ArticlePDF AvailableLiterature Review

Aminoglycosides: An Overview

Authors:

Abstract and Figures

Aminoglycosides are natural or semisynthetic antibiotics derived from actinomycetes. They were among the first antibiotics to be introduced for routine clinical use and several examples have been approved for use in humans. They found widespread use as first-line agents in the early days of antimicrobial chemotherapy, but were eventually replaced in the 1980s with cephalosporins, carbapenems, and fluoroquinolones. Aminoglycosides synergize with a variety of other antibacterial classes, which, in combination with the continued increase in the rise of multidrug-resistant bacteria and the potential to improve the safety and efficacy of the class through optimized dosing regimens, has led to a renewed interest in these broad-spectrum and rapidly bactericidal antibacterials. © 2016, Cold Spring Harbor Laboratory Press. All rights reserved.
Content may be subject to copyright.
Aminoglycosides: An Overview
Kevin M. Krause,1Alisa W. Serio,1Timothy R. Kane,1and Lynn E. Connolly1,2
1
Achaogen, South San Francisco, California 94080
2
Department of Medicine, Division of Infectious Diseases, University of California, San Francisco,
San Francisco, California 94143
Correspondence: kevinmichaelkrause@gmail.com
Aminoglycosides are natural or semisynthetic antibiotics derived from actinomycetes.
They were among the first antibiotics to be introduced for routine clinical use and several
examples have been approved for use in humans. They found widespread use as first-line
agents in the early days of antimicrobial chemotherapy, but were eventually replaced in the
1980s with cephalosporins, carbapenems, and fluoroquinolones. Aminoglycosides syner-
gize with a variety of other antibacterial classes, which, in combination with the continued
increase in the rise of multidrug-resistant bacteria and the potential to improve the safety and
efficacy of the classthrough optimized dosing regimens, has led to a renewed interest in these
broad-spectrum and rapidly bactericidal antibacterials.
Aminoglycosides are potent, broad-spectrum
antibiotics that act through inhibition of
protein synthesis. The class has been a corner-
stone of antibacterial chemotherapy since strep-
tomycin (Fig. 1) was first isolated from Strepto-
myces griseus and introduced into clinical use
in 1944. Several other members of the class
(Fig. 1) were introduced over the intervening
years including neomycin (1949, S. fradiae),
kanamycin (1957, S. kanamyceticus), gentami-
cin (1963, Micromonospora purpurea), netilmi-
cin (1967, derived from sisomicin), tobramycin
(1967, S. tenebrarius), and amikacin (1972, de-
rived from kanamycin). A shift away from sys-
temic use of the class began in the 1980s with
the availability of the third-generation cepha-
losporins, carbapenems, and fluoroquinolones,
which were perceived to be less toxic and/or
provide broader coverage than the aminoglyco-
sides. However, increasing resistance to these
classes of drugs, combined with more exten-
sive knowledge of the basis of aminoglycoside
resistance, has led to renewed interest in the
legacy aminoglycosides and the development
of novel aminoglycosides such as arbekacin
and plazomicin (Fig. 4). These latter agents
were designed to overcome common amino-
glycoside resistance mechanisms thereby main-
taining potency against multidrug-resistant
(MDR) pathogens.
Additionally, improved dosing schemes
have been developed that attempt to reduce
aminoglycoside toxicity while maintaining effi-
cacy. Specifically, clinical studies have reported
a lower incidence of nephrotoxicity with once-
daily dosing (Nicolau et al. 1995). These data
Editors: Lynn L. Silver and Karen Bush
Additional Perspectives on Antibiotics and Antibiotic Resistance available at www.perspectivesinmedicine.org
Copyright #2016 Cold Spring Harbor Laboratory Press; all rights reserved; doi: 10.1101/cshperspect.a027029
Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029
1
www.perspectivesinmedicine.org
are consistent with the concept that higher dos-
es given less often should reduce the risk of tox-
icity while maintaining and possibly enhancing
efficacy (Drusano et al. 2007). The advantages
of once-daily dosing of aminoglycosides are
now widely accepted and, for many infection
types, this dosing schedule has become the stan-
dard of care (Avent et al. 2011). Additionally,
inhaled delivery of aminog lycosides has become
an area of renewed interest because it allows for
greater local exposure within the lungs with re-
duced systemic toxicity. Inhaled tobramycin is
available in the European Union (EU) and the
United States for the treatment of patients with
chronic Pseudomonas aeruginosa lung infection
associated with cystic fibrosis (CF), and inhaled
amikacin and arbekacin are in development for
potential use in CF and acute respiratory tract
infections.
SPECTRUM OF ACTIVITY
Aminoglycosides are active against various
Gram-positive and Gram-negative organisms.
Streptomycin Apramycin
Amikacin Neomycin B
Gentamicin C1a
Tobramycin
HO
HO
HO
OH
OH
OH
OH
NH
NH
HN
NH2
H2N
N
O
O
O
O
O
H
N
H
HO
HO O
OO
O
H
H
HN
OH
OH
OH
OH
NH2
H2NNH2
H2N
HO
HO
H2NH2N
H2N
NH2
NH2
O
O
OH
OH
OH
O
O
H2NH2NNH2
NH2HO
OH
HN
O
OO
O
OH
H2N
H2N
OH
NH2
H2NNH
O
OH
OH
OH
OH
HO
HO
HO
O
O
O
O
HO OH
OH
OH
OH
HO
HO
O
O
O
O
O
O
H2NH2N
NH2
NH2
NH2
NH2
O
Figure 1. Structures of representative aminoglycosides, including the atypical aminoglycosides streptomycin and
apramycin, 4,6-substituted AGs tobramycin, gentamcin, and amikacin, and the 4,5-substituted AG neomycin.
The deoxystreptamine or streptidine rings are in bold.
K.M. Krause et al.
2Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029
www.perspectivesinmedicine.org
Aminoglycosides are particularly potent against
members of the Enterobacteriaceae family, in-
cluding Escherichia coli,Klebsiella pneumoniae
and K.oxytoca,Enterobacter cloacae and E.
aerogenes,Providencia spp., Proteus spp., Mor-
ganella spp., and Serratia spp. (Ristuccia and
Cunha 1985; Aggen et al. 2010; Landman et al.
2010). Furthermore, aminoglycosides are ac-
tive against Yersinia pestis (Heine 2015) and
Francisella tularensis (Ika
¨heimo et al. 2000),
the causative agents of plague and tularemia,
respectively. The class also has good activity
against Staphylococcus aureus, including meth-
icillin-resistant and vancomycin-intermediate
and -resistant isolates, P. aeruginosa and to a
lesser extent Acinetobacter baumannii (Ristuc-
cia and Cunha 1985; Karlowsky et al. 2003;
Aggen et al. 2010; Landman et al. 2011).
Many Mycobacterium spp. are also susceptible
to aminoglycosides including Mycobacterium
tuberculosis,M. fortuitum,M. chelonae, and
M. avium (Swenson et al. 1985; Ho et al.
1997).
Active electron transport is required for ami-
noglycoside uptake into cells, so the class inher-
ently lacks activity against anaerobic bacteria
(Kislak 1972; Martin et al. 1972; Ramirez and
Tolmasky 2010). Aminoglycosides are also inac-
tive against most Burkholderia spp. and Steno-
trophomonas spp. as well as Streptococcus spp.
and Enterococcus spp. (Brogden et al. 1976; Va-
kulenko and Mobashery 2003; Brooke 2012;
Podnecky et al. 2015).
Contemporary large-scale surveillance pro-
grams provide an understanding of current
aminoglycoside susceptibility among impor-
tant pathogens associated with common infec-
tion types. A recent surveillance study of Gram-
negative organisms isolated from patients
hospitalized in intensive care units (ICUs) in
the United States and the EU found that ami-
kacin and gentamicin showed good activity
against key Gram-negative pathogens (Sader
et al. 2014). In the United States, 99.5% and
87.9% of E. coli isolates were susceptible to ami-
kacin and gentamicin, respectively, according to
the Clinical and Laboratory Standards Institute
(CLSI) criteria. Likewise, 97.3% and 87.2% of
E. coli isolates from the EU were susceptible to
amikacin and gentamicin, respectively, accord-
ing to the European Committee on Antimicro-
bial Susceptibility Testing (EUCAST) criteria.
Among Klebsiella spp., 94.8% and 92.7% of
U.S. isolates and 90.5% and 83.3% of EU
isolates were susceptible to amikacin and gen-
tamicin, respectively. Amikacin was one of
the few agents that retained activity against P.
aeruginosa (97.3% susceptibility in the United
States and 84.9% in the EU according to CLSI
and EUCAST criteria, respectively). Similarly,
aminoglycoside susceptibility rates were high
among isolates collected from U.S. medical cen-
ters in 2012 (Sader et al. 2015). In this study,
CLSI-based susceptibility rates were 99.0%,
88.2%, and 86.3% among E. coli and 88.2%,
89.2%, and 82.4% against K. pneumoniae for
amikacin, gentamicin, and tobramycin, respec-
tively. However, activity was reduced among
K. pneumoniae carbapenemase (KPC)-produc-
ing K. pneumoniae, which are often resistant to
multiple classes of drugs, for amikacin (42.5%
susceptible), gentamicin (50.0% susceptible),
and tobramycin (25.0% susceptible). Amikacin
was the most active agent tested against P. aer-
uginosa (98% susceptible), followed closely by
tobramycin (90% susceptible) and gentamicin
(88.0% susceptible). Broad potency was also
observed against S. aureus (96.0%, 95.0%, and
76.0% for amikacin, gentamicin, and tobramy-
cin, respectively), but these compounds were
less active against A. baumannii (58.0% sus-
ceptible).
The broad-spectrum activity of amino-
glycosides is enhanced in vitro through synergy
with other classes of antimicrobials. This phe-
nomenon, in which the combined effect of
two antimicrobial agents is greater than the
sum of their individual effects, is particularly
well characterized between aminoglycosides
and cell-wall-active agents such as b-lactams.
In vitro synergy between aminoglycosides
and b-lactams has been observed in both
Gram-negative and -positive organisms, in-
cluding wild-type and MDR isolates, using a
variety of methodologies (Eliopoulos and Eli-
opoulos 1988). These in vitro observations
have helped to stimulate the use of aminogly-
coside-containing combination therapy in the
Aminoglycosides: An Overview
Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029 3
www.perspectivesinmedicine.org
treatment of a number of infection types (see
below).
MECHANISM OF ACTION
Aminoglycosides inhibit protein synthesis by
binding, with high affinity, to the A-site on
the 16S ribosomal RNA of the 30S ribosome
(Kotra et al. 2000). Although aminoglycoside
class members have a different specificity for
different regions on the A-site, all alter its con-
formation. As a result of this interaction, the
antibiotic promotes mistranslation by inducing
codon misreading on delivery of the aminoacyl
transfer RNA. This results in error prone pro-
tein synthesis, allowing for incorrect amino
acids to assemble into a polypeptide that is sub-
sequently released to cause damage to the cell
membrane and elsewhere (Davis et al. 1986;
Mingeot-Leclercq et al. 1999; Ramirez and
Tolmasky 2010; Wilson 2014). Some aminogly-
cosides can also impact protein synthesis by
blocking elongation or by directly inhibiting
initiation (Davis 1987; Kotra et al. 2000; Wilson
2014). The exact mechanism of binding and the
subsequent downstream effects varies by chem-
ical structure, but all aminoglycosides are rap-
idly bactericidal (Davis 1987; Mingeot-Leclercq
et al. 1999) and typically produce a prolonged
postantibiotic effect (PAE) (Zhanel et al. 1991)
The PAE has been shown to be directly related
to the length of time that the bacteria take
to recover from the inhibition of protein syn-
thesis (Stubbings et al. 2006). It is hypothesized
that this is dependent on the eventual disasso-
ciation of the antibiotic from its target and exit
from the cell.
Aminoglycosides are characterized by a core
structure of amino sugars connected via glyco-
sidic linkages to a dibasic aminocyclitol, which
is most commonly 2-deoxystreptamine (Min-
geot-Leclercq et al. 1999). Aminoglycosides
are broadly classified into four subclasses based
on the identity of the aminocyclitol moiety:
(1) no deoxystreptamine (e.g., streptomycin,
which has a streptidine ring); (2) a mono-sub-
stituted deoxystreptamine ring (e.g., apramy-
cin); (3) a 4,5-di-substituted deoxystreptamine
ring (e.g., neomycin, ribostamycin); or (4) a
4,6-di-substituted deoxystreptamine ring (e.g.,
gentamicin, amikacin, tobramycin, and plazo-
micin) (Magnet and Blanchard 2005; Wachino
and Arakawa 2012). Examples of each subclass
are shown in Figure 1. The core structure is
decorated with a variety of amino and hydroxyl
substitutions that have a direct influence on
the mechanisms of action and susceptibility
to various aminoglycoside-modifying enzymes
(AMEs) associated with each of the aminogly-
cosides.
Aminoglycoside entry into bacterial cells is
comprised of three distinct stages, the first of
which increases permeability of the bacterial
membrane, whereas the second and third are
energy-dependent. The first stage involves elec-
trostatic binding of the polycationic aminogly-
coside to the negatively charged components
of the bacterial membrane, such as the phos-
pholipids and teichoic acids of Gram-positive
organisms and the phospholipids and lipopoly-
saccharide (LPS) of Gram-negative organisms,
followed by displacement of magnesium ions
(Davis 1987; Taber et al. 1987; Ramirez and Tol-
masky 2010). These cations are responsible for
cross bridging and stabilization of the lipid
components of the bacterial membrane and
their removal leads to disruption of the outer
membrane, enhanced permeability, and initia-
tion of aminoglycoside uptake (Hancock et al.
1981, 1991; Hancock 1984; Ramirez and Tol-
masky 2010). This phenomenon facilitates
entry into the cytoplasm via a slow, energy-de-
pendent, electron-transport-mediated process
(Kislak 1972; Martin et al. 1972; Davis 1987;
Taber et al. 1987; Ramirez and Tolmasky
2010). Inhibition of protein synthesis and mis-
translation of proteins occurs once aminogly-
coside molecules access the cytoplasm. These
mistranslated proteins insert into and cause
damage to the cytoplasmic membrane itself
and facilitate subsequent aminoglycoside entry
(Nichols and Young 1985; Davis et al. 1986).
This then leads to rapid uptake of additional
aminoglycoside molecules into the cytoplasm,
increased inhibition of protein synthesis, mis-
translation, and accelerated cell death (Davis
et al. 1986; Davis 1987; Taber et al. 1987; Ra-
mirez and Tolmasky 2010).
K.M. Krause et al.
4Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029
www.perspectivesinmedicine.org
MECHANISMS OF AMINOGLYCOSIDE
RESISTANCE
Aminoglycoside resistance takes many different
forms including enzymatic modification, target
site modification via an enzyme or chromosom-
al mutation, and efflux. Each of these mecha-
nisms has varying effects on different members
of the class and often multiple mechanisms are
involved in any given resistant isolate. Resis-
tance to aminoglycosides via target site muta-
tions has not been observed because nearly all
prokaryotes, with the exception of Mycobacteri-
um spp. (Bercovier et al. 1986) and Borrelia spp.
(Schwartz et al. 1992), encode multiple copies
of rRNA. Although contemporary large-scale
surveillance programs provide an understand-
ing of phenotypic aminoglycoside resistance
among important pathogens, these studies
have generally not focused on the epidemiology
of specific resistance mechanisms (Jones et al.
2014; Sader et al. 2015).
Enzymatic Drug Modification
AMEs are often found on plasmids containing
multiple resistance elements, including other
AMEs or b-lactamases. The mobility of these
enzymes might be tied to their origins, which
has been hypothesized to be via horizontal gene
transfer from the actinomycetes responsible
for the natural production of aminoglycosides
(Shaw et al. 1993; Ramirez and Tolmasky 2010).
More than 100 AMEs have been described and
are broadly categorized into three groups based
on their ability to acetylate, phosphorylate, or
adenylate amino or hydroxyl groups found at
various positions around the aminoglycoside
core scaffold (Ramirez and Tolmasky 2010).
These modifications decrease the binding affin-
ity of the drug for its target and lead to a loss in
antibacterial potency (Llano-Sotelo et al. 2002).
These three families of AMEs include ami-
noglycoside N-acetyltransferases (abbreviated
AACs), aminoglycoside O-nucleotidyltransfer-
ases (ANTs), and aminoglycoside O-phospho-
transferases (APHs). The classes are further di-
vided into subtypes according to the position
on the aminoglycoside that the enzyme modi-
fies followed by a Roman numeral and, in some
cases, a letter when multiple enzymes exist that
modify the same position (Shaw et al. 1993;
Ramirez and Tolmasky 2010). The major sites
of aminoglycoside modification for the most
common AMEs are shown for kanamycin A in
Figure 2.
Aminoglycoside Acetyltransferases
The aminoglycoside acetyltransferases or AACs
comprise the largest group of AMEs. They are
part of the GCN5-related N-acetyltransferase
(GNAT) superfamily of 10,000 described pro-
teins (Ramirez and Tolmasky 2010). These en-
zymes acetylate amino groups found at various
positions on the aminoglycoside scaffold in
an acetyl-CoA-dependent reaction (Fig. 3A).
There are four main subclasses of this group
OH
OH
OH
HO HO
O
H2NH2NNH2
N1
6′′
4′′
3′′
2
5
AAC(3)AAC(6)
H2N
OO
O
OH
HO
ANT(4)
APH(3)ANT(2′′)
Figure 2. Sites of chemical modification by representative aminoglycoside-modifying enzymes (AMEs) on
kanamycin A.
Aminoglycosides: An Overview
Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029 5
www.perspectivesinmedicine.org
of enzymes whose nomenclature is derived
from the specific amino group that is modified.
Specifically, AAC(1) and AAC(3) target the ami-
no groups found at position 1 and 3, respective-
ly, of the 2-deoxystreptamine ring, whereas
AAC(20) and AAC(60) target amino groups
found at the 20and 60position of the 2,6-di-
deoxy-2,6-diaminoglucose ring (Mingeot-Le-
clercq et al. 1999; Wright and Thompson 1999;
Azucena and Mobashery 2001; Magnet and
Blanchard 2005).
A representative of the AAC family
(AAC(60)-IV) was the first AME found in bac-
teria to be described in the literature (Okamoto
and Suzuki 1965). Since then, .70 members
of the AAC family have been described. Fre-
quently observed class members found among
Gram-negative bacteria include the AAC(60)-1
enzyme that leads to amikacin, netilmicin, and
tobramycin resistance, AAC(3)-IIa, which is
responsible for resistance to gentamicin, tobra-
mycin, and netilmicin, and AAC(3)-I, which
modifies gentamicin (Shaw et al. 1993; Cas-
tanheira et al. 2015). Less common are the
AAC(60)-APH(200) hybrid enzyme responsible
for high-level aminoglycoside resistance in
Enterococcus faecalis (Culebras and Martı
´nez
1999), the chromosomally encoded AAC(60)-
Ii enzyme responsible for intrinsic resistance
to aminoglycosides among E. faecium (Costa
et al. 1993) and the chromosomal AAC(20) en-
zyme found in Mycobacterium spp. and Provi-
OH
OH
O
O
O
OOO
HO
HN
OH
HN
OH
O
O
O
OH
NH2
H2NH2N
A
B
C
H2NNH2
NH2
NH2Acetyl-CoA CoA
HO
HN
NH
NH2
ATP ADP
HO
HO
O
OO
O
NH
O
OH
OH
OH
OH
OH
O
O
P
O
O
O
O
O
H2NH2NH2N
NH2
H2N
H2N
H2N
H2N
H2N
H2N
H2NH2N
H2N
NH2
NH2
HO
HO HO
OH
OH
OH
NN
N
NO
O
O
O
OO
O
O
O
P
HO
HO
HO
OH
OH
OH
OH
OH
ATP PPi
ANT(2′′)
OH
O
OO
O
H2N
O
O
OH
HO HO
HO
OH OH
APH(3)
AAC(3)
OH
Figure 3. Examples of aminoglycoside modification by AMEs. (A) An example of chemical modification of
gentamicin catalyzed by the aminoglycoside acetyltransferase AAC(3). (B) An example of chemical modification
catalyzed by the aminoglycoside phosphotransferase APH(30) on amikacin. (C) Adenylation of the 200 hydroxyl
of kanamycin A catalyzed by the aminoglycoside nucleotidyltransferase ANT(200).
K.M. Krause et al.
6Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029
www.perspectivesinmedicine.org
dencia stuartii (Rather et al. 1993; Aı
´nsa et al.
1997; Macinga and Rather 1999; Ramirez and
Tolmasky 2010). Last, a variant of AAC(60)-Ib
that has acquired the ability to modify fluoro-
quinolones without significantly altering its
activity against aminoglycosides has been iden-
tified in clinical isolates of Gram-negative bac-
teria (Robicsek et al. 2006; Strahilevitz et al.
2009).
Aminoglycoside Phosphotransferases
The second largest group of AMEs is the APHs.
This structurally diverse group of enzymes acts
like kinases in that they catalyze the ATP-depen-
dent phosphorylation of hydroxyl groups found
on aminoglycosides (Fig. 3B). APH enzymes
are functionally and structurally similar to the
serine–threonine and tyrosine kinases found in
eukaryotes (Wright and Thompson 1999). The
modifications made by these enzymes lower
binding affinity to the target by decreasing the
hydrogen bonding potential of aminoglycoside
hydroxyl groups with important rRNA residues.
Most of the .30 described APH enzymes
belong to the APH(30) subfamily (Kim and
Mobashery 2005), although variants that target
the 200 hydroxyl also exist (Ramirez and Tolma-
sky 2010). These enzymes are found in diverse
groups of Gram-negative bacteria, although
APH(30)-IIIa was discovered in S. aureus and
Enterococcus spp. All members of the family
lead to kanamycin and neomycin resistance
with various members of the family also able
to modify a variety of other aminoglycosides,
including amikacin and gentamicin B (Shaw
et al. 1993).
Aminoglycoside Nucleotidyltransferases
The final group of AMEs is the ANTs. These en-
zymes act by adding AMP from an ATP donor
to hydroxyl groups at the 200,3
00,4
0, 6, and 9
positions (Fig. 3C). The most clinically relevant
members of the class include ANT(200 ) and
ANT(40) (Kotra et al. 2000; Magnet and Blan-
chard 2005), which were first described in
K. pneumoniae and S. aureus, respectively (Ben-
veniste and Davies 1971; Le Goffic et al. 1976).
ANT(200) broadly effects the activity of 4,6-
di-substituted aminoglycosides (Gates 1988),
whereas ANT(40) targets kanamycin A, B, and
C, gentamicin A, amikacin, tobramycin, and
neomycin B and C. Other members of this class
include ANT(300), ANT(6), and ANT(9), which
confer resistance to streptomycin and spectino-
mycin (Hollingshead and Vapnek 1985; Mur-
phy 1985; Ounissi et al. 1990).
16S rRNA Methylation
Target site modification leading to aminogly-
coside resistance occurs via the action of 16S
rRNA methyltransferases (RMTs). These en-
zymes modify specific rRNA nucleotide resi-
dues in a manner that blocks aminoglycosides
from effectively binding to their target (Beau-
clerk and Cundliffe 1987; Cundliffe 1989; Wa-
chino and Arakawa 2012). There are two general
classes of RMTs that are characterized by the
specific nucleotide residues that they modify.
These include enzymes that render bacteria
resistant to 4,6-di-substituted aminoglycosides
via methylation of the N7 position of nucleo-
tide G
1405
(Thompson et al. 1985; Beauclerk
and Cundliffe 1987) and those that affect both
4,6- and 4,5-di-substituted aminoglycosides
through methylation of the N1 position of nu-
cleotide A
1408
(Skeggs et al. 1985; Beauclerk and
Cundliffe 1987; Mingeot-Leclercq et al. 1999).
The first clinical case of a pathogen with
an RMT as a mechanis m of aminoglycoside re-
sistance was reported in a P. aeruginosa isolate
from Japan in 2003 (Yokoyama et al. 2003). This
isolate contained a plasmid-encoded RMT,
named RmtA for ribosomal methyltransferase
A. Subsequently, several additional plasmid-
borne RMTs, encoded by the genes armA,
rmtB1, rmtB2,rmtC,rmtD,rmtD2,rmtE,
rmtF,rmtG, and rmtH, have emerged in clinical
isolates that show high-level resistance to mul-
tiple aminoglycosides (Yokoyama et al. 2003;
Doi et al. 2004; Yan et al. 2004; Yamane et al.
2005; Wachino et al. 2006; Wachino and Ara-
kawa 2012). These enzymes modify the G
1405
nucleotide and, thus, impact the activity of all
4,6-di-substituted aminoglycosides (i.e., ami-
kacin, gentamicin, and tobramycin). In 2007,
Aminoglycosides: An Overview
Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029 7
www.perspectivesinmedicine.org
the enzyme NpmAwas discovered encoded on a
plasmid in an aminoglycoside-resistant E. coli
clinical isolate, also from Japan (Wachino et al.
2007). This methyltransferase modifies the
A
1408
nucleotide and, thus, impacts 4,6- and
4,5-di-substituted as well as monosubstituted
aminoglycosides, conferring pan-aminoglyco-
side resistance. To date, there has been only
one additional report of this enzyme in a clinical
isolate (Al Sheikh et al. 2014).
Efflux-Mediated Resistance
Several members of the resistance– nodula-
tion–division (RND) family of efflux systems
have been shown to be involved in intrinsic ami-
noglycoside resistance in various pathogens
(Aires et al. 1999; Westbrock-Wadman et al.
1999; Rosenberg et al. 2000; Magnet et al. 2001;
Hocquet et al. 2003; Islam et al. 2004). In the
opportunistic pathogen P. aeruginosa, intrinsic
low-level resistance to aminoglycosides, tetracy-
cline and erythromycin, is mediated by the ex-
pression of the multiple efflux (Mex) XY-OprM
system. MexXY orthologs are found in other
species of bacteria. For example, members of
the Burkholderia cenopacia complex are often
intrinsically resistant to aminoglycosides via
RND efflux pumps (Buroni et al. 2009). A ho-
mologous transporter in E. coli, AcrD, partici-
pates in efflux of aminoglycosides (Rosenberg
et al. 2000) as does the Acinetobacter drug efflux
(Ade) ABC efflux system in A. baumannii
(Magnet et al. 2001) and the major facilitator
superfamily (MFS) in Mycobacteria spp.
Molecular Epidemiology of Aminoglycoside
Resistance Mechanisms
Comprehensive molecular data regarding the
current prevalence of specific aminoglycoside
resistance mechanisms among common patho-
gens is sparse. One recent study evaluated the
aminoglycoside resistance mechanisms found
among 200 Gram-negative bacilli isolates se-
lected at random from an extensive culture col-
lection (Castanheira et al. 2015). Ninety-nine
Enterobacteriaceae (91.9% AME-positive), 49
A. baumannii (79.6% AME-positive), and 52
P. aeruginosa (63.5% AME positive) with a va-
riety of aminoglycoside resistance profiles were
included. A diversity of AME genes were iden-
tified with the most prevalent being aac(60)-Ib
(n¼75), ant(30)-Ia (n¼51), and aac(3)-IIa
(n¼45), and with 26 isolates harboring more
than one AME gene. In addition, 21 of the iso-
lates were found to carry an RMT including
12 Enterobacteriaceae, seven A. baumannii, and
two P. aeruginosa. Many of these isolates were
also resistant to other common antibiotics used
to treat Gram-negative infections, highlight-
ing the ability of these resistance mechanisms
to spread through conjugation of plasmids and
nonreplicative transposons among bacteria
(Courvalin 1994; Waters 1999; Dzidic and Be-
dekovic
´2003; Feizabadi et al. 2004).
AMEs are commonly found in association
with other key resistance elements, such as
carbapenemases and extended spectrum b-lac-
tamases (ESBLs). Among 50 carbapenem-resis-
tant K. pneumoniae clinical isolates from two
U.S. medical centers (80% possessing KPC-2,
10% possessing KPC-3 with all KPC
þ
isolates
also expressing TEM-1 and SHV-12), 98% of
isolates expressed at least one AME (Alma-
ghrabi et al. 2014). Specifically, 98% were pos-
itive for aac(60)-Ib, 56% positive for aph(30)-Ia,
38% positive for aac(30)-IV, and 2% positive for
ant(200)-Ia. Overall, 40%, 98%, and 16% of the
strains were nonsusceptible to gentamicin, to-
bramycin, and amikacin, respectively. Plazomi-
cin, a novel aminoglycoside in clinical develop-
ment that was designed to evade AME-based
resistance, had MICs that ranged from 0.25
to 1 mg/mL against this set of isolates. AME
characterization in 330 aminoglycoside-resis-
tant clinical Enterobacteriaceae isolates from
Spain revealed the presence of aph(300)-Ib and
ant(300)-Ia genes in 65.4% and 37.5% of the
isolates, consistent with 92% phenotypic resis-
tance to streptomycin found in this strain col-
lection (Miro
´et al. 2013). These isolates were
resistant to other aminoglycosides to varying
degrees, including gentamicin (18.4%), tobra-
mycin (16.9%), and amikacin (1.5%), indicat-
ing the presence of other AMEs; aph(30)-Ia was
found in 13.9% of isolates, aac(3)-IIa in 12.4%,
aac(60)-Ib in 4.2%, ant(200)-Ia in 3.6%, and
K.M. Krause et al.
8Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029
www.perspectivesinmedicine.org
aph(300 )-IIa 1.2% (Miro
´et al. 2013). Under-
scoring the association of AMEs with resistance
elements to other key antibiotic classes, many of
these isolates also produced ESBLs and were
resistant to the fluoroquinolones.
Similar to the AMEs, comprehensive data
describing the prevalence and molecular epide-
miology of RMTs is lacking. A recent literature
review described reports of widespread, global
dissemination of genes encoding RMTs among
isolates from both human and livestock sources
(Wachino and Arakawa 2012). Despite the
widespread detection of RMTs across many re-
gions, the prevalence appears to vary by region
with the highest rates reported in Asia. A
SENTRY surveillance study from the Asia-Pa-
cific region (2007–2008) detected genes encod-
ing RMTs in 6.9% (China), 10.5% (India), 1.5%
(Hong Kong), 6.1% (Korea), and 5.0% (Tai-
wan) of Enterobacteriaceae isolates (Bell et al.
2010). Lower rates (1.3%) of RMTs among
Enterobacteriaceae have been reported from
single institution or local studies from Euro-
pean medical centers (Wachino and Arakawa
2012). Similar to the association between
AMEs and other key resistance mechanisms,
an association between RMTs and specific
b-lactamases has been described. Seventy-six
percent of the RMT containing isolates de-
scribed in the SENTRY surveillance study above
also possessed a CTX-M ESBL (Bell et al. 2010),
and RMTs are frequently found in associa-
tion with the New Delhi metallo-b-lactamase
(NDM) (Berc¸ot et al. 2011; Livermore et al.
2011; Mushtaq et al. 2011; Poirel et al. 2014).
AGENTS IN DEVELOPMENT
Plazomicin is a new aminoglycoside that was
specifically engineered to be resistant to the ac-
tion of the AMEs that are prevalent in key
Gram-negative pathogens (Armstrong and Mil-
ler 2010). It is synthesized from a sisomicin
scaffold that is intrinsically refractory to mod-
ification by APH(30)-III, -VI, and -VII and
ANT(40), which confer amikacin resistance be-
cause of an absence of the 30- and 40-OH groups.
Modification at the N-1 position via addition
of a hydroxylaminobutyric acid substituent
sterically hinders the action of the AAC(3),
ANT(200), and APH(200 ) enzymes, which confer
resistance to gentamicin and tobramycin. Final-
ly, addition of a hydroxyethyl substituent at the
60position inhibits the action of the AAC(60)
enzymes, which confer resistance to a broad
range of agents, including amikacin, tobramy-
cin, and gentamicin (Fig. 4). Importantly, these
modifications to sisomicin do not reduce in-
trinsic potency as has been associated with pre-
vious efforts to protect the 60position and,
as predicted, lead to improved activity against
Enterobacteriaceae (MIC
90
2mg/L) that are
resistant to currently available aminoglycosides
(Nagabhushan et al. 1982; Aggen et al. 2010).
Plazomicin retains vulnerability to modifica-
tion by AAC(20)-I, a chromosomal AME found
in P. stuartii and some mycobacterial species.
However, this enzyme is rare, has not been
found on a mobile element and has not been
shown to have clinical relevance in Mycobac-
terium spp. In addition, plazomicin, like all
Arbekacin
Plazomicin
O
O
HN
O
O
OH
HO
O
HO
NH
OH
OH
HN H2N
NH2
NH2
O
OO
O
OH
O
OH
OH
OH
HO
H2N
H2N
H2N
NH2
NH
NH2
Figure 4. Structures of plazomicin and arbekacin.
Aminoglycosides: An Overview
Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029 9
www.perspectivesinmedicine.org
4,6-linked aminoglycosides, is inactive against
isolates that produce RMTs. As described
above, this mechanism of resistance frequently
travels on mobile genetic elements with NDM-
positive Enterobacteriaceae, and, thus, plazo-
micin is not active against many isolates har-
boring this enzyme (Berc¸ot et al. 2011; Liver-
more et al. 2011; Mushtaq et al. 2011; Poirel
et al. 2014).
Similar to plazomicin, arbekacin, a semi-
synthetic derivative of dibekacin, was specifi-
cally engineered to overcome the action of
a subset of clinically important AMEs (Fig.
4). Arbekacin is stable to the action of
AMEs commonly found in methicillin-resis-
tant S. aureus (MRSA), such as APH, ANT,
and AAC and possesses potent activity against
this clinically important pathogen (Matsu-
moto 2014). Although arbekacin has been
approved for use in Japan since 1990 for the
treatment of sepsis and pneumonia caused
by MRSA, this molecule also retains potent
activity against key Gram-negative pathogens,
including MDR strains, and has more recently
gained attention as a potential therapy for in-
fections caused by these organisms.In a recent
surveillance study of isolates from hospitalized
patients with pneumonia, arbekacin was the
most potent aminoglycoside tested against
ESBL-producing E. coli and was slightly more
active than amikacin and tobramycin against
ESBL- and KPC-expressing K. pneumoniae.
Similarly, arbekacin possessed greater activity
against P. aeruginosa and Acinetobacter spp.,
including MDR isolates, than the other ami-
noglycosides tested (amikacin, gentamicin,
and tobramcin) (Sader et al. 2015). Because
of its broad-spectrum in vitro activity against
both MDR Gram-positive and Gram-negative
pathogens, arbekacin is currently under de-
velopment as an inhalational agent for the
treatment of mechanically ventilated patients
with bacterial pneumonia (clinicaltrials.gov/
ct2/show/NCT02459158) and is under study
at Walter Reed Army Medical Hospital for
the treatment of patients with infections caused
by MDR organisms that have limited treat-
ment options (clinicaltrials.gov/ct2/show/
NCT01659515).
PHARMACOKINETICS AND
PHARMACODYNAMICS
Aminoglycosides are poorly absorbed via the
gastrointestinal (GI) tract and are, thus, admin-
istered via the intravenous or intramuscular
route (Ramirez and Tolmasky 2010; Craig
2011). The volume of distribution for members
of the aminoglycoside class approaches total
body volume, indicating a broad distribution
into tissues, including the lung (Simon et al.
1973). This feature has led to extensive use
of aminoglycosides as part of combination
regimens for the treatment of pneumonia. Ami-
noglycosides are rapidly cleared through the
urinary tract, which also makes these drugs ide-
al for the treatment of urinary tract infections
(Lode et al. 1976; Ramirez and Tolmasky 2010;
Craig 2011).
The relationship between aminoglycoside
pharmacokinetics (PKs) and pharmacodynam-
ics (PDs) has been studied extensively in mice.
The PK/PD variable that is most often correlat-
ed with efficacy of aminoglycosides is the ratio
of area under the concentration– time curve
(AUC) to MIC, although peak concentration
also appears to play a role. The magnitude of
the PK/PD target for aminoglycosides is not as
well defined as it is for other antibiotic classes
because of, until recently, the lack of develop-
ment of new aminoglycosides. Available data
suggests that significant variations in the PK/
PD target exist between species and body site
of infection. For example, an AUC/MIC target
of 100 was reportedly associated with a 1- to 2-
log
10
kill in the mouse neutropenic thigh infec-
tion model with amikacin and K. pneumoniae
(Craig 2011), whereas this same group reported
better efficacy at the same dose level and w ith the
same strain in the mouse lung infection model
(Craig et al. 1991). These results were potentially
a result of a longer measured PAE in the lung
compared with the thigh (Craig et al. 1991).
AUC /MIC thresholds have been correlated
to efficacy in patients in whom extensive PK
sampling was also conducted. These case re-
ports and reviews describe a variety of different
ways that the AUC/MIC ratio might be used to
predict outcomes in patients. Smith et al. (2001)
K.M. Krause et al.
10 Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029
www.perspectivesinmedicine.org
reviewed 23 patients with intra-abdominal in-
fections (n¼16) or lower respiratory infec-
tions (n¼7) treated with tobramycin infused
.30 min to achieve a C
max
of between 4 and
10 mg/L. A PK/PD model constructed from
these patients data showed that improved effi-
cacy was observed when an AUC/MIC ratio of
110 was achieved compared with AUC/MIC
ratios of ,110 (80% clinical cure rate compared
with 47%). Other investigators have found sim-
ilar correlations between exposure and efficacy.
Jacobs (2001) describes the need to achieve an
AUC /MIC ratio of 25 for less severe infections
or in immune competent patients and a ratio of
at least 100 in patients with severe infections
and/or those that are immune-compromised.
Zelenitsky et al. (2003) found significantly im-
proved outcomes in patients with bacteremia
treated with gentamicin when AUC/MIC ratios
exceeded 70 compared with AUC/MIC ratios
,70 (90% cure vs. 45.5%, respectively). How-
ever, a much stronger correlation with outcome
was noted for C
max
/MIC with 84% and 90%
clinical cure rates for C
max
/MIC values of 4.8 or
8 compared with 0% clinical cure for C
max
/
MIC ,2.9.
There is substantial evidence that adminis-
tering larger doses of aminoglycosides less fre-
quently may be associated with improved out-
comes compared with providing the same total
daily dose over more frequent dosing schedules
(Barclay et al. 1999). This dosing approach,
known as once daily or extended interval dos-
ing, takes advantage of three features of amino-
glycosides—concentration-dependent killing,
rapid elimination, and a prolonged PAE. Larger
doses are thought to increase target body site
concentrations to improve PD while minimiz-
ing potential toxicity through allowance for a
period of time in which there is little or no
drug in circulation. The PAE properties of the
class allow for an extended period of killing after
the drug is cleared from the body and before the
next dose is administered. A number of meta-
analyses of results from clinical trials comparing
once versus multiple daily administration of
aminoglycosides have been published (Barclay
et al. 1999). Overall, the results of these studies
suggest that once daily dosing is associated with
reduced nephrotoxicity and equal, if not slightly
improved, efficacy.
Therapeutic drug management (TDM), the
clinical practice of measuring drug concentra-
tions at designated intervals for use in optimiz-
ing individualized dosage regimens, has further
improved the safety profile of aminoglycosides.
Older studies using multiple daily doses of these
drugs have shown nephrotoxicity rates of 10%
to 20% (Humes et al. 1982; Moore et al. 1984),
whereas lower rates of nephrotoxicity ranging
from 0% to 14% have been reported with
once-daily dosing regimens with dose adjust-
ment guided by the use of TDM (Prins et al.
1993; Nicolau et al. 1995; Murry et al. 1999;
Rybak et al. 1999; Buijk et al. 2002). The advan-
tages of once-daily or extended interval dosing
of aminoglycosides combined with the use of
TDM are now widely accepted and, for many
infection types, this dosing approach has be-
come the standard of care (Avent et al. 2011).
CLINICAL USES OF AMINOGLYCOSIDES
The spectrum of activity, rapid bactericidal ac-
tivity, and favorable chemical and pharmacoki-
netic properties of aminoglycosides make them
a clinically useful class of drugs. Aminoglyco-
sides are used as single agents and in combina-
tion with other antibiotics in both empirical
and definitive therapy for a broad range of in-
dications (Avent et al. 2011; Jackson et al. 2013).
In patients with serious infections caused
by Gram-negative pathogens, the receipt of em-
piric combination therapy containing at least
one antimicrobial agent to which the pathogen
is susceptible leads to lower mortality and
improved outcomes (Tamma et al. 2012). In ad-
dition to helping ensure that the pathogen is
adequately covered by at least one active drug,
the use of empiric aminoglycoside-b-lactam
combination therapy has also been theorized
to contribute to improved outcomes by taking
advantage of the in vitro synergy observed be-
tween these classes and to prevent the emergence
of resistance (Pankuch et al. 2010; Le et al. 2011).
Although clinical data to support the latter two
theoretical benefits of combination therapy are
conflicting (Tamma et al. 2012), aminoglyco-
Aminoglycosides: An Overview
Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029 11
www.perspectivesinmedicine.org
sides are often combined with b-lactams for the
empirical treatment of severe sepsis and certain
nosocomial infections in patients with a high
risk of mortality or when there is concern that
the causative pathogen may be resistant to more
commonly used agents (American Thoracic
Society; Infectious Diseases Society of America
2005; Dellinger et al. 2013). More recently, ami-
noglycosides have increasingly become impor-
tant components of therapy for patients infected
with MDR pathogens, such as carbapenem-
resistant Enterobacteriaceae (CRE), for whom
few treatment options remain. In a retrospective
cohort study of 50 cases of sepsis caused by car-
bapenem- and colistin-resistant K. pneumoniae,
the receipt of gentamicin as part of definitive
therapy was associated with lower mortality
compared with the receipt of non-gentamicin-
containing regimens (Gonzalez-Padilla et al.
2015). The novel aminoglycoside plazomicin
(see section on Agents in Development) is cur-
rently under phase 3 study for the treatment of
serious infections caused by CRE (clinicaltrials
.gov/ct2/show/NCT01970371).
Aminoglycosides are also an important
component of combination therapy for multi-
drug-resistant tuberculosis (MDR-TB) and cer-
tain non-tuberculous mycobacterial (NTM)
infections. Current MDR-TB treatment guide-
lines recommend inclusion of one of the follow-
ing agents during the intensive phase of therapy:
amikacin, kanamycin, streptomycin, or capreo-
mycin, a cyclic peptide antibiotic that is often
considered as an aminoglycoside because of
its mechanism of action. Each of these agents
possesses potent bactericidal activity against
M. tuberculosis (Ho et al. 1997) and the choice
of agent depends on previous injectable use (if
any) and the likelihood of resistance. A meta-
analysis including 32 studies with .9000 treat-
ment episodes did not reveal any clear differenc-
es in efficacy among the available agents (World
Health Organization 2011). Similar to treat-
ment of MDR-TB, combination therapy for
patients with fibrocavitary, severe nodular/
bronchiectatic or macrolide-resistant lung dis-
ease because of the M. avium complex general-
ly includes amikacin or streptomycin (Griffith
et al. 2007). Among the rapidly growing myco-
bacteria, amikacin is the preferred agent for in-
fections because of M. fortuitum or M. abscessus,
whereas tobramycin is the most active agent
against M. chelonae (Griffith et al. 2007).
Aminoglycosides remain the preferred
therapy for certain zoonotic infections such as
plague and tularemia. Although streptomycin
has traditionally been the agent of choice for
these infection types, gentamicin is now widely
used because of the broader availability of this
agent as well as data suggesting similar efficacy
to streptomycin (Mwengee et al. 2006; Snowden
and Stovall 2011). Aminoglycosides are bacter-
icidal against these organisms and the use of
bacteriostatic agents, such as doxycycline or
chloramphenicol has led to treatment failures
(Dennis et al. 2001; Snowden and Stovall 2011).
Inhaled tobramycin therapy in CF patients
with chronic lung infection caused by P. aerugi-
nosa has been shown to improve respiratory
function, decrease hospitalizations, and reduce
systemic antibiotic use (Ramseyet al. 1999), and
has contributed to a significant increase in sur-
vival for these patients (Sawicki et al. 2012).
Inhaled aminoglycosides, alone or as part of
combination therapy, are currently under eval-
uation as adjunctive agents for treatment of ad-
ditional respiratory infection types including
chronic lung infections associated with non-
CF bronchiectasis and refractory NTM infec-
tions of the lung, and for the prevention and/
or treatment of ventilator associated infec-
tions, tracheobronchitis (VAT), and pneumonia
(VAP). Regarding the latter indication, although
definitive data from large randomized trials is
not available, a number of small studies focused
on the prevention or treatment of VAP have
provided encouraging results. A meta-analysis
of eight comparative trials of prophylactic aero-
solized antibiotics (four of which used inhaled
tobramycin or gentamicin) found that ICU-ac-
quired pneumonia was less common in the
group of patients that received antibiotic pro-
phylaxis compared with those who received no
prophylaxis (Falagas et al. 2006). Similarly, a
meta-analysis of five comparative trials of ad-
junctive aerosolized antibiotics in the treatment
of VAP, each of which evaluated an inhaled ami-
noglycoside versus placebo or no therapy, re-
K.M. Krause et al.
12 Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029
www.perspectivesinmedicine.org
vealed that administration of aerosolized anti-
biotics was associated with better treatment
success compared with control in both the
intent-to-treat and clinically evaluable popula-
tions (Ioannidou et al. 2007). The relative inef-
ficiency of drug delivery through the ventilator
circuit is a significant challenge with the use of
aerosolized antibiotics for the treatment of ven-
tilator-associated infections. To overcome this
issue, BAY41-6551, an investigational drug– de-
vice combination of amikacin specifically for-
mulated for inhalation, has been developed.
This drug–device combination is currently
under phase 3 study as adjunctive therapy in
intubated and mechanically ventilated patients
with Gram-negative pneumonia (clinicaltrials
.gov/ct2/show/NCT01799993 and clinicaltrials
.gov/ct2/show/results/NCT00805168).
Because of their poor oral bioavailability,
aminoglycosides are a key element of oropha-
ryngeal or gut decolonization/decontamina-
tion regimens, including those targeting MDR
pathogens. The purpose of these decontamina-
tion regimens is to eradicate potential pathogens
from the oropharynx and digestive tract of pa-
tients at risk for nosocomial or postoperative
infections. Selective digestive decontamination
(SDD) consists of the oropharyngeal and gastric
administration of non-absorbable antibiotics
lacking anaerobic activity (often a polymyxin,
an aminoglycoside, and amphotericin) along
with a short course of systemic antibiotic thera-
py, whereas selective oropharyngeal decon-
tamination (SOD) consists of application of
non-absorbable antibiotics to the oropharynx
alone. More than 50 randomized studies and
10 meta-analyses of SDD/SOD have been pub-
lished. Overall, these data suggest that SDD/
SOD are associated with improved survival in
ICU patients (Price et al. 2014) and SDD is as-
sociated with a reduction in the rate of post-
operative infection, including anastomotic leak-
age, in patients undergoing elective GI surgery
(Abis et al. 2013; Roos et al. 2013). Despite these
successes in the use of SOD and SDD to improve
patient outcomes in the setting of low levels
of antibiotic resistance, controversy remains re-
garding their effectiveness in the setting of high
levels of antibiotic resistance as well as their im-
pact on antibiotic resistance. No relationship
between the use of SDD or SOD and the devel-
opment of antimicrobial resistance has been
shown in individual studies or meta-analyses
in the setting of low antibiotic resistance (Dane-
man et al. 2013; Plantinga and Bonten 2015)
and an international multicenter study of the
effects of SDD and SOD on ICU-level antibiotic
resistance in countries with higher levels of re-
sistance is currently ongoing (clinicaltrials.gov/
ct2/show/NCT02208154).
The ability of paromomycin to bind to eu-
karyotic ribosomes has led to the use of this
agent in the treatment of protozoal infections.
Like other aminoglycosides, oral paromomycin
is poorly absorbed and may be used for the
treatment of noninvasive amebiasis, crypto-
sporidiosis, trichomoniasis, and giardiasis in
patients in whom other agents are contraindi-
cated (Stover et al. 2012). More recently, this
agent has been used to treat both cutaneous
and visceral leishmaniasis. Topical paromomy-
cin yielded a significantly higher cure rate com-
pared with control therapy in patients with
cutaneous leishmaniasis caused by Leishmania
major (Ben Salah et al. 2013). Intramuscular
paromomycin monotherapy was noninferior
to standard amphotericin B therapy in a ran-
domized control trial in patients with visceral
disease in India (Sundar et al. 2007) and short-
course combination regimens containing this
agent were also noninferior to standard therapy
with fewer adverse events (Sundar et al. 2011).
CONCLUSIONS
The aminoglycosides are a critical component
of the current antibacterial arsenal. Their broad
spectrum of activity, rapid bactericidal action,
and favorable chemical and pharmacokinetic
properties make them a clinically useful class
of drugs across numerous infection types, in-
cluding certain protozoal infections. The use
of aminoglycosides waned as a result of the
emergence of other classes of broad-spectrum
agents with improved safety profiles, but the
emergence of MDR pathogens has led to re-
newed interest in this class of drugs. Improved
understanding of the drivers of toxicity and
Aminoglycosides: An Overview
Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029 13
www.perspectivesinmedicine.org
efficacy has led to the implementation of opti-
mized dosing regimens that improve safety
while maintaining efficacy. Increased under-
standing of key aminoglycoside resistance
mechanisms combined with innovative me-
dicinal chemistry approaches have led to the
synthesis and development of novel agents
specifically designed to evade resistance while
maintaining potency against fully susceptible
isolates. Given the dearth of new agents in the
antibiotic pipeline and the ever-increasing spec-
ter of resistance, further optimization of the
aminoglycoside scaffold to generate new agents
with superior potency against MDR pathogens
as well as an improved safety profile is warranted.
REFERENCES
Abis GS, Stockmann HB, van Egmond M, Bonjer HJ, Van-
denbroucke-Grauls CM, Oosterling SJ. 2013. Selective
decontamination of the digestive tract in gastrointestinal
surgery: Useful in infection prevention? A systematic re-
view. J Gastrointest Surg 17: 2172–2178.
Aggen JB, Armstrong ES, Goldblum AA, Dozzo P, Linsell
MS, Gliedt MJ, Hildebrandt DJ, Feeney LA, Kubo A,
Matias RD, et al. 2010. Synthesis and spectrum of the
neoglycoside ACHN-490. Antimicrob Agents Chemother
54: 4636–4642.
´nsa JA, Pe
´rez E, Pelicic V, Berthet FX, Gicquel B, Martı
´nC.
1997. Aminoglycoside 20-N-acetyltransferase genes are
universally present in mycobacteria: Characterization of
the aac(20)-Ic gene from Mycobacterium tuberculosis and
the aac(20)-Id gene from Mycobacterium smegmatis.Mol
Microbiol 24: 431– 441.
Aires JR, Ko
¨hler T, Nikaido H, Ple
´siat P. 1999. Involvement
of an active efflux system in the natural resistance of
Pseudomonas aeruginosa to aminoglycosides. Antimicrob
Agents Chemother 43: 2624– 2628.
Almaghrabi R, Clancy CJ, Doi Y, Hao B, Chen L, Shields
RK, Press EG, Iovine NM, Townsend BM, Wagener MM,
et al. 2014. Carbapenem-resistant Klebsiella pneumoniae
strains exhibit diversity in aminoglycoside-modifying
enzymes, which exert differing effects on plazomicin
and other agents. Antimicrob Agents Chemother 58:
4443– 4451.
Al Sheikh YA, Marie MA, John J, Krishnappa LG, Dabwab
KH. 2014. Prevalence of 16S rRNA methylase genes
among b-lactamase-producing Enterobacteriaceae clini-
cal isolates in Saudi Arabia. Libyan J Med 9: 24432.
American Thoracic Society; Infectious Diseases Society
of America. 2005. Guidelines for the management
of adults with hospital-acquired, ventilator-associated,
and healthcare-associated pneumonia. Am J Respir Crit
Care Med 171: 388– 416.
Armstrong ES, Miller GH. 2010. Combating evolution with
intelligent design: The neoglycoside ACHN-490. Curr
Opin Microbiol 13: 565– 573.
Avent ML, Rogers BA, Cheng AC, Paterson DL. 2011. Cur-
rent use of aminoglycosides: Indications, pharmacoki-
netics and monitoring for toxicity. Intern Med J 41:
441–449.
Azucena E, Mobashery S. 2001. Aminoglycoside-modifying
enzymes: Mechanisms of catalytic processes and inhibi-
tion. Drug Resist Updat 4: 106–117.
Barclay ML, Kirkpatrick CM, Begg EJ. 1999. Once daily
aminoglycoside therapy. Is it less toxic than multiple dai-
ly doses and how should it be monitored? Clin Pharma-
cokinet 36: 89–98.
Beauclerk AA, Cundliffe E. 1987. Sites of action of two ri-
bosomal RNA methylases responsible for resistance to
aminoglycosides. J Mol Biol 193: 661– 671.
Bell J, Andersson P, Jones R, Turnidge J. 2010. 16S rRNA
methylase containing Enterobacteriaceae in the SENTRY
Asia-Pacific region frequently harbour plasmid-mediat-
ed quinolone resistance CTXM types. 20th European
Congress of Clinical Microbiology and Infectious Diseases
(ECCMID), Abstract O559. Vienna, April 10–13.
Ben Salah A, Ben Messaoud N, Guedri E, Zaatour A, Ben
Alaya N, Bettaieb J, Gharbi A, Belhadj Hamida N, Bouk-
thir A, Chlif S, et al. 2013. Topical paromomycin with or
without gentamicin for cutaneous leishmaniasis. N Engl J
Med 368: 524– 532.
Benveniste R, Davies J. 1971. R-factor mediated gentamicin
resistance: A new enzyme which modifies aminoglyco-
side antibiotics. FEBS Lett 14: 293–296.
Berc¸ ot B, Poirel L, Nordmann P. 2011. Updated multiplex
polymerase chain reaction for detection of 16S rRNA
methylases: High prevalence among NDM-1 producers.
Diagn Microbiol Infect Dis 71: 442– 445.
Bercovier H, Kafri O, Sela S. 1986. Mycobacteria possess a
surprisingly small number of ribosomal RNA genes in
relation to the size of their genome. Biochem Biophys
Res Commun 136: 1136– 1141.
Brogden RN, Pinder RM, SawyerPR, Speight TM, Avery GS.
1976. Tobramycin: A review of its antibacterial and phar-
macokinetic properties and therapeutic use. Drugs 12:
166–200.
Brooke JS. 2012. Stenotrophomonas maltophilia: An emerg-
ing global opportunistic pathogen. Clin Microbiol Rev 25:
2–41.
Buijk SE, Mouton JW, Gyssens IC, Verbrugh HA, Bruining
HA. 2002. Experience with a once-daily dosing program
of aminoglycosides in critically ill patients. Intensive Care
Med 28: 936– 942.
Buroni S, Pasca MR, Flannagan RS, Bazzini S, Milano A,
Bertani I, Venturi V, Valvano MA, Riccardi G. 2009. As-
sessment of three resistance-nodulation-cell division
drug efflux transporters of Burkholderia cenocepacia in
intrinsic antibiotic resistance. BMC Microbiol 9: 200.
Castanheira M, Costello SE, Jones RN, Mendes RE. 2015.
Prevalence of aminoglycoside resistance genes among
contemporary Gram-negative resistant isolates collected
worldwide. 25th European Congress of Clinical Microbri-
ology and Infectious Diseases (ECCMID), Abstract O011.
Copenhagen, April 25– 28.
Cattoir V, Nordmann P. 2009. Plasmid-mediated quinolone
resistance in Gram-negative bacterial species: An update.
Curr Med Chem 16: 1028–1046.
K.M. Krause et al.
14 Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029
www.perspectivesinmedicine.org
Costa Y, Galimand M, Leclercq R, DuvalJ, Courvalin P. 1993.
Characterization of the chromosomal aac(60)-Ii gene
specific for Enterococcus faecium.Antimicrob Agents Che-
mother 37: 1896– 1903.
Courvalin P. 1994. Transfer of antibiotic resistance genes be-
tween Gram-positive and Gram-negative bacteria. Anti-
microb Agents Chemother 38: 1447– 1451.
Craig WA. 2011. Optimizing aminoglycoside use. Crit Care
Clin 27: 107–121.
Craig WA, Redington J, Ebert SC. 1991. Pharmacodynamics
of amikacin in vitro and in mouse thigh and lung infec-
tions. J Antimicrob Chemother 27: 29– 40.
Culebras E, Martı
´nez JL. 1999. Aminoglycoside resistance
mediated by the bifunctional enzyme 60-N-aminogly-
coside acetyltransferase-200-O-aminoglycoside phospho-
transferase. Front Biosci 4: D1– D8.
Cundliffe E. 1989. How antibiotic-producing organisms
avoid suicide. Annu Rev Microbiol 43: 207– 233.
Daneman N, Sarwar S, Fowler RA, Cuthbertson BH,
Group SCS. 2013. Effect of selective decontamination
on antimicrobial resistance in intensive care units: A
systematic review and meta-analysis. Lancet Infect Dis
13: 328– 341.
Davis BD. 1987. Mechanismof bactericidal action of amino-
glycosides. Microbiol Rev 51: 341– 350.
Davis BD, Chen LL, Tai PC. 1986. Misread protein creates
membrane channels: An essential step in the bactericidal
action of aminoglycosides. Proc Natl Acad Sci 83: 6164–
6168.
Dellinger RP, Levy MM, Rhodes A, Annane D, Gerlach H,
Opal SM, Sevransky JE, Sprung CL, Douglas IS, Jaeschke
R, et al. 2013. Surviving sepsis campaign: International
guidelines for management of severe sepsis and septic
shock: 2012. Crit Care Med 41: 580–637.
Dennis DT, Inglesby TV, Henderson DA, Bartlett JG, Ascher
MS, Eitzen E, Fine AD, Friedlander AM, Hauer J, Layton
M, et al. 2001. Tularemia as a biological weapon: Medical
and public health management. JAMA 285: 2763– 2773.
Doi Y, Yokoyama K, Yamane K, Wachino J, Shibata N, Yagi T,
Shibayama K, Kato H, Arakawa Y. 2004. Plasmid-medi-
ated 16S rRNA methylase in Serratia marcescens confer-
ring high-level resistance to aminoglycosides. Antimicrob
Agents Chemother 48: 491– 496.
Drusano GL, Ambrose PG, Bhavnani SM, Bertino JS, Naf-
ziger AN, Louie A. 2007. Back to the future: Using amino-
glycosides again and how to dose them optimally. Clin
Infect Dis 45: 753– 760.
Dzidic S, Bedekovic
´V. 2003. Horizontal gene transfer-
emerging multidrug resistance in hospital bacteria. Acta
Pharmacol Sin 24: 519– 526.
Eliopoulos GM, Eliopoulos CT. 1988. Antibiotic combina-
tions: Should they be tested? Clin Microbiol Rev 1: 139–
156.
Falagas ME, Siempos II, Bliziotis IA, Michalopoulos A.
2006. Administration of antibiotics via the respirato-
ry tract for the prevention of ICU-acquired pneumonia:
A meta-analysis of comparative trials. Crit Care 10: R123.
Feizabadi MM, Asadi S, Zohari M, Gharavi S, Etemadi G.
2004. Genetic characterization of high-level gentamicin-
resistant strains of Enterococcus faecalis in Iran. Can J
Microbiol 50: 869– 872.
Gonzalez-Padilla M, Torre-Cisneros J, Rivera-Espinar F,
Pontes-Moreno A, Lo
´pez-Cerero L, Pascual A, Natera
C, Rodrı
´guez M, Salcedo I, Rodrı
´guez-Lo
´pez F, et al.
2015. Gentamicin therapy for sepsis due to carbape-
nem-resistant and colistin-resistant Klebsiella pneumo-
niae.J Antimicrob Chemother 70: 905– 913.
Griffith DE, Aksamit T, Brown-Elliott BA, Catanzaro A,
Daley C, Gordin F, Holland SM, Horsburgh R, Huitt G,
Iademarco MF, et al. 2007. An official ATS/IDSA state-
ment: Diagnosis, treatment, and prevention of nontu-
berculous mycobacterial diseases. Am J Respir Crit Care
Med 175: 367– 416.
Hancock RE. 1984. Alterations in outer membrane perme-
ability. Annu Rev Microbiol 38: 237– 264.
Hancock RE, Raffle VJ, Nicas TI. 1981. Involvement of the
outer membrane in gentamicin and streptomycin uptake
and killing in Pseudomonas aeruginosa.Antimicrob
Agents Chemother 19: 777– 785.
Hancock RE, Farmer SW, Li ZS, Poole K. 1991. Interaction
of aminoglycosides with the outer membranes and puri-
fied lipopolysaccharide and OmpF porin of Escherichia
coli.Antimicrob Agents Chemother 35: 1309–1314.
Ho YI, Chan CY, Cheng AF. 1997. In-vitro activities of ami-
noglycoside-aminocyclitols against mycobacteria. J Anti-
microb Chemother 40: 27– 32.
Hocquet D, Vogne C, El Garch F, Vejux A, Gotoh N, Lee A,
Lomovskaya O, Ple
´siat P. 2003. MexXY-OprM efflux
pump is necessary for adaptive resistance of Pseudomonas
aeruginosa to aminoglycosides. Antimicrob Agents Che-
mother 47: 1371–1375.
Hollingshead S, Vapnek D. 1985. Nucleotide sequence anal-
ysis of a gene encoding a streptomycin/spectinomycin
adenylyltransferase. Plasmid 13: 17– 30.
Humes HD, Weinberg JM, Knauss TC. 1982. Clinical and
pathophysiologic aspects of aminoglycoside nephrotoxi-
city. Am J Kidney Dis 2: 5– 29.
Ika
¨heimo I, Syrja
¨la
¨H, Karhukorpi J, Schildt R, Koskela M.
2000. In vitro antibiotic susceptibility of Francisella tu-
larensis isolated from humans and animals. J Antimicrob
Chemother 46: 287– 290.
Ioannidou E, Siempos II, Falagas ME. 2007. Administration
of antimicrobials via the respiratory tract for the treat-
ment of patients with nosocomial pneumonia: A meta-
analysis. J Antimicrob Chemother 60: 1216– 1226.
Islam S, Jalal S, Wretlind B. 2004. Expression of the MexXY
efflux pump in amikacin-resistant isolates of Pseudomo-
nas aeruginosa.Clin Microbiol Infect 10: 877– 883.
Jackson J, Chen C, Buising K. 2013. Aminoglycosides: How
should we use them in the 21st century? Curr Opin Infect
Dis 26: 516– 525.
Jones RN, Flonta M, Gurler N, Cepparulo M, Mendes RE,
Castanheira M. 2014. Resistance surveillance program
report for selected European nations (2011). Diagn Mi-
crobiol Infect Dis 78: 429– 436.
Karlowsky JA, Draghi DC, Jones ME, Thornsberry C, Fried-
land IR, Sahm DF. 2003. Surveillance for antimicrobial
susceptibility among clinical isolates of Pseudomonas aer-
uginosa and Acinetobacter baumannii from hospitalized
patients in the United States, 1998 to 2001. Antimicrob
Agents Chemother 47: 1681– 1688.
Aminoglycosides: An Overview
Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029 15
www.perspectivesinmedicine.org
Kim C, Mobashery S. 2005. Phosphoryl transfer by amino-
glycoside 30-phosphotransferases and manifestation of
antibiotic resistance. Bioorg Chem 33: 149– 158.
Kislak JW. 1972. The susceptibility of Bacteroides fragilis to
24 antibiotics. J Infect Dis 125: 295–299.
Kotra LP, Haddad J, Mobashery S. 2000. Aminoglycosides:
Perspectives on mechanisms of action and resistance
and strategies to counter resistance. Antimicrob Agents
Chemother 44: 3249– 3256.
Landman D, Babu E, Shah N, Kelly P, Ba
¨cker M, Bratu S,
Quale J. 2010. Activityof a novel aminoglycoside, ACHN-
490, against clinical isolates of Escherichia coli and
Klebsiella pneumoniae from New York City. J Antimicrob
Chemother 65: 2123– 2127.
Landman D, Kelly P, Ba
¨cker M, Babu E, Shah N, Bratu S,
Quale J. 2011. Antimicrobial activity of a novel amino-
glycoside, ACHN-490, against Acinetobacter baumannii
and Pseudomonas aeruginosa from New York City. J Anti-
microb Chemother 66: 332– 334.
Le J, McKee B, Srisupha-Olarn W, Burgess DS. 2011. In vitro
activity of carbapenems alone and in combination with
amikacin against KPC-producing Klebsiella pneumoniae.
J Clin Med Res 3: 106– 110.
Le Goffic F, Baca B, Soussy CJ, Dublanchet A, Duval J. 1976.
ANT(40)I: A new aminoglycoside nucleotidyltransferase
found in “staphylococcus aureus.Ann Microbiol (Paris)
127: 391– 399 (author’s transl.).
Livermore DM, Mushtaq S, Warner M, Zhang JC, Maharjan
S, Doumith M, Woodford N. 2011. Activity of amino-
glycosides, including ACHN-490, against carbapenem-
resistant Enterobacteriaceae isolates. J Antimicrob Che-
mother 66: 48–53.
Llano-Sotelo B, Azucena EF, Kotra LP, Mobashery S, Chow
CS. 2002. Aminoglycosides modified by resistance en-
zymes display diminished binding to the bacterial ribo-
somal aminoacyl-tRNA site. Chem Biol 9: 455– 463.
Lode H, Grunert K, Koeppe P, Langmaack H. 1976. Phar-
macokinetic and clinical studies with amikacin, a new
aminoglycoside antibiotic. J Infect Dis 134: S316– S322.
Macinga DR, Rather PN. 1999. The chromosomal 20-N-
acetyltransferase of Providencia stuartii: Physiological
functions and genetic regulation. Front Biosci 4: D132–
D140.
Magnet S, Blanchard JS. 2005. Molecular insights into ami-
noglycoside action and resistance. Chem Rev 105: 477
498.
Magnet S, Courvalin P, Lambert T. 2001. Resistance-nodu-
lation-cell division-type efflux pump involved in amino-
glycoside resistance in Acinetobacter baumannii strain
BM4454. Antimicrob Agents Chemother 45: 3375–3380.
Martin WJ, Gardner M, Washington JA. 1972. In vitro an-
timicrobial susceptibility of anaerobic bacteria isolated
from clinical specimens. Antimicrob Agents Chemother
1: 148–158.
Matsumoto T. 2014. Arbekacin: Another novel agent for
treating infections due to methicillin-resistant Staphylo-
coccus aureus and multidrug-resistant Gram-negative
pathogens. Clin Pharmacol 6: 139– 148.
Mingeot-Leclercq MP, Glupczynski Y, Tulkens PM. 1999.
Aminoglycosides: Activity and resistance. Antimicrob
Agents Chemother 43: 727– 737.
Miro
´E, Gru
¨nbaum F, Go
´mez L, Rivera A, Mirelis B, Coll P,
Navarro F. 2013. Characterization of aminoglycoside-
modifying enzymes in enterobacteriaceae clinical strains
and characterization of the plasmids implicated in their
diffusion. Microb Drug Resist 19: 94– 99.
Moore RD, Smith CR, Lipsky JJ, Mellits ED, Lietman PS.
1984. Risk factors for nephrotoxicity in patients treated
with aminoglycosides. Ann Intern Med 100: 352– 357.
Murphy E. 1985. Nucleotide sequence of a spectinomycin
adenyltransferase AAD(9) determinant from Staphylo-
coccus aureus and its relationship to AAD(300)(9). Mol
Gen Genet 200: 33–39.
Murry KR, McKinnon PS, Mitrzyk B,Rybak MJ. 1999. Phar-
macodynamic characterization of nephrotoxicity associ-
ated with once-daily aminoglycoside. Pharmacotherapy
19: 1252–1260.
Mushtaq S, Irfan S, Sarma JB, Doumith M, Pike R, Pitout
J, Livermore DM, Woodford N. 2011. Phylogenetic di-
versity of Escherichia coli strains producing NDM-
type carbapenemases. J Antimicrob Chemother 66: 2002–
2005.
Mwengee W, Butler T,Mgema S, Mhina G, Almasi Y,Bradley
C, Formanik JB, RochesterCG. 2006. Treatmentof plague
with gentamicin or doxycycline in a randomized clinical
trial in Tanzania. Clin Infect Dis 42: 614–621.
Nagabhushan T, Miller G, Weinstein M. 1982. Structure–
activity relationships in aminoglycoside-aminocyclitol
antibiotics. In The aminoglycosides: Microbiology, clinical
use and toxicology (ed. Whelton A, Neu HC), pp. 3– 27.
Marcel Dekker, New York.
Nichols WW, Young SN. 1985. Respiration-dependent up-
take of dihydrostreptomycin by Escherichia coli. Its irre-
versible nature and lack of evidence for a uniport process.
Biochem J 228: 505–512.
Nicolau DP, Belliveau PP, Nightingale CH, Quintiliani R,
Freeman CD. 1995. Implementation of a once-daily ami-
noglycoside program in a large community-teaching
hospital. Hosp Pharm 30: 674–676, 679–680.
Okamoto S, Suzuki Y. 1965. Chloramphenicol-, dihydros-
treptomycin-, and kanamycin-inactivating enzymes
from multiple drug-resistant Escherichia coli carrying
episome “R”. Nature 208: 1301– 1303.
Ounissi H, Derlot E, Carlier C, Courvalin P. 1990. Gene
homogeneity for aminoglycoside-modifying enzymes
in Gram-positive cocci. Antimicrob Agents Chemother
34: 2164–2168.
Pankuch GA, Seifert H, Appelbaum PC. 2010. Activity of
doripenem with and without levofloxacin, amikacin,
and colistin against Pseudomonas aeruginosa and Aci-
netobacter baumannii.Diagn Microbiol Infect Dis 67:
191–197.
Plantinga NL, Bonten MJ. 2015. Selective decontamination
and antibiotic resistance in ICUs. Crit Care 19: 259.
Podnecky NL, Rhodes KA, Schweizer HP. 2015. Efflux
pump-mediated drug resistance in Burkholderia.Front
Microbiol 6: 305.
Poirel L, Savov E, Nazli A, Trifonova A, Todorova I, Gergova
I, Nordmann P. 2014. Outbreak caused by NDM-1- and
RmtB-producing Escherichia coli in Bulgaria. Antimicrob
Agents Chemother 58: 2472– 2474.
K.M. Krause et al.
16 Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029
www.perspectivesinmedicine.org
Price R, MacLennan G, Glen J; SuDDICU Collaboration.
2014. Selective digestive or oropharyngeal decontamina-
tion and topical oropharyngeal chlorhexidine for preven-
tion of death in general intensive care: Systematic review
and network meta-analysis. BMJ 348: g2197.
Prins JM, Bu
¨ller HR, Kuijper EJ, Tange RA, Speelman P.
1993. Once versus thrice daily gentamicin in patients
with serious infections. Lancet 341: 335–339.
Ramirez MS, Tolmasky ME. 2010. Aminoglycoside modify-
ing enzymes. Drug Resist Updat 13: 151–171.
Ramsey BW, Pepe MS, Quan JM, Otto KL, Montgomery AB,
Williams-Warren J, Vasiljev-K M, Borowitz D, Bowman
CM, Marshall BC, et al. 1999. Intermittent administra-
tion of inhaled tobramycin in patients with cystic fibrosis.
Cystic Fibrosis Inhaled Tobramycin Study Group. N Engl
JMed340: 23–30.
Rather PN, Orosz E, Shaw KJ, Hare R, Miller G. 1993.
Characterization and transcriptional regulation of the
20-N-acetyltransferase gene from Providencia stuartii.J
Bacteriol 175: 6492– 6498.
Ristuccia AM, Cunha BA. 1985. An overview of amikacin.
Ther Drug Monit 7: 12– 25.
Robicsek A, Strahilevitz J, Jacoby GA, Macielag M, Abbanat
D, Park CH, Bush K, Hooper DC. 2006. Fluoroquino-
lone-modifying enzyme: A new adaptation of a common
aminoglycoside acetyltransferase. Nat Med 12: 83– 88.
Roos D, DijksmanLM, Tijssen JG, Gouma DJ, Gerhards MF,
Oudemans-van Straaten HM. 2013. Systematic review of
perioperative selective decontamination of the digestive
tract in elective gastrointestinal surgery. Br J Surg 100:
1579– 1588.
Rosenberg EY, Ma D, Nikaido H. 2000. AcrD of Escherichia
coli is an aminoglycoside efflux pump. J Bacteriol 182:
1754– 1756.
Rybak MJ, Abate BJ, Kang SL, Ruffing MJ, Lerner SA, Dru-
sano GL. 1999. Prospective evaluation of the effect of an
aminoglycoside dosing regimen on rates of observed
nephrotoxicity and ototoxicity. Antimicrob Agents Che-
mother 43: 1549– 1555.
Sader HS, Farrell DJ, Flamm RK, Jones RN. 2014. Antimi-
crobial susceptibility of Gram-negative organisms isolat-
ed from patients hospitalised with pneumonia in U.S.
and European hospitals: Results from the SENTRY Anti-
microbial Surveillance Program, 2009– 2012. Int J Anti-
microb Agents 43: 328– 334.
Sader HS, Rhomberg PR, Farrell DJ, Jones RN. 2015. Arbe-
kacin activity against contemporary clinical bacteria iso-
lated from patients hospitalized with pneumonia. Anti-
microb Agents Chemother 59: 3263– 3270.
Sawicki GS, Signorovitch JE, Zhang J, Latremouille-Viau D,
von Wartburg M, Wu EQ, Shi L. 2012. Reducedmortalit y
in cystic fibrosis patients treated with tobramycin inha-
lation solution. Pediatr Pulmonol 47: 44–52.
Schwartz JJ, Gazumyan A, Schwartz I. 1992. rRNA gene
organization in the Lyme disease spirochete, Borrelia
burgdorferi.J Bacteriol 174: 3757–3765.
Shaw KJ, Rather PN, Hare RS, Miller GH. 1993. Molecular
genetics of aminoglycoside resistance genes and familial
relationships of the aminoglycoside-modifying enzymes.
Microbiol Rev 57: 138– 163.
Simon VK, Mo
¨singer EU, Malerczy V. 1973. Pharmacoki-
netic studies of tobramycin and gentamicin. Antimicrob
Agents Chemother 3: 445– 450.
Skeggs PA, Thompson J, Cundliffe E. 1985. Methylation of
16S ribosomal RNA and resistance to aminoglycoside
antibiotics in clones of Streptomyces lividans carrying
DNA from Streptomyces tenjimariensis.Mol Gen Genet
200: 415– 421.
Snowden J, Stovall S. 2011. Tularemia: Retrospective review
of 10 years’ experience in Arkansas. Clin Pediatr (Phila)
50: 64–68.
Stover KR, Riche DM, Gandy CL, Henderson H. 2012. What
would we do without metronidazole? Am J Med Sci 343:
316–319.
Strahilevitz J, Jacoby GA, Hooper DC, Robicsek A. 2009.
Plasmid-mediated quinolone resistance: A multifaceted
threat. Clin Microbiol Rev 22: 664–689.
Stubbings W, Bostock J, Ingham E, Chopra I. 2006. Mech-
anisms of the post-antibiotic effects induced by rifampi-
cin and gentamicin in Escherichia coli.J Antimicrob Che-
mother 58: 444–448.
Sundar S, Jha TK, Thakur CP, Sinha PK, Bhattacharya SK.
2007. Injectable paromomycin for visceral leishmaniasis
in India. N Engl J Med 356: 2571 –2581.
Sundar S, Sinha PK, Rai M, Verma DK, Nawin K, Alam S,
Chakravarty J, Vaillant M, VermaN, Pandey K, et al. 2011.
Comparison of short-course multidrug treatment with
standard therapy for visceral leishmaniasis in India: An
open-label, non-inferiority, randomised controlled trial.
Lancet 377: 477–486.
Swenson JM, Wallace RJ, Silcox VA, Thornsberry C. 1985.
Antimicrobial susceptibility of five subgroups of Myco-
bacterium fortuitum and Mycobacterium chelonae.Anti-
microb Agents Chemother 28: 807– 811.
Taber HW, Mueller JP, Miller PF, Arrow AS. 1987. Bacterial
uptake of aminoglycoside antibiotics. Microbiol Rev 51:
439–457.
Tamma PD, Cosgrove SE, Maragakis LL. 2012. Combination
therapy for treatment of infections with Gram-negative
bacteria. Clin Microbiol Rev 25: 450– 470.
Thompson J, Skeggs PA, Cundliffe E. 1985. Methylation of
16S ribosomal RNA and resistance to the aminoglycoside
antibiotics gentamicin and kanamycin determined by
DNA from the gentamicin-producer, Micromonospora
purpurea.Mol Gen Genet 201: 168– 173.
Vakulenko SB, Mobashery S. 2003. Versatility of aminogly-
cosides and prospects for their future. Clin Microbiol Rev
16: 430– 450.
Wachino J, Arakawa Y. 2012. Exogenously acquired 16S
rRNA methyltransferases found in aminoglycoside-re-
sistant pathogenic Gram-negative bacteria: An update.
Drug Resist Updat 15: 133–148.
Wachino J, Yamane K, Shibayama K, Kurokawa H, Shibata
N, Suzuki S, Doi Y, Kimura K, Ike Y, Arakawa Y. 2006.
Novel plasmid-mediated 16S rRNA methylase, RmtC,
found in a Proteus mirabilis isolate demonstrating ex-
traordinary high-level resistance against various amino-
glycosides. Antimicrob Agents Chemother 50: 178–184.
Wachino J, Shibayama K, Kurokawa H, Kimura K,
Yamane K, Suzuki S, Shibata N, Ike Y, Arakawa Y.
2007. Novel plasmid-mediated 16S rRNA m1A1408
Aminoglycosides: An Overview
Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029 17
www.perspectivesinmedicine.org
methyltransferase, NpmA, found in a clinically isolat-
ed Escherichia coli strain resistant to structurally di-
verse aminoglycosides. Antimicrob Agents Chemother
51: 4401–4409.
Waters VL. 1999. Conjugative transfer in the dissemination
of b-lactam and aminoglycoside resistance. Front Biosci
4: D433–D456.
Westbrock-Wadman S, Sherman DR, Hickey MJ, Coulter
SN, Zhu YQ, Warrener P, Nguyen LY, Shawar RM, Folger
KR, Stover CK. 1999. Characterization of a Pseudomonas
aeruginosa efflux pump contributing to aminoglycoside
impermeability. Antimicrob Agents Chemother 43: 2975–
2983.
Wilson DN. 2014. Ribosome-targeting antibiotics and
mechanisms of bacterial resistance. Nat Rev Microbiol
12: 35–48.
World Health Organization. 2011. Guidelines for the pro-
grammatic management of drug-resistant tuberculosis:
2011 update. World Health Organization, Geneva.
Wright GD, Thompson PR. 1999. Aminoglycoside phos-
photransferases: Proteins, structure, and mechanism.
Front Biosci 4: D9– D21.
Yamane K, Wachino J, Doi Y, Kurokawa H, Arakawa Y. 2005.
Global spread of multiple aminoglycoside resistance
genes. Emerg Infect Dis 11: 951– 953.
Yan JJ, Wu JJ, Ko WC, Tsai SH, Chuang CL, Wu HM, Lu YJ,
Li JD. 2004. Plasmid-mediated 16S rRNA methylases
conferring high-level aminoglycoside resistance in Es-
cherichia coli and Klebsiella pneumoniae isolates from
two Taiwanese hospitals. J Antimicrob Chemother 54:
1007–1012.
Yokoyama K, Doi Y, Yamane K, Kurokawa H, Shibata N,
Shibayama K, Yagi T, Kato H, Arakawa Y. 2003. Acquisi-
tion of 16S rRNA methylase gene in Pseudomonas aeru-
ginosa.Lancet 362: 1888–1893.
Zhanel GG, Hoban DJ, Harding GK. 1991. The postanti-
biotic effect: A review of in vitro and in vivo data. DICP
25: 153– 163.
K.M. Krause et al.
18 Cite this article as Cold Spring Harb Perspect Med 2016;6:a027029
www.perspectivesinmedicine.org
... In ESKAPE pathogens, aminoglycoside resistance takes place through aminoglycoside modifying enzymes (AMEs). AMEs are reported to be found on the plasmid and their transmission occurs through horizontal gene transfer (Krause et al. 2016). A decrease in antibiotic potency arises due to the binding of AMEs to the aminoglycosides and rendering changes to its amino and hydroxyl groups (Ramirez and Tolmasky 2010). ...
Article
Full-text available
Antimicrobial resistance is a huge challenge for the global health system, bringing about massive mortalities throughout the world. On the landscape of infectious diseases, the ESKAPE (Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter species) organisms have emerged as grave threats, leading to millions of life-threatening infections. These pathogens have developed numerous mechanisms of resistance against conventional antibiotic agents such as penicillins, cephalosporins, glycopeptides, etc. The prospective research is pursuing novel methodologies for combating antimicrobial resistance in ESKAPE bacteria. This review depicts strategies other than conventional antibiotic agents to tackle the rising tendencies of antimicrobial resistance in ESKAPE pathogens. Many in vitro and in vivo studies have established the effectiveness of phage therapy and endolysins in countering the dominance of ESKAPE pathogens. Antimicrobial peptides, photodynamic therapy with mesoporous nanoparticles, and some other strategies have also demonstrated their potency to kill ESKAPE bacteria. Despite their fruitfulness, many challenges and limitations are also associated with these approaches. This review aims to summarize the progress and advancement in countering the antimicrobial resistance in ESKAPE pathogens. This review also recommends conducting further research to better understand the technicalities of these strategies and to further refine them to compose these methods as promising options for antimicrobial therapy against ESKAPE pathogens.
... Aminoglycosides were amongst the first broad-spectrum antibiotics used for clinical purposes and are now garnering renewed interest due to the growing global incidence of antimicrobial resistance [51]. The antagonism of the COMS modality by aminoglycosides has importance beyond that of in vitro electromagnetic regenerative paradigms, as they are used across a broad array of in vitro and in vivo medical paradigms that are responsive to diverse developmental stimuli transduced by TRPC1 [3]. ...
Article
Full-text available
Concurrent optical and magnetic stimulation (COMS) combines extremely low-frequency electromagnetic and light exposure for enhanced wound healing. We investigated the potential mechanistic synergism between the magnetic and light components of COMS by comparing their individual and combined cellular responses. Lone magnetic field exposure produced greater enhancements in cell proliferation than light alone, yet the combined effects of magnetic fields and light were supra-additive of the individual responses. Reactive oxygen species were incrementally reduced by exposure to light, magnetics fields, and their combination, wherein statistical significance was only achieved by the combined COMS modality. By contrast, ATP production was most greatly enhanced by magnetic exposure in combination with light, indicating that mitochondrial respiratory efficiency was improved by the combination of magnetic fields plus light. Protein expression pertaining to cell proliferation was preferentially enhanced by the COMS modality, as were the protein levels of the TRPC1 cation channel that had been previously implicated as part of a calcium-mitochondrial signaling axis invoked by electromagnetic exposure and necessary for proliferation. These results indicate that light facilitates functional synergism with magnetic fields that ultimately impinge on mitochondria-dependent developmental responses. Aminoglycoside antibiotics (AGAs) have been previously shown to inhibit TRPC1-mediated magnetotransduction, whereas their influence over photomodulation has not been explored. Streptomycin applied during exposure to light, magnetic fields, or COMS reduced their respective proliferation enhancements, whereas streptomycin added after the exposure did not. Magnetic field exposure and the COMS modality were capable of partially overcoming the antagonism of proliferation produced by strep-tomycin treatment, whereas light alone was not. The antagonism of photon-electromagnetic effects by streptomycin implicates TRPC1-mediated calcium entry in both magnetotransduction and pho-tomodulation. Avoiding the prophylactic use of AGAs during COMS therapy will be crucial for maintaining clinical efficacy and is a common concern in most other electromagnetic regenerative paradigms. Citation: Iversen, J.N.; Fröhlich, J.; Tai, Y.K.; Franco-Obregón, A.
... Among aminoglycoside antibiotics, neomycin has been widely used to control bacterial and viral infections in the aquaculture industry. For example, neomycin is effective against Aeromonas species and cyprinid herpesvirus 2 infections in fish [11][12][13] and has demonstrated efficacy in treating E. tardainduced infections [14,15]. Despite the current susceptibility of most E. tarda strains to neomycin [16,17], the substantial use of this antibiotic in aquaculture increases the development of resistance, potentially compromising its effectiveness. ...
Article
Full-text available
The global surge in multidrug-resistant bacteria owing to antibiotic misuse and overuse poses considerable risks to human and animal health. With existing antibiotics losing their effectiveness and the protracted process of developing new antibiotics, urgent alternatives are imperative to curb disease spread. Notably, improving the bactericidal effect of antibiotics by using non-antibiotic substances has emerged as a viable strategy. Although reduced nicotinamide adenine dinucleotide (NADH) may play a crucial role in regulating bacterial resistance, studies examining how the change of metabolic profile and bacterial resistance following by exogenous administration are scarce. Therefore, this study aimed to elucidate the metabolic changes that occur in Edwardsiella tarda (E. tarda), which exhibits resistance to various antibiotics, following the exogenous addition of NADH using metabolomics. The effects of these alterations on the bactericidal activity of neomycin were investigated. NADH enhanced the effectiveness of aminoglycoside antibiotics against E. tarda ATCC15947, achieving bacterial eradication at low doses. Metabolomic analysis revealed that NADH reprogrammed the ATCC15947 metabolic profile by promoting purine metabolism and energy metabolism, yielding increased adenosine triphosphate (ATP) levels. Increased ATP levels played a crucial role in enhancing the bactericidal effects of neomycin. Moreover, exogenous NADH promoted the bactericidal efficacy of tetracyclines and chloramphenicols. NADH in combination with neomycin was effective against other clinically resistant bacteria, including Aeromonas hydrophila, Vibrio parahaemolyticus, methicillin-resistant Staphylococcus aureus, and Listeria monocytogenes. These results may facilitate the development of effective approaches for preventing and managing E. tarda–induced infections and multidrug resistance in aquaculture and clinical settings.
... Therefore, increasing efforts have been made in recent years to develop small molecules as readthrough agents [25]. A well-known class of small molecules that promote translational readthrough are aminoglycosides [26], which have historically been used as antibiotics [27]. Ataluren, an oxadiazole discovered to suppress Duchenne muscular dystrophy nonsense mutations, is also renowned [28]. ...
Article
Full-text available
Nonsense mutations are genetic mutations that create premature termination codons (PTCs), leading to truncated, defective proteins in diseases such as cystic fibrosis, neurofibromatosis type 1, Dravet syndrome, Hurler syndrome, Beta thalassemia, inherited bone marrow failure syndromes, Duchenne muscular dystrophy, and even cancer. These mutations can also trigger a cellular surveillance mechanism known as nonsense-mediated mRNA decay (NMD) that degrades the PTC-containing mRNA. The activation of NMD can attenuate the consequences of truncated, defective, and potentially toxic proteins in the cell. Since approximately 20% of all single-point mutations are disease-causing nonsense mutations, it is not surprising that this field has received significant attention, resulting in a remarkable advancement in recent years. In fact, since our last review on this topic, new examples of nonsense suppression approaches have been reported, namely new ways of promoting the translational readthrough of PTCs or inhibiting the NMD pathway. With this review, we update the state-of-the-art technologies in nonsense suppression, focusing on novel modalities with therapeutic potential, such as small molecules (readthrough agents, NMD inhibitors, and molecular glue degraders); antisense oligonucleotides; tRNA suppressors; ADAR-mediated RNA editing; targeted pseudouridylation; and gene/base editing. While these various modalities have significantly advanced in their development stage since our last review, each has advantages (e.g., ease of delivery and specificity) and disadvantages (manufacturing complexity and off-target effect potential), which we discuss here.
... The aminoglycosides were amongst the first broad-spectrum antibiotics used for clinical purposes that are now garnering renewed interest due to the growing global incidence of 10 antimicrobial resistance [51]. The antagonism of the COMS modality by the aminoglycosides will have importance beyond that of in vitro electromagnetic regenerative paradigms as they are used across a broad array of in vitro and in vivo medical paradigms that are responsive to diverse developmental stimuli transduced by TRPC1 [3]. ...
Preprint
Full-text available
Concurrent Optical and Magnetic Stimulation (COMS) combines extremely low frequency electromagnetic and light stimulation modalities for enhanced wound healing. In this report, we investigated the potential mechanistic synergism between magnetic and light components by comparing their individual and combined cellular responses. Although each individual stimulation mode was capable of stimulating cell proliferation, magnetic fields exposure alone produced greater proliferation enhancement than light alone, yet their combined effects of were supra-additive of the individual responses. Reactive oxygen species were incrementally reduced by exposure to light, magnetics fields and their combination, where statistical significance was ultimately achieved. On the other hand, ATP production was significantly enhanced in response to magnetic exposure alone or in combination with light. These results suggest that mitochondrial respiratory efficiency was preferentially improved by the combination of the two stimuli. Protein expression related to cell proliferation was preferentially enhanced by the combinatory COMS modality as were the protein levels of the TRPC1 cation channel previously implicated as part of a calcium-mitochondrial signaling axis invoked by electromagnetic exposure and involved in proliferation. These results corroborate that light facilitates functional synergism with magnetic fields that ultimately impinge on mitochondria-dependent developmental responses. The aminoglycoside antibiotics (AGAs) have been previously shown to antagonize TRPC1-mediated magnetotransduction. Streptomycin applied during the exposure to any of the stimulation modes reduced proliferation enhancement, whereas streptomycin added after the exposure did not. Notably, the COMS modality and magnetic field exposure were capable of partially overcoming streptomycin antagonism of proliferation, whereas light alone was not. Transductional antagonism of combinatorial photonelectromagnetic effects by streptomycin thus aligns with previous studies showing that the AGAs block TRPC1-mediated magnetotransduction when applied during exposure. The prophylactic usage of the AGAs (neomycin) should be avoided during COMS therapy to maintain clinical efficacy and would be a common concern with most other electromagnetic regenerative paradigms.
... To investigate the effect of IAA on antibiotic resistance in A153, we determined the MIC values of various antibiotics that operate with different mechanisms of action, namely, ampicillin, chloramphenicol, gentamicin, kanamycin, nalidixic acid, streptomycin, rifampicin, and tetracycline. We found that IAA treatment enhances resistance to gentamicin and kanamycin (Table S6), two aminogly coside antibiotics that inhibit protein synthesis by binding to the 30S ribosomal subunit (57). Subsequent experiments revealed that IAA significantly increased resistance to ampicillin in minimal medium agar plates by at least an order of magnitude ( Fig. 5), which correlates with the increased expression of the β-lactamase encoding gene The complete list of differentially expressed genes is provided in Table S1. ...
Article
The communication between plants and their microbiota is highly dynamic and involves a complex network of signal molecules. Among them, the auxin indole-3-acetic acid (IAA) is a critical phytohormone that not only regulates plant growth and development, but is emerging as an important inter- and intra-kingdom signal that modulates many bacterial processes that are important during interaction with their plant hosts. However, the corresponding signaling cascades remain largely unknown. Here, we advance our understanding of the largely unknown mechanisms by which IAA carries out its regulatory functions in plant-associated bacteria. We showed that IAA caused important changes in the global transcriptome of the rhizobacterium Serratia plymuthica and multidisciplinary approaches revealed that IAA sensing interferes with the signaling mediated by other pivotal plant-derived signals such as amino acids and 4-hydroxybenzoic acid. Exposure to IAA caused large alterations in the transcript levels of genes involved in amino acid metabolism, resulting in significant metabolic alterations. IAA treatment also increased resistance to toxic aromatic compounds through the induction of the AaeXAB pump, which also confers resistance to IAA. Furthermore, IAA promoted motility and severely inhibited biofilm formation; phenotypes that were associated with decreased c-di-GMP levels and capsule production. IAA increased capsule gene expression and enhanced bacterial sensitivity to a capsule-dependent phage. Additionally, IAA induced the expression of several genes involved in antibiotic resistance and led to changes in the susceptibility and responses to antibiotics with different mechanisms of action. Collectively, our study illustrates the complexity of IAA-mediated signaling in plant-associated bacteria. IMPORTANCE Signal sensing plays an important role in bacterial adaptation to ecological niches and hosts. This communication appears to be particularly important in plant-associated bacteria since they possess a large number of signal transduction systems that respond to a wide diversity of chemical, physical, and biological stimuli. IAA is emerging as a key inter- and intra-kingdom signal molecule that regulates a variety of bacterial processes. However, despite the extensive knowledge of the IAA-mediated regulatory mechanisms in plants, IAA signaling in bacteria remains largely unknown. Here, we provide insight into the diversity of mechanisms by which IAA regulates primary and secondary metabolism, biofilm formation, motility, antibiotic susceptibility, and phage sensitivity in a biocontrol rhizobacterium. This work has important implications for our understanding of bacterial ecology in plant environments and for the biotechnological and clinical applications of IAA, as well as related molecules.
Article
Background Tuberculosis (TB), a global infectious threat, has seen a concerning rise in aminoglycoside-resistant Mycobacterium tuberculosis (M.tb) strains. The potential role of capsule proteins remains largely unexplored. This layer acts as the primary barrier for tubercle bacilli, attempting to infiltrate host cells and subsequent disease development. Methods The study aims to bridge this gap by investigating the differentially expressed capsule proteins in aminoglycoside-resistant M.tb clinical isolates compared with drug-sensitive isolates employing two-dimensional gel electrophoresis, mass spectrometry, and bioinformatic approaches. Results We identified eight proteins that exhibited significant upregulation in aminoglycoside-resistant isolates. Protein Rv3029c and Rv2110c were associated with intermediary metabolism and respiration; Rv2462c with cell wall and cell processes; Rv3804c with lipid metabolism; Rv2416c and Rv2623 with virulence and detoxification/adaptation; Rv0020c with regulatory functions; and Rv0639 with information pathways. Notably, the Group-based Prediction System for Prokaryotic Ubiquitin-like Protein (GPS-PUP) algorithm identified potential pupylation sites within all proteins except Rv3804c. Interactome analysis using the STRING 12.0 database revealed potential interactive partners for these proteins, suggesting their involvement in aminoglycoside resistance. Molecular docking studies revealed suitable binding between amikacin and kanamycin drugs with Rv2462c, Rv3804c, and Rv2623 proteins. Conclusion As a result, our findings illustrate the multifaceted nature of aminoglycoside resistance in M.tb and the importance of understanding how capsule proteins play a role in counteracting drug efficacy. Identifying the role of these proteins in drug resistance is crucial for developing more effective treatments and diagnostics for TB.
Article
Aminoglycosides are crucial antibiotics facing challenges from bacterial resistance. This study addresses the importance of aminoglycoside modifying enzymes in the context of escalating resistance. Drawing upon over two decades of structural data in the Protein Data Bank, we focused on two key antibiotics, neomycin B and kanamycin A, to explore how the aminoglycoside structure is exploited by this family of enzymes. A systematic comparison across diverse enzymes and the RNA A‐site target identified common characteristics in the recognition mode, while assessing the adaptability of neomycin B and kanamycin A in various environments.
Article
Full-text available
Selective digestive decontamination (SDD) and selective oropharyngeal decontamination (SOD) have been associated with reduced mortality and lower ICU-acquired bacteremia and ventilator-associated pneumonia rates in areas with low levels of antibiotic resistance. However, the effect of selective decontamination (SDD/SOD) in areas where multidrug-resistant Gram-negative bacteria are endemic is less clear. It will be important to determine whether SDD/SOD improves patient outcome in such settings and how these measures affect the epidemiology of multidrug-resistant Gram-negative bacteria. Here we review the current evidence on the effects of SDD/SOD on antibiotic resistance development in individual ICU patients as well as the effect on ICU ecology, the latter including both ICU-level antibiotic resistance and antibiotic resistance development during long-term use of SDD/SOD.
Article
Full-text available
Several members of the genus Burkholderia are prominent pathogens. Infections caused by these bacteria are difficult to treat because of significant antibiotic resistance. Virtually all Burkholderia species are also resistant to polymyxin, prohibiting use of drugs like colistin that are available for treatment of infections caused by most other drug resistant Gram-negative bacteria. Despite clinical significance and antibiotic resistance of Burkholderia species, characterization of efflux pumps lags behind other non-enteric Gram-negative pathogens such as Acinetobacter baumannii and Pseudomonas aeruginosa. Although efflux pumps have been described in several Burkholderia species, they have been best studied in B. cenocepacia and B. pseudomallei. As in other non-enteric Gram-negatives, efflux pumps of the resistance nodulation cell division (RND) family are the clinically most significant efflux systems in these two species. Several efflux pumps were described in B. cenocepacia, which when expressed confer resistance to clinically significant antibiotics, including aminoglycosides, chloramphenicol, fluoroquinolones, and tetracyclines. Three RND pumps have been characterized in B. pseudomallei, two of which confer either intrinsic or acquired resistance to aminoglycosides, macrolides, chloramphenicol, fluoroquinolones, tetracyclines, trimethoprim, and in some instances trimethoprim+sulfamethoxazole. Several strains of the host-adapted B. mallei, a clone of B. pseudomallei, lack AmrAB-OprA and are therefore aminoglycoside and macrolide susceptible. B. thailandensis is closely related to B. pseudomallei, but non-pathogenic to humans. Its pump repertoire and ensuing drug resistance profile parallels that of B. pseudomallei. An efflux pump in B. vietnamiensis plays a significant role in acquired aminoglycoside resistance. Summarily, efflux pumps are significant players in Burkholderia drug resistance.
Article
Full-text available
Arbekacin is a broad-spectrum aminoglycoside licensed for systemic use in Japan and under clinical development as an inhalation solution in the United States (USA). We evaluated the occurrence of organisms isolated from pneumonias in USA hospitalized patients (PHP), including ventilator-associated pneumonia (VAP), and the in vitro activity of arbekacin. Organism frequency was evaluated from a collection of 2,203 bacterial isolates (339 from VAP) consecutively collected from 25 medical centers in 2012 through the SENTRY Antimicrobial Surveillance Program. Arbekacin activity was tested against 904 isolates from PHP collected in 2012 from 62 USA medical centers and 303 multidrug-resistant (MDR) organisms collected worldwide in 2009 and 2010 from various infection types. Susceptibility to arbekacin and comparator agents was evaluated by the reference broth microdilution method. The four most common organisms from PHP were Staphylococcus aureus, Pseudomonas aeruginosa, Klebsiella spp. and Enterobacter spp. The highest arbekacin MIC value among S. aureus from PHP (43% MRSA) was 4 μg/ml. Against P. aeruginosa from PHP, only one strain had an arbekacin MIC of >16 μg/ml (MIC50/90, 1/4 μg/ml), and susceptibility rates for gentamicin, tobramycin and amikacin were 88.0, 90.0 and 98.0%, respectively. Arbekacin (MIC50, 2 μg/ml) and tobramycin (MIC50, 4 μg/ml) were the most potent aminoglycosides tested against Acinetobacter baumannii. Against Enterobacteriaceae from PHP, arbekacin and gentamicin (MIC50/90, 0.25-1/1-8 μg/ml for both compounds) were generally more potent than tobramycin (MIC50/90, 0.25-2/1-32 μg/ml) and amikacin (MIC50/90, 1-2/2-32 μg/ml). Arbekacin also demonstrated potent in vitro activity against a worldwide collection of well characterized MDR Gram-negative and MRSA strains. Copyright © 2015, American Society for Microbiology. All Rights Reserved.
Article
Objective: To provide an update to the "Surviving Sepsis Campaign Guidelines for Management of Severe Sepsis and Septic Shock," last published in 2008. Design: A consensus committee of 68 international experts representing 30 international organizations was convened. Nominal groups were assembled at key international meetings (for those committee members attending the conference). A formal conflict of interest policy was developed at the onset of the process and enforced throughout. The entire guidelines process was conducted independent of any industry funding. A stand-alone meeting was held for all subgroup heads, co- and vice-chairs, and selected individuals. Teleconferences and electronic-based discussion among subgroups and among the entire committee served as an integral part of the development. Methods: The authors were advised to follow the principles of the Grading of Recommendations Assessment, Development and Evaluation (GRADE) system to guide assessment of quality of evidence from high (A) to very low (D) and to determine the strength of recommendations as strong (1) or weak (2). The potential drawbacks of making strong recommendations in the presence of low-quality evidence were emphasized. Some recommendations were ungraded (UG). Recommendations were classified into three groups: 1) those directly targeting severe sepsis; 2) those targeting general care of the critically ill patient and considered high priority in severe sepsis; and 3) pediatric considerations. Results: Key recommendations and suggestions, listed by category, include: early quantitative resuscitation of the septic patient during the first 6 hrs after recognition (1C); blood cultures before antibiotic therapy (1C); imaging studies performed promptly to confirm a potential source of infection (UG); administration of broad-spectrum antimicrobials therapy within 1 hr of recognition of septic shock (1B) and severe sepsis without septic shock (1C) as the goal of therapy; reassessment of antimicrobial therapy daily for de-escalation, when appropriate (1B); infection source control with attention to the balance of risks and benefits of the chosen method within 12 hrs of diagnosis (1C); initial fluid resuscitation with crystalloid (1B) and consideration of the addition of albumin in patients who continue to require substantial amounts of crystalloid to maintain adequate mean arterial pressure (2C) and the avoidance of hetastarch formulations (1C); initial fluid challenge in patients with sepsis-induced tissue hypoperfusion and suspicion of hypovolemia to achieve a minimum of 30 mL/kg of crystalloids (more rapid administration and greater amounts of fluid may be needed in some patients) (1C); fluid challenge technique continued as long as hemodynamic improvement, as based on either dynamic or static variables (UG); norepinephrine as the first-choice vasopressor to maintain mean arterial pressure ≥ 65 mm Hg (1B); epinephrine when an additional agent is needed to maintain adequate blood pressure (2B); vasopressin (0.03 U/min) can be added to norepinephrine to either raise mean arterial pressure to target or to decrease norepinephrine dose but should not be used as the initial vasopressor (UG); dopamine is not recommended except in highly selected circumstances (2C); dobutamine infusion administered or added to vasopressor in the presence of a) myocardial dysfunction as suggested by elevated cardiac filling pressures and low cardiac output, or b) ongoing signs of hypoperfusion despite achieving adequate intravascular volume and adequate mean arterial pressure (1C); avoiding use of intravenous hydrocortisone in adult septic shock patients if adequate fluid resuscitation and vasopressor therapy are able to restore hemodynamic stability (2C); hemoglobin target of 7-9 g/dL in the absence of tissue hypoperfusion, ischemic coronary artery disease, or acute hemorrhage (1B); low tidal volume (1A) and limitation of inspiratory plateau pressure (1B) for acute respiratory distress syndrome (ARDS); application of at least a minimal amount of positive end-expiratory pressure (PEEP) in ARDS (1B); higher rather than lower level of PEEP for patients with sepsis-induced moderate or severe ARDS (2C); recruitment maneuvers in sepsis patients with severe refractory hypoxemia due to ARDS (2C); prone positioning in sepsis-induced ARDS patients with a PaO2/FIO2 ratio of ≤ 100 mm Hg in facilities that have experience with such practices (2C); head-of-bed elevation in mechanically ventilated patients unless contraindicated (1B); a conservative fluid strategy for patients with established ARDS who do not have evidence of tissue hypoperfusion (1C); protocols for weaning and sedation (1A); minimizing use of either intermittent bolus sedation or continuous infusion sedation targeting specific titration endpoints (1B); avoidance of neuromuscular blockers if possible in the septic patient without ARDS (1C); a short course of neuromuscular blocker (no longer than 48 hrs) for patients with early ARDS and a Pao2/Fio2 < 150 mm Hg (2C); a protocolized approach to blood glucose management commencing insulin dosing when two consecutive blood glucose levels are > 180 mg/dL, targeting an upper blood glucose ≤ 180 mg/dL (1A); equivalency of continuous veno-venous hemofiltration or intermittent hemodialysis (2B); prophylaxis for deep vein thrombosis (1B); use of stress ulcer prophylaxis to prevent upper gastrointestinal bleeding in patients with bleeding risk factors (1B); oral or enteral (if necessary) feedings, as tolerated, rather than either complete fasting or provision of only intravenous glucose within the first 48 hrs after a diagnosis of severe sepsis/septic shock (2C); and addressing goals of care, including treatment plans and end-of-life planning (as appropriate) (1B), as early as feasible, but within 72 hrs of intensive care unit admission (2C). Recommendations specific to pediatric severe sepsis include: therapy with face mask oxygen, high flow nasal cannula oxygen, or nasopharyngeal continuous PEEP in the presence of respiratory distress and hypoxemia (2C), use of physical examination therapeutic endpoints such as capillary refill (2C); for septic shock associated with hypovolemia, the use of crystalloids or albumin to deliver a bolus of 20 mL/kg of crystalloids (or albumin equivalent) over 5 to 10 mins (2C); more common use of inotropes and vasodilators for low cardiac output septic shock associated with elevated systemic vascular resistance (2C); and use of hydrocortisone only in children with suspected or proven "absolute"' adrenal insufficiency (2C). Conclusions: Strong agreement existed among a large cohort of international experts regarding many level 1 recommendations for the best care of patients with severe sepsis. Although a significant number of aspects of care have relatively weak support, evidence-based recommendations regarding the acute management of sepsis and septic shock are the foundation of improved outcomes for this important group of critically ill patients.
Article
Objectives: Antimicrobial therapy for sepsis caused by carbapenem- and colistin-resistant Klebsiella pneumoniae is not well established. We hypothesized that the early use of gentamicin in cases due to susceptible organisms would decrease the crude mortality rate of this infection. Methods: This retrospective cohort study examined 50 cases of sepsis caused by carbapenem-resistant K. pneumoniae occurring between June 2012 and February 2013 during an outbreak of K. pneumoniae ST512 producing KPC-3, SHV-11 and TEM-1. Survival curves categorized by the use of gentamicin were constructed using the Kaplan-Meier method and compared using the log-rank test. Eight multivariate models using Cox regression were designed to study the risk factors for mortality and test the hypothesis. Results: The 30 day crude mortality rate was 38%. The use of targeted gentamicin was associated with reduced mortality (20.7% versus 61.9%, P = 0.02). In all multivariate regression models, the use of gentamicin was independently associated with lower mortality until Day 30 (HR 0.17-0.29, P = 0.03-0.002 depending on the model) after controlling for other potential confounding variables such as age, optimal treatment, renal function, severity of infection, underlying disease, use of tigecycline and previous hospitalization. Conclusions: Gentamicin reduced the mortality from sepsis caused by this K. pneumoniae ST512 clone producing KPC-3, SHV-11 and TEM-1.