ArticlePDF Available

Plectasin, a Fungal Defensin, Targets the Bacterial Cell Wall Precursor Lipid II

Authors:

Abstract

Defining Defensins' Mode of Action Defensins are antimicrobial host defense peptides that play a role in innate immunity. Many such peptides act by disrupting the bacterial membrane; however, Schneider et al. (p. 1168 ) now show that the fungal defensin, plecstasin, targets cell wall biosynthesis. Biochemical studies identified Lipid II as the cellular target of plecstasin and the residues involved in complex formation were identified using NMR spectroscopy and computational modeling. Initial studies identified two defensins from invertebrates that also target Lipid II. Plecstasin is active against some drug-resistant Gram-positive bacteria, and its action against a validated target makes it a promising lead for further drug development.
1
Plectasin, a fungal defensin, targets the bacterial cell wall
precursor Lipid II
Metazoan and fungal host defence peptides act as specific inhibitors of bacterial
peptidoglycan biosynthesis
Tanja Schneider1, Thomas Kruse5, Reinhard Wimmer3, Imke Wiedemann1, Vera Sass1, Ulrike
Pag1, Andrea Jansen1, Allan K Nielsen2, Per H Mygind2, Dorotea S Raventós2, Søren Neve2,
Birthe Ravn2, Alexandre MJJ Bonvin4, Leonardo De Maria2, Anders S Andersen2,5, Lora K
Gammelgaard2, Hans-Georg Sahl1 & Hans-Henrik Kristensen2*
1Institute for Medical Microbiology, Immunology and Parasitology – Pharmaceutical
Microbiology Section, University of Bonn, D-53115 Bonn, Germany
2Novozymes AS, DK-2880 Bagsvaerd, Denmark
3Department of Biotechnology, Chemistry and Environmental Engineering, Aalborg
University, DK-9000 Aalborg, Denmark
4Department of Chemistry, Faculty of Science, Utrecht University, 3584 CH Utrecht, the
Netherlands
5Statens Serum Institut, 2300 Copenhagen S, Denmark
*To whom correspondence should be addressed: hahk@novozymes.com (HHK)
2
Host defence peptides such as defensins are components of innate immunity and have
retained antibiotic activity throughout evolution. Their activity is thought to be due to
amphipathic structures which enable binding and disruption of microbial cytoplasmic
membranes. Contrary to this, we show here that plectasin, a fungal defensin, acts by
directly binding the bacterial cell wall precursor Lipid II. A wide range of genetic and
biochemical approaches identify cell wall biosynthesis as the pathway targeted by
plectasin. In vitro assays for cell wall synthesis identified Lipid II as the specific cellular
target. Consistently, binding studies confirmed the formation of an equimolar
stoichiometric complex between Lipid II and plectasin. Furthermore, key residues in
plectasin involved in complex formation were identified using NMR spectroscopy and
computational modelling.
3
Plectasin is a 40 amino acid residue fungal defensin produced by the saprophytic ascomycete
Pseudoplectania nigrella (1). Plectasin shares primary structural features with defensins from
spiders, scorpions, dragonflies and mussels and folds into a cystine-stabilized alpha-beta-
structure (CSαβ). In vitro and in animal models of infection, plectasin is potently active
against drug-resistant gram-positive bacteria such as streptococci, while the antibacterial
spectrum of an improved derivative, NZ2114 (2), also includes staphylococci such as
methicillin-resistant Staphylococcus aureus (MRSA).
Here we set out to determine the molecular target and specific mechanism by which
plectasin kills bacteria. While many host defence peptides (HDPs) act on and disintegrate the
bacterial membrane several observations suggested that this is not the case for plectasin.
Growth kinetic measurements of the gram-positive bacterium Bacillus subtilis exposed
to plectasin clearly demonstrated that plectasin exhibited kinetic behaviour similar to cell
wall-interfering agents (e.g. vancomycin, penicillin and bacitracin) and not to the rapidly-lytic
membrane-active agents (e.g. polymyxin and novispirin) or non-lytic antibiotics with
replication (ciprofloxacin), transcription (rifampicin) or protein translation (kanamycin,
tetracycline) as their primary target (Fig. 1A). Consistent with this, killing kinetics indicated
that over a period of approximately one generation time (0.5 hours), treated cells were unable
to multiply, but remained viable (Fig. 1B insert), before the number of colony-forming units
decreased (Fig. 1B). Next, the effect of plectasin on macromolecular biosynthesis pathways
was investigated. The incorporation of radiolabelled isoleucine into protein and of thymidine
into nucleic acids was not affected whereas glucosamine, an essential precursor of bacterial
peptidoglycan, was no longer incorporated (Fig. 1C). Finally, treatment of B. subtilis with
plectasin induced severe cell shape deformations as visualized by phase contrast microscopy
(Fig. S1). These characteristics are all typical for compounds interfering with cell-wall
biosynthesis rather than for membrane disintegration (3, 4). Consistently, neither pore
formation as measured by K+ efflux (Fig. 1E) nor changes in membrane potential using TPP+
or DiBAC4 (Fig. S2AB), nor carboxy-fluorescein efflux from liposomes were detected (Fig.
S2C). Thus, despite its amphipathic nature, plectasin does not compromise membrane
integrity reducing the risk of unspecific toxicity.
We obtained further support for the cell-wall interfering activity using DNA
microarrays to compare the transcriptional responses of plectasin-treated cells with response
patterns obtained for a range of reference antibiotics. For both B. subtilis 168 and S. aureus
SG511, we found that the transcriptional profiles overlapped those of established cell wall
4
biosynthesis inhibitors such as vancomycin and bacitracin (5, 6, 7, 8) (Fig. S3; Tables S1 and
S2).
The biosynthesis of bacterial cell walls requires a number of steps (9). Initially the N-
acetylmuramic acid-pentapeptide (MurNAc-pentapeptide), a major constituent of the cell wall
building block is produced in the cytoplasm as an UDP-activated precursor before it is
transferred onto a membrane carrier, bactoprenolphosphate (reaction I, Fig 2B). The resulting
membrane-anchored precursor Lipid I is then further modified to the structural cell wall
subunit, Lipid II (reaction II). In some gram-positive bacteria, Lipid II (Fig. 2A) is further
decorated by an interpeptide bridge (a pentaglycine peptide in case of S. aureus (10), reaction
III), before it gets translocated across the cytoplasmic membrane to the outside, where it is
incorporated into the peptidoglycan polymer through the activity of transglycosylases and
transpeptidases (reactions IV). We analysed the intracellular pool of cell wall precursors by
reverse HPLC and mass spectrometry and found accumulation of the soluble molecule, UDP-
MurNAc-pentapeptide in plectasin-treated cells (Fig. 1D), suggesting that one of the later,
membrane-associated or extracellular processes may be targeted by plectasin.
We then analysed the effect of plectasin on the membrane-bound steps of cell wall
biosynthesis in vitro. Cytoplasmic membranes with associated CW biosynthesis apparatus
were isolated and incubated with plectasin and radiolabelled substrates necessary for Lipid II
formation. Using thin layer chromatography and subsequent scintillation counting we found
the overall synthesis reaction to be strongly inhibited (Fig. 2C). For a more detailed analysis,
we cloned the individual CW biosynthesis genes from S. aureus, expressed them in
Escherichia coli and analysed the activity of the purified enzymes in the presence of plectasin
by measuring the amount of product formed. These enzymes included MraY (Fig. 2B,
reaction I); MurG (Fig. 2B, reaction II); FemXAB (Fig. 2B, reaction III); and PBP2 (Fig. 2B,
reaction IV). Whereas the MraY reaction was not affected by plectasin, we found the MurG,
FemX and PBP2 reactions to be inhibited in a dose-dependent fashion (Fig.2C). For these
three enzymes, Lipid I (MurG) or Lipid II (FemX and PBP2) are substrates and significant
inhibition of the reactions was only observed when plectasin was added in equimolar
concentrations with respect to Lipid I or Lipid II (Fig. 2C). Thus plectasin, similar to
glycopeptide antibiotics (e.g. vancomycin (11, 12)) and lantibiotics (13, 14), may form a
stoichiometric complex with the substrate rather than inhibiting the enzyme. To further
validate this we incubated either Lipid I or II with plectasin in various molar ratios and used
thin layer chromatography to analyse the migration behaviour. Free Lipid I and II as well as
free peptide were found to migrate to defined positions in the chromatogram, while the Lipid
5
I/II-plectasin complex remained at the start point (Fig. 2D). Only at an equimolar ratio, was
neither free Lipid I/II nor free peptide detectable, indicating the formation of a 1:1
stoichiometric complex.
We further analysed the interaction of both Lipid I and II with plectasin using a
liposome system with membranes composed of phosphatidylcholine and Lipid II (0.2 or 0.5
mol%) and 14C-labelled plectasin. We found the maximum number of plectasin molecules that
bound to liposomes to approximately match the number of Lipid II molecules available on the
liposome surface (Fig. S4). Using Scatchard plot analysis, we determined an equilibrium
binding constant of 1.8x10-7 mol for Lipid II and 1.1x10-6 mol for Lipid I, suggesting that the
second sugar in Lipid II, the N-acetyl glucosamine contributes to the stability of the complex.
To gain further insight into the structural nature of the plectasin/Lipid II interaction at
the membrane interface we measured chemical shifts changes for 15N-labelled plectasin.
Heteronuclear single quantum coherence (HSQC) nuclear magnetic resonance (NMR) spectra
were measured either in solution or upon binding membrane-mimicking
dodecylphosphocholine (DPC) micelles (Fig. S5, Fig. S6). Fitting the binding data to a
Langmuir isotherm yielded a free enthalpy of binding ΔG=-27±1 kJ/mol (Fig. S7). Backbone
HN and N atoms of ten residues (G6, W8, D9, A31, K32, G33, G34, F35, V36 and C37),
which in the tertiary structure all locate to one end of plectasin, exhibited marked changes in
chemical shifts (Δδobs>0.15 ppm) (Fig. 3A; residues labelled yellow), suggesting an
orientation in which one end of plectasin specifically is located in the membrane interface.
To identify the residues on plectasin that bind Lipid II, we then titrated plectasin
bound to DPC with Lipid II. With increasing concentrations of Lipid II, another set of NMR
signals appeared and became stronger, whereas the NMR signals of apo-plectasin bound to
DPC micelles became weaker, until they disappeared at equimolar concentrations of plectasin
and Lipid II, supporting the 1:1 binding stoichiometry found by TLC. Addition of extra
plectasin to the mixture brought the signals of apo-plectasin forward again, further addition of
Lipid II to equimolarity led to the disappearance of the signals again. From a 3D-HNCA
spectrum we could assign backbone HN, N and Cα signals of the plectasin:Lipid II:DPC
complex . The strongest changes in chemical shift (Δδobs>0.22 ppm) were obtained for amino
acids F2, C4, D12, Y29, A31, G33, C37 and K38 (Fig. S5, Fig. S6). Most of these residues
localize in a coherent patch in close proximity to the residues affected by binding to DPC (Fig.
3B, residues labelled magenta). A31, G33 and C37 exert chemical shift-changes both upon
addition of DPC and Lipid II. To further verify this, site-saturated mutagenesis (where a given
6
amino acid is changed to each of the other 19 natural amino acids) was carried out at all
positions in plectasin except the 6 cysteines. The mutant libraries were expressed in S.
cerevisiae and 400-600 transformants of each position tested for activity against S. aureus in a
plate overlay assay. No amino acid substitutions at positions D12, Y29 or G33 resulted in
activity against S. aureus, whereas only the very conservative mutations of A31 to G and K38
to R resulted in activity against S. aureus. At other amino acid positions not involved in DPC
or Lipid II binding, a wide range of non-homologous amino acid substitutions gave rise to
plectasin variants retaining antimicrobial activity.
To visualize the complex between Lipid II and plectasin, docking studies using the
GOLD and HADDOCK programs were performed (15, 16). In accordance with the NMR data,
evidence in favour of a primary binding site involving the interaction of the pyro-phosphate
moiety of Lipid II with the amide protons F2, G3, C4 and C37 of plectasin via hydrogen
bonding was obtained (Fig. 3C). Several of the other large chemical shift changes are present
in residues involved in secondary structure interactions (e.g. formation of beta-sheets), which
most likely undergo structural changes upon binding to the target. Taken together, these data
strongly support a model in which plectasin gains affinity and specificity through binding to
the solvent-exposed part of Lipid II while the hydrophobic part of plectasin is located in the
membrane interface. Thus, plectasin shares functional features with the lantibiotic nisin in that
for both peptides the pyrophosphate moiety is most relevant for binding of Lipid II, although
nisin inserts deeply into the membrane bilayer, forming pores and causing major
delocalization of Lipid II (17,18).
To test whether inhibition of CW biosynthesis is restricted to plectasin or represents a
general feature we tested a series of defensin peptides from other fungi, mollusc and
arthropods for Lipid II binding and inhibition of the overall Lipid II synthesis and FemX
reaction (Fig. S8A). Two fungal defensins, oryzeasin (from Aspergillus oryzea) and eurocin
(Eurotium amstelodami) did inhibit the enzymatic reactions and bind to Lipid II in
stoichiometric numbers as did the two defensins from invertebrates, lucifensin from maggots
of the blowfly Lucilia sericata and gallicin from the mussel Mytilus galloprovinciali (Fig.
S8BCD). In contrast, heliomicin from the tobacco budworm Heliothis virescens which share
the conserved cysteine pattern did not show affinity for Lipid II and had no activity in these
assays. These data clearly demonstrate that among the host defence peptides of eukaryotic
organisms specific inhibitors of CW biosynthesis can be found which directly target Lipid II,
“the bacterial Achilles’ heel” for antibiotic attack (19).
7
Vancomycin, one of the very few remaining drugs for the treatment of multi-resistant
gram-positive infections, has been shown to predominantly bind the D-alanyl-D-alanine (D-
ala-D-ala) part of the pentapeptide in Lipid II (11) (Fig. 2A). However, high-level
vancomycin resistance has been observed in both enterococci (VRE) and staphylococci
(VRSA). Importantly, there is no cross-resistance between vancomycin and plectasin, and in
contrast to vancomycin, plectasin is not competitively inhibited by the presence of the D-ala-
D-ala ligand (Fig. S9). This further demonstrates that the primary interactions to Lipid II
differ between plectasin and vancomycin and taken together these results suggest that future
development of true cross-resistance between vancomycin and plectasin is unlikely.
Plectasin and its improved derivatives such as NZ2114 possess a range of features -
such as potent activity in vitro under physiological conditions and in animal models of
infection, low potential for unwanted toxicities, extended serum stability and in vivo half-life,
and cost-effective large-scale manufacturing – which combined with a validated microbial
target make it a promising lead for further drug development.
8
Reference List
1. P. H. Mygind et al., Nature 437, 975 (2005).
2. D. Andes, W. Craig, L. A. Nielsen, H. H. Kristensen, Antimicrob. Agents Chemother. 53,
3003 (2009).
3. M. Zasloff, Nature 415, 389 (2002).
4. R. E. Hancock, H. G. Sahl, Nat. Biotechnol. 24, 1551 (2006).
5. B. Hutter et al., Antimicrob. Agents Chemother. 48, 2838 (2004).
6. F. McAleese et al., J. Bacteriol. 188, 1120 (2006).
7. M. Cao, T. Wang, R. Ye, J. D. Helmann, Mol. Microbiol. 45, 1267 (2002).
8. T. Mascher, N. G. Margulis, T. Wang, R. W. Ye, J. D. Helmann, Mol. Microbiol. 50,
1591 (2003).
9. H. J. van Heijenoort, Microbiol. Mol. Biol. Rev. 71, 620 (2007).
10. T. Schneider et al., Mol. Microbiol. 53, 675 (2004).
11. P. E. Reynolds, Eur. J. Clin. Microbiol. Infect. Dis. 8, 943 (1989).
12. N. E. Allen, T. I. Nicas, FEMS Microbiol. Rev. 26, 511 (2003).
13. H. Brotz et al., Mol. Microbiol. 30, 317 (1998).
14. J. M. Willey, W. A. van der Donk, Annu. Rev. Microbiol. 61, 477 (2007).
15. G. Jones, P. Willett, R. C. Glen, A. R. Leach, R. Taylor, J. Mol. Biol. 267, 727 (1997).
16. C. Dominguez, R. Boelens, A. M. Bonvin, J. Am. Chem. Soc. 125, 1731 (2003).
17. S. T. Hsu et al., Nat. Struct. Mol. Biol. 11, 963 (2004).
18. H. E. Hasper et al., Science 313, 1636 (2006).
19. E. Breukink, B. de Kruijff, Nat. Rev. Drug Discov. 5, 321 (2006).
20. We thank Michaele Josten, Annette Hansen and Marianne R Markvardsen for expert
technical assistance and acknowledge the Carlsberg Research Center for use of the 800MHz
NMR spectrometer and the Obel Foundation for supporting the NMR laboratory at Aalborg
University. HGS acknowledges financial support by the German Research Foundation (SA
292/10-2 and SA 292/13-1), the BMBF (SkinStaph) and by the BONFOR programme of the
Medical Faculty, University of Bonn. AMJJB acknowledges financial support from the
Netherlands Organization for Scientific Research (VICI grant #700.56.442). ASA
acknowledges financial support from The Danish Research council for Technology and
Production (274-05-0435). Competing interest statement: Allan K Nielsen, Dorothea S
Raventós, Søren Neve, Birthe Ravn, Leonardo De Maria, Anders S Andersen & Hans-Henrik
Kristensen are employees of Novozymes. The authors declare they have no other competing
9
financial interest. DNA microarray data can be accessed through ArrayExpress, accession
number: E-MTAB-60. NMR assignment of 1H, 15N and 13C atoms of plectasin have been
deposited in the BioMagResBank (accession no. 16739)
Supporting Online Material
Materials and Methods
Figs S1 to S9
Tables S1 to S2
References
10
FIGURE LEGENDS
Fig. 1.
Effect of Plectasin on intact cells. (A) Classification of antimicrobial compounds using optical
density measurements. Growth kinetic measurements of B. subtilis exposed to plectasin or
various antibiotics with known cellular targets. 2-4 times the minimal inhibitory concentration
(MIC) of the respective compounds were used. Plectasin (black) falls into the cluster of cell
wall biosynthesis inhibiting antibiotics (red colours). (B) Killing kinetics of plectasin;
Staphylococcus simulans 22 treated with plectasin at 2 × MIC (open diamonds) and 4 x MIC
(squares); control without peptide (triangles). Insert shows a similar experiment with more
time points within the first 60 minutes demonstrating the absence of killing in the first 30 min
of treatment (C) Impact of plectasin on macromolecular biosynthesis in B. subtilis 168.
Incorporation of [14C]-thymidine into nucleic acids, of L-[14C]-isoleucine into protein and of
[3H]–glucosamine in cell wall was measured in untreated controls (squares) and plectasin
treated cells (open circles); glucosamine incorporation into cell wall material was selectively
inhibited. (D) Intracellular accumulation of the ultimate soluble cell wall precursor UDP-
MurNAc-pentapeptide in vancomycin-treated (dotted line) and plectasin-treated (dashed line)
cells of S. simulans 22. Cells were treated for 30 min with plectasin or vancomycin, which is
known to form a complex with Lipid II1. Treated cells were extracted with boiling water and
the intracellular nucleotide pool analyzed by reversed HPLC. UDP-MurNAc-pentapeptide
was identified by mass spectrometry using the negative mode and 1 mg/ml 6-aza-2-
thiothymine (in 50% (v/v) ethanol/20 mM ammonium citrate) as matrix; the calculated
monoisotopic mass is 1149.35; in addition to the singly-charged ion, the mono- and di-sodium
salts are detected. (E) Plectasin is unable to form pores in the cytoplasmic membrane of S.
simulans 22. Potassium efflux from living cells was monitored with a potassium-sensitive
electrode. Ion leakage is expressed relative to the total amount of potassium released after
addition of 1 µM of the pore forming lantibiotic nisin (100%, open diamonds). Plectasin was
added at 0.2 µM (triangles) and 1 µM (open triangles); controls without peptide antibiotics
(squares).
Fig. 2.
Inhibition of membrane associated cell wall biosynthesis steps. (A) Structure of the cell wall
precursor Lipid II. (B) The membrane-bound steps of cell wall precursor biosynthesis and
11
bactoprenol (C55P) carrier cycling in staphylococci. Cell wall biosynthesis starts in the
cytoplasm with the formation of the soluble precursor UDP-MurNAc-pentapeptide (UDP-
MurNAc-pp). This precursor is linked to the membrane carrier bactoprenolphosphate (C55P)
by MraY yielding Lipid I (step I). Lipid II is formed by MurG which adds N-acetyl-
glucosamine (GlcNAc) (step II). When the interpeptide bridge, which only occurs in some
Gram-positive bacteria, is accomplished (step III), the monomeric peptidoglycan unit is
translocated across the cytoplasmic membrane to the outside and incorporated into the cell
wall (step IV). (C) Inhibition of membrane associated steps of cell wall biosynthesis by
plectasin. In all tests, plectasin was added in molar ratios of 0.1 to 1 with respect to the
amount of the appropriate lipid substrate C55P, Lipid I or Lipid II used in the individual test
system. The amount of reaction products synthesized in the absence of plectasin was taken as
100%. Product analysis was done by thin layer chromatography (see D) and subsequent
scintillation counting of stained and excised product-containing bands; radiolabelling was
based on [3H]-labelled C55P (for Lipid I), [14C]-GlcNAc for Lipid II and [14C]-glycine for
Lipid II-Gly1. Error bars represents +/- SD and the experiments were repeated at least three
times. Technical details on the assays, the cloning and purification of the enzymes are given
in Materials and Methods. (D) Estimation of the stoichiometry of plectasin:Lipid II binding.
Lipid II was incubated in the presence of plectasin at the molar concentration ratios indicated.
The stable complex of plectasin with the Lipid II remains at the application spot whereas both
components migrate to the sites indicated. At a molar ratio of 1:1 neither free Lipid II nor free
plectasin were observed. Data obtained with Lipid I were comparable (not shown).
Fig. 3.
NMR-based model of the plectasin/Lipid II-complex. (A) Surface representation of plectasin
with the residues showing significant chemical shift perturbations upon binding to DPC
micelles indicated in yellow. (B) Surface representation of plectasin with the residues
showing significant chemical shift perturbations upon Lipid II titration shown in magenta. (C)
Detailed view of the pyrophosphate binding pocket. In this proposed HADDOCK-generated
model the pyrophosphate moiety forms hydrogen bonds to F2, G3, C4 and C27 and the D-γ-
Glutamate of Lipid II forms a salt bridge with the N-terminus of plectasin and the side-chain
of His18.
12
Figure 1. Effect of plectasin on intact cells
B
E
C
0.2 µM plectasin
0
20
40
60
80
100
050 100 150 200 250 300 [sec]
% potassium release
addition of
peptide
no peptide antibiotic
nisin [1 µM]
plectasin [1 x MIC]
plectasin [5 x MIC]
A
contro
l
0010 20 30[min]
14C-isoleucine
0
0.5
1.0
1.5
2.0
2.5
CPM x 103
0
0.5
1.0
1.5
2.0
2.5
010 20 30[min]
CPM x 103
14C-thymidine
0
5
10
15
20
25
010 20 30[min]
3H-glucosamine
CPM x 103
010 20 30 40 [min]
101
102
103
104
105
106
107
108
109
CFU/ml
120 240
plectasin-
treated
D
UDP-MurNAc-
pentapeptide
[min]
010 20 30
untreated
vancomycin-
treated
0
0.1
0.2
A260
m/z
1148.723
1170.624
1192.444
0
1000
2000
3000
4000
intensity [a.u.]
1100 1200
0 1 2 3 4 5 6
2x MIC 4x MIC control
[h]
log cfu/ mL
9
4
5
6
7
8
AB
O
O
OPO
O
O
P
O
OO
2
8
L-Ala
D-Glu
L-Lys
D-Ala
D-Ala
O
O
HO
CH2
NH
CO
CH3
HO
HO
CH
CO
CH2
OH
NH
CO
CH3
H3C
bactoprenol
(C55)
pentapeptide-
sidechain
MurNAc
GlcNAc
Gly-Gly-Gly-Gly-Gly-
interpeptide-
bridge
plectasin
Lipid II
lipid II : plectasin ratio
1:0
1:0.1
1:0.25
1:0.5
1:1
1:2
0:1
C
D
Figure 2. Inhibition of membrane associated cell wall biosynthesis steps
UDP
UMP UDP
MraY MurG FemXAB
PBP
III III
IV
cytoplasm
peptidoglycan
PP
P
P
P
P
P
P
P
P
P
UDP
I
MurNAc
GlcNAc
C55-P
Ppentaglycine-
interpeptidebridge
pentapeptide-
sidechain
Pi
10
20
30
40
50
60
70
80
90
100
product formed (%)
substrate
used
product
formed
overall
reaction
(I-II)
MraY
(I)
MurG
(II)
FemX
(III)
PBP2
(IV)
C55P C55P lipid I lipid II lipid II
lipid II lipid I lipid II lipid II-Gly1 C55PP
plectasin :
substrate ratio 0.1:1 0.25:1 0.5:1
1:1 nisin
Lipid II/ plectasin
complex
Figure 3. NMR-based model of the plectasin/Lipid II-complex
... M i c h a e l S u p p l e m e n t 3 2 11 easy starting that during the pandemic 13 , but now we are in a phase where we are deep into collecting data, and one way of collecting the data is doing such witness seminars where we bring together people who are involved in stories that relate to the interest of our project. ...
... The idea is that it's not an interview; it's just trying to remember what happened back then, how you felt as a researcher, what your perspectives were, and a little bit about the background. 13 Covid-19 short for "Coronavirus Disease 2019", is a highly contagious respiratory illness caused by the novel coronavirus SARS-CoV-2. It was first identified in Wuhan, China, in late 2019 and has since spread globally, leading to a worldwide pandemic. ...
... 54 Investigator, The Lundquist Institute. Professor of Medicine, David Geffen School of Medicine at UCLA. https://lundquist.org/arnold-bayer-md. 55 Tanja Schneider was first author of a Plectasin mode of action paper (13). 56 Hans-Georg Sahl comments here: actually, I do not remember that we had any difficulties about showing Lipid II binding .. and we never were disappointed. ...
Article
Full-text available
The seminar "Plectasin's Odyssey: From Hopeful Beginnings to Untimely End," held on the 16th of November 2023 at University of Copenhagen, Denmark, gives us an overview of the challenges in developing new antibiotics inside the antimicrobial pipeline. The seminar focused on the story of Plectasin, the first antimicrobial peptide found in a fungus, discovered by Novozymes back in 2002. A group of twelve people was invited, including key figures of the original discovery team of Plectasin, as well as academic historians and researchers. The story of Plectasin highlighted how complex and challenging it is to bring a new drug into pharmaceutical development. The participants reflected on the initial excitement of discovering Plectasin and its potential as a good antimicrobial candidate. Subsequently, discussions focused on market obstacles, including the crucial roles of regulatory and economic stakeholders. The issue of a visible gap between the health system’s needs and the pharmaceutical industry’s focus, the importance of innovative funding and development models, and the potential for repurposing shelved antimicrobial candidates were addressed. However, it was agreed that publication of the discovery in prestigious journals such as Nature and Science added to the perceived competence by the public of Novozymes as a company. In summary, a common thread in the discussion was the urgent need for a shift towards more collaborative, community-focused, and sustainably financed strategies in antibiotic development. Such a change is essential if we are to stand a chance in the global fight against antimicrobial resistance (AMR).
... M i c h a e l S u p p l e m e n t 3 2 11 easy starting that during the pandemic 13 , but now we are in a phase where we are deep into collecting data, and one way of collecting the data is doing such witness seminars where we bring together people who are involved in stories that relate to the interest of our project. ...
... The idea is that it's not an interview; it's just trying to remember what happened back then, how you felt as a researcher, what your perspectives were, and a little bit about the background. 13 Covid-19 short for "Coronavirus Disease 2019", is a highly contagious respiratory illness caused by the novel coronavirus SARS-CoV-2. It was first identified in Wuhan, China, in late 2019 and has since spread globally, leading to a worldwide pandemic. ...
... 54 Investigator, The Lundquist Institute. Professor of Medicine, David Geffen School of Medicine at UCLA. https://lundquist.org/arnold-bayer-md. 55 Tanja Schneider was first author of a Plectasin mode of action paper (13). 56 Hans-Georg Sahl comments here: actually, I do not remember that we had any difficulties about showing Lipid II binding .. and we never were disappointed. ...
Book
The seminar "Plectasin's Odyssey: From Hopeful Beginnings to Untimely End," held on the 16th of November 2023 at University of Copenhagen, Denmark, gives us an overview of the challenges in developing new antibiotics inside the antimicrobial pipeline. The seminar focused on the story of Plectasin, the first antimicrobial peptide found in a fungus, discovered by Novozymes back in 2002. A group of twelve people was invited, including key figures of the original discovery team of Plectasin, as well as academic historians and researchers. The story of Plectasin highlighted how complex and challenging it is to bring a new drug into pharmaceutical development. The participants reflected on the initial excitement of discovering Plectasin and its potential as a good antimicrobial candidate. Subsequently, discussions focused on market obstacles, including the crucial roles of regulatory and economic stakeholders. The issue of a visible gap between the health system’s needs and the pharmaceutical industry’s focus, the importance of innovative funding and development models, and the potential for repurposing shelved antimicrobial candidates were addressed. However, it was agreed that publication of the discovery in prestigious journals such as Nature and Science added to the perceived competence by the public of Novozymes as a company. In summary, a common thread in the discussion was the urgent need for a shift towards more collaborative, community-focused, and sustainably financed strategies in antibiotic development. Such a change is essential if we are to stand a chance in the global fight against antimicrobial resistance (AMR).
... These cationic proteins exhibit broad-spectrum antibacterial activity, disrupting bacterial membranes [153][154][155]. Defensins, in particular, hinder bacterial cell wall synthesis through interactions with lipid II [156]. Moreover, they activate CCR6-positive dendritic cells, neutralize exotoxins, and regulate water and salt absorption by IECs [157][158][159]. ...
Article
Full-text available
The intestinal lumen acts as a critical interface connecting the external environment with the body’s internal state. It’s essential to prevent the passage of harmful antigens and bacteria while facilitating nutrient and water absorption. The intestinal barriers encompass microbial, mechanical, immunological, and chemical elements, working together to maintain intestinal balance. Numerous studies have associated m6A modification with intestinal homeostasis. This review comprehensively outlines potential mechanisms through which m6A modification could initiate, exacerbate, or sustain barrier damage from an intestinal perspective. The pivotal role of m6A modification in preserving intestinal equilibrium provides new insights, guiding the exploration of m6A modification as a target for optimizing preventive and therapeutic strategies for intestinal homeostasis.
... Although the known species are not rich in Pseudoplectania, this genus has broad research prospects. For example, the Plectasin, isolated from P. nigrella, is the first defensin to be isolated from a fungus, implying that the genus may be a new resource to help solve the problem of bacterial resistance (Mygind et al. 2005, Schneider et al. 2010, Li et al. 2017. Some species, such as P. melaena, are sometimes considered as indicator organisms of ecological damage (Holec & Kříž 2013, Krisai-Greilhuber 2019. ...
Article
A cup fungus producing ascomata on senescing to dead rhizomes of bamboo in East China is described as Pseudoplectania mystica sp. nov. (Ascomycota, Pezizales) based on morphological, ecological and phylogenetic evidence. Some sequences of unidentified endophytic fungi uploaded to the GenBank database by other researchers were also identified as P. mystica in our phylogeny. By checking the original literature of these sequences, P. mystica was further found to act as an endophyte in a broad range of host plants including Aegiceras, Cinnamomum, Dendrobium and Lindera, probably lurking in the tissues of the host plants as invisible mycelium in most cases and only producing visible ascomata following senescence of the host tissues. Such ecology and a combination of all its morphological characteristics make P. mystica recognizable from other known species of Pseudoplectania. This study contributes to the understanding of the species diversity and lifecycle of Pseudoplectania.
Article
Full-text available
Utviklingen av nye legemidler er langvarig, komplisert, kostbar og oftest avhengig av den private farmasøytiske industri. For antibiotika har dette vist seg å være særlig uheldig. Plectasin – et lovende peptid – ble stoppet på målstreken på grunn av firmaets markedsoverveielser. Andre utviklingsstrategier synes nødvendige hvis man skal kunne møte samfunnets behov for antibiotika. Historien om Plectasin ble belyst ved et medisinhistorisk aktørseminar, der involverte reflekterte over hva som hadde skjedd.
Article
Full-text available
Antimicrobial resistance is a leading cause of mortality, calling for the development of new antibiotics. The fungal antibiotic plectasin is a eukaryotic host defence peptide that blocks bacterial cell wall synthesis. Here, using a combination of solid-state nuclear magnetic resonance, atomic force microscopy and activity assays, we show that plectasin uses a calcium-sensitive supramolecular killing mechanism. Efficient and selective binding of the target lipid II, a cell wall precursor with an irreplaceable pyrophosphate, is achieved by the oligomerization of plectasin into dense supra-structures that only form on bacterial membranes that comprise lipid II. Oligomerization and target binding of plectasin are interdependent and are enhanced by the coordination of calcium ions to plectasin’s prominent anionic patch, causing allosteric changes that markedly improve the activity of the antibiotic. Structural knowledge of how host defence peptides impair cell wall synthesis will likely enable the development of superior drug candidates.
Chapter
The escalated increase in morbidity and mortality in the recent era has resulted due to the impact of antimicrobial resistance (AMR) among microorganisms. There could be more than the predicted deaths resulting due to cancer as reported by the United Nations Environment Program, which estimates ten million annual deaths from AMR by 2050. Thus, there is a dire need to develop new antimicrobial agents and strategies to address such a pressing issue. Naturally obtained small multifunctional supramolecular peptides and antimicrobial peptides (AMPs) are found in all living organisms as an innate immune system. Therefore, to combat AMR, a multimodal target is a prerequisite. Such AMP-based nanomaterials would not only have an antimicrobial activity but would also have the potential to reinvigorate old drugs. In this chapter, we will emphasize the overview of the importance and progress of AMPs, the mechanism of action of AMPs, the technique of designing AMP-based nanomaterials, and the application and classification of such AMPs. A brief discussion has also been provided on the challenges of developing AMP-based nanomaterials and strategies to overcome them. Finally, the chapter summarizes future prospects and major potentials in harnessing antimicrobial activities via AMP-based nanomaterials.
Article
This review summarizes the medium-sized antimicrobial peptides discovered in the last three decades (1993 to the end of 2022) and highlights the novel antibacterial mechanisms as well as part of the structure–activity relationships.
Article
Full-text available
Lantibiotics are polycyclic peptides containing unusual amino acids, which have binding specificity for bacterial cells, targeting the bacterial cell wall component lipid II to form pores and thereby lyse the cells. Yet several members of these lipid II–targeted lantibiotics are too short to be able to span the lipid bilayer and cannot form pores, but somehow they maintain their antibacterial efficacy. We describe an alternative mechanism by which members of the lantibiotic family kill Gram-positive bacteria by removing lipid II from the cell division site (or septum) and thus block cell wall synthesis.
Article
Full-text available
Bacillus subtilis encodes seven extracytoplasmic function (ECF) sigma factors. The σW regulon includes functions involved in detoxification and protection against antimicrobials, whereas σM is essential for growth at high salt concentrations. We now report that antibiotics that inhibit cell wall biosynthesis induce both σW and σM regulons as monitored using DNA microarrays. Induction of selected σW -dependent genes was confirmed using lacZ reporter fusions and Northern blot analysis. The ability of vancomycin to induce the σW regulon is dependent on both σW and the cognate anti- σ , RsiW, but is independent of the transition state regulator AbrB. These results suggest that the membrane-localized RsiW anti- σW factor mediates the transcriptional response to cell wall stress. Our findings are consistent with the idea that one function of ECF σ factors is to coordinate antibiosis stress responses and cell envelope homeostasis.
Article
Full-text available
NZ2114 is a novel plectasin derivative with potent activity against gram-positive bacteria, including multiply drug-resistant strains. We used the neutropenic murine thigh infection model to characterize the time course of antimicrobial activity of NZ2114 and determine which pharmacokinetic/pharmacodynamic (PK/PD) index and magnitude best correlated with efficacy. Serum drug levels following administration of three fourfold-escalating single-dose levels of NZ2114 were measured by microbiologic assay. Single-dose time-kill studies following doses of 10, 40, and 160 mg/kg of body weight demonstrated concentration-dependent killing over the dose range (0.5 to 3.7 log10 CFU/thigh) and prolonged postantibiotic effects (3 to 15 h) against both Staphylococcus aureus and Streptococcus pneumoniae. Mice had 106.3 to 106.8 CFU/thigh of strains of S. pneumoniae or S. aureus at the start of therapy when treated for 24 h with 0.625 to 160 mg/kg/day of NZ2114 fractionated for 4-, 6-, 12-, and 24-h dosing regimens. Nonlinear regression analysis was used to determine which PK/PD index best correlated with microbiologic efficacy. Efficacies of NZ2114 were similar among the dosing intervals (P = 0.99 to 1.0), and regression with the 24-h area under the concentration-time curve (AUC)/MIC index was strong (R2, 0.90) for both S. aureus and S. pneumoniae. The maximum concentration of drug in serum/MIC index regression was also strong for S. pneumoniae (R2, 0.96). Studies to identify the PD target for NZ2114 utilized eight S. pneumoniae and six S. aureus isolates and an every-6-h regimen of drug (0.156 to 160 mg/kg/day). Treatment against S. pneumoniae required approximately twofold-less drug for efficacy in relationship to the MIC than did treatment against S. aureus. The free drug 24-h AUCs/MICs necessary to produce a stasis effect were 12.3 ± 6.7 and 28.5 ± 11.1 for S. pneumoniae and S. aureus, respectively. The 24-h AUC/MIC associated with a 1-log killing endpoint was only 1.6-fold greater than that needed for stasis. Resistance to other antimicrobial classes did not impact the magnitude of the PD target required for efficacy. The PD target in this model should be considered in the design of clinical trials with this novel antibiotic.
Article
Full-text available
Bacillus subtilis encodes seven extracytoplasmic function (ECF) sigma factors. The sigma(W) regulon includes functions involved in detoxification and protection against antimicrobials, whereas sigma(M) is essential for growth at high salt concentrations. We now report that antibiotics that inhibit cell wall biosynthesis induce both sigma(W) and sigma(M) regulons as monitored using DNA microarrays. Induction of selected sigma(W)-dependent genes was confirmed using lacZ reporter fusions and Northern blot analysis. The ability of vancomycin to induce the sigma(W) regulon is dependent on both sigma(W) and the cognate anti-sigma, RsiW, but is independent of the transition state regulator AbrB. These results suggest that the membrane-localized RsiW anti-sigma(W) factor mediates the transcriptional response to cell wall stress. Our findings are consistent with the idea that one function of ECF sigma factors is to coordinate antibiosis stress responses and cell envelope homeostasis.
Article
It is generally assumed that type A lantibiotics primarily kill bacteria by permeabilization of the cytoplasmic membrane. As previous studies had demonstrated that nisin interacts with the membrane-bound peptidoglycan precursors lipid I and lipid II, we presumed that this interaction could play a role in the pore formation process of lantibiotics. Using a thin-layer chromatography system, we found that only nisin and epidermin, but not Pep5, can form a complex with [14C]-lipid II. Lipid II was then purified from Micrococcus luteus and incorporated into carboxyfluorescein-loaded liposomes made of phosphatidylcholine and cholesterol (1:1). Liposomes supplemented with 0.05 or 0.1 mol% of lipid II did not release any marker when treated with Pep5 or epilancin K7 (peptide concentrations of up to 5 mol% were tested). In contrast, as little as 0.01 mol% of epidermin and 0.1 mol% of nisin were sufficient to induce rapid marker release; phosphatidylglycerol-containing liposomes were even more susceptible. Controls with moenomycin-, undecaprenol- or dodecaprenolphosphate-doped liposomes demonstrated the specificity of the lantibiotics for lipid II. These results were correlated with intact cells in an in vivo model. M. luteus and Staphylococcus simulans were depleted of lipid II by preincubation with the lipopeptide ramoplanin and then tested for pore formation. When applied in concentrations below the minimal inhibitory concentration (MIC) and up to 5–10 times the MIC, the pore formation by nisin and epidermin was blocked; at higher concentrations of the lantibiotics the protective effect of ramoplanin disappeared. These results demonstrate that, in vitro and in vivo, lipid II serves as a docking molecule for nisin and epidermin, but not for Pep5 and epilancin K7, and thereby facilitates the formation of pores in the cytoplasmic membrane.
Article
Glycopeptide antibiotics, including vancomycin and teicoplanin, are large, rigid molecules that inhibit a late stage in bacterial cell wall peptidoglycan synthesis. The three-dimensional structure contains a cleft into which peptides of highly specific configuration (L-aa-D-aa-D-aa) can fit: such sequences are found only in bacterial cell walls, hence glycopeptides are selectively toxic. Glycopeptides interact with peptides of this conformation by hydrogen bonding, forming stable complexes. As a result of binding to L-aa-D-Ala-D-Ala groups in wall intermediates, glycopeptides inhibit, apparently by steric hindrance, the formation of the backbone glycan chains (catalysed by peptidoglycan polymerase) from the simple wall subunits as they are extruded through the cytoplasmic membrane. The subsequent transpeptidation reaction that imparts rigidity to the cell wall is also thus inhibited. This unique mechanism of action, involving binding of the bulky inhibitor to the substrate outside the membrane so that the active sites of two enzymes cannot align themselves correctly, renders the acquisition of resistance to the glycopeptide antibiotics more difficult than that to the majority of the other antibiotic groups.
Article
Prediction of small molecule binding modes to macromolecules of known three-dimensional structure is a problem of paramount importance in rational drug design (the "docking" problem). We report the development and validation of the program GOLD (Genetic Optimisation for Ligand Docking). GOLD is an automated ligand docking program that uses a genetic algorithm to explore the full range of ligand conformational flexibility with partial flexibility of the protein, and satisfies the fundamental requirement that the ligand must displace loosely bound water on binding. Numerous enhancements and modifications have been applied to the original technique resulting in a substantial increase in the reliability and the applicability of the algorithm. The advanced algorithm has been tested on a dataset of 100 complexes extracted from the Brookhaven Protein DataBank. When used to dock the ligand back into the binding site, GOLD achieved a 71% success rate in identifying the experimental binding mode.
Article
Multicellular organisms live, by and large, harmoniously with microbes. The cornea of the eye of an animal is almost always free of signs of infection. The insect flourishes without lymphocytes or antibodies. A plant seed germinates successfully in the midst of soil microbes. How is this accomplished? Both animals and plants possess potent, broad-spectrum antimicrobial peptides, which they use to fend off a wide range of microbes, including bacteria, fungi, viruses and protozoa. What sorts of molecules are they? How are they employed by animals in their defence? As our need for new antibiotics becomes more pressing, could we design anti-infective drugs based on the design principles these molecules teach us?
Article
The structure determination of protein-protein complexes is a rather tedious and lengthy process, by both NMR and X-ray crystallography. Several methods based on docking to study protein complexes have also been well developed over the past few years. Most of these approaches are not driven by experimental data but are based on a combination of energetics and shape complementarity. Here, we present an approach called HADDOCK (High Ambiguity Driven protein-protein Docking) that makes use of biochemical and/or biophysical interaction data such as chemical shift perturbation data resulting from NMR titration experiments or mutagenesis data. This information is introduced as Ambiguous Interaction Restraints (AIRs) to drive the docking process. An AIR is defined as an ambiguous distance between all residues shown to be involved in the interaction. The accuracy of our approach is demonstrated with three molecular complexes. For two of these complexes, for which both the complex and the free protein structures have been solved, NMR titration data were available. Mutagenesis data were used in the last example. In all cases, the best structures generated by HADDOCK, that is, the structures with the lowest intermolecular energies, were the closest to the published structure of the respective complexes (within 2.0 A backbone RMSD).
Article
In response to sublethal concentrations of antibiotics, bacteria often induce an adaptive response that can contribute to antibiotic resistance. We report the response of Bacillus subtilis to bacitracin, an inhibitor of cell wall biosynthesis found in its natural environment. Analysis of the global transcriptional profile of bacitracin-treated cells reveals a response orchestrated by two alternative sigma factors (sigmaB and sigmaM) and three two-component systems (YvqEC, YvcPQ and BceRS). All three two-component systems are located next to target genes that are strongly induced by bacitracin, and the corresponding histidine kinases share an unusual topology: they lack about 100 amino acids in their extracellular sensing domain, which is almost entirely buried in the cytoplasmic membrane. Sequence analysis indicates that this novel N-terminal sensing domain is a characteristic feature of a subfamily of histidine kinases, found almost entirely in Gram-positive bacteria and frequently linked to ABC transporters. A systematic mutational analysis of bacitracin-induced genes led to the identification of a new bacitracin-resistance determinant, bceAB, encoding a putative ABC transporter. The bcrC bacitracin resistance gene, which is under the dual control of sigmaX and sigmaM, was also induced by bacitracin. By comparing the bacitracin and the vancomycin stimulons, we can differentiate between loci induced specifically by bacitracin and those that are induced by multiple cell wall-active antibiotics.