You are on page 1of 11

Article

pubs.acs.org/jmc

Siderophore Receptor-Mediated Uptake of Lactivicin Analogues in


Gram-Negative Bacteria
Jeremy Starr,#,† Matthew F. Brown,#,† Lisa Aschenbrenner,§ Nicole Caspers,¶ Ye Che,⧧
Brian S. Gerstenberger,† Michael Huband,§ John D. Knafels,¶ M. Megan Lemmon,§ Chao Li,†
Sandra P. McCurdy,§ Eric McElroy,† Mark R. Rauckhorst,† Andrew P. Tomaras,§ Jennifer A. Young,†
Richard P. Zaniewski,§ Veerabahu Shanmugasundaram,⧧ and Seungil Han*,#,¶

Medicinal Chemistry, ⧧Computational Chemistry, §Antibacterials Research Unit, and ¶Structural Biology, Pfizer Global Research and
Development, Eastern Point Road, Groton, Connecticut 06340, United States
*
S Supporting Information

ABSTRACT: Multidrug-resistant Gram-negative pathogens


are an emerging threat to human health, and addressing this
challenge will require development of new antibacterial agents.
This can be achieved through an improved molecular
understanding of drug−target interactions combined with
enhanced delivery of these agents to the site of action. Herein
we describe the first application of siderophore receptor-
mediated drug uptake of lactivicin analogues as a strategy that
enables the development of novel antibacterial agents against
clinically relevant Gram-negative bacteria. We report the first
crystal structures of several sideromimic conjugated compounds bound to penicillin binding proteins PBP3 and PBP1a from
Pseudomonas aeruginosa and characterize the reactivity of lactivicin and β-lactam core structures. Results from drug sensitivity
studies with β-lactamase enzymes are presented, as well as a structure-based hypothesis to reduce susceptibility to this enzyme
class. Finally, mechanistic studies demonstrating that sideromimic modification alters the drug uptake process are discussed.

■ INTRODUCTION
Since Fleming’s landmark publication of the discovery of
outer membrane of Gram-negative bacteria to gain access to its
target in the periplasmic space. Classically, a β-lactam with
penicillin in 1929,1 the β-lactam antibiotic class has produced Gram-negative activity exhibits specific structural and phys-
more drugs than any other chemotype and has had a profound icochemical properties carefully tuned to enable passage
impact on extending human life. Despite the development of through porin channels while not engaging efflux trans-
new anti-infective drugs, bacterial infections continue to porters.10 Inhibition of PBPs then leads to the introduction
represent a major global health threat, especially infections of flaws in the peptidoglycan layer, resulting in loss of structural
with serious Gram-negative pathogens such as Pseudomonas integrity, osmotic control, morphological changes, and eventual
aeruginosa, Acinetobacter baumannii, Klebsiella pneumoniae, and cell lysis. In response to drug pressure, bacteria utilize a variety
Escherichia coli.2 Particularly vulnerable are hospitalized patients of drug-neutralizing mechanisms, including the production of
with comorbidities or weakened immune systems who are at β-lactamase (BLA) enzymes. To combat this resistance
high risk of infection with multidrug-resistant variants. β- mechanism clinically, β-lactams are often administered with
Lactams continue to be utilized as the foundation of treatment BLA inhibitors. However, inhibitor-resistant BLA variants are
for multiple conditions arising from these infections, such as prevalent and contribute to significant treatment failure.
ventilator-associated pneumonia, intra-abdominal infections, Therefore, novel PBP inhibitors with reduced susceptibility to
and bacteremia,3−5 but their effectiveness continues to diminish BLAs are critical to the development of next-generation
due to ever increasing prevalence of drug-resistant strains.6,7 antibacterial therapies. It is noteworthy that while decades of
There is a critical need for new therapeutic options for these research across the pharmaceutical industry has generated
difficult to treat infections.8 dozens of marketed β-lactam drugs, no other PBP-inhibiting
Penicillin binding proteins (PBPs) are a large class of chemotype has thus far reached the market.
enzymes ubiquitously expressed across the spectrum of Bacteria have developed a highly effective mechanism to
prokaryotic pathogens and are the biochemical targets of β- acquire iron, a critical element, from their host organism. The
lactam drugs.9 Most PBPs are bound in the peptidoglycan layer process involves the release of siderophores with extremely
and, in the case of Gram-positive bacteria, are readily accessible
to drugs capable of transiting the relatively permeable cell wall. Received: February 10, 2014
By contrast, an inhibitor must pass through the formidable Published: April 2, 2014

© 2014 American Chemical Society 3845 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855
Journal of Medicinal Chemistry Article

Figure 1. Lactivicin (1) and comparator agents.

high affinity for iron, which can effectively compete with the could positively influence the discovery of novel, next-
host’s endogenous iron binding proteins. Following secretion generation Gram-negative antibacterial agents.
and chelation of iron, the siderophore complexes are actively
retrieved through binding to specialized bacterial receptors. β-
Lactams conjugated to structurally simple, small molecule
■ RESULTS AND DISCUSSION
Relative Reactivity of Lactivicin and Monobactam
siderophore mimics (hereafter referred to as “sideromimics”) Analogues. Critical to the mechanism of β-lactam drugs is the
such as catechols and hydroxypyridones have been described presence of an electrophilic carbonyl that covalently modifies a
which utilize this internalization process, leading to active highly conserved PBP active site serine hydroxyl group leading
cellular uptake. Despite extensive industrial and academic to the formation of an acyl enzyme complex that is significantly
research efforts utilizing this “Trojan horse” approach to lower in energy due to relief of ring strain. By comparison, a
enhance intracellular delivery of β-lactams,11−20 a similar monocyclic five-membered lactam or larger homologue would
approach has not been reported for non-β-lactam PBP not be expected to exhibit a strain-induced irreversibility due to
inhibitors. Here we describe the application of the sideromimic the exceptional stability of such lactams, while a highly strained
conjugation strategy to lactivicins, a non-β-lactam natural three-membered ring lactam would be too reactive in water to
product class of PBP inhibitors of microbial origin (Figure survive en route to its bacterial target. However, the primacy of
1).21−23 We report that lactivicin−sideromimic conjugates four-membered ring lactams as PBP inhibitors has been
provide an improved PBP profile relative to aztreonam, with challenged.24−26 For example, the lactivicin class is unusual
potent inhibition of PBP3 and 1a as well as enhanced potency because it is the only known naturally occurring PBP inhibitor
vs PBP1b. This expanded PBP inhibitory profile may eventually class that does not contain a β-lactam. Instead of a β-lactam
prove to be an advantage for the lactivicin drug class. We report ring, lactivicin (1, Figure 1) and its derivatives contain
the first X-ray crystal structures of sideromimic-conjugated cycloserine (a five-membered lactam) and γ-lactone (a five-
lactivicin analogues bound to Pseudomonas aeruginosa PBP3 membered cyclic ester) rings. Recently published X-ray
(Pae PBP3) and PBP1a (Pae PBP1a), as well as the related β- crystallographic data supports the covalent mechanism of
lactam sideromimic-conjugate 4 (BAL30072) (Figure 1)16,17 action of lactivicins wherein the carbonyl of the cycloserine ring
bound to Pae PBP3 and discuss novel insights gained from the reacts with the Gram-positive Streptococcus pneumoniae active
effectiveness of these compounds. We report the sensitivity of site serine hydroxyl to form a covalent bond analogous to that
formed by a typical β-lactam drug (Figure 2A).27
the lactivicin sideromimic conjugates to β-lactamase enzymes,
Nucleophilic substitution at the carbonyl group of an amide
and based on structure-derived knowledge, we propose that
usually occurs in a stepwise manner initiated by the formation
optimal attachment and positioning of the sideromimic moiety
of a tetrahedral intermediate as the rate-determining step
enhances intrinsic activity and could lead to reduced (Figure 2A). Here the relative reactivity of covalent inhibitors
susceptibility to drug-degrading BLA enzymes. In addition, to such as β-lactams and lactivicins was theoretically characterized
aid analogue design and improve the molecular understanding by molecular orbital calculations as the activation barrier (ΔG⧧)
of covalent bond formation, relative reactivities of lactivicin and of the transition state structure in an aqueous solution with a
monobactam analogues are characterized using a computational simple nucleophile (CH3O−) representing the active site serine.
method which defines the activation barrier of a transition-state While this approach does not factor in the multiple hydrogen-
structure derived from the reaction of the electrophilic carbonyl bonding interactions which stabilize the transition state in the
with a simple model nucleophile, CH3O−. We have also oxyanion hole of the protein, it can be useful in the design of
characterized the role of bacterial siderophore receptors in the covalent inhibitors. For example, previously reported semi-
uptake of the new phthalimide-conjugated lactivicin analogue empirical calculations (CNDO/2) by Boyd et al. described an
17 and have discovered that it utilizes a broader set of Ton-B- optimal reactivity range for β-lactams which provides effective
dependent siderophore receptors as compared to related β- covalent inhibition of PBPs combined with the hydrolytic
lactam sideromimic conjugates. Taken together, these results stability needed to provide antibacterial activity in an aqueous
3846 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855
Journal of Medicinal Chemistry Article

environment.28 Here, we compare ΔG⧧ values, derived from


high-level DFT calculations, for a number of reported lactivicin
analogues to evaluate the influence of ring strain and electron-
withdrawing substituents on reactivity and subsequent
antibacterial activity (Figure 2B).
While the expanded lactam ring size of lactivicins relative to
β-lactams such as aztreonam (Figure 1) may be expected to
lead to reduced reactivity, the calculated activation barriers for
each are quite similar (ΔG⧧ for 1 and monobactam 9 = 31.3
and 30.7 kcal/mol, Figure 2B). Our assumption is that the
reduced ring strain of lactivicins relative to monobactams is
overcome by the electronic activation provided by the
appended lactone moiety and the cycloserine ring oxygen,
thus providing similar ΔG⧧ values, sufficient PBP reactivity, and
the resulting antibacterial activity. In addition, lactivicins are
known to undergo a ring-chain tautomerization (i.e., 5 to 6)
which leads to equilibration of the γ-lactone chiral center
(Figure 2A).29 The ring-opened tautomer 6 may be expected to
be a highly reactive species, and calculations support this
statement as virtually no barrier exists, suggesting that a
spontaneous reaction would occur with the nucleophile
CH3O−. Nevertheless, the fact that lactivicin analogues exhibit
sufficient chemical stability to provide antibacterial activity in an
aqueous environment suggests this may not be the case. Harada
et al. provide an explanation suggesting that the imminium ion
6 may be stabilized by the tethered carboxylate(s).29 So while
the existence of tautomer 6 is necessary to enable the reported
lactone equilibration process, its impact on overall chemical
stability and reactivity with the PBP active site serine remains
unknown. The potential of tautomer 6 to react with PBP is
captured in Figure 2a; however, the calculations described
above are based solely on the transition of the ring-closed
Figure 2. (A) Proposed mechanisms of acylation of lactivicin
analogues. (B) The activation barrier (ΔG⧧) of forming the tautomer 5 to 7.
transition-state structure derived from the reaction of the electrophilic A number of research groups have reported efforts to modify
carbonyl with a simple nucleophile CH3O− (DFT calculations). the lactivicin core, most of which lead to a significant or
complete loss of antibacterial activity. These results can be
rationalized with calculated ΔG⧧ values. For example, the
expanded lactam ring analogue 10 was reported to be inactive

Table 1. MICs and Pae PBP IC50 Data for Lactivicin Analogues

MIC (μg/mL)a Pae PBP IC50 (μM)


a
compound Eco EC-28 Pae PA01 Pae 1091-05 Kpn KP-3700 Aba AB-3167 1a 1b 2 3 nb
aztreonam 0.125 4 4 >64 32 3.34 ± 0.57 2.87 ± 0.64 >300 0.008 ± 0.004 8
13 NTc 32 32 4 32 0.046 1.23 33.3 0.092 1
14 2 8 4 32 0.5 0.046 2.46 22.2 0.14 1
15R >64 >64 >64 >64 >64 0.11 ± 0.032 2.44 ± 0.03 47.1 ± 23.5 0.11 ± 0.03 2
15S >64 >64 >64 >64 >64 0.065 ± 0.033 1.74 ± 0.87 33.3 0.065 ± 0.033 2
16R 1 16 2 16 0.5 0.046 0.82 6.41 ± 5.23 0.065 ± 0.033 2
16S 1 2 64 8 0.5 0.046 0.82 3.7 0.092 2
17 0.25 0.5 0.5 0.5 0.06 0.030 0.27 ± 0.39 3.7 ± 1.8 0.046 2

a
Eco: E. coli, Pae: Pseudomonas, Kpn: Klebsiella pneumoniae, Aba: Acinetobacter baumannii. bn = number of replicates for PBP IC50 determination.
c
NT: not tested.

3847 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855


Journal of Medicinal Chemistry Article

Figure 3. Interactions of compound 14 and 4 in the active site of the Pae PBP3. (A) Active site of Pae PBP3 bound to compound 14 (navy). (B)
Active site of Pae PBP3 bound to 4 colored in green. (C) Molecular surface of the Pae PBP3 in the active site region in complex with compound 4
(green). The residues of hydrophobic aromatic wall (Tyr-Tyr-Phe) and Val333 are shown in orange. (D) Molecular surface of the Pae PBP3 active
site in complex with aztreonam in cyan (PDB code: 3PBS). (E) Molecular surface of the Pae PBP3 active site in complex with 3 in pink (PDB code:
3PBT).

vs a panel for Gram-positive and negative bacteria.30 Ring sideromimic-containing C4 side chain of compound 14 is
expansion would be expected to reduce reactivity, and this is identical to that found in the monocyclic β-lactam analogue 4,
reflected in the 2.7 kcal/mol increase in the ΔG⧧ value for 10 vs which is currently in clinical evaluation for the treatment of
1 (Figure 2B). In addition, a number of reports describe efforts Gram-negative infections. Compound 14 provided our first
to replace the lactone N-activating group. For example, 11 opportunity to examine the covalently bound conformation of a
contains the N-sulfate activating group common to the lactivicin−sideromimic analogue with its biological target and
monobactam drug class. The significantly lower ΔG⧧ of this enable comparison to the related monocarbam analogue 4.
analogue (24.6 kcal/mol) as compared to 1 (31.3 kcal/mol) or Complexes of Pae PBP3 with 14 and 4. The 2.3 Å
monobactam 9 (30.7 kcal/mol) is suggestive of high level resolution crystal structure of Pae PBP3 with 14 revealed that
reactivity and poor hydrolytic stability, which may explain the both the L-cycloserine and γ-lactone rings are open and the
reported lack of antibacterial activity.31 Similarly, the low antibiotic is covalently linked to the nucleophilic hydroxyl
calculated ΔG⧧ (19.3 kcal/mol) for compound 12 explains the group of Ser294 (Figure 3A). The carboxylic acid group of 14
reported lack of antibacterial activity and poor aqueous resulting from the opening of the lactone ring is stabilized by a
stability.32 While the utilization of calculated ΔG⧧ values has H-bonding interaction with conserved residues Ser349, Ser485,
great potential to enable the rational design of next-generation and Thr487. The interaction of the core of 14 with Pae PBP3 is
lactivicin (or β-lactam) core structures, our initial efforts similar to the published S. pneumoniae PBP1b−lactivicin
maintained the cycloserine-lactone core structure found in the complex which involves all three signature motifs (SXXK,
lactivicin natural product (vide infra). SXN, and KSGT).27 Importantly, in the structure of the
Initial Application of Sideromimic-Conjugated Lacti- complex of 14, the electron densities corresponding to the
vicins: Oxime-Linked Dihydroxypyridone 14. Lactivicin ethylene carboxylate moiety resulting from the opening of the
lead optimization efforts reported by Takeda describe a number lactone ring and that of the dihydroxypyridone sideromimic
of potent analogues, including 13 (Table 1), possessing a fairly moiety are weak and discontinuous, suggesting that these side
balanced spectrum of activity toward Gram-negative pathogens chains are not only flexible but also do not form strong
and in vivo efficacy.33−35 However, when evaluated against a interactions with Pae PBP3 (Figure S1, Supporting Informa-
contemporary panel of clinically relevant Gram-negative tion). These flexible side-chain conformations suggested that
bacteria, weak whole cell activity (MIC) was observed for 13 the potency of 14 could be further improved by structural
vs P. aeruginosa and A. baumannii strains (Table 1). As modifications to enable interaction with the unique Tyr-Tyr-
discussed earlier, conjugation of sideromimics such as catechols Phe aromatic wall present in Pae PBP3 (vide infra).
and hydroxypyridones to β-lactams has been shown to impart Given the structural similarities between 14 and 4, a 2.0 Å
improved cellular potency via utilization of the bacterial iron resolution crystal structure of the Pae PBP3 in complex with 4
uptake process. Recent examples include 3 (MC-1) and 4, both was prepared to enable direct comparison of the two series
of which employ hydroxypyridones as the sideromimics (Figure (Figure 3B). While not well-resolved, the 4-dihydroxypyridone
1).16,18 Incorporation of a hydroxypyridone moiety in lactivicin, moiety is positioned near a hydrophobic pocket formed by
as in compound 14, quickly demonstrated the potential of the Tyr503, Tyr532, and Phe533 and the N-hydroxy group appears
siderophore receptor uptake strategy, providing improved to be within hydrogen bonding distance of the phenolic
MICs in P. aeruginosa and A. baumannii strains while retaining oxygens in both Tyr532 and Tyr503. The low resolution in this
PBP inhibitory potency similar to comparator analogue 13 region of the structure limits the ability to confidently define
which lacks a sideromimic (Table 1). Furthermore, the meaningful drug−protein interactions. This is different from
3848 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855
Journal of Medicinal Chemistry Article

Figure 4. Crystal structure of Pae PBP1a in complex with compound 17 and comparison of the structures with Pae PBP3 and Aba PBP3. (A)
Overall structure of Pae PBP1a complexed with compound 17. The bound compound 17 is shown as spheres. (B) Active site of Pae PBP1a. The
catalytic Ser461 is shown in red. The loop connecting β3 and β4 is shown in cyan. (C) Overlap of proteins from 3−Pae PBP3 (orange) and 17−
PBP1a (cyan) structures, illustrating that PBP3 Phe533 occupies the same general space as PBP1a Tyr733. (D) Comparison of the extended loop
connecting β3 and β4 in apo-Pae PBP3 (green) and apo-Aba PBP3 (pink).

the published complexes of aztreonam and monocarbam 3 with no meaningful whole cell activity against any of the Gram-
Pae PBP3 in which the C4-linked gem-dimethyl group common negative pathogens of interest despite both having PBP
to these compounds clearly interacts favorably with the inhibitory potency comparable to the Takeda prototype
hydrophobic pocket composed of Tyr503-Tyr532-Phe533 and compound 13 as well as the hydroxypyridone-conjugate 14
the oxime-linked carboxylate forms a salt bridge with Arg489 (Table 1). The increased size and lipophilicity of the Cbz group
(Figure 3C−E).36 Based on these favorable interactions, a was likely deleterious to passive transit through porins and
design strategy was developed to incorporate this isobutyric suggested that structurally similar compounds would gain little
acid side chain into the lactivicin analogues while exploring an exposure to the periplasm without facilitated transport.
alternate attachment point of the sideromimic group. Furthermore, on the basis of the similar PBP activity, there
Linkage of the Sideromimic to the α-Lactone was little apparent preference for one configuration of the
Position. Of particular interest were α-amido-γ-lactones such amido substituent over the other.
as 15−17 (Table 1). A sideromimic attachment at the lactone Replacement of the Cbz groups in 15R and 15S with a
α-position would potentially position this substituent in the hydroxypyridone sideromimic provided the diastereomeric
semiopen channel distal from the active site serine where compounds 16R and 16S. With siderophore receptor facilitated
additional ligand−protein interactions might be possible. This transport now enabled, both analogues exhibited improved
channel has been exploited previously in the design of MICs relative to 15R and 15S, despite only limited differences
monocarbams such as 336 (Figure 3E) and related mono- in PBP inhibition relative to the comparators lacking a
bactams.19,20 A priori, predicting the optimal stereochemistry of sideromimic, providing additional evidence that sideromimics
the lactone α-amido connection was challenging. The available play a critical role in cell entry in this series. Again, despite a
data from the crystal structure of P. aeruginosa PBP3 covalently difference in stereochemistry of sideromimic attachment, the
bound to 14 showed only the ring-opened adduct, as expected, PBP inhibitory activity exhibited little differentiation between
and could not inform about a preferred binding orientation of the diastereomers. However, fairly significant differences in P.
the ring-closed species. This analysis was further complicated aeruginosa cellular potency were observed for the 16R/S isomer
by the dynamic stereochemistry of the tertiary γ-carbon of the pair. The reason for this is currently unknown.
lactone resulting from the ring-opening equilibration process In an effort to further optimize drug PBP3 enzyme
described in Figure 2A. Therefore, exemplars of both interactions, a design strategy to exploit potential aryl−aryl
configurations of the α-lactone substituent were prepared and interactions between the sideromimic and the Tyr503, Tyr532,
evaluated. The resulting compounds (15R and 15S) bearing a and Phe533 aromatic residues led to analogues bearing phenyl
benzyloxycarbonyl (Cbz) group (not a sideromimic) exhibited rings. From this effort, compound 17 was found to exhibit both
3849 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855
Journal of Medicinal Chemistry Article

the lowest MICs and PBP IC50s to date for the series, perhaps of 17 bound to Pae PBP3 was not prepared, comparison of the
due to an improved interaction between the phthalimide previously reported 3−Pae PBP3 structure36 with the 17−Pae
phenyl ring and the PBP active sites and to the competence of PBP1a structure demonstrates a similar positioning of the PBP3
the phthalimide as a substrate for active transport via Phe533 and PBP1a Tyr733 side chains, suggesting that
siderophore receptors (vide infra). Interestingly, attachment analogous π-stacking interactions with the phthalimido group
of a 4,5-dihydroxyphthalimide to a cephalosporin core has been of 17 may occur in both PBP1a and PBP3 (Figure 4C).
reported to provide potent antibacterial analogues as well, Evidence Supporting Siderophore Receptor-Medi-
suggesting that this sideromimic may have general utility for ated Drug Uptake. P. aeruginosa strains are able to express
drug transport.37 an impressive range of Ton-B-dependent siderophore recep-
Monobactam antiobiotics such as aztreonam typically tors.39 This high level of functional redundancy is critical to
demonstrate potent inhibition of PBP3, with lesser activity vs ensure that adequate levels of iron can be obtained in a variety
PBP1a and 1b and no relevant activity vs PBP2 (Table 1). of environments. Previous work from our group has suggested a
While selective PBP3 inhibitors like aztreonam can successfully role for the siderophore receptors PiuA and PirA for the uptake
treat susceptible infections, enhanced potency vs other PBPs of a variety of hydroxypyridone-conjugated β-lactams.20,40
could lead to improved antibacterial performance. On the basis Because analogues 14 and 16R/S incorporate the same
of initial data, it would appear that the lactivicin drug class hydroxypyridone sideromimics, our assumption is that they
provides an improved PBP profile relative to aztreonam, utilize a similar uptake process; however, this remains to be
providing potent inhibition of PBP3 and 1a as well as enhanced determined. Because lactivicin analogue 17 incorporates a
potency vs PBP1b and 2 (Table 1). While the impact of this structurally unique phthalimide sideromimic, a study was
expanded PBP inhibitory profile remains to be determined, it conducted to determine which receptors are involved in
may eventually prove to be an advantage for the lactivicin drug cellular drug uptake. Whole cell activity was recorded for 17 vs
class. the wild-type P. aeruginosa strain PAO1 as well as an isogenic
Class A Pae PBP1a Structure and Its Complex with a panel of strains lacking one or more siderophore receptors
Phthalimide Sideromimic, Compound 17. The potent (Table 2). The study was conducted with both standard
PBP1a inhibitory activity of phthalimide 17 catalyzed crystallo-
graphic studies to investigate its binding pose with this class A Table 2. Pae PAO1 Isogenic Siderophore Receptor Mutant
PBP. Following an extensive effort with multiple constructs, Panel
Pae PBP1a crystals were obtained after controlled proteolysis
MICs (μg/mL)a
with trypsin (see Experimental Section for details). The
crystallized Pae PBP1a is composed of four major polypeptides, 17
including: Pro47-Met64, Thr256-Leu494, Thr506-Val609, MHB/low- aztreonam low-
Glu652-Glu792. In Pae PBP1a, the initial transmembrane strains MHB low-Fe Fe Fe
helix is followed by an N-terminal transglycosylase (TG) PAO1 0.5(1) ≤0.03(1) 16 4(1)
domain connected through a β-rich linker to a C-terminal piuA 8(16) 0.125(≥4) 64 4(1)
transpeptidase (TP) domain. The entire TG domain is absent pirA 0.5(1) ≤0.03(1) 16 4(1)
in the Pae PBP1a structure due to intrinsic flexibility and tryptic f pvA 16(32) ≤0.03(1) 512 2(0.5)
cleavage. The interdomain linker contains a six-stranded β- f ptA 4(8) 0.06(≥2) 64 4(1)
sheet where four strands are from TG and TP domains. The pfeA 0.5(1) ≤0.03(1) 16 4(1)
TP domain in the center of Pae PBP1a (residues 292−762) piuA fpvA 16(32) 0.25(≥8) 64 4(1)
shares a similar overall fold with other transpeptidases and piuA fptA 8(16) 0.5(≥16) 16 4(1)
serine β-lactamases. In addition to TP and TG domains that are piuA pirA 16(32) 0.25(≥8) 64 4(1)
commonly found in bifunctional PBPs, Pae PBP1a contains an piuA pfeA 8(16) 0.125(≥4) 64 4(1)
OB (oligonucleotide/oligosaccharide binding)-fold domain pirA pfeA 0.5(1) ≤0.03(1) 16 4(1)
inserted in the TP domain between the first α-helix and the piuA pfeA pirA 16(32) 0.25(≥8) 64 4(1)
first β-strand, also seen in the A. baumannii PBP1a (Aba a
Data in parentheses = strain MIC/PAO1 MIC.
PBP1a) structure (Figure 4A).38 The overall Pae PBP1a
structure can be superimposed onto the Aba PBP1a structure Mueller Hinton Broth (MHB) as well as a modified broth
with an rmsd of 1.5 Å for 384 Cα atoms. depleted of iron to mimic the low iron in vivo environment to
In complex with 17, the electron density of the P. aeruginosa determine if the drug uptake mechanism varies based on
PBP1a active site unambiguously revealed a covalent acyl- availability of iron. A significant shift (>4×) in MIC for a
enzyme interaction (Figure S1, Supporting Information). The receptor deficient strain vs the parent strain PAO1 is indicative
phthalimide ring and the isobutyrate groups of 17 are well of the involvement of the deleted receptor(s) in drug uptake.
defined, and the electron density is continuous between the For example, under standard conditions (MHB), PiuA, FpvA,
active site Ser461 hydroxyl group and the connecting lactivicin and FptA single mutants all provided an 8-fold or greater shift
acyl carbon, with the acyl carbonyl oxygen lying in the oxyanion in MIC relative to PAO1 when treated with 17, as did a number
hole defined by the Ser461 and Thr698 main chain amides. The of double mutant strains. However, under low iron conditions,
carboxylate at the γ-lactone position in 17 is anchored by H- only the PiuA single mutant showed a modest MIC shift (4-
bonds with the side chains of Thr696 and Thr698, and a water- fold), with no relevant shift observed for the FpvA or FptA
mediated H-bond with the Gly735 backbone. Though the single mutants, suggesting an adaption in receptor expression
ethylene carboxylate resulting from lactone ring opening does and/or function has occurred with differing iron levels, which is
not show a clear interaction with PBP1a, the group does a known phenomenon.41 More robust MIC shifts, especially in
provide rigidity to the phthalimide ring, facilitating a π-stacking the low iron media, were observed for a number of the double
interaction with Tyr733 (Figure 4B). While a crystal structure mutant strains, suggesting that multiple receptors including
3850 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855
Journal of Medicinal Chemistry Article

Table 3. Isogenic Library of E. coli Expressing β-Lactamases


MIC (μg/mL)
BLA-expressed ampicillin cefepime 3 4 13 14 17
DH5α (parent strain) 2 0.03 0.125 0.125 0.25 1 0.03
pUCP26 (empty vector) 2 0.03 0.125 0.125 0.25 0.5 0.03
SHV-5 >64 4 0.5 >64 0.5 32 0.03
SHV-12 >64 0.5 0.25 >64 0.25 16 0.015
TEM-1 >64 0.125 0.03 0.03 0.25 1 0.06
TEM-24 >64 0.5 0.125 >64 0.5 4 0.06
CTX-M-15 >64 8 0.25 1 0.5 4 0.015
VIM-2 >64 0.06 0.015 0.03 0.25 0.5 0.015
NDM-1 >64 8 0.03 0.06 0.5 1 0.03
OXA-24/40 32 0.06 0.25 0.125 1 1 0.03
OXA-58 >64 0.03 0.125 0.125 0.25 1 NTa
KPC-3 64 0.125 0.06 0.125 0.5 1 0.03
a
NT: not tested.

PiuA, PirA, FpvA, FptA, and perhaps others are involved in the any β-lactam or related drug. Table 3 reports the data generated
uptake of 17. The PiuA and PirA receptors are known to for 3, 4, cefepime, ampicillin, and lactivicin analogues 13, 14,
recognize catechol-containing siderophores, such as enter- and 17. Results are reported as MICs, and susceptibility to a
obactin, while FpvA and FptA are involved in the uptake of the particular BLA can be inferred from a >4-fold MIC shift of the
major P. aeruginosa siderophores pyoverdine and pyochelin, corresponding mutant E. coli strain relative to the empty vector
respectively. In comparison to the previously reported results control strain. As expected, the positive control drug, ampicillin,
for hydroxypyridone-containing β-lactams,20,40 compound 17 was hydrolyzed and thus inactivated by the full range of BLAs,
appears to utilize a broader set of siderophore receptors, which and the more robust cephalosporin control, cefepime, displayed
could translate to a therapeutic advantage in vivo where susceptibility to SHV-5, CTX-M-15, and NDM-1 while
siderophore receptor expression and function is expected to resisting hydrolysis by the other BLAs in the panel. As
vary depending on the availability of iron. In addition, 17 previously reported, the monocarbam 3 exhibited significant
appears to provide a relative potency advantage in the low iron stability toward inactivation by a broad range of BLAs with only
media, and the magnitude of the activity difference observed in borderline susceptibility to SHV-5.40 In general, the lactivicin
normal vs low-iron conditions for 17 (16−512-fold) is analogues also provide broad stability to BLAs similarly to 3,
significantly greater than that reported for the pyridone- including the strain expressing the New Delhi metallo-β-
conjugated β-lactams. (0.5−32-fold).20,40 lactamase (NDM-1). The original Takeda compound 13
While these preliminary in vitro results with phthalimide- lacking a sideromimic moiety exhibited good stability across
conjugated lactivicin analogue 17 may appear promising, a the entire panel, with only the MIC vs the OXA-24/40
recent report demonstrates that the gold standard in vitro expressing strain being 4-fold higher than the empty vector
potency (i.e., MIC) and frequency of resistance (FOR) control strain. The phthalimide-conjugated lactivicin 17 also
methodologies utilized ubiquitously in antibacterial drug performed well, consistently inhibiting the BLA mutant strains
research fail to predict in vivo efficacy for the related at MICs within 2-fold of the empty vector control including the
monobactam−sideromimic conjugate, 2 (MB-1) (Figure 1).42 OXA-24/40 BLA-expressed E.coli strain. In contrast, both 4 and
The lack of efficacy appears to involve an adaptive response of the related lactivicin derivative 14 displayed liabilities, notably
bacteria when exposed to 2, leading to modified expression and with respect to SHV-5 and SHV-12, and to a lesser extent,
utilization of siderophores and siderophore receptors, thus CTX-M-15 and TEM-24. Given the significantly different core
reducing the effectiveness of 2 in murine P. aeruginosa infection structure of 14 compared to 4, the similar BLA liabilities may
model studies. The authors describe modified in vitro reside with the common placement of the dihydroxypyridone
methodologies which correlate nicely with 2 in vivo outcomes, moiety shared by these compounds. When considered
and it remains to be determined if the modified siderophore alongside the original Takeda lead 13, the dramatic increase
receptor repertoire of phthalimide 17 conveys any advantage in susceptibility of 14 to SHV-5 and SHV-12 further suggests
over 2 in these assays. that this placement of the sideromimic enhances susceptibility
Profiling β-Lactamase Susceptibility. Given that lactivi- to some BLAs and may potentially limit utility against
cins inactivate PBPs in a fashion analogous to β-lactams, it is Enterobacteriaceae carrying these common extended spectrum
not surprising that they have been reported to be ineffective in β-lactamases (ESBLs). By installing the known isobutryic acid
killing bacteria harboring certain BLA enzymes.23,43 Therefore, C4 group found in 3, 13, and aztreonam, and by modifying the
to assess the relative BLA susceptibility for new lactivicin sideromimic point of attachment, compound 17 demonstrates a
analogues, an isogenic panel of E. coli strains was assembled significant advantage over 4 and 14 with regard to BLA
wherein each strain was engineered to express a single, specific, susceptibility in this panel. In addition, the 64-fold improve-
commonly occurring BLA.40 While this panel was not ment in potency for 17 vs 14 when tested against the K.
developed to match wild-type expression levels of BLAs or pneumoniae strain KP-3700 may provide additional evidence for
assess the impact of BLAs working in concert, evaluation of the BLA susceptibility discussion above as this strain is known
whole cell activity (i.e., MICs) against this panel as compared to to express SHV-5 (Table 1). Bonomo et al. recently published a
the parent E. coli strain harboring the empty vector (pUCP26) crystal structure of a preacylation complex of the BLA inhibitor
provides a snapshot of the relative risk of BLA susceptibility of sulbactam bound to a S70C variant of the SHV-1 enzyme.44
3851 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855
Journal of Medicinal Chemistry Article

Modeling of the ring-closed analogues 14 and 17 in this crystal bound to Pae PBP3 and Pae PBP1a. In the course of this work,
structure lends further support to the hypothesis that we identified compound 17, a novel phthalimide-conjugated
sideromimic placement could impact BLA susceptibility, as lactivicin analogue. Efforts to understand the Ton-B-dependent
14 can be accommodated in the active site with the siderophore receptors involved in the uptake of compound 17
sideromimic moiety pointing toward solvent, while preserving suggest that it may utilize a broader set of receptors than related
key interactions such as proximity to nucleophile and hydroxypyridone-conjugated β-lactams, potentially providing a
positioning in the oxyanion hole. However, the phthalimide unique advantage in uptake and Gram-negative effectiveness.
group of compound 17 appears to clash with the protein due to Further profiling of 17 for susceptibility against BLAs indicates
the limited space available in this region, while maintaining the potential for generally low susceptibility, and evaluation vs
other key interactions and thereby leading to poor binding PBPs suggests the potential for a broader inhibitory profile
(Figure 5). While a more definitive BLA susceptibility relative to monobactams. Overall, the results demonstrate that
the lactivicin−sideromimic conjugate class merits additional
research in the quest to find new effective antibacterial agents.

■ EXPERIMENTAL SECTION
Cloning, Expression, and Purification of Pae PBP3 and
PBP1a. Pae PBP3 was prepared as discussed previously.36 PBP1a
(residues 36−822) from PAO1 genomic DNA was cloned into the
pET28a vector (Novagen) with an N-terminal His6 fusion tag. Protein
was expressed in E. coli BL21 (Gold) cells grown in autoinduction
media (Novagen) overnight at 25 °C. Cell pellets were lysed in 5
volumes (w/v) of B-PER protein extraction reagent (Thermo
Scientific Pierce) supplemented with 0.4 M NaCl, 10 mM MgCl2,
Complete EDTA-free protease inhibitor tablets (Roche), and
Benzonase (Novagen). The lysate was clarified by centrifuging at 4
°C for 1 h in a Sorvall SS-34 rotor at 30 000g and loaded onto two
tandem 5 mL HisTrap FF crude columns (GE Healthcare)
equilibrated in buffer A (25 mM TrisHCl, 0.4 M NaCl, pH 8.0)
with 20 mM imidazole pH 8.0 (Emerald Bio) added. After the
columns were washed with 20 column volumes of buffer, PBP1a was
eluted with a linear gradient to 250 mM imidazole over 20 column
volumes. The eluted peak was concentrated and loaded onto a HiLoad
Superdex 200 16/60 gel filtration column (GE Healthcare)
equilibrated in buffer A. The peak corresponding to monomeric
PBP1a (data not shown) was pooled and concentrated to 20 mg/mL
using an Amicon Ultra 30 kDa MWCO centrifugal concentrator.
Aliquots were flash-frozen in liquid nitrogen and stored at −80 °C.
PBP1a was digested in a series of 84 μL reactions at a concentration of
2.67 mg/mL in 100 mM Tris-HCl, 100 mM NaCl, 10 mM CaCl2, and
1.1 μg of TPCK trypsin (Thermo Scientific Pierce). After 2 h at 4 °C,
the digestions were stopped by adding PMSF to 10 mM. Ten
digestion reactions were pooled and loaded onto a Superdex 200 HR
10/30 column equilibrated in buffer A. Fractions with digested PBP1a
were pooled, concentrated to 14 mg/mL, flash-frozen in liquid
Figure 5. Proposed models of compound 14 and 17 in the active site nitrogen, and stored at −80 °C.
of crystal structure of S70C SHV-1 β-lactamase. The S70C mutation Crystallization and Data Collection of Pae PBP3 and PBP1a.
was designed to affect the reactivity of the catalytic residue to allow for Crystals of apo-PBP3 were obtained with a reservoir solution
capture of the preacylation complex. (A) Model of S70C SHV-1 β- containing 30% PEG 4000, 0.2 M MgCl2, and 0.1 M Tris pH 8.5.
lactamase with compound 14. The volume occupied by the ring-closed The compound was soaked into the crystals. Soaks were performed
form of the analogue is shown as a blue wire surface and the active site with 1 mM compound in reservoir solution for 1 day. Crystals were
of SHV-1 as an indigo solid surface. (B) Model of compound 17. The cryoprotected by dragging the crystals through MiTeGen’s LV
volume occupied by the ring-closed form of the analogue is shown as a CryoOil (MiTeGen, LLC) and flash frozen in liquid nitrogen.
pink wire surface and the active site of SHV-1 as an indigo solid PBP1a was crystallized by sitting drop vapor diffusion at room
surface. The arrow points to a region in the active site where there is temperature using equal volumes of 10 mg/mL protein and well
significant steric clash with compound 17. solution. Crystallization conditions were screened using Crystal Screen
HT (Hampton Research). Crystals grew from condition 11 (0.1 M
sodium citrate tribasic dihydrate pH 5.6, 1.0 M ammonium phosphate
assessment would require evaluation against an expanded monobasic) and reached their full size (∼100 μm) in 2−3 weeks.
panel of clinical isolates known to express specific BLAs, these Crystals were soaked for 4 h in a solution containing 0.1 M sodium
initial results may provide useful guidance for future drug citrate tribasic dihydrate pH 5.6, 1.2 M ammonium phosphate
optimization efforts. monobasic, and 2 mM compound 17. The soaked crystals were


cryoprotected in 0.07 M sodium citrate tribasic dihydrate pH 5.6, 0.7
CONCLUSION M ammonium phosphate monobasic, and 35% glycerol and flash-
frozen in liquid nitrogen.
We have successfully applied the sideromimic-conjugation Structure Determination. Data were processed using the
strategy to lactivicins, providing new analogues with enhanced HKL2000 software suite.45 The structures of PBP3− and PBP1a−
in vitro Gram-negative activity. Analogue design was signifi- inhibitor complexes were solved by molecular replacement methods
cantly aided by learnings derived from structures of inhibitors with the CCP4 version of PHASER46 using Acinetobacter baumannii

3852 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855


Journal of Medicinal Chemistry


Article

PBP1a (PDB code: 3UDF) and Pae PBP3 (PDB code: 3PBN) as a ASSOCIATED CONTENT
search model. After molecular replacement, maximum likelihood-based
refinement of the atomic position and temperature factors were *
S Supporting Information

performed with autoBUSTER47 and the atomic model was built with The experimental details regarding the synthesis of PBP
the program COOT.48 The stereochemical quality of the final model inhibitors and reactivity modeling. This material is available
was assessed with PROCHECK.49 Crystallographic statistics for the free of charge via the Internet at http://pubs.acs.org.
final models are shown in Table S1, Supporting Information. Figures Accession Codes
were prepared with PYMOL (www.pymol.org).
Reactivity Modeling. Density functional theory (DFT) calcu- The atomic coordinates and structure factors have been
lations, full geometry optimizations, and frequency analyses were deposited in the Protein Data Bank with entry codes 4OOL,
carried out with B3LYP hybrid functional and triple-ζ 6-311+G(d,p) 4OOM, and 4OON.
basis set using the Gaussian 09 package.50 A scaling factor of 0.9877
was used to correct zero-point vibrational energies.51 Implicit solvation
using SMD polarizable continuum model of Cramer and Thrular was
utilized.52 Transition state optimizations were conducted for additions
■ AUTHOR INFORMATION
Corresponding Author
of a simple model nucleophile (CH3O−), as a computational surrogate *Phone: (860)-686-1788. E-mail: seungil.han@pfizer.com.
for the catalytic serine residue, to the electrophilic carbonyl. The Author Contributions
nature of stationary points was checked by means of frequency #
J.S, M.F.B., and S.H. contributed equally.
calculations, and transition states were further verified by IRC
calculations.53 The activation barrier (ΔG⧧) determined at room Notes
temperature (298.15 K) was used for ranking relative reactivity. The authors declare no competing financial interest.


MICs, PBP IC50s, PAO1 Isogenic Siderophore Receptor
Mutant Panel, and E. coli Isogenic BLA Panel. The minimum ACKNOWLEDGMENTS
inhibitory concentration (MIC) values were determined using the
broth microdilution protocol according to the methods of the Clinical We thank Mark Plummer, Mark Mitton-Fry, John Mueller, Paul
and Laboratory Standards Institute (CLSI).54 The experimental Miller, and Mark Noe for helpful discussions. Our appreciation
methods utilized to generate the P. aeruginosa siderophore receptor is also extended to Jinshan (Mike) Chen for managing external
KO data as well as the E. coli isogenic BLA panel have been described chemistry resources. Finally, we thank Juergen Bulitta for
in detail,40 as has the method utilized to generate PBP IC50 values.20 providing helpful comments regarding this manuscript.


Chemistry. Detailed synthetic procedures and spectral character-
ization for all compounds are provided in the Supporting Information.
The methods of purification and purity determination of compounds ABBREVIATIONS USED
13, 14, 15R/S, 16R/S, and 17 are as follows: MIC, minimum inhibitory concentration; PBP, penicillin
Method A. Phenomenex Max-RP, C18, 150 × 21.2 mm, 5 μm; binding protein; MHB, Mueller−Hinton broth; Pae, Pseudo-
mobile phase A: 0.1% formic acid in water (v/v); mobile phase B: monas aeruginosa; Kpn, Klebsiella pneumoniae; Eco, Escherichia
0.1% formic acid in methanol (v/v); gradient: 5% B for 1.50 min, 5% coli; Aba, Acinetobacter baumannii; BLA, β-lactamase; FOR,
to 100% B over 8.5 min 100% B for 1 min; flow rate: 28 mL/minute. frequency of resistance; TG, transglycosylase; TP, trans-
Temperature: not controlled; detection: DAD 210−360 nm; MS (+)
range 150−700 m/z; injection volume: 10 μL.
peptidase
Method B. Phenomenex Gemini-NX, 4.6 mm × 50 mm, C18, 3 μm,
110A; mobile phase A: 0.1% formic acid in water (v/v); mobile phase
B: 0.1% formic acid in acetonitrile (v/v); gradient: 0% to 100% B over
■ REFERENCES
(1) Fleming, A. The antibacterial action of cultures of a Penicillium,
4.1 min, 100% B 0.4 min; flow rate: 1.5 mL/min. Temperature: 60 °C.; with special reference to their use in the isolation of B. influenzae. Br. J.
detection: 200−450 nm; MS (+) range 100−1200 m/z; injection Exp. Pathol. 1929, 10, 226−236.
volume: 5 μL; instrument: column oven from Agilent Technologies, (2) Boucher, H. W.; Talbot, G. H.; Bradley, J. S.; Edwards, J. E.;
Wilmington, DE; autosampler and MS detector from Waters Gilbert, D.; Rice, L. B.; Scheld, M.; Spellberg, B.; Bartlett, J. Bad bugs,
Corporation, Milford, MA; ELS detector from Varian medical devices, no drugs: no ESKAPE! An update from the Infectious Diseases Society
Palo Alto, CA. of America. Clin. Infect. Dis. 2009, 48, 1−12.
Method C. Waters Acqity HSS T3, C18, 2.1 × 50 mm, 1.7 μm; (3) Solomkin, J. S.; Mazuski, J. E.; Bradley, J. S.; Rodvold, K. A.;
mobile phase A: 0.1% formic acid in water (v/v); mobile phase B: Goldstein, E. J. C.; Baron, E. J.; O’Neill, P. J.; Chow, A. W.; Dellinger,
0.1% formic acid in acetonitrile (v/v); gradient: 5% B over 0.1 min, 5% E. P.; Eachempati, S. R.; Gorbach, S.; Hilfiker, M.; May, A. K.;
to 95% B over 2.5 min, 95% B 0.35 min; flow rate: 1.25 mL/min. Nathens, A. B.; Sawyer, R. G.; Bartlett, J. G. Diagnosis and
Temperature: 60 °C.; detection: 200−450 nm; MS (+) range 100− management of complicated intra-abdominal infection in adults and
2000 m/z; injection volume: 5 μL; instrument: Waters Acquity UPLC. children: guidelines by the Surgical Infection Society and the
Compound 13. Purification method A, purity determination Infectious Diseases Society of America. Clin. Infect. Dis. 2009, 50,
method B. The compound was found to have a purity > 95%. 133−164.
Compound 14. Crude product was triturated with diethyl ether (4) Anonymous. Guidelines for the management of adults with
followed by 4:1 dichloromethane: diethyl ether. Purity determination hospital-acquired, ventilator-associated, and healthcare-associated
method B. The compound was found to have a purity > 95%. pneumonia. Am. J. Respir. Crit. Care Med. 2005, 171, 388-416.
Compound 15R. Purification method A, purity determination (5) Dellinger, R. P.; Levy, M. M.; Rhodes, A.; Annane, D.; Gerlach,
method B. The compound was found to have a purity = 87%. H.; Opal, S. M.; Sevransky, J. E.; Sprung, C. L.; Douglas, I. S.; Jaeschke,
Compound 15S. Purification method A, purity determination R.; Osborn, T. M.; Nunnally, M. E.; Townsend, S. R.; Reinhart, K.;
method B. The compound was found to have a purity = 94%. Kleinpell, R. M.; Angus, D. C.; Deutschman, C. S.; Machado, F. R.;
Compound 16R. Purification method A, purity determination Rubenfeld, G. D.; Webb, S. A.; Beale, R. J.; Vincent, J. L.; Moreno, R.
method B. The compound was found to have a purity > 95%. Surviving sepsis campaign: International guidelines for management of
Compound 16S. Purification method A, purity determination severe sepsis and septic shock: 2012. Crit. Care Med. 2013, 41, 580−
method B. The compound was found to have a purity = 84%. 637.
Compound 17. Purification method A, purity determination (6) Livermore, D. M. Current epidemiology and growing resistance
method C. The compound was found to have a purity > 95%. of Gram-negative pathogens. Korean J. Intern. Med. 2012, 27, 128−142.

3853 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855


Journal of Medicinal Chemistry Article

(7) Patel, G.; Bonomo, R. A. Status report on carbapenemases: (22) Nozaki, Y.; Katayama, N.; Ono, H.; Tsubotani, S.; Harada, S.;
challenges and prospects. Expert Rev. Anti-Infect. Ther. 2011, 9, 555− Okazaki, H.; Nakao, Y. Binding of a non-beta-lactam antibiotic to
570. penicillin-binding proteins. Nature 1987, 325, 179−180.
(8) Obrecht, D.; Bernardini, F.; Dale, G.; Dembowsky, K. Emerging (23) Nozaki, Y.; Katayama, N.; Harada, S.; Ono, H.; Okazaki, H.
new therapeutics against key Gram-negative pathogens. Annu. Rep. Lactivicin, a naturally occurring non-β-lactam antibiotic having β-
Med. Chem. 2011, 46, 245−262. lactam-like action: biological activities and mode of action. J. Antibiot.
(9) Waxman, D. J.; Strominger, J. L. Penicillin-binding proteins and 1989, 42, 84−93.
the mechanism of action of β-lactam antibiotics. Annu. Rev. Biochem. (24) Aszodi, J.; Rowlands, D. A.; Mauvais, P.; Collette, P.; Bonnefoy,
1983, 52, 825−869. A.; Lampilas, M. Design and synthesis of bridged γ-lactams as
(10) O’Shea, R.; Moser, H. E. Physicochemical properties of analogues of β-lactam antibiotics. Bioorg. Med. Chem. Lett. 2004, 14,
antibacterial compounds: implications for drug discovery. J. Med. 2489−2492.
Chem. 2008, 51, 2871−2878. (25) Coleman, K. Diazabicyclooctanes (DBOs): A potent new class
(11) Wencewicz, T. A.; Moellmann, U.; Long, T. E.; Miller, M. J. Is of non-β-lactam β-lactamase inhibitors. Curr. Opin. Microbiol. 2011, 14,
drug release necessary for antimicrobial activity of siderophore-drug 550−555.
conjugates? Syntheses and biological studies of the naturally occurring (26) Zervosen, A.; Sauvage, E.; Frere, J.-M.; Charlier, P.; Luxen, A.
salmycin “Trojan Horse” antibiotics and synthetic desferridanoxamine- Development of new drugs for an old target - the penicillin binding
proteins. Molecules 2012, 17, 12478−12505.
antibiotic conjugates. BioMetals 2009, 22, 633−648.
(27) Macheboeuf, P.; Fischer, D. S.; Brown, T., Jr.; Zervosen, A.;
(12) Ji, C.; Juarez-Hernandez, R. E.; Miller, M. J. Exploiting bacterial
Luxen, A.; Joris, B.; Dessen, A.; Schofield, C. J. Structural and
iron acquisition: siderophore conjugates. Future Med. Chem. 2012, 4,
mechanistic basis of penicillin-binding protein inhibition by lactivicins.
297−313.
Nat. Chem. Biol. 2007, 3, 565−569.
(13) Sykes, R. B.; Koster, W. H.; Bonner, D. P. The new
(28) Boyd, D. B.; Herron, D. K.; Lunn, W. H. W.; Spitzer, W. A.
monobactams: Chemistry and biology. J. Clin. Pharmacol. 1988, 28, Electronic structures of cephalosporins and penicillins. 11. Parabolic
113−119. relationships between antibacterial activity of cephalosporins and β-
(14) Barbachyn, M. R.; Tuominen, T. C. Synthesis and structure- lactam reactivity predicted from molecular orbital calculations. J. Am.
activity relationships of monocarbams leading to U-78608. J. Antibiot. Chem. Soc. 1980, 102, 1812−1814.
1990, 43, 1199−1203. (29) Harada, S.; Tsubotani, S.; Hida, T.; Koyama, K.; Kondo, M.;
(15) Arnould, J. C.; Boutron, P.; Pasquet, M. J. Synthesis and Ono, H. Chemistry of a new antibiotic: Lactivicin. Tetrahedron 1988,
antibacterial activity of C-4 substituted monobactams. Eur. J. Med. 44, 6589−6606.
Chem. 1992, 27, 131−140. (30) Wolfe, S.; Wilson, M.-C.; Cheng, M.-H.; Shustov, G. V.; Akuche,
(16) Page, M. G. P.; Dantier, C.; Desarbre, E. In vitro properties of C. I. Cyclic hydroxamates, especially multiply substituted [1,2]-
BAL30072, a novel siderophore sulfactam with activity against oxazinan-3-ones. Can. J. Chem. 2003, 81, 937−960.
multiresistant Gram-negative bacilli. Antimicrob. Agents Chemother. (31) Baldwin, J. E.; Ng, S. C.; Pratt, A. J. Synthesis of phenoxyacetyl-
2010, 54, 2291−2302. N-sulphonyl cycloserine. Tetrahedron Lett. 1987, 28, 4319−4320.
(17) Mushtaq, S.; Warner, M.; Livermore, D. Activity of the (32) Ueda, Y.; Crast, L. B., Jr.; Mikkilineni, A. B.; Partyka, R. A.
siderophore monobactam BAL30072 against multiresistant non- Synthesis of phenoxyacetyl-N-(hydroxydioxocyclobutenyl)-
fermenters. J. Antimicrob. Chemother. 2010, 65, 266−270. cycloserines. Tetrahedron Lett. 1991, 32, 3767−3770.
(18) Flanagan, M. E.; Brickner, S. J.; Lall, M.; Casavant, J.; (33) Natsugari, H.; Kawano, Y.; Morimoto, A.; Yoshioka, K.; Ochiai,
Deschenes, L.; Finegan, S. M.; George, D. M.; Granskog, K.; M. Synthesis of lactivicin and its derivatives. J. Chem. Soc., Chem.
Hardink, J. R.; Huband, M. D.; Hoang, T.; Lamb, L.; Marra, A.; Commun. 1987, 62−63.
Mitton-Fry, M.; Mueller, J. P.; Mullins, L. M.; Noe, M. C.; O’Donnell, (34) Tamura, N.; Matsushita, Y.; Yoshioka, K.; Ochiai, M. Synthesis
J. P.; Pattavina, D.; Penzien, J. B.; Schuff, B. P.; Sun, J.; Whipple, D. A.; of lactivicin analogs. Tetrahedron 1988, 44, 3231−3240.
Young, J.; Gootz, T. D. Preparation, Gram-negative antibacterial (35) Tamura, N.; Matsushita, Y.; Kawano, Y.; Yoshioka, K. Synthesis
activity, and hydrolytic stability of novel siderophore-conjugated and antibacterial activity of lactivicin derivatives. Chem. Pharm. Bull.
monocarbam diols. ACS Med. Chem. Lett. 2011, 2, 385−390. 1990, 38, 116−122.
(19) Mitton-Fry, M. J.; Arcari, J. T.; Brown, M. F.; Casavant, J. M.; (36) Han, S.; Zaniewski, R. P.; Marr, E. S.; Lacey, B. M.; Tomaras, A.
Finegan, S. M.; Flanagan, M. E.; Gao, H.; George, D. M.; P.; Evdokimov, A.; Miller, J. R.; Shanmugasundaram, V. Structural
Gerstenberger, B. S.; Han, S.; Hardink, J. R.; Harris, T. M.; Hoang, basis for effectiveness of siderophore-conjugated monocarbams against
T.; Huband, M. D.; Irvine, R.; Lall, M. S.; Megan Lemmon, M.; Li, C.; clinically relevant strains of Pseudomonas aeruginosa. Proc. Natl. Acad.
Lin, J.; McCurdy, S. P.; Mueller, J. P.; Mullins, L.; Niosi, M.; Noe, M. Sci. U. S.A. 2010, 107, 22002−22007.
C.; Pattavina, D.; Penzien, J.; Plummer, M. S.; Risley, H.; Schuff, B. P.; (37) Baudart, M. G.; Hennequin, L. F. Synthesis and biological
activity of C-3′ ortho-dihydroxyphthalimido cephalosporins. J. Antibiot.
Shanmugasundaram, V.; Starr, J. T.; Sun, J.; Winton, J.; Young, J. A.
1993, 46, 1458−1470.
Novel monobactams utilizing a siderophore uptake mechanism for the
(38) Han, S.; Caspers, N.; Zaniewski, R. P.; Lacey, B. M.; Tomaras,
treatment of gram-negative infections. Bioorg. Med. Chem. Lett. 2012,
A. P.; Feng, X.; Geoghegan, K. F.; Shanmugasundaram, V. Distinctive
22, 5989−5994. attributes of β-lactam target proteins in Acinetobacter baumannii
(20) Brown, M. F.; Mitton-Fry, M. J.; Arcari, J. T.; Barham, R.; relevant to development of new antibiotics. J. Am. Chem. Soc. 2011,
Casavant, J.; Gerstenberger, B. S.; Han, S.; Hardink, J. R.; Harris, T. 133, 20536−20545.
M.; Hoang, T.; Huband, M. D.; Lall, M. S.; Lemmon, M. M.; Li, C.; (39) Winsor, G. L.; Lam, D. K. W.; Fleming, L.; Lo, R.; Whiteside, M.
Lin, J.; McCurdy, S. P.; McElroy, E.; McPherson, C.; Marr, E. S.; D.; Yu, N. Y.; Hancock, R. E. W.; Brinkman, F. S. L. Pseudomonas
Mueller, J. P.; Mullins, L.; Nikitenko, A. A.; Noe, M. C.; Penzien, J.; genome database: Improved comparative analysis and population
Plummer, M. S.; Schuff, B. P.; Shanmugasundaram, V.; Starr, J. T.; genomics capability for Pseudomonas genomes. Nucleic Acids Res.
Sun, J.; Tomaras, A.; Young, J. A.; Zaniewski, R. P. Pyridone- 2011, 39, D596−D600.
conjugated monobactam antibiotics with Gram-negative activity. J. (40) McPherson, C. J.; Aschenbrenner, L. M.; Lacey, B. M.; Fahnoe,
Med. Chem. 2013, 56, 5541−5552. K. C.; Lemmon, M. M.; Finegan, S. M.; Tadakamalla, B.; O’Donnell, J.
(21) Harada, S.; Tsubotani, S.; Hida, T.; Ono, H.; Okazaki, H. P.; Mueller, J. P.; Tomaras, A. P. Clinically relevant Gram-negative
Structure of lactivicin, an antibiotic having a new nucleus and similar resistance mechanisms have no effect on the efficacy of MC-1, a novel
biological activities to β-lactam antibiotics. Tetrahedron Lett. 1986, 27, siderophore-conjugated monocarbam. Antimicrob. Agents Chemother.
6229−6232. 2012, 56, 6334−6342.

3854 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855


Journal of Medicinal Chemistry Article

(41) Carpenter, B. M.; Whitmire, J. M.; Merrell, D. S. This is not


your mother’s repressor: The complex role of Fur in pathogenesis.
Infect. Immun. 2009, 77, 2590−2601.
(42) Tomaras, A. P.; Crandon, J. L.; McPherson, C. J.; Banevicius, M.
A.; Finegan, S. M.; Irvine, R. L.; Brown, M. F.; O’Donnell, J. P.;
Nicolau, D. P. Adaptation-based resistance to siderophore-conjugated
antibacterial agents by Pseudomonas aeruginosa. Antimicrob. Agents
Chemother. 2013, 57, 4197−4207.
(43) Brown, T., Jr.; Charlier, P.; Herman, R.; Schofield, C. J.;
Sauvage, E. Structural basis for the interaction of lactivicins with serine
β-lactamases. J. Med. Chem. 2010, 53, 5890−5894.
(44) Rodkey, E. A.; Drawz, S. M.; Sampson, J. M.; Bethel, C. R.;
Bonomo, R. A.; van den Akker, F. Crystal structure of a preacylation
complex of the β-lactamase inhibitor sulbactam bound to a
sulfenamide bond-containing thiol-β-lactamase. J. Am. Chem. Soc.
2012, 134, 16798−16804.
(45) Otwinowski, Z.; Minor, W. Processing of X-ray diffraction data
collected in oscillation mode. Methods Enzymol. 1997, 276, 307−326.
(46) McCoy, A. J.; Grosse-Kunstleve, R. W.; Adams, P. D.; Winn, M.
D.; Storoni, L. C.; Read, R. J. Phaser crystallographic software. J. Appl.
Crystallogr. 2007, 40, 658−674.
(47) Blanc, E.; Roversi, P.; Vonrhein, C.; Flensburg, C.; Lea, S. M.;
Bricogne, G. Refinement of severely incomplete structures with
maximum likelihood in BUSTER-TNT. Acta Crystallogr., Sect. D: Biol.
Crystallogr. 2004, 60, 2210−2221.
(48) Emsley, P.; Cowtan, K. Coot: model-building tools for
molecular graphics. Acta Crystallogr., Sect. D.: Biol. Crystallogr. 2004,
60, 2126−2132.
(49) Laskowski, R. A.; Moss, D. S.; Thornton, J. M. Main-chain bond
lengths and bond angles in protein structures. J. Mol. Biol. 1993, 231,
1049−1067.
(50) Frisch, M. J.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.;
Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson,
G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov,
A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.;
Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda,
Y.; Kitao, O.; Nakai, H.; Vreven, T.; J. A. Montgomery, J.; Peralta, J. E.;
Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.;
Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.;
Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega,
N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.;
Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.;
Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.;
Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.;
Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .;
Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09,
Revision C.1; Gaussian, Inc., Wallingford, CT, 2009.
(51) Andersson, M. P.; Uvdal, P. New scale factors for harmonic
vibrational frequencies using the B3LYP density functional method
with the triple-ζ basis set 6−311+G(d,p). J. Phys. Chem. A 2005, 109,
2937−2941.
(52) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal
solvation model based on solute electron density and on a continuum
model of the Solvent defined by the bulk dielectric constant and
atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378−6396.
(53) Gonzalez, C.; Schlegel, H. B. Reaction path following in mass-
weighted internal coordinates. J. Phys. Chem. 1990, 94, 5523−5527.
(54) Clinical Laboratory Standards Institute (CLSI). Methods for
dilution antimicrobial susceptibility tests for bacteria that grow
aerobically: approved standard. CLSI, M07, 2009a. CLSI. Performance
standards for antimicrobial susceptibility testing: 19th informational
supplement. CLSI, M100, 2009b.

3855 dx.doi.org/10.1021/jm500219c | J. Med. Chem. 2014, 57, 3845−3855

You might also like