Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Dec 1.
Published in final edited form as: Nat Chem Biol. 2020 Nov 16;16(12):1293–1302. doi: 10.1038/s41589-020-00674-6

Defining New Chemical Space for Drug Penetration into Gram-Negative Bacteria

Shibin Zhao 1,, Justyna W Adamiak 2,, Vincent Bonifay 2, Jitender Mehla 2, Helen I Zgurskaya 2,*, Derek S Tan 1,3,*
PMCID: PMC7897441  NIHMSID: NIHMS1649258  PMID: 33199906

Abstract

We live in the era of antibiotic resistance, and this problem will progressively worsen if no new solutions emerge. In particular, Gram-negative pathogens present both biological and chemical challenges that hinder the discovery of new antibacterial drugs. First, these bacteria are protected from a variety of structurally diverse drugs by a low-permeability barrier comprised of two membranes with distinct permeability properties, in addition to active drug efflux, making this cell envelope impermeable to most compounds. Second, chemical libraries currently used in drug discovery contain few compounds that can penetrate Gram-negative bacteria. As a result, intensive screening campaigns have led to few successes, highlighting the need for new approaches to identify regions of chemical space that are specifically relevant to antibacterial drug discovery. Herein, we provide an overview of emerging insights into this problem and outline a general approach to addressing it using prospective analysis of chemical libraries for the ability of compounds to accumulate in Gram-negative bacteria. The overall goal is to develop robust cheminformatic tools to predict Gram-negative permeation and efflux, which can then be used to guide medicinal chemistry campaigns and the design of antibacterial discovery libraries.

Introduction

Drug-resistant bacteria pose an increasing threat to public health1-3. In particular, Gram-negative pathogens pose a serious and urgent problem, comprising four out of the six ESKAPE pathogens that have high intrinsic antibiotic resistance and can acquire additional resistance through mutations and horizontal gene transfer4,5. Infections caused by these pathogens are difficult to treat using existing drugs and, of the handful of new drugs in clinical development for these indications, none represents a novel structural class or mechanism of action6. Thus, a robust pipeline of new drugs must be developed to keep pace with the continuing emergence of drug-resistant pathogens.

Discovery of new antibiotics against drug-resistant Gram-negative pathogens is hindered by both biological and chemical challenges. The major biological challenge is the two-membrane cell envelope of Gram-negative bacteria, which presents a low-permeability barrier that protects the cells against a variety of structurally diverse drugs (Fig. 1)7-12. The outer membrane (OM) restricts penetration of large and hydrophobic compounds. Conversely, the cytoplasmic or inner membrane (IM) limits hydrophilic compounds from crossing8,9. Trans-envelope efflux pumps extrude molecules that have successfully penetrated through the membranes. This combination of orthogonal membranes and efflux pumps creates a barrier that is impermeable to most compounds13-15. This biological challenge is then the basis for the major chemical challenge, which is the identification of new molecules that can penetrate both membranes and avoid efflux. Existing screening collections have provided few such compounds16,17 and our limited knowledge of the chemical properties required for Gram-negative penetration makes it difficult to design molecules with improved permeability9,11,18. To address this critical problem, we propose herein a comprehensive approach to understanding the biological and chemical factors that influence small-molecule permeation and efflux in Gram-negative bacteria, with the ultimate goal of developing robust, predictive cheminformatic models that can be used to design libraries and individual molecules with improved properties.

Fig. 1 ∣. The Gram-negative cell envelope and pathways of drug fluxes across it.

Fig. 1 ∣

The envelope includes a network of lipopolysaccharides that limits penetration of large and hydrophobic compounds (green C) (1), porin channels that permit passage of small, hydrophilic molecules (red C) (2), and trans-envelope efflux pumps that capture their substrates from the periplasm (3) or cytosol (4) and pump them out of the cell. The kinetic scheme describes four compartments: outside the cell (O), within the outer membrane (M), in the periplasm (P), and in the cytoplasm (I). Active efflux is approximated as a Michaelis–Menten process, where V is the maximum achievable velocity, P is the concentration of drug in the periplasm, and KM is the Michaelis constant. The binding to the membrane is postulated to be saturable, with the maximal flux F. The degree of saturation is denoted as ϕ, while k1 through k4 are microscopic rate constants14. Additional transporters can be readily integrated into this model13.

Recent advances in understanding the Gram-negative permeability barrier

Emerging hierarchy of rules of permeation.

Recent efforts have generated significant advances in rationalization of compound permeation across Gram-negative cell envelope, leading to emergence of a hierarchy of rules of permeation12. At the highest level, permeation into Gram-negative bacteria differs from that in Gram-positive bacteria because of the two-membrane cell envelope, with active efflux acting across both membranes (Fig. 1). The OM of Gram-negative bacteria is an asymmetric lipid bilayer consisting of an inner leaflet of phospholipids and an outer leaflet of glycolipids, primarily lipopolysaccharide (LPS)19,20. The glycolipids of different bacterial species vary in structure, resulting in membranes with differing permeability properties21. The OM also contains various porin proteins, which provide channels through which small, polar molecules can pass7,22. Furthermore, although multidrug efflux pumps are present in the cytoplasmic membranes of both bacterial phyla, active efflux across the low permeability barrier of the OM in Gram-negative cells is the most effective14,23. These efflux pumps establish non-linear, synergistic relationships between permeation across the OM and active efflux from both the periplasm and cytoplasm in Gram-negative bacteria, rendering the Gram-negative cell envelope virtually impermeable to certain classes of molecules13,14.

At the next level, the OM and active efflux exclude compounds based on different molecular and physicochemical principles. Separation of the contributions of the OM barrier and active efflux to intracellular accumulation, and hence to the activities of antibiotics, can be achieved by breaching the OM barrier or changing the expression of efflux pumps. However, this is not a trivial task and a variety of approaches have been used. Permeabilization using cationic peptides such as polymyxins is effective against susceptible Gram-negatives but fails against intrinsically resistant species such as Burkholderia spp. and resistant clinical isolates that contain lipid A modifications24. Alternatively, the structure of the OM bilayer can be compromised by genetic mutations, but analysis of the results is complicated by pleiotropic effects of these mutations25. More recently, hyperporinated strains that produce an inducible large pore in the OM have been constructed, in which the OM is permeable to a variety of compounds15,26. On the other hand, efflux efficiency can be controlled by inactivation or overproduction of efflux pumps. The downside of this approach is that, in some species such as Acinetobacter spp., Burkholderia spp., and others, efflux pumps are tightly integrated in cell physiology, and changes in their activities may affect multiple pathways27,28. Nevertheless, by modulating the OM barrier through hyperporination and deletion of efflux pumps, it has been shown that antibiotics cluster into different groups, depending on the main permeation barrier that they encounter while acting on cells, and that the OM barrier and efflux pumps recognize different properties in compounds15. Furthermore, at least in the case of Escherichia coli, these properties are orthogonal, with efflux pumps most efficiently recognizing compounds that can permeate the OM.

At the lower levels, bacterial species-specific differences in the composition and properties of the OM and efflux pumps also affect drug permeation. Although the basic architecture and composition of the cell envelope is conserved across different Gram-negative bacteria, different evolutionary paths and environmental conditions have led to significant variability in their properties29. A well-recognized example is the permeability barrier of Pseudomonas aeruginosa, which has only one-eighth the permeability of a typical E. coli cell8. In this case, the repertoires of efflux pumps (12 RND [resistance-nodulation-division] pumps in P. aeruginosa vs. four in E. coli) and OM porins (P. aeruginosa lacks general 600-Da cut-off porins that are abundant in E. coli) define the differences in permeability barriers. In addition, the structures of lipids, especially in the OM, define the differences in permeation in Acinetobacter spp. and Burkholderia spp.

Permeation through general porins of E. coli.

Predictive models based on physicochemical features of compounds and their interactions with E. coli porins have been reported recently22,30 and success stories in using these models to optimize the activity of compounds against Gram-negative bacteria have also emerged31-36. Porin-mediated permeation is a multivariate, dynamic process that is governed by multi-step interactions and physicochemical properties of both the antibiotic and the channel. These interactions have been studied in detail using electrophysiological analyses in reconstituted lipid bilayers37,38 Quantification of these intermolecular interactions is essential for revealing “rules” of permeation. Recently, a quantitative model was developed to predict the permeability of molecules through the OM based exclusively on structural features of E. coli general porins39. In general, antibiotic permeation through the constriction zone (the narrowest part of the channel) of a porin requires the molecule to adopt a conformation that is compatible with both the shape and functional groups presented by the channel. The model integrates the size of the molecule and its electrostatic interactions with the porin through its charge and dipole moment. These properties have been identified as top predictors of the antibacterial activities of β-lactam and fluoroquinolone antibiotics and their dependence on permeation in E. coli40, and in intracellular accumulation experiments discussed below31,41.

New kinetic models of drug permeation against active efflux in the context of two membranes.

The synergy between the OM permeability barrier and active efflux was recognized in early studies based on multiplicative changes in the activities of antibiotics in bacterial strains with compromised OM barriers or efflux systems42. However, kinetic models that integrate both factors in the permeation across the Gram-negative cell envelopes have only recently been developed and validated13,14 (Fig. 1). These models predict non-linear relationships between the extracellular concentration of a compound and its intracellular accumulation, based on the rate of compound permeation across the OM and the efficiency of its efflux. The system is controlled by two kinetic parameters: a barrier constant and an efflux constant that account for the efficiency of these two processes, respectively. Multiple transporters acting across either one or both membranes can be readily integrated into the model and their contributions to the intracellular accumulation can be predicted.

The quantitative kinetic models also revealed basic rules that define how the kinetic parameters are integrated in a multi-transporter system functioning in the context of two membranes13,14. The level of saturation and final accumulation of compounds depend on their mode of interaction with and passage across the permeability barriers. Amphiphilic compounds that permeate the membrane barriers by transmembrane diffusion are expected to diffuse slowly across the OM, whereas hydrophilic molecules that permeate through the water-filled porins are instead limited by the IM. The presence of multiple drug transporters located in same membrane (inner or outer) decide the kinetic outcome of drug translocation. Modeling and experimental data show that transporters located in the same membrane function in an additive manner. In contrast, a multiplicative response was observed between transporters acting across different membranes. A synergy could be predicted between efflux transporters and the transmembrane flux/diffusion operating at the same membrane. These kinetic models can now guide the development of experimental techniques to quantify the intracellular accumulation of compounds.

Despite these advances, only qualitative matches have been found between minimum inhibitory concentration (MIC) values of antibiotics predicted by mathematical models and those determined experimentally, and between antibacterial activity and experimentally determined intracellular concentrations across several series of antibiotics13,43,44. Further efforts are needed to develop kinetic models that can predict antibacterial activity based on intracellular accumulation of compounds13.

The need to define new Gram-negative chemical space

Leading antibacterial drug discovery groups in industry have reported very low success rates in large screening campaigns for novel antibiotics, particularly for Gram-negative bacteria16,17. These studies also revealed the difficulties in converting hits identified in biochemical assays, and in whole-cell assays using permeability- and efflux-compromised strains, to lead compounds with antibacterial activity against wild-type bacteria, highlighting the need for a better understanding of permeability and efflux of small-molecule antibiotics.

Several studies have highlighted differences in physicochemical properties between known antibiotics and other classes of drugs45,46, as well as hits from screening campaigns47. However, general and robust rules that can be used to predict Gram-negative permeation and efflux have not yet been developed9,11,18. Such rules would provide a critical means to design compounds for increased accumulation in Gram-negative bacteria, both in the context of medicinal chemistry optimization of screening hits as well as in the earlier-stage design of screening libraries tailored for antibacterial drug discovery.

Bacterial accumulation studies of antibiotics.

Toward this goal, numerous studies have sought to correlate physicochemical properties with bacterial compound accumulation (Table 1). Many of the earliest studies focused on OM permeability of β-lactam antibiotics in E. coli, as periplasmic cleavage by β-lactamase enzymes can be detected with an iodometric assay and used to infer the periplasmic concentration of the drug48. These studies identified inverse correlations between hydrophobicity and OM diffusion49,50. Similar correlations were also found for penetration through individual porins in E. coli51 and when reconstituted in proteoliposomes52. For the OmpF and OmpC porins, increased penetration for zwitterionic analogues and decreased penetration for dianionic analogues was also noted. Analogues with especially bulky or protruding side chains penetrated less effectively than their hydrophobicity would predict, suggestive of inability to pass through the porin constriction zone.

Table 1 ∣.

Selected studies evaluating correlations between physicochemical properties and compound accumulation in bacteria.

Compounds tested (year) Bacteria Detection
method
Correlations identified
8 β-lactams (1977)49 E. coli β-lactamase/ iodometric hydrophobicity [P7.4a] (−)
8 cephalosporins (3 matched pairs) (1982)50 E. coli M. morganii β-lactamase/iodometric hydrophobicity [Rmb] (−)
11 cephalosporins (1983)51 E. coli (OmpF, OmpC, PhoE) β-lactamase/iodometric hydrophobicity [log Pc] (−)
zwitterionic (+; OmpF, OmpC)
dianionic (−; OmpF, OmpC)
36 β-lactams (1985)52 proteoliposomes (OmpF, OmpC) liposome swelling (OD400) hydrophobicity [log P] (−)
zwitterionic (+)
dianionic (−)
steric bulk (−)
5 tetracyclines (1992)53 E. coli reporter gene hydrophobicity [log KDd] (−)
11 fluoroquinolones (1992)56 E. coli P. aeruginosa intrinsic fluorescence hydrophobicity [log D7.2e] (−)
inhibition x accum. ≈ activity
S. aureus hydrophobicity [log D7.2e] (+)
8 fluoroquinolones (1992)57 E. coli P. aeruginosa intrinsic fluorescence hydrophobicity [Papp-7.2e] (−)
efflux inhibition [CCCPf] (+)
S. aureus hydrophobicity [Papp-7.2e] (+)
15 quinolones (1993)58 E. coli P. aeruginosa intrinsic fluorescence not hydrophobicity [P7.0g]
not molecular weight
S. aureus not hydrophobicity [P7.0g]
not molecular weight
10 fluoroquinolones (2001)59 E. coli intrinsic fluorescence not hydrophobicity [log Pg]
not molecular weight
S. aureus not hydrophobicity [log Pg]
molecular weight (+)
10 sulfamoyladenosines (2014)46 E. coli LC-MS/MS hydrophobicity [log D7.4h, log Pi] (+)
polarity [log Sj, relPSAk] (−)
B. subtilis hydrophobicity (log Pi) (+)
polarity (log Sj, relPSAk) (−)
surface areal (+)
rotatable bonds (+)
M. smegmatis ring count (+)
sp3 fractionm (−)
ring complexityn (−)
100 + 54 + 49 natural product-derived compounds (2017)31 E. coli LC-MS/MS primary amine (+)
globularity (−)
rotatable bonds (−)
amphiphilic moment [vsurf_Ao] (+)
20 pyrimidines + 80 pyridopyrimidinones (2018)43 E. coli LC-MS/MS not molecular weight
hydrophobicity (log D7.4p) (+)
positive charge (+)
a

Distribution coefficient in isobutanol–0.02 M phopshate buffer (pH 7.4).

b

Reverse-phase thin-layer chromatography retention parameter in 0.05 M Na2HPO4–KH2PO4 buffer (pH 7.0) at 0% acetone (fit of data with 0–6% acetone).

c

Partition coefficient in 1-octanol–water.

d

Partition coefficient in chloroform–buffer (pH 7.0).

e

Distribution coefficient in 1-octanol–0.1 M phosphate buffer (pH 7.2).

f

Carbonyl cyanide m-chlorophenyl hydrazone.

g

Partition coefficient in 1-octanol–0.1 M sodium phosphate buffer (pH 7.0).

h

Distribution coefficient in 1-octanol–buffer (pH 7.4), calculated (Instant JChem).

i

Partition coefficient in 1-octanol–water, calculated (Instant JChem and Virtual Computational Chemistry Lab).

j

Solubility constant in water, calculated (Virtual Computational Chemistry Lab).

k

Relative polar surface area, calculated (Instant JChem).

l

van der Waals surface area, calculated (Instant JChem).

m

sp3 carbons ÷ carbons.

n

Rings per ring system.

o

Length of vector pointing from center of hydrophobic domain to center of hydrophilic domain, calculated (Molecular Operating Environment).

p

Distribution coefficient in 1-octanol–buffer (pH 7.4), calculated (program not specified).

Moving beyond β-lactam antibiotics, investigation of intracellular accumulation of tetracyclines in E. coli using a reporter gene system also revealed an inverse correlation with hydrophobicity53. Studies of quinolone antibiotics in E. coli using radiolabeling or intrinsic fluorescence provided some of the first detailed insights into accumulation kinetics, revealing an initial rapid accumulation phase followed by a second, slower phase54,55.

In an early effort to correlate biochemical inhibition, compound accumulation, and antibacterial activity, further studies of fluoroquinolones identified moderate inverse correlations between hydrophobicity and accumulation in E. coli and P. aeruginosa, but a moderate positive correlation in the Gram-positive bacteria Staphylococcus aureus56. Notably, correlations between biochemical inhibition of DNA gyrase and MIC values could be improved by correcting for the accumulation level in E. coli and P. aeruginosa. This provided important proof of concept for integration of accumulation data with biochemical inhibition to predict antibacterial activity. A contemporaneous study found similar correlations for fluoroquinolone accumulation in these three bacteria and also tested the impact of carbonyl cyanide m-chlorophenyl hydrazone (CCCP), an inhibitor of energy-dependent efflux, finding increased accumulation in both of the Gram-negative bacteria57. However, a subsequent study found no correlations between hydrophobicity or molecular weight with accumulation or antibacterial activity of other quinolones in any of the three bacteria58. Moreover, later work with novel fluoroquinolones also showed no correlations in E. coli, nor any correlation between accumulation and antibacterial activity59. These divergent results highlight the challenges in identifying simple correlations between individual physicochemical parameters and bacterial accumulation and in extending those correlations to predict antibacterial activity.

Notably, later work highlighted two important approaches to probing the roles of other barrier components in compound accumulation studies, using E. coli porin mutants to dissect the effects of porins on OM passage60, and E. coli spheroplasts to differentiate between OM and IM effects61. More recently, titratable induction of P. aeruginosa OprD porin expression in E. coli, in which the three major porins had been knocked out (ompA, ompC, ompF), was used to infer OprD-mediated OM permeability of a panel of carbapenems from the proportionality of antibacterial activity to the level of porin expression41.

Bacterial accumulation studies of diverse compounds.

One of the main constraints of these early efforts may, ironically, have been their focus on antibiotics. This inherently limited the diversity of structures evaluated, thus potentially limiting the opportunity to identify general correlations between physicochemical properties and bacterial accumulation. In addition, studies that rely upon radiolabeling or intrinsic fluorescence are severely limited in the range of chemical structures that can be used. To address these limitations, two recent studies have used prospective, activity-independent analysis of diverse chemical libraries in bacterial compound accumulation assays. These compounds were not designed to have specific targets or antibacterial activity, but rather to cover a wide area of chemical space. This structure-agnostic analysis was enabled by the use of liquid chromatography–tandem mass spectrometry (LC-MS/MS) to quantitate intracellular compound concentrations in bacterial lysates62.

In the first of these two studies, a general platform for prospective compound accumulation studies was reported46. Initially, concentration-, time-, and efflux-dependent accumulation of salicyl-AMS (5´-O-[N-salicylsulfamoyl]adenosine), a siderophore biosynthesis inhibitor63, was assessed in E. coli. Accumulation at a single timepoint was then determined for a panel of 10 acyl-AMS analogues with diverse physicochemical properties in E. coli, Bacillus subtilis (Gram-positive), and Mycobacterium smegmatis, in the presence or absence of efflux pump inhibitors. Cheminformatic analysis using a set of 20 calculated physicochemical parameters identified significant positive correlations between accumulation in E. coli or B. subtilis and hydrophobicity, as well as inverse correlations with polarity and solubility. Additional positive correlations with B. subtilis accumulation were identified for surface area and rotatable bonds. For M. smegmatis accumulation, ring counts correlated positively while sp3-hybridized carbon fraction and ring complexity correlated negatively. Several significant correlations were also identified between physicochemical properties and efflux sensitivity. Subsequently, based on correlations identified in E. coli, two new salicyl-AMS analogues were designed and shown to exhibit higher accumulation as predicted. This provided proof-of-concept for the potential utility of this platform in developing predictive rules for use in antibiotic medicinal chemistry.

More recently, this approach was extended to a larger analysis of a diversity library of 100 compounds built primarily around five natural product-derived scaffolds31,64. Notably, only 12 positively-charged compounds accumulated significantly in E. coli, and within this subset, no correlation was observed with hydrophobicity or molecular weight, in contrast to some early studies of quinolones53,57. Subsequent synthesis and evaluation of designed analogues revealed a requirement for unhindered primary amines. Cheminformatic analysis using 297 calculated molecular descriptors led to a random forest classification model in which rotatable bonds and globularity were negative predictors of accumulation while amphiphilic moment (distance between hydrophobic and hydrophilic portions of a molecule) was a positive predictor. Evaluation of selected compounds in E. coli protoplasts, stripped of their OM and peptidoglycan layer, indicated limited differences in accumulation, suggesting that the OM represented the major selection filter for these compounds.

Notably, this study did not identify linear relationships between individual properties and accumulation, but rather established a set of threshold guidelines for accumulation, based on an unhindered primary amine, presence of some degree of hydrophobicity, rigidity, and low globularity. These were later codified into “eNTRy rules” (http://entry-way.org), representing presence of a primary amine (N), low three-dimensionality (T; globularity ≤0.25), and ≤5 rotatable bonds (R). These guidelines were then leveraged to design a new primary amine-containing analogue of deoxynybomycin, an antibiotic that is restricted to Gram-positive bacteria, which had increased accumulation and antibacterial activity against the Gram-negative bacteria E. coli, Acinetobacter baumannii, Klebsiella pneumoniae, and Enterobacter cloacae. Notably, however, the analogue showed only modest activity against P. aeruginosa, indicating the need for further species-specific analyses. While correlations of antibacterial activity with positive charge are well-established in antibiotic medicinal chemistry41,51,52 and the impact of molecular shape in porin passage through the outer membrane has also been suggested previously52, this seminal study demonstrated the utility of large, prospective, activity-independent analyses of diversity libraries in the systematic identification of physicochemical properties that are predictive of accumulation and can be used in medicinal chemistry optimization. This dataset also served as the basis for the computational model for OM porin passage described above39. Moreover, the eNTRy rules have been successfully applied to develop Gram-negative active analogues of the Gram-positive-restricted antibiotics Debio-1452, an inhibitor of the enoyl-acyl carrier reductase enzyme FabI32, and ribocil, an inhibitor of the E. coli riboflavin riboswitch33.

In another recent study using LC-MS/MS detection, accumulation of >100 NAD-dependent DNA ligase (LigA) inhibitors was tested in E. coli43. Consistent with several earlier small studies58,59, they found poor correlation between the overall level of bacteria-associated compound and antibacterial activity in compounds with matched biochemical activities. This suggested the importance of considering subcellular localization as well as accumulation kinetics in efforts to correlate biochemical inhibition, compound accumulation, and antibacterial activity.

A roadmap for future efforts

The studies highlighted above have provided useful insights into physicochemical properties that impact compound accumulation in Gram-negative bacteria and highlighted the challenges of correlating intracellular accumulation with antibacterial activity. This provides a roadmap for further efforts to develop robust, predictive tools for compound accumulation to support antibacterial drug discovery. However, several challenges remain:

  1. Bespoke chemical libraries must be designed and synthesized to probe chemical space systematically.

  2. Various bacterial species have considerable differences between their uptake and efflux pathways that must be assessed individually.

  3. Combined effects of OM, IM, and efflux on overall accumulation and subcellular localization must be dissected.

  4. Time and concentration dependence of accumulation kinetics may need to be considered to develop robust models.

  5. Accumulation assays that are both high-throughput and structure agnostic are needed.

  6. Cheminformatic models that accommodate multivariate and non-linear impacts of physicochemical properties on accumulation and kinetics must be developed.

We envision that future efforts should continue to involve large-scale, prospective analyses of designed chemical libraries for accumulation in multiple pathogenic Gram-negative bacteria, integrating assessments of OM, IM, and efflux contributions to accumulation, and considering both steady-state and kinetic accumulation data, as well as antibacterial activity where relevant, to develop robust cheminformatic models to predict accumulation.

With these considerations in mind, our labs at Memorial Sloan Kettering Cancer Center and the University of Oklahoma, with collaborators at Merck Research Labs, have launched one such multidisciplinary program called SPEAR-GN (Small-molecule Penetration and Efflux in Antibiotic-Resistant Gram-Negative bacteria) to pursue this approach (Fig. 2). The name draws analogy between the issues of permeation and efflux to equipment used in spearfishing, in which the spear must penetrate its target and is also barbed to prevent it from falling out. This work requires contributions from disparate disciplines, including diversity-oriented synthesis, medicinal chemistry, molecular microbiology, high-throughput assay technologies, analytical chemistry, biophysical modeling, and cheminformatics. Importantly, many key enabling technologies have been developed in recent years to make this feasible, and these are discussed in further detail below.

Fig. 2 ∣. Comprehensive approach to developing cheminformatic tools to predict Gram-negative bacterial compound accumulation.

Fig. 2 ∣

Chemical libraries are designed to probe regions of chemical space and to assess impacts of scaffolds, regiochemistry, and substituents. For each bacterial species, four isogenic strains are used – wild-type, hyperporinated, efflux-deficient, and doubly-compromised – to dissect the contributions of the OM, IM, and efflux to net accumulation (blue = native porins, green = engineered large-bore porins, purple = efflux pumps). Compounds are screened in high-throughput assays and analyzed by high-throughput detection methods, such as solid-phase extraction–mass spectrometry. Steady-state analyses are performed for large collections of compounds while detailed kinetic analyses are carried out for smaller subsets of compounds to enable calculation of kinetic parameters. Regression and classification approaches are used to develop cheminformatic models that correlate physiochemical properties with accumulation, which are then tested further in additional analogues.

Evolution of enabling technologies

Bespoke chemical libraries.

Previous analyses have demonstrated the need to explore new regions of chemical space tailored specifically for antibacterial drug discovery16,17,45-47. Indeed, the “eNTRy” rules described above were formulated using a diversity library based on novel scaffolds31,64. Future efforts should continue to expand the range of scaffolds tested, as well as use a priori cheminformatic analysis to ensure that the libraries span a wide range of properties. Syntheses of such libraries at a scale sufficient for comprehensive analysis is now readily feasible following over two decades of work in the field of diversity-oriented synthesis65,66.

In our work, we are also using a common set of substituents across different scaffolds and regioisomeric positions to assess the generality of models we derive from these libraries. In the limit, it may be possible to develop a universal model that is applicable to all chemotypes, but this is viewed as unlikely, given that penetration through OM porins and efflux by various pumps are protein-mediated processes that have some degree of substrate specificity. A more likely scenario is that models will have some aspects that are generally applicable across multiple chemotypes, while other aspects are chemotype-specific. In the other extreme, it may be necessary to develop a specific model for every chemotype. However, even if this is the case, generalizable platforms such as that in the SPEAR-GN project may enable relatively rapid development of such models for individual, high-value chemotypes.

Species-specific contributions of OM, IM, and efflux.

Previous large compound accumulation screens have given limited consideration to species-specific effects and subcellular localization31,43,46. As a consequence, the resulting models do not always translate between different bacterial species31. Inability of a drug to access the same subcellular compartment as its target may also contribute to poor correlations between accumulation and antibacterial activity9,43. Future efforts should incorporate such considerations directly into screening and modeling. Toward this end, we are conducting our studies in isogenic strains of E. coli, P. aeruginosa, and A. baumannii that allow cross-species comparisons and enable the relative roles of OM and IM passage and efflux to be differentiated, thus also allowing inference of subcellular localization. These four-strain sets are comprised of the wild-type strain, a hyperporinated strain engineered with an inducible large-bore porin that renders the OM freely permeable, a pan-efflux knockout strain, and a doubly compromised strain that is both hyperporinated and efflux deficient and provides direct assessment of IM permeability14,15,26,40. The hyperporinated strain provides advantages over spheroplasts or protoplasts, as it more closely approximates the normal cell envelope structure and is also readily grown in culture prior to induction of pore expression. Use of efflux mutants is preferred over small-molecule efflux inhibitors that can have pluripotent effects and alter cell envelope structure. This approach also provides practical advantages over subcellular fractionation, which is experimentally demanding and relatively low-throughput67. By comparing results across the four strains in each bacterial species, it is possible to determine the relative contributions of the OM, IM, and efflux to accumulation, thereby allowing identification the relevant barrier(s) and the development of predictive models for each barrier.

Consideration of accumulation kinetics.

Early studies of antibiotic accumulation kinetics generally showed a rapid initial uptake phase, followed by a slower second phase54-57. More recent large compound accumulation screens have used a single timepoint, assumed to approximate a steady-state concentration while minimizing effects on bacterial growth31,43,46. However, we have reported detailed kinetic modeling of compound accumulation in Gram-negative bacteria indicating that differential effects of OM and IM passage, efflux, and target binding result in a complex kinetic landscape that is dependent upon both time and concentration14. Thus, in addition to these conventional steady-state analyses, which can be done for hundreds to thousands of compounds, we are also testing representative subsets of compounds at multiple timepoints and concentrations to enable determination of kinetic parameters. This may provide predictive models that are more robust and quantitative than the steady-state-based models developed to date.

High-throughput accumulation assays.

Recent large compound accumulation screens have relied upon manual, tube-based assays that limit throughput31,43,46. Analysis of larger libraries, across multiple strains and species, using multiple timepoints and concentrations will require a high-throughput assay that incorporates compound incubation, bacterial lysis, lysate extraction, and sample analysis, ideally in a structure-agnostic fashion. Recently, a 96-well plate assay was developed using a filter plate to separate cells from unbound compound, and leveraging automated solid-phase extraction (SPE) technology for rapid sample cleanup and MS analysis44. Use of the SPE-MS technology allowed each sample to be analyzed in as little as 9 sec. We are now using a similar approach in our own work. Of note, an alternative approach that quantitates the remaining extracellular compound has been reported68. While this approach avoids the lysis step, its dynamic range is likely limited by the small intracellular volume of the cells compared to the extracellular solution.

Importantly, a kaleidoscope of technologies has been developed to analyze bacterial compound accumulation (Table 2). Each of these approaches has advantages and limitations with respect to chemical structural requirements, sensitivity, throughput, and other considerations. Radiolabeing provides quantitative, sensitive measurements54, but limits structural diversity and throughput for practical reasons. Numerous studies have used intrinsic fluorescence55, allowing for more convenient and higher-throughput measurements, as well as increased resolution down to the single-cell or subcellular levels69-75. However, this also presents significant structural constraints and the alternative approach of labeling non-fluorescent compounds with a fluorophore introduces a large substituent that is likely to perturb accumulation. MS-based approaches in use in our project provide an attractive combination of structure independence, sensitivity, and throughput, although they require significant investments in instrumentation and extensive experimental optimization. LC-MS/MS with electrospray ionization (ESI) integrates chromatographic sample separation, is versatile in detecting a variety of chemotypes, and provides good limits of detection62. Matrix-assisted laser desorption/ionization (MALDI) has also been used to ‘image’ antibiotics in lungs of tuberculosis-infected animals76,77, and it is possible that such an approach could be adapted to higher-throughput formats using cell culture samples. As noted above, the advent of automated SPE systems44 has increased sample processing throughput considerably. Analysis speeds may be further increased to the ultrahigh-throughput regime through integration of microfluidic technologies78,79 with mass spectrometric methods of compatible speed, such as desorption electrospray ionization (DESI)80,81 and acoustic droplet ejection (ADE)82, although such direct sampling of biological matrices without sample cleanup also results in decreased sensitivity. Other approaches have also been reported recently for more specialized applications, including Raman spectroscopy71,83, which allows label-free detection of antibiotics and metabolites inside intact bacterial cells; electrophysiology84,85, which can be used to study antibiotic transport dynamics in whole cells and artificial membranes, and biomimetic membrane models85,86, which allow detailed studies of the kinetics and mechanisms of passive diffusion.

Table 2 ∣.

Methods to measure intracellular compound accumulation in bacteria.

Method Advantages Limitations
radioactivity54 sensitive, quantitative radioactive material, requires radiolabel that may alter activity and permeation, low-throughput, limited structural diversity
fluorimetry69,70 (including fluorescence microscopy71,72, spectrofluorimetry73, flow cytometry74,75) ease of use, high-throughput, kinetics down to ms, resolution down to single cell level requires intrinsic fluorescence or fluorescent label that may alter activity and permeation, limited structural diversity
LC-MS/MSa,62 label-free, quantitative to nM concentrations, in-line sample separation low-throughput (5–20 min/sample)
SPE-MSb,44 label-free, high-throughput (<10 s/sample), in-line sample clean up extensive optimization required, complex matrix can suppress signal
MALDI-MSc,76,77 label-free, high temporal and spatial (25 μm) resolution of drug distribution requires appropriate matrix, moderate limit of detection
microfluidic ESI-MSd,78,79 label-free, high-throughput (45 s/sample), small sample size no sample separation, frequent clogging, reproduciblity of microchip fabrication
DESI-MSe,80,81 label-free, ambient ionization, non-destructive, ultrahigh-throughput (<1 s/sample) semi-quantitative, moderate limit of detection
ADE-MSf,82 label-free, ultrahigh-throughput (<1 s/sample) no sample clean up, complex matrix will suppress signal
Raman spectroscopy71,83 label-free, in cellulo analysis, high-throughput cellular background interference, requires chemical reaction of analyte
electrophysiology84,85 provides insights into roles of membrane potential dynamics experimental complexity, low-throughput
biomimetic membrane models85,86 realistic model systems, sensitive, kinetics down to ms concentration gradients due to reduced mass transport close to membrane may impact accumulation
a

LC-MS/MS = liquid chromatography–tandem mass spectrometry (triple quadrupole).

b

SPE-MS = solid-phase extraction–mass spectrometry.

c

MALDI-MS = matrix-assisted laser desorption/ionisation mass spectrometry.

d

ESI = electrospray ionization.

e

DESI = desorption electrospray ionization.

f

ADE = acoustic droplet ejection.

Cheminformatic modeling approaches.

The ultimate goal of these studies is to develop predictive cheminformatic models that can be used to guide antibacterial drug discovery library design and medicinal chemistry efforts. Both regression and classification approaches have been used previously, including linear regression analysis, Pearson pairwise correlation coefficient analysis, principle component analysis, and random forest classification31,46,53,56. Notably, random forest classification can consider non-linear correlations and interactions between multiple parameters, which may be important for identifying useful relationships between physicochemical properties and bacterial accumulation. We envision that, as in QSAR (quantitative structure–activity relationship) modeling, the models are likely to be more useful to chemists if they are readily interpretable, rather than derived from a black box as in the case of neural networks. These models will be developed to predict overall accumulation as well as individual kinetic parameters where sufficient data is available. The models will then be tested by designing and testing new analogues predicted to have increased or decreased overall accumulation or kinetic rates for individual barriers, as demonstrated previously by us and others31,46. Validated models can then be used to inform subsequent medicinal chemistry efforts on individual scaffolds and broader efforts to design novel libraries for antibacterial drug discovery.

Conclusion

Recent efforts to screen collections of diverse compounds in a prospective, activity-independent fashion have demonstrated the utility of these approaches in identifying physicochemical properties that correlate with accumulation and can be used in the design of antibacterial drug discovery libraries as well as in medicinal chemistry campaigns31,46. We envision that future efforts in this area should continue to address the six challenges highlighted above. Studies should continue to expand the use of diverse, fit-for-purpose chemical libraries that probe broad regions of chemical space. Such libraries will certainly be distinct from existing screening collections that have evolved largely from campaigns against human protein targets, and may also be distinct from known classes of antibiotics. Analyses must be carried out across multiple bacterial species to differentiate between cheminformatic insights that are generally applicable and others that are species-specific. It will also be important to assess how broadly insights derived from laboratory model strains can be applied across various clinical isolates, including antibiotic-resistant strains. Subcellular localization must also be considered in assay designs to ensure that the derived models predict concentrations at the relevant site of action. Differentiation of impacts of the individual barriers (OM, IM, efflux) may also allow separate models to be developed for each and may facilitate medicinal chemistry optimization efforts. Consideration of accumulation kinetics may also provide more robust models than steady-state measurements, although this will need to be balanced judiciously with the necessary reduction in compound throughput to ensure that sufficient data is available for cheminformatic modeling. To accommodate the daunting number of experiments envisioned above, there remains a great need for development of accumulation assays and quantitative detection methods that are both high-throughput and structure-agnostic. Ideally, these technologies should be broadly accessible to the scientific community with respect to both technical complexity and cost to ensure widespread uptake and maximal utility. Finally, any cheminformatic models developed must be readily interpretable in terms of structural and physicochemical properties in order to be useful to medicinal chemists. Regression models that correlate accumulation or kinetic parameters with specific properties, and classification models that identify relevant structural features will be most useful in guiding the design of new compounds. Conversely, although complex, black box models may be more accurate and robust, they are difficult to implement effectively in the context of prospective medicinal chemistry design.

In our own SPEAR-GN program, these efforts are well underway, including cheminformatic design and synthesis of new libraries, engineering of isogenic four-strain sets, development of a 96-well plate accumulation assay, integration with SPE-MS analysis, development of a computational tool for kinetic modeling, and completion of pilot assays using a panel of known antibiotics. This sets the stage for forthcoming large-scale screening to enable the development and validation of predictive cheminformatic models to guide antibacterial drug discovery efforts in the future.

After more than 40 years of study, the problem of predicting small-molecule permeability in Gram-negative bacteria remains largely unsolved. However, while the complexity of the Gram-negative cell envelope poses a formidable challenge, recent advances suggest that this problem is becoming increasingly tractable. Solving it will require a concerted, sustained, and multidisciplinary effort across the scientific community, and would have widespread and lasting impacts on public health.

Acknowledgments

We thank A. Duerfeldt and V. Rybenkov (University of Oklahoma); C. Balibar, D. McLaren, B. Sherborne, B. Squadroni, M. Tudor, and S. Walker (Merck Research Labs); H. Voss (Weill Cornell Medicine); and the entire SPEAR-GN team for helpful discussions. Financial support from the National Institutes of Health (R01 AI136795 to D.S.T. and H.I.Z.; R01 GM100477 and R01 AI118224 to D.S.T.; R01 AI136799 to H.I.Z.; and MSK CCSG P30 CA008748 to C. B. Thompson) is gratefully acknowledged.

Footnotes

Competing Interest Statement

Merck is a collaborating institution on the SPEAR-GN project and has provided in-kind support to the labs of D.S.T. and H.I.Z. D.S.T. serves on the External Advisory Board of the Institute for Research in Biomedicine, Barcelona; is a shareholder and has been a paid consultant and speaker for Merck; and has been a paid consultant or speaker for Eli Lilly, Elsevier, Emerson Collective, and Venenum Biosciences. H.I.Z. has been a paid speaker for Genentech and Novartis Research Institutes.

References

  • 1.Centers for Disease Control and Prevention, “Antibiotic Resistance Threats in the United States, 2019. ” https://www.cdc.gov/drugresistance/pdf/threats-report/2019-ar-threats-report-508.pdf.
  • 2.World Health Organization, “Antimicrobial Resistance: Global Report on Surveillance 2014. ” http://www.who.int/drugresistance/documents/surveillancereport/en/.
  • 3.The Pew Charitable Trusts, “A Scientific Roadmap for Antibiotic Discovery.” 2016; https://www.pewtrusts.org/en/research-and-analysis/reports/2016/05/a-scientific-roadmap-for-antibiotic-discovery.
  • 4.Rice LB Federal funding for the study of antimicrobial resistance in nosocomial pathogens: No ESKAPE. J. Infect. Dis 197, 1079–1081 (2008). [DOI] [PubMed] [Google Scholar]
  • 5.De Oliveira DMP et al. Antimicrobial resistance in ESKAPE pathogens. Clin. Microbiol. Rev 33, e00181 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.The Pew Charitable Trusts, “Tracking the Global Pipeline of Antibiotics in Development, April 2020. ” https://www.pewtrusts.org/en/research-and-analysis/issue-briefs/2020/04/tracking-the-global-pipeline-of-antibiotics-in-development.
  • 7.Nikaido H Molecular basis of bacterial outer membrane permeability revisited. Microbiol. Mol. Biol. Rev 67, 593–656 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Zgurskaya HI, Lopez CA & Gnanakaran S Permeability barrier of Gram-negative cell envelopes and approaches to bypass It. ACS Infect. Dis 1, 512–522 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Silver LL A Gestalt approach to Gram-negative entry. Bioorg. Med. Chem 24, 6379–6389 (2016). [DOI] [PubMed] [Google Scholar]
  • 10.Masi M, Refregiers M, Pos KM & Pages J-M Mechanisms of envelope permeability and antibiotic influx and efflux in Gram-negative bacteria. Nat. Microbiol 2, 17001 (2017). [DOI] [PubMed] [Google Scholar]
  • 11.Tommasi R, Iyer R & Miller AA Antibacterial drug discovery: Some assembly required. ACS Infect. Dis 4, 686–695 (2018). [DOI] [PubMed] [Google Scholar]
  • 12.Zgurskaya HI & Rybenkov VV Permeability barriers of Gram-negative pathogens. Ann. N. Y. Acad. Sci 1459, 5–18 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Saha P, Sikdar S, Krishnamoorthy G, Zgurskaya HI & Rybenkov VV Drug permeation against efflux by two transporters. ACS Infect. Dis 6, 747–758 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Westfall DA et al. Bifurcation kinetics of drug uptake by Gram-negative bacteria. PLoS One 12, e0184671 (2017).A kinetic model that accurately describes small-molecule permeation of the Gram-negative envelope was developed. The model predicted complex, non-linear patterns dependent upon concentration and time that were validated experimentally.
  • 15.Krishnamoorthy G et al. Synergy between active efflux and outer membrane diffusion defines rules of antibiotic permeation into Gram-negative bacteria. mBio 8, e01172 (2017).Analysis of the activity of antibiotics in 4-strain sets of wild-type, hyperporinated, efflux-deficient, and doubly-compromised A. baumannii, P. aerguinosa, and Burkholderia spp. revealed a synergistic relationship between efflux and the permeability barrier in protecting these bacteria from structurally-diverse antibiotics.
  • 16.Payne DJ, Gwynn MN, Holmes DJ & Pompliano DL Drugs for bad bugs: confronting the challenges of antibacterial discovery. Nat. Rev. Drug Discovery 6, 29–40 (2007). [DOI] [PubMed] [Google Scholar]
  • 17.Tommasi R, Brown DG, Walkup GK, Manchester JI & Miller AA ESKAPEing the labyrinth of antibacterial discovery. Nat. Rev. Drug Discovery 14, 529–542 (2015). [DOI] [PubMed] [Google Scholar]
  • 18.Lewis K Antibiotics: Recover the lost art of drug discovery. Nature (London, U. K.) 485, 439–440 (2012). [DOI] [PubMed] [Google Scholar]
  • 19.Kamio Y & Nikaido H Outer membrane of Salmonella typhimurium: Accessibility of phospholipid head groups to phospholipase C and cyanogen bromide activated dextran in the external medium. Biochemistry 15, 2561–2570 (1976). [DOI] [PubMed] [Google Scholar]
  • 20.Raetz CR & Whitfield C Lipopolysaccharide endotoxins. Annu. Rev. Biochem 71, 635–700 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Henderson JC et al. The power of asymmetry: Architecture and assembly of the Gram-negative outer membrane lipid bilayer. Annu. Rev. Microbiol 70, 255–278 (2016). [DOI] [PubMed] [Google Scholar]
  • 22.Vergalli J et al. Porins and small-molecule translocation across the outer membrane of Gram-negative bacteria. Nat. Rev. Microbiol 18, 164–176 (2020). [DOI] [PubMed] [Google Scholar]
  • 23.Zgurskaya HI, Rybenkov VV, Krishnamoorthy G & Leus IV Trans-envelope multidrug efflux pumps of Gram-negative bacteria and their synergism with the outer membrane barrier. Res. Microbiol 169, 351–356 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Poirel L, Jayol A & Nordmann P Polymyxins: Antibacterial activity, susceptibility testing, and resistance mechanisms encoded by plasmids or chromosomes. Clin. Microbiol. Rev 30, 557–596 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Pagnout C et al. Pleiotropic effects of rfa-gene mutations on Escherichia coli envelope properties. Sci. Rep 9, 9696 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Krishnamoorthy G et al. Breaking the permeability barrier of Escherichia coli by controlled hyperporination of the outer membrane. Antimicrob. Agents Chemother 60, 7372–7381 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Leus IV et al. Substrate specificities and efflux efficiencies of RND efflux pumps of Acinetobacter baumannii. J Bacteriol 200, e00049 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Krishnamoorthy G et al. Efflux pumps of Burkholderia thailandensis control the permeability barrier of the outer membrane. Antimicrob. Agents Chemother 63, e00956 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Sohlenkamp C & Geiger O Bacterial membrane lipids: Diversity in structures and pathways. FEMS Microbiol. Rev 40, 133–159 (2016). [DOI] [PubMed] [Google Scholar]
  • 30.Richter MF & Hergenrother PJ The challenge of converting Gram-positive-only compounds into broad-spectrum antibiotics. Ann. N. Y. Acad. Sci 1435, 18–38 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Richter MF et al. Predictive compound accumulation rules yield a broad-spectrum antibiotic. Nature (London, U. K.) 545, 299–304 (2017).Screening of a library of 100 diverse, natural product-derived compounds for compound accumulation in E. coli identified a primary amine, amphiphilic moment, low globularity, and rigidity as factors associated with accumulation, enabling design of a Gram-negative-active analogue of a Gram-positive-restricted antibiotic.
  • 32.Parker EN et al. Implementation of permeation rules leads to a FabI inhibitor with activity against Gram-negative pathogens. Nat. Microbiol 5, 67–75 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Motika SE et al. Gram-negative antibiotic active through inhibition of an essential riboswitch. J. Am. Chem. Soc 142, 10856–10862 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Isabella VM et al. Toward the rational design of carbapenem uptake in Pseudomonas aeruginosa. Chem. Biol 22, 535–547 (2015). [DOI] [PubMed] [Google Scholar]
  • 35.Li Y et al. First-generation structure-activity relationship studies of 2,3,4,9-tetrahydro-1H-carbazol-1-amines as CpxA phosphatase inhibitors. Bioorg. Med. Chem. Lett 29, 1836–1841 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Green AT et al. Discovery of multidrug efflux pump inhibitors with a novel chemical scaffold. Biochim. Biophys. Acta Gen. Subj 1864, 129546 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Nestorovich EM, Danelon C, Winterhalter M & Bezrukov SM Designed to penetrate: Time-resolved interaction of single antibiotic molecules with bacterial pores. Proc. Natl. Acad. Sci. U. S. A 99, 9789–9794 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Wang J, Terrasse R, Bafna JA, Benier L & Winterhalter M Electrophysiological characterization of transport across outer-membrane channels from Gram-negative bacteria in presence of lipopolysaccharides. Angew. Chem., Int. Ed 59, 8517–8521 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Acosta-Gutierrez S et al. Getting drugs into Gram-negative bacteria: Rational rules for permeation through general porins. ACS Infect. Dis 4, 1487–1498 (2018). [DOI] [PubMed] [Google Scholar]
  • 40.Cooper SJ et al. Molecular properties that define the activities of antibiotics in Escherichia coli and Pseudomonas aeruginosa. ACS Infect. Dis 4, 1223–1234 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Iyer R et al. Whole-cell-based assay to evaluate structure permeation relationships for carbapenem passage through the Pseudomonas aeruginosa porin OprD. ACS Infect. Dis 3, 310–319 (2017). [DOI] [PubMed] [Google Scholar]
  • 42.Nikaido H Prevention of drug access to bacterial targets: Permeability barriers and active efflux. Science 264, 382–388 (1994). [DOI] [PubMed] [Google Scholar]
  • 43.Iyer R et al. Evaluating LC-MS/MS to measure accumulation of compounds within bacteria. ACS Infect. Dis 4, 1336–1345 (2018).Screening of a library of >100 DNA ligase inhibitors for compound accumulation in E. coli showed poor correlation between overall bacteria-associated compound levels and antibacterial activity in compounds with matched biochemical activities, highlighting the importance of subcellular localization.
  • 44.Widya M et al. Development and optimization of a higher-throughput bacterial compound accumulation assay. ACS Infect. Dis 5, 394–405 (2019). [DOI] [PubMed] [Google Scholar]
  • 45.O’Shea R & Moser HE Physicochemical properties of antibacterial compounds: Implications for drug discovery. J. Med. Chem 51, 2871–2878 (2008). [DOI] [PubMed] [Google Scholar]
  • 46.Davis TD, Gerry CJ & Tan DS General platform for systematic quantitative evaluation of small-molecule permeability in bacteria. ACS Chem. Biol 9, 2535–2544 (2014).A platform for prospective, activity-independent analysis of compound accumulation in bacteria was developed. Analysis of a panel of acyl sulfamoyladenosines identified physicochemical properties that correlate with accumulation, enabling design of analogues with increased accumulation.
  • 47.Brown DG, May-Dracka TL, Gagnon MM & Tommasi R Trends and exceptions of physical properties on antibacterial activity for Gram-positive and Gram-negative pathogens. J. Med. Chem 57, 10144–10161 (2014). [DOI] [PubMed] [Google Scholar]
  • 48.Novick RP Microiodometric assay for penicillinase. Biochem. J 83, 236–240 (1962). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Zimmermann W & Rosselet A Function of the outer membrane of Escherichia coli as a permeability barrier to beta-lactam antibiotics. Antimicrob. Agents Chemother 12, 368–372 (1977). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Murakami K & Yoshida T Penetration of cephalosporins and corresponding 1-oxacephalosporins through the outer layer of Gram-negative bacteria and its contribution to antibacterial activity. Antimicrob. Agents Chemother 21, 254–258 (1982). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Nikaido H, Rosenberg EY & Foulds J Porin channels in Escherichia coli: Studies with beta-lactams in intact cells. J Bacteriol 153, 232–240 (1983). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Yoshimura F & Nikaido H Diffusion of β-lactam antibiotics through the porin channels of Escherichia coli K-12. Antimicrob. Agents Chemother 27, 84–92 (1985). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Chopra I & Hacker K Uptake of minocycline by Escherichia coli. J. Antimicrob. Chemother 29, 19–25 (1992). [DOI] [PubMed] [Google Scholar]
  • 54.Diver JM, Piddock LJV & Wise R The accumulation of five quinolone antibacterial agents by Escherichia coli. J. Antimicrob. Chemother 25, 319–333 (1990). [DOI] [PubMed] [Google Scholar]
  • 55.Mortimer PGS & Piddock LJV A comparison of methods used for measuring the accumulation of quinolones by Enterobacteriaceae, Pseudomonas aeruginosa and Staphylococcus aureus. J. Antimicrob. Chemother 28, 639–653 (1991). [DOI] [PubMed] [Google Scholar]
  • 56.Bazile S, Moreau N, Bouzard D & Essiz M Relationship among antibacterial activity, inhibition of DNA gyrase, and intracellular accumulation of 11 fluoroquinolones. Antimicrob. Agents Chemother 36, 2622–2627 (1992).Seminal early study demonstrating integration of compound accumulation levels with biochemical inhibitory activity to predict antibacterial activity in E. coli and P. aeruginosa.
  • 57.McCaffrey C, Bertasso A, Pace J & Georgopapadakou NH Quinolone accumulation in Escherichia coli, Pseudomonas aeruginosa, and Staphylococcus aureus. Antimicrob. Agents Chemother 36, 1601–1605 (1992). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Asuquo AE & Piddock LJ Accumulation and killing kinetics of fifteen quinolones for Escherichia coli, Staphylococcus aureus and Pseudomonas aeruginosa. J Antimicrob Chemother 31, 865–880 (1993). [DOI] [PubMed] [Google Scholar]
  • 59.Piddock LJV, Jin YF & Griggs DJ Effect of hydrophobicity and molecular mass on the accumulation of fluoroquinolones by Staphylococcus aureus. J. Antimicrob. Chemother 47, 261–270 (2001). [DOI] [PubMed] [Google Scholar]
  • 60.Mortimer PGS & Piddock LJV The accumulation of five antibacterial agents in porin-deficient mutants of Escherichia coli. J. Antimicrob. Chemother 32, 195–213 (1993). [DOI] [PubMed] [Google Scholar]
  • 61.Piddock LJV, Jin YF, Ricci V & Asuquo AE Quinolone accumulation by Pseudomonas aeruginosa, Staphylococcus aureus and Escherichia coli. J. Antimicrob. Chemother 43, 61–70 (1999). [DOI] [PubMed] [Google Scholar]
  • 62.Cai H, Rose K, Liang L-H, Dunham S & Stover C Development of a liquid chromatography/mass spectrometry-based drug accumulation assay in Pseudomonas aeruginosa. Anal. Biochem 385, 321–325 (2009). [DOI] [PubMed] [Google Scholar]
  • 63.Ferreras JA, Ryu J-S, Di Lello F, Tan DS & Quadri LEN Small-molecule inhibition of siderophore biosynthesis in Mycobacterium tuberculosis and Yersinia pestis. Nat. Chem. Biol 1, 29–32 (2005). [DOI] [PubMed] [Google Scholar]
  • 64.Huigens RW et al. A ring-distortion strategy to construct stereochemically complex and structurally diverse compounds from natural products. Nat. Chem 5, 195–202 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.O’Connor CJ, Beckmann HSG & Spring DR Diversity-oriented synthesis: Producing chemical tools for dissecting biology. Chem. Soc. Rev 41, 4444–4456 (2012). [DOI] [PubMed] [Google Scholar]
  • 66.Gerry CJ & Schreiber SL Recent achievements and current trajectories of diversity-oriented synthesis. Curr. Opin. Chem. Biol 56, 1–9 (2020). [DOI] [PubMed] [Google Scholar]
  • 67.Prochnow H et al. Subcellular quantification of uptake in Gram-negative bacteria. Anal. Chem. (Washington, DC, U. S.) 91, 1863–1872 (2019). [DOI] [PubMed] [Google Scholar]
  • 68.Zhou Y et al. Thinking outside the “bug”: A unique assay to measure intracellular drug penetration in Gram-negative bacteria. Anal. Chem. (Washington, DC, U. S.) 87, 3579–3584 (2015). [DOI] [PubMed] [Google Scholar]
  • 69.Blair JMA & Piddock LJV How to measure export via bacterial multidrug resistance efflux pumps. mBio 7, e00840 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Stone MRL, Butler MS, Phetsang W, Cooper MA & Blaskovich MAT Fluorescent antibiotics: New research tools to fight antibiotic resistance. Trends Biotechnol. 36, 523–536 (2018). [DOI] [PubMed] [Google Scholar]
  • 71.Tanner L, Denti P, Wiesner L & Warner DF Drug permeation and metabolism in Mycobacterium tuberculosis: Prioritising local exposure as essential criterion in new TB drug development. IUBMB Life 70, 926–937 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Kaščáková S, Maigre L, Chevalier J, Réfrégiers M & Pagès J-M Antibiotic transport in resistant bacteria: Synchrotron UV fluorescence microscopy to determine antibiotic accumulation with single cell resolution. PLoS One 7, e38624 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Vergalli J et al. Spectrofluorimetric quantification of antibiotic drug concentration in bacterial cells for the characterization of translocation across bacterial membranes. Nat. Protoc 13, 1348–1361 (2018). [DOI] [PubMed] [Google Scholar]
  • 74.Paixão L et al. Fluorometric determination of ethidium bromide efflux kinetics in Escherichia coli. J. Biol. Eng 3, 18 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Whittle EE et al. Flow cytometric analysis of efflux by dye accumulation. Front. Microbiol 10, 2319 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Prideaux B et al. High-sensitivity MALDI-MRM-MS imaging of moxifloxacin distribution in tuberculosis-infected rabbit lungs and granulomatous lesions. Anal. Chem 83, 2112–2118 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Prideaux B et al. Mass spectrometry imaging of levofloxacin distribution in TB-infected pulmonary lesions by MALDI-MSI and continuous liquid microjunction surface sampling. Int. J. Mass Spectrom 377, 699–708 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Van Pelt CK et al. A fully automated nanoelectrospray tandem mass spectrometric method for analysis of Caco-2 samples. Rapid Commun. Mass Spectrom 17, 1573–1578 (2003). [DOI] [PubMed] [Google Scholar]
  • 79.Chen CY, Lam BL & Bhattacharya SK Mass spectrometric analyses of phospholipids in the S334ter-3 rat model of retinal degeneration. Mol. Vis 20, 1605–1611 (2014). [PMC free article] [PubMed] [Google Scholar]
  • 80.Yan C et al. Real-time screening of biocatalysts in live bacterial colonies. J. Am. Chem. Soc 139, 1408–1411 (2017). [DOI] [PubMed] [Google Scholar]
  • 81.Wleklinski M et al. High throughput reaction screening using desorption electrospray ionization mass spectrometry. Chem. Sci 9, 1647–1653 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Sinclair I et al. Novel acoustic loading of a mass spectrometer: Toward next-generation high-throughput MS screening. J. Lab. Autom 21, 19–26 (2016). [DOI] [PubMed] [Google Scholar]
  • 83.Heidari Torkabadi H et al. Following drug uptake and reactions inside Escherichia coli cells by Raman microspectroscopy. Biochemistry 53, 4113–4121 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Stratford JP et al. Electrically induced bacterial membrane-potential dynamics correspond to cellular proliferation capacity. Proc. Natl. Acad. Sci. USA 116, 9552–9557 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Cama J, Henney AM & Winterhalter M Breaching the barrier: Quantifying antibiotic permeability across Gram-negative bacterial membranes. J. Mol. Biol 431, 3531–3546 (2019). [DOI] [PubMed] [Google Scholar]
  • 86.Kuhn P, Eyer K, Allner S, Lombardi D & Dittrich PS A microfluidic vesicle screening platform: Monitoring the lipid membrane permeability of tetracyclines. Anal. Chem 83, 8877–8885 (2011). [DOI] [PubMed] [Google Scholar]

RESOURCES