You are on page 1of 15

Chapter 8

Biochemical Logic of Antibiotic Inactivation


and Modification

Vanessa D’Costa and Gerard D. Wright

1 Introduction ability. In Table 1 we itemize representative enzymes and


mechanisms, differentiating between mechanisms that mod-
ify the antibiotic (e.g., acylation, phosphorylation), and those
Bacterial resistance to antibiotics manifests itself in both
that essentially cause irreversible destruction (e.g., hydroly-
general and specific protection mechanisms. Consequently,
sis). A general observation evident from Table 1 is that most
the characteristics of resistance can be paralleled to those of
antibiotics that are either natural products or are based on
the mammalian immune response. Antibiotic resistance can
natural product chemical scaffolds are more susceptible to
be differentiated into: (1) nonspecific mechanisms that con-
some form of enzyme-based inactivation, while antibiotics
fer general innate immunity to a class of antibiotics (e.g.,
of synthetic origin (e.g., fluoroquinolones) are not (however,
broad spectrum efflux mechanisms, target modification), and
enzyme-based inactivation of certain fluoroquinolones has
(2) highly precise responses that include selective enzyme-
been reported (2) ). These relatively enzyme-impervious
based mechanisms that mirror the acquired immune response
antibiotics are nonetheless still susceptible to resistance
with respect to target specificity and potency. Bacteria deploy
mechanisms, often substrates for efflux pumps.
both types of mechanisms in response to the presence of
Walsh described the cellular impact and rationale of bio-
cytotoxic antibiotics.
chemical reactions as “molecular logic” (3) and this termi-
Although antibiotic resistance via target modification or
nology works very well in dissecting mechanisms of
efflux mechanisms results in the survival of the resistant
antibiotic resistance. Thus enzyme-catalyzed antibiotic resis-
organism, the concentration of antibiotic that the bacte-
tance is functionally and structurally linked to the mode of
rium is exposed to remains unaffected. Thus, other proximal
action of these agents. For example, modification of key
susceptible organisms can still be targeted by the antimicro-
functional groups on an antibiotic can sterically or electroni-
bial agent. In contrast, enzyme-catalyzed detoxification of
cally block interaction with target (see Sect. 3 below for
antibiotics effectively (and often irreversibly) lowers the con-
examples). This review describes mechanisms of antibiotic
centration of the drug and as a result has the potential for a
destruction and modification resulting in resistance in the
much broader impact on microbial growth. The presence of
context of the mode of action of the antibiotic. Our aim is not
an antibiotic-resistant microbe can, at least in theory, promote
to provide a comprehensive examination of the details of all
the growth of adjacent bacteria that otherwise would be sus-
known resistance mechanisms but rather to focus on selected
ceptible to the antibiotic by inactivating the drug in the local
examples to decode the molecular basis and biological
environment. This can occur even if it is the susceptible organ-
impact of these inactivation strategies.
ism, not the resistant strain, which is the cause of infection.
As a result, enzyme-catalyzed antibiotic inactivation can have
a significant and broad impact on antimicrobial therapy.
Since the first reports of penicillin inactivating strains of 2 Destruction of Antibiotics
bacteria in the early 1940s (1), virtually all antibiotics have
been shown to be modified or destroyed by a cadre of
We classify antibiotic destruction as a mechanism that results
enzymes with hydrolytic, chemical group transfer or redox
in either ablation of a key reactive centre or massive structural
rearrangement that is not readily reversed under normal
G.D. Wright ( ) physiological conditions. Hydrolysis of the reactive β-lactam
M.G. DeGroote Institute for Infectious Disease Research,
Antimicrobial Research Centre, Department of Biochemistry and
ring of penicillin and cephalosporin antibiotics by
Biomedical Sciences, McMaster University, Hamilton, ON, Canada β-lactamases is an example of the first class, and linearization
wrightge@mcmaster.ca of the cyclic depsipeptide of Type B streptogramins by Vgb

D.L. Mayers (ed.), Antimicrobial Drug Resistance, 81


DOI 10.1007/978-1-59745-180-2_8, © Humana Press, a part of Springer Science+Business Media, LLC 2009
82 V. D’Costa and G.D. Wright

Table 1 Survey of enzymatic mechanisms of antibiotic resistance


Antibiotic destruction
Antibiotic Mechanisms Enzyme(s)
β-Lactams Hydrolysis β-Lactamase
Macrolides Hydrolysis Macrolide esterase
Type B streptogramins C–O-bond cleavage Vgb lyase
Tetracyclines Mono-oxidation TetX
Fosfomycin Hydrolysis Epoxidase
Thiol transfer Thiol transferase
Antibiotic modification
Antibiotic Mechanisms Enzyme
Aminoglycosides Acylation Acetyltransferase
Phosphorylation Kinase
Adenylylation AMP-transferase
Macrolides Phosphorylation Kinase
Glycosylation UDP-glucosyl
transferase
Lincosamides Adenylylation AMP-transferase
Rifamycin Glycosylation TDP-glucosyl
transferase
Phosphorylation Kinase
Chloramphenicol Acylation Acetyltransferase
Type A streptogramins Acylation Acetyltransferase

lyase is an example of the second. In all classes, the action of Fig. 1 Structure of the bacterial peptidoglycan unit. Peptidoglycan
resistance enzymes tactically impacts the mode of action of consists of repeating disaccharyl units (N-acetylglucosamine-N-
the affected antibiotics to disrupt their biological activity. acetylmuramic acid), to which a pentapeptide is linked each
N-acetylmuramic acid. Crosslinking between adjacent pentapeptides
Examples of each class are discussed below. provides rigidity to the bacterial cell

2.1 b-Lactam Antibiotics

The β-lactams remain some of the most successful and


widely used antibiotics in modern chemotherapy. These
natural products and their semi-synthetic derivatives act by
covalently modifying so-called penicillin binding proteins
(PBPs) (4). PBPs include membrane-associated enzymes
important in bacterial peptidoglycan assembly and mainte- Fig. 2 Comparison of the β-lactam penicillin and the d-Ala-d-Ala
nance. Covalent modification of this subclass of PBPs by peptidoglycan terminus
β-lactams blocks their enzymatic activity thereby inhibiting
cell wall metabolism, which results in impaired wall integ-
rity and cell death. PBPs include the transpeptidases and
dd-carboxypeptidases that act on the pentapeptide portion bond of the pentapeptide (Fig. 1). Strominger noted
of the peptidoglycan repeating unit that consists of the 40 years ago that the β-lactam antibiotics sterically and
disaccharyl unit, N-acetylglucosamine-N-acetylmuramic electronically mimic the acyl-d-Ala-d-Ala terminus of the
acid, to which a d-Ala-d-Ala terminating pentapeptide is pentapeptide (Fig. 2) (5). This model overlaps the highly
linked through the lactyl group of N-acetylmuramic acid strained (and thus chemically reactive) β-lactam ring over
(Fig. 1). Transpeptidases and dd-carboxypeptidases use the scissile d-Ala-d-Ala peptide bond. Attack of the nucleo-
canonical Ser hydrolase chemistry to either rigidify the cell philic Ser hydroxyl onto the β-lactam ring carbonyl opens
wall by synthesizing interstrand peptidoglycan crosslinks the cyclic structure and generates a covalent intermediate
between the d-Ala-d-Ala termini of adjacent peptidoglycan that is resistant to hydrolysis (Fig. 3), thereby chemically
strands (transpeptidases) or control cell wall strength and titrating PBPs into inactive complexes and shutting down
flexibility by cleaving the terminal d-Ala-d-Ala peptide cell wall synthesis.
8 Biochemical Logic of Antibiotic Inactivation and Modification 83

In order to overcome the action of cytotoxic β-lactams, employed by metallo-proteases to cleave the reactive
bacteria have evolved secreted enzymes that hydrolytically β-lactam ring.
cleave the β-lactam ring of penicillins and cephalosporins (6)
(Fig. 4). The molecular logic of this resistance mechanism
therefore involves the destruction of the reactive “warhead”
of the β-lactam antibiotics, thereby eliminating the essential 2.2 Fosfomycin
chemical structure necessary for PBP inactivation.
These hydrolytic enzymes, appropriately named Destruction of a reactive chemical warhead is also employed
β-lactamases, fall into two general structural classes: Ser by enzymes that inactivate fosfomycin. The key structural
β-lactamases and metallo-β-lactamases (Fig. 4). The former element of this antibiotic is a reactive epoxide that is attacked
group share structural homology with the dd-carboxypeptidases by its intracellular target, the cell wall biosynthetic enzyme
and operate by similar Ser hydrolase chemistry. However the MurA (Figs. 5 and 6a). This enzyme is essential for synthe-
hydrolytic step, which is slow in PBPs, is fast in β-lactamases, sis of N-acetylmuramic acid and covalent modification of a
resulting in highly efficient detoxification of the antibiotics. key Cys residue by fosfomycin efficiently inactivates the
Metallo-β-lactamases adopt analogous hydrolytic chemistry enzyme.

Fig. 3 Mechanism of action of β-lactam antibiotics on bacterial trans- β-lactam ring. The active site machinery of transpeptidases or dd-
peptidases and dd-carboxypeptidases. Nucleophilic attack of the PBP carboxypeptidases is effectively captured as the subsequent covalent
Ser hydroxyl on the β-lactam ring carbonyl results in an opening of the intermediate cannot be hydrolyzed

Fig. 4 Mechanisms of enzymatic inactivation of β-lactam antibiotics. which is quickly hydrolyzed. (b) Metallo-β-lactamases utilize a bound
β-lactamases catalyze the hydrolytic cleavage of β-lactam rings. (a) Zn2+ to activate water for hydrolytic attack of the β-lactam ring
Serine-β-lactamases form a transient enzyme-antibiotic intermediate,
84 V. D’Costa and G.D. Wright

The bacterial countermeasure to inactivate this antibi- vicinal diol (8) (Fig. 6b). The second is via a thiol-
otic is an epoxide ring opening reaction using one of two dependent ring opening by enzymes that use abundant
distinct chemical tactics. The first, catalyzed by FosX, is a intracellular thiols such as glutathione (FosA) (9) (Fig. 6c)
metal-dependent hydrolytic process that generates the and cysteine (FosB) (10) (Fig. 6d). Either strategy results
in efficient destruction of the antibiotic’s reactive centre,
thereby blocking its action on the target MurA.

2.3 Macrolide Antibiotics

The macrolide antibiotics include natural products such as


erythromycin and semi-synthetic derivatives (e.g., clarithro-
mycin). These antibiotics are assembled via a polyketide
assembly line, cyclized to form a macrolactone ring struc-
ture, and subsequently modified by glycosylation to generate
a mature antibiotic (11). Macrolides inhibit bacterial transla-
tion by binding to the large ribosomal subunit in the vicin-
ity of the peptide exit tunnel (12). This interaction requires
an intact cyclic macrolide ring and in most cases the amino
sugar desosamine (Fig. 7).
Enzymatic resistance to macrolide antibiotics occurs
either by modification of the desosamine sugar (see Sect. 3
below) or by linearization of the macrolactone ring (Fig. 8).
The latter mechanism is catalyzed by esterases that hydro-
lytically cleave the lactone resulting in ring opening and
consequently the inability to effectively bind to the peptide
Fig. 5 Interactions of fosfomycin with its bacterial target, MurA.
exit tunnel. The erythromycin esterases EreA and EreB
Fosfomycin forms a covalent bond with MurA’s active site Cys.
Additional interactions with MurA are designated as arrows. MurA have been identified in E. coli integrons and R-plasmids
residues are labeled in grey. Adapted from (7) (14–17).

Fig. 6 Mechanism of fosfomycin action and


inactivation. (a) Fosfomycin targets the active-site
Cys residue of MurA, forming a covalent
intermediate. (b) FosX-mediated inactivation of
fosfomycin results in the formation of a diol.
(c) The product of FosA-mediated fosfomycin
inactivation is a glutathione-fosfomycin adduct.
(d) FosB-mediated resistance to fosfomycin
results in a ring-opened inactivated product with
free Cys
8 Biochemical Logic of Antibiotic Inactivation and Modification 85

Fig. 7 Interactions of the macrolide


erythromycin with the bacterial ribosomal
RNA. Key 23S rRNA residues are shown
in grey and the interactions are designated
as arrows. The hydroxyl group that serves
as a site of inactivation interacts with
A2058. Adapted from (13)

Fig. 8 Inactivation of the macrolide


erythromycin by hydrolysis.
Macrolides can be inactivated by
hydrolysis of the macrolactone ring.
This reaction is mediated by the
esterase Ere

2.4 Type B Streptogramins occurs through a ring-opening reaction catalyzed by the


enzyme Vgb. Vgb, originally identified in streptogramin
The streptogramins are natural product inhibitors of bacte- resistant Staphylococcus aureus, cleaves the cyclic peptide
rial translation that consist of two structurally distinct (21), resulting in depsipeptide linearization. The resulting
classes, denoted as Type A and Type B (18). Type B strepto- structure no longer exhibits affinity for the bacterial ribo-
gramins are cyclic depsipeptides that, like macrolides, bind some, mirroring the biochemical logic of macrolide
to a region of the bacterial ribosome’s peptide exit tunnel esterases. However, the mechanism of ring opening is quite
(19, 20) (Fig. 9). Type A streptogramins are mixed distinct. Rather than causing a hydrolytic reaction at the
peptide-polyketide antibiotics that bind to the peptidyl- thermodynamically vulnerable ester bond of Type B strep-
transferase centre of the ribosome. Enzymatic resistance to togramins, Vgb catalyzes a lyase reaction that results in a
Type A streptogramins occurs via an acetyltransfer mecha- ring opening of the peptide by a C–O cleavage strategy
nism, while enzymatic resistance to Type B streptogramins (22) (Fig. 10).
86 V. D’Costa and G.D. Wright

2.5 Tetracycline efflux and ribosomal protection (23, 25). However, an enzy-
matic mechanism of tetracycline resistance, originally dis-
covered in Bacteroides (26), has been identified that
The tetracycline antibiotics have found extensive clinical
inactivates the antibiotic via an oxygen-dependent process.
use for almost half a century. This class of antibiotics binds
Purification of the enzyme that catalyzes this reaction, TetX,
divalent metals and acts by blocking bacterial translation by
followed by careful analysis of the products of the reaction
binding to the small ribosomal subunit (23) (Fig. 11). The
showed that the enzyme first facilitates mono-hydroxylation
principal mechanisms of clinical tetracycline resistance are
of the antibiotic at position 11a, effectively disrupting the
essential metal-binding site on the molecule (27) (Fig. 12).
Furthermore, this step triggers a nonenzymatic decomposi-
tion of the antibiotic to a form of unknown structure that
turned the growth media black. This enzyme is also capable
of mono-hydroxylation of the latest generation of tetracy-
cline antibiotics, the glycylcyclines, resulting in resistance,
but not the subsequent nonenzymatic decomposition of the
antibiotic (28).

3 Antibiotic Modification

The most diverse class of resistance enzymes catalyzes the


covalent modification of antibiotics. This strategy confers
resistance by means of group transfer and includes both O-
and N-acetylation, O-phosphorylation, O-nucleotidylylation,
O-ribosylation, and O-glycosylation. Covalent modification
of antibiotics by this class of enzymes does not destroy the
essential active warheads of the compounds, as described in
the previous section, but rather obstructs interaction of the
antimicrobial with its target. This is accomplished by func-
tionally derivatizing the antibiotic at structural location(s)
that play an essential role in binding with the target. By doing
Fig. 9 Streptogramin B interactions with bacterial ribosome.
Quinupristin, a Type B Streptogramin, binds to the bacterial ribosome’s so, key interactions (e.g., hydrogen bonding, ionic inter-
polypeptide exit tunnel. Key interactions with the 23S rRNA are desig- actions, steric complementarity, etc.) are disrupted by the
nated as arrows. Adapted from (19, 20) introduction of the modifying group, resulting in an overall

Fig. 10 Vgb-catalyzed inactivation of the type B streptogramin quinupristin. Quinupristin undergoes a ring-opening elimination reaction, result-
ing in an inactive derivative
8 Biochemical Logic of Antibiotic Inactivation and Modification 87

Fig. 11 Interactions of tetracycline with the


bacterial 16S rRNA. Interactions are
designated by grey arrows and key ribosomal
RNA residues are indicated in grey. Adapted
from (24)

Fig. 12 TetX-mediated
inactivation of tetracycline.
TetX catalyzes the
hydroxylation of the
antibiotic, which interferes
with the metal-binding site
required for activity.
Adapted from (27)

decrease in the affinity of the antibiotic derivative for its tar- 3.1 Aminoglycosides
get in comparison to the unmodified counterpart.
This antibiotic inactivation tactic requires the presence of The aminoglycoside class of antibiotics is a diverse group of
a co-substrate for enzyme activity, such as acetyl-CoA, ATP hydrophilic aminocyclitols modified by amino and neutral
or UDP-glucose. Consequently, enzyme activity is localized sugars that consist of both natural products and their
to the bacterial cytosol. The inactivation products are com- semi-synthetic derivatives. Polycationic aminoglycoside
monly stable in the cellular environment, thus the reactions antibiotics, as previously mentioned, act by interacting with
are considered to be irreversible in the absence of an enzyme the 16S rRNA region of the bacterial ribosome’s A-site,
that counteracts the reaction. However it is conceivable that impairing its decoding mechanism and consequently result-
the presence of such reversing enzymes (e.g., phosphatases, ing in a misreading of the mRNA (29–32). X-Ray crystallo-
acylases) can undo resistance in vivo. graphic studies of aminoglycoside antibiotics and the small
88 V. D’Costa and G.D. Wright

ribosomal subunit or fragments of the 16S rRNA reveal that 3.1.1 Aminoglycoside Acetyltransferases
interactions between aminoglycosides and the ribosome span (AAC Family)
the entire length of the antibiotic (24, 33–36). The primary
mode of interaction is through predicted hydrogen bonding Aminoglycoside acetyltransferases (AACs) utilize intracellu-
and ionic contacts between the antibiotic amino and hydroxyl lar acetyl-CoA as a co-substrate, catalyzing the formation of
groups and the 16S rRNA (Fig. 13). a biologically stable amide with the aminoglycoside. Although
The most prevalent mode of clinically relevant aminogly- AACs primarily modify amino groups (N-acetylation),
coside resistance is via enzymatic modification (38). Three O-acetylation has been documented with the acetyltransferase
classes of enzymes, whose reactions differ with respect to domain of the bifunctional enzyme AAC(6′)-APH(2′′) (42)
the functional group transferred and the acceptor site, are res- and the mycobacterial enzyme AAC(2′)-Ic (43).
ponsible for aminoglycoside modification. Aminoglycoside AACs are members of the GCN5 superfamily of proteins
acetyltransferases (AACs) modify amino groups, aminogly- (44, 45). Although all enzymes of this class do not exhibit
coside phosphotransferases (APHs) target hydroxyl groups, significant primary sequence homology or conserved cata-
and aminoglycoside nucleotidyltransferases (ANTs) modify lytic residues, analysis of available X-ray crystal structures
hydroxyl groups (Fig. 14). There are numerous examples of of four enzymes (AAC(6′Ii), AAC(3)-Ia, AAC(2′)-Ic, and
each group and the genes encoding aminoglycoside-modify- AAC(6′)-Iy) indicates that the aminoglycoside binding
ing enzymes are commonly located on mobile genetic pocket commonly contains a highly negatively charged sur-
elements such as plasmids or transposons, although some face to accommodate the polycationic antibiotic (45–48).
have been identified within chromosomal DNA (39–41). AACs are further classified based on the site of acetyla-
The action of all three classes of modifying enzyme changes tion along the aminoglycoside structure. By convention, the
the electronic properties of the antibiotic, in addition to its position along the amino sugar/aminocyclitol targeted is
size and structure. These alterations result in steric and indicated in brackets, and the amino sugar/aminocyclitol
electronic clashes between the modified antibiotic and the modified is designated in brackets after the position of attack.
16S rRNA, impairing efficient binding and resulting in For example, in Fig. 14, AAC (3) indicates acetylation of the
resistance. 3-position of the central aminocyclitol moiety, the term (2′)

Fig. 13 Interactions of the aminoglycoside


gentamicin C1a with the bacterial 16S rRNA. Key
16S rRNA residues are shown in grey and the
interactions are designated as arrows. Adapted
from (37)
8 Biochemical Logic of Antibiotic Inactivation and Modification 89

Fig. 14 Inactivation of the amino-


glycoside gentamicin C1a by
aminoglycoside-modifying enzymes.
Aminoglycosides can be modified by
the addition of acetyl groups,
phosphate groups or AMP moieties.
These enzymatic reactions are
catalyzed by aminoglycoside
acetyltransferases (AACs), phospho-
transferases (APHs), and nucleotidyl-
transferases (ANTs) respectively.
(a) Sites of aminoglycoside
inactivation. Groups targeted are
labeled by the corresponding
resistance enzymes. (b) The products
of aminoglycoside acetyltransferases,
phosphotransferases, and
nucleotidyltransferases

suggests modification of the 2-position of the 4-substituent 3.1.3 Aminoglycoside Nucleotidyltransferases


diaminohexose, and (2′′) indicates modification of the (ANT Family)
2-position of the 6-substituent aminohexose.
Modification of aminoglycosides by AACs results in neu- Aminoglycoside nucleotidyltransferases utilize the co-substrate
tralization of the positive charge on the target amino group, ATP to transfer an AMP moiety to selected aminoglycoside
eliminating key ionic interactions and sterically blocking hydroxyl groups. This class of inactivating enzymes has been
interaction with the 16S rRNA. identified in some Gram-positive bacterial isolates, as well as
a broad range of Gram-negatives (40).
ANTs display very little primary sequence homology,
3.1.2 Aminoglycoside Phosphotransferases however they exhibit a common core signature region (49).
(APH Family) The enzyme ANT(4′)-Ia has been crystallized and its atomic
structure determined alone and in complex with the sub-
Aminoglycoside phosphotransferases catalyze the phosphory- strates kanamycin and a nonhydrolyzable ATP analogue (52,
lation of specific aminoglycoside hydroxyl residues (Fig. 14), 53). Although the primary sequence homology is only 10%,
using intracellular ATP as a phosphate donor. Classification of the putative active site was determined to be structurally
phosphotransferases is based on the site of action, analogous equivalent to that of rat DNA-polymerase β, one of the small-
to the system described above for acetyltransferases. The APH est and simplest of the polymerases (54), and catalyzes a
enzymes are subdivided into seven classes, based on their site similar chemical reaction.
of action on the aminoglycoside: APH(2′′), APH(3′), APH(3′′), Paralleling the strategies of the other classes of aminogly-
APH(4), APH(6), APH(7′′), and APH(9). There exists very coside modifying enzymes, the action of ANTs causes a
little primary sequence homology among the subclasses of change in antibiotic structure that results in both a steric and
APHs; however common signature sequences and residues electronic clash between the antibiotic and its target. This
essential for catalysis are evident (49). theme and molecular logic finds other examples in antibiotic
The largest subclass of APHs modifies the 3′-hydroxyl of resistance as outlined below.
the aminoglycoside and is consequently called APH(3′) (49).
Crystal structure analysis of the enzyme APH(3′)-IIIa bound
to ADP has established a remarkable similarity to known
protein kinases, despite the low primary sequence similarity 3.2 Macrolides
(50). This may be evidence that APH(3′) and protein kinases
evolved from a common ancestor. Macrolide antibiotics are a large class of antibiotics that
Modification of aminoglycosides by APH-catalyzed include both natural products and semi-synthetic derivatives.
phosphorylation results in changes in overall charge and size Most macrolides are derived from bacterial fermentation
of the antibiotic. This results in electronic and steric clashes products, particularly from species of the actinomycete genus
with the 16S rRNA and a 103-fold impairment of binding to Streptomyces. Erythromycin was the first member of this
the target 16S rRNA (51). class to be identified (1952), a natural product of Streptomyces
90 V. D’Costa and G.D. Wright

erythrae (now known as Saccharopolyspora erythrae). The hydroxyl residue of the desoamine sugar plays a key role in
name macrolide is derived from the macrolactone ring that the interaction of the macrolide with its target rRNA.
characterizes the class, which can consist of 14–16 members The second mode of enzymatic macrolide inactivation
and is commonly attached to one or two sugar moieties. occurs by modification of this essential desosamine sugar
Macrolides have found an important role in the treatment of (Fig. 15). Modification of the 2′ hydroxyl residue can occur
clinical pathogens. Since their introduction in the 1950s, efforts by either phosphorylation or glycosylation. This hydroxyl
to expand the spectrum of activity and deal with the inevitable group, as mentioned, plays an important role in macrolide-
resistance that followed have resulted in a number of different target binding, serving as a multiple contact site of hydrogen
classes of derivatives. Azalides incorporate an endocyclic bonding with the 23S rRNA (Fig. 7). Modification of the
nitrogen into the macrolactone ring. Azithromycin, the first antibiotic at this site therefore results in loss of vital struc-
azalide approved for clinical use, exhibits increased potency tural connections with the target and also results in steric
against a number of Gram-negative organisms, as well as a impairment of complex formation.
longer apparent half-life. Ketolides, which have a keto group
in place of the l-cladinose in the 3-position, exhibit increased
activity against a number of macrolide resistant strains. 3.2.1 Macrolide Kinases (Mph Family)
Macrolides, as described previously, act by binding with
the 23S rRNA of the bacterial 50S ribosomal subunit adja- Clinical resistance to macrolides has been documented by
cent to the peptide exit tunnel, blocking polymerization at means of phosphate transfer from ATP by a family of
the peptidyltransferase centre and inducing premature pep- macrolide-inactivating phosphotransferases, encoded by the
tide dissociation (13, 20). Interactions with the ribosomal mph genes (Fig. 15). These enzymes have been identified in
RNA occur primarily through hydrogen bonding, as shown both Gram-positive and Gram-negative pathogens (55–57).
with erythromycin in Fig. 7. Much of the hydrogen bonding Members of the Mph class of resistance enzymes appear to
ability of macrolides can be attributed to their hydroxyl and be extremely diverse with respect to the nucleotide sequences
amino groups, which interact with the nitrogenous bases or that encode the enzymes. The gene mphA exhibits a 66% G + C
backbone phosphate groups of the rRNA. As shown, the content, uncharacteristically high for the organism it was

Fig. 15 Inactivation of the


macrolide erythromycin by
Mph and Mgt. Macrolides
can be modified by the
addition of phosphate and
glucose moieties. The
hydroxyl group targeted and
the subsequent modifica-
tions are labeled in grey
8 Biochemical Logic of Antibiotic Inactivation and Modification 91

originally identified in (E. coli, G + C content approximately


50%) (58). The sequences of mphB and mphC, conversely,
display a G + C content of only 38%. The structure of these
enzymes has yet to be elucidated, however they share canoni-
cal phosphate transfer residues with APHs and probably
resemble these aminoglycoside resistance enzymes.

3.2.2 Macrolide Glycosyltransferases (Mgt Family)

Resistance to macrolides in antibiotic-producing strains of


bacteria as well as other soil-dwelling organisms is com-
monly accomplished by intracellular glycosylation of the
antibiotic prior to export. This is catalyzed by a class of
enzymes called macrolide glycosyltransferases (Fig. 15).
Members of this class include Mgt, from the nonmacrolide-
producing Streptomyces lividans (59, 60), as well as OleD
(61) and GimA (62) from the macrolide-producing S. antibi-
oticus and S. ambofaciens respectively.
Members of the Mgt family are extremely similar with
respect to both DNA and primary amino acid sequences,
however each enzyme appears to display a unique substrate
specificity in vitro (63).
Fig. 16 Interactions of the rifampicin with the bacterial β-subunit of
RNA polymerase. Key amino acid residues are shown in grey and the
interactions are designated as arrows. Adapted from (64)
3.3 Rifamycins

The rifamycin family of antibiotics includes semisynthetic


derivatives of a natural product synthesized by the actinomy-
cete Amycolatopsis mediterranei. Rifampicin was first intro- 3.3.1 ADP-Ribosyltransferases (ARR Family)
duced in 1968, but the most widely used member of the group
is rifamycin, which has become an integral component of the Although both eukaryotic and prokaryotic proteins are com-
multiantibiotic gold-standard treatment for Mycobacterium monly modified by means of ADP-ribosyl transfer, this
tuberculosis infections. mechanism of antibiotic resistance has so far only been doc-
Rifamycins target the bacterial β-subunit of RNA poly- umented for the rifamycin class. Resistance to rifampicin by
merase. The crystal structure of rifampicin bound to the RNA means of ADP-ribosylation has been documented in numer-
polymerase of Thermus aquaticus has been determined to ous nontuberculosis Mycobacterium strains, such as M. smeg-
3.3Å (64). Twelve amino acid residues were shown to asso- matis. Modification is due to a unique ADP-ribosyltransferase
ciate closely with rifampicin, six of which participate in known as ARR (65). Another inactivating ribosyltransferase
hydrogen bonding, as shown in Fig. 16. The majority of these (ARR-2) with 55% identity to ARR has been identified in a
interactions occur at four crucial hydroxyl residues on the multidrug resistant integron in a Gram-negative Acinetobacter
rifampicin molecule including a key interaction between the strain (66).
hydroxyl group at position 23 and the amide of Phe394. The resistance enzymes from the ARR family are
Resistance to rifampicin commonly occurs through characteristically small (approximately 200 amino acids) and
amino acid mutations in the RNA polymerase β-subunit. do not display sequence similarity to protein ADP-ribosyl
However, inactivating enzymes have also evolved to modify transferases. They target the hydroxyl residue at position 23
the antibiotic (Fig. 17). Group transfer can result in ADP- of rifampicin (Fig. 17), and utilize nicotinamide adenine
ribosylation, phosphate addition, and glycosylation of the dinucleotide (NAD+) as a donor for the ADP-ribosyl moiety.
rifampicin’s 23-hydroxyl. Through the addition of a bulky It has also been shown that this ADP-ribosylated antibiotic
functional group, rifampicin’s tight binding to its target is can undergo subsequent decomposition to release the ADP
impaired. moiety (65, 67, 68).
92 V. D’Costa and G.D. Wright

Fig. 17 Inactivation of rifampicin. Rifampicin can be enzymatically modified by the addition of ADP-ribose, glucose, and phosphate moieties.
The hydroxyl group targeted and the subsequent modifications are labeled in grey

3.3.2 Rifampicin Kinases “warhead” or “active site” of the antibiotic (e.g., cleavage of
the β-lactam ring by β-lactamases), or mechanisms that
Inactivation of rifampicin by phosphorylation (Fig. 17) has modify key structural elements that are essential for binding
been documented by species of Nocardia (69, 70), Rhodo- of the antibiotic to target (e.g., phosphorylation of aminogly-
coccus (71), as well as Bacillus (72). The kinases responsible cosides). The molecular logic of these approaches is revealed
for this inactivation have yet to be identified or studied. with knowledge of the interaction of the active antibiotic
Phosphorylation of rifampicin’s hydroxyl at position 23 logi- with its cellular target. Study of enzymatic resistance there-
cally impedes interaction with the RNA polymerase target, fore not only can inform on molecular aspects of antibiotic-
although little has been done to elucidate the details of this target interactions, but can serve to guide target identification
mechanism. where this is not yet known.
Another spin-off of the study of these mechanisms is the
opportunity to develop strategies to overcome the resistance
3.3.3 Rifampicin Glycosyltransferases activity. For example, the observation that aminoglycosides
were inactivated by phosphorylation of the hydroxyl group at
Glycosylation of the 23-position of rifampicin has also been position 3′ of the 6′-aminohexose ring guided the development
reported in Nocardia species (69, 73) (Fig. 17). Glycosyla- of antibiotics such as tobramycin, which lack this hydroxyl and
tion at this position prevents hydrogen bonding with the which were consequently resistant to this mechanism. A second
23-hydroxyl, hindering effective target binding to RNA approach is to develop inhibitors of resistance enzymes. This
polymerase β. The genes encoding the enzymes that catalyze strategy has been very successful in the β-lactam arena where
these reactions have yet to be elucidated. combinations of an antibiotic and a resistance enzyme inhibi-
tor, such as amoxicillin and clavulanic acid respectively
(Augmentin), have emerged as billion-dollar drugs.
Finally, antibiotic-modifying enzymes also have the
4 Summary and Conclusions opportunity to be exploited as novel reagents in antibiotic
semi-synthesis as protecting agents. In some cases, antibiotic-
Bacteria use enzymes to strategically incapacitate and neu- modifying proteins are employed by antibiotic-producing
tralize antibiotics. Tactically this includes deployment of bacteria as a means of self-protection. For example, during
mechanisms that either destroy the essential chemical streptomycin biosynthesis in Streptomyces griseus, the
8 Biochemical Logic of Antibiotic Inactivation and Modification 93

enzyme StrA modifies mature antibiotic to the inactive 14. Arthur, M., Autissier, D. & Courvalin, P. (1986). Analysis of the
6-phosphoderivative. Export of this “pro-drug” is followed nucleotide sequence of the ereB gene encoding the erythromycin
esterase type II. Nucleic Acids Res 14, 4987–4999
by unmasking of the cytotoxic agent by an extracellular 15. Biskri, L. & Mazel, D. (2003). Erythromycin esterase gene ere(A)
phosphatase (74). These enzymes could serve as reagents to is located in a functional gene cassette in an unusual class 2 inte-
chemically protect and deprotect sensitive structural elements gron. Antimicrob Agents Chemother 47, 3326–3331
in the synthesis of libraries of semi-synthetic antibiotics. 16. Ounissi, H. & Courvalin, P. (1985). Nucleotide sequence of the
gene ereA encoding the erythromycin esterase in Escherichia coli.
Enzymatic resistance therefore provides both challenges Gene 35, 271–278
and opportunities in new drug development. Through a com- 17. Plante, I., Centron, D. & Roy, P. H. (2003). An integron cassette
bination of rigorous biochemical analysis and parallel efforts encoding erythromycin esterase, ere(A), from Providencia stuartii.
in the determination of enzyme structure and target identifi- J Antimicrob Chemother 51, 787–790
18. Mukhtar, T. A. & Wright, G. D. (2005). Streptogramins, oxazolidi-
cation, new approaches that circumvent these selective and nones, and other inhibitors of bacterial protein synthesis. Chem Rev
potent agents can be developed to extend antibiotic lifetime 105, 529–542
and efficacy. 19. Harms, J. M., Schlunzen, F., Fucini, P., Bartels, H. & Yonath, A.
(2004). Alterations at the peptidyl transferase centre of the ribos-
ome induced by the synergistic action of the streptogramins dalfo-
pristin and quinupristin. BMC Biol 2, 4
20. Tu, D., Blaha, G., Moore, P. B. & Steitz, T. A. (2005). Structures
References of MLSBK antibiotics bound to mutated large ribosomal subu-
nits provide a structural explanation for resistance. Cell 121,
257–270
1. Abraham, E. P. & Chain, E. (1940). An enzyme from bacteria able 21. Allignet, J., Loncle, V., Mazodier, P. & el Solh, N. (1988).
to destroy penicillin. Nature 146, 837 Nucleotide sequence of a staphylococcal plasmid gene, vgb, encod-
2. Robicsek, A., Strahilevitz, J., Jacoby, G. A., Macielag, M., ing a hydrolase inactivating the B components of virginiamycin-
Abbanat, D., Hye Park, C., Bush, K. & Hooper, D. C. (2006). like antibiotics. Plasmid 20, 271–275
Fluoroquinolone-modifying enzyme: a new adaptation of a com- 22. Mukhtar, T. A., Koteva, K. P., Hughes, D. W. & Wright, G. D.
mon aminoglycoside acetyltransferase. Nat Med 12, 83–88 (2001). Vgb from Staphylococcus aureus inactivates strepto-
3. Walsh, C. T. (2003). Antibiotics Action, Origins, Resistance, ASM gramin B antibiotics by an elimination mechanism not hydrolysis.
Press, Washington, DC Biochemistry 40, 8877–8886
4. Fisher, J. F., Meroueh, S. O. & Mobashery, S. (2005). Bacterial 23. Chopra, I., Hawkey, P. M. & Hinton, M. (1992). Tetracyclines, molec-
resistance to beta-lactam antibiotics: compelling opportunism, ular and clinical aspects. J Antimicrob Chemother 29, 245–277
compelling opportunity. Chem Rev 105, 395–424 24. Brodersen, D. E., Clemons, W. M., Jr., Carter, A. P., Morgan-
5. Tipper, D. J. & Strominger, J. L. (1965). Mechanism of action of Warren, R. J., Wimberly, B. T. & Ramakrishnan, V. (2000). The
penicillins: a proposal based on their structural similarity to acyl- structural basis for the action of the antibiotics tetracycline,
d-alanyl-d-alanine. Proc Natl Acad Sci USA 54, 1133–1141 pactamycin, and hygromycin B on the 30 S ribosomal subunit. Cell
6. Bush, K., Jacoby, G. A. & Medeiros, A. A. (1995). A functional 103, 1143–1154
classification scheme for beta-lactamases and its correlation with 25. Poole, K. (2005). Efflux-mediated antimicrobial resistance. J Antimi-
molecular structure. Antimicrob Agents Chemother 39, 1211–1233 crob Chemother 56, 20–51
7. Skarzynski, T., Mistry, A., Wonacott, A., Hutchinson, S. E., Kelly, 26. Speer, B. S. & Salyers, A. A. (1988). Characterization of a novel
V. A. & Duncan, K. (1996). Structure of UDP-N-acetylglucosamine tetracycline resistance that functions only in aerobically grown
enolpyruvyl transferase, an enzyme essential for the synthe- Escherichia coli. J Bacteriol 170, 1423–1429
sis of bacterial peptidoglycan, complexed with substrate UDP- 27. Yang, W., Moore, I. F., Koteva, K. P., Bareich, D. C., Hughes,
N-acetylglucosamine and the drug fosfomycin. Structure 4, D. W. & Wright, G. D. (2004). TetX is a flavin-dependent monoox-
1465–1474 ygenase conferring resistance to tetracycline antibiotics. J Biol
8. Fillgrove, K. L., Pakhomova, S., Newcomer, M. E. & Armstrong, Chem 279, 52346–52352
R. N. (2003). Mechanistic diversity of fosfomycin resistance in 28. Moore, I. F., Hughes, D. W. & Wright, G. D. (2005). Tigecycline
pathogenic microorganisms. J Am Chem Soc 125, 15730–15731 is modified by the flavin-dependent monooxygenase TetX.
9. Bernat, B. A., Laughlin, L. T. & Armstrong, R. N. (1997). Biochemistry 44, 11829–11835
Fosfomycin resistance protein (FosA) is a manganese metal- 29. Davies, J. & Davis, B. D. (1968). Misreading of ribonucleic acid
loglutathione transferase related to glyoxalase I and the extradiol code words induced by aminoglycoside antibiotics. The effect of
dioxygenases. Biochemistry 36, 3050–3055 drug concentration. J Biol Chem 243, 3312–3316
10. Cao, M., Bernat, B. A., Wang, Z., Armstrong, R. N. & Helmann, J. D. 30. Edelmann, P. & Gallant, J. (1977). Mistranslation in E. coli. Cell
(2001). FosB, a cysteine-dependent fosfomycin resistance protein 10, 131–137
under the control of sigma(W), an extracytoplasmic-function sigma 31. Moazed, D. & Noller, H. F. (1987). Interaction of antibiotics with
factor in Bacillus subtilis. J Bacteriol 183, 2380–2383 functional sites in 16S ribosomal RNA. Nature 327, 389–394
11. Katz, L. & McDaniel, R. (1999). Novel macrolides through genetic 32. Woodcock, J., Moazed, D., Cannon, M., Davies, J. & Noller, H. F.
engineering. Med Res Rev 19, 543–558 (1991). Interaction of antibiotics with A- and P-site-specific bases
12. Schlunzen, F., Zarivach, R., Harms, J., Bashan, A., Tocilj, A., in 16S ribosomal RNA. EMBO J 10, 3099–3103
Albrecht, R., Yonath, A. & Franceschi, F. (2001). Structural basis 33. Carter, A. P., Clemons, W. M., Brodersen, D. E., Morgan-Warren,
for the interaction of antibiotics with the peptidyl transferase centre R. J., Wimberly, B. T. & Ramakrishnan, V. (2000). Functional
in eubacteria. Nature 413, 814–821 insights from the structure of the 30 S ribosomal subunit and its
13. Hansen, J. L., Ippolito, J. A., Ban, N., Nissen, P., Moore, P. B. & interactions with antibiotics. Nature 407, 340–348
Steitz, T. A. (2002). The structures of four macrolide antibiotics 34. Ogle, J. M., Brodersen, D. E., Clemons, W. M., Jr., Tarry, M.
bound to the large ribosomal subunit. Mol Cell 10, 117–128 J., Carter, A. P. & Ramakrishnan, V. (2001). Recognition of
94 V. D’Costa and G.D. Wright

cognate transfer RNA by the 30 S ribosomal subunit. Science 292, 53. Sakon, J., Liao, H. H., Kanikula, A. M., Benning, M. M.,
897–902 Rayment, I. & Holden, H. M. (1993). Molecular structure of
35. Vicens, Q. & Westhof, E. (2002). Crystal structure of a complex kanamycin nucleotidyltransferase determined to 3.0-Å resolution.
between the aminoglycoside tobramycin and an oligonucleotide Biochemistry 32, 11977–11984
containing the ribosomal decoding a site. Chem Biol 9, 747–755 54. Holm, L. & Sander, C. (1995). DNA polymerase beta belongs to an
36. Vicens, Q. & Westhof, E. (2003). Crystal structure of geneticin ancient nucleotidyltransferase superfamily. Trends Biochem Sci 20,
bound to a bacterial 16S ribosomal RNA A site oligonucleotide. J 345–347
Mol Biol 326, 1175–1188 55. Matsuoka, M., Endou, K., Kobayashi, H., Inoue, M. & Nakajima, Y.
37. Hobbie, S. N., Pfister, P., Brull, C., Westhof, E. & Bottger, E. C. (1998). A plasmid that encodes three genes for resistance to mac-
(2005). Analysis of the contribution of individual substituents rolide antibiotics in Staphylococcus aureus. FEMS Microbiol Lett
in 4,6-aminoglycoside-ribosome interaction. Antimicrob Agents 167, 221–227
Chemother 49, 5112–5118 56. Noguchi, N., Katayama, J. & O’Hara, K. (1996). Cloning
38. Wright, G. D. (1999). Aminoglycoside-modifying enzymes. Curr and nucleotide sequence of the mphB gene for macrolide
Opin Microbiol 2, 499–503 2′-phosphotransferase II in Escherichia coli. FEMS Microbiol Lett
39. Magnet, S. & Blanchard, J. S. (2005). Molecular insights into 144, 197–202
aminoglycoside action and resistance. Chem Rev 105, 477–498 57. O’Hara, K., Kanda, T., Ohmiya, K., Ebisu, T. & Kono, M. (1989).
40. Shaw, K. J., Rather, P. N., Hare, R. S. & Miller, G. H. (1993). Purification and characterization of macrolide 2′-phosphotransferase
Molecular genetics of aminoglycoside resistance genes and from a strain of Escherichia coli that is highly resistant to erythro-
familial relationships of the aminoglycoside-modifying enzymes. mycin. Antimicrob Agents Chemother 33, 1354–1357
Microbiol Rev 57, 138–163 58. Noguchi, N., Emura, A., Matsuyama, H., O’Hara, K., Sasatsu, M. &
41. Wright, G. D., Berghuis, A. M. & Mobashery, S. (1998). Kono, M. (1995). Nucleotide sequence and characterization of
Aminoglycoside antibiotics. Structures, functions, and resistance. erythromycin resistance determinant that encodes macrolide
Adv Exp Med Biol 456, 27–69 2′-phosphotransferase I in Escherichia coli. Antimicrob Agents
42. Daigle, D. M., Hughes, D. W. & Wright, G. D. (1999). Prodigious Chemother 39, 2359–2363
substrate specificity of AAC(6′)-APH(2′′), an aminoglycoside 59. Jenkins, G. & Cundliffe, E. (1991). Cloning and characterization of
antibiotic resistance determinant in enterococci and staphylococci. two genes from Streptomyces lividans that confer inducible resist-
Chem Biol 6, 99–110 ance to lincomycin and macrolide antibiotics. Gene 108, 55–62
43. Hegde, S. S., Javid-Majd, F. & Blanchard, J. S. (2001). 60. Cundliffe, E. (1992). Glycosylation of macrolide antibiotics in
Overexpression and mechanistic analysis of chromosomally extracts of Streptomyces lividans. Antimicrob Agents Chemother
encoded aminoglycoside 2′-N-acetyltransferase (AAC(2′)-Ic) from 36, 348–352
Mycobacterium tuberculosis. J Biol Chem 276, 45876–45881 61. Quiros, L. M., Aguirrezabalaga, I., Olano, C., Mendez, C. &
44. Vetting, M. W., LP, S. d. C., Yu, M., Hegde, S. S., Magnet, S., Salas, J. A. (1998). Two glycosyltransferases and a glycosidase are
Roderick, S. L. & Blanchard, J. S. (2005). Structure and functions involved in oleandomycin modification during its biosynthesis by
of the GNAT superfamily of acetyltransferases. Arch Biochem Streptomyces antibioticus. Mol Microbiol 28, 1177–1185
Biophys 433, 212–226 62. Gourmelen, A., Blondelet-Rouault, M. H. & Pernodet, J. L. (1998).
45. Wybenga-Groot, L. E., Draker, K., Wright, G. D. & Berghuis, Characterization of a glycosyltransferase inactivating macrolides,
A. M. (1999). Crystal structure of an aminoglycoside 6′-N- encoded by gimA from Streptomyces ambofaciens. Antimicrob
acetyltransferase: defining the GCN5-related N-acetyltransferase Agents Chemother 42, 2612–2619
superfamily fold. Struct Fold Des 7, 497–507 63. Yang, M., Proctor, M. R., Bolam, D. N., Errey, J. C., Field, R. A.,
46. Vetting, M. W., Hegde, S. S., Javid-Majd, F., Blanchard, J. S. & Gilbert, H. J. & Davis, B. G. (2005). Probing the breadth of mac-
Roderick, S. L. (2002). Aminoglycoside 2′-N-acetyltransferase rolide glycosyltransferases: in vitro remodeling of a polyketide
from Mycobacterium tuberculosis in complex with coenzyme A antibiotic creates active bacterial uptake and enhances potency.
and aminoglycoside substrates. Nat Struct Biol 9, 653–658 J Am Chem Soc 127, 9336–9337
47. Vetting, M. W., Magnet, S., Nieves, E., Roderick, S. L. & 64. Campbell, E. A., Korzheva, N., Mustaev, A., Murakami, K., Nair, S.,
Blanchard, J. S. (2004). A bacterial acetyltransferase capable of Goldfarb, A. & Darst, S. A. (2001). Structural mechanism for
regioselective N-acetylation of antibiotics and histones. Chem Biol rifampicin inhibition of bacterial RNA polymerase. Cell 104,
11, 565–573 901–912
48. Wolf, E., Vassilev, A., Makino, Y., Sali, A., Nakatani, Y. & Burley, S. K. 65. Quan, S., Venter, H. & Dabbs, E. R. (1997). Ribosylative inac-
(1998). Crystal structure of a GCN5-related N-acetyltransferase: tivation of rifampin by Mycobacterium smegmatis is a principal
Serratia marcescens aminoglycoside 3-N-acetyltransferase. Cell contributor to its low susceptibility to this antibiotic. Antimicrob
94, 439–449 Agents Chemother 41, 2456–2460
49. Wright, G. D. & Thompson, P. R. (1999). Aminoglycoside phos- 66. Houang, E. T., Chu, Y. W., Lo, W. S., Chu, K. Y. & Cheng, A. F.
photransferases: proteins, structure, and mechanism. Front Biosci (2003). Epidemiology of rifampin ADP-ribosyltransferase (arr-2)
4, D9–D21 and metallo-beta-lactamase (blaIMP-4) gene cassettes in class 1
50. Hon, W. C., McKay, G. A., Thompson, P. R., Sweet, R. M., Yang, integrons in Acinetobacter strains isolated from blood cultures in
D. S., Wright, G. D. & Berghuis, A. M. (1997). Structure of an 1997 to 2000. Antimicrob Agents Chemother 47, 1382–1390
enzyme required for aminoglycoside antibiotic resistance reveals 67. Morisaki, N., Hashimoto, Y., Furihata, K., Imai, T., Watanabe, K.,
homology to eukaryotic protein kinases. Cell 89, 887–895 Mikami, Y., Yazawa, K., Ando, A., Nagata, Y. & Dabbs, E. R.
51. Llano-Sotelo, B., Azucena, E. F., Jr., Kotra, L. P., Mobashery, S. & (2000). Structures of ADP-ribosylated rifampicin and its metabo-
Chow, C. S. (2002). Aminoglycosides modified by resistance lite: intermediates of rifampicin-ribosylation by Mycobacterium
enzymes display diminished binding to the bacterial ribosomal smegmatis DSM43756. J Antibiot (Tokyo) 53, 269–275
aminoacyl-tRNA site. Chem Biol 9, 455–463 68. Morisaki, N., Kobayashi, H., Iwasaki, S., Furihata, K., Dabbs, E. R.,
52. Pedersen, L. C., Benning, M. M. & Holden, H. M. (1995). Yazawa, K. & Mikami, Y. (1995). Structure determination of ribo-
Structural investigation of the antibiotic and ATP-binding sites in sylated rifampicin and its derivative: new inactivated metabolites of
kanamycin nucleotidyltransferase. Biochemistry 34, 13305–13311 rifampicin by mycobacterial strains. J Antibiot (Tokyo) 48, 1299–1303
8 Biochemical Logic of Antibiotic Inactivation and Modification 95

69. Morisaki, N., Iwasaki, S., Yazawa, K., Mikami, Y. & Maeda, A. (1993). 72. Dabbs, E. R., Yazawa, K., Mikami, Y., Miyaji, M., Morisaki, N.,
Inactivated products of rifampicin by pathogenic Nocardia spp.: struc- Iwasaki, S. & Furihata, K. (1995). Ribosylation by mycobacterial
tures of glycosylated and phosphorylated metabolites of rifampicin strains as a new mechanism of rifampin inactivation. Antimicrob
and 3-formylrifamycin SV. J Antibiot (Tokyo) 46, 1605–1610 Agents Chemother 39, 1007–1009
70. Yazawa, K., Mikami, Y., Maeda, A., Morisaki, N. & Iwasaki, S. 73. Yazawa, K., Mikami, Y., Maeda, A., Akao, M., Morisaki, N. &
(1994). Phosphorylative inactivation of rifampicin by Nocardia Iwasaki, S. (1993). Inactivation of rifampin by Nocardia brasilien-
otitidiscaviarum. J Antimicrob Chemother 33, 1127–1135 sis. Antimicrob Agents Chemother 37, 1313–1317
71. Tanaka, Y., Yazawa, K., Dabbs, E. R., Nishikawa, K., Komaki, H., 74. Piepersberg, W. (1997). Molecular biology, biochemistry, and
Mikami, Y., Miyaji, M., Morisaki, N. & Iwasaki, S. (1996). fermentation of aminoglycoside antibiotics. In: Biotechnology
Different rifampicin inactivation mechanisms in Nocardia and of Industrial Antibiotics, 2nd edn. (Strohl, W., ed.), pp. 81–163.
related taxa. Microbiol Immunol 40, 1–4 Marcel Dekker, New York

You might also like