SlideShare a Scribd company logo
1 of 11
Download to read offline
Review
Chemical Biology Tools
for Examining the Bacterial Cell Wall
Ashley R. Brown,1 Rebecca A. Gordon,2,3 Stephen N. Hyland,1 M. Sloan Siegrist,2,3 and Catherine L. Grimes1,4,*
1Department of Chemistry and Biochemistry, University of Delaware, Newark, DE 19716, USA
2Department of Microbiology, University of Massachusetts, Amherst, MA 01003-9298, USA
3Molecular and Cellular Biology Graduate Program, University of Massachusetts, Amherst 01003-9298, USA
4Department of Biological Sciences, University of Delaware, Newark, DE 19716, USA
*Correspondence: cgrimes@udel.edu
https://doi.org/10.1016/j.chembiol.2020.07.024
SUMMARY
Bacteria surround themselves with cell walls to maintain cell rigidity and protect against environmental in-
sults. Here we review chemical and biochemical techniques employed to study bacterial cell wall biogenesis.
Recent advances including the ability to isolate critical intermediates, metabolic approaches for probe incor-
poration, and isotopic labeling techniques have provided critical insight into the biochemistry of cell walls.
Fundamental manuscripts that have used these techniques to discover cell wall-interacting proteins, flip-
pases, and cell wall stoichiometry are discussed in detail. The review highlights that these powerful methods
and techniques have exciting potential to identify and characterize new targets for antibiotic development.
INTRODUCTION
The bacterial cell wall is arguably just as important to human
health as it is to the survival of the bacterium. The complex poly-
mers that comprise the cell wall provide bacteria with strength
and a barrier to the outside world, allowing them to thrive in a
multitude of environments, including the human body. Humans
have taken advantage of natural product antibiotics, often pro-
duced by bacteria themselves, to target bacterial polymers,
yielding some of the most widely used antibiotics to date (Cho-
pra and Roberts, 2001). In this review, we focus on understand-
ing the bacterial cell wall in the context of the threat that antibiotic
resistance poses to society. With the announcement that many
major pharmaceutical companies will no longer fund research
programs in this critical area (Hu, 2018), the onus falls on curious,
determined academics to identify new targets for antibiotics and
new opportunities to combat resistance. Here we highlight
recently reported chemical and biochemical approaches to
study bacterial cell wall biosynthesis and maintenance both in
the laboratory and in the clinic, focusing on the use of antibiotics.
In some cases, antibiotics are part of the toolkit that enables bio-
logical insight. In other cases, the biological insights gleaned
from applying the tools enable (or have the potential to enable)
target discovery. Studies of the bacterial cell wall and the antibi-
otics that corrupt it are iterative and promote both mechanistic
insights and translational applications.
THE TARGET AND ITS BASICS
Scientists have used small molecules to study bacteria since the
19th
century when Christian Gram treated cells with crystal violet
and realized that bacteria could be canonically divided into
two general classes: Gram-positive and Gram-negative
(Gram, 1884). In the present day, sophisticated tools exist to
visualize the cell wall and to dissect its composition and biosyn-
thesis at the molecular level (Hsu et al., 2019; Kocaoglu and Carl-
son, 2016; Radkov et al., 2018; Siegrist et al., 2015; Taguchi
et al., 2019a). Here we will review a subset of the new methods,
but first offer a brief introduction to cell walls, noting the many
recent, detailed reviews on these structures (e.g., Radkov
et al., 2018) and their biosynthesis and maintenance (Taguchi
et al., 2019a).
The bacterial cell envelope is a complex structure that pro-
vides protection from the external environment, maintains cell
shape, and provides resistance to chemical, physical, and bio-
logical damage (Figure 1A; Vollmer et al., 2008). Nearly all bacte-
rial envelopes have a peptidoglycan (PG) cell wall layer lying just
outside the plasma membrane. PG biosynthesis is a highly
conserved process in bacteria, starting with UDP-N-acetyl-
glucosamine (UDP-GlcNAc) conversion into UDP-N-acetyl-mur-
amic acid (UDP-MurNAc), as the first committed step in PG syn-
thesis (Figure 1B). The subsequent steps involve the addition of
amino acids (commonly L-Ala, D-g-Glu, L-Lys [Gram-positive] or
meso-diaminopimelic acid [mDAP; Gram-negative and myco-
bacteria], and D-Ala-D-Ala) to the UDP-MurNAc lactate moiety.
Variation exists within the stem peptide depending on species.
For example, Mycobacterium leprae can utilize Gly in place of
L-Ala at the one position (Draper et al., 1987), many Gram-posi-
tive bacteria and mycobacteria amidate the second position to
post-biosynthetically generate D-g-Gln, and spirochetes include
L-Orn at the third position (Schleifer and Kandler, 1972; Vollmer
et al., 2008). These and additional variations are highlighted in
Vollmer et al. (2008). Stem peptide ligation is followed by the
transfer of this PG building block to phosphate polyprenyl to
make Lipid I. GlcNAc is added to Lipid I to make Lipid II as the
final PG precursor. Lipid II is subsequently flipped across the
ll
1052 Cell Chemical Biology 27, August 20, 2020 ª 2020 Elsevier Ltd.
membrane by the MurJ flippase (Ruiz, 2008; Sham et al., 2014)
as the complete PG subunit. Class A penicillin-binding proteins
(aPBPs), L,D-transpeptidases, and shape, elongation, division,
and sporulation (SEDS) proteins in complex with class B PBPs
(bPBPs) assemble the PG through two different enzymatic reac-
tions, transglycosylation and transpeptidation (Cho et al., 2016;
Emami et al., 2017; Meeske et al., 2016; Taguchi et al., 2019b).
PG transglycosylases (TGs) link the sugar backbone of the PG
subunit to the next unit through a b-1,4 glycosidic linkage
(Vollmer et al., 2008). PG transpeptidases (TPs) most commonly
connect the fourth amino acid of one peptide chain to the third
amino acid (such as mDAP or L-Lys) of an adjacent strand
yielding 3–4 cross-linked PG (Vollmer and Seligman, 2010).
However, there are more variations that feature diverse connec-
tivity—3-3, 2-4, and 1-3 linkages—and bridge lengths across
species. While the biosynthesis of Lipid II is generally conserved
across bacteria, the polymerization of the monomer marks the
beginning of the diversification process (Egan et al., 2020; Typas
et al., 2011). Depending on the activity and the protein-protein in-
teractors of the TGs/TPs, a variety of shapes are formed (Do
et al., 2020; Salama, 2020). For example, in H. Pylori, the TG
and TP activity is spatially and temporally regulated along the
cell wall, which greatly influences the shape. This diversity of
Gram-positive and Gram-negative cells is further enhanced by
the inclusion of lipids in the cell wall.
In Gram-positive bacteria, the PG layer is significantly thicker
than that of Gram-negative (Vollmer and Seligman, 2010) and
features lipoteichoic and wall teichoic acids (WTAs) anchored
to the cytoplasmic membrane and the MurNAc 6-OH, respec-
tively. The WTA is assembled largely on the cytoplasmic face.
Upon delivery to the extracellular surface, it is anchored to PG
by the family of LytR-CpsA-Psr (LCP) enzymes (Kawai et al.,
2011; Li et al., 2020; Schaefer et al., 2017). Teichoic acid (TA) el-
ements are essential to the virulence of pathogenic bacteria.
They permit bacterial cell adhesion to host cells, in addition to
controlling cell wall remodeling by autolysins (Brown et al., 2013).
Gram-negative bacteria have a thinner PG but have additional
structural support and protection due to the asymmetric outer
membrane consisting of phospholipids and lipopolysaccharides
(LPS) (Rojas et al., 2018). LPS is made of three components: lipid
A, an oligosaccharide core, and the O-antigen. Toward the end
of the lipid A construction, the core oligosaccharide is added,
making the lipid oligosaccharide intermediate. The O-antigen is
synthesized separately and bound to lipooligosaccharide
(LOS), to make LPS, before transport to the outer membrane
(Simpson and Trent, 2019). LPS is not anchored to the PG but
rather inserted into the outer membrane by complex cellular ma-
chinery and hydrophobically adhered via Lipid A.
There is a third class of bacteria that has elements of both
Gram-positives and Gram-negatives; it is sometimes referred
Figure 1. Bacterial Cell Wall Basics
(A) Structural features of the cell wall of Gram-positives, Gram-negatives, and mycobacteria. PG, peptidoglycan; PM, plasma membrane; LPS, lipopolysac-
charide; MM, mycomembrane; OM, outer membrane; AG, arabinogalactan.
(B) Biosynthetic steps of peptidoglycan (PG): PG biosynthesis occurs in 14 unique biochemical steps, starting with UDP-GlcNAc conversion into UDP-MurNAc.
UDP-MurNAc is transferred to the lipid carrier and, subsequently, glycosylated by MurG. This process culminates in Lipid II being flipped across the membrane by
MurJ. Transglycosylases (TGs) and transpeptidases (TPs) incorporate the PG subunits into the growing PG polymer. Ultimately, peptidoglycan is altered through
the addition of small and large molecules post-synthetically that are not represented here, such as wall teichoic acids (WTAs).
ll
Cell Chemical Biology 27, August 20, 2020 1053
Review
to as Gram-indeterminate. M. tuberculosis and other members
of the Corynebacterineae suborder are phylogenetically related
to Gram-positive bacteria, but possess a waxy, relatively imper-
meable outer membrane-like structure (Figure 1A; Alderwick
et al., 2015). The core cell wall has a PG layer of intermediate
thickness that is covalently bound to the branched arabinogalac-
tan (AG). The AG layer acts as a scaffold for a covalently bound,
inner leaflet of mycolic acids. The outer leaflet possesses more
mycolic acids, along with intercalated glycolipids such as treha-
lose monomycolates (TMMs) and trehalose dimycolate (TDM)
(Dulberger et al., 2020). The three classes of bacteria discussed
here contain some similarities in the composition of their cell wall
(i.e., Lipid II; Figure 1B) and differences (i.e., lipid modifications).
It has been proposed that some of the lipid modifications that
help to differentiate Gram-positive, Gram-negative, and Gram-
indeterminate bacteria could provide mechanisms to uniquely
target specific bacteria with antibiotic therapy (Jackson et al.,
2013; Kuhn, 2019).
Bacterial cell walls have historically made an excellent target for
antibiotics. There are multiple ways to kill a bacterial cell, inhibiting
DNA replication and halting protein synthesis (Walsh, 2003). How-
ever, the bacterial cell wall is an especially attractive target
because it is unique to bacteria, meaning that human cells do
not contain the biochemical machinery that is required to build
and maintain it. In addition, the biosynthesis is largely conserved
(Figure 1B), allowing the development of broad-spectrum antibi-
otics. Unfortunately, bacteria have developed resistance to nearly
every cell wall-targeting antibiotic, including the well-known b-lac-
tams, such as penicillin and methicillin, and the antibiotics of ‘‘last
resort,’’ the glycopeptides (Kahne et al., 2005). Modes of resis-
tance can be conferred by, but not limited to, the expression of
insensitive PBPs, production of b-lactamases, activity of efflux
pumps, and alterations to the cell wall (Walsh, 2000). For a more
extensive review of these and other modes of resistance, the au-
thors refer the reader to Walsh (2000). In recent years, there has
been an emergence of multi-drug-resistant bacteria, particularly
in the ESKAPE pathogens (Enterococcus faecium, Staphylo-
coccus aureus, Klebsiella pneumoniae, Acinetobacter baumanii,
Pseudomonas aeruginosa, and Enterobacter species) (Boucher
et al., 2009; Santajit and Indrawattana, 2016). These bacteria,
together with tuberculosis (Bloom and Murray, 1992; Porter and
McAdam, 1994) and gonorrhea (Unemo and Shafer, 2014), pose
a viable threat with a lack of new antimicrobial compounds
(CDC, 2017). Within 1 year after the first clinical use of a natural
or synthetic antibiotic, resistance can develop (Walsh, 2003).
Therefore, there is a continuous demand for the development of
novel antibiotics via rational design, high-throughput screening ef-
forts, and medicinal chemistry campaigns.
In order to identify new bactericidal, cell wall-acting com-
pounds, peptidoglycan biosynthetic pathways (Figure 1) that
are less understood and not currently targeted by mainstream
antibiotics must be interrogated with appropriate tools. Fluores-
cent probes have permitted the close study of bacterial struc-
tures, and more recently probes such as fluorescent D-amino
acids (FDAAs) have enabled methods to screen for novel antibi-
otics and to identify their mode of action (Culp et al., 2020). There
are many excellent reviews that highlight direct imaging of bac-
terial cell walls to study the structure, biosynthesis, and dy-
namics of this polymer (Kocaoglu and Carlson, 2016; Radkov
et al., 2018; Siegrist et al., 2015; Taguchi et al., 2019a). Here
we focus on recent work that merges chemical and biochemical
methods including immunoblotting, photo-crosslinking, spec-
troscopy, and radiolabeling to probe PG biosynthetic pathways
and composition that can help inform the development of novel
antimicrobial therapeutics.
USING ANTIBIOTICS TO ISOLATE CELL WALL
INTERMEDIATES
In order to study bacterial cell wall biosynthetic enzymes, one
must have access to PG precursor substrates. A common strat-
egy is to use bacterial cell wall inhibitors such as vancomycin and
moenomycin to cause a buildup of intermediates. Such a tactic
was used by Strominger and Park in the 1950s to identify the
UDP-MurNAc intermediate, known as Park’s nucleotide, using
penicillin (Figure 1B) (Park, 1952; Strominger et al., 1959). The
lipid-linked precursor, Lipid II, was long sought in PG biosyn-
thesis (Figure 1B) as it is difficult to synthesize, and its scarcity
made in vitro studies of PG polymerization nearly impossible
(Lazar and Walker, 2002; Ye et al., 2001). Lipid II is the result of
the b-1,4 glycosylation of Lipid I by MurG with the monosaccha-
ride unit, GlcNAc (Figure 1B). This event produces a disaccha-
ride pentapeptide containing an undecaprenyl pyrophosphate
group. A major feat from the Walker Laboratory in 1999 showed
that Lipid I derivatives could be transformed with the glycosyl-
transferase MurG to form native Lipid II as well as a variety of de-
rivatives (Ha et al., 1999; Men et al., 1998). This in vitro enzymatic
conversion provided the first reliable method to produce work-
able quantities of Lipid II. Access to Lipid II, in turn, enabled
the mechanism of various glycopeptide antibiotics to be studied.
For example, vancomycin, as well as other glycopeptides such
as dalbavancin (Leimkuhler et al., 2005), has been shown to
inhibit TGs directly (Chen et al., 2003).
In 2017, a method to accumulate and isolate Lipid II directly
from cells was developed (Figure 2A) (Qiao et al., 2017). Qiao,
Kahne, and Walker developed a two-step extraction protocol of
Lipid II. Using a similar strategy as Strominger in 1958
and armed with the knowledge that PG TGs are responsible
for the last steps of PG biosynthesis (Figure 1B), antibiotics
that target these later steps were used to isolate native Lipid
II (Figure 2A). Cells treated with moenomycin or vancomycin,
which inhibit late-stage biosynthetic enzymes in the outer
leaflet of the plasma membrane (Chen et al., 2003; Kahne
et al., 2005; Taylor et al., 2006), accumulate a significant
amount of Lipid II (Figure 2A). However, when cells were
treated with a sublethal dose of fosfomycin, an antibiotic that
inhibits the first committed step in PG biosynthesis (Falagas
et al., 2016), Lipid II was undetectable (Qiao et al., 2017). The
ability to alter the amounts of PG biosynthetic intermediates
is a powerful tool for studying both upstream and downstream
effects of peptidoglycan synthesis, as evidenced by experi-
ments that have used the Lipid II isolation method to screen
for novel antibiotics, assess the function of critical PG biosyn-
thetic enzymes, and identify new enzymes involved in PG
biosynthesis (as discussed below).
To interrogate the production of Lipid II in a cellular context,
the Walker and Kahne labs have developed an easily accessible
technique to label Lipid II from bacterial cultures (Figure 2B)
ll
1054 Cell Chemical Biology 27, August 20, 2020
Review
(Qiao et al., 2014). The method involves detection of Lipid II from
growing bacterial cells by treating the intermediate with peni-
cillin-binding protein 4 (PBP4) to allow for the installation of a
biotin tag. This low molecular weight PBP isolated from
S. aureus is a promiscuous transpeptidase that, in addition to
catalyzing bonds between adjacent muropeptides, can ex-
change the terminal D-Ala of the muropeptide with various D-
amino acids. In vitro, this technique can switch the terminal D-
Ala for biotinylated D-Lys (BDL); biotinylated Lipid II is then
detectable via western blot (Figure 2B). This technique provides
Figure 2. Biochemical and Chemical Biology Techniques for Intergorating Bacterial Peptidolgycans
(A) Two-step extraction method for isolating Lipid II from bacterial cultures.
(B) PBP4 transpeptidase mediated terminal D-Ala exchange with unnatural amino acids. Depicted is the incorporation of biotin-D-Lys (BDL) to the stem peptide of
Lipid II. Lipid II consists of a diphosphate (P) disaccharide backbone, GlcNAc (G), and MurNAc (M), with a pentapeptide chain: alanine (A), glutamate (E), and
lysine (K).
(C) Substituted cysteine accessibility method (SCAM) utilizes single cysteine mutations (orange circle) in a protein of interest in conjunction with two cysteine
reactive reagents: MTSES and NEM. MTSES cannot penetrate the membrane and will only react with periplasmic cysteines. NEM can penetrate the membrane
and react with both periplasmic and cytoplasmic cysteine residues. Using this method with a cysteine mutant library, it can provide topological information of a
protein, in this case a transmembrane protein.
(D) MurJ-pBPA photocrosslinking assay for detection of Lipid II/MurJ adducts. MurJ encoded with single pBPA mutations undergoes crosslinking with Lipid II
upon UV activation. After SDS-PAGE and electroblotting, crosslinked Lipid II is biotinylated in-gel via a BDL exchange reaction and then detected via blotting with
streptavidin-HRP.
(E) Subsequent lysis and a click reaction to attach a fluorophore allow for analysis of mycolate-protein interactions via metabolic incorporation of a bifunctional
TMM analog. N-x-AlkTMM-C15 is metabolically incorporated into the mycobacterial mycomembrane (MM). Covalent crosslinks with MM-associated proteins
are induced by UV activation. Subsequent lysis and a click reaction to attach a fluorophore allow for analysis by a variety of techniques.
ll
Cell Chemical Biology 27, August 20, 2020 1055
Review
researchers with a streamlined process to obtain functionalized
Lipid II analogs that can be selectively detected, thus generating
a valuable assay for studying PG biosynthetic machinery and
screening antibiotics (Cochrane and Lohans, 2020).
In sum, there are at least three methods to access Lipid II and
its derivatives that have been reported: in vitro biochemical
methods using MurG, isolation of Lipid II from bacterial cultures,
and Lipid II biotinylation. Collectively, these methods have been
applied to diverse bacterial species: S. aureus, Bacillus subtilis,
Escherichia coli (Qiao et al., 2017), and M. smegmatis (Garcı́a-
Heredia et al., 2018) have all been successfully used for the Lipid
II isolation and/or visualization, despite variation in PG penta-
peptide chains (Qiao et al., 2017). Access to these substrates
will be important to fully understand the structural diversity pre-
sent in bacterial cell walls and to characterize pools of PG inter-
mediates during growth. These methods have been useful in
studying the mechanism of action of antibiotics, as mentioned
above with glycopeptides and more recently with lysobactin
(Lee et al., 2016). Research groups have also used them to study
the mechanisms of enzymes involved in PG biosynthesis. For
example, multiple labs have used the tools to biochemically
characterize the TG activity of the SEDS proteins, which have
been proposed as new potential targets for antibiotics (Cho
et al., 2016; Emami et al., 2017; Meeske et al., 2016; Rohs
et al., 2018; Sjodt et al., 2018; Taguchi et al., 2019b). The
methods of isolation and detection (Figures 2A and 2B) have
been used in combination in vitro to study the order of addition
of WTA precursors to PG intermediates (Figure 3B; this work
Figure 3. Critical Details for Understanding Peptidoglycan Isotopic Probe Incoroporation
(A) Cell wall depictions of S. aureus strains utilized to determine characteristic peaks of PG and TA.
(B) Assessing the ability of un-cross-linked and cross-linked PG as a substrate for WTA transfer by LcpB. Un-cross-linked PG oligomers featuring a pentaglycine
chain are bioenzymatically prepared with a SgtB mutant. The oligomers are then either modified with WTA via LcpB (top) or cross-linked with PBP4 (bottom).
Subsequently, these molecules are either subjected to transpeptidation with PBP4 or a ligation reaction with LcpB.
ll
1056 Cell Chemical Biology 27, August 20, 2020
Review
will be discussed in detail in Radiolabeling). Finally, these
methods have been used to identify and study the mechanism
of Lipid II transport (Rubino et al., 2020) as discussed in detail
in the next section.
BIOCHEMICAL TOOLS TO INVESTIGATE THE PG
ENZYMES
For a long time, it was not understood how Lipid II translocates
across the lipid bilayer into the periplasmic space for incorpora-
tion into the PG meshwork. Multiple proteins were evoked to
perform this function, including the SEDS protein FtsW, which
was later shown to be a transglycosylase (Egan et al., 2020;
Ruiz, 2016; Young, 2014). MurJ was discovered by Ruiz and col-
leagues to be the elusive flippase (Ruiz, 2008; Sham et al., 2014).
In a series of elegant experiments utilizing substituted cysteine
accessibility method (SCAM) (Figure 2C) and the protein toxin
ColM (Touzé et al., 2012), Ruiz and collaborators were able to
create an assay that provided context-dependent monitoring
of Lipid II movement. Using an extensive single Cys mutant
MurJ library, the authors mapped the topology of MurJ, which
was then used in parallel with protein modeling to predict
possible dynamic conformations (Butler et al., 2013). Building
on these findings, Sham et al. used an in vivo assay with the addi-
tion of sulfonating reagents, ColM, or both in either wild-type
MurJ or the reversible MurJ mutant to formalize MurJ as the Lipid
II flippase (Sham et al., 2014).
With the knowledge that MurJ was the flippase, the biochem-
ical details of its activity were investigated using the Lipid II as-
says discussed above. MurJ proved challenging to study
because, as a flippase, it does not structurally alter the Lipid II
precursor when it transports. However, by using the Lipid II
detection method (Figure 2B), it was possible to monitor
changes in Lipid II’s movement. In 2018, Rubino et al. showed
that treatment with a protonophore to disperse the proton-
motive force caused Lipid II accumulation in E. coli, suggesting
that the transportation of Lipid II is coupled to an electrochemical
gradient. The effect of the protonophore mimicked controls
known to disrupt MurJ activity. Furthermore, they characterized
the conformation of MurJ when the membrane potential is dissi-
pated by probing individual cysteine residues in MurJ (Rubino
et al., 2018).
To determine what residues of MurJ are involved in transport,
photocrosslinking experiments were used to tether interacting
partners. Photocrosslinking has become an invaluable method
to determine protein-substrate interactions (Lancia et al., 2014;
Parker and Pratt, 2020; Wu and Kohler, 2019). Unnatural amino
acid incorporation can be used to install a photoactivatable
p-benzoyl-L-phenylalanine (pBPA) in the protein of interest—in
this case, MurJ. Excitation of pBPA using 350–365 nm light leads
to a reactive diradical that forms a C-H bond to any vicinal func-
tional group within a 3.1 Å reactivity radius (Lancia et al., 2014). In
2012, Okuda et al. utilized this method to determine specific sites
where LPS interacts with the LPS transport (Lpt) machinery
(Okuda et al., 2012). Briefly, they incorporated pBPA residues
in several Lpt transport proteins. After UV activation, photo-
crosslinked adducts were identified using immunoblotting with
LPS-specific antibodies. This method allowed them to take
chemical snapshots of LPS transport and determine that shut-
tling of LPS across the periplasm is accomplished through cyto-
plasmic ATP hydrolysis. The Ruiz group, in collaboration with the
Kahne lab, applied this methodology to probe MurJ flippase ac-
tivity (Figure 2D) (Rubino et al., 2020). They hypothesized that
pBPA incorporated into MurJ would prompt crosslinking to its
natural substrate Lipid II. However, the lack of antibodies for
Lipid II made it necessary to implement a method that allowed
for the detection of the crosslinked adduct. PBP4 was used to
incorporate a BDL after photocrosslinking Lipid II to MurJ-
pBPA mutants to allow for detection (Figure 2D). They applied
this protocol to investigate the role of three essential arginine
residues, located in the central cavity of MurJ, that were previ-
ously proposed to be key in recognizing the pyrophosphate of
Lipid II (Kuk et al., 2019). Using single and multiple Arg/Ala mu-
tants, they observed similar levels of crosslinking compared to
wild-type MurJ. However, these mutants displayed impaired
ability to flip Lipid II. This methodology permitted the observation
of the intermediate transport steps in living cells and provided
direct, biochemical evidence that the conserved arginine resi-
dues control Lipid II movement through MurJ. Thus, through
the combination of Lipid II chemical probes, genetic tools, and
biochemical conversions, the function of MurJ has been identi-
fied and the biochemical mechanisms of Lipid II transport are
rapidly being unveiled. This also highlights MurJ as an exciting
target for antibiotic development. The ability to track Lipid II in
cellular biochemistry assays was critical because it yielded
detailed biochemical information (i.e., protein residues, mem-
brane potential) that would not have been possible with other
methods.
METABOLIC INCORPORATION OF
PHOTOCROSSLINKING SUGARS
The proteins that bind the cell wall and its associated glycocon-
jugates are also potential antibiotic targets. In contrast to pro-
tein-mediated interactions, glycan recognition events are often
weak and short-lived. Additionally, glycans are not directly
genetically encoded and their biosynthesis is complex, so it is
challenging to use standard genetic engineering methods to
tag them. The incorporation of functionalized metabolites has al-
lowed a way to bypass these challenges (Campbell et al., 2007).
Several groups have introduced unnatural sugars containing
photoactivatable crosslinkers to the cell by hijacking carbohy-
drate metabolic pathways and capitalizing on enzyme promiscu-
ity (Tanaka and Kohler, 2008; Yu et al., 2012). The mycomem-
brane (Figure 1A, ‘‘MM’’) is attached to the mycobacterial cell
wall via AG and is a barrier to environmental, immune, and anti-
biotic insults. However, its protein composition has eluded
classic biochemical techniques for a long time, in part because
of the difficulty of cleanly separating the covalently bound myco-
membrane from other layers of the complex mycobacterial enve-
lope. Kavunja et al. recently developed the first photocrosslink-
ing probes for the mycomembrane to analyze mycolate-protein
interactions in vivo (Figure 2E) (Kavunja et al., 2020). They syn-
thesized a TMM analog that specifically incorporates into the
TDM portion of the mycomembrane via previously reported
conserved, substrate-promiscuous Ag85 mycoloyltransferases
(Fiolek et al., 2019). This analog contains a bifunctional linker
bearing a photoactivatable diazirine group and a clickable alkyne
ll
Cell Chemical Biology 27, August 20, 2020 1057
Review
handle. After metabolic incorporation into the cell surface, myco-
bacteria were irradiated with UV light. The diazirine cross-linked
with neighboring proteins that were then enriched following click
ligation to an azide-fluorophore-biotin. This method allowed for
the identification of both known and previously undetectable my-
comembrane-resident proteins, as well as tracking them by in-
gel fluorescence. Similar techniques are likely to be useful in
future studies to interrogate different layers of the cell envelope,
including PG—especially as DeMeester and coworkers have
shown that PG precursors containing the diazirine cross-linkers
at the C2 position of MurNAc are accepted by the PG biosyn-
thetic enzymes (DeMeester et al., 2018). Previously, Sarkar
et al. exploited the MurF ligation process to insert a D-Ala- that
was biofunctionalized with an alkyne and photocrosslinking han-
dles into Lipid II to mine the protein-interacting partners (Sarkar
et al., 2016). Targeted acquisition of cell envelope interactomes
may reveal new potential targets for antibiotic therapies (Kavunja
et al., 2020).
SPECTROSCOPIC METHODS TO STUDY THE
CELL WALL
The stable isotopes 13
C and 15
N are effective as probes to eluci-
date structural features of the bacterial cell wall at the molecular
level using spectroscopic methods (Kim et al., 2014, 2015; Yang
et al., 2017). In this strategy, the macromolecules of biological in-
terest are unperturbed and uniform enrichment enhances signal
output of the NMR spectrum in a non-destructive manner (Ny-
gaard et al., 2015). Romaniuk and Cegelski utilized 13
C and
15
N cross-polarization magic angle spinning (CP/MAS) solid-
state NMR (ssNMR) to characterize the composition of uniformly
labeled PG and WTA in S. aureus (Romaniuk and Cegelski,
2018). This technique affords an alternative to solution-based
analytical methods that do not permit full characterization of
the highly insoluble material. Using this approach, spectra of pu-
rified PG and WTA isolates from wild-type and DtarO (a mutant
that is unable to synthesize WTA), respectively, were first used
to identify and quantify characteristic carbon peaks from each
component (Figure 3A). Since the sum of the two spectra repro-
duced the peak intensities of the intact cell wall sample, they
were able to quickly determine the composition ratio of TA to
PG. The masses for TA and PG calculated by this method
were consistent with those determined by phosphate analysis,
a more traditional but labor-intensive method. Subsequently,
the relative composition ratios of the two components in both
the stationary and the exponential phases were determined.
In the stationary phase, PG thickness increased while WTA
decreased. This finding was confirmed with selective labeling
using either D-[15
N]Ala (WTA) or [15
N]Gly (PG) in 15
N CP/MAS.
With these baselines in hand, they were able to validate that
this analytical method is amenable to determining the composi-
tion levels of TA and PG of cell walls with the antibiotic tunicamy-
cin, which inhibits WTA growth. Cegelski et al. have used similar
strategies with ssNMR to establish that vancomycin primarily
targets transglycosylation over transpeptidation using uniformly
13
C- and 15
N-labeled amino acids in S. aureus (Cegelski et al.,
2002). Changes to the D-alanine-pentaglycyl bridge-links were
unperturbed in the presence of vancomycin, suggesting that
transpeptidation is unaffected. Instead, this supports the idea
that vancomycin blocks transglycosylation and impedes translo-
cation of Lipid II into the periplasm. Rotational echo double reso-
nance (REDOR) and CP/MAS ssNMR have also been used to
discern perturbations to PG and WTA of S. aureus treated with
the cyclic decapeptide amphomycin, a drug that is effective
against superbugs such as multi-drug-resistant S. aureus and
vancomycin-resistant enterococci by targeting bactoprenol-
phosphate. Using REDOR, Singh et al. observed 15
N shifts in
bridge-links between Gly and L-Lys and the free side chain amine
of lysine, indicating that the compound induced PG thinning, the
accumulation of Park’s nucleotide, and a decrease in alanylation
of WTA (Singh et al., 2016). These data suggested that ampho-
mycin acts on the cell wall prior to transglycosylation. Overall,
CP/MAS and REDOR analysis of the bacterial cell wall permits
rapid assessment of whole-cell composition and can be applied
to monitor and determine the compositional perturbations
caused by antibiotics as showcased in the studies above.
In another example, Calabretta et al. employed 13
C radiola-
beled lipid-linked arabinofuranose donors to study Gram-inde-
terminate bacteria. These probes were biosynthetically incorpo-
rated into the arabinan layer of Corynebacterium glutamicum
and M. smegmatis strains unable to produce this polymer due
to genetic mutation or treatment with benzothiazinone antibi-
otics (Calabretta et al., 2019). This procedure circumvents the
requirement of metabolic processing of cell wall probes prior
to incorporation into the cell wall. Analysis of the soluble arabinan
isolated from these models retained characteristic peaks with 2D
NMR experiments (1
H-13
C HSQC, 1
H-13
C HMBC, and 1
H-13
C
HSQC-TOCSY). Though this study did not use intact cells to
gain structural details, it yielded a powerful tool that can be
applied to examine structural changes in the lesser-understood
mycobacterial cell envelope. Antibiotics are critically needed
against mycobacteria to combat public health threats such as
tuberculosis, which the Centers for Disease Control (CDC) has
recently identified as a top AMR threat (Bloom and Murray,
1992; CDC, 2017).
RADIOLABELING
Although NMR methods are useful in determining composition
of the cell wall, they do not inform on its biosynthesis. This can
be achieved using radioisotopes, as bacterial incorporation
of extremely sensitive ionizing radionuclides permits submicro-
molar (high fM to pM) detection of cell wall products and inter-
mediates of biosynthetic pathways in vivo. This sensitivity is
also useful in translational applications for disease diagnosis.
Here we discuss the use of 14
C and 11
C radionuclides in bac-
terial cell surface structures to discern mechanistic details of
PG synthesis and structure (long-lived) in addition to usage
as a diagnostic imaging tool of bacterial infection in vivo
(short-lived).
Long-Lived Radioisotopes
Elucidating some of the lesser-understood biosynthetic pro-
cesses has the potential to provide new targets for antimicro-
bials. TAs are important for the structure and virulence of
Gram-positive bacteria (Brown et al., 2013), but far less is known
about their synthesis compared to that of PG. Recently, Schaefer
et al. investigated the order and process by which WTA are
ll
1058 Cell Chemical Biology 27, August 20, 2020
Review
appended to PG. This is useful, as methicillin-resistant S. aureus
(MRSA) is resensitized to b-lactams when WTA biosynthesis is
inhibited (Campbell et al., 2011; Farha et al., 2013). They found
that WTA is ligated to un-crosslinked PG oligomers and that liga-
tion preference is due to steric restrictions of the LcpB WTA
ligase binding site (Schaefer et al., 2018). With isolated Lipid II
from S. aureus, featuring a pentaglycine chain for downstream
cross-linking, the group bioenzymatically prepared un-cross-
linked PG fragments (2–10 carbohydrates in length) with a trans-
glycosylase mutant (SgtBY181D
) that releases premature PG olig-
omers (Figure 3B). These fragments were then subjected to
transpeptidation with PBP4 followed by treatment with LcpB
(WTA ligase) and 14
C radiolabeled LIIA
WTA
, a truncated radiola-
beled WTA precursor, or just the latter to assess substrate pref-
erence. Detection by polyacrylamide gel electrophoresis (PAGE)
autoradiography showed that the cross-linked substrate is un-
able to be ligated to WTA, whereas the un-cross-linked PG frag-
ments serve as an acceptable substrate for the transfer of WTA
by LcpB. Additional experiments evaluating the structural re-
quirements for recognition co-crystalized the homolog TagT
from B. subtilis with either LIA
WTA
or LIIA
WTA
, which indicated a
putative binding groove narrow in size. The steric restrictions
of this site prompted Schaefer and coworkers to explore if the
stem peptide was a necessary feature of recognition by employ-
ing chitin, deacetylated chitin, and cellulose-based oligosaccha-
rides as probes of LcpB and TagT. Results indicated that the
stem peptide is not a necessary structural feature for recognition
of the transfer substrate; however, acetylation of the C2 position
of MurNAc is required. These radioisotopic probes enabled
Schaefer and coworkers to visually determine if PG intermedi-
ates were modified by WTA ligase in this enzymatic assay and
suggest the ligation order of WTA precursors to peptidoglycan
intermediates. With this fundamental information, new antibiotic
targets can be established as a means to stymie the ligation pro-
cess and ultimately disrupt the integrity of the cell wall of Gram-
positive bacteria.
Short-Lived Radioisotopes
Positron emission tomography (PET) radioisotopes such as 11
C,
13
N, 15
O, and 18
F are extremely sensitive tools that permit
whole-body imaging of physiological processes in deep-lying
tissues and organs with low nanogram quantities of probe. Un-
like stable isotopes, the radiation produced upon decay is
detectable outside of the human body. [18
F]-Fluorine deoxyglu-
cose (FDG) is a clinical PET probe frequently used to image cells
with higher energy requirements, such as oncogenic cells and
sites of inflammation (Zhuang and Alavi, 2002, Seminars in Nu-
clear Medicine, conference). [18
F]FDG accumulates because it
cannot be further metabolized after sequestration into the cell.
Since this probe is not bacteria-specific, improved imaging
technologies are required to detect active bacterial infections
in vivo (Ordonez and Jain, 2018, Seminars in Nuclear Medicine,
conference). To address this issue, Rosenberg, Ohliger, and
Wilson have developed clinically relevant 11
C radiotracers that
can distinguish between sterile inflammation and active infec-
tion by metabolically incorporating into both clinically relevant
pathogenic Gram-positive and Gram-negative bacteria (Neu-
mann et al., 2017; Parker et al., 2020; Stewart et al., 2020), uti-
lizing D-amino acid probe incorporation (de Pedro et al., 1997).
High incorporation of D-[14
C]Met in both S. aureus and E. coli
in the stem peptide of PG led to the development of a 11
C-syn-
thetic probe from a D-homocysteinethiolactone precursor
(Figure 4). Earlier experimentation had shown that exogenous
D-Met is amply integrated into the PG polymer of stationary bac-
teria (Caparrós et al., 1992; Lam et al., 2009). Moreover, a robust
synthetic strategy and efficient labeling probe for D-Ala and
D-Glu were not readily available. Using the L-[11
C]-Met tracer
for imaging in mice modeling myositis, they were able to detect
and differentiate between bacterial infection and sterile inflam-
mation, modeled by heat-killed bacteria, in the deltoid muscle
with great sensitivity. These probes are likely to be especially
useful for regions of the body that are normally sterile such as
the musculoskeletal, biliary, and nervous systems, as there will
not be significant background from the commensals. The syn-
thetic efficiency of the probe was later improved for use in the
clinical setting by using automated synthesis (Neumann et al.,
2017; Parker et al., 2020; Stewart et al., 2020). Most recently,
Parker et al. expanded the tracers to include amino acids
canonically incorporated into the stem peptide, D-[3-11
C]-Ala
and D-[3-11
C]-Ala-D-Ala probes (Figure 4). In vivo studies utilized
the D-Ala probe, since D-Ala showed anywhere from 2-3 times
greater uptake in E. coli and S. aureus over the dipeptide. More-
over, it was determined that the D-Ala probe generated high
levels of incorporation with the previous panel of bacteria tested
in the 2019 paper and that these probes were highly selective for
bacterial cells over mammalian cells. Infectious imaging chal-
lenge experiments also confirmed the ability of the D-Ala probe
to distinguish between live bacterial infection and sterile
Figure 4. Radiolabels in Living Bacteria’s Peptioglycans
Incorporation of radiolabeled probes, D-[methyl-11
C]-Met or D-[3-11
C]-Ala, post-biosynthetically via transpeptidases in clinically relevant bacteria to assess use
for in vivo labeling of active bacterial infections.
ll
Cell Chemical Biology 27, August 20, 2020 1059
Review
inflammation in mouse models. The tracer was able to monitor
cell wall-acting antibiotic (ampicillin) treatment of an E. coli bac-
terial infection, image infection in the intervertebral space, and
proficiently screen for pneumonia in mice. Thus, this class of
radiolabeled probes yielded an indispensable bacteria-specific
imaging strategy for pathogenic bacteria in difficult-to-access
physiological niches in clinical settings compared to currently
approved methods.
CONCLUSION
Biochemists have been using small molecules to study bacterial
cell walls since the 1800s. As improved probes have become
available, the ability to study the biochemical phenomena has
exploded. Here we have highlighted examples of chemical
probes that were used to reveal fundamental cell wall biochem-
istry or, in some cases, to identify bacterial pathology in a living
host. There is a rich intersection between chemical biology and
cell walls. It is this intersection that will be powerful in tackling
the problem of antibiotic resistance, as a means to identify
new targets and to characterize the mechanism of action for
promising compounds. We highlighted the recent advances in
the field such as monitoring PG precursors to elucidate MurJ,
TG, and TP activity; establishing new protein targets for anti-
biotic development through photocrosslinking techniques;
applying NMR spectroscopy to capture native cell wall compo-
sition; and demonstrating a highly sensitive method of detection
of cellular processes using radiolabels in vitro and in vivo. In
conjunction with current visualization and imaging techniques,
chemical, biochemical, and hybrid methods have the potential
to provide both mechanistic and structural information for valu-
able targets.
ACKNOWLEDGMENTS
This project was supported by a grant from the Glycoscience Common Fund
(U01 CA221230) and the Delaware COBRE program, as well as a grant from
the National Institute of General Medical Sciences-NIGMS (P20GM104316).
C.L.G. is a Pew Biomedical Scholar, Sloan Scholar, and Camille Dreyfus
Scholar and thanks the Pew Foundation, the Sloan Foundation for Science
Advancement, and the Dreyfus Foundation for support. M.S.S. thanks NIH
R21 AI144748 and NIH DP2 AI138238 for financial support. S.H. thanks the
NIH Chemistry-Biology Interface Program (GM133395) and the University of
Delaware Dissertation Fellowship for funding support.
DECLARATION OF INTERESTS
C.L.G. and M.S.S. have pending patents and patents relevant to this work.
REFERENCES
Alderwick, L.J., Harrison, J., Lloyd, G.S., and Birch, H.L. (2015). The mycobac-
terial cell wall—peptidoglycan and arabinogalactan. Cold Spring Harb. Per-
spect. Med. 5, a021113.
Bloom, B.R., and Murray, C.J.L. (1992). Tuberculosis: commentary on a ree-
mergent killer. Science 257, 1055–1064.
Boucher, H.W., Talbot, G.H., Bradley, J.S., Edwards, J.E., Gilbert, D., Rice,
L.B., Scheld, M., Spellberg, B., and Bartlett, J. (2009). Bad bugs, no drugs:
no ESKAPE! An update from the Infectious Diseases Society of America.
Clin. Infect. Dis. 48, 1–12.
Brown, S., Santa Maria, J.P., Jr., and Walker, S. (2013). Wall teichoic acids of
gram-positive bacteria. Annu. Rev. Microbiol. 67, 313–336.
Butler, E.K., Davis, R.M., Bari, V., Nicholson, P.A., and Ruiz, N. (2013). Struc-
ture-function analysis of MurJ reveals a solvent-exposed cavity containing res-
idues essential for peptidoglycan biogenesis in Escherichia coli. J. Bacteriol.
195, 4639–4649.
Calabretta, P.J., Hodges, H.L., Kraft, M.B., Marando, V.M., and Kiessling, L.L.
(2019). Bacterial cell wall modification with a glycolipid substrate. J. Am.
Chem. Soc. 141, 9262–9272.
Campbell, C.T., Sampathkumar, S.-G., and Yarema, K.J. (2007). Metabolic
oligosaccharide engineering: perspectives, applications, and future direc-
tions. Mol. Biosyst. 3, 187–194.
Campbell, J., Singh, A.K., Santa Maria, J.P., Jr., Kim, Y., Brown, S., Swoboda,
J.G., Mylonakis, E., Wilkinson, B.J., and Walker, S. (2011). Synthetic lethal
compound combinations reveal a fundamental connection between wall tei-
choic acid and peptidoglycan biosyntheses in Staphylococcus aureus. ACS
Chem. Biol. 6, 106–116.
Caparrós, M., Pisabarro, A.G., and de Pedro, M.A. (1992). Effect of D-amino
acids on structure and synthesis of peptidoglycan in Escherichia coli.
J. Bacteriol. 174, 5549–5559.
CDC (2017). Drug-resistant TB (Centers for Disease Control and Prevention).
https://www.cdc.gov/tb/topic/drtb/default.htm.
Cegelski, L., Kim, S.J., Hing, A.W., Studelska, D.R., O’Connor, R.D., Mehta,
A.K., and Schaefer, J. (2002). Rotational-echo double resonance characteriza-
tion of the effects of vancomycin on cell wall synthesis in Staphylococcus
aureus. Biochemistry 41, 13053–13058.
Chen, L., Walker, D., Sun, B., Hu, Y., Walker, S., and Kahne, D. (2003). Vanco-
mycin analogues active against vanA-resistant strains inhibit bacterial trans-
glycosylase without binding substrate. Proc. Natl. Acad. Sci. USA 100,
5658–5663.
Cho, H., Wivagg, C.N., Kapoor, M., Barry, Z., Rohs, P.D.A., Suh, H., Marto,
J.A., Garner, E.C., and Bernhardt, T.G. (2016). Bacterial cell wall biogenesis
is mediated by SEDS and PBP polymerase families functioning semi-autono-
mously. Nat. Microbiol. 1, 16172.
Chopra, I., and Roberts, M. (2001). Tetracycline antibiotics: mode of action,
applications, molecular biology, and epidemiology of bacterial resistance. Mi-
crobiol. Mol. Biol. Rev. 65, 232–260.
Cochrane, S.A., and Lohans, C.T. (2020). Breaking down the cell wall: strate-
gies for antibiotic discovery targeting bacterial transpeptidases. Eur. J. Med.
Chem. 194, 112262.
Culp, E.J., Waglechner, N., Wang, W., Fiebig-Comyn, A.A., Hsu, Y.-P., Koteva,
K., Sychantha, D., Coombes, B.K., Van Nieuwenhze, M.S., Brun, Y.V., and
Wright, G.D. (2020). Evolution-guided discovery of antibiotics that inhibit pepti-
doglycan remodelling. Nature 578, 582–587.
de Pedro, M.A., Quintela, J.C., Höltje, J.V., and Schwarz, H. (1997). Murein
segregation in Escherichia coli. J. Bacteriol. 179, 2823–2834.
DeMeester, K.E., Liang, H., Jensen, M.R., Jones, Z.S., D’Ambrosio, E.A.,
Scinto, S.L., Zhou, J., and Grimes, C.L. (2018). Synthesis of functionalized
N-acetyl muramic acids to probe bacterial cell wall recycling and biosynthesis.
J. Am. Chem. Soc. 140, 9458–9465.
Do, T., Page, J.E., and Walker, S. (2020). Uncovering the activities, biological
roles, and regulation of bacterial cell wall hydrolases and tailoring enzymes.
J. Biol. Chem. 295, 3347–3361.
Draper, P., Kandler, O., and Darbre, A. (1987). Peptidoglycan and arabinoga-
lactan of Mycobacterium leprae. J. Gen. Microbiol. 133, 1187–1194.
Dulberger, C.L., Rubin, E.J., and Boutte, C.C. (2020). The mycobacterial cell
envelope - a moving target. Nat. Rev. Microbiol. 18, 47–59.
Egan, A.J.F., Errington, J., and Vollmer, W. (2020). Regulation of peptidoglycan
synthesis and remodelling. Nat. Rev. Microbiol. 18, 446–460.
Emami, K., Guyet, A., Kawai, Y., Devi, J., Wu, L.J., Allenby, N., Daniel, R.A.,
and Errington, J. (2017). RodA as the missing glycosyltransferase in Bacillus
subtilis and antibiotic discovery for the peptidoglycan polymerase pathway.
Nat. Microbiol. 2, 16253.
Falagas, M.E., Vouloumanou, E.K., Samonis, G., and Vardakas, K.Z. (2016).
Fosfomycin. Clin. Microbiol. Rev. 29, 321–347.
ll
1060 Cell Chemical Biology 27, August 20, 2020
Review
Farha, M.A., Leung, A., Sewell, E.W., D’Elia, M.A., Allison, S.E., Ejim, L., Per-
eira, P.M., Pinho, M.G., Wright, G.D., and Brown, E.D. (2013). Inhibition of
WTA synthesis blocks the cooperative action of PBPs and sensitizes MRSA
to b-lactams. ACS Chem. Biol. 8, 226–233.
Fiolek, T.J., Banahene, N., Kavunja, H.W., Holmes, N.J., Rylski, A.K., Pohane,
A.A., Siegrist, M.S., and Swarts, B.M. (2019). Engineering the mycomembrane
of live mycobacteria with an expanded set of trehalose monomycolate ana-
logues. ChemBioChem 20, 1282–1291.
Garcı́a-Heredia, A., Pohane, A.A., Melzer, E.S., Carr, C.R., Fiolek, T.J., Run-
dell, S.R., Lim, H.C., Wagner, J.C., Morita, Y.S., Swarts, B.M., and Siegrist,
M.S. (2018). Peptidoglycan precursor synthesis along the sidewall of pole-
growing mycobacteria. eLife 7, e37243.
Gram, C. (1884). The differential staining of Schizomycetes in tissue sections
and in dried preparations. Fortschr. Med. 2, 185–189.
Ha, S., Chang, E., Lo, M.C., Men, H., Park, P., Ge, M., and Walker, S. (1999).
The kinetic characterization of Escherichia coli MurG using synthetic substrate
analogues. J. Am. Chem. Soc. 121, 8415–8426.
Hsu, Y.-P., Booher, G., Egan, A., Vollmer, W., and VanNieuwenhze, M.S.
(2019). d-amino acid derivatives as in situ probes for visualizing bacterial pepti-
doglycan biosynthesis. Acc. Chem. Res. 52, 2713–2722.
Hu, C. (2018). Pharmaceutical companies are backing away from a
growing threat that could kill 10 million people a year by 2050 (Business In-
sider). https://www.businessinsider.com/major-pharmaceutical-companies-
dropping-antibiotic-projects-superbugs-2018-7.
Jackson, M., McNeil, M.R., and Brennan, P.J. (2013). Progress in targeting cell
envelope biogenesis in Mycobacterium tuberculosis. Future Microbiol. 8,
855–875.
Kahne, D., Leimkuhler, C., Lu, W., and Walsh, C. (2005). Glycopeptide and lip-
oglycopeptide antibiotics. Chem. Rev. 105, 425–448.
Kavunja, H.W., Biegas, K.J., Banahene, N., Stewart, J.A., Piligian, B.F., Groe-
nevelt, J.M., Sein, C.E., Morita, Y.S., Niederweis, M., Siegrist, M.S., and
Swarts, B.M. (2020). Photoactivatable glycolipid probes for identifying myco-
late-protein interactions in live mycobacteria. J. Am. Chem. Soc. 142,
7725–7731.
Kawai, Y., Marles-Wright, J., Cleverley, R.M., Emmins, R., Ishikawa, S., Ku-
wano, M., Heinz, N., Bui, N.K., Hoyland, C.N., Ogasawara, N., et al. (2011).
A widespread family of bacterial cell wall assembly proteins. EMBO J. 30,
4931–4941.
Kim, S.J., Singh, M., Sharif, S., and Schaefer, J. (2014). Cross-link formation
and peptidoglycan lattice assembly in the FemA mutant of Staphylococcus
aureus. Biochemistry 53, 1420–1427.
Kim, S.J., Chang, J., and Singh, M. (2015). Peptidoglycan architecture of
Gram-positive bacteria by solid-state NMR. Biochim. Biophys. Acta 1848 (1
Pt B), 350–362.
Kocaoglu, O., and Carlson, E.E. (2016). Progress and prospects for small-
molecule probes of bacterial imaging. Nat. Chem. Biol. 12, 472–478.
Kuhn, A. (2019). The bacterial cell wall and membrane-a treasure chest for anti-
biotic targets. In Bacterial Cell Walls and Membranes, A. Kuhn, ed.
(Springer), pp. 1–5.
Kuk, A.C.Y., Hao, A., Guan, Z., and Lee, S.-Y. (2019). Visualizing conformation
transitions of the Lipid II flippase MurJ. Nat. Commun. 10, 1736.
Lam, H., Oh, D.-C., Cava, F., Takacs, C.N., Clardy, J., de Pedro, M.A., and
Waldor, M.K. (2009). D-amino acids govern stationary phase cell wall remod-
eling in bacteria. Science 325, 1552–1555.
Lancia, J.K., Nwokoye, A., Dugan, A., Joiner, C., Pricer, R., and Mapp, A.K.
(2014). Sequence context and crosslinking mechanism affect the efficiency
of in vivo capture of a protein-protein interaction. Biopolymers 101, 391–397.
Lazar, K., and Walker, S. (2002). Substrate analogues to study cell-wall biosyn-
thesis and its inhibition. Curr. Opin. Chem. Biol. 6, 786–793.
Lee, W., Schaefer, K., Qiao, Y., Srisuknimit, V., Steinmetz, H., M€
uller, R.,
Kahne, D., and Walker, S. (2016). The mechanism of action of lysobactin.
J. Am. Chem. Soc. 138, 100–103.
Leimkuhler, C., Chen, L., Barrett, D., Panzone, G., Sun, B., Falcone, B.,
Oberth€
ur, M., Donadio, S., Walker, S., and Kahne, D. (2005). Differential inhibi-
tion of Staphylococcus aureus PBP2 by glycopeptide antibiotics. J. Am.
Chem. Soc. 127, 3250–3251.
Li, F.K.K., Rosell, F.I., Gale, R.T., Simorre, J.P., Brown, E.D., and Strynadka,
N.C.J. (2020). Crystallographic analysis of Staphylococcus aureus LcpA, the
primary wall teichoic acid ligase. J. Biol. Chem. 295, 2629–2639.
Meeske, A.J., Riley, E.P., Robins, W.P., Uehara, T., Mekalanos, J.J., Kahne,
D., Walker, S., Kruse, A.C., Bernhardt, T.G., and Rudner, D.Z. (2016). SEDS
proteins are a widespread family of bacterial cell wall polymerases. Nature
537, 634–638.
Men, H.B., Park, P., Ge, M., and Walker, S. (1998). Substrate synthesis and ac-
tivity assay for MurG. J. Am. Chem. Soc. 120, 2484–2485.
Neumann, K.D., Villanueva-Meyer, J.E., Mutch, C.A., Flavell, R.R., Blecha,
J.E., Kwak, T., Sriram, R., VanBrocklin, H.F., Rosenberg, O.S., Ohliger, M.A.,
and Wilson, D.M. (2017). Imaging active infection in vivo using D-amino acid
derived PET radiotracers. Sci. Rep. 7, 7903.
Nygaard, R., Romaniuk, J.A.H., Rice, D.M., and Cegelski, L. (2015). Spectral
snapshots of bacterial cell-wall composition and the influence of antibiotics
by whole-cell NMR. Biophys. J. 108, 1380–1389.
Okuda, S., Freinkman, E., and Kahne, D. (2012). Cytoplasmic ATP hydrolysis
powers transport of lipopolysaccharide across the periplasm in E. coli. Sci-
ence 338, 1214–1217.
Park, J.T. (1952). Uridine-50
-pyrophosphate derivatives. II. A structure com-
mon to three derivatives. J. Biol. Chem. 194, 885–895.
Parker, C.G., and Pratt, M.R. (2020). Click chemistry in proteomic investiga-
tions. Cell 180, 605–632.
Parker, M.F.L., Luu, J.M., Schulte, B., Huynh, T.L., Stewart, M.N., Sriram, R.,
Yu, M.A., Jivan, S., Turnbaugh, P.J., Flavell, R.R., et al. (2020). Sensing living
bacteria in vivo using d-alanine-derived 11
C radiotracers. ACS Cent. Sci. 6,
155–165.
Porter, J.D.H., and McAdam, K.P. (1994). The re-emergence of tuberculosis.
Annu. Rev. Public Health 15, 303–323.
Qiao, Y., Lebar, M.D., Schirner, K., Schaefer, K., Tsukamoto, H., Kahne, D.,
and Walker, S. (2014). Detection of lipid-linked peptidoglycan precursors by
exploiting an unexpected transpeptidase reaction. J. Am. Chem. Soc. 136,
14678–14681.
Qiao, Y., Srisuknimit, V., Rubino, F., Schaefer, K., Ruiz, N., Walker, S., and
Kahne, D. (2017). Lipid II overproduction allows direct assay of transpeptidase
inhibition by b-lactams. Nat. Chem. Biol. 13, 793–798.
Radkov, A.D., Hsu, Y.-P., Booher, G., and VanNieuwenhze, M.S. (2018). Imag-
ing bacterial cell wall biosynthesis. Annu. Rev. Biochem. 87, 991–1014.
Rohs, P.D.A., Buss, J., Sim, S.I., Squyres, G.R., Srisuknimit, V., Smith, M.,
Cho, H., Sjodt, M., Kruse, A.C., Garner, E.C., et al. (2018). A central role for
PBP2 in the activation of peptidoglycan polymerization by the bacterial cell
elongation machinery. PLoS Genet. 14, e1007726.
Rojas, E.R., Billings, G., Odermatt, P.D., Auer, G.K., Zhu, L., Miguel, A., Chang,
F., Weibel, D.B., Theriot, J.A., and Huang, K.C. (2018). The outer membrane is
an essential load-bearing element in Gram-negative bacteria. Nature 559,
617–621.
Romaniuk, J.A.H., and Cegelski, L. (2018). Peptidoglycan and teichoic acid
levels and alterations in Staphylococcus aureus by cell-wall and whole-cell nu-
clear magnetic resonance. Biochemistry 57, 3966–3975.
Rubino, F.A., Kumar, S., Ruiz, N., Walker, S., and Kahne, D.E. (2018). Mem-
brane potential is required for MurJ function. J. Am. Chem. Soc. 140,
4481–4484.
Rubino, F.A., Mollo, A., Kumar, S., Butler, E.K., Ruiz, N., Walker, S., and Kahne,
D.E. (2020). Detection of transport intermediates in the peptidoglycan flippase
MurJ identifies residues essential for conformational cycling. J. Am. Chem.
Soc. 142, 5482–5486.
Ruiz, N. (2008). Bioinformatics identification of MurJ (MviN) as the peptido-
glycan lipid II flippase in Escherichia coli. Proc. Natl. Acad. Sci. USA 105,
15553–15557.
ll
Cell Chemical Biology 27, August 20, 2020 1061
Review
Ruiz, N. (2016). Lipid flippases for bacterial peptidoglycan biosynthesis. Lipid
Insights 8 (Suppl 1), 21–31.
Salama, N.R. (2020). Cell morphology as a virulence determinant: lessons from
Helicobacter pylori. Curr. Opin. Microbiol. 54, 11–17.
Santajit, S., and Indrawattana, N. (2016). Mechanisms of antimicrobial resis-
tance in ESKAPE pathogens. BioMed Res. Int. 2016, 2475067.
Sarkar, S., Libby, E.A., Pidgeon, S.E., Dworkin, J., and Pires, M.M. (2016).
In vivo probe of Lipid II-interacting proteins. Angew. Chem. Int. Ed. Engl. 55,
8401–8404.
Schaefer, K., Matano, L.M., Qiao, Y., Kahne, D., and Walker, S. (2017). In vitro
reconstitution demonstrates the cell wall ligase activity of LCP proteins. Nat.
Chem. Biol. 13, 396–401.
Schaefer, K., Owens, T.W., Kahne, D., and Walker, S. (2018). Substrate pref-
erences establish the order of cell wall assembly in Staphylococcus aureus.
J. Am. Chem. Soc. 140, 2442–2445.
Schleifer, K.H., and Kandler, O. (1972). Peptidoglycan types of bacterial cell
walls and their taxonomic implications. Bacteriol. Rev. 36, 407–477.
Sham, L.-T., Butler, E.K., Lebar, M.D., Kahne, D., Bernhardt, T.G., and Ruiz, N.
(2014). Bacterial cell wall. MurJ is the flippase of lipid-linked precursors for
peptidoglycan biogenesis. Science 345, 220–222.
Siegrist, M.S., Swarts, B.M., Fox, D.M., Lim, S.A., and Bertozzi, C.R. (2015).
Illumination of growth, division and secretion by metabolic labeling of the bac-
terial cell surface. FEMS Microbiol. Rev. 39, 184–202.
Simpson, B.W., and Trent, M.S. (2019). Pushing the envelope: LPS modifica-
tions and their consequences. Nat. Rev. Microbiol. 17, 403–416.
Singh, M., Chang, J., Coffman, L., and Kim, S.J. (2016). Solid-state NMR char-
acterization of amphomycin effects on peptidoglycan and wall teichoic acid
biosyntheses in Staphylococcus aureus. Sci. Rep. 6, 31757.
Sjodt, M., Brock, K., Dobihal, G., Rohs, P.D.A., Green, A.G., Hopf, T.A.,
Meeske, A.J., Srisuknimit, V., Kahne, D., Walker, S., et al. (2018). Structure
of the peptidoglycan polymerase RodA resolved by evolutionary coupling
analysis. Nature 556, 118–121.
Stewart, M.N., Parker, M.F.L., Jivan, S., Luu, J.M., Huynh, T.L., Schulte, B.,
Seo, Y., Blecha, J.E., Villanueva-Meyer, J.E., Flavell, R.R., et al. (2020). High
enantiomeric excess in-loop synthesis of d-[methyl-11
C]methionine for use
as a diagnostic positron emission tomography radiotracer in bacterial infec-
tion. ACS Infect. Dis. 6, 43–49.
Strominger, J.L., Park, J.T., and Thompson, R.E. (1959). Composition of the
cell wall of Staphylococcus aureus: its relation to the mechanism of action of
penicillin. J. Biol. Chem. 234, 3263–3268.
Taguchi, A., Kahne, D., and Walker, S. (2019a). Chemical tools to characterize
peptidoglycan synthases. Curr. Opin. Chem. Biol. 53, 44–50.
Taguchi, A., Welsh, M.A., Marmont, L.S., Lee, W., Sjodt, M., Kruse, A.C.,
Kahne, D., Bernhardt, T.G., and Walker, S. (2019b). FtsW is a peptidoglycan
polymerase that is functional only in complex with its cognate penicillin-bind-
ing protein. Nat. Microbiol. 4, 587–594.
Tanaka, Y., and Kohler, J.J. (2008). Photoactivatable crosslinking sugars for
capturing glycoprotein interactions. J. Am. Chem. Soc. 130, 3278–3279.
Taylor, J.G., Li, X., Oberth€
ur, M., Zhu, W., and Kahne, D.E. (2006). The total
synthesis of moenomycin A. J. Am. Chem. Soc. 128, 15084–15085.
Touzé, T., Barreteau, H., El Ghachi, M., Bouhss, A., Barnéoud-Arnoulet, A.,
Patin, D., Sacco, E., Blanot, D., Arthur, M., Duché, D., et al. (2012). Colicin
M, a peptidoglycan lipid-II-degrading enzyme: potential use for antibacterial
means? Biochem. Soc. Trans. 40, 1522–1527.
Typas, A., Banzhaf, M., Gross, C.A., and Vollmer, W. (2011). From the regula-
tion of peptidoglycan synthesis to bacterial growth and morphology. Nat. Rev.
Microbiol. 10, 123–136.
Unemo, M., and Shafer, W.M. (2014). Antimicrobial resistance in Neisseria
gonorrhoeae in the 21st century: past, evolution, and future. Clin. Microbiol.
Rev. 27, 587–613.
Vollmer, W., and Seligman, S.J. (2010). Architecture of peptidoglycan: more
data and more models. Trends Microbiol. 18, 59–66.
Vollmer, W., Blanot, D., and de Pedro, M.A. (2008). Peptidoglycan structure
and architecture. FEMS Microbiol. Rev. 32, 149–167.
Walsh, C. (2000). Molecular mechanisms that confer antibacterial drug resis-
tance. Nature 406, 775–781.
Walsh, C. (2003). Where will new antibiotics come from? Nat. Rev. Microbiol.
1, 65–70.
Wu, H., and Kohler, J. (2019). Photocrosslinking probes for capture of carbo-
hydrate interactions. Curr. Opin. Chem. Biol. 53, 173–182.
Yang, H., Singh, M., Kim, S.J., and Schaefer, J. (2017). Characterization of the
tertiary structure of the peptidoglycan of Enterococcus faecalis. Biochim. Bio-
phys. Acta Biomembr. 1859, 2171–2180.
Ye, X.-Y., Lo, M.-C., Brunner, L., Walker, D., Kahne, D., and Walker, S. (2001).
Better substrates for bacterial transglycosylases. J. Am. Chem. Soc. 123,
3155–3156.
Young, K.D. (2014). Microbiology. A flipping cell wall ferry. Science 345,
139–140.
Yu, S.-H., Boyce, M., Wands, A.M., Bond, M.R., Bertozzi, C.R., and Kohler,
J.J. (2012). Metabolic labeling enables selective photocrosslinking of O-
GlcNAc-modified proteins to their binding partners. Proc. Natl. Acad. Sci.
USA 109, 4834–4839.
ll
1062 Cell Chemical Biology 27, August 20, 2020
Review

More Related Content

Similar to nadira.pdf

Microbiology antibiotic
Microbiology   antibioticMicrobiology   antibiotic
Microbiology antibioticMBBS IMS MSU
 
TB DRUG TARGETS
TB DRUG TARGETSTB DRUG TARGETS
TB DRUG TARGETSebushura
 
S. pyogenes, its virulence, antibiotic, phytochemicals
S. pyogenes, its virulence, antibiotic, phytochemicalsS. pyogenes, its virulence, antibiotic, phytochemicals
S. pyogenes, its virulence, antibiotic, phytochemicalsUniversité Laval
 
On the Modulation of Biocompatibility of Hydrogels with Collagen and Guar Gum...
On the Modulation of Biocompatibility of Hydrogels with Collagen and Guar Gum...On the Modulation of Biocompatibility of Hydrogels with Collagen and Guar Gum...
On the Modulation of Biocompatibility of Hydrogels with Collagen and Guar Gum...IIJSRJournal
 
ESKAPE Pathogens.pptx
ESKAPE Pathogens.pptxESKAPE Pathogens.pptx
ESKAPE Pathogens.pptxMayuriRani3
 
2014 E cells and mats 27 332-349
2014 E cells and mats 27 332-3492014 E cells and mats 27 332-349
2014 E cells and mats 27 332-349Helen Cox
 
Bacterial cytology cell wall and cell membrane
Bacterial cytology   cell wall and cell membraneBacterial cytology   cell wall and cell membrane
Bacterial cytology cell wall and cell membraneVishrut Ghare
 
Antimicrob. agents chemother. 1999-mingeot-leclercq-727-37
Antimicrob. agents chemother. 1999-mingeot-leclercq-727-37Antimicrob. agents chemother. 1999-mingeot-leclercq-727-37
Antimicrob. agents chemother. 1999-mingeot-leclercq-727-37microbiologia.dad
 
Understanding the Potency of β-Lactam Antibiotics
Understanding the Potency of β-Lactam AntibioticsUnderstanding the Potency of β-Lactam Antibiotics
Understanding the Potency of β-Lactam AntibioticsSAYAN DAS
 
Gram ve bacilla
Gram   ve bacillaGram   ve bacilla
Gram ve bacillashanonps
 
Metabolic co-dependence gives rise to collective oscillations within biofilms.
Metabolic co-dependence gives rise to collective oscillations within biofilms.Metabolic co-dependence gives rise to collective oscillations within biofilms.
Metabolic co-dependence gives rise to collective oscillations within biofilms.Kazuya Horibe
 
Presentazione ii anno campana
Presentazione ii anno campanaPresentazione ii anno campana
Presentazione ii anno campanalab13unisa
 
Receptor Sub-types _Pharmacodynamics_Pharmacology.
Receptor Sub-types _Pharmacodynamics_Pharmacology.Receptor Sub-types _Pharmacodynamics_Pharmacology.
Receptor Sub-types _Pharmacodynamics_Pharmacology.Megh Vithalkar
 
Bacteria cell structure.pptx
 Bacteria cell structure.pptx Bacteria cell structure.pptx
Bacteria cell structure.pptxShayanAhmad48
 

Similar to nadira.pdf (20)

Microbiology antibiotic
Microbiology   antibioticMicrobiology   antibiotic
Microbiology antibiotic
 
TB DRUG TARGETS
TB DRUG TARGETSTB DRUG TARGETS
TB DRUG TARGETS
 
S. pyogenes, its virulence, antibiotic, phytochemicals
S. pyogenes, its virulence, antibiotic, phytochemicalsS. pyogenes, its virulence, antibiotic, phytochemicals
S. pyogenes, its virulence, antibiotic, phytochemicals
 
On the Modulation of Biocompatibility of Hydrogels with Collagen and Guar Gum...
On the Modulation of Biocompatibility of Hydrogels with Collagen and Guar Gum...On the Modulation of Biocompatibility of Hydrogels with Collagen and Guar Gum...
On the Modulation of Biocompatibility of Hydrogels with Collagen and Guar Gum...
 
Bacteriology
BacteriologyBacteriology
Bacteriology
 
Bacteriology
BacteriologyBacteriology
Bacteriology
 
ESKAPE Pathogens.pptx
ESKAPE Pathogens.pptxESKAPE Pathogens.pptx
ESKAPE Pathogens.pptx
 
2014 E cells and mats 27 332-349
2014 E cells and mats 27 332-3492014 E cells and mats 27 332-349
2014 E cells and mats 27 332-349
 
Bacterial cytology cell wall and cell membrane
Bacterial cytology   cell wall and cell membraneBacterial cytology   cell wall and cell membrane
Bacterial cytology cell wall and cell membrane
 
Antimicrob. agents chemother. 1999-mingeot-leclercq-727-37
Antimicrob. agents chemother. 1999-mingeot-leclercq-727-37Antimicrob. agents chemother. 1999-mingeot-leclercq-727-37
Antimicrob. agents chemother. 1999-mingeot-leclercq-727-37
 
Wheat gluten and celiac disease 2013
Wheat gluten and celiac disease 2013Wheat gluten and celiac disease 2013
Wheat gluten and celiac disease 2013
 
Understanding the Potency of β-Lactam Antibiotics
Understanding the Potency of β-Lactam AntibioticsUnderstanding the Potency of β-Lactam Antibiotics
Understanding the Potency of β-Lactam Antibiotics
 
Project
ProjectProject
Project
 
Gram ve bacilla
Gram   ve bacillaGram   ve bacilla
Gram ve bacilla
 
Acta Crystalo D
Acta Crystalo DActa Crystalo D
Acta Crystalo D
 
Metabolic co-dependence gives rise to collective oscillations within biofilms.
Metabolic co-dependence gives rise to collective oscillations within biofilms.Metabolic co-dependence gives rise to collective oscillations within biofilms.
Metabolic co-dependence gives rise to collective oscillations within biofilms.
 
Presentazione ii anno campana
Presentazione ii anno campanaPresentazione ii anno campana
Presentazione ii anno campana
 
Receptor Sub-types _Pharmacodynamics_Pharmacology.
Receptor Sub-types _Pharmacodynamics_Pharmacology.Receptor Sub-types _Pharmacodynamics_Pharmacology.
Receptor Sub-types _Pharmacodynamics_Pharmacology.
 
Bacteria cell structure.pptx
 Bacteria cell structure.pptx Bacteria cell structure.pptx
Bacteria cell structure.pptx
 
Probiotics and adhesion paper
Probiotics and adhesion paperProbiotics and adhesion paper
Probiotics and adhesion paper
 

Recently uploaded

microwave assisted reaction. General introduction
microwave assisted reaction. General introductionmicrowave assisted reaction. General introduction
microwave assisted reaction. General introductionMaksud Ahmed
 
Accessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impactAccessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impactdawncurless
 
CARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptxCARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptxGaneshChakor2
 
Science 7 - LAND and SEA BREEZE and its Characteristics
Science 7 - LAND and SEA BREEZE and its CharacteristicsScience 7 - LAND and SEA BREEZE and its Characteristics
Science 7 - LAND and SEA BREEZE and its CharacteristicsKarinaGenton
 
Solving Puzzles Benefits Everyone (English).pptx
Solving Puzzles Benefits Everyone (English).pptxSolving Puzzles Benefits Everyone (English).pptx
Solving Puzzles Benefits Everyone (English).pptxOH TEIK BIN
 
_Math 4-Q4 Week 5.pptx Steps in Collecting Data
_Math 4-Q4 Week 5.pptx Steps in Collecting Data_Math 4-Q4 Week 5.pptx Steps in Collecting Data
_Math 4-Q4 Week 5.pptx Steps in Collecting DataJhengPantaleon
 
How to Make a Pirate ship Primary Education.pptx
How to Make a Pirate ship Primary Education.pptxHow to Make a Pirate ship Primary Education.pptx
How to Make a Pirate ship Primary Education.pptxmanuelaromero2013
 
Presiding Officer Training module 2024 lok sabha elections
Presiding Officer Training module 2024 lok sabha electionsPresiding Officer Training module 2024 lok sabha elections
Presiding Officer Training module 2024 lok sabha electionsanshu789521
 
Crayon Activity Handout For the Crayon A
Crayon Activity Handout For the Crayon ACrayon Activity Handout For the Crayon A
Crayon Activity Handout For the Crayon AUnboundStockton
 
Introduction to AI in Higher Education_draft.pptx
Introduction to AI in Higher Education_draft.pptxIntroduction to AI in Higher Education_draft.pptx
Introduction to AI in Higher Education_draft.pptxpboyjonauth
 
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...EduSkills OECD
 
Introduction to ArtificiaI Intelligence in Higher Education
Introduction to ArtificiaI Intelligence in Higher EducationIntroduction to ArtificiaI Intelligence in Higher Education
Introduction to ArtificiaI Intelligence in Higher Educationpboyjonauth
 
Micromeritics - Fundamental and Derived Properties of Powders
Micromeritics - Fundamental and Derived Properties of PowdersMicromeritics - Fundamental and Derived Properties of Powders
Micromeritics - Fundamental and Derived Properties of PowdersChitralekhaTherkar
 
Paris 2024 Olympic Geographies - an activity
Paris 2024 Olympic Geographies - an activityParis 2024 Olympic Geographies - an activity
Paris 2024 Olympic Geographies - an activityGeoBlogs
 
Sanyam Choudhary Chemistry practical.pdf
Sanyam Choudhary Chemistry practical.pdfSanyam Choudhary Chemistry practical.pdf
Sanyam Choudhary Chemistry practical.pdfsanyamsingh5019
 
MENTAL STATUS EXAMINATION format.docx
MENTAL     STATUS EXAMINATION format.docxMENTAL     STATUS EXAMINATION format.docx
MENTAL STATUS EXAMINATION format.docxPoojaSen20
 
Mastering the Unannounced Regulatory Inspection
Mastering the Unannounced Regulatory InspectionMastering the Unannounced Regulatory Inspection
Mastering the Unannounced Regulatory InspectionSafetyChain Software
 
Concept of Vouching. B.Com(Hons) /B.Compdf
Concept of Vouching. B.Com(Hons) /B.CompdfConcept of Vouching. B.Com(Hons) /B.Compdf
Concept of Vouching. B.Com(Hons) /B.CompdfUmakantAnnand
 

Recently uploaded (20)

microwave assisted reaction. General introduction
microwave assisted reaction. General introductionmicrowave assisted reaction. General introduction
microwave assisted reaction. General introduction
 
Accessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impactAccessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impact
 
CARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptxCARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptx
 
Science 7 - LAND and SEA BREEZE and its Characteristics
Science 7 - LAND and SEA BREEZE and its CharacteristicsScience 7 - LAND and SEA BREEZE and its Characteristics
Science 7 - LAND and SEA BREEZE and its Characteristics
 
Solving Puzzles Benefits Everyone (English).pptx
Solving Puzzles Benefits Everyone (English).pptxSolving Puzzles Benefits Everyone (English).pptx
Solving Puzzles Benefits Everyone (English).pptx
 
_Math 4-Q4 Week 5.pptx Steps in Collecting Data
_Math 4-Q4 Week 5.pptx Steps in Collecting Data_Math 4-Q4 Week 5.pptx Steps in Collecting Data
_Math 4-Q4 Week 5.pptx Steps in Collecting Data
 
How to Make a Pirate ship Primary Education.pptx
How to Make a Pirate ship Primary Education.pptxHow to Make a Pirate ship Primary Education.pptx
How to Make a Pirate ship Primary Education.pptx
 
TataKelola dan KamSiber Kecerdasan Buatan v022.pdf
TataKelola dan KamSiber Kecerdasan Buatan v022.pdfTataKelola dan KamSiber Kecerdasan Buatan v022.pdf
TataKelola dan KamSiber Kecerdasan Buatan v022.pdf
 
Staff of Color (SOC) Retention Efforts DDSD
Staff of Color (SOC) Retention Efforts DDSDStaff of Color (SOC) Retention Efforts DDSD
Staff of Color (SOC) Retention Efforts DDSD
 
Presiding Officer Training module 2024 lok sabha elections
Presiding Officer Training module 2024 lok sabha electionsPresiding Officer Training module 2024 lok sabha elections
Presiding Officer Training module 2024 lok sabha elections
 
Crayon Activity Handout For the Crayon A
Crayon Activity Handout For the Crayon ACrayon Activity Handout For the Crayon A
Crayon Activity Handout For the Crayon A
 
Introduction to AI in Higher Education_draft.pptx
Introduction to AI in Higher Education_draft.pptxIntroduction to AI in Higher Education_draft.pptx
Introduction to AI in Higher Education_draft.pptx
 
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
 
Introduction to ArtificiaI Intelligence in Higher Education
Introduction to ArtificiaI Intelligence in Higher EducationIntroduction to ArtificiaI Intelligence in Higher Education
Introduction to ArtificiaI Intelligence in Higher Education
 
Micromeritics - Fundamental and Derived Properties of Powders
Micromeritics - Fundamental and Derived Properties of PowdersMicromeritics - Fundamental and Derived Properties of Powders
Micromeritics - Fundamental and Derived Properties of Powders
 
Paris 2024 Olympic Geographies - an activity
Paris 2024 Olympic Geographies - an activityParis 2024 Olympic Geographies - an activity
Paris 2024 Olympic Geographies - an activity
 
Sanyam Choudhary Chemistry practical.pdf
Sanyam Choudhary Chemistry practical.pdfSanyam Choudhary Chemistry practical.pdf
Sanyam Choudhary Chemistry practical.pdf
 
MENTAL STATUS EXAMINATION format.docx
MENTAL     STATUS EXAMINATION format.docxMENTAL     STATUS EXAMINATION format.docx
MENTAL STATUS EXAMINATION format.docx
 
Mastering the Unannounced Regulatory Inspection
Mastering the Unannounced Regulatory InspectionMastering the Unannounced Regulatory Inspection
Mastering the Unannounced Regulatory Inspection
 
Concept of Vouching. B.Com(Hons) /B.Compdf
Concept of Vouching. B.Com(Hons) /B.CompdfConcept of Vouching. B.Com(Hons) /B.Compdf
Concept of Vouching. B.Com(Hons) /B.Compdf
 

nadira.pdf

  • 1. Review Chemical Biology Tools for Examining the Bacterial Cell Wall Ashley R. Brown,1 Rebecca A. Gordon,2,3 Stephen N. Hyland,1 M. Sloan Siegrist,2,3 and Catherine L. Grimes1,4,* 1Department of Chemistry and Biochemistry, University of Delaware, Newark, DE 19716, USA 2Department of Microbiology, University of Massachusetts, Amherst, MA 01003-9298, USA 3Molecular and Cellular Biology Graduate Program, University of Massachusetts, Amherst 01003-9298, USA 4Department of Biological Sciences, University of Delaware, Newark, DE 19716, USA *Correspondence: cgrimes@udel.edu https://doi.org/10.1016/j.chembiol.2020.07.024 SUMMARY Bacteria surround themselves with cell walls to maintain cell rigidity and protect against environmental in- sults. Here we review chemical and biochemical techniques employed to study bacterial cell wall biogenesis. Recent advances including the ability to isolate critical intermediates, metabolic approaches for probe incor- poration, and isotopic labeling techniques have provided critical insight into the biochemistry of cell walls. Fundamental manuscripts that have used these techniques to discover cell wall-interacting proteins, flip- pases, and cell wall stoichiometry are discussed in detail. The review highlights that these powerful methods and techniques have exciting potential to identify and characterize new targets for antibiotic development. INTRODUCTION The bacterial cell wall is arguably just as important to human health as it is to the survival of the bacterium. The complex poly- mers that comprise the cell wall provide bacteria with strength and a barrier to the outside world, allowing them to thrive in a multitude of environments, including the human body. Humans have taken advantage of natural product antibiotics, often pro- duced by bacteria themselves, to target bacterial polymers, yielding some of the most widely used antibiotics to date (Cho- pra and Roberts, 2001). In this review, we focus on understand- ing the bacterial cell wall in the context of the threat that antibiotic resistance poses to society. With the announcement that many major pharmaceutical companies will no longer fund research programs in this critical area (Hu, 2018), the onus falls on curious, determined academics to identify new targets for antibiotics and new opportunities to combat resistance. Here we highlight recently reported chemical and biochemical approaches to study bacterial cell wall biosynthesis and maintenance both in the laboratory and in the clinic, focusing on the use of antibiotics. In some cases, antibiotics are part of the toolkit that enables bio- logical insight. In other cases, the biological insights gleaned from applying the tools enable (or have the potential to enable) target discovery. Studies of the bacterial cell wall and the antibi- otics that corrupt it are iterative and promote both mechanistic insights and translational applications. THE TARGET AND ITS BASICS Scientists have used small molecules to study bacteria since the 19th century when Christian Gram treated cells with crystal violet and realized that bacteria could be canonically divided into two general classes: Gram-positive and Gram-negative (Gram, 1884). In the present day, sophisticated tools exist to visualize the cell wall and to dissect its composition and biosyn- thesis at the molecular level (Hsu et al., 2019; Kocaoglu and Carl- son, 2016; Radkov et al., 2018; Siegrist et al., 2015; Taguchi et al., 2019a). Here we will review a subset of the new methods, but first offer a brief introduction to cell walls, noting the many recent, detailed reviews on these structures (e.g., Radkov et al., 2018) and their biosynthesis and maintenance (Taguchi et al., 2019a). The bacterial cell envelope is a complex structure that pro- vides protection from the external environment, maintains cell shape, and provides resistance to chemical, physical, and bio- logical damage (Figure 1A; Vollmer et al., 2008). Nearly all bacte- rial envelopes have a peptidoglycan (PG) cell wall layer lying just outside the plasma membrane. PG biosynthesis is a highly conserved process in bacteria, starting with UDP-N-acetyl- glucosamine (UDP-GlcNAc) conversion into UDP-N-acetyl-mur- amic acid (UDP-MurNAc), as the first committed step in PG syn- thesis (Figure 1B). The subsequent steps involve the addition of amino acids (commonly L-Ala, D-g-Glu, L-Lys [Gram-positive] or meso-diaminopimelic acid [mDAP; Gram-negative and myco- bacteria], and D-Ala-D-Ala) to the UDP-MurNAc lactate moiety. Variation exists within the stem peptide depending on species. For example, Mycobacterium leprae can utilize Gly in place of L-Ala at the one position (Draper et al., 1987), many Gram-posi- tive bacteria and mycobacteria amidate the second position to post-biosynthetically generate D-g-Gln, and spirochetes include L-Orn at the third position (Schleifer and Kandler, 1972; Vollmer et al., 2008). These and additional variations are highlighted in Vollmer et al. (2008). Stem peptide ligation is followed by the transfer of this PG building block to phosphate polyprenyl to make Lipid I. GlcNAc is added to Lipid I to make Lipid II as the final PG precursor. Lipid II is subsequently flipped across the ll 1052 Cell Chemical Biology 27, August 20, 2020 ª 2020 Elsevier Ltd.
  • 2. membrane by the MurJ flippase (Ruiz, 2008; Sham et al., 2014) as the complete PG subunit. Class A penicillin-binding proteins (aPBPs), L,D-transpeptidases, and shape, elongation, division, and sporulation (SEDS) proteins in complex with class B PBPs (bPBPs) assemble the PG through two different enzymatic reac- tions, transglycosylation and transpeptidation (Cho et al., 2016; Emami et al., 2017; Meeske et al., 2016; Taguchi et al., 2019b). PG transglycosylases (TGs) link the sugar backbone of the PG subunit to the next unit through a b-1,4 glycosidic linkage (Vollmer et al., 2008). PG transpeptidases (TPs) most commonly connect the fourth amino acid of one peptide chain to the third amino acid (such as mDAP or L-Lys) of an adjacent strand yielding 3–4 cross-linked PG (Vollmer and Seligman, 2010). However, there are more variations that feature diverse connec- tivity—3-3, 2-4, and 1-3 linkages—and bridge lengths across species. While the biosynthesis of Lipid II is generally conserved across bacteria, the polymerization of the monomer marks the beginning of the diversification process (Egan et al., 2020; Typas et al., 2011). Depending on the activity and the protein-protein in- teractors of the TGs/TPs, a variety of shapes are formed (Do et al., 2020; Salama, 2020). For example, in H. Pylori, the TG and TP activity is spatially and temporally regulated along the cell wall, which greatly influences the shape. This diversity of Gram-positive and Gram-negative cells is further enhanced by the inclusion of lipids in the cell wall. In Gram-positive bacteria, the PG layer is significantly thicker than that of Gram-negative (Vollmer and Seligman, 2010) and features lipoteichoic and wall teichoic acids (WTAs) anchored to the cytoplasmic membrane and the MurNAc 6-OH, respec- tively. The WTA is assembled largely on the cytoplasmic face. Upon delivery to the extracellular surface, it is anchored to PG by the family of LytR-CpsA-Psr (LCP) enzymes (Kawai et al., 2011; Li et al., 2020; Schaefer et al., 2017). Teichoic acid (TA) el- ements are essential to the virulence of pathogenic bacteria. They permit bacterial cell adhesion to host cells, in addition to controlling cell wall remodeling by autolysins (Brown et al., 2013). Gram-negative bacteria have a thinner PG but have additional structural support and protection due to the asymmetric outer membrane consisting of phospholipids and lipopolysaccharides (LPS) (Rojas et al., 2018). LPS is made of three components: lipid A, an oligosaccharide core, and the O-antigen. Toward the end of the lipid A construction, the core oligosaccharide is added, making the lipid oligosaccharide intermediate. The O-antigen is synthesized separately and bound to lipooligosaccharide (LOS), to make LPS, before transport to the outer membrane (Simpson and Trent, 2019). LPS is not anchored to the PG but rather inserted into the outer membrane by complex cellular ma- chinery and hydrophobically adhered via Lipid A. There is a third class of bacteria that has elements of both Gram-positives and Gram-negatives; it is sometimes referred Figure 1. Bacterial Cell Wall Basics (A) Structural features of the cell wall of Gram-positives, Gram-negatives, and mycobacteria. PG, peptidoglycan; PM, plasma membrane; LPS, lipopolysac- charide; MM, mycomembrane; OM, outer membrane; AG, arabinogalactan. (B) Biosynthetic steps of peptidoglycan (PG): PG biosynthesis occurs in 14 unique biochemical steps, starting with UDP-GlcNAc conversion into UDP-MurNAc. UDP-MurNAc is transferred to the lipid carrier and, subsequently, glycosylated by MurG. This process culminates in Lipid II being flipped across the membrane by MurJ. Transglycosylases (TGs) and transpeptidases (TPs) incorporate the PG subunits into the growing PG polymer. Ultimately, peptidoglycan is altered through the addition of small and large molecules post-synthetically that are not represented here, such as wall teichoic acids (WTAs). ll Cell Chemical Biology 27, August 20, 2020 1053 Review
  • 3. to as Gram-indeterminate. M. tuberculosis and other members of the Corynebacterineae suborder are phylogenetically related to Gram-positive bacteria, but possess a waxy, relatively imper- meable outer membrane-like structure (Figure 1A; Alderwick et al., 2015). The core cell wall has a PG layer of intermediate thickness that is covalently bound to the branched arabinogalac- tan (AG). The AG layer acts as a scaffold for a covalently bound, inner leaflet of mycolic acids. The outer leaflet possesses more mycolic acids, along with intercalated glycolipids such as treha- lose monomycolates (TMMs) and trehalose dimycolate (TDM) (Dulberger et al., 2020). The three classes of bacteria discussed here contain some similarities in the composition of their cell wall (i.e., Lipid II; Figure 1B) and differences (i.e., lipid modifications). It has been proposed that some of the lipid modifications that help to differentiate Gram-positive, Gram-negative, and Gram- indeterminate bacteria could provide mechanisms to uniquely target specific bacteria with antibiotic therapy (Jackson et al., 2013; Kuhn, 2019). Bacterial cell walls have historically made an excellent target for antibiotics. There are multiple ways to kill a bacterial cell, inhibiting DNA replication and halting protein synthesis (Walsh, 2003). How- ever, the bacterial cell wall is an especially attractive target because it is unique to bacteria, meaning that human cells do not contain the biochemical machinery that is required to build and maintain it. In addition, the biosynthesis is largely conserved (Figure 1B), allowing the development of broad-spectrum antibi- otics. Unfortunately, bacteria have developed resistance to nearly every cell wall-targeting antibiotic, including the well-known b-lac- tams, such as penicillin and methicillin, and the antibiotics of ‘‘last resort,’’ the glycopeptides (Kahne et al., 2005). Modes of resis- tance can be conferred by, but not limited to, the expression of insensitive PBPs, production of b-lactamases, activity of efflux pumps, and alterations to the cell wall (Walsh, 2000). For a more extensive review of these and other modes of resistance, the au- thors refer the reader to Walsh (2000). In recent years, there has been an emergence of multi-drug-resistant bacteria, particularly in the ESKAPE pathogens (Enterococcus faecium, Staphylo- coccus aureus, Klebsiella pneumoniae, Acinetobacter baumanii, Pseudomonas aeruginosa, and Enterobacter species) (Boucher et al., 2009; Santajit and Indrawattana, 2016). These bacteria, together with tuberculosis (Bloom and Murray, 1992; Porter and McAdam, 1994) and gonorrhea (Unemo and Shafer, 2014), pose a viable threat with a lack of new antimicrobial compounds (CDC, 2017). Within 1 year after the first clinical use of a natural or synthetic antibiotic, resistance can develop (Walsh, 2003). Therefore, there is a continuous demand for the development of novel antibiotics via rational design, high-throughput screening ef- forts, and medicinal chemistry campaigns. In order to identify new bactericidal, cell wall-acting com- pounds, peptidoglycan biosynthetic pathways (Figure 1) that are less understood and not currently targeted by mainstream antibiotics must be interrogated with appropriate tools. Fluores- cent probes have permitted the close study of bacterial struc- tures, and more recently probes such as fluorescent D-amino acids (FDAAs) have enabled methods to screen for novel antibi- otics and to identify their mode of action (Culp et al., 2020). There are many excellent reviews that highlight direct imaging of bac- terial cell walls to study the structure, biosynthesis, and dy- namics of this polymer (Kocaoglu and Carlson, 2016; Radkov et al., 2018; Siegrist et al., 2015; Taguchi et al., 2019a). Here we focus on recent work that merges chemical and biochemical methods including immunoblotting, photo-crosslinking, spec- troscopy, and radiolabeling to probe PG biosynthetic pathways and composition that can help inform the development of novel antimicrobial therapeutics. USING ANTIBIOTICS TO ISOLATE CELL WALL INTERMEDIATES In order to study bacterial cell wall biosynthetic enzymes, one must have access to PG precursor substrates. A common strat- egy is to use bacterial cell wall inhibitors such as vancomycin and moenomycin to cause a buildup of intermediates. Such a tactic was used by Strominger and Park in the 1950s to identify the UDP-MurNAc intermediate, known as Park’s nucleotide, using penicillin (Figure 1B) (Park, 1952; Strominger et al., 1959). The lipid-linked precursor, Lipid II, was long sought in PG biosyn- thesis (Figure 1B) as it is difficult to synthesize, and its scarcity made in vitro studies of PG polymerization nearly impossible (Lazar and Walker, 2002; Ye et al., 2001). Lipid II is the result of the b-1,4 glycosylation of Lipid I by MurG with the monosaccha- ride unit, GlcNAc (Figure 1B). This event produces a disaccha- ride pentapeptide containing an undecaprenyl pyrophosphate group. A major feat from the Walker Laboratory in 1999 showed that Lipid I derivatives could be transformed with the glycosyl- transferase MurG to form native Lipid II as well as a variety of de- rivatives (Ha et al., 1999; Men et al., 1998). This in vitro enzymatic conversion provided the first reliable method to produce work- able quantities of Lipid II. Access to Lipid II, in turn, enabled the mechanism of various glycopeptide antibiotics to be studied. For example, vancomycin, as well as other glycopeptides such as dalbavancin (Leimkuhler et al., 2005), has been shown to inhibit TGs directly (Chen et al., 2003). In 2017, a method to accumulate and isolate Lipid II directly from cells was developed (Figure 2A) (Qiao et al., 2017). Qiao, Kahne, and Walker developed a two-step extraction protocol of Lipid II. Using a similar strategy as Strominger in 1958 and armed with the knowledge that PG TGs are responsible for the last steps of PG biosynthesis (Figure 1B), antibiotics that target these later steps were used to isolate native Lipid II (Figure 2A). Cells treated with moenomycin or vancomycin, which inhibit late-stage biosynthetic enzymes in the outer leaflet of the plasma membrane (Chen et al., 2003; Kahne et al., 2005; Taylor et al., 2006), accumulate a significant amount of Lipid II (Figure 2A). However, when cells were treated with a sublethal dose of fosfomycin, an antibiotic that inhibits the first committed step in PG biosynthesis (Falagas et al., 2016), Lipid II was undetectable (Qiao et al., 2017). The ability to alter the amounts of PG biosynthetic intermediates is a powerful tool for studying both upstream and downstream effects of peptidoglycan synthesis, as evidenced by experi- ments that have used the Lipid II isolation method to screen for novel antibiotics, assess the function of critical PG biosyn- thetic enzymes, and identify new enzymes involved in PG biosynthesis (as discussed below). To interrogate the production of Lipid II in a cellular context, the Walker and Kahne labs have developed an easily accessible technique to label Lipid II from bacterial cultures (Figure 2B) ll 1054 Cell Chemical Biology 27, August 20, 2020 Review
  • 4. (Qiao et al., 2014). The method involves detection of Lipid II from growing bacterial cells by treating the intermediate with peni- cillin-binding protein 4 (PBP4) to allow for the installation of a biotin tag. This low molecular weight PBP isolated from S. aureus is a promiscuous transpeptidase that, in addition to catalyzing bonds between adjacent muropeptides, can ex- change the terminal D-Ala of the muropeptide with various D- amino acids. In vitro, this technique can switch the terminal D- Ala for biotinylated D-Lys (BDL); biotinylated Lipid II is then detectable via western blot (Figure 2B). This technique provides Figure 2. Biochemical and Chemical Biology Techniques for Intergorating Bacterial Peptidolgycans (A) Two-step extraction method for isolating Lipid II from bacterial cultures. (B) PBP4 transpeptidase mediated terminal D-Ala exchange with unnatural amino acids. Depicted is the incorporation of biotin-D-Lys (BDL) to the stem peptide of Lipid II. Lipid II consists of a diphosphate (P) disaccharide backbone, GlcNAc (G), and MurNAc (M), with a pentapeptide chain: alanine (A), glutamate (E), and lysine (K). (C) Substituted cysteine accessibility method (SCAM) utilizes single cysteine mutations (orange circle) in a protein of interest in conjunction with two cysteine reactive reagents: MTSES and NEM. MTSES cannot penetrate the membrane and will only react with periplasmic cysteines. NEM can penetrate the membrane and react with both periplasmic and cytoplasmic cysteine residues. Using this method with a cysteine mutant library, it can provide topological information of a protein, in this case a transmembrane protein. (D) MurJ-pBPA photocrosslinking assay for detection of Lipid II/MurJ adducts. MurJ encoded with single pBPA mutations undergoes crosslinking with Lipid II upon UV activation. After SDS-PAGE and electroblotting, crosslinked Lipid II is biotinylated in-gel via a BDL exchange reaction and then detected via blotting with streptavidin-HRP. (E) Subsequent lysis and a click reaction to attach a fluorophore allow for analysis of mycolate-protein interactions via metabolic incorporation of a bifunctional TMM analog. N-x-AlkTMM-C15 is metabolically incorporated into the mycobacterial mycomembrane (MM). Covalent crosslinks with MM-associated proteins are induced by UV activation. Subsequent lysis and a click reaction to attach a fluorophore allow for analysis by a variety of techniques. ll Cell Chemical Biology 27, August 20, 2020 1055 Review
  • 5. researchers with a streamlined process to obtain functionalized Lipid II analogs that can be selectively detected, thus generating a valuable assay for studying PG biosynthetic machinery and screening antibiotics (Cochrane and Lohans, 2020). In sum, there are at least three methods to access Lipid II and its derivatives that have been reported: in vitro biochemical methods using MurG, isolation of Lipid II from bacterial cultures, and Lipid II biotinylation. Collectively, these methods have been applied to diverse bacterial species: S. aureus, Bacillus subtilis, Escherichia coli (Qiao et al., 2017), and M. smegmatis (Garcı́a- Heredia et al., 2018) have all been successfully used for the Lipid II isolation and/or visualization, despite variation in PG penta- peptide chains (Qiao et al., 2017). Access to these substrates will be important to fully understand the structural diversity pre- sent in bacterial cell walls and to characterize pools of PG inter- mediates during growth. These methods have been useful in studying the mechanism of action of antibiotics, as mentioned above with glycopeptides and more recently with lysobactin (Lee et al., 2016). Research groups have also used them to study the mechanisms of enzymes involved in PG biosynthesis. For example, multiple labs have used the tools to biochemically characterize the TG activity of the SEDS proteins, which have been proposed as new potential targets for antibiotics (Cho et al., 2016; Emami et al., 2017; Meeske et al., 2016; Rohs et al., 2018; Sjodt et al., 2018; Taguchi et al., 2019b). The methods of isolation and detection (Figures 2A and 2B) have been used in combination in vitro to study the order of addition of WTA precursors to PG intermediates (Figure 3B; this work Figure 3. Critical Details for Understanding Peptidoglycan Isotopic Probe Incoroporation (A) Cell wall depictions of S. aureus strains utilized to determine characteristic peaks of PG and TA. (B) Assessing the ability of un-cross-linked and cross-linked PG as a substrate for WTA transfer by LcpB. Un-cross-linked PG oligomers featuring a pentaglycine chain are bioenzymatically prepared with a SgtB mutant. The oligomers are then either modified with WTA via LcpB (top) or cross-linked with PBP4 (bottom). Subsequently, these molecules are either subjected to transpeptidation with PBP4 or a ligation reaction with LcpB. ll 1056 Cell Chemical Biology 27, August 20, 2020 Review
  • 6. will be discussed in detail in Radiolabeling). Finally, these methods have been used to identify and study the mechanism of Lipid II transport (Rubino et al., 2020) as discussed in detail in the next section. BIOCHEMICAL TOOLS TO INVESTIGATE THE PG ENZYMES For a long time, it was not understood how Lipid II translocates across the lipid bilayer into the periplasmic space for incorpora- tion into the PG meshwork. Multiple proteins were evoked to perform this function, including the SEDS protein FtsW, which was later shown to be a transglycosylase (Egan et al., 2020; Ruiz, 2016; Young, 2014). MurJ was discovered by Ruiz and col- leagues to be the elusive flippase (Ruiz, 2008; Sham et al., 2014). In a series of elegant experiments utilizing substituted cysteine accessibility method (SCAM) (Figure 2C) and the protein toxin ColM (Touzé et al., 2012), Ruiz and collaborators were able to create an assay that provided context-dependent monitoring of Lipid II movement. Using an extensive single Cys mutant MurJ library, the authors mapped the topology of MurJ, which was then used in parallel with protein modeling to predict possible dynamic conformations (Butler et al., 2013). Building on these findings, Sham et al. used an in vivo assay with the addi- tion of sulfonating reagents, ColM, or both in either wild-type MurJ or the reversible MurJ mutant to formalize MurJ as the Lipid II flippase (Sham et al., 2014). With the knowledge that MurJ was the flippase, the biochem- ical details of its activity were investigated using the Lipid II as- says discussed above. MurJ proved challenging to study because, as a flippase, it does not structurally alter the Lipid II precursor when it transports. However, by using the Lipid II detection method (Figure 2B), it was possible to monitor changes in Lipid II’s movement. In 2018, Rubino et al. showed that treatment with a protonophore to disperse the proton- motive force caused Lipid II accumulation in E. coli, suggesting that the transportation of Lipid II is coupled to an electrochemical gradient. The effect of the protonophore mimicked controls known to disrupt MurJ activity. Furthermore, they characterized the conformation of MurJ when the membrane potential is dissi- pated by probing individual cysteine residues in MurJ (Rubino et al., 2018). To determine what residues of MurJ are involved in transport, photocrosslinking experiments were used to tether interacting partners. Photocrosslinking has become an invaluable method to determine protein-substrate interactions (Lancia et al., 2014; Parker and Pratt, 2020; Wu and Kohler, 2019). Unnatural amino acid incorporation can be used to install a photoactivatable p-benzoyl-L-phenylalanine (pBPA) in the protein of interest—in this case, MurJ. Excitation of pBPA using 350–365 nm light leads to a reactive diradical that forms a C-H bond to any vicinal func- tional group within a 3.1 Å reactivity radius (Lancia et al., 2014). In 2012, Okuda et al. utilized this method to determine specific sites where LPS interacts with the LPS transport (Lpt) machinery (Okuda et al., 2012). Briefly, they incorporated pBPA residues in several Lpt transport proteins. After UV activation, photo- crosslinked adducts were identified using immunoblotting with LPS-specific antibodies. This method allowed them to take chemical snapshots of LPS transport and determine that shut- tling of LPS across the periplasm is accomplished through cyto- plasmic ATP hydrolysis. The Ruiz group, in collaboration with the Kahne lab, applied this methodology to probe MurJ flippase ac- tivity (Figure 2D) (Rubino et al., 2020). They hypothesized that pBPA incorporated into MurJ would prompt crosslinking to its natural substrate Lipid II. However, the lack of antibodies for Lipid II made it necessary to implement a method that allowed for the detection of the crosslinked adduct. PBP4 was used to incorporate a BDL after photocrosslinking Lipid II to MurJ- pBPA mutants to allow for detection (Figure 2D). They applied this protocol to investigate the role of three essential arginine residues, located in the central cavity of MurJ, that were previ- ously proposed to be key in recognizing the pyrophosphate of Lipid II (Kuk et al., 2019). Using single and multiple Arg/Ala mu- tants, they observed similar levels of crosslinking compared to wild-type MurJ. However, these mutants displayed impaired ability to flip Lipid II. This methodology permitted the observation of the intermediate transport steps in living cells and provided direct, biochemical evidence that the conserved arginine resi- dues control Lipid II movement through MurJ. Thus, through the combination of Lipid II chemical probes, genetic tools, and biochemical conversions, the function of MurJ has been identi- fied and the biochemical mechanisms of Lipid II transport are rapidly being unveiled. This also highlights MurJ as an exciting target for antibiotic development. The ability to track Lipid II in cellular biochemistry assays was critical because it yielded detailed biochemical information (i.e., protein residues, mem- brane potential) that would not have been possible with other methods. METABOLIC INCORPORATION OF PHOTOCROSSLINKING SUGARS The proteins that bind the cell wall and its associated glycocon- jugates are also potential antibiotic targets. In contrast to pro- tein-mediated interactions, glycan recognition events are often weak and short-lived. Additionally, glycans are not directly genetically encoded and their biosynthesis is complex, so it is challenging to use standard genetic engineering methods to tag them. The incorporation of functionalized metabolites has al- lowed a way to bypass these challenges (Campbell et al., 2007). Several groups have introduced unnatural sugars containing photoactivatable crosslinkers to the cell by hijacking carbohy- drate metabolic pathways and capitalizing on enzyme promiscu- ity (Tanaka and Kohler, 2008; Yu et al., 2012). The mycomem- brane (Figure 1A, ‘‘MM’’) is attached to the mycobacterial cell wall via AG and is a barrier to environmental, immune, and anti- biotic insults. However, its protein composition has eluded classic biochemical techniques for a long time, in part because of the difficulty of cleanly separating the covalently bound myco- membrane from other layers of the complex mycobacterial enve- lope. Kavunja et al. recently developed the first photocrosslink- ing probes for the mycomembrane to analyze mycolate-protein interactions in vivo (Figure 2E) (Kavunja et al., 2020). They syn- thesized a TMM analog that specifically incorporates into the TDM portion of the mycomembrane via previously reported conserved, substrate-promiscuous Ag85 mycoloyltransferases (Fiolek et al., 2019). This analog contains a bifunctional linker bearing a photoactivatable diazirine group and a clickable alkyne ll Cell Chemical Biology 27, August 20, 2020 1057 Review
  • 7. handle. After metabolic incorporation into the cell surface, myco- bacteria were irradiated with UV light. The diazirine cross-linked with neighboring proteins that were then enriched following click ligation to an azide-fluorophore-biotin. This method allowed for the identification of both known and previously undetectable my- comembrane-resident proteins, as well as tracking them by in- gel fluorescence. Similar techniques are likely to be useful in future studies to interrogate different layers of the cell envelope, including PG—especially as DeMeester and coworkers have shown that PG precursors containing the diazirine cross-linkers at the C2 position of MurNAc are accepted by the PG biosyn- thetic enzymes (DeMeester et al., 2018). Previously, Sarkar et al. exploited the MurF ligation process to insert a D-Ala- that was biofunctionalized with an alkyne and photocrosslinking han- dles into Lipid II to mine the protein-interacting partners (Sarkar et al., 2016). Targeted acquisition of cell envelope interactomes may reveal new potential targets for antibiotic therapies (Kavunja et al., 2020). SPECTROSCOPIC METHODS TO STUDY THE CELL WALL The stable isotopes 13 C and 15 N are effective as probes to eluci- date structural features of the bacterial cell wall at the molecular level using spectroscopic methods (Kim et al., 2014, 2015; Yang et al., 2017). In this strategy, the macromolecules of biological in- terest are unperturbed and uniform enrichment enhances signal output of the NMR spectrum in a non-destructive manner (Ny- gaard et al., 2015). Romaniuk and Cegelski utilized 13 C and 15 N cross-polarization magic angle spinning (CP/MAS) solid- state NMR (ssNMR) to characterize the composition of uniformly labeled PG and WTA in S. aureus (Romaniuk and Cegelski, 2018). This technique affords an alternative to solution-based analytical methods that do not permit full characterization of the highly insoluble material. Using this approach, spectra of pu- rified PG and WTA isolates from wild-type and DtarO (a mutant that is unable to synthesize WTA), respectively, were first used to identify and quantify characteristic carbon peaks from each component (Figure 3A). Since the sum of the two spectra repro- duced the peak intensities of the intact cell wall sample, they were able to quickly determine the composition ratio of TA to PG. The masses for TA and PG calculated by this method were consistent with those determined by phosphate analysis, a more traditional but labor-intensive method. Subsequently, the relative composition ratios of the two components in both the stationary and the exponential phases were determined. In the stationary phase, PG thickness increased while WTA decreased. This finding was confirmed with selective labeling using either D-[15 N]Ala (WTA) or [15 N]Gly (PG) in 15 N CP/MAS. With these baselines in hand, they were able to validate that this analytical method is amenable to determining the composi- tion levels of TA and PG of cell walls with the antibiotic tunicamy- cin, which inhibits WTA growth. Cegelski et al. have used similar strategies with ssNMR to establish that vancomycin primarily targets transglycosylation over transpeptidation using uniformly 13 C- and 15 N-labeled amino acids in S. aureus (Cegelski et al., 2002). Changes to the D-alanine-pentaglycyl bridge-links were unperturbed in the presence of vancomycin, suggesting that transpeptidation is unaffected. Instead, this supports the idea that vancomycin blocks transglycosylation and impedes translo- cation of Lipid II into the periplasm. Rotational echo double reso- nance (REDOR) and CP/MAS ssNMR have also been used to discern perturbations to PG and WTA of S. aureus treated with the cyclic decapeptide amphomycin, a drug that is effective against superbugs such as multi-drug-resistant S. aureus and vancomycin-resistant enterococci by targeting bactoprenol- phosphate. Using REDOR, Singh et al. observed 15 N shifts in bridge-links between Gly and L-Lys and the free side chain amine of lysine, indicating that the compound induced PG thinning, the accumulation of Park’s nucleotide, and a decrease in alanylation of WTA (Singh et al., 2016). These data suggested that ampho- mycin acts on the cell wall prior to transglycosylation. Overall, CP/MAS and REDOR analysis of the bacterial cell wall permits rapid assessment of whole-cell composition and can be applied to monitor and determine the compositional perturbations caused by antibiotics as showcased in the studies above. In another example, Calabretta et al. employed 13 C radiola- beled lipid-linked arabinofuranose donors to study Gram-inde- terminate bacteria. These probes were biosynthetically incorpo- rated into the arabinan layer of Corynebacterium glutamicum and M. smegmatis strains unable to produce this polymer due to genetic mutation or treatment with benzothiazinone antibi- otics (Calabretta et al., 2019). This procedure circumvents the requirement of metabolic processing of cell wall probes prior to incorporation into the cell wall. Analysis of the soluble arabinan isolated from these models retained characteristic peaks with 2D NMR experiments (1 H-13 C HSQC, 1 H-13 C HMBC, and 1 H-13 C HSQC-TOCSY). Though this study did not use intact cells to gain structural details, it yielded a powerful tool that can be applied to examine structural changes in the lesser-understood mycobacterial cell envelope. Antibiotics are critically needed against mycobacteria to combat public health threats such as tuberculosis, which the Centers for Disease Control (CDC) has recently identified as a top AMR threat (Bloom and Murray, 1992; CDC, 2017). RADIOLABELING Although NMR methods are useful in determining composition of the cell wall, they do not inform on its biosynthesis. This can be achieved using radioisotopes, as bacterial incorporation of extremely sensitive ionizing radionuclides permits submicro- molar (high fM to pM) detection of cell wall products and inter- mediates of biosynthetic pathways in vivo. This sensitivity is also useful in translational applications for disease diagnosis. Here we discuss the use of 14 C and 11 C radionuclides in bac- terial cell surface structures to discern mechanistic details of PG synthesis and structure (long-lived) in addition to usage as a diagnostic imaging tool of bacterial infection in vivo (short-lived). Long-Lived Radioisotopes Elucidating some of the lesser-understood biosynthetic pro- cesses has the potential to provide new targets for antimicro- bials. TAs are important for the structure and virulence of Gram-positive bacteria (Brown et al., 2013), but far less is known about their synthesis compared to that of PG. Recently, Schaefer et al. investigated the order and process by which WTA are ll 1058 Cell Chemical Biology 27, August 20, 2020 Review
  • 8. appended to PG. This is useful, as methicillin-resistant S. aureus (MRSA) is resensitized to b-lactams when WTA biosynthesis is inhibited (Campbell et al., 2011; Farha et al., 2013). They found that WTA is ligated to un-crosslinked PG oligomers and that liga- tion preference is due to steric restrictions of the LcpB WTA ligase binding site (Schaefer et al., 2018). With isolated Lipid II from S. aureus, featuring a pentaglycine chain for downstream cross-linking, the group bioenzymatically prepared un-cross- linked PG fragments (2–10 carbohydrates in length) with a trans- glycosylase mutant (SgtBY181D ) that releases premature PG olig- omers (Figure 3B). These fragments were then subjected to transpeptidation with PBP4 followed by treatment with LcpB (WTA ligase) and 14 C radiolabeled LIIA WTA , a truncated radiola- beled WTA precursor, or just the latter to assess substrate pref- erence. Detection by polyacrylamide gel electrophoresis (PAGE) autoradiography showed that the cross-linked substrate is un- able to be ligated to WTA, whereas the un-cross-linked PG frag- ments serve as an acceptable substrate for the transfer of WTA by LcpB. Additional experiments evaluating the structural re- quirements for recognition co-crystalized the homolog TagT from B. subtilis with either LIA WTA or LIIA WTA , which indicated a putative binding groove narrow in size. The steric restrictions of this site prompted Schaefer and coworkers to explore if the stem peptide was a necessary feature of recognition by employ- ing chitin, deacetylated chitin, and cellulose-based oligosaccha- rides as probes of LcpB and TagT. Results indicated that the stem peptide is not a necessary structural feature for recognition of the transfer substrate; however, acetylation of the C2 position of MurNAc is required. These radioisotopic probes enabled Schaefer and coworkers to visually determine if PG intermedi- ates were modified by WTA ligase in this enzymatic assay and suggest the ligation order of WTA precursors to peptidoglycan intermediates. With this fundamental information, new antibiotic targets can be established as a means to stymie the ligation pro- cess and ultimately disrupt the integrity of the cell wall of Gram- positive bacteria. Short-Lived Radioisotopes Positron emission tomography (PET) radioisotopes such as 11 C, 13 N, 15 O, and 18 F are extremely sensitive tools that permit whole-body imaging of physiological processes in deep-lying tissues and organs with low nanogram quantities of probe. Un- like stable isotopes, the radiation produced upon decay is detectable outside of the human body. [18 F]-Fluorine deoxyglu- cose (FDG) is a clinical PET probe frequently used to image cells with higher energy requirements, such as oncogenic cells and sites of inflammation (Zhuang and Alavi, 2002, Seminars in Nu- clear Medicine, conference). [18 F]FDG accumulates because it cannot be further metabolized after sequestration into the cell. Since this probe is not bacteria-specific, improved imaging technologies are required to detect active bacterial infections in vivo (Ordonez and Jain, 2018, Seminars in Nuclear Medicine, conference). To address this issue, Rosenberg, Ohliger, and Wilson have developed clinically relevant 11 C radiotracers that can distinguish between sterile inflammation and active infec- tion by metabolically incorporating into both clinically relevant pathogenic Gram-positive and Gram-negative bacteria (Neu- mann et al., 2017; Parker et al., 2020; Stewart et al., 2020), uti- lizing D-amino acid probe incorporation (de Pedro et al., 1997). High incorporation of D-[14 C]Met in both S. aureus and E. coli in the stem peptide of PG led to the development of a 11 C-syn- thetic probe from a D-homocysteinethiolactone precursor (Figure 4). Earlier experimentation had shown that exogenous D-Met is amply integrated into the PG polymer of stationary bac- teria (Caparrós et al., 1992; Lam et al., 2009). Moreover, a robust synthetic strategy and efficient labeling probe for D-Ala and D-Glu were not readily available. Using the L-[11 C]-Met tracer for imaging in mice modeling myositis, they were able to detect and differentiate between bacterial infection and sterile inflam- mation, modeled by heat-killed bacteria, in the deltoid muscle with great sensitivity. These probes are likely to be especially useful for regions of the body that are normally sterile such as the musculoskeletal, biliary, and nervous systems, as there will not be significant background from the commensals. The syn- thetic efficiency of the probe was later improved for use in the clinical setting by using automated synthesis (Neumann et al., 2017; Parker et al., 2020; Stewart et al., 2020). Most recently, Parker et al. expanded the tracers to include amino acids canonically incorporated into the stem peptide, D-[3-11 C]-Ala and D-[3-11 C]-Ala-D-Ala probes (Figure 4). In vivo studies utilized the D-Ala probe, since D-Ala showed anywhere from 2-3 times greater uptake in E. coli and S. aureus over the dipeptide. More- over, it was determined that the D-Ala probe generated high levels of incorporation with the previous panel of bacteria tested in the 2019 paper and that these probes were highly selective for bacterial cells over mammalian cells. Infectious imaging chal- lenge experiments also confirmed the ability of the D-Ala probe to distinguish between live bacterial infection and sterile Figure 4. Radiolabels in Living Bacteria’s Peptioglycans Incorporation of radiolabeled probes, D-[methyl-11 C]-Met or D-[3-11 C]-Ala, post-biosynthetically via transpeptidases in clinically relevant bacteria to assess use for in vivo labeling of active bacterial infections. ll Cell Chemical Biology 27, August 20, 2020 1059 Review
  • 9. inflammation in mouse models. The tracer was able to monitor cell wall-acting antibiotic (ampicillin) treatment of an E. coli bac- terial infection, image infection in the intervertebral space, and proficiently screen for pneumonia in mice. Thus, this class of radiolabeled probes yielded an indispensable bacteria-specific imaging strategy for pathogenic bacteria in difficult-to-access physiological niches in clinical settings compared to currently approved methods. CONCLUSION Biochemists have been using small molecules to study bacterial cell walls since the 1800s. As improved probes have become available, the ability to study the biochemical phenomena has exploded. Here we have highlighted examples of chemical probes that were used to reveal fundamental cell wall biochem- istry or, in some cases, to identify bacterial pathology in a living host. There is a rich intersection between chemical biology and cell walls. It is this intersection that will be powerful in tackling the problem of antibiotic resistance, as a means to identify new targets and to characterize the mechanism of action for promising compounds. We highlighted the recent advances in the field such as monitoring PG precursors to elucidate MurJ, TG, and TP activity; establishing new protein targets for anti- biotic development through photocrosslinking techniques; applying NMR spectroscopy to capture native cell wall compo- sition; and demonstrating a highly sensitive method of detection of cellular processes using radiolabels in vitro and in vivo. In conjunction with current visualization and imaging techniques, chemical, biochemical, and hybrid methods have the potential to provide both mechanistic and structural information for valu- able targets. ACKNOWLEDGMENTS This project was supported by a grant from the Glycoscience Common Fund (U01 CA221230) and the Delaware COBRE program, as well as a grant from the National Institute of General Medical Sciences-NIGMS (P20GM104316). C.L.G. is a Pew Biomedical Scholar, Sloan Scholar, and Camille Dreyfus Scholar and thanks the Pew Foundation, the Sloan Foundation for Science Advancement, and the Dreyfus Foundation for support. M.S.S. thanks NIH R21 AI144748 and NIH DP2 AI138238 for financial support. S.H. thanks the NIH Chemistry-Biology Interface Program (GM133395) and the University of Delaware Dissertation Fellowship for funding support. DECLARATION OF INTERESTS C.L.G. and M.S.S. have pending patents and patents relevant to this work. REFERENCES Alderwick, L.J., Harrison, J., Lloyd, G.S., and Birch, H.L. (2015). The mycobac- terial cell wall—peptidoglycan and arabinogalactan. Cold Spring Harb. Per- spect. Med. 5, a021113. Bloom, B.R., and Murray, C.J.L. (1992). Tuberculosis: commentary on a ree- mergent killer. Science 257, 1055–1064. Boucher, H.W., Talbot, G.H., Bradley, J.S., Edwards, J.E., Gilbert, D., Rice, L.B., Scheld, M., Spellberg, B., and Bartlett, J. (2009). Bad bugs, no drugs: no ESKAPE! An update from the Infectious Diseases Society of America. Clin. Infect. Dis. 48, 1–12. Brown, S., Santa Maria, J.P., Jr., and Walker, S. (2013). Wall teichoic acids of gram-positive bacteria. Annu. Rev. Microbiol. 67, 313–336. Butler, E.K., Davis, R.M., Bari, V., Nicholson, P.A., and Ruiz, N. (2013). Struc- ture-function analysis of MurJ reveals a solvent-exposed cavity containing res- idues essential for peptidoglycan biogenesis in Escherichia coli. J. Bacteriol. 195, 4639–4649. Calabretta, P.J., Hodges, H.L., Kraft, M.B., Marando, V.M., and Kiessling, L.L. (2019). Bacterial cell wall modification with a glycolipid substrate. J. Am. Chem. Soc. 141, 9262–9272. Campbell, C.T., Sampathkumar, S.-G., and Yarema, K.J. (2007). Metabolic oligosaccharide engineering: perspectives, applications, and future direc- tions. Mol. Biosyst. 3, 187–194. Campbell, J., Singh, A.K., Santa Maria, J.P., Jr., Kim, Y., Brown, S., Swoboda, J.G., Mylonakis, E., Wilkinson, B.J., and Walker, S. (2011). Synthetic lethal compound combinations reveal a fundamental connection between wall tei- choic acid and peptidoglycan biosyntheses in Staphylococcus aureus. ACS Chem. Biol. 6, 106–116. Caparrós, M., Pisabarro, A.G., and de Pedro, M.A. (1992). Effect of D-amino acids on structure and synthesis of peptidoglycan in Escherichia coli. J. Bacteriol. 174, 5549–5559. CDC (2017). Drug-resistant TB (Centers for Disease Control and Prevention). https://www.cdc.gov/tb/topic/drtb/default.htm. Cegelski, L., Kim, S.J., Hing, A.W., Studelska, D.R., O’Connor, R.D., Mehta, A.K., and Schaefer, J. (2002). Rotational-echo double resonance characteriza- tion of the effects of vancomycin on cell wall synthesis in Staphylococcus aureus. Biochemistry 41, 13053–13058. Chen, L., Walker, D., Sun, B., Hu, Y., Walker, S., and Kahne, D. (2003). Vanco- mycin analogues active against vanA-resistant strains inhibit bacterial trans- glycosylase without binding substrate. Proc. Natl. Acad. Sci. USA 100, 5658–5663. Cho, H., Wivagg, C.N., Kapoor, M., Barry, Z., Rohs, P.D.A., Suh, H., Marto, J.A., Garner, E.C., and Bernhardt, T.G. (2016). Bacterial cell wall biogenesis is mediated by SEDS and PBP polymerase families functioning semi-autono- mously. Nat. Microbiol. 1, 16172. Chopra, I., and Roberts, M. (2001). Tetracycline antibiotics: mode of action, applications, molecular biology, and epidemiology of bacterial resistance. Mi- crobiol. Mol. Biol. Rev. 65, 232–260. Cochrane, S.A., and Lohans, C.T. (2020). Breaking down the cell wall: strate- gies for antibiotic discovery targeting bacterial transpeptidases. Eur. J. Med. Chem. 194, 112262. Culp, E.J., Waglechner, N., Wang, W., Fiebig-Comyn, A.A., Hsu, Y.-P., Koteva, K., Sychantha, D., Coombes, B.K., Van Nieuwenhze, M.S., Brun, Y.V., and Wright, G.D. (2020). Evolution-guided discovery of antibiotics that inhibit pepti- doglycan remodelling. Nature 578, 582–587. de Pedro, M.A., Quintela, J.C., Höltje, J.V., and Schwarz, H. (1997). Murein segregation in Escherichia coli. J. Bacteriol. 179, 2823–2834. DeMeester, K.E., Liang, H., Jensen, M.R., Jones, Z.S., D’Ambrosio, E.A., Scinto, S.L., Zhou, J., and Grimes, C.L. (2018). Synthesis of functionalized N-acetyl muramic acids to probe bacterial cell wall recycling and biosynthesis. J. Am. Chem. Soc. 140, 9458–9465. Do, T., Page, J.E., and Walker, S. (2020). Uncovering the activities, biological roles, and regulation of bacterial cell wall hydrolases and tailoring enzymes. J. Biol. Chem. 295, 3347–3361. Draper, P., Kandler, O., and Darbre, A. (1987). Peptidoglycan and arabinoga- lactan of Mycobacterium leprae. J. Gen. Microbiol. 133, 1187–1194. Dulberger, C.L., Rubin, E.J., and Boutte, C.C. (2020). The mycobacterial cell envelope - a moving target. Nat. Rev. Microbiol. 18, 47–59. Egan, A.J.F., Errington, J., and Vollmer, W. (2020). Regulation of peptidoglycan synthesis and remodelling. Nat. Rev. Microbiol. 18, 446–460. Emami, K., Guyet, A., Kawai, Y., Devi, J., Wu, L.J., Allenby, N., Daniel, R.A., and Errington, J. (2017). RodA as the missing glycosyltransferase in Bacillus subtilis and antibiotic discovery for the peptidoglycan polymerase pathway. Nat. Microbiol. 2, 16253. Falagas, M.E., Vouloumanou, E.K., Samonis, G., and Vardakas, K.Z. (2016). Fosfomycin. Clin. Microbiol. Rev. 29, 321–347. ll 1060 Cell Chemical Biology 27, August 20, 2020 Review
  • 10. Farha, M.A., Leung, A., Sewell, E.W., D’Elia, M.A., Allison, S.E., Ejim, L., Per- eira, P.M., Pinho, M.G., Wright, G.D., and Brown, E.D. (2013). Inhibition of WTA synthesis blocks the cooperative action of PBPs and sensitizes MRSA to b-lactams. ACS Chem. Biol. 8, 226–233. Fiolek, T.J., Banahene, N., Kavunja, H.W., Holmes, N.J., Rylski, A.K., Pohane, A.A., Siegrist, M.S., and Swarts, B.M. (2019). Engineering the mycomembrane of live mycobacteria with an expanded set of trehalose monomycolate ana- logues. ChemBioChem 20, 1282–1291. Garcı́a-Heredia, A., Pohane, A.A., Melzer, E.S., Carr, C.R., Fiolek, T.J., Run- dell, S.R., Lim, H.C., Wagner, J.C., Morita, Y.S., Swarts, B.M., and Siegrist, M.S. (2018). Peptidoglycan precursor synthesis along the sidewall of pole- growing mycobacteria. eLife 7, e37243. Gram, C. (1884). The differential staining of Schizomycetes in tissue sections and in dried preparations. Fortschr. Med. 2, 185–189. Ha, S., Chang, E., Lo, M.C., Men, H., Park, P., Ge, M., and Walker, S. (1999). The kinetic characterization of Escherichia coli MurG using synthetic substrate analogues. J. Am. Chem. Soc. 121, 8415–8426. Hsu, Y.-P., Booher, G., Egan, A., Vollmer, W., and VanNieuwenhze, M.S. (2019). d-amino acid derivatives as in situ probes for visualizing bacterial pepti- doglycan biosynthesis. Acc. Chem. Res. 52, 2713–2722. Hu, C. (2018). Pharmaceutical companies are backing away from a growing threat that could kill 10 million people a year by 2050 (Business In- sider). https://www.businessinsider.com/major-pharmaceutical-companies- dropping-antibiotic-projects-superbugs-2018-7. Jackson, M., McNeil, M.R., and Brennan, P.J. (2013). Progress in targeting cell envelope biogenesis in Mycobacterium tuberculosis. Future Microbiol. 8, 855–875. Kahne, D., Leimkuhler, C., Lu, W., and Walsh, C. (2005). Glycopeptide and lip- oglycopeptide antibiotics. Chem. Rev. 105, 425–448. Kavunja, H.W., Biegas, K.J., Banahene, N., Stewart, J.A., Piligian, B.F., Groe- nevelt, J.M., Sein, C.E., Morita, Y.S., Niederweis, M., Siegrist, M.S., and Swarts, B.M. (2020). Photoactivatable glycolipid probes for identifying myco- late-protein interactions in live mycobacteria. J. Am. Chem. Soc. 142, 7725–7731. Kawai, Y., Marles-Wright, J., Cleverley, R.M., Emmins, R., Ishikawa, S., Ku- wano, M., Heinz, N., Bui, N.K., Hoyland, C.N., Ogasawara, N., et al. (2011). A widespread family of bacterial cell wall assembly proteins. EMBO J. 30, 4931–4941. Kim, S.J., Singh, M., Sharif, S., and Schaefer, J. (2014). Cross-link formation and peptidoglycan lattice assembly in the FemA mutant of Staphylococcus aureus. Biochemistry 53, 1420–1427. Kim, S.J., Chang, J., and Singh, M. (2015). Peptidoglycan architecture of Gram-positive bacteria by solid-state NMR. Biochim. Biophys. Acta 1848 (1 Pt B), 350–362. Kocaoglu, O., and Carlson, E.E. (2016). Progress and prospects for small- molecule probes of bacterial imaging. Nat. Chem. Biol. 12, 472–478. Kuhn, A. (2019). The bacterial cell wall and membrane-a treasure chest for anti- biotic targets. In Bacterial Cell Walls and Membranes, A. Kuhn, ed. (Springer), pp. 1–5. Kuk, A.C.Y., Hao, A., Guan, Z., and Lee, S.-Y. (2019). Visualizing conformation transitions of the Lipid II flippase MurJ. Nat. Commun. 10, 1736. Lam, H., Oh, D.-C., Cava, F., Takacs, C.N., Clardy, J., de Pedro, M.A., and Waldor, M.K. (2009). D-amino acids govern stationary phase cell wall remod- eling in bacteria. Science 325, 1552–1555. Lancia, J.K., Nwokoye, A., Dugan, A., Joiner, C., Pricer, R., and Mapp, A.K. (2014). Sequence context and crosslinking mechanism affect the efficiency of in vivo capture of a protein-protein interaction. Biopolymers 101, 391–397. Lazar, K., and Walker, S. (2002). Substrate analogues to study cell-wall biosyn- thesis and its inhibition. Curr. Opin. Chem. Biol. 6, 786–793. Lee, W., Schaefer, K., Qiao, Y., Srisuknimit, V., Steinmetz, H., M€ uller, R., Kahne, D., and Walker, S. (2016). The mechanism of action of lysobactin. J. Am. Chem. Soc. 138, 100–103. Leimkuhler, C., Chen, L., Barrett, D., Panzone, G., Sun, B., Falcone, B., Oberth€ ur, M., Donadio, S., Walker, S., and Kahne, D. (2005). Differential inhibi- tion of Staphylococcus aureus PBP2 by glycopeptide antibiotics. J. Am. Chem. Soc. 127, 3250–3251. Li, F.K.K., Rosell, F.I., Gale, R.T., Simorre, J.P., Brown, E.D., and Strynadka, N.C.J. (2020). Crystallographic analysis of Staphylococcus aureus LcpA, the primary wall teichoic acid ligase. J. Biol. Chem. 295, 2629–2639. Meeske, A.J., Riley, E.P., Robins, W.P., Uehara, T., Mekalanos, J.J., Kahne, D., Walker, S., Kruse, A.C., Bernhardt, T.G., and Rudner, D.Z. (2016). SEDS proteins are a widespread family of bacterial cell wall polymerases. Nature 537, 634–638. Men, H.B., Park, P., Ge, M., and Walker, S. (1998). Substrate synthesis and ac- tivity assay for MurG. J. Am. Chem. Soc. 120, 2484–2485. Neumann, K.D., Villanueva-Meyer, J.E., Mutch, C.A., Flavell, R.R., Blecha, J.E., Kwak, T., Sriram, R., VanBrocklin, H.F., Rosenberg, O.S., Ohliger, M.A., and Wilson, D.M. (2017). Imaging active infection in vivo using D-amino acid derived PET radiotracers. Sci. Rep. 7, 7903. Nygaard, R., Romaniuk, J.A.H., Rice, D.M., and Cegelski, L. (2015). Spectral snapshots of bacterial cell-wall composition and the influence of antibiotics by whole-cell NMR. Biophys. J. 108, 1380–1389. Okuda, S., Freinkman, E., and Kahne, D. (2012). Cytoplasmic ATP hydrolysis powers transport of lipopolysaccharide across the periplasm in E. coli. Sci- ence 338, 1214–1217. Park, J.T. (1952). Uridine-50 -pyrophosphate derivatives. II. A structure com- mon to three derivatives. J. Biol. Chem. 194, 885–895. Parker, C.G., and Pratt, M.R. (2020). Click chemistry in proteomic investiga- tions. Cell 180, 605–632. Parker, M.F.L., Luu, J.M., Schulte, B., Huynh, T.L., Stewart, M.N., Sriram, R., Yu, M.A., Jivan, S., Turnbaugh, P.J., Flavell, R.R., et al. (2020). Sensing living bacteria in vivo using d-alanine-derived 11 C radiotracers. ACS Cent. Sci. 6, 155–165. Porter, J.D.H., and McAdam, K.P. (1994). The re-emergence of tuberculosis. Annu. Rev. Public Health 15, 303–323. Qiao, Y., Lebar, M.D., Schirner, K., Schaefer, K., Tsukamoto, H., Kahne, D., and Walker, S. (2014). Detection of lipid-linked peptidoglycan precursors by exploiting an unexpected transpeptidase reaction. J. Am. Chem. Soc. 136, 14678–14681. Qiao, Y., Srisuknimit, V., Rubino, F., Schaefer, K., Ruiz, N., Walker, S., and Kahne, D. (2017). Lipid II overproduction allows direct assay of transpeptidase inhibition by b-lactams. Nat. Chem. Biol. 13, 793–798. Radkov, A.D., Hsu, Y.-P., Booher, G., and VanNieuwenhze, M.S. (2018). Imag- ing bacterial cell wall biosynthesis. Annu. Rev. Biochem. 87, 991–1014. Rohs, P.D.A., Buss, J., Sim, S.I., Squyres, G.R., Srisuknimit, V., Smith, M., Cho, H., Sjodt, M., Kruse, A.C., Garner, E.C., et al. (2018). A central role for PBP2 in the activation of peptidoglycan polymerization by the bacterial cell elongation machinery. PLoS Genet. 14, e1007726. Rojas, E.R., Billings, G., Odermatt, P.D., Auer, G.K., Zhu, L., Miguel, A., Chang, F., Weibel, D.B., Theriot, J.A., and Huang, K.C. (2018). The outer membrane is an essential load-bearing element in Gram-negative bacteria. Nature 559, 617–621. Romaniuk, J.A.H., and Cegelski, L. (2018). Peptidoglycan and teichoic acid levels and alterations in Staphylococcus aureus by cell-wall and whole-cell nu- clear magnetic resonance. Biochemistry 57, 3966–3975. Rubino, F.A., Kumar, S., Ruiz, N., Walker, S., and Kahne, D.E. (2018). Mem- brane potential is required for MurJ function. J. Am. Chem. Soc. 140, 4481–4484. Rubino, F.A., Mollo, A., Kumar, S., Butler, E.K., Ruiz, N., Walker, S., and Kahne, D.E. (2020). Detection of transport intermediates in the peptidoglycan flippase MurJ identifies residues essential for conformational cycling. J. Am. Chem. Soc. 142, 5482–5486. Ruiz, N. (2008). Bioinformatics identification of MurJ (MviN) as the peptido- glycan lipid II flippase in Escherichia coli. Proc. Natl. Acad. Sci. USA 105, 15553–15557. ll Cell Chemical Biology 27, August 20, 2020 1061 Review
  • 11. Ruiz, N. (2016). Lipid flippases for bacterial peptidoglycan biosynthesis. Lipid Insights 8 (Suppl 1), 21–31. Salama, N.R. (2020). Cell morphology as a virulence determinant: lessons from Helicobacter pylori. Curr. Opin. Microbiol. 54, 11–17. Santajit, S., and Indrawattana, N. (2016). Mechanisms of antimicrobial resis- tance in ESKAPE pathogens. BioMed Res. Int. 2016, 2475067. Sarkar, S., Libby, E.A., Pidgeon, S.E., Dworkin, J., and Pires, M.M. (2016). In vivo probe of Lipid II-interacting proteins. Angew. Chem. Int. Ed. Engl. 55, 8401–8404. Schaefer, K., Matano, L.M., Qiao, Y., Kahne, D., and Walker, S. (2017). In vitro reconstitution demonstrates the cell wall ligase activity of LCP proteins. Nat. Chem. Biol. 13, 396–401. Schaefer, K., Owens, T.W., Kahne, D., and Walker, S. (2018). Substrate pref- erences establish the order of cell wall assembly in Staphylococcus aureus. J. Am. Chem. Soc. 140, 2442–2445. Schleifer, K.H., and Kandler, O. (1972). Peptidoglycan types of bacterial cell walls and their taxonomic implications. Bacteriol. Rev. 36, 407–477. Sham, L.-T., Butler, E.K., Lebar, M.D., Kahne, D., Bernhardt, T.G., and Ruiz, N. (2014). Bacterial cell wall. MurJ is the flippase of lipid-linked precursors for peptidoglycan biogenesis. Science 345, 220–222. Siegrist, M.S., Swarts, B.M., Fox, D.M., Lim, S.A., and Bertozzi, C.R. (2015). Illumination of growth, division and secretion by metabolic labeling of the bac- terial cell surface. FEMS Microbiol. Rev. 39, 184–202. Simpson, B.W., and Trent, M.S. (2019). Pushing the envelope: LPS modifica- tions and their consequences. Nat. Rev. Microbiol. 17, 403–416. Singh, M., Chang, J., Coffman, L., and Kim, S.J. (2016). Solid-state NMR char- acterization of amphomycin effects on peptidoglycan and wall teichoic acid biosyntheses in Staphylococcus aureus. Sci. Rep. 6, 31757. Sjodt, M., Brock, K., Dobihal, G., Rohs, P.D.A., Green, A.G., Hopf, T.A., Meeske, A.J., Srisuknimit, V., Kahne, D., Walker, S., et al. (2018). Structure of the peptidoglycan polymerase RodA resolved by evolutionary coupling analysis. Nature 556, 118–121. Stewart, M.N., Parker, M.F.L., Jivan, S., Luu, J.M., Huynh, T.L., Schulte, B., Seo, Y., Blecha, J.E., Villanueva-Meyer, J.E., Flavell, R.R., et al. (2020). High enantiomeric excess in-loop synthesis of d-[methyl-11 C]methionine for use as a diagnostic positron emission tomography radiotracer in bacterial infec- tion. ACS Infect. Dis. 6, 43–49. Strominger, J.L., Park, J.T., and Thompson, R.E. (1959). Composition of the cell wall of Staphylococcus aureus: its relation to the mechanism of action of penicillin. J. Biol. Chem. 234, 3263–3268. Taguchi, A., Kahne, D., and Walker, S. (2019a). Chemical tools to characterize peptidoglycan synthases. Curr. Opin. Chem. Biol. 53, 44–50. Taguchi, A., Welsh, M.A., Marmont, L.S., Lee, W., Sjodt, M., Kruse, A.C., Kahne, D., Bernhardt, T.G., and Walker, S. (2019b). FtsW is a peptidoglycan polymerase that is functional only in complex with its cognate penicillin-bind- ing protein. Nat. Microbiol. 4, 587–594. Tanaka, Y., and Kohler, J.J. (2008). Photoactivatable crosslinking sugars for capturing glycoprotein interactions. J. Am. Chem. Soc. 130, 3278–3279. Taylor, J.G., Li, X., Oberth€ ur, M., Zhu, W., and Kahne, D.E. (2006). The total synthesis of moenomycin A. J. Am. Chem. Soc. 128, 15084–15085. Touzé, T., Barreteau, H., El Ghachi, M., Bouhss, A., Barnéoud-Arnoulet, A., Patin, D., Sacco, E., Blanot, D., Arthur, M., Duché, D., et al. (2012). Colicin M, a peptidoglycan lipid-II-degrading enzyme: potential use for antibacterial means? Biochem. Soc. Trans. 40, 1522–1527. Typas, A., Banzhaf, M., Gross, C.A., and Vollmer, W. (2011). From the regula- tion of peptidoglycan synthesis to bacterial growth and morphology. Nat. Rev. Microbiol. 10, 123–136. Unemo, M., and Shafer, W.M. (2014). Antimicrobial resistance in Neisseria gonorrhoeae in the 21st century: past, evolution, and future. Clin. Microbiol. Rev. 27, 587–613. Vollmer, W., and Seligman, S.J. (2010). Architecture of peptidoglycan: more data and more models. Trends Microbiol. 18, 59–66. Vollmer, W., Blanot, D., and de Pedro, M.A. (2008). Peptidoglycan structure and architecture. FEMS Microbiol. Rev. 32, 149–167. Walsh, C. (2000). Molecular mechanisms that confer antibacterial drug resis- tance. Nature 406, 775–781. Walsh, C. (2003). Where will new antibiotics come from? Nat. Rev. Microbiol. 1, 65–70. Wu, H., and Kohler, J. (2019). Photocrosslinking probes for capture of carbo- hydrate interactions. Curr. Opin. Chem. Biol. 53, 173–182. Yang, H., Singh, M., Kim, S.J., and Schaefer, J. (2017). Characterization of the tertiary structure of the peptidoglycan of Enterococcus faecalis. Biochim. Bio- phys. Acta Biomembr. 1859, 2171–2180. Ye, X.-Y., Lo, M.-C., Brunner, L., Walker, D., Kahne, D., and Walker, S. (2001). Better substrates for bacterial transglycosylases. J. Am. Chem. Soc. 123, 3155–3156. Young, K.D. (2014). Microbiology. A flipping cell wall ferry. Science 345, 139–140. Yu, S.-H., Boyce, M., Wands, A.M., Bond, M.R., Bertozzi, C.R., and Kohler, J.J. (2012). Metabolic labeling enables selective photocrosslinking of O- GlcNAc-modified proteins to their binding partners. Proc. Natl. Acad. Sci. USA 109, 4834–4839. ll 1062 Cell Chemical Biology 27, August 20, 2020 Review