplanin, J Antimicrob Chemother, 27 Suppl B (1991) 69–73. [DOI] [PubMed] [Google Scholar]
- [213].Rowland M, Clinical pharmacokinetics of teicoplanin, Clin Pharmacokinet, 18 (1990) 184–209. [DOI] [PubMed] [Google Scholar]
- [214].Printsevskaya SS, Pavlov AY, Olsufyeva EN, Mirchink EP, Preobrazhenskaya MN, Role of the glycopeptide framework in the antibacterial activity of hydrophobic derivatives of glycopeptide antibiotics, J Med Chem, 46 (2003) 1204–1209. [DOI] [PubMed] [Google Scholar]
- [215].Leadbetter MR, Adams SM, Bazzini B, Fatheree PR, Karr DE, Krause KM, Lam BM, Linsell MS, Nodwell MB, Pace JL, Quast K, Shaw JP, Soriano E, Trapp SG, Villena JD, Wu TX, Christensen BG, Judice JK, Hydrophobic vancomycin derivatives with improved ADME properties: Discovery of telavancin (TD-6424), J Antibiot (Tokyo), 57 (2004) 326–336. [DOI] [PubMed] [Google Scholar]
- [216].Higgins DL, Chang R, Debabov DV, Leung J, Wu T, Krause KM, Sandvik E, Hubbard JM, Kaniga K, Schmidt DE Jr., Gao Q, Cass RT, Karr DE, Benton BM, Humphrey PP, Telavancin, a multifunctional lipoglycopeptide, disrupts both cell wall synthesis and cell membrane integrity in methicillin-resistant Staphylococcus aureus, Antimicrob Agents Chemother, 49 (2005) 1127–1134. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [217].Zhanel GG, Trapp S, Gin AS, DeCorby M, Lagacé-Wiens PR, Rubinstein E, Hoban DJ, Karlowsky JA, Dalbavancin and telavancin: Novel lipoglycopeptides for the treatment of Gram-positive infections, Expert Rev Anti Infect Ther, 6 (2008) 67–81. [DOI] [PubMed] [Google Scholar]
- [218].Laohavaleeson S, Kuti JL, Nicolau DP, Telavancin: A novel lipoglycopeptide for serious Gram-positive infections, Expert Opin Investig Drugs, 16 (2007) 347–357. [DOI] [PubMed] [Google Scholar]
- [219].Malabarba A, Goldstein BP, Origin, structure, and activity in vitro and in vivo of dalbavancin, J Antimicrob Chemother, 55 Suppl 2 (2005) ii15–20. [DOI] [PubMed] [Google Scholar]
- [220].Smith JR, Roberts KD, Rybak MJ, Dalbavancin: A novel lipoglycopeptide antibiotic with extended activity against Gram-positive infections, Infect Dis Ther, 4 (2015) 245–258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [221].Nagarajan R, Marcel Dekker, Glycopeptide antibiotics. Drugs Pharm Sci, 63 (1994) 63–104. [Google Scholar]
- [222].Butler MS, Hansford KA, Blaskovich MA, Halai R, Cooper MA, Glycopeptide antibiotics: Back to the future, J Antibiot (Tokyo), 67 (2014) 631–644. [DOI] [PubMed] [Google Scholar]
- [223].Zhanel GG, Calic D, Schweizer F, Zelenitsky S, Adam H, Lagacé-Wiens PR, Rubinstein E, Gin AS, Hoban DJ, Karlowsky JA, New lipoglycopeptides: A comparative review of dalbavancin, oritavancin and telavancin, Drugs, 70 (2010) 859–886. [DOI] [PubMed] [Google Scholar]
- [224].Allen NE, Nicas TI, Mechanism of action of oritavancin and related glycopeptide antibiotics, FEMS Microbiol Rev, 26 (2003) 511–532. [DOI] [PubMed] [Google Scholar]
- [225].Cooper RD, Snyder NJ, Zweifel MJ, Staszak MA, Wilkie SC, Nicas TI, Mullen DL, Butler TF, Rodriguez MJ, Huff BE, Thompson RC, Reductive alkylation of glycopeptide antibiotics: Synthesis and antibacterial activity, J Antibiot (Tokyo), 49 (1996) 575–581. [DOI] [PubMed] [Google Scholar]
- [226].Belley A, McKay GA, Arhin FF, Sarmiento I, Beaulieu S, Fadhil I, Parr TR Jr., Moeck G, Oritavancin disrupts membrane integrity of Staphylococcus aureus and vancomycin-resistant enterococci to effect rapid bacterial killing, Antimicrob Agents Chemother, 54 (2010) 5369–5371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [227].Reynolds PE, Structure, biochemistry and mechanism of action of glycopeptide antibiotics, Eur J Clin Microbiol Infect Dis, 8 (1989) 943–950. [DOI] [PubMed] [Google Scholar]
- [228].Courvalin P, Vancomycin resistance in Gram-positive cocci, Clin Infect Dis, 42 Suppl 1 (2006) S25–34. [DOI] [PubMed] [Google Scholar]
- [229].Mühlberg E, Umstätter F, Kleist C, Domhan C, Mier W, Uhl P, Renaissance of vancomycin: Approaches for breaking antibiotic resistance in multidrug-resistant bacteria, Can J Microbiol, 66 (2020) 11–16. [DOI] [PubMed] [Google Scholar]
- [230].Cooper MA, Williams DH, Binding of glycopeptide antibiotics to a model of a vancomycin-resistant bacterium, Chem Biol, 6 (1999) 891–899. [DOI] [PubMed] [Google Scholar]
- [231].Yang Z, Vorpagel ER, Laskin J, Influence of the charge state on the structures and interactions of vancomycin antibiotics with cell-wall analogue peptides: Experimental and theoretical studies, Chemistry, 15 (2009) 2081–2090. [DOI] [PubMed] [Google Scholar]
- [232].Rekharsky M, Hesek D, Lee M, Meroueh SO, Inoue Y, Mobashery S, Thermodynamics of interactions of vancomycin and synthetic surrogates of bacterial cell wall, J Am Chem Soc, 128 (2006) 7736–7737. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [233].Beauregard DA, Williams DH, Gwynn MN, Knowles DJ, Dimerization and membrane anchors in extracellular targeting of vancomycin group antibiotics, Antimicrob Agents Chemother, 39 (1995) 781–785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [234].Sieradzki K, Tomasz A, Inhibition of cell wall turnover and autolysis by vancomycin in a highly vancomycin-resistant mutant of Staphylococcus aureus, J Bacteriol, 179 (1997) 2557–2566. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [235].Peschel A, Vuong C, Otto M, Götz F, The D-Alanine residues of Staphylococcus aureus teichoic acids alter the susceptibility to vancomycin and the activity of autolytic enzymes, Antimicrob Agents Chemother, 44 (2000) 2845–2847. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [236].Jordan DC, Inniss WE, Selective inhibition of ribonucleic acid synthesis in Staphylococcus aureus by vancomycin, Nature, 184(Suppl 24) (1959) 1894–1895. [DOI] [PubMed] [Google Scholar]
- [237].Kim A, Kuti JL, Nicolau DP, Review of dalbavancin, a novel semisynthetic lipoglycopeptide, Expert Opin Investig Drugs, 16 (2007) 717–733. [DOI] [PubMed] [Google Scholar]
- [238].Belley A, Neesham-Grenon E, McKay G, Arhin FF, Harris R, Beveridge T, Parr TR Jr., Moeck G, Oritavancin kills stationary-phase and biofilm Staphylococcus aureus cells In vitro, Antimicrob Agents Chemother, 53 (2009) 918–925. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [239].Ashford PA, Bew SP, Recent advances in the synthesis of new glycopeptide antibiotics, Chem Soc Rev, 41 (2012) 957–978. [DOI] [PubMed] [Google Scholar]
- [240].Saravolatz LD, Stein GE, Johnson LB, Telavancin: A novel lipoglycopeptide, Clin Infect Dis, 49 (2009) 1908–1914. [DOI] [PubMed] [Google Scholar]
- [241].Patti GJ, Kim SJ, Yu TY, Dietrich E, Tanaka KS, Parr TR Jr., Far AR, Schaefer J, Vancomycin and oritavancin have different modes of action in Enterococcus faecium, J Mol Biol, 392 (2009) 1178–1191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [242].Das B, Sarkar C, Schachter J, Oritavancin - A new semisynthetic lipoglycopeptide agent to tackle the challenge of resistant Gram-positive pathogens, Pak J Pharm Sci, 26 (2013) 1045–1055. [PubMed] [Google Scholar]
- [243].Syed YY, Scott LJ, Oritavancin: A review in acute bacterial skin and skin structure infections, Drugs, 75 (2015) 1891–1902. [DOI] [PubMed] [Google Scholar]
- [244].Morrissey I, Seifert H, Canton R, Nordmann P, Stefani S, Macgowan A, Janes R, Knight D, Activity of oritavancin against methicillin-resistant staphylococci, vancomycin-resistant enterococci and β-haemolytic streptococci collected from western European countries in 2011, J Antimicrob Chemother, 68 (2013) 164–167. [DOI] [PubMed] [Google Scholar]
- [245].Mendes RE, Woosley LN, Farrell DJ, Sader HS, Jones RN, Oritavancin activity against vancomycin-susceptible and vancomycin-resistant Enterococci with molecularly characterized glycopeptide resistance genes recovered from bacteremic patients, 2009–2010, Antimicrob Agents Chemother, 56 (2012) 1639–1642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [246].Leclercq R, Derlot E, Duval J, Courvalin P, Plasmid-mediated resistance to vancomycin and teicoplanin in Enterococcus faecium, N Engl J Med, 319 (1988) 157–161. [DOI] [PubMed] [Google Scholar]
- [247].Uttley AH, Collins CH, Naidoo J, George RC, Vancomycin-resistant enterococci, Lancet, 1 (1988) 57–58. [DOI] [PubMed] [Google Scholar]
- [248].Cattoir V, Leclercq R, Twenty-five years of shared life with vancomycin-resistant enterococci: Is it time to divorce, J Antimicrob Chemother, 68 (2013) 731–742. [DOI] [PubMed] [Google Scholar]
- [249].Howden BP, Davies JK, Johnson PD, Stinear TP, Grayson ML, Reduced vancomycin susceptibility in Staphylococcus aureus, including vancomycin-intermediate and heterogeneous vancomycin-intermediate strains: Resistance mechanisms, laboratory detection, and clinical implications, Clin Microbiol Rev, 23 (2010) 99–139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [250].Sato T, Wada T, Shinagawa M, Fukushima Y, Nakajima C, Suzuki Y, Takahashi S, Yokota SI, Emergence of vancomycin- and teicoplanin-resistant Enterococcus faecium via vanD5-harbouring large genomic island, J Antimicrob Chemother, 75 (2020) 2411–2415. [DOI] [PubMed] [Google Scholar]
- [251].Wagner TM, Janice J, Sivertsen A, Sjögren I, Sundsfjord A, Hegstad K, Alternative vanHAX promoters and increased vanA-plasmid copy number resurrect silenced glycopeptide resistance in Enterococcus faecium, J Antimicrob Chemother, 76 (2021) 876–882. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [252].Bugg TD, Wright GD, Dutka-Malen S, Arthur M, Courvalin P, Walsh CT, Molecular basis for vancomycin resistance in Enterococcus faecium BM4147: Biosynthesis of a depsipeptide peptidoglycan precursor by vancomycin resistance proteins VanH and VanA, Biochemistry, 30 (1991) 10408–10415. [DOI] [PubMed] [Google Scholar]
- [253].Périchon B, Courvalin P, VanA-type vancomycin-resistant Staphylococcus aureus, Antimicrob Agents Chemother, 53 (2009) 4580–4587. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [254].Stogios PJ, Savchenko A, Molecular mechanisms of vancomycin resistance, Protein Sci, 29 (2020) 654–669. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [255].McGuinness WA, Malachowa N, DeLeo FR, Vancomycin Resistance in Staphylococcus aureus Yale J Biol Med, 90 (2017) 269–281. [PMC free article] [PubMed] [Google Scholar]
- [256].Roch M, Clair P, Renzoni A, Reverdy ME, Dauwalder O, Bes M, Martra A, Freydière AM, Laurent F, Reix P, Dumitrescu O, Vandenesch F, Exposure of Staphylococcus aureus to subinhibitory concentrations of β-lactam antibiotics induces heterogeneous vancomycin-intermediate Staphylococcus aureus, Antimicrob Agents Chemother, 58 (2014) 5306–5314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [257].Kim JW, Lee KJ, Single-nucleotide polymorphisms in a vancomycin-resistant Staphylococcus aureus strain based on whole-genome sequencing, Arch Microbiol, 202 (2020) 2255–2261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [258].Weigel LM, Clewell DB, Gill SR, Clark NC, McDougal LK, Flannagan SE, Kolonay JF, Shetty J, Killgore GE, Tenover FC, Genetic analysis of a high-level vancomycin-resistant isolate of Staphylococcus aureus, Science, 302 (2003) 1569–1571. [DOI] [PubMed] [Google Scholar]
- [259].Chang S, Sievert DM, Hageman JC, Boulton ML, Tenover FC, Downes FP, Shah S, Rudrik JT, Pupp GR, Brown WJ, Cardo D, Fridkin SK, Infection with vancomycin-resistant Staphylococcus aureus containing the vanA resistance gene, N Engl J Med, 348 (2003) 1342–1347. [DOI] [PubMed] [Google Scholar]
- [260].Billot-Klein D, Blanot D, Gutmann L, van Heijenoort J, Association constants for the binding of vancomycin and teicoplanin to N-Acetyl-D-Alanyl-D-Alanine and N-Acetyl-D-Alanyl-D-Serine, Biochem J, 304 (Pt 3) (1994) 1021–1022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [261].Kosowska-Shick K, Clark C, Pankuch GA, McGhee P, Dewasse B, Beachel L, Appelbaum PC, Activity of telavancin against staphylococci and enterococci determined by MIC and resistance selection studies, Antimicrob Agents Chemother, 53 (2009) 4217–4224. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [262].Goldstein BP, Draghi DC, Sheehan DJ, Hogan P, Sahm DF, Bactericidal activity and resistance development profiling of dalbavancin, Antimicrob Agents Chemother, 51 (2007) 1150–1154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [263].Arthur M, Depardieu F, Reynolds P, Courvalin P, Moderate-level resistance to glycopeptide LY333328 mediated by genes of the vanA and vanB clusters in Enterococci, Antimicrob Agents Chemother, 43 (1999) 1875–1880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [264].Vimberg V, Zieglerová L, Buriánková K, Branny P, Balíková Novotná G, VanZ reduces the binding of lipoglycopeptide antibiotics to Staphylococcus aureus and Streptococcus pneumoniae cells, Front Microbiol, 11 (2020) 566. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [265].Perl TM, The threat of vancomycin resistance, Am J Med, 106 (1999) 26S–37S; discussion 48S-52S. [DOI] [PubMed] [Google Scholar]
- [266].Wood MJ, The comparative efficacy and safety of teicoplanin and vancomycin, J Antimicrob Chemother, 37 (1996) 209–222. [DOI] [PubMed] [Google Scholar]
- [267].Cavalcanti AB, Goncalves AR, Almeida CS, Bugano DD, Silva E, Teicoplanin versus vancomycin for proven or suspected infection, Cochrane Database Syst Rev, (2010) Cd007022. [DOI] [PubMed] [Google Scholar]
- [268].Peng Y, Ye X, Li Y, Bu T, Chen X, Bi J, Zhou J, Yao Z, Teicoplanin as an effective alternative to vancomycin for treatment of MRSA infection in Chinese population: A meta-analysis of randomized controlled trials, Plos One, 8 (2013) e79782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [269].Loo VG, Lavallée J, McAlear D, Robson HG, Activity of vancomycin, teicoplanin and cephalosporins against penicillin-susceptible and penicillin-intermediate Streptococcus pneumoniae, Can J Infect Dis, 6 (1995) 83–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [270].Knudsen JD, The importance of pharmacodynamic properties in treatment of penicillin resistant Streptococcus pneumoniae, Dan Med Bull, 47 (2000) 313–327. [PubMed] [Google Scholar]
- [271].Felmingham D, Foxall P, O’Hare M, Grüneberg R, The bactericidal activity of vancomycin and teicoplanin against Streptococcus pneumoniae, Scand J Infect Dis Suppl, 72 (1990) 20–25. [PubMed] [Google Scholar]
- [272].Brauers J, Kresken M, Menke A, Orland A, Weiher H, Morrissey I, Bactericidal activity of daptomycin, vancomycin, teicoplanin and linezolid against Staphylococcus aureus, Enterococcus faecalis and Enterococcus faecium using human peak free serum drug concentrations, Int J Antimicrob Agents, 29 (2007) 322–325. [DOI] [PubMed] [Google Scholar]
- [273].Zeng D, Debabov D, Hartsell TL, Cano RJ, Adams S, Schuyler JA, McMillan R, Pace JL, Approved glycopeptide antibacterial drugs: Mechanism of action and resistance, Cold Spring Harb Perspect Med, 6 (2016) a026989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [274].Löwdin E, Odenholt I, Cars O, In vitro studies of pharmacodynamic properties of vancomycin against Staphylococcus aureus and Staphylococcus epidermidis, Antimicrob Agents Chemother, 42 (1998) 2739–2744. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [275].Larsson AJ, Walker KJ, Raddatz JK, Rotschafer JC, The concentration-independent effect of monoexponential and biexponential decay in vancomycin concentrations on the killing of Staphylococcus aureus under aerobic and anaerobic conditions, J Antimicrob Chemother, 38 (1996) 589–597. [DOI] [PubMed] [Google Scholar]
- [276].Greenberg RN, Benes CA, Time-kill studies with oxacillin, vancomycin, and teicoplanin versus Staphylococcus aureus, J Infect Dis, 161 (1990) 1036–1037. [DOI] [PubMed] [Google Scholar]
- [277].Peetermans WE, Hoogeterp JJ, Hazekamp-van Dokkum AM, van den Broek P, Mattie H, Antistaphylococcal activities of teicoplanin and vancomycin in vitro and in an experimental infection, Antimicrob Agents Chemother, 34 (1990) 1869–1874. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [278].Ravizzola G, Pirali F, Foresti I, Turano A, Comparison of the in vitro antibacterial activity of teicoplanin and vancomycin against Gram-positive cocci, Drugs Exp Clin Res, 13 (1987) 225–229. [PubMed] [Google Scholar]
- [279].Lee DG, Murakami Y, Andes DR, Craig WA, Inoculum effects of ceftobiprole, daptomycin, linezolid, and vancomycin with Staphylococcus aureus and Streptococcus pneumoniae at inocula of 10(5) and 10(7) CFU injected into opposite thighs of neutropenic mice, Antimicrob Agents Chemother, 57 (2013) 1434–1441. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [280].Rio-Marques L, Hartke A, Bizzini A, The effect of inoculum size on selection of in vitro resistance to vancomycin, daptomycin, and linezolid in methicillin-resistant Staphylococcus aureus, Microb Drug Resist, 20 (2014) 539–543. [DOI] [PubMed] [Google Scholar]
- [281].LaPlante KL, Rybak MJ, Impact of high-inoculum Staphylococcus aureus on the activities of nafcillin, vancomycin, linezolid, and daptomycin, alone and in combination with gentamicin, in an in vitro pharmacodynamic model, Antimicrob Agents Chemother, 48 (2004) 4665–4672. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [282].Ambrose PG, Drusano GL, Craig WA, In vivo activity of oritavancin in animal infection models and rationale for a new dosing regimen in humans, Clin Infect Dis, 54 Suppl 3 (2012) S220–228. [DOI] [PubMed] [Google Scholar]
- [283].Andes D, Craig WA, In vivo pharmacodynamic activity of the glycopeptide dalbavancin, Antimicrob Agents Chemother, 51 (2007) 1633–1642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [284].Biavasco F, Vignaroli C, Lupidi R, Manso E, Facinelli B, Varaldo PE, In vitro antibacterial activity of LY333328, a new semisynthetic glycopeptide, Antimicrob Agents Chemother, 41 (1997) 2165–2172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [285].Greenwood D, Bidgood K, Turner M, A comparison of the responses of staphylococci and streptococci to teicoplanin and vancomycin, J Antimicrob Chemother, 20 (1987) 155–164. [DOI] [PubMed] [Google Scholar]
- [286].Thabaut A, Meyran M, Activité in vitro de la Teichomycine, de la Vancomycine et de la N-Formimidoyl-Thienamycine sur Staphylococcus aureus, Med Mal Infect, 13 (1983) 272–277. [Google Scholar]
- [287].Odenholt I, Löwdin E, Cars O, In vitro studies of the pharmacodynamics of teicoplanin against Staphylococcus aureus, Staphylococcus epidermidis and Enterococcus faecium, Clin Microbiol Infect, 9 (2003) 930–937. [DOI] [PubMed] [Google Scholar]
- [288].Cooper MA, Jin YF, Ashby JP, Andrews JM, Wise R, In-vitro comparison of the post-antibiotic effect of vancomycin and teicoplanin, J Antimicrob Chemother, 26 (1990) 203–207. [DOI] [PubMed] [Google Scholar]
- [289].Hanberger H, Nilsson LE, Maller R, Isaksson B, Pharmacodynamics of daptomycin and vancomycin on Enterococcus faecalis and Staphylococcus aureus demonstrated by studies of initial killing and postantibiotic effect and influence of Ca2+ and albumin on these drugs, Antimicrob Agents Chemother, 35 (1991) 1710–1716. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [290].Shaw JP, Seroogy J, Kaniga K, Higgins DL, Kitt M, Barriere S, Pharmacokinetics, serum inhibitory and bactericidal activity, and safety of telavancin in healthy subjects, Antimicrob Agents Chemother, 49 (2005) 195–201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [291].Matzke GR, Zhanel GG, Guay DR, Clinical pharmacokinetics of vancomycin, Clin Pharmacokinet, 11 (1986) 257–282. [DOI] [PubMed] [Google Scholar]
- [292].Shi Y, Zhang J, Yu J, Cao G, Zhang Y, The pharmacokinetics of levofloxacin and absolute bioavailability study of levofloxacin tablet in Chinese volunteers, Acta Pharmacol Sin, 27 (2006) 215–215. [Google Scholar]
- [293].Cao G, Shi Y, Zhang J, Yu J, Guo B, An evaluation of the pharmacokinetics of single and multiple doses of gemifloxacin in Chinese healthy subjects, Acta Pharmacol Sin, 27 (2006) 216–216. [Google Scholar]
- [294].Follin SL, Mueller BA, Scott MK, Carfagna MA, Kraus MA, Falsely elevated serum vancomycin concentrations in hemodialysis patients, Am J Kidney Dis, 27 (1996) 67–74. [DOI] [PubMed] [Google Scholar]
- [295].Launay-Vacher V, Izzedine H, Baumelou A, Deray G, FHD: An index to evaluate drug elimination by hemodialysis, Am J Nephrol, 25 (2005) 342–351. [DOI] [PubMed] [Google Scholar]
- [296].Bourguignon L, Cazaubon Y, Debeurme G, Loue C, Ducher M, Goutelle S, Pharmacokinetics of vancomycin in elderly patients aged over 80 years, Antimicrob Agents Chemother, 60 (2016) 4563–4567. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [297].Sosnin N, Curtis N, Cranswick N, Chiletti R, Gwee A, Vancomycin is commonly under-dosed in critically ill children and neonates, Br J Clin Pharmacol, 85 (2019) 2591–2598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [298].Byrne CJ, Roberts JA, McWhinney B, Ryder SA, Fennell JP, O’Byrne P, Deasy E, Egan S, Desmond R, Enright H, D’Arcy DM, McHugh J, Population pharmacokinetics of teicoplanin and attainment of pharmacokinetic/pharmacodynamic targets in adult patients with haematological malignancy. Clin Microbiol Infect. 9 (2017) 23. [DOI] [PubMed] [Google Scholar]
- [299].Smithers JA, Thompson GA, Kenny MT, Dulworth JK, Kulmala HK, Lewis EW, Ruberg SJ, Shapiro BE, Keane WF, Halstenson CE, Applicability of teicoplanin dosage adjustment guidelines for renally impaired patients over the range of 3 to 30 mg kg-1, Biopharm Drug Dispos, 13 (1992) 571–581. [DOI] [PubMed] [Google Scholar]
- [300].Rodvold KA, Blum RA, Fischer JH, Zokufa HZ, Rotschafer JC, Crossley KB, Riff LJ, Vancomycin pharmacokinetics in patients with various degrees of renal function, Antimicrob Agents Chemother, 32 (1988) 848–852. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [301].Rotschafer JC, Crossley K, Zaske DE, Mead K, Sawchuk RJ, Solem LD, Pharmacokinetics of vancomycin: Observations in 28 patients and dosage recommendations, Antimicrob Agents Chemother, 22 (1982) 391–394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [302].Tarral E, Jehl F, Tarral A, Simeoni U, Monteil H, Willard D, Geisert J, Pharmacokinetics of teicoplanin in children, J Antimicrob Chemother, 21 Suppl A (1988) 47–51. [DOI] [PubMed] [Google Scholar]
- [303].James A, Koren G, Milliken J, Soldin S, Prober C, Vancomycin pharmacokinetics and dose recommendations for preterm infants, Antimicrob Agents Chemother, 31 (1987) 52–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [304].Bailie GR, Neal D, Vancomycin ototoxicity and nephrotoxicity. A review, Med Toxicol Adverse Drug Exp, 3 (1988) 376–386. [DOI] [PubMed] [Google Scholar]
- [305].Aljutayli A, Marsot A, Nekka F, An update on population pharmacokinetic analyses of vancomycin, part I: In Adults, Clin Pharmacokinet, 59 (2020) 671–698. [DOI] [PubMed] [Google Scholar]
- [306].Tang Z, Guan J, Li J, Yu Y, Qian M, Cao J, Shuai W, Jiao Z, Determination of vancomycin exposure target and individualised dosing recommendations for neonates: Model-informed precision dosing, Int J Antimicrob Agents, (2021) 106300. [DOI] [PubMed] [Google Scholar]
- [307].Byrne CJ, Roberts JA, McWhinney B, Ryder SA, Fennell JP, O’Byrne P, Deasy E, Egan S, Desmond R, Enright H, D’Arcy DM, McHugh J, Population pharmacokinetics of teicoplanin and attainment of pharmacokinetic/pharmacodynamic targets in adult patients with haematological malignancy, Clin Microbiol Infect, 23 (2017) 674. [DOI] [PubMed] [Google Scholar]
- [308].Ogawa R, Kobayashi S, Sasaki Y, Makimura M, Echizen H, Population pharmacokinetic and pharmacodynamic analyses of teicoplanin in Japanese patients with systemic MRSA infection, Int J Clin Pharmacol Ther, 51 (2013) 357–366. [DOI] [PubMed] [Google Scholar]
- [309].Soy D, López E, Ribas J, Teicoplanin population pharmacokinetic analysis in hospitalized patients, Ther Drug Monit, 28 (2006) 737–743. [DOI] [PubMed] [Google Scholar]
- [310].Kanazawa N, Matsumoto K, Ikawa K, Fukamizu T, Shigemi A, Yaji K, Shimodozono Y, Morikawa N, Takeda Y, Yamada K, An initial dosing method for teicoplanin based on the area under the serum concentration time curve required for MRSA eradication, J Infect Chemother, 17 (2011) 297–300. [DOI] [PubMed] [Google Scholar]
- [311].Cazaubon Y, Venisse N, Mimoz O, Maire P, Ducher M, Bourguignon L, Goutelle S, Population pharmacokinetics of teicoplanin administered by subcutaneous or intravenous route and simulation of optimal loading dose regimen, J Antimicrob Chemother, 72 (2017) 2804–2812. [DOI] [PubMed] [Google Scholar]
- [312].Byrne CJ, Parton T, McWhinney B, Fennell JP, O’Byrne P, Deasy E, Egan S, Enright H, Desmond R, Ryder SA, D’Arcy DM, McHugh J, Roberts JA, Population pharmacokinetics of total and unbound teicoplanin concentrations and dosing simulations in patients with haematological malignancy, J Antimicrob Chemother, 73 (2018) 995–1003. [DOI] [PubMed] [Google Scholar]
- [313].Seki M, Yabuno K, Miyawaki K, Miwa Y, Tomono K, Loading regimen required to rapidly achieve therapeutic trough plasma concentration of teicoplanin and evaluation of clinical features, Clin Pharmacol, 4 (2012) 71–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [314].Matsumoto K, Kanazawa N, Watanabe E, Yokoyama Y, Fukamizu T, Shimodozono Y, Maeda C, Yasuda T, Kakihana Y, Ikawa K, Morikawa N, Takeda Y, Development of initial loading procedure for teicoplanin in critically ill patients with severe infections, Biol Pharm Bull, 36 (2013) 1024–1026. [DOI] [PubMed] [Google Scholar]
- [315].Wilson AP, Clinical pharmacokinetics of teicoplanin, Clin Pharmacokinet, 39 (2000) 167–183. [DOI] [PubMed] [Google Scholar]
- [316].Attwood RJ, LaPlante KL, Telavancin: A novel lipoglycopeptide antimicrobial agent, Am J Health Syst Pharm, 64 (2007) 2335–2348. [DOI] [PubMed] [Google Scholar]
- [317].Sun HK, Duchin K, Nightingale CH, Shaw JP, Seroogy J, Nicolau DP, Tissue penetration of telavancin after intravenous administration in healthy subjects, Antimicrob Agents Chemother, 50 (2006) 788–790. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [318].Lodise TP Jr., Gotfried M, Barriere S, Drusano GL, Telavancin penetration into human epithelial lining fluid determined by population pharmacokinetic modeling and Monte Carlo simulation, Antimicrob Agents Chemother, 52 (2008) 2300–2304. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [319].Gotfried MH, Shaw JP, Benton BM, Krause KM, Goldberg MR, Kitt MM, Barriere SL, Intrapulmonary distribution of intravenous telavancin in healthy subjects and effect of pulmonary surfactant on in vitro activities of telavancin and other antibiotics, Antimicrob Agents Chemother, 52 (2008) 92–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [320].Stucki A, Gerber P, Acosta F, Cottagnoud M, Cottagnoud P, Efficacy of telavancin against penicillin-resistant pneumococci and Staphylococcus aureus in a rabbit meningitis model and determination of kinetic parameters, Antimicrob Agents Chemother, 50 (2006) 770–773. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [321].Barcia-Macay M, Lemaire S, Mingeot-Leclercq MP, Tulkens PM, Van Bambeke F, Evaluation of the extracellular and intracellular activities (human THP-1 macrophages) of telavancin versus vancomycin against methicillin-susceptible, methicillin-resistant, vancomycin-intermediate and vancomycin-resistant Staphylococcus aureus, J Antimicrob Chemother, 58 (2006) 1177–1184. [DOI] [PubMed] [Google Scholar]
- [322].Duchin K, Shaw J, Spencer E, et al. Single dose pharmacokinetics (PK) of telavancin (TLV) in subjects with renal dysfunction [abstract no P1027]. Clin Microbiol Infect,10 (2004) 276. [Google Scholar]
- [323].Rubino CM, Van Wart SA, Bhavnani SM, Ambrose PG, McCollam JS, Forrest A, Oritavancin population pharmacokinetics in healthy subjects and patients with complicated skin and skin structure infections or bacteremia, Antimicrob Agents Chemother, 53 (2009) 4422–4428. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [324].Fetterly GJ, Ong CM, Bhavnani SM, Loutit JS, Porter SB, Morello LG, Ambrose PG, Nicolau DP, Pharmacokinetics of oritavancin in plasma and skin blister fluid following administration of a 200-milligram dose for 3 days or a single 800-milligram dose, Antimicrob Agents Chemother, 49 (2005) 148–152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [325].Cavaleri M, Riva S, Valagussa A, Guanci M, Colombo L, Dowell J, Stogniew M, Pharmacokinetics and excretion of dalbavancin in the rat, J Antimicrob Chemother, 55 Suppl 2 (2005) ii31–35. [DOI] [PubMed] [Google Scholar]
- [326].Nicolau DP, Sun HK, Seltzer E, Buckwalter M, Dowell JA, Pharmacokinetics of dalbavancin in plasma and skin blister fluid, J Antimicrob Chemother, 60 (2007) 681–684. [DOI] [PubMed] [Google Scholar]
- [327].Leighton A, Gottlieb AB, Dorr MB, Jabes D, Mosconi G, VanSaders C, Mroszczak EJ, Campbell KC, Kelly E, Tolerability, pharmacokinetics, and serum bactericidal activity of intravenous dalbavancin in healthy volunteers, Antimicrob Agents Chemother, 48 (2004) 940–945. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [328].Bhavnani SM, Owen JS, Loutit JS, Porter SB, Ambrose PG, Pharmacokinetics, safety, and tolerability of ascending single intravenous doses of oritavancin administered to healthy human subjects, Diagn Microbiol Infect Dis, 50 (2004) 95–102. [DOI] [PubMed] [Google Scholar]
- [329].Sun H, Maderazo EG, Krusell AR, Serum protein-binding characteristics of vancomycin, Antimicrob Agents Chemother, 37 (1993) 1132–1136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [330].Albrecht LM, Rybak MJ, Warbasse LH, Edwards DJ, Vancomycin protein binding in patients with infections caused by Staphylococcus aureus, DICP, 25 (1991) 713–715. [DOI] [PubMed] [Google Scholar]
- [331].Chen Y, Norris RL, Schneider JJ, Ravenscroft PJ, The influence of vancomycin concentration and the pH of plasma on vancomycin protein binding, J Pharmacol Toxicol Methods, 28 (1992) 57–60. [DOI] [PubMed] [Google Scholar]
- [332].Ackerman BH, Taylor EH, Olsen KM, Abdel-Malak W, Pappas AA, Vancomycin serum protein binding determination by ultrafiltration, Drug Intell Clin Pharm, 22 (1988) 300–303. [DOI] [PubMed] [Google Scholar]
- [333].Butterfield JM, Patel N, Pai MP, Rosano TG, Drusano GL, Lodise TP, Refining vancomycin protein binding estimates: Identification of clinical factors that influence protein binding, Antimicrob Agents Chemother, 55 (2011) 4277–4282. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [334].Kees MG, Wicha SG, Seefeld A, Kees F, Kloft C, Unbound fraction of vancomycin in intensive care unit patients, J Clin Pharmacol, 54 (2014) 318–323. [DOI] [PubMed] [Google Scholar]
- [335].Oyaert M, Spriet I, Allegaert K, Smits A, Vanstraelen K, Peersman N, Wauters J, Verhaegen J, Vermeersch P, Pauwels S, Factors impacting unbound vancomycin concentrations in different patient populations, Antimicrob Agents Chemother, 59 (2015) 7073–7079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [336].Assandri A, Bernareggi A, Binding of teicoplanin to human serum albumin, Eur J Clin Pharmacol, 33 (1987) 191–195. [DOI] [PubMed] [Google Scholar]
- [337].Bernareggi A, Borgonovi M, Del Favero A, Rosina R, GavanAghi L, Teicoplanin binding in plasma following administration of increasing intravenous doses to healthy volunteers, Eur J Drug Metab Pharmacokinet, Spec No 3 (1991) 256–260. [PubMed] [Google Scholar]
- [338].Dorr MB, Jabes D, Cavaleri M, Dowell J, Mosconi G, Malabarba A, White RJ, Henkel TJ, Human pharmacokinetics and rationale for once-weekly dosing of dalbavancin, a semi-synthetic glycopeptide, J Antimicrob Chemother, 55 Suppl 2 (2005) ii25–30. [DOI] [PubMed] [Google Scholar]
- [339].Stein GE, Wells EM, The importance of tissue penetration in achieving successful antimicrobial treatment of nosocomial pneumonia and complicated skin and soft-tissue infections caused by methicillin-resistant Staphylococcus aureus: Vancomycin and linezolid, Curr Med Res Opin, 26 (2010) 571–588. [DOI] [PubMed] [Google Scholar]
- [340].de Jesús Valle MJ, López FG, Hurlé AD, Navarro AS, Pulmonary versus systemic delivery of antibiotics: Comparison of vancomycin dispositions in the isolated rat lung, Antimicrob Agents Chemother, 51 (2007) 3771–3774. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [341].Skhirtladze K, Hutschala D, Fleck T, Thalhammer F, Ehrlich M, Vukovich T, Müller M, Tschernko EM, Impaired target site penetration of vancomycin in diabetic patients following cardiac surgery, Antimicrob Agents Chemother, 50 (2006) 1372–1375. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [342].Albanèse J, Léone M, Bruguerolle B, Ayem ML, Lacarelle B, Martin C, Cerebrospinal fluid penetration and pharmacokinetics of vancomycin administered by continuous infusion to mechanically ventilated patients in an intensive care unit, Antimicrob Agents Chemother, 44 (2000) 1356–1358. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [343].Blassmann U, Hope W, Roehr AC, Frey OR, Vetter-Kerkhoff C, Thon N, Briegel J, Huge V, CSF penetration of vancomycin in critical care patients with proven or suspected ventriculitis: A prospective observational study, J Antimicrob Chemother, 74 (2019) 991–996. [DOI] [PubMed] [Google Scholar]
- [344].Beach JE, Perrott J, Turgeon RD, Ensom MHH, Penetration of Vancomycin into the Cerebrospinal Fluid: A Systematic Review, Clin Pharmacokinet, 56 (2017) 1479–1490. [DOI] [PubMed] [Google Scholar]
- [345].Kalil AC, Metersky ML, Klompas M, Muscedere J, Sweeney DA, Palmer LB, Napolitano LM, O’Grady NP, Bartlett JG, Carratalà J, El Solh AA, Ewig S, Fey PD, File TM Jr., Restrepo MI, Roberts JA, Waterer GW, Cruse P, Knight SL, Brozek JL, Management of adults with hospital-acquired and ventilator-associated pneumonia: 2016 clinical practice guidelines by the Infectious Diseases Society of America and the American Thoracic Society, Clin Infect Dis, 63 (2016) e61–e111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [346].Lamer C, de Beco V, Soler P, Calvat S, Fagon JY, Dombret MC, Farinotti R, Chastre J, Gibert C, Analysis of vancomycin entry into pulmonary lining fluid by bronchoalveolar lavage in critically ill patients, Antimicrob Agents Chemother, 37 (1993) 281–286. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [347].Georges H, Leroy O, Alfandari S, Guery B, Roussel-Delvallez M, Dhennain C, Beaucaire G, Pulmonary disposition of vancomycin in critically ill patients, Eur J Clin Microbiol Infect Dis, 16 (1997) 385–388. [DOI] [PubMed] [Google Scholar]
- [348].Cruciani M, Gatti G, Lazzarini L, Furlan G, Broccali G, Malena M, Franchini C, Concia E, Penetration of vancomycin into human lung tissue, J Antimicrob Chemother, 38 (1996) 865–869. [DOI] [PubMed] [Google Scholar]
- [349].Lodise TP, Drusano GL, Butterfield JM, Scoville J, Gotfried M, Rodvold KA, Penetration of vancomycin into epithelial lining fluid in healthy volunteers, Antimicrob Agents Chemother, 55 (2011) 5507–5511. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [350].Rodvold K GM, Loutit JS, Porter SB, Plasma and intrapulmonary concentrations of oritavancin and vancomycin in normal healthy adults [abstract O257]. In: Program and abstracts of the 14th European Congress of Clinical Microbiology and Infectious Diseases (Prague). Basel: European Society of Clinical Microbiology and Infectious Diseases, (2004) 44. [Google Scholar]
- [351].Kiem S, Schentag JJ, Interpretation of antibiotic concentration ratios measured in epithelial lining fluid, Antimicrob Agents Chemother, 52 (2008) 24–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [352].Huang DB, File TM Jr., Torres A, Shorr AF, Wilcox MH, Hadvary P, Dryden M, Corey GR, A phase II randomized, double-blind, multicenter study to evaluate efficacy and safety of intravenous iclaprim versus vancomycin for the treatment of nosocomial pneumonia suspected or confirmed to be due to Gram-positive pathogens, Clin Ther, 39 (2017) 1706–1718. [DOI] [PubMed] [Google Scholar]
- [353].A study of AeroVanc for the treatment of MRSA infection in CF patients, https://clinicaltrials.gov/ct2/show/NCT03181932.
- [354].Waterer G, Lord J, Hofmann T, Jouhikainen T, Phase I, dose-escalating study of the safety and pharmacokinetics of inhaled dry-powder vancomycin (Aerovanc) in volunteers and patients with cystic fibrosis: a new approach to therapy for methicillin-resistant Staphylococcus aureus, Antimicrob Agents Chemother, 3 (2020) 64: e01776–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [355].Shoji H, Maeda M, Shirakura T, Takuma T, Ugajin K, Fukuchi K, Ishino K, Niki Y, More accurate measurement of vancomycin minimum inhibitory concentration indicates poor outcomes in meticillin-resistant Staphylococcus aureus bacteraemia, Int J Antimicrob Agents, 46 (2015) 532–537. [DOI] [PubMed] [Google Scholar]
- [356].Exner K, Lang E, Borsche A, Lemperle G, Efficacy, tolerability and pharmacokinetics of teicoplanin in patients undergoing breast surgery, Eur J Surg Suppl, (1992) 33–38. [PubMed] [Google Scholar]
- [357].Nehrer S, Thalhammer F, Schwameis E, Breyer S, Kotz R, Teicoplanin in the prevention of infection in total hip replacement, Arch Orthop Trauma Surg, 118 (1998) 32–36. [DOI] [PubMed] [Google Scholar]
- [358].Cruciani M, Navarra A, Di Perri G, Andreoni M, Danzi MC, Concia E, Bassetti D, Evaluation of intraventricular teicoplanin for the treatment of neurosurgical shunt infections, Clin Infect Dis, 15 (1992) 285–289. [DOI] [PubMed] [Google Scholar]
- [359].Mercier E, Darrouzain F, Montharu J, Guillon A, Diot P, Paintaud G, Vecellio L, Lung and serum teicoplanin concentration after aerosol and intravenous administration in a rat model, J Aerosol Med Pulm Drug Deliv, 27 (2014) 306–312. [DOI] [PubMed] [Google Scholar]
- [360].Mimoz O, Rolland D, Adoun M, Marchand S, Breilh D, Brumpt I, Debaene B, Couet W, Steady-state trough serum and epithelial lining fluid concentrations of teicoplanin 12 mg/kg per day in patients with ventilator-associated pneumonia, Intensive Care Med, 32 (2006) 775–779. [DOI] [PubMed] [Google Scholar]
- [361].Open-label, Dose-escalation Study to Evaluate the Pharmacokinetics of Inhaled Teicoplanin in Cystic Fibrosis Patients, https://clinicaltrials.gov/ct2/show/NCT04176328.
- [362].Goldschmidt E, Rasmussen J, Chabot JD, Gandhoke G, Luzzi E, Merlotti L, Proni R, Loresi M, Hamilton DK, Okonkwo DO, Kanter AS, Gerszten PC, The effect of vancomycin powder on human dural fibroblast culture and its implications for dural repair during spine surgery, J Neurosurg Spine, 25 (2016) 665–670. [DOI] [PubMed] [Google Scholar]
- [363].Nau R, Blei C, Eiffert H, Intrathecal antibacterial and antifungal therapies, Clin Microbiol Rev, 33 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [364].Zhang Q, Chen H, Zhu C, Chen F, Sun S, Liang N, Zheng W, Efficacy and safety of intrathecal meropenem and vancomycin in the treatment of postoperative intracranial infection in patients with severe traumatic brain injury, Exp Ther Med, 17 (2019) 4605–4609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [365].Moise-Broder PA, Forrest A, Birmingham MC, Schentag JJ, Pharmacodynamics of vancomycin and other antimicrobials in patients with Staphylococcus aureus lower respiratory tract infections, Clin Pharmacokinet, 43 (2004) 925–942. [DOI] [PubMed] [Google Scholar]
- [366].Kullar R, Davis SL, Levine DP, Rybak MJ, Impact of vancomycin exposure on outcomes in patients with methicillin-resistant Staphylococcus aureus bacteremia: Support for consensus guidelines suggested targets, Clin Infect Dis, 52 (2011) 975–981. [DOI] [PubMed] [Google Scholar]
- [367].Brown J, Brown K, Forrest A, Vancomycin AUC24/MIC ratio in patients with complicated bacteremia and infective endocarditis due to methicillin-resistant Staphylococcus aureus and its association with attributable mortality during hospitalization, Antimicrob Agents Chemother, 56 (2012) 634–638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [368].Holmes NE, Turnidge JD, Munckhof WJ, Robinson JO, Korman TM, O’Sullivan MV, Anderson TL, Roberts SA, Warren SJ, Gao W, Howden BP, Johnson PD, Vancomycin AUC/MIC ratio and 30-day mortality in patients with Staphylococcus aureus bacteremia, Antimicrob Agents Chemother, 57 (2013) 1654–1663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [369].Gawronski KM, Goff DA, Brown J, Khadem TM, Bauer KA, A stewardship program’s retrospective evaluation of vancomycin AUC24/MIC and time to microbiological clearance in patients with methicillin-resistant Staphylococcus aureus bacteremia and osteomyelitis, Clin Ther, 35 (2013) 772–779. [DOI] [PubMed] [Google Scholar]
- [370].Mogle BT, Steele JM, Seabury RW, Dang UJ, Kufel WD, Implementation of a two-point pharmacokinetic AUC-based vancomycin therapeutic drug monitoring approach in patients with methicillin-resistant Staphylococcus aureus bacteraemia, Int J Antimicrob Agents, 52 (2018) 805–810. [DOI] [PubMed] [Google Scholar]
- [371].Le J, Ny P, Capparelli E, Lane J, Ngu B, Muus R, Romanowski G, Vo T, Bradley J, Pharmacodynamic characteristics of nephrotoxicity associated with vancomycin use in children, J Pediatr Infect Dis Soc, 4 (2015) e109–116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [372].Rybak M, Lomaestro B, Rotschafer JC, Moellering R Jr., Craig W, Billeter M, Dalovisio JR, Levine DP, Therapeutic monitoring of vancomycin in adult patients: A consensus review of the American Society of Health-System Pharmacists, the Infectious Diseases Society of America, and the Society of Infectious Diseases Pharmacists, Am J Health Syst Pharm, 66 (2009) 82–98. [DOI] [PubMed] [Google Scholar]
- [373].Singh NB, Yim J, Jahanbakhsh S, Sakoulas G, Rybak MJ, Impact of cefazolin co-administration with vancomycin to reduce development of vancomycin-intermediate Staphylococcus aureus, Diagn Microbiol Infect Dis, 91 (2018) 363–370. [DOI] [PubMed] [Google Scholar]
- [374].Vidaillac C, Gardete S, Tewhey R, Sakoulas G, Kaatz GW, Rose WE, Tomasz A, Rybak MJ, Alternative mutational pathways to intermediate resistance to vancomycin in methicillin-resistant Staphylococcus aureus, J Infect Dis, 208 (2013) 67–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [375].Rybak MJ, Le J, Lodise TP, Levine DP, Bradley JS, Liu C, Mueller BA, Pai MP, Wong-Beringer A, Rotschafer JC, Rodvold KA, Maples HD, Lomaestro BM, Therapeutic monitoring of vancomycin for serious methicillin-resistant Staphylococcus aureus infections: A revised consensus guideline and review by the American Society of Health-System Pharmacists, the Infectious Diseases Society of America, the Pediatric Infectious Diseases Society, and the Society of Infectious Diseases Pharmacists, Am J Health Syst Pharm, 77 (2020) 835–864. [DOI] [PubMed] [Google Scholar]
- [376].Patel N, Pai MP, Rodvold KA, Lomaestro B, Drusano GL, Lodise TP, Vancomycin: We can’t get there from here, Clin Infect Dis, 52 (2011) 969–974. [DOI] [PubMed] [Google Scholar]
- [377].Heil EL, Claeys KC, Mynatt RP, Hopkins TL, Brade K, Watt I, Rybak MJ, Pogue JM, Making the change to area under the curve-based vancomycin dosing, Am J Health Syst Pharm, 75 (2018) 1986–1995. [DOI] [PubMed] [Google Scholar]
- [378].Pai MP, Neely M, Rodvold KA, Lodise TP, Innovative approaches to optimizing the delivery of vancomycin in individual patients, Adv Drug Deliv Rev, 77 (2014) 50–57. [DOI] [PubMed] [Google Scholar]
- [379].Pais GM, Avedissian SN, O’Donnell JN, Rhodes NJ, Lodise TP, Prozialeck WC, Lamar PC, Cluff C, Gulati A, Fitzgerald JC, Downes KJ, Zuppa AF, Scheetz MH, Comparative performance of urinary biomarkers for vancomycin-induced kidney injury according to timeline of injury, Antimicrob Agents Chemother, 7 (2019) 63: e00079–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [380].O’Donnell JN, Rhodes NJ, Lodise TP, Prozialeck WC, Miglis CM, Joshi MD, Venkatesan N, Pais G, Cluff C, Lamar PC, Briyal S, Day JZ, Gulati A, Scheetz MH, 24-Hour pharmacokinetic relationships for vancomycin and novel urinary biomarkers of acute kidney injury, Antimicrob Agents Chemother, 11 (2017) 61: e00416–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [381].Avedissian SN, Pais G, Liu J, O’Donnell JN, Lodise TP, Neely M, Prozialeck WC, Lamar PC, Becher L, Scheetz MH, The pharmacodynamic-toxicodynamic relationship of AUC and C (max) in vancomycin-induced kidney injury in an animal model, Antimicrob Agents Chemother, 3 (2021) 65: e01945–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [382].Zhu LL, Zhou Q, Optimal infusion rate in antimicrobial therapy explosion of evidence in the last five years, Infect Drug Resist, 11 (2018) 1105–1117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [383].Álvarez R, López Cortés LE, Molina J, Cisneros JM, Pachón J, Optimizing the Clinical Use of Vancomycin, Antimicrob Agents Chemother, 60 (2016) 2601–2609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [384].Girand HL, Continuous Infusion Vancomycin in Pediatric Patients: A Critical Review of the Evidence, J Pediatr Pharmacol Ther, 25 (2020) 198–214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [385].King A, Phillips I, Kaniga K, Comparative in vitro activity of telavancin (TD-6424), a rapidly bactericidal, concentration-dependent anti-infective with multiple mechanisms of action against Gram-positive bacteria, J Antimicrob Chemother, 53 (2004) 797–803. [DOI] [PubMed] [Google Scholar]
- [386].Odenholt I, Löwdin E, Cars O, Pharmacodynamic effects of telavancin against methicillin-resistant and methicillin-susceptible Staphylococcus aureus strains in the presence of human albumin or serum and in an in vitro kinetic model, Antimicrob Agents Chemother, 51 (2007) 3311–3316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [387].MacGowan AP, Noel AR, Tomaselli S, Elliott HC, Bowker KE, Pharmacodynamics of telavancin studied in an in vitro pharmacokinetic model of infection, Antimicrob Agents Chemother, 55 (2011) 867–873. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [388].Lepak AJ, Zhao M, Andes DR, Comparative pharmacodynamics of telavancin and vancomycin in the neutropenic murine thigh and lung infection models against Staphylococcus aureus, Antimicrob Agents Chemother, 61 (2017): e00281–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [389].Lepak A, Marchillo K, VanHecker J, Andes D, Impact of glycopeptide resistance in Staphylococcus aureus on the dalbavancin in vivo pharmacodynamic target, Antimicrob Agents Chemother, 59 (2015) 7833–7836. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [390].Boylan CJ, Campanale K, Iversen PW, Phillips DL, Zeckel ML, Parr TR Jr., Pharmacodynamics of oritavancin (LY333328) in a neutropenic-mouse thigh model of Staphylococcus aureus infection, Antimicrob Agents Chemother, 47 (2003) 1700–1706. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [391].Okusanya OO, Lehoux D, van Wart S, Rafai Far A, Forrest A, Moeck G, et al. , editors. Pharmacokinetics (PK) and pharmacokinetics–pharmacodynamics (PK–PD) of oritavancin (ORI) against Staphylococcus aureus (SA) in a neutropenic murine thigh-infection model. 49th interscience conference on Antimicrob Agents Chemother; 2009. San Francisco, CA. [Google Scholar]
- [392].Bhavnani SM, Passarell JA, Owen JS, Loutit JS, Porter SB, Ambrose PG, Pharmacokinetic-pharmacodynamic relationships describing the efficacy of oritavancin in patients with Staphylococcus aureus bacteremia, Antimicrob Agents Chemother, 50 (2006) 994–1000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [393].Bruniera FR, Ferreira FM, Saviolli LR, Bacci MR, Feder D, da Luz Gonçalves Pedreira M, Sorgini Peterlini MA, Azzalis LA, Campos Junqueira VB, Fonseca FL, The use of vancomycin with its therapeutic and adverse effects: A review, Eur Rev Med Pharmacol Sci, 19 (2015) 694–700. [PubMed] [Google Scholar]
- [394].Harding I, Sorgel F, Comparative pharmacokinetics of teicoplanin and vancomycin, J Chemother, 12 Suppl 5 (2000) 15–20. [DOI] [PubMed] [Google Scholar]
- [395].Sin Yan Chui CO, Turner Sean, Coulthard Kingsley, McKinnon Ross, Incidence of vancomycin-induced red man syndrome in a women’s and children’s hospital, J Pharm Pract Res, 37 (2007) 124–126. [Google Scholar]
- [396].Polk RE, Anaphylactoid reactions to glycopeptide antibiotics, J Antimicrob Chemother, 27 Suppl B (1991) 17–29. [DOI] [PubMed] [Google Scholar]
- [397].Sidi V, Roilides E, Bibashi E, Gompakis N, Tsakiri A, Koliouskas D, Comparison of efficacy and safety of teicoplanin and vancomycin in children with antineoplastic therapy-associated febrile neutropenia and Gram-positive bacteremia, J Chemother, 12 (2000) 326–331. [DOI] [PubMed] [Google Scholar]
- [398].Svetitsky S, Leibovici L, Paul M, Comparative efficacy and safety of vancomycin versus teicoplanin: Systematic review and meta-analysis, Antimicrob Agents Chemother, 53 (2009) 4069–4079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [399].Wood MJ, Comparative safety of teicoplanin and vancomycin, J Chemother, 12 Suppl 5 (2000) 21–25. [DOI] [PubMed] [Google Scholar]
- [400].Maher ER, Hollman A, Grüneberg RN, Teicoplanin-induced ototoxicity in Down’s syndrome, Lancet, 1 (1986) 613. [DOI] [PubMed] [Google Scholar]
- [401].Bibler MR, Frame PT, Hagler DN, Bode RB, Staneck JL, Thamlikitkul V, Harris JE, Haregewoin A, Bullock WE Jr., Clinical evaluation of efficacy, pharmacokinetics, and safety of teicoplanin for serious Gram-positive infections, Antimicrob Agents Chemother, 31 (1987) 207–212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [402].Lanvers-Kaminsky C, Zehnhoff-Dinnesen AA, Parfitt R, Ciarimboli G, Drug-induced ototoxicity: Mechanisms, Pharmacogenetics, and protective strategies, Clin Pharmacol Ther, 101 (2017) 491–500. [DOI] [PubMed] [Google Scholar]
- [403].Martin JH, Norris R, Barras M, Roberts J, Morris R, Doogue M, Jones GR, Therapeutic monitoring of vancomycin in adult patients: A consensus review of the American Society of Health-System Pharmacists, the Infectious Diseases Society of America, and the Society of Infectious Diseases Pharmacists, Clin Biochem Rev, 31 (2010) 21–24. [PMC free article] [PubMed] [Google Scholar]
- [404].Humphrey C, Veve MP, Walker B, Shorman MA, Long-term vancomycin use had low risk of ototoxicity, PLoS One, 14 (2019) e0224561. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [405].Jeffres MN, The whole price of vancomycin: Toxicities, troughs, and time, drugs, 77 (2017) 1143–1154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [406].Bugano DD, Cavalcanti AB, Goncalves AR, Almeida CS, Silva E, Cochrane meta-analysis: Teicoplanin versus vancomycin for proven or suspected infection, Einstein (Sao Paulo), 9 (2011) 265–282. [DOI] [PubMed] [Google Scholar]
- [407].Marre R, Schulz E, Anders T, Sack K, Renal tolerance and pharmacokinetics of vancomycin in rats, J Antimicrob Chemother, 14 (1984) 253–260. [DOI] [PubMed] [Google Scholar]
- [408].Vaidya VS, Ozer JS, Dieterle F, Collings FB, Ramirez V, Troth S, Muniappa N, Thudium D, Gerhold D, Holder DJ, Bobadilla NA, Marrer E, Perentes E, Cordier A, Vonderscher J, Maurer G, Goering PL, Sistare FD, Bonventre JV, Kidney injury molecule-1 outperforms traditional biomarkers of kidney injury in preclinical biomarker qualification studies, Nat Biotechnol, 28 (2010) 478–485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [409].Fuchs TC, Frick K, Emde B, Czasch S, von Landenberg F, Hewitt P, Evaluation of novel acute urinary rat kidney toxicity biomarker for subacute toxicity studies in preclinical trials, Toxicol Pathol, 40 (2012) 1031–1048. [DOI] [PubMed] [Google Scholar]
- [410].Nishino Y, Takemura S, Minamiyama Y, Hirohashi K, Ogino T, Inoue M, Okada S, Kinoshita H, Targeting superoxide dismutase to renal proximal tubule cells attenuates vancomycin-induced nephrotoxicity in rats, Free Radic Res, 37 (2003) 373–379. [DOI] [PubMed] [Google Scholar]
- [411].Oktem F, Arslan MK, Ozguner F, Candir O, Yilmaz HR, Ciris M, Uz E, In vivo evidences suggesting the role of oxidative stress in pathogenesis of vancomycin-induced nephrotoxicity: Protection by erdosteine, Toxicology, 215 (2005) 227–233. [DOI] [PubMed] [Google Scholar]
- [412].King DW, Smith MA, Proliferative responses observed following vancomycin treatment in renal proximal tubule epithelial cells, Toxicol In Vitro, 18 (2004) 797–803. [DOI] [PubMed] [Google Scholar]
- [413].van Hal SJ, Paterson DL, Lodise TP, Systematic review and meta-analysis of vancomycin-induced nephrotoxicity associated with dosing schedules that maintain troughs between 15 and 20 milligrams per liter, Antimicrob Agents Chemother, 57 (2013) 734–744. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [414].Sinha Ray A, Haikal A, Hammoud KA, Yu AS, Vancomycin and the risk of AKI: A systematic review and meta-analysis, Clin J Am Soc Nephrol, 11 (2016) 2132–2140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [415].Cosgrove SE, Vigliani GA, Fowler VG Jr., Abrutyn E, Corey GR, Levine DP, Rupp ME, Chambers HF, Karchmer AW, Boucher HW, Initial low-dose gentamicin for Staphylococcus aureus bacteremia and endocarditis is nephrotoxic, Clin Infect Dis, 48 (2009) 713–721. [DOI] [PubMed] [Google Scholar]
- [416].Carreno JJ, Kenney RM, Divine G, Vazquez JA, Davis SL, Randomized controlled trial to determine the efficacy of early switch from vancomycin to vancomycin alternatives as a strategy to prevent nephrotoxicity in patients with multiple risk factors for adverse renal outcomes (STOP-NT), Ann Pharmacother, 51 (2017) 185–193. [DOI] [PubMed] [Google Scholar]
- [417].Aljefri DM, Avedissian SN, Rhodes NJ, Postelnick MJ, Nguyen K, Scheetz MH, Vancomycin area under the curve and acute kidney injury: A meta-analysis, Clin Infect Dis, 69 (2019) 1881–1887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [418].Neely MN, Youn G, Jones B, Jelliffe RW, Drusano GL, Rodvold KA, Lodise TP, Are vancomycin trough concentrations adequate for optimal dosing, Antimicrob Agents Chemother, 58 (2014) 309–316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [419].Suzuki Y, Kawasaki K, Sato Y, Tokimatsu I, Itoh H, Hiramatsu K, Takeyama M, Kadota J, Is peak concentration needed in therapeutic drug monitoring of vancomycin? A pharmacokinetic-pharmacodynamic analysis in patients with methicillin-resistant Staphylococcus aureus pneumonia, Chemotherapy, 58 (2012) 308–312. [DOI] [PubMed] [Google Scholar]
- [420].Bamgbola O, Review of vancomycin-induced renal toxicity: An update, Ther Adv Endocrinol Metab, 7 (2016) 136–147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [421].Pryka RD, Vancomycin serum concentration monitoring: A continued debate, Ann Pharmacother, 28 (1994) 1397–1399. [DOI] [PubMed] [Google Scholar]
- [422].Kufel WD, Seabury RW, Mogle BT, Beccari MV, Probst LA, Steele JM, Readiness to implement vancomycin monitoring based on area under the concentration-time curve: A cross-sectional survey of a national health consortium, Am J Health Syst Pharm, 76 (2019) 889–894. [DOI] [PubMed] [Google Scholar]
- [423].He N, Su S, Ye Z, Du G, He B, Li D, Liu Y, Yang K, Zhang X, Zhang Y, Chen X, Chen Y, Chen Z, Dong Y, Du G, Gu J, Guo D, Guo R, Hu X, Jiao Z, Li H, Liu G, Li Z, Lv Y, Lu W, Miao L, Qu J, Sun T, Tong R, Wang L, Wang M, Wang R, Wen A, Wu J, Wu X, Xu Y, Yang Y, Yang F, Zhan S, Zhang B, Zhang C, Zhang H, Zhang J, Zhang J, Zhang J, Zhang W, Zhao L, Zhao L, Zhao R, Zhao W, Zhao Z, Zhou W, Zeng XT, Zhai S, Evidence-based guideline for therapeutic drug monitoring of vancomycin: 2020 update by the division of therapeutic drug monitoring, Chinese Pharmacological Society, Clin Infect Dis, 71 (2020) S363–S371. [DOI] [PubMed] [Google Scholar]
- [424].Dilworth TJ, Schulz LT, Rose WE, Vancomycin advanced therapeutic drug monitoring: An exercise in futility or virtuous endeavor to improve drug efficacy and safety, Clin Infect Dis, (2020). [DOI] [PubMed] [Google Scholar]
- [425].Crass RL, Rodvold KA, Mueller BA, Pai MP, Renal dosing of antibiotics: Are we jumping the gun, Clin Infect Dis, 68 (2019) 1596–1602. [DOI] [PubMed] [Google Scholar]
- [426].Lodise TP, Lomaestro B, Graves J, Drusano GL, Larger vancomycin doses (at least four grams per day) are associated with an increased incidence of nephrotoxicity, Antimicrob Agents Chemother, 52 (2008) 1330–1336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [427].Pai MP, Drug dosing based on weight and body surface area: Mathematical assumptions and limitations in obese adults, Pharmacotherapy, 32 (2012) 856–868. [DOI] [PubMed] [Google Scholar]
- [428].Khare M, Azim A, Kneese G, Haag M, Weinstein K, Rhee KE, Foster BA, Vancomycin dosing in children with overweight or obesity: A systematic review and meta-analysis, Hosp Pediatr, 10 (2020) 359–368. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [429].Goulenok T, Fantin B, Antimicrobial treatment of febrile neutropenia: Pharmacokinetic-pharmacodynamic considerations, Clin Pharmacokinet, 52 (2013) 869–883. [DOI] [PubMed] [Google Scholar]
- [430].Lortholary O, Lefort A, Tod M, Chomat AM, Darras-Joly C, Cordonnier C, Pharmacodynamics and pharmacokinetics of antibacterial drugs in the management of febrile neutropenia, Lancet Infect Dis, 8 (2008) 612–620. [DOI] [PubMed] [Google Scholar]
- [431].He N, Dong F, Liu W, Zhai S, A systematic review of vancomycin dosing in patients with hematologic malignancies or neutropenia, Infect Drug Resist, 13 (2020) 1807–1821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [432].Haeseker MB, Croes S, Neef C, Bruggeman CA, Stolk LM, Verbon A, Vancomycin dosing in neutropenic patients, PLoS One, 9 (2014) e112008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [433].Hirai K, Ishii H, Shimoshikiryo T, Shimomura T, Tsuji D, Inoue K, Kadoiri T, Itoh K, Augmented renal clearance in patients with febrile neutropenia is associated with increased risk for subtherapeutic concentrations of vancomycin, Ther Drug Monit, 38 (2016) 706–710. [DOI] [PubMed] [Google Scholar]
- [434].Fernández de Gatta MD, Calvo MV, Hernández JM, Caballero D, San Miguel JF, Domínguez-Gil A, Cost-effectiveness analysis of serum vancomycin concentration monitoring in patients with hematologic malignancies, Clin Pharmacol Ther, 60 (1996) 332–340. [DOI] [PubMed] [Google Scholar]
- [435].Hahn A, Frenck RW Jr., Allen-Staat M, Zou Y, Vinks AA, Evaluation of target attainment of vancomycin area under the curve in children with methicillin-resistant Staphylococcus aureus bacteremia, Ther Drug Monit, 37 (2015) 619–625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [436].Buckel WR, Ghobrial S, Tamma PD, Milstone AM, Zhao Y, Hsu AJ, Risk factors for non-therapeutic initial steady-state vancomycin trough concentrations in children and adolescents receiving high empiric doses of intravenous vancomycin, Paediatr Drugs, 19 (2017) 43–51. [DOI] [PubMed] [Google Scholar]
- [437].McNeil JC, Kaplan SL, Vancomycin therapeutic drug monitoring in children: New recommendations, similar challenges, J Pediatr Pharmacol Ther, 25 (2020) 472–475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [438].Shen K, Fan Y, Yang M, Chen Y, Tao J, Lu G, Zhang H, Huang Q, Zhang J, Modeling approach to optimizing dose regimen of vancomycin for Chinese pediatric patients with Gram-positive bacterial infections, Frontiers in pharmacology, 12 (2021) 648668. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [439].Cheong JY, Makmor-Bakry M, Lau CL, Abdul Rahman R, The relationship between trough concentration of vancomycin and effect on methicillin-resistant Staphylococcus aureus in critically ill patients, S Afr Med J, 102 (2012) 616–619. [DOI] [PubMed] [Google Scholar]
- [440].Wesolek JL, McNorton K, Delgado G Jr., Giuliano CA, Effect of vancomycin initial dosing on time to systemic inflammatory response syndrome resolution in patients with methicillin-resistant Staphylococcus aureus bacteremia, J Chemother, 30 (2018) 101–106. [DOI] [PubMed] [Google Scholar]
- [441].USP. Vancomycin Injection. FDA 2017 NDA 050671; 2017. https://www.accessdata.fda.gov/drugsatfda_docs/label/2017/050671s024lbl.pdf.
- [442].Matsumoto K, Takesue Y, Ohmagari N, Mochizuki T, Mikamo H, Seki M, Takakura S, Tokimatsu I, Takahashi Y, Kasahara K, Okada K, Igarashi M, Kobayashi M, Hamada Y, Kimura M, Nishi Y, Tanigawara Y, Kimura T, Practice guidelines for therapeutic drug monitoring of vancomycin: A consensus review of the Japanese Society of Chemotherapy and the Japanese Society of Therapeutic Drug Monitoring, J Infect Chemother, 19 (2013) 365–380. [DOI] [PubMed] [Google Scholar]
- [443].Mei H, Wang J, Che H, Wang R, Cai Y, The clinical efficacy and safety of vancomycin loading dose: A systematic review and meta-analysis, Medicine (Baltimore), 98 (2019) e17639. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [444].Watanabe E, Matsumoto K, Ikawa K, Yokoyama Y, Shigemi A, Enoki Y, Umezaki Y, Nakamura K, Ueno K, Terazono H, Morikawa N, Takeda Y, Pharmacokinetic/pharmacodynamic evaluation of teicoplanin against Staphylococcus aureus in a murine thigh infection model, J Glob Antimicrob Resist, 24 (2020) 83–87. [DOI] [PubMed] [Google Scholar]
- [445].Ramos-Martín V, Johnson A, McEntee L, Farrington N, Padmore K, Cojutti P, Pea F, Neely MN, Hope WW, Pharmacodynamics of teicoplanin against MRSA, J Antimicrob Chemother, 72 (2017) 3382–3389. [DOI] [PubMed] [Google Scholar]
- [446].Matsumoto K, Watanabe E, Kanazawa N, Fukamizu T, Shigemi A, Yokoyama Y, Ikawa K, Morikawa N, Takeda Y, Pharmacokinetic/pharmacodynamic analysis of teicoplanin in patients with MRSA infections, Clin Pharmacol, 8 (2016) 15–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [447].Hagihara M, Umemura T, Kimura M, Mori T, Hasegawa T, Mikamo H, Exploration of optimal teicoplanin dosage based on pharmacokinetic parameters for the treatment of intensive care unit patients infected with methicillin-resistant Staphylococcus aureus, J Infect Chemother, 18 (2012) 10–16. [DOI] [PubMed] [Google Scholar]
- [448].Kuti JL, Kiffer CR, Mendes CM, Nicolau DP, Pharmacodynamic comparison of linezolid, teicoplanin and vancomycin against clinical isolates of Staphylococcus aureus and coagulase-negative staphylococci collected from hospitals in Brazil, Clin Microbiol Infect, 14 (2008) 116–123. [DOI] [PubMed] [Google Scholar]
- [449].Harding I, MacGowan AP, White LO, Darley ES, Reed V, Teicoplanin therapy for Staphylococcus aureus septicaemia: Relationship between pre-dose serum concentrations and outcome, J Antimicrob Chemother, 45 (2000) 835–841. [DOI] [PubMed] [Google Scholar]
- [450].Ahn BJ, Yim DS, Lee DG, Kwon JC, Kim SH, Choi SM, Teicoplanin dosing strategy for treatment of Staphylococcus aureus in Korean patients with neutropenic fever, Yonsei Med J, 52 (2011) 616–623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [451].Zhang T, Sun D, Shu Z, Duan Z, Liu Y, Du Q, Zhang Y, Dong Y, Wang T, Hu S, Cheng H, Dong Y, Population pharmacokinetics and model-based dosing optimization of teicoplanin in pediatric patients, Front Pharmacol, 11 (2020) 594562. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [452].Yamada T, Kubota T, Yonezawa M, Nishio H, Kanno S, Yano T, Kobayashi D, Egashira N, Takada H, Hara T, Masuda S, Evaluation of teicoplanin trough values after the recommended loading dose in children with associated safety analysis, Pediatr Infect Dis J, 36 (2017) 398–400. [DOI] [PubMed] [Google Scholar]
- [453].Sun D, Zhang T, Mi J, Dong Y, Liu Y, Zhang Y, Zhang D, Wang T, Cheng H, Dong Y, Therapeutic drug monitoring and nephrotoxicity of teicoplanin therapy in Chinese children: A retrospective study, Infect Drug Resist, 13 (2020) 4105–4113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [454].Ramos-Martín V, Neely MN, Padmore K, Peak M, Beresford MW, Turner MA, Paulus S, López-Herce J, Hope WW, Tools for the individualized therapy of teicoplanin for neonates and children, Antimicrob Agents and Chemother, 61 (2017) e00707–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [455].Huber FM, Pieper RL, Tietz AJ, The formation of daptomycin by supplying decanoic acid to streptomyces-roseosporus cultures producing the antibiotic complex a21978c, J Biotechnol, 7 (1988) 283–292. [Google Scholar]
- [456].Manfredi R, Sabbatani S, Novel pharmaceutical molecules against emerging resistant Gram-positive cocci, Braz J Infect Dis, 14 (2010) 96–108. [DOI] [PubMed] [Google Scholar]
- [457].Bradley J, Glasser C, Patino H, Arnold SR, Arrieta A, Congeni B, Daum RS, Kojaoghlanian T, Yoon M, Anastasiou D, Wolf DJ, Bokesch P, Daptomycin for complicated skin infections: A randomized trial, Pediatrics, 139 (2017) e20162477. [DOI] [PubMed] [Google Scholar]
- [458].Tally FP, DeBruin MF, Development of daptomycin for Gram-positive infections, J Antimicrob Chemother, 46 (2000) 523–526. [DOI] [PubMed] [Google Scholar]
- [459].a.f.C.P. Merck Sharp & Dohme Corp., LLC, Cubicin, Daptomycin for injection. Application Number: NDA 21–572/S-053 https://www.accessdata.fda.gov/drugsatfda_docs/nda/2016/021572Orig1s053.pdf.
- [460].Debono M, Barnhart M, Carrell CB, Hoffmann JA, Occolowitz JL, Abbott BJ, Fukuda DS, Hamill RL, Biemann K, Herlihy WC, A21978C, a complex of new acidic peptide antibiotics: Isolation, chemistry, and mass spectral structure elucidation, J Antibiot (Tokyo), 40 (1987) 761–777. [DOI] [PubMed] [Google Scholar]
- [461].Allen NE, Hobbs JN, Alborn WE Jr., Inhibition of peptidoglycan biosynthesis in Gram-positive bacteria by LY146032, Antimicrob Agents Chemother, 31 (1987) 1093–1099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [462].Debono M, Abbott BJ, Molloy RM, Fukuda DS, Hunt AH, Daupert VM, Counter FT, Ott JL, Carrell CB, Howard LC, et al. , Enzymatic and chemical modifications of lipopeptide antibiotic A21978C: The synthesis and evaluation of daptomycin (LY146032), J Antibiot (Tokyo), 41 (1988) 1093–1105. [DOI] [PubMed] [Google Scholar]
- [463].Cottagnoud P, Daptomycin: A new treatment for insidious infections due to Gram-positive pathogens, Swiss Med Wkly, 138 (2008) 93–99. [DOI] [PubMed] [Google Scholar]
- [464].Vilhena C, Bettencourt A, Daptomycin: A review of properties, clinical use, drug delivery and resistance, Mini Rev Med Chem, 12 (2012) 202–209. [DOI] [PubMed] [Google Scholar]
- [465].Hachmann AB, Sevim E, Gaballa A, Popham DL, Antelmann H, Helmann JD, Reduction in membrane phosphatidylglycerol content leads to daptomycin resistance in Bacillus subtilis, Antimicrob Agents Chemother, 55 (2011) 4326–4337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [466].Müller A, Wenzel M, Strahl H, Grein F, Saaki TNV, Kohl B, Siersma T, Bandow JE, Sahl HG, Schneider T, Hamoen LW, Daptomycin inhibits cell envelope synthesis by interfering with fluid membrane microdomains, Proc Natl Acad Sci U S A, 113 (2016) E7077–e7086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [467].Baltz RH, Daptomycin: Mechanisms of action and resistance, and biosynthetic engineering, Curr Opin Chem Biol, 13 (2009) 144–151. [DOI] [PubMed] [Google Scholar]
- [468].Silverman JA, Perlmutter NG, Shapiro HM, Correlation of daptomycin bactericidal activity and membrane depolarization in Staphylococcus aureus, Antimicrob Agents Chemother, 47 (2003) 2538–2544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [469].Taylor R, Butt K, Scott B, Zhang T, Muraih JK, Mintzer E, Taylor S, Palmer M, Two successive calcium-dependent transitions mediate membrane binding and oligomerization of daptomycin and the related antibiotic A54145, Biochim Biophys Acta, 1858 (2016) 1999–2005. [DOI] [PubMed] [Google Scholar]
- [470].Grein F, Müller A, Scherer KM, Liu X, Ludwig KC, Klöckner A, Strach M, Sahl HG, Kubitscheck U, Schneider T, Ca(2+)-Daptomycin targets cell wall biosynthesis by forming a tripartite complex with undecaprenyl-coupled intermediates and membrane lipids, Nat Commun, 11 (2020) 1455. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [471].Algorri M, Wong-Beringer A, Antibiotics differentially modulate lipoteichoic acid-mediated host immune response, Antibiotics (Basel), 9 (2020) 573. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [472].Hayden MK, Rezai K, Hayes RA, Lolans K, Quinn JP, Weinstein RA, Development of Daptomycin resistance in vivo in methicillin-resistant Staphylococcus aureus, J Clin Microbiol, 43 (2005) 5285–5287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [473].Ye Y, Xia Z, Zhang D, Sheng Z, Zhang P, Zhu H, Xu N, Liang S, Multifunctional Pharmaceutical Effects of the Antibiotic Daptomycin, Biomed Res Int, 2019 (2019) 8609218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [474].Ledger EVK, Pader V, Edwards AM, Enterococcus faecalis and pathogenic streptococci inactivate daptomycin by releasing phospholipids, Microbiology (Reading), 163 (2017) 1502–1508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [475].Pader V, Hakim S, Painter KL, Wigneshweraraj S, Clarke TB, Edwards AM, Staphylococcus aureus inactivates daptomycin by releasing membrane phospholipids, Nat Microbiol, 2 (2016) 16194. [DOI] [PubMed] [Google Scholar]
- [476].Mishra NN, Tran TT, Seepersaud R, Garcia-de-la-Maria C, Faull K, Yoon A, Proctor R, Miro JM, Rybak MJ, Bayer AS, Arias CA, Sullam PM, Perturbations of phosphatidate cytidylyltransferase (CdsA) mediate daptomycin resistance in Streptococcus mitis/oralis by a novel mechanism, Antimicrob Agents Chemother, 61 (2017) e02435–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [477].Heidary M, Khosravi AD, Khoshnood S, Nasiri MJ, Soleimani S, Goudarzi M, Daptomycin J Antimicrob Chemother, 73 (2018) 1–11. [DOI] [PubMed] [Google Scholar]
- [478].Safdar N, Andes D, Craig WA, In vivo pharmacodynamic activity of daptomycin, Antimicrob Agents Chemother, 48 (2004) 63–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [479].Xu X, Khadzhynov D, Peters H, Chaves RL, Hamed K, Levi M, Corti N, Population pharmacokinetics of daptomycin in adult patients undergoing continuous renal replacement therapy, Br J Clin Pharmacol, 83 (2017) 498–509. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [480].Rybak MJ, The efficacy and safety of daptomycin: First in a new class of antibiotics for Gram-positive bacteria, Clinical microbiology and infection: The official publication of the European Society of Clinical Microbiology and Infectious Diseases, 12 Suppl 1 (2006) 24–32. [DOI] [PubMed] [Google Scholar]
- [481].Bayer AS, Xiong YQ, Redeploying β-Lactams against Staphylococcus aureus: Repurposing with a purpose, J Infect Dis, 215 (2017) 11–13. [DOI] [PubMed] [Google Scholar]
- [482].Osorio C, Garzón L, Jaimes D, Silva E, Bustos RH, Impact on antibiotic resistance, therapeutic success, and control of side effects in therapeutic drug monitoring (TDM) of daptomycin: a scoping review, Antibiotics (Basel, Switz), 10 (2021) 263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [483].Cojutti PG, Candoni A, Ramos-Martin V, Lazzarotto D, Zannier ME, Fanin R, Hope W, Pea F, Population pharmacokinetics and dosing considerations for the use of daptomycin in adult patients with haematological malignancies, J Antimicrob Chemother, 72 (2017) 2342–2350. [DOI] [PubMed] [Google Scholar]
- [484].Galar A, Muñoz P, Valerio M, Cercenado E, García-González X, Burillo A, Sánchez-Somolinos M, Juárez M, Verde E, Bouza E, Current use of daptomycin and systematic therapeutic drug monitoring: Clinical experience in a tertiary care institution, Int J Antimicrob Agents, 53 (2019) 40–48. [DOI] [PubMed] [Google Scholar]
- [485].Karageorgos SA, Miligkos M, Dakoutrou M, Tsioutis C, Clinical effectiveness, safety profile, and pharmacokinetics of daptomycin in pediatric patients: A systematic review, J Pediatr Infect Dis Soc, 5 (2016) 446–457. [DOI] [PubMed] [Google Scholar]
- [486].Louie A, Kaw P, Liu W, Jumbe N, Miller MH, Drusano GL, Pharmacodynamics of daptomycin in a murine thigh model of Staphylococcus aureus infection, Antimicrob Agents Chemother, 45 (2001) 845–851. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [487].Abdul-Aziz MH, Alffenaar JC, Bassetti M, Bracht H, Dimopoulos G, Marriott D, Neely MN, Paiva JA, Pea F, Sjovall F, Timsit JF, Udy AA, Wicha SG, Zeitlinger M, De Waele JJ, Roberts JA, Antimicrobial therapeutic drug monitoring in critically ill adult patients: A position paper, Intensive Care Med, 46 (2020) 1127–1153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [488].Jager NG, van Hest RM, Lipman J, Taccone FS, Roberts JA, Therapeutic drug monitoring of anti-infective agents in critically ill patients, Expert Rev Clin Phar, 9 (2016) 961–979. [DOI] [PubMed] [Google Scholar]
- [489].Miyadera Y, Naito T, Yamada T, Kawakami J, Simple LC-MS/MS methods using core-shell octadecylsilyl microparticulate for the quantitation of total and free daptomycin in human plasma, Ther Drug Monit, 40 (2018) 589–595. [DOI] [PubMed] [Google Scholar]
- [490].Falcone M, Russo A, Cassetta MI, Lappa A, Tritapepe L, d’Ettorre G, Fallani S, Novelli A, Venditti M, Variability of pharmacokinetic parameters in patients receiving different dosages of daptomycin: Is therapeutic drug monitoring necessary?, J Infect Chemother, 19 (2013) 732–739. [DOI] [PubMed] [Google Scholar]
- [491].D’Avolio A, Pensi D, Baietto L, Pacini G, Di Perri G, De Rosa FG, Daptomycin pharmacokinetics and pharmacodynamics in septic and critically ill patients, Drugs, 76 (2016) 1161–1174. [DOI] [PubMed] [Google Scholar]
- [492].Reiber C, Senn O, Müller D, Kullak-Ublick GA, Corti N, Therapeutic drug monitoring of daptomycin: A retrospective monocentric analysis, Ther Drug Monit, 37 (2015) 634–640. [DOI] [PubMed] [Google Scholar]
- [493].Soraluce A, Asín-Prieto E, Rodríguez-Gascón A, Barrasa H, Maynar J, Carcelero E, Soy D, Isla A, Population pharmacokinetics of daptomycin in critically ill patients, Int J Antimicrob Agents, 52 (2018) 158–165. [DOI] [PubMed] [Google Scholar]
