ArticlePDF AvailableLiterature Review

An overview of the antimicrobial resistance mechanisms of bacteria

Authors:

Abstract and Figures

Resistance to antimicrobial agents has become a major source of morbidity and mortality worldwide. When antibiotics were first introduced in the 1900's, it was thought that we had won the war against microorganisms. It was soon discovered however, that the microorganisms were capable of developing resistance to any of the drugs that were used. Apparently most pathogenic microorganisms have the capability of developing resistance to at least some antimicrobial agents. The main mechanisms of resistance are: limiting uptake of a drug, modification of a drug target, inactivation of a drug, and active efflux of a drug. These mechanisms may be native to the microorganisms, or acquired from other microorganisms. Understanding more about these mechanisms should hopefully lead to better treatment options for infective diseases, and development of antimicrobial drugs that can withstand the microorganisms attempts to become resistant.
Content may be subject to copyright.
AIMS Microbiology, 4(3): 482501.
DOI: 10.3934/microbiol.2018.3.482
Received: 18 April 2018
Accepted: 13 June 2018
Published: 26 June 2018
http://www.aimspress.com/journal/microbiology
Review
An overview of the antimicrobial resistance mechanisms of bacteria
Wanda C Reygaert*
Department of Biomedical Sciences, Oakland University William Beaumont School of Medicine,
Rochester, MI, USA
* Correspondence: Email: reygaert@oakland.edu; Tel: +2483702709.
Abstract: Resistance to antimicrobial agents has become a major source of morbidity and mortality
worldwide. When antibiotics were first introduced in the 1900’s, it was thought that we had won the
war against microorganisms. It was soon discovered however, that the microorganisms were capable
of developing resistance to any of the drugs that were used. Apparently most pathogenic
microorganisms have the capability of developing resistance to at least some antimicrobial agents.
The main mechanisms of resistance are: limiting uptake of a drug, modification of a drug target,
inactivation of a drug, and active efflux of a drug. These mechanisms may be native to the
microorganisms, or acquired from other microorganisms. Understanding more about these
mechanisms should hopefully lead to better treatment options for infective diseases, and
development of antimicrobial drugs that can withstand the microorganisms attempts to become
resistant.
Keywords: antimicrobial resistance; β-lactamase; MRSA; ESBL; CRE
1. Introduction
With the discovery of antibiotics, the healthcare community thought that the battle with
infectious diseases was won. However, now that so many bacteria have become resistant to multiple
antimicrobial agents, the war has seemingly escalated in favor of the bacteria. Infectious diseases are
currently a significant cause of morbidity and mortality worldwide. An assessment of these diseases
by the World Health Organization (WHO) found that lower respiratory infection, diarrheal diseases,
HIV/AIDS, and malaria are in the top ten contributors to morbidity and mortality [1]. The advent of
antimicrobial resistance has added significantly to the impact of infectious diseases, in number of
483
AIMS Microbiology Volume 4, Issue 3, 482501.
infections, as well as added healthcare costs. Even though we have a very large number of
antimicrobial agents from which to choose for potential infection therapy, there is documented
antimicrobial resistance to all of these, and this resistance occurs shortly after a new drug is okayed
for use. These concerns prompted the WHO to launch a Global Action Plan on antimicrobial
resistance in 2015 [2].
Antimicrobial agents can be divided into groups based on the mechanism of antimicrobial
activity. The main groups are: agents that inhibit cell wall synthesis, depolarize the cell membrane,
inhibit protein synthesis, inhibit nuclei acid synthesis, and inhibit metabolic pathways in bacteria.
Table 1 gives examples of drugs from each of these groups. It would seem that with such a wide
range of mechanisms we would have better control over the organisms. Unfortunately, improper
stewardship of antimicrobial agents has helped lead to the tremendous resistance issue that we now
face. Factors that have contributed to the growing resistance problem include: increased consumption
of antimicrobial drugs, both by humans and animals; and improper prescribing of antimicrobial
therapy. Overuse of many common antimicrobials agents by physicians may occur because the
choice of drug is based on a combination of low cost and low toxicity [3]. There may also be
improper prescribing of antimicrobials drugs, such as the initial prescription of a broad-spectrum
drug that is unnecessary, or ultimately found to be ineffective for the organism(s) causing the
infection [4]. The danger is that excessive use of antibiotics in humans leads to emergence of
resistant organisms [5,6]. In addition, prior use of antimicrobial drugs puts a patient at risk for
infection with a drug resistant organism, and those patients with the highest exposure to
antimicrobials are most often those who are infected with resistant bacteria [3,7].
For many years antibiotics have been used for treating or preventing disease in raising food
animals. The animal feed often contains antibiotics in amounts that range from below therapeutic
levels to full therapeutic levels, and the antibiotics used come from most of the antimicrobial classes
used in humans. There is evidence to support the idea that feeding antibiotics to animals may result
in development of antimicrobial resistant organisms, and that those resistant organisms may be
transferred to the humans who consume those animals [8,9]. The antimicrobial resistance patterns
seen in the animals reflects the types and amounts of antibiotics given to the animals. The
transmission of antimicrobial resistance from the animals to humans may occur in various ways, with
the direct oral route being the most common (includes eating meat plus ingestion of feces in
contaminated food or water). Another common route is from direct contact with the animals by
humans [9].
Continued increases in antimicrobial resistance have led to fewer treatment options for patients,
and an associated increase in morbidity and mortality. The result is that now we are facing more
severe infections needing more extensive treatment, and longer courses of illness often requiring
extended hospitalization. This has dramatically increased the healthcare costs associated with these
infections. The CDC has reported that a conservative estimate is that over 2 million people in the U.S
become ill each year with antimicrobial resistant infections, resulting in more than 23,000
deaths [10]. The costs attributed to these resistant infections ranges from nearly $7,000 to more than
$29,000 per patient [11]. Studies on the healthcare costs for methicillin-resistant Staphylococcus
aureus (MRSA) infections alone show that in the U.S. the costs are over $18,000 per case, in
Germany the costs are nearly €9,000 per case, and in Switzerland there is an average added cost of
over 100,000 Swiss francs per case [1214]. Various methods of antimicrobial stewardship have
been suggested to stem the increases in resistance. One method involves the use of diversity in
484
AIMS Microbiology Volume 4, Issue 3, 482501.
antimicrobial use. This refers to various components such as not giving a single drug, but using two
or more drugs, either alternatively or concurrently, preferably using drugs with different mechanisms
of action [15,16].
Table 1. Antimicrobial groups based on mechanism of action.
Mechanism of Action
Antimicrobial Groups
Inhibit Cell Wall Synthesis
β-Lactams
Carbapenems
Cephalosporins
Monobactams
Penicillins
Glycopeptides
Depolarize Cell Membrane
Lipopeptides
Inhibit Protein Synthesis
Bind to 30S Ribosomal Subunit
Aminoglycosides
Tetracyclines
Bind to 50S Ribosomal Subunit
Chloramphenicol
Lincosamides
Macrolides
Oxazolidinones
Streptogramins
Inhibit Nucleic Acid Synthesis
Quinolones
Fluoroquinolones
Inhibit Metabolic Pathways
Sulfonamides
Trimethoprim
2. Persistence versus resistance
Before discussing the various aspects of antimicrobial resistance, it would be helpful to
distinguish resistance from persistence. If a bacterium is resistant to a certain antimicrobial agent,
then all of the daughter cells would also be resistant (unless additional mutations occurred in the
meantime). Persistence, however, describes bacterial cells that are not susceptible to the drug, but do
not possess resistance genes. The persistence is undoubtedly due to the fact that some cells in a
bacterial population may be in stationary growth phase (dormant); and most antimicrobial agents
485
AIMS Microbiology Volume 4, Issue 3, 482501.
have no effect on cells that are not actively growing and dividing. These persister cells occur at a rate
of around 1% in a culture that is in stationary phase [17,18]. Figure 1 shows the difference between
persistent and resistant bacterial cells.
Figure 1. Resistance vs. persistence. When bacterial cells are exposed to an
antimicrobial agent there are two possible scenarios. There may be cells present that are
resistant to the antimicrobial agent (A). The non-resistant cells are killed, leaving only
the resistant cells. When the resistant cells are regrown, all of the cells in the culture will
be resistant. The other possibility is that there may be persister cells (dormant, not
resistant) present (B). The non-persister cells are killed, leaving only the persister cells.
When the persister cells are regrown, those cells not in a dormant state will still be
susceptible to the antimicrobial agent.
3. Origins of resistance
Bacteria as a group or species are not necessarily uniformly susceptible or resistant to any
particular antimicrobial agent. Levels of resistance may vary greatly within related bacterial groups.
Susceptibility and resistance are usually measured as a function of minimum inhibitory concentration
(MIC), the minimal concentration of drug that will inhibit growth of the bacteria. The susceptibility
is actually a range of the average MICs for any given drug across the same bacterial species. If that
average MIC for a species is in the resistant part of the range, the species is considered to have
intrinsic resistance to that drug. Bacteria may also acquire resistance genes from other related organisms,
and the level of resistance will vary depending on the species and the genes acquired [19,20].
3.1. Natural resistance
Natural resistance may be intrinsic (always expressed in the species), or induced (the genes are
naturally occurring in the bacteria, but are only expressed to resistance levels after exposure to an
antibiotic). Intrinsic resistance may be defined as a trait that is shared universally within a bacterial
species, is independent of previous antibiotic exposure, and not related to horizontal gene
486
AIMS Microbiology Volume 4, Issue 3, 482501.
transfer [20,21]. The most common bacterial mechanisms involved in intrinsic resistance are reduced
permeability of the outer membrane (most specifically the lipopolysaccharide, LPS, in gram negative
bacteria) and the natural activity of efflux pumps. Multidrug-efflux pumps are also a common
mechanism of induced resistance [21,22]. Table 2 shows some examples of bacteria with intrinsic
antimicrobial resistance.
Table 2. Examples of bacteria with intrinsic resistance.
Organism
Intrinsic resistance
Bacteroides (anaerobes)
aminoglycosides, many β-lactams, quinolones
All gram positives
aztreonam
Enterococci
aminoglycosides, cephalosporins, lincosamides
Listeria monocytogenes
cephalosporins
All gram negatives
glycopeptides, lipopeptides
Escherichia coli
macrolides
Klebsiella spp.
ampicillin
Serratia marcescens
macrolides
Pseudomonas aeruginosa
sulfonamides, ampicillin, 1st and 2nd generation cephalosporins, chloramphenicol, tetracycline
Stenotrophomonas maltophilia
aminoglycosides, β-lactams, carbapenems, quinolones
Acinetobacter spp.
ampicillin, glycopeptides
3.2. Acquired resistance
Acquisition of genetic material that confers resistance is possible through all of the main routes
by which bacteria acquire any genetic material: transformation, transposition, and conjugation
(all termed horizontal gene transferHGT); plus, the bacteria may experience mutations to its own
chromosomal DNA. The acquisition may be temporary or permanent. Plasmid-mediated
transmission of resistance genes is the most common route for acquisition of outside genetic material;
bacteriophage-borne transmission is fairly rare. Certain bacteria such as Acinetobacter spp. are
naturally competent, and therefore capable of acquiring genetic material directly from the outside
environment. Internally, insertion sequences and integrins may move genetic material around, and
stressors (starvation, UV radiation, chemicals, etc.) on the bacteria are common causes of genetic
mutations (substitutions, deletions etc.). Bacteria have an average mutation rate of 1 for every 106 to
109 cell divisions, and most of these mutations will be deleterious to the cell [19,23]. Mutations that
aid in antimicrobial resistance usually only occur in a few types of genes; those encoding drug
targets, those encoding drug transporters, those encoding regulators that control drug transporters,
and those encoding antibiotic-modifying enzymes [20]. In addition, many mutations that confer
antimicrobial resistance do so at a cost to the organism. For example, in the acquisition of resistance
to methicillin in Staphylococcus aureus, the growth rate of the bacteria is significantly decreased [24].
One huge conundrum of antimicrobial resistance is that the use of these drugs leads to increased
resistance. Even the use of low or very low concentrations of antimicrobials (sub-inhibitory) can lead
to selection of high-level resistance in successive bacterial generations, may select for bacteria that
are hypermutatable strains (increase the mutation rate), may increase the ability to acquire resistance
to other antimicrobial agents, and may promote the movement of mobile genetic elements [25].
487
AIMS Microbiology Volume 4, Issue 3, 482501.
4. Mechanisms of resistance
Antimicrobial resistance mechanisms fall into four main categories: (1) limiting uptake of a
drug; (2) modifying a drug target; (3) inactivating a drug; (4) active drug efflux. Intrinsic resistance
may make use of limiting uptake, drug inactivation, and drug efflux; acquired resistance mechanisms
used may be drug target modification, drug inactivation, and drug efflux. Because of differences in
structure, etc., there is variation in the types of mechanisms used by gram negative bacteria versus
gram positive bacteria. Gram negative bacteria make use of all four main mechanisms, whereas gram
positive bacteria less commonly use limiting the uptake of a drug (don’t have an LPS outer
membrane), and don’t have the capacity for certain types of drug efflux mechanisms (refer to the
drug efflux pumps later in this manuscript) [26,27]. Figure 2 illustrates the general antimicrobial
resistance mechanisms.
Figure 2. General antimicrobial resistance mechanisms.
4.1. Limiting drug uptake
As already mentioned, there is a natural difference in the ability of bacteria to limit the uptake
of antimicrobial agents. The structure and functions of the LPS layer in gram negative bacteria
provides a barrier to certain types of molecules. This gives those bacteria innate resistance to certain
groups of large antimicrobial agents [28]. The mycobacteria have an outer membrane that has a high
lipid content, and so hydrophobic drugs such as rifampicin and the fluoroquinolones have an easier
access to the cell, but hydrophilic drugs have limited access [29,30].
Bacteria that lack a cell wall, such as Mycoplasma and related species, are therefore intrinsically
resistant to all drugs that target the cell wall including β-lactams and glycopeptides [31]. Gram
positive bacteria do not possess an outer membrane, and restricting drug access is not as prevalent. In
the enterococci, the fact that polar molecules have difficulty penetrating the cell wall gives intrinsic
resistance to aminoglycosides. Another gram positive bacteria, Staphylococcus aureus, recently has
developed resistance to vancomycin. Of the two mechanisms that S. aureus uses against vancomycin,
488
AIMS Microbiology Volume 4, Issue 3, 482501.
a yet unexplained mechanism allows the bacteria to produce a thickened cell wall which makes it
difficult for the drug to enter the cell, and provides an intermediate resistance to vancomycin. These
strains are designated as VISA strains [30,32].
In those bacteria with large outer membranes, substances often enter the cell through porin
channels. The porin channels in gram negative bacteria generally allow access to hydrophilic
molecules [28,33]. There are two main ways in which porin changes can limit drug uptake: a
decrease in the number of porins present, and mutations that change the selectivity of the porin
channel [29]. Members of the Enterobacteriaceae are known to become resistant due to reducing the
number of porins (and sometime stopping production entirely of certain porins). As a group, these
bacteria reduce porin number as a mechanism for resistance to carbapenems [34,35]. Mutations that
cause changes within the porin channel have been seen in E. aerogenes which then become resistant
to imipenem and certain cephalosporins, and in Neisseria gonorrhoeae which then become resistant
to β-lactams and tetracycline [33,36].
Another widely seen phenomenon in bacterial colonization is the formation of a biofilm by a
bacterial community. These biofilms may contain a predominant organism (such as by Pseudomonas
aeruginosa in the lung), or may consist of a wide variety of organisms, as seen in the biofilm
community of normal flora in the gut. For pathogenic organisms, formation of a biofilm protects the
bacteria from attack by the host immune system, plus provides protection from antimicrobial agents.
The thick, sticky consistency of the biofilm matrix which contains polysaccharides, and proteins and
DNA from the resident bacteria, makes it difficult for antimicrobial agents to reach the bacteria.
Thus, to be effective, much higher concentrations of the drugs are necessary. In addition the bacterial
cells in the biofilm tend to be sessile (slow metabolism rate, slow cell division), so antimicrobials
that target growing, dividing bacterial cells have little effect. An important observation about
biofilms is that it is likely that horizontal transfer of genes is facilitated by the proximity of the
bacterial cells. That means that sharing of antimicrobial resistance genes is potentially easier for
these bacterial communities [3739].
4.2. Modification of drug targets
There are multiple components in the bacterial cell that may be targets of antimicrobial agents;
and there are just as many targets that may be modified by the bacteria to enable resistance to those
drugs. One mechanism of resistance to the β-lactam drugs used almost exclusively by gram positive
bacteria is via alterations in the structure and/or number of PBPs (penicillin-binding proteins). PBPs
are transpeptidases involved in the construction of peptidoglycan in the cell wall. A change in the
number (increase in PBPs that have a decrease in drug binding ability, or decrease in PBPs with
normal drug binding) of PBPs impacts the amount of drug that can bind to that target. A change in
structure (e.g. PBP2a in S. aureus by acquisition of the mecA gene) may decrease the ability of the
drug to bind, or totally inhibit drug binding [24,40].
The glycopeptides (e.g. vancomycin) also work by inhibiting cell wall synthesis, and
lipopeptides (e.g. daptomycin) work by depolarizing the cell membrane. Gram negative bacteria
(thick LPS layer) have intrinsic resistance to these drugs [41]. Resistance to vancomycin has become
a major issue in the enterococci (VREvancomycin-resistant enterococci) and in Staphylococcus
aureus (MRSA). Resistance is mediated through acquisition of van genes which results in changes in the
structure of peptidoglycan precursors that cause a decrease in the binding ability of vancomycin [21,40].
489
AIMS Microbiology Volume 4, Issue 3, 482501.
Daptomycin requires the presence of calcium for binding. Mutations in genes (e.g. mprF) change the
charge of the cell membrane surface to positive, inhibiting the binding of calcium, and therefore,
daptomycin [4244].
Resistance to drugs that target the ribosomal subunits may occur via ribosomal mutation
(aminoglycosides, oxazolidinones), ribosomal subunit methylation (aminoglycosides, macrolides
gram positive bacteria, oxazolidinones, streptogramins) most commonly involving erm genes, or
ribosomal protection (tetracyclines). These mechanisms interfere with the ability of the drug to bind
to the ribosome. The level of drug interference varies greatly among these mechanisms [4547].
For drugs that target nucleic acid synthesis (fluoroquinolones), resistance is via modifications
in DNA gyrase (gram negative bacteriae.g. gyrA) or topoisomerase IV (gram positive bacteria
e.g. grlA). These mutations cause changes in the structure of gyrase and topoisomerase which
decrease or eliminate the ability of the drug to bind to these components [48,49].
For the drugs that inhibit metabolic pathways, resistance is via mutations in enzymes (DHPS
dihydropteroate synthase, DHFRdihydrofolate reductase) involved in the folate biosynthesis
pathway and/or overproduction of resistant DHPS and DHFR enzymes (sulfonamidesDHPS,
trimethoprimDHFR). The sulfonamides and trimethoprim bind to their respective enzymes due to
their being structural analogs of the natural substrates (sulfonamidesp-amino-benzoic acid,
trimethoprimdihydrofolate). The action of these drugs is through competitive inhibition by binding
in the active site of the enzymes. Mutations in these enzymes are most often located in or near the
active site, and resulting structural changes in the enzyme interfere with drug binding while still
allowing the natural substrate to bind [50,51].
4.3. Drug inactivation
There are two main ways in which bacteria inactivate drugs; by actual degradation of the drug,
or by transfer of a chemical group to the drug. The β-lactamases are a very large group of drug
hydrolyzing enzymes. Another drug that can be inactivated by hydrolyzation is tetracycline, via the
tetX gene [45,52].
Drug inactivation by transfer of a chemical group to the drug most commonly uses transfer of
acetyl, phosphoryl, and adenyl groups. There are a large number of transferases that have been
identified. Acetylation is the most diversely used mechanism, and is known to be used against the
aminoglycosides, chloramphenicol, the streptogramins, and the fluoroquinolones. Phosphorylation
and adenylation are known to be used primarily against the aminoglycosides [5255].
4.4. β-lactamases
The most widely used group of antimicrobial agents are the β-lactam drugs. The members of
this drug group all share a specific core structure which consists of a four-sided β-lactam ring.
Resistance to the β-lactam drugs occurs through three general mechanisms: (1) preventing the
interaction between the target PBP and the drug, usually by modifying the ability of the drug to bind
to the PBP (this is mediated by alterations to existing PBPs or acquisition of other PBPs; (2) the
presence of efflux pumps that can extrude β-lactam drugs; (3) hydrolysis of the drug by β-lactamase
enzymes [56,57].
490
AIMS Microbiology Volume 4, Issue 3, 482501.
The β-lactamases (originally called penicillinases and cephalosporinases) inactivate β-lactam
drugs by hydrolyzing a specific site in the β-lactam ring structure, causing the ring to open. The
open-ring drugs are not able to bind to their target PBP proteins. The known β-lactamases are wide-
spread, and the group contains enzymes that are able to inactivate any of the current β-lactam drugs.
The production of β-lactamases is the most common resistance mechanism used by gram negative
bacteria against β-lactam drugs, and the most important resistance mechanism against penicillin and
cephalosporin drugs [45,58].
The β-lactamase enzymes are classified based on their molecular structure and/or functional
characteristics. Structurally they are placed into four main categories (A, B, C, or D). There are three
functional groupings based on the substrate specificity: the cephalosporinases, the serine
β-lactamases, and the metallo (zinc-dependent) β-lactamases. These enzymes may also be commonly
known by their enzyme family; for example: the TEM (named after the first patient) family, the SHV
(sulphydryl variable) family, and the CTX (preferentially hydrolyze cefotaxime) family. Gram
negative bacteria may produce β-lactamases from all four structural groups. The β-lactamases found
in gram positive bacteria are mainly from group A, with some from group B [5962].
These enzymes may be innately found on the bacterial chromosome or may be acquired via a
plasmid. Many members of the Enterobacteriaceae family of gram negative bacteria possess
chromosomal β-lactamase genes. Other gram negative bacteria that possess these include Aeromonas
spp., Acinetobacter spp., and Pseudomonas spp. Plasmid-carried β-lactamase genes are most
commonly found in the Enterobacteriaceae, but may also be found in some species of gram positive
bacteria such as Staphylococcus aureus, Enterococcus faecalis, and Enterococcus faecium [26,59].
The first β-lactamase to be characterized was from E. coli and is chromosomally encoded by the
ampC gene (so named for ampicillin resistance). This gene is constitutively expressed at a low level,
but mutations may result in overexpression of the gene. The AmpC β-lactamases are most effective
against the penicillins and some first generation cephalosporins. There are also many plasmid-borne
β-lactamases which carry a variety of bla genes -lactamase genes). If these β-lactamases confer
resistance to later generation cephalosporins, they were designated as ESBLs, and include members
of the TEM, SHV, CTX-M, and OXA enzyme families. The largest group is the CTX-Ms, which are
most commonly found in E. coli, especially UTI isolates. The ESBL producers may also be resistant to
multiple drug classes, but are generally sensitive to β-lactamase inhibitors. The β-lactamase inhibitors
are structurally similar to β-lactamases, have weak antimicrobial ability alone, but work synergistically
in combination with a β-lactam drug. Commonly used β-lactamase inhibitor/drug pairings include
amoxicillin/clavulanic acid, ampicillin/sulbactam, and piperacillin/tazobactam [56,59,60,6366].
Recently there has been an emergence of β-lactamases that are active against the carbapenems
(carbapenemases), and are found primarily in the Enterobacteriaceae. There are two types of
carbapenemases; the Klebsiella pneumoniae carbapenemases (KPCs), and those designated as
Carbapenem-Resistant Enterobacteriaceae (CRE) enzymes. The KPCs belong to the serine Class A
(functional group 2f) β-lactamases, are resistant to all β-lactam drugs, but may still be affected by
β-lactamase inhibitors. In bacteria that are CRE strains the carbapenemases are all
metallo-β-lactamases (MBLs) in Class B, functional group 3a, and are capable of hydrolyzing all
β-lactam drugs, but are not inactivated by β-lactamase inhibitors. The most widely distributed CREs
are the IMP-1 (for imipenem resistance) and VIM-1 (Verona integron encoded MBL) types. A new
MBL has recently been identified, mainly in strains of E. coli. It has been designated as NDM-1
491
AIMS Microbiology Volume 4, Issue 3, 482501.
(New Delhi MBL). Infections caused by CRE strains have been associated with in-hospital mortality
of up to 71% [5658,67,68].
There is a lot of emphasis on the development of more effective β-lactamase inhibitor drug
combinations, especially in an effort to combat the CRE strains. One newer β-lactamase/drug
combination is ceftolozane/tazobactam, which is mainly used against P. aeruginosa, and shows
promise against gram negative ESBL producing strains. There are also newer β-lactamase inhibitors
which do not have a structure similar to the β-lactam drugs. The first one of these to be approved for
use is avibactam, and it has been approved for use with ceftazidime against gram negative bacteria.
In addition, avibactam is being tested for use with aztreonam against CREs. Another β-lactamase
inhibitor which in non β-lactam structured is vaborbactam. It was approved for use with meropenem
in 2017 against gram negative bacteria causing complicated urinary tract infections (cUTIs).
Unfortunately, so far none of the newer combination drugs is designed to combat the CREs directly.
The metallo-β-lactamases are proving difficult to defeat as these enzymes comprise 3 groups that
vary greatly in structure and mechanisms [6971].
4.5. Drug efflux
Bacteria possess chromosomally encoded genes for efflux pumps. Some are expressed
constitutively, and others are induced or overexpressed (high-level resistance is usually via a
mutation that modifies the transport channel) under certain environmental stimuli or when a suitable
substrate is present. The efflux pumps function primarily to rid the bacterial cell of toxic substances,
and many of these pumps will transport a large variety of compounds (multi-drug [MDR] efflux
pumps). The resistance capability of many of these pumps is influenced by what carbon source is
available [28,72].
Most bacteria possess many different types of efflux pumps. There are five main families of
efflux pumps in bacteria classified based on structure and energy source: the ATP-binding cassette
(ABC) family, the multidrug and toxic compound extrusion (MATE) family, the small multidrug
resistance (SMR) family, the major facilitator superfamily (MFS), and the resistance-nodulation-cell
division (RND) family. Most of these efflux pump families are single-component pumps which
transport substrates across the cytoplasmic membrane. The RND family are multi-component pumps
(found almost exclusively in gram negative bacteria) that function in association with a periplasmic
membrane fusion protein (MFP) and an outer membrane protein (OMP-porin) to efflux substrate
across the entire cell envelope [28,29,73,74]. There are instances where other efflux family members
act with other cellular components as multicomponent pumps in gram negative bacteria. One
member of the ABC family, MacB, works as a tripartite pump (MacAB-TolC) to extrude macrolide
drugs. A member of the MFS, EmrB, works as a tripartite pump (EmrAB-TolC) to extrude nalidixic
acid in E. coli [75,76]. Figure 3 shows the basic structure of the various efflux pump families.
Efflux pumps found in gram positive bacteria may confer intrinsic resistance because of being
encoded on the chromosome. These pumps include members of the MATE and MFS families and
efflux fluoroquinolones. There are also gram positive efflux pumps known to be carried on plasmids.
Currently, the characterized pumps in gram positive bacteria are from the MFS family [7780].
Efflux pumps found in gram negative bacteria are widely distributed and may come from all five of
the families, with the most clinically significant pumps belonging to the RND family [28,79].
492
AIMS Microbiology Volume 4, Issue 3, 482501.
Figure 3. General structure of main efflux pump families.
4.6. ABC transporter family
The ABC efflux family contains both uptake and efflux transport systems. The members of this
family are unique in that they use energy derived from ATP hydrolysis. These pumps transport
amino acids, drugs, ions, polysaccharides, proteins, and sugars. Bacterial ABC transporters usually
are made up of six transmembrane segments (TMS) consisting of α-helices, function in the
membrane in pairs, either as homodimers or heterodimers, and work in conjunction with cytoplasmic
ATPases. These pumps have fairly specific substrates, and there are very few found in clinically
significant bacteria. One notable ABC pump is found in Vibrio cholerae (VcaM), and is capable of
transporting fluoroquinolones and tetracycline [29,81,82].
4.7. MATE transporter family
The MATE efflux family use a Na+ gradient as the energy source, and efflux cationic dyes, and
most efflux fluoroquinolone drugs. Some MATE pumps have also been shown to efflux some
aminoglycosides. Other substrates for these pumps may have unrelated chemical structures. These
pumps are made up of twelve TMS. Very few of these have been characterized in bacteria, and most
are found in gram negative organisms. The first to be characterized was the NorM pump from
chromosomal DNA in Vibrio parahaemolyticus. Other clinically significant bacteria that have NorM
pumps include Neisseria gonorrhoeae and Neisseria meningitidis [73,83,84].
4.8. SMR transporter family
The SMR efflux family are energized by the proton-motive force (H+), are hydrophobic, and
efflux mainly lipophilic cations, so may have a very narrow substrate range. The genes for these
pumps have been found in chromosomal DNA and on plasmids and transposable elements. These
493
AIMS Microbiology Volume 4, Issue 3, 482501.
pumps are made up of four TMS and function as asymmetrical homotetramers. Drug efflux has only
been seen in a few of these pumps, and these most commonly confer resistance to β-lactams and
some aminoglycosides. Examples of SMR pumps are seen in Staphylococcus epidermidis (the SMR
pump which transports ampicillin, erythromycin and tetracycline) and Escherichia coli (the EmeR
pump which transports vancomycin, erythromycin, and tetracycline) [28,29,85,86].
4.9. MFS transporter family
The MFS efflux family catalyze transport via solute/cation (H+ or Na+) symport or solute/H+
antiport. They are involved in the transport of anions, drugs (e.g. macrolides and tetracycline),
metabolites (e.g. bile salts), and sugars. The MFS pumps have the greatest substrate diversity as a
group, yet individually tend to be substrate specific. Examples of this substrate specificity include
Acinetobacter baumannii having separate MFS pumps for erythromycin (SmvA) and
chloramphenicol (CraA and CmlA), and Escherichia coli having separate MFS pumps for
macrolides (MefB), fluoroquinolones (QepA), and trimethoprim (Fsr). There are rare examples of
MFS pumps with a slightly broader substrate specificity, such as in the NorA pump in
Staphylococcus aureus which transports fluoroquinolones and chloramphenicol (these antimicrobials
are the most commonly transported by MFS pumps), or the S. aureus LmrS pump which transports
linezolid, erythromycin, chloramphenicol, and trimethoprim. These pumps are made up of twelve or
fourteen TMS, and over 1,000 have been sequenced in bacteria. Most MFS pumps have been found on
bacterial chromosomes, and nearly 50% of the efflux pumps in E. coli are MFS pumps [28,29,45,87].
4.10. RND transporter family
The RND efflux family members catalyze substrate efflux via a substrate/H+ antiport
mechanism, and are found in numerous gram negative bacteria. They are involved in the efflux of
antibiotics (all are multi-drug transporters), detergents, dyes, heavy metals, solvents, and many other
substrates. Some of these pumps may be drug or drug class specific (Tet pumptetracycline; Mef
pumpmacrolides). Many other RND pumps are capable of transporting a wide range of drugs, such
as the MexAB-OprM pump in Pseudomonas aeruginosa that confers intrinsic resistance to β-lactams,
chloramphenicol, tetracycline, trimethoprim, sulfamethoxazole, and some fluoroquinolones. These
pumps are complex multi-component pumps generally made up of twelve TMS and contain two
large periplasmic loops between TMS 1 and 2, and TMS 7 and 8. In order to function, these pumps
will connect to an OMP and that connection is stabilized by MFPs. Interestingly, these pumps share a
high degree of homology among the RND members. The genes for the RND pumps are generally
organized as an operon. In many, the gene organization is as follows: the gene for the regulator
(which may be transcribed in the opposite direction to the other genes) is adjacent to the MFP gene,
which is adjacent to the main pump gene, and then the OMP gene. Probably the most widely studied
RND pump is the AcrAB-TolC pump in Escherichia coli, which confers resistance to penicillins,
chloramphenicol, macrolides, fluoroquinolones, and tetracycline. The AcrB pump protein contains
two binding pockets which allow the binding of substrates of varying size and chemical
properties [28,29,52,73,74,79,82,88].
Table 3 shows a summary of the antimicrobial resistance mechanisms that are used against the
various drugs.
494
AIMS Microbiology Volume 4, Issue 3, 482501.
Table 3. Antimicrobial resistance mechanisms.
Drug
Drug Uptake Limitation
Drug Inactivation
Efflux Pumps
β-Lactams
Decreased numbers of porins, no outer cell
wall
Gram pos, gram neg—β-lactamases
RND
Carbapenems
Changed selectivity of porin
Cephalosporins
Changed selectivity of porin
Monobactams
Penicillins
Glycopeptides
Thickened cell wall, no outer cell wall
Lipopeptides
Aminoglycosides
Cell wall polarity
Aminoglycoside modifying enzymes, acetylation, phosphorylation,
adenylation
RND
Tetracyclines
Decreased numbers of porins
Antibiotic modification, oxidation
MFS, RND
Chloramphenicol
Acetylation of drug
MFS, RND
Lincosamides
ABC, RND
Macrolides
ABC, MFS, RND
Oxazolidinones
RND
Continued on next page
495
AIMS Microbiology Volume 4, Issue 3, 482501.
Drug
Drug Uptake Limitation
Drug Inactivation
Efflux Pumps
Streptogramins
ABC
Fluoroquinolones
Acetylation of drug
MATE, MFS,
RND
Sulfonamides
RND
Trimethoprim
RND
ABCATP binding cassette family, DHFRdihydrofolate reductase, DHPSdihydropteroate synthase, MATEmultidrug and toxic compound extrusion family, MFSmajor facilitator superfamily,
PBPpenicillin-binding protein, RNDresistance-nodulation-cell division family.
4.11. Impact of antimicrobial resistance for individual bacteria
It is vitally important that we have a clear picture of how many of these resistance mechanisms individual bacteria may have in their arsenals. An
excellent and important example of this is MRSA. The increase in costs for MRSA infections was mentioned previously [1214]. These increased costs
are affected by excess length of hospital stay, increases in number of tests needed, and increased medical and rehabilitation services provided. We also
need to think about the impact on morbidity and mortality caused by MRSA, including significant increases in disease complications. The methicillin
susceptible Staphylococcus aureus (MSSA) and MRSA strains possess the same large number of virulence factors including surface molecules that
promote colonization, and secreted molecules that allow invasion of and damage to host cells. These virulence factors assist the bacteria in causing
multiple types of infections. Since MRSA is well known for infections of skin and related tissues, it is easier to spread the infection from person-to-
person, especially in hospital settings. It has been estimated that the mortality rate for MRSA infections is 23 times higher than that for MSSA strains.
In addition, MRSA strains are frequently multidrug resistant, which limits the impact of available antimicrobial therapy [24,61]. Table 4 is a summary
of the types of resistance mechanisms that S. aureus has in place [61]. There are of course, many pathogens that have similarly diverse arsenals (e.g.
Escherichia coli and Klebsiella pneumoniae) and are becoming resistant to most of the antimicrobial agents available.
496
AIMS Microbiology Volume 4, Issue 3, 482501.
Table 4. Antimicrobial resistance mechanisms in Staphylococcus aureus.
Resistance Mechanism
Antimicrobial Agents
Limiting Drug Uptake
Glycopeptides
Modification of Drug Target
β-lactams
Glycopeptides
Lipopeptides
Aminoglycosides
Tetracyclines
Macrolides
Lincosamides
Oxazolidinones
Streptogramins
Fluoroquinolones
Metabolic Pathway Inhibitors
Inactivation of Drug
β-lactams
Chloramphenicol
Active Drug Efflux
Tetracyclines
Fluoroquinolones
5. Conclusions
The reality is that bacterial are very versatile and adaptive. In order to survive they need to be
capable of dealing with toxic substances. Free living bacteria need to be able to survive toxic attacks
and waste products from other organisms. It should come as no surprise that the bacteria that infect
humans are able to defend themselves against antimicrobial agents. With the alarming increase in
antimicrobial resistance, it is imperative that we find ways to combat these pathogens. Unfortunately,
there is no easy (or cheap, probably) answer to this dilemma. Perhaps we need to rethink how we
design new antimicrobial agents; or maybe start looking to natural substances for clues on what
could be used in this fight.
The mechanisms described here are as varied as are the bacteria themselves. These bacterial
weapons pretty much cover all of the antimicrobial agents that we have, and there are probably more
resistance mechanisms out there that we have not yet characterized. The outlook for fighting
microorganisms might seem to be a little bleak. In 2010 the Infectious Diseases Society of America
(ISDA) requested that by 2020 there would be FDA approval of 10 novel antibiotics. As of 2016,
8 new drugs had been approved, but only one of these is a novel antibiotic. The median time in the
approval pipeline for these drugs was 6.2 years, and the cost per dose of these drugs ranges from
nearly $2,000 to nearly $4,200 [89]. So we will need to work hard, and work quickly to find
remedies for this pressing problem.
Conflict of interest
The author declares that there are no conflicts of interest in this paper.
497
AIMS Microbiology Volume 4, Issue 3, 482501.
References
1. World Health Organization (2014) World Health Statistics 2014.
2. World Health Organization (2015) Global action plan on antimicrobial resistance.
3. Griffith M, Postelnick M, Scheetz M (2012) Antimicrobial stewardship programs: methods of
operation and suggested outcomes. Expert Rev Anti-Infe 10: 63–73.
4. Yu VL (2011) Guidelines for hospital-acquired pneumonia and health-care-associated
pneumonia: a vulnerability, a pitfall, and a fatal flaw. Lancet Infect Dis 11: 248–252.
5. Goossens H (2009) Antibiotic consumption and link to resistance. Clin Microbiol Infec 15 3:12–
15.
6. Pakyz AL, MacDougall C, Oinonen M, et al. (2008) Trends in antibacterial use in US academic
health centers: 2002 to 2006. Arch Intern Med 168: 2254–2260.
7. Tacconelli E (2009) Antimicrobial use: risk driver of multidrug resistant microorganisms in
healthcare settings. Curr Opin Infect Dis 22: 352–358.
8. Landers TF, Cohen B, Wittum TE, et al. (2012) A review of antibiotic use in food animals:
perspective, policy, and potential. Public Health Rep 127: 4–22.
9. Wegener HC (2012) Antibiotic resistance—Linking human and animal health, In: Improving
food safety through a One Health approach, Washington: National Academy of Sciences, 331–
349.
10. Centers for Disease Control and Prevention (CDC) (2013) Antibiotic resistance threats in the
United States, 2013, U.S, Department of Health and Human Services, CS239559-B.
11. Maragakis LL, Perencevich EN, Cosgrove SE (2008) Clinical and economic burden of
antimicrobial resistance. Expert Rev Anti-Infe 6: 751–763.
12. Filice GA, Nyman JA, Lexau C, et al. (2010) Excess costs and utilization associated with
methicillin resistance for patients with Staphylococcus aureus infection. Infect Cont Hosp Ep
31: 365–373.
13. bner C, Hübner NO, Hopert K, et al. (2014) Analysis of MRSA-attributed costs of
hospitalized patients in Germany. Eur J Clin Microbiol 33: 1817–1822.
14. Macedo-Viñas M, De Angelis G, Rohner P, et al. (2013) Burden of methicillin-resistant
Staphylococcus aureus infections at a Swiss University hospital: excess length of stay and costs.
J Hosp Infect 84: 132–137.
15. Pakyz A, Powell JP, Harpe SE, et al. (2008) Diversity of antimicrobial use and resistance in 42
hospitals in the United States. Pharmacotherapy 28: 906–912.
16. Sandiumenge
development of antimicrobial resistance. J Antimicrob Chemoth 57: 1197–1204.
17. Wood TK, Knabel SJ, Kwan BW (2013) Bacterial persister cell formation and dormancy. Appl
Environ Microbiol 79: 7116–7121.
18. Keren I, Kaldalu N, Spoering A, et al. (2004) Persister cells and tolerance to antimicrobials.
FEMS Microbiol Lett 230: 13–18.
19. Coculescu BI (2009) Antimicrobial resistance induced by genetic changes. J Med Life 2: 114–
123.
20. Martinez JL (2014) General principles of antibiotic resistance in bacteria. Drug Discov Today
11: 33–39.
A, Diaz E, Rodriguez A, et al. (2006) Impact of diversity of antibiotic use on the
498
AIMS Microbiology Volume 4, Issue 3, 482501.
21. Cox G, Wright GD (2013) Intrinsic antibiotic resistance: mechanisms, origins, challenges and
solutions. Int J Med Microbiol 303: 287–292.
22. Fajardo A, Martinez-Martin N, Mercadillo M, et al. (2008) The neglected intrinsic resistome of
bacterial pathogens. PLoS One 3: e1619.
23. Davies J, Davies D (2010) Origins and evolution of antibiotic resistance. Microbiol Mol Biol
Rev 74: 417–433.
24. Reygaert WC (2009) Methicillin-resistant Staphylococcus aureus (MRSA): molecular aspects of
antimicrobial resistance and virulence. Clin Lab Sci 22: 115–119.
25. Blázquez J, Couce A, Rodríguez-Beltrán J, et al. (2012) Antimicrobials as promoters of genetic
variation. Curr Opin Microbiol 15: 561–569.
26. Chancey ST, Zähner D, Stephens DS (2012) Acquired inducible antimicrobial resistance in
Gram-positive bacteria. Future Microbiol 7: 959–978.
27. Mahon CR, Lehman DC, Manuselis G (2014) Antimicrobial agent mechanisms of action and
resistance, In: Textbook of Diagnostic Microbiology, St. Louis: Saunders, 254–273.
28. Blair JM, Richmond GE, Piddock LJ (2014) Multidrug efflux pumps in Gram-negative bacteria
and their role in antibiotic resistance. Future Microbiol 9: 1165–1177.
29. Kumar A, Schweizer HP (2005) Bacterial resistance to antibiotics: active efflux and reduced
uptake. Adv Drug Deliver Rev 57: 1486–1513.
30. Lambert PA (2002) Cellular impermeability and uptake of biocides and antibiotics in gram-
positive bacteria and mycobacteria. J Appl Microbiol 92: 46S–54S.
31. Bébéar CM, Pereyre S (2005) Mechanisms of drug resistance in Mycoplasma pneumoniae. Curr
Drug Targets 5: 263–271.
32. Miller WR, Munita JM, Arias CA (2014) Mechanisms of antibiotic resistance in enterococci.
Expert Rev Anti-Infe 12: 1221–1236.
33. Gill MJ, Simjee S, Al-Hattawi K, et al. (1998) Gonococcal resistance to β-lactams and
tetracycline involves mutation in loop 3 of the porin encoded at the penB locus. Antimicrob
Agents Ch 42: 2799–2803.
34. Cornaglia G, Mazzariol A, Fontana R, et al. (1996) Diffusion of carbapenems through the outer
membrane of enterobacteriaceae and correlation of their activities with their periplasmic
concentrations. Microb Drug Resist 2: 273–276.
35. Chow JW, Shlaes DM (1991) Imipenem resistance associated with the loss of a 40 kDa outer
membrane protein in Enterobacter aerogenes. J Antimicrob Chemoth 28: 499–504.
36. Thiolas A, Bornet C, Davin-Régli A, et al. (2004) Resistance to imipenem, cefepime, and
cefpirome associated with mutation in Omp36 osmoporin of Enterobacter aerogenes. Biochem
Bioph Res Co 317: 851–856.
37. Mah TF (2012) Biofilm-specific antibiotic resistance. Future Microbiol 7: 1061–1072.
38. Soto SM (2013) Role of efflux pumps in the antibiotic resistance of bacteria embedded in a
biofilm. Virulence 4: 223–229.
39. Van Acker H, Van Dijck P, Coenye T (2014) Molecular mechanisms of antimicrobial tolerance
and resistance in bacterial and fungal biofilms. Trends Microbiol 22: 326–333.
40. Beceiro A, Tomás M, Bou G (2013) Antimicrobial resistance and virulence: a successful or
deleterious association in the bacterial world? Clin Microbiol Rev 26: 185–230.
41. Randall CP, Mariner KR, Chopra I, et al. (2013) The target of daptomycin is absent form
Escherichia coli and other gram-negative pathogens. Antimicrob Agents Ch 57: 637–639.
499
AIMS Microbiology Volume 4, Issue 3, 482501.
42. Yang SJ, Kreiswirth BN, Sakoulas G, et al. (2009) Enhanced expression of dltABCD is
associated with development of daptomycin nonsusceptibility in a clinical endocarditis isolate of
Staphylococcus aureus. J Infect Dis 200: 1916–1920.
43. Mishra NN, Bayer AS, Weidenmaier C, et al. (2014) Phenotypic and genotypic characterization
of daptomycin-resistant methicillin-resistant Staphylococcus aureus strains: relative roles of
mprF and dlt operons. PLoS One 9: e107426.
44. Stefani S, Campanile F, Santagati M, et al. (2015) Insights and clinical perspectives of
daptomycin resistance in Staphylococcus aureus: a review of the available evidence. Int J
Antimicrob Agents 46: 278–289.
45. Kumar S, Mukherjee MM, Varela MF (2013) Modulation of bacterial multidrug resistance
efflux pumps of the major facilitator superfamily. Int J Bacteriol.
46. Roberts MC (2003) Tetracycline therapy: update. Clin Infect Dis 36: 462–467.
47. Roberts MC (2004) Resistance to macrolide, lincosamide, streptogramin, ketolide, and
oxazolidinone antibiotics. Mol Biotechnol 28: 47–62.
48. Hawkey PM (2003) Mechanisms of quinolone action and microbial response. J Antimicrob
Chemoth 1: 28–35.
49. Redgrave LS, Sutton SB, Webber MA, et al. (2014) Fluoroquinolone resistance: mechanisms,
impact on bacteria, and role in evolutionary success. Trends Microbiol 22: 438–445.
50. Huovinen P, Sundström L, Swedberg G, et al. (1995) Trimethoprim and sulfonamide resistance.
Antimicrob Agents Ch 39: 279–289.
51. Vedantam G, Guay GG, Austria NE, et al. (1998) Characterization of mutations contributing to
sulfathiazole resistance in Escherichia coli. Antimicrob Agents Ch 42: 88–93.
52. Blair JM, Webber MA, Baylay AJ, et al. (2015) Molecular mechanisms of antibiotic resistance.
Nat Rev Microbiol 13: 42–51.
53. Ramirez MS, Tolmasky ME (2010) Aminoglycoside modifying enzymes. Drug Resist Update
13: 151–171.
54. Robicsek A, Strahilevitz J, Jacoby GA, et al. (2006) Fluoroquinolone-modifying enzyme: a new
adaptation of a common aminoglycoside acetyltransferase. Nat Med 12: 83–88.
55. Schwarz S, Kehrenberg C, Doublet B, et al. (2004) Molecular basis of bacterial resistance to
chloramphenicol and florfenicol. FEMS Microbiol Rev 28: 519–542.
56. Pfeifer Y, Cullik A, Witte W (2010) Resistance to cephalosporins and carbapenems in Gram-
negative bacterial pathogens. Int J Med Microbiol 300: 371–379.
57. Bush K, Bradford PA (2016) β-Lactams and β-lactamase inhibitors: an overview. CSH Perspect
Med 6: a02527.
58. Bush K, Jacoby GA (2010) Updated functional classification of β-lactamases. Antimicrob
Agents Ch 54: 969–976.
59. Schultsz C, Geerlings S (2012) Plasmid-mediated resistance in Enterobacteriaceae. Drugs 72:
1–16.
60. Bush K (2013) Proliferation and significance of clinically relevant β-lactamases. Ann NY Acad
Sci 1277: 84–90.
61. Reygaert WC (2013) Antimicrobial resistance mechanisms of Staphylococcus aureus, In:
Microbial pathogens and strategies for combating them: science, technology and education,
Spain: Formatex, 297–310.
500
AIMS Microbiology Volume 4, Issue 3, 482501.
62. Toth M, Antunes NT, Stewart NK, et al. (2016) Class D β-lactamases do exist in Gram-positive
bacteria. Nat Chem Biol 12: 9–14.
63. Jacoby GA (2009) AmpC β-lactamases. Clin Microbiol Rev 22: 161–182.
64. Thomson KS (2010) Extended-spectrum-β-lactamase, AmpC, and carbapenemase issues. J Clin
Microbiol 48: 1019–1025.
65. Lahlaoui H, Khalifa ABH, Mousa MB (2014) Epidemiology of Enterobacteriaceae producing
CTX-M type extended spectrum β-lactamase (ESBL). Med Maladies Infect 44: 400–404.
66. Bevan ER, Jones AM, Hawkey PM (2017) Global epidemiology of CTX-M β-lactamases:
temporal and geographical shifts in genotype. J Antimicrob Chemoth 72: 2145–2155.
67. Bajaj P, Singh NS, Virdi JS (2016) Escherichia coli β-lactamases: what really matters. Front
Microbiol 7: 417.
68. Friedman ND, Tomkin E, Carmeli Y (2016) The negative impact of antibiotic resistance. Clin
Microbiol Infect 22: 416–422.
69. Zhanel GG, Lawson CD, Adam H, et al. (2013) Ceftazidime-Avibactam: a novel
cephalosporin/β-lactamase inhibitor combination. Drugs 73: 159–177.
70. Bush K (2018) Game changers: new β-lactamase inhibitor combinations targeting antibiotic
resistance in gram-negative bacteria. ACS Infect Dis 4: 84–87.
71. Docquier JD, Mangani S (2018) An update on β-lactamase inhibitor discovery and development.
Drug Resist Update 36: 13–29.
72. Villagra NA, Fuentes JA, Jofré MR, et al. (2012) The carbon source influences the efflux pump-
mediated antimicrobial resistance in clinically important Gram-negative bacteria. J Antimicrob
Chemoth 67: 921–927.
73. Piddock LJ (2006) Clinically relevant chromosomally encoded multidrug resistance efflux
pumps in bacteria. Clin Microbiol Rev 19: 382–402.
74. Poole K (2007) Efflux pumps as antimicrobial resistance mechanisms. Ann Med 39: 162–176.
75. Tanabe M, Szakonyi G, Brown KA, et al. (2009) The multidrug resistance efflux complex,
EmrAB from Escherichia coli forms a dimer in vitro. Biochem Bioph Res Co 380: 338–342.
76. Jo I, Hong S, Lee M, et al. (2017) Stoichiometry and mechanistic implications of the MacAB-
TolC tripartite efflux pump. Biochem Bioph Res Co 494: 668–673.
77. Jonas BM, Murray BE, Weinstock GM (2001) Characterization of emeA, a norA homolog and
multidrug resistance efflux pump, in Enterococcus faecalis. Antimicrob Agents Ch 45: 3574–
3579.
78. Truong-Bolduc QC, Dunman PM, Strahilevitz J, et al. (2005) MgrA is a multiple regulator of
two new efflux pumps in Staphylococcus aureus. J Bacteriol 187: 2395–2405.
79. Kourtesi C, Ball AR, Huang YY, et al. (2013) Microbial efflux systems and inhibitors:
approaches to drug discovery and the challenge of clinical implementation. Open Microbiol J 7:
34–52.
80. Costa SS, Viveiros M, Amaral L, et al. (2013) Multidrug efflux pumps in Staphylococcus
aureus: an update. Open Microbiol J 7: 59–71.
81. Lubelski J, Konings WN, Driessen AJ (2007) Distribution and physiology of ABC-type
transporters contributing to multidrug resistance in bacteria. Microbiol Mol Biol Rev 71: 463–
476.
82. Putman M, van Veen HW, Konings WN (2000) Molecular properties of bacterial multidrug
transporters. Microbiol Mol Biol Rev 64: 672–693.
501
AIMS Microbiology Volume 4, Issue 3, 482501.
83. Kuroda T, Tsuchiya T (2009) Multidrug efflux transporters in the MATE family. BBA-Proteins
Proteom 1794: 763–768.
84. Rouquette-Loughlin, C, Dunham SA, Kuhn M, et al. (2003) The NorM efflux pump of Neisseria
gonorrhoeae and Neisseria meningitidis recognizes antimicrobial cationic compounds. J
Bacteriol 185: 1101–1106.
85. Bay DC, Rommens KL, Turner RJ (2008) Small multidrug resistance proteins: a multidrug
transporter family that continues to grow. BBA-Biomembranes 1778: 1814–1838.
86. Yerushalmi H, Lebendiker M, Schuldiner S (1995) EmrE, an Escherichia coli 12-kDa multidrug
transporter, exchanges toxic cations and H+ and is soluble in organic solvents. J Biol Chem 270:
6856–6863.
87. Collu F, Cascella M (2013) Multidrug resistance and efflux pumps: insights from molecular
dynamics simulations. Curr Top Med Chem 13: 3165–3183.
88. Martinez JL, Sánchez MB, Martinez-Solano L, et al. (2009) Functional role of bacterial
multidrug efflux pumps in microbial natural ecosystems. FEMS Microbiol Rev 33: 430–449.
89. Deak D, Outterson K, Powers JH, et al. (2016) Progress in the fight against multidrug-resistant
bacteria? A review of U.S. Food and Drug Administration-approved antibiotics, 2010-2015. Ann
Intern Med 165: 363–372.
© 2018 the Author(s), licensee AIMS Press. This is an open access
article distributed under the terms of the Creative Commons
Attribution License (http://creativecommons.org/licenses/by/4.0)
... Nevertheless, the rise and dispersion of antibiotic-resistant bacteria poses a challenge to global public health (Neu, 1992), and indeed by 2050, an annual mortality toll of 10 million people has been predicted (Balouiri et al., 2016). Different antibiotic resistance mechanisms have been described that are inherent to microorganisms or can be acquired genetically by horizontal gene transfer (Reygaert, 2018). In contrast to other environments, the diversity and relative abundance of ARGs in the atmosphere are understudied Zhou et al., 2023). ...
... A total of 152 different ARGs were found only at the Hospital's main entrance. No ARG was common and shared in all analysed samples, and only the multidrug efflux pump acrB, a widely studied transporter with a high capacity for transporting a vast range of substrates (including antibiotics) (Reygaert, 2018), was shared by two (hospital main entrance and WWTP in situ) out of the four analysed samples ( Figure 4B; Table S3). Regarding the ARG airborne microbial vectors, in the main entrance of the hospital, multiple bacterial species were found carrying ARGs, such as Moraxella, Staphylococcus, Micrococcus, and so forth ( Figure 5). ...
Article
Full-text available
Human activities are a significant contributor to the spread of antibiotic resistance genes (ARGs), which pose a serious threat to human health. These ARGs can be transmitted through various pathways, including air, within the context of One Health. This study used metagenomics to monitor the resistomes in urban air from two critical locations: a wastewater treatment plant and a hospital, both indoor and outdoor. The presence of cell‐like structures was confirmed through fluorescence microscopy. The metagenomic analysis revealed a wide variety of ARGs and a high diversity of antibiotic‐resistant bacteria in the airborne particles collected. The wastewater treatment plant showed higher relative abundances with 32 ARG hits per Gb and m³, followed by the main entrance of the hospital (indoor) with ≈5 ARG hits per Gb and m³. The hospital entrance exhibited the highest ARG richness, with a total of 152 different ARGs classified into nine categories of antibiotic resistance. Common commensal and pathogenic bacteria carrying ARGs, such as Moraxella, Staphylococcus and Micrococcus, were detected in the indoor airborne particles of the hospital. Interestingly, no ARGs were shared among all the samples analysed, indicating a highly variable dynamic of airborne resistomes. Furthermore, the study found no ARGs in the airborne viral fractions analysed, suggesting that airborne viruses play a negligible role in the dissemination of ARGs.
... Shortly after the discovery and release of the first antibiotic -penicillin, [2], penicillinresistant Staphylococcus aureus strains emerged, paving the way to spread of antibacterial resistance, a current worldwide public health problem [3]. Several factors contribute to the origin of antibacterial resistance, such as conjugation, transduction, transformation [4][5][6], acquired genetic mutations [7], intrinsic natural resistance [8], etc. Lebanon shares along with other Eastern Mediterranean countries (low-income countries) a prevalence of resistance higher than several European countries, due to misuse of antibiotic in animal agriculture [9] that spread antibiotic resistant to humans [10], and inadequate wastewater treatment and disposal [11]. The inadequate wastewater treatment of hospitals, industrial and farm sewage lead to an increase in the rates of antibiotic resistance in E. coli. ...
... AUBRB 02 stability was tested against a wide range of pH's (2)(3)(4)(5)(6)(7)(8)(9)(10)(11)(12)(13). Briefly, phage lysates were suspended in adjusted LB broth pH, to yield a pH range of 2-13, then incubated at 37 o C for 2 hours. ...
Preprint
Full-text available
We obtained a new and unique Escherichia phage, AUBRB02, from sewage water in Beirut, Lebanon, as part of this research. AUBRB02 has an incubation period of around 45 minutes, a lysis period of about 10 minutes, and a burst size of around 30 plaque-forming units per cell. The phage exhibited strong biological stability over a pH range of 5.0–9.0 and temperatures ranging from 4°C to 60°C. AUBRB02 was found to have a genome size of 166,871 base pairs and a G+C content of 35.47% using whole-genome sequencing. A comparative analysis revealed that AUBRB02, a newly found phage, shares 93% intergenomic similarity to closest relative in refseq. Functional annotation revealed the presence of 10 tRNA and 262 coding sequences, out of which 123 are categorized as putative proteins. These results indicate that AUBRB02 is a highly infectious virus that belongs to the Tequatrovirus genus. This study is significant reference information that can be used in the development of phage therapy. IMPORTANCE Escherichia coli , a gram-negative bacterium, is a widely distributed pathogen in the natural environment and a frequent cause of illnesses. The extensive utilization of antibiotics has resulted in a rise of clinically resistant strains, posing a substantial obstacle to antimicrobial therapy. This urgent circumstance highlights the necessity for antibiotic substitutes to combat E. coli infections. In this context, we introduce AUBRB02, a novel Escherichia phage isolated from an untreated sewage source in Beirut. Our findings indicate that AUBRB02 is highly lytic, stable against extreme culturing conditions, and has a biofilm elimination capability.
... In particular, virulent strains of Escherichia coli, Acinetobacter baumannii, and Salmonella enterica are prevalent in several bird species inhabiting urban environments (Borges et al. 2017;Morakchi et al. 2017;Murray et al. 2020) and agricultural fields (Dahiru and Enabulele 2015;Rivadeneira et al. 2016;Navarro-Gonzalez et al. 2020). Importantly, these pathogens have developed multidrug resistance (Singh and Mustapha 2013;Reygaert 2018;Martín-Maldonado et al. 2020;Huo et al. 2022). According to the World Health Organization (2013), enteropathogenic E. coli and non-typhoid S. enterica, which cause diarrheal diseases, are responsible for the deaths of approximately 37,000 and 59,000 humans, respectively (Havelaar et al. 2015). ...
Article
Full-text available
The gut microbiome of wild birds contributes to host fitness by supporting nutrient absorption, toxin processing, and immune function. It also fights bacterial pathogens through competitive exclusion and the production of antimicrobial metabolites. This study analyzed the in vitro antagonistic activity of bacteria isolated from the feces of the violet-crowned hummingbird (Ramsomyia violiceps) against strains of Bacillus spp., Escherichia coli, Salmonella enterica, and Acinetobacter baumannii. Mist nets were placed in three parks within the Guadalajara Metropolitan Area. Fecal samples were collected from captured R. violiceps and inoculated into culture media. Bacteria exhibiting antagonist activity were identified using molecular techniques that targeted the V1-V9 region of the 16S rRNA gene. The gut strains Bacillus sp. 1, Bacillus sp. 2, B. altitudinis, B. thuringiensis, and B. subtilis exhibited antagonistic activity against Bacillus cereus, B. tequilensis, and A. baumannii. Pseudomonas putida M5 antagonized Bacillus spp., E. coli, S. enterica, and A. baumannii. This result indicates that some Bacillus spp. and Pseudomonas spp. in the cultivable bacterial assembly of the gut of R. violiceps produce secondary metabolites that can inhibit the growth of both Gram-positive and Gram-negative strains. Since diet plays a determining role in the gut bacterial assemblage of birds, our results suggest that the strains that showed antagonistic activity in vitro could be related to the nectar consumed by the hummingbird. This may help promote the synthesis of antimicrobial compounds as a resistance mechanism.
... 1,2 The prevalence of HAIs is so elevated in modern medicine that the U.S. Centers for Disease Control and Prevention estimate that 1 in 31 healthcare patients currently suffer from an HAI on any chosen day. 3 Moreover, many of these infections are caused by antimicrobial-resistant microorganisms that have adapted through horizontal gene transfer or persister cell survival to no longer respond to standard antibiotic treatment. [4][5][6] The timeline between antibiotic development and documented resistance is diminishing as the medical world struggles to meet infectiontreatment demand. One class of antibiotics, lipoglycopeptides, was first marketed in 2009 and showed promise initially, as it had the dual action of inhibiting peptidoglycan synthesis and destabilizing the bacterial cell membrane. ...
Article
Full-text available
The development of drug‐resistant microorganisms is taking a heavy toll on the biomedical world. Clinical infections are costly and becoming increasingly dangerous as bacteria that once responded to standard antibiotic treatment are developing resistance mechanisms that require innovative treatment strategies. Nitric oxide (NO) is a gaseous molecule produced endogenously that has shown potent antibacterial capabilities in numerous research studies. Its multimechanistic antibacterial methods prevent the development of resistance and have shown potential as an alternative to antibiotics. However, there has yet to be a direct comparison study evaluating the antibacterial properties of NO against antibiotic susceptible and antibiotic‐resistant clinically isolated bacterial strains. Herein, standardized lab and clinically isolated drug‐resistant bacterial strains are compared side‐by‐side for growth and viability following treatment with NO released from S‐nitrosoglutathione (GSNO), an NO donor molecule. Evaluation of growth kinetics revealed complete killing of E. coli lab and clinical strains at 17.5 mM GSNO, though 15 mM displayed >50% killing and significantly reduced metabolic activity, with greater dose dependence for membrane permeability. Clinical P. aeruginosa showed greater susceptibility to GSNO during growth curve studies, but metabolic activity and membrane permeability demonstrated similar effects for 12.5 mM GSNO treatment of lab and clinical strains. MRSA lab and clinical strains exhibited total killing at 17.5 mM treatment, though metabolic activity was decreased, and membrane permeation began at 12.5 mM for both strains. Lastly, both S. epidermidis strains were killed by 15 mM GSNO, with sensitivities in metabolic activity and membrane permeability at 12.5 mM GSNO. The mirrored antibacterial effects seen by the lab and clinical strains of two Gram‐negative and two Gram‐positive bacteria reveal the translational success of NO as an antibacterial therapy and potential alternative to standard antibiotic treatment.
... This can be achieved by extruding antibiotic molecules through efflux pumps and reducing the permeability of the bacterial outer membrane. In contrast, acquired resistance occurs due to the acquisition of external resistance genetic determinants or mutations in chromosomal genes [11]. Acquired resistance caused the bacteria to resist a particular antibiotic's activity to which it was previously susceptible. ...
Article
Full-text available
Antibiotic resistance is a significant global health concern, with multidrug-resistant bacteria emerging in hospitals and community settings. This highlights the presence of antibiotic-resistance genes beyond hospital environments. Bacterial adaptations in response to selective pressures drive the growth of antibiotic resistance. To survive these challenges, bacteria develop various defense mechanisms, including chemical modification of antibiotics, enzyme-catalyzed degradation, altered permeability, efflux, mutation of target sites, and biofilm formation. As a result, bacteria become resistant to most currently available antibiotics. This review provides insights into the molecular mechanisms of antibiotic resistance, which can improve strategies for combating resistance and developing new therapeutic approaches to counter multidrug-resistant bacterial infections.
... Staphylococcal peptidoglycan (PG), the major component of the cell wall, features pentaglycine cross-bridges of nascent peptides that provide mechanical strength and flexibility for bacterial growth in the presence of countering osmotic pressure [33]. Impeding peptidoglycan biosynthesis or destabilizing its integrity can arrest cell growth, as bacterial biological pathways are intricately interlinked, and disruption in one system inevitably affects numerous other functional mechanisms with the cell [34]. In this study, the exposure of MRSA to crude plant extracts resulted in a significantly thickening of the bacterial cell wall. ...
Article
Full-text available
Background Rising resistance to antimicrobials, particularly in the case of methicillin-resistant Staphylococcus aureus (MRSA), represents a formidable global health challenge. Consequently, it is imperative to develop new antimicrobial solutions. This study evaluated 68 Chinese medicinal plants renowned for their historical applications in treating infectious diseases. Methods The antimicrobial efficacy of medicinal plants were evaluated by determining their minimum inhibitory concentration (MIC) against MRSA. Safety profiles were assessed on human colorectal adenocarcinoma (Caco-2) and hepatocellular carcinoma (HepG2) cells. Mechanistic insights were obtained through fluorescence and transmission electron microscopy (FM and TEM). Synergistic effects with vancomycin were investigated using the Fractional Inhibitory Concentration Index (FICI). Results Rheum palmatum L., Arctium lappa L. and Paeonia suffructicosaas Andr. have emerged as potential candidates with potent anti-MRSA properties, with an impressive low MIC of 7.8 µg/mL, comparable to the 2 µg/mL MIC of vancomycin served as the antibiotic control. Crucially, these candidates demonstrated significant safety profiles when evaluated on Caco-2 and HepG2 cells. Even at 16 times the MIC, the cell viability ranged from 83.3% to 95.7%, highlighting their potential safety. FM and TEM revealed a diverse array of actions against MRSA, such as disrupting the cell wall and membrane, interference with nucleoids, and inducing morphological alterations resembling pseudo-multicellular structures in MRSA. Additionally, the synergy between vancomycin and these three plant extracts was evident against MRSA (FICI < 0.5). Notably, aqueous extract of R. palmatum at 1/4 MIC significantly reduced the vancomycin MIC from 2 µg/mL to 0.03 µg/mL, making a remarkable 67-fold decrease. Conclusions This study unveil new insights into the mechanistic actions and pleiotropic antibacterial effectiveness of these medicinal plants against resistant bacteria, providing robust evidence for their potential use as standalone or in conjunction with antibiotics, to effectively combat antimicrobial resistance, particularly against MRSA.
... However, bacteria have adapted to evolve various biochemical mechanisms to evade these antimicrobial targets, such as the generation of biofilms, excretion of efflux pumps, changes in the permeability of their membrane, and cell wall. Resistant pathogens can destroy the drug target through enzymatic inactivation, such as β-lactamase enzymes that damage β-lactam antibiotics [7][8][9]. An example of such antibiotics is amoxicillin, a penicillin derivative whose primary function is to inhibit the connection between the linear peptidoglycan chains forming the bacteria cell wall [10]. ...
Article
Full-text available
Antibacterial resistance has been associated with significant morbidity and mortality, leading to increased costs due to prolonged hospitalization. In this study, we report the synthesis of amoxicillin-functionalized gold nanoparticles (AuNPs@amox) through the chemical reduction of [0.01 M] HAuCl4 with [0.001 M] of amoxicillin in a single step. Amoxicillin, a β-lactam antibiotic can reduce to Au⁰ through the protonated amino group present in its molecular structure. Multidrug-resistant bacteria, Staphylococcus aureus and Acinetobacter baumannii were used. Both were isolated from clinical samples and were resistant to Amoxicillin, so it was chosen for functionalization to evaluate its effect on the resistant bacteria. It was obtaining AuNPs@amox with a bimodal size of 9 ± 2 nm, smaller ones of 1 ± 0.6 nm, and amoxicillin loading of 27.5% on their surface. X-ray photoelectron spectroscopy (XPS) analysis revealed an electrostatic bond between sulfur and the amino group of the antibiotic. Colloidal stability was evaluated through zeta potential and hydrodynamic diameter in trypticase soy broth (0.5X), resulting in -14.1 mV and 138 ± 10 nm, respectively. The AuNPS@amox exhibits antibacterial properties against the resistant strains Acinetobacter baumannii and Methicillin-resistant S. aureus (MRSA), inhibiting up to 77% and 80%, respectively. Upon internalization of AuNPs@amox into the cell, it induces ruptures in the cell wall, leading to cell lysis. It also disrupted the production of biofilms at 10 µg/mL. This work aims to contribute to developing a system that improves drug transport in resistant bacteria—generating high inhibition and significant damage to the resistance mechanisms of resistant bacteria, using very low concentrations to mitigate side effects.
Article
Full-text available
The tendency toward natural herbal products has increased due to the antibiotic resistance developed by microorganisms and the severe side effects of antibiotics commonly used in infectious diseases worldwide. Although antimicrobial studies have been conducted with several species of the Iris genus, this study is the first in the literature to be performed with Iris persica L. subsp. persica aqueous and methanol extracts. In this study, the phenolic content of I. persica was determined by LC–MS/MS analysis, the in vitro antimicrobial activity of I. persica aqueous and methanol extracts was examined, and this study was supported by in silico analysis. Consequently, methanol and aqueous extracts were observed to have inhibitory effects against all tested microorganisms except Candida krusei. Although the MIC values of aqueous extract and methanol extract against Staphylococcus aureus and Klebsiella pneumoniae are the same (22.5 and 11.25 mg/mL, respectively), the inhibitory effect of aqueous extract is generally more potent (MIC value is 11.25 mg/mL for Candida parapsilosis and other bacterial species, and 90 mg/mL for Candida albicans and Candida tropicalis) than that of methanol extract. In silico results showed that hydroxybenzaldeyde, vanillin, resveratrol, isoquercitrin, kaempferol‐3‐glucoside, fisetin, and luteolin were more prone to antifungal activity. Hence, shikimic, gallic, protocatechuic, vanillic, caffeic, o‐coumaric, trans‐ferulic, sinapic acids, and hesperidin were more prone to antibacterial activity. In vitro and in silico results show that the antibacterial activity of our extracts may be higher than the antifungal activity. This preliminary study indicates the anti‐infective potential of I. persica extracts and their usability in medicine and pharmacology.
Article
Stenotrophomonas maltophilia, Achromobacter xylosoxidans, and Burkholderia cenocepacia are considered emerging pathogens classified as a public health problem due to extensive antimicrobial resistance. Therefore, the discovery of new therapeutic strategies has become crucial. This study aimed to evaluate the antimicrobial activity of gallic acid and methyl gallate against non-fermenting bacteria. The study included five clinical isolates of Stenotrophomonas maltophilia, Achromobacter xylosoxidans, and Burkholderia cenocepacia. The minimum inhibitory concentrations of gallic acid and methyl gallate were determined by the broth microdilution method. Growth curves, metabolic activity, and biofilm formation of each bacterial strain in the presence or absence of phenolic compounds were performed. Finally, the therapeutic efficacy of the compounds was evaluated using an in vivo model. Gallic acid and methyl gallate showed antibacterial activity against bacterial strains in a concentration range of 64 to 256 µg/mL, both compounds reduced bacterial growth and metabolic activity of the strains, even at subinhibitory concentrations. Only, methyl gallate exhibited activity to inhibit the formation of bacterial biofilms. Moreover, gallic acid and methyl gallate increased larval survival by up to 60% compared to 30% survival of untreated larvae in a bacterial infection model in Galleria mellonella. Our results highlight the potential of gallic acid and methyl gallate as therapeutic alternatives for infections by emerging non-fermentative bacteria.
Article
Full-text available
A weak antibiotic pipeline and the increase in drug-resistant pathogens have led to calls for more new antibiotics. Eight new antibiotics were approved by the U.S. Food and Drug Administration (FDA) between January 2010 and December 2015: ceftaroline, fidaxomicin, bedaquiline, dalbavancin, tedizolid, oritavancin, ceftolozane-tazobactam, and ceftazidime-avibactam. This study evaluates the development course and pivotal trials of these antibiotics for their innovativeness, development process, documented patient outcomes, and cost. Data sources were FDA approval packages and databases (January 2010 to December 2015); the Red Book (Truven Health Analytics); Orange Book: Approved Drug Products with Therapeutic Equivalence Evaluations (FDA); and supplementary information from company filings, press releases, and media reports. Four antibiotics were approved for acute bacterial skin and skin-structure infection. Seven had similar mechanisms of action to those of previously approved drugs. Six were initially developed by small to midsized companies, and 7 are currently marketed by 1 of 3 large companies. The drugs spent a median of 6.2 years in clinical trials (interquartile range [IQR], 5.4 to 8.8 years) and 8 months in FDA review (IQR, 7.5 to 8 months). The median number of patients enrolled in the pivotal trials was 666 (IQR, 553 to 739 patients; full range, 44 to 1005 patients), and median trial duration was 18 months (IQR, 15 to 22 months). Seven drugs were approved on the basis of pivotal trials evaluating noninferiority. One drug demonstrated superiority on an exploratory secondary end point, 2 showed decreased efficacy in patients with renal insufficiency, and 1 showed increased mortality compared with older drugs. Seven of the drugs are substantially more expensive than their trial comparators. Limitations are that future research may show benefit to patients; new drugs from older classes may show superior effectiveness in specific patient populations; and initial U.S. prices for each new antibiotic were obtained from public sources. Recently marketed antibiotics are more expensive but have been approved without evidence of clinical superiority.
Article
Full-text available
Escherichia coli strains belonging to diverse pathotypes have increasingly been recognized as a major public health concern. The β-lactam antibiotics have been used successfully to treat infections caused by pathogenic E. coli. However, currently, the utility of β-lactams is being challenged severely by a large number of hydrolytic enzymes – the β-lactamases expressed by bacteria. The menace is further compounded by the highly flexible genome of E. coli, and propensity of resistance dissemination through horizontal gene transfer and clonal spread. Successful management of infections caused by such resistant strains requires an understanding of the diversity of β-lactamases, their unambiguous detection, and molecular mechanisms underlying their expression and spread with regard to the most relevant information about individual bacterial species. Thus, this review comprises first such effort in this direction for E. coli, a bacterial species known to be associated with production of diverse classes of β-lactamases. The review also highlights the role of commensal E. coli as a potential but under-estimated reservoir of β-lactamases-encoding genes.
Article
Full-text available
Production of β-lactamases of one of four molecular classes (A, B, C and D) is the major mechanism of bacterial resistance to β-lactams, the largest class of antibiotics, which have saved countless lives since their inception 70 years ago. Although several hundred efficient class D enzymes have been identified in Gram-negative pathogens over the last four decades, none have been reported in Gram-positive bacteria. Here we demonstrate that efficient class D β-lactamases capable of hydrolyzing a wide array of β-lactam substrates are widely disseminated in various species of environmental Gram-positive organisms. Class D enzymes of Gram-positive bacteria have a distinct structural architecture and employ a unique substrate-binding mode that is quite different from that of all currently known class A, C and D β-lactamases. These enzymes thus constitute a previously unknown reservoir of novel antibiotic-resistance enzymes.
Article
Recent regulatory approvals for the β-lactam inhibitor combinations of ceftazidime–avibactam and meropenem–vaborbactam have provided two novel therapeutic options for the treatment of multidrug-resistant infections caused by Gram-negative bacteria. Most importantly, these combination agents have satisfied an important medical need related to antibiotic-resistant Klebsiella pneumoniae that produce serine carbapenemases, especially the Klebsiella pneumoniae carbapenemase (KPC) enzymes. Both combinations contain non-β-lactam β-lactamase inhibitors of novel chemical classes not previously developed as antibacterial agents, the diazabicyclooctanes and cyclic boronic acid derivatives. Their rapid development and approval programs have spurred a number of similar inhibitor combinations that will need to differentiate themselves for commercial success. Gaps still exist for the treatment of infections caused by multidrug-resistant Pseudomonas aeruginosa, Acinetobacter spp., and metallo-β-lactamase-producing pathogens. Overall, the new β-lactamase inhibitor combinations have infused new life into the search for new antibacterial agents to treat multidrug-resistant bacteria.
Article
Antibiotic resistance, and the emergence of pan-resistant clinical isolates, seriously threatens our capability to treat bacterial diseases, including potentially deadly hospital-acquired infections. This growing issue certainly requires multiple adequate responses, including the improvement of both diagnosis methods and use of antibacterial agents, and obviously the development of novel antibacterial drugs, especially active against Gram-negative pathogens, which represent an urgent medical need. Considering the clinical relevance of both β-lactam antibiotics and β-lactamase-mediated resistance, the discovery and development of combinations including a β-lactamase inhibitor seems to be particularly attractive, despite being extremely challenging due to the enormous diversity, both structurally and mechanistically, of the potential β-lactamase targets. This review will cover the evolution of currently available β-lactamase inhibitors along with the most recent research leading to new β-lactamase inhibitors of potential clinical interest or already in the stage of clinical development.
Article
The MacAB-TolC tripartite efflux pump is involved in resistance to macrolide antibiotics and secretion of protein toxins in many Gram-negative bacteria. The pump spans the entire cell envelope and operates by expelling substances to extracellular space. X-ray crystal and electron microscopic structures have revealed the funnel-like MacA hexamer in the periplasmic space and the cylindrical TolC trimer. Nonetheless, the inner membrane transporter MacB still remains ambiguous in terms of its oligomeric state in the functional complex. In this study, we purified a stable binary complex using a fusion protein of MacA and MacB of Escherichia coli, and then supplemented MacA to meet the correct stoichiometry between the two proteins. The result demonstrated that MacB is a homodimer in the complex, which is consistent with results from the recent complex structure using cryo-electron microscopy single particle analysis. Structural comparison with the previously reported MacB periplasmic domain structure suggests a molecular mechanism for regulation of the activity of MacB via an interaction between the MacB periplasmic domain and MacA. Our results provide a better understanding of the tripartite pumps at the molecular level.
Article
Globally, rates of ESBL-producing Enterobacteriaceae are rising. We undertook a literature review, and present the temporal trends in blaCTX-M epidemiology, showing that blaCTX-M-15 and blaCTX-M-14 have displaced other genotypes in many parts of the world. Explanations for these changes can be attributed to: (i) horizontal gene transfer (HGT) of plasmids; (ii) successful Escherichia coli clones; (iii) ESBLs in food animals; (iv) the natural environment; and (v) human migration and access to basic sanitation. We also provide explanations for the changing epidemiology of blaCTX-M-2 and blaCTX-M-27. Modifiable anthropogenic factors, such as poor access to basic sanitary facilities, encourage the spread of blaCTX-M and other antimicrobial resistance (AMR) genes, such as blaNDM, blaKPC and mcr-1. We provide further justification for novel preventative and interventional strategies to reduce transmission of these AMR genes.
Article
β-Lactams are the most widely used class of antibiotics. Since the discovery of benzylpenicillin in the 1920s, thousands of new penicillin derivatives and related β-lactam classes of cephalosporins, cephamycins, monobactams, and carbapenems have been discovered. Each new class of β-lactam has been developed either to increase the spectrum of activity to include additional bacterial species or to address specific resistance mechanisms that have arisen in the targeted bacterial population. Resistance to β-lactams is primarily because of bacterially produced β-lactamase enzymes that hydrolyze the β-lactam ring, thereby inactivating the drug. The newest effort to circumvent resistance is the development of novel broad-spectrum β-lactamase inhibitors that work against many problematic β-lactamases, including cephalosporinases and serine-based carbapenemases, which severely limit therapeutic options. This work provides a comprehensive overview of β-lactam antibiotics that are currently in use, as well as a look ahead to several new compounds that are in the development pipeline.
Article
Antibacterial therapy is one of the most important medical developments of the 20th century; however, the spread of resistance in healthcare settings and in the community threatens the enormous gains made by the availability of antibiotic therapy. Infections caused by resistant bacteria lead to up to two-fold higher rates of adverse outcomes compared with similar infections caused by susceptible strains. These adverse outcomes may be clinical or economic and reflect primarily the failure or delay of antibiotic treatment. The magnitude of these adverse outcomes will be more pronounced as disease severity, strain virulence, or host vulnerability increases. The negative impacts of antibacterial resistance can be measured at the patient level by increased morbidity and mortality, at the healthcare level by increased resource utilization, higher costs, and reduced hospital activity and at the society level by antibiotic treatment guidelines favouring increasingly broad-spectrum empiric therapy. In this review we will discuss the negative impact of antibiotic resistance on patients, the healthcare system and society.