ArticlePDF Available

Interaction of the Cyclic Antimicrobial Cationic Peptide Bactenecin with the Outer and Cytoplasmic Membrane

Authors:

Abstract and Figures

Bactenecin, a 12-amino acid cationic antimicrobial peptide from bovine neutrophils, has two cysteine residues, which form one disulfide bond, making it a cyclic molecule. To study the importance of the disulfide bond, a linear derivative Bac2S was made and the reduced form (linear bactenecin) was also included in this study. Circular dichroism spectroscopy showed that bactenecin existed as a type I beta-turn structure regardless of its environment, while the reduced form and linear bactenecin adopted different conformations according to the lipophilicity of the environment. Bactenecin was more active against the Gram-negative wild type bacteria Escherichia coli, Pseudomonas aeruginosa, and Salmonella typhimurium than its linear derivative and reduced form, while all three peptides were equally active against the outer membrane barrier-defective mutants of the first two bacteria. Only the two linear peptides showed activity against the Gram-positive bacteria Staphylococcus epidermidis and Enterococcus facaelis. Bactenecin interacted well with the outer membrane and its higher affinity for E. coli UB1005 lipopolysaccharide and improved ability to permeabilize the outer membrane seemed to account for its better antimicrobial activity against Gram-negative bacteria. The interaction of bactenecin with the cytoplasmic membrane was determined by its ability to dissipate the membrane potential by using the fluorescence probe 3, 3-dipropylthiacarbocyanine and an outer membrane barrier-defective mutant E. coli DC2. It was shown that the linear derivative and reduced form were able to dissipate the membrane potential at much lower concentrations than bactenecin despite the similar minimal inhibitory concentrations of all three against this barrier-defective mutant.
Content may be subject to copyright.
Manhong Wu and Robert E. W. Hancock
and Cytoplasmic Membrane
OuterCationic Peptide Bactenecin with the
Interaction of the Cyclic Antimicrobial
STRUCTURE:
PROTEIN CHEMISTRY AND
doi: 10.1074/jbc.274.1.29
1999, 274:29-35.J. Biol. Chem.
http://www.jbc.org/content/274/1/29Access the most updated version of this article at
.JBC Affinity SitesFind articles, minireviews, Reflections and Classics on similar topics on the
Alerts:
When a correction for this article is posted
When this article is cited
to choose from all of JBC's e-mail alertsClick here
http://www.jbc.org/content/274/1/29.full.html#ref-list-1
This article cites 32 references, 15 of which can be accessed free at
at Stanford University on October 14, 2013http://www.jbc.org/Downloaded from at Stanford University on October 14, 2013http://www.jbc.org/Downloaded from at Stanford University on October 14, 2013http://www.jbc.org/Downloaded from at Stanford University on October 14, 2013http://www.jbc.org/Downloaded from at Stanford University on October 14, 2013http://www.jbc.org/Downloaded from at Stanford University on October 14, 2013http://www.jbc.org/Downloaded from at Stanford University on October 14, 2013http://www.jbc.org/Downloaded from at Stanford University on October 14, 2013http://www.jbc.org/Downloaded from
Interaction of the Cyclic Antimicrobial Cationic Peptide Bactenecin
with the Outer and Cytoplasmic Membrane*
(Received for publication, April 24, 1998, and in revised form, July 19, 1998)
Manhong Wu‡ and Robert E. W. Hancock§
From the Department of Microbiology and Immunology, University of British Columbia, Vancouver,
British Columbia V6T 1Z3, Canada
Bactenecin, a 12-amino acid cationic antimicrobial
peptide from bovine neutrophils, has two cysteine resi-
dues, which form one disulfide bond, making it a cyclic
molecule. To study the importance of the disulfide bond,
a linear derivative Bac2S was made and the reduced
form (linear bactenecin) was also included in this study.
Circular dichroism spectroscopy showed that bactene-
cin existed as a type I
b
-turn structure regardless of its
environment, while the reduced form and linear bacte-
necin adopted different conformations according to the
lipophilicity of the environment. Bactenecin was more
active against the Gram-negative wild type bacteria
Escherichia coli, Pseudomonas aeruginosa, and Salmo-
nella typhimurium than its linear derivative and re-
duced form, while all three peptides were equally active
against the outer membrane barrier-defective mutants
of the first two bacteria. Only the two linear peptides
showed activity against the Gram-positive bacteria
Staphylococcus epidermidis and Enterococcus facaelis.
Bactenecin interacted well with the outer membrane
and its higher affinity for E. coli UB1005 lipopolysaccha-
ride and improved ability to permeabilize the outer
membrane seemed to account for its better antimicro-
bial activity against Gram-negative bacteria. The inter-
action of bactenecin with the cytoplasmic membrane
was determined by its ability to dissipate the membrane
potential by using the fluorescence probe 3,3-dipropy-
lthiacarbocyanine and an outer membrane barrier-de-
fective mutant E. coli DC2. It was shown that the linear
derivative and reduced form were able to dissipate the
membrane potential at much lower concentrations than
bactenecin despite the similar minimal inhibitory con-
centrations of all three against this barrier-defective
mutant.
Polycationic antimicrobial peptides have been found in a
variety of sources, including humans, mammals, plants, in-
sects, and bacteria (1). The primary structures of these posi-
tively charged molecules are highly diverse, yet their secondary
structures share the common feature of amphipathicity (2).
a
-Helical peptides, including cecropins (3) and
b
-sheet pep-
tides, including defensins (4), have been studied extensively. It
has been proposed (1, 2) that cationic peptides first interact
with bacteria by binding to their negatively charged surfaces,
and for Gram-negative bacteria they act as outer membrane
permeabilizers. Their interactions with the cytoplasmic mem-
brane have been proposed to lead to the disruption of mem-
brane structure (5), resulting in dissipation of the transmem-
brane potential (6) and eventual cell death.
Recently, a few cationic peptides with only one disulfide bond
forming a looped structure have been identified (7–11). One of
them, bactenecin (also called dodecapeptide), was found in
bovine neutrophils (12). It has 12 amino acids, including four
arginine residues and two cysteine residues and is the smallest
known cationic antimicrobial peptide. The two cysteine resi-
dues form a disulfide bond to make bactenecin a loop molecule.
Bactenecin was previously found to be active against Esche-
richia coli and Staphylococcus aureus, and strongly cytotoxic
for rat embryonic neurons, fetal rat astrocytes, and human
glioblastoma (13). However, little is known about its antimi-
crobial mechanism and whether it shares the common killing
mechanism of other antimicrobial peptides or if it has a distinct
mode of action due to its unique compact structure (cf. the silk
moth peptide cecropin, which is a 26-amino acid amphipathic
a
-helix). Its small size and only single disulfide bond also
makes bactenecin an interesting candidate for research and
drug development. The aim of this study was to investigate how
bactenecin interacts with and kills microoganisms. Interest-
ingly we found a rather distinct spectrum of activity for bacte-
necin compared with its linear form.
MATERIALS AND METHODS
Bacterial Strains and Chemicals—Bacterial strains for antimicrobial
activity testing included E. coli UB1005 and its antibiotic supersuscep-
tible derivative DC2 (14), Pseudomonas aeruginosa K799 and its anti-
biotic-supersusceptible derivative Z61 (15), Salmonella typhimurium
14028s (16), S. aureus ATCC25923, and clinical isolates of Staphylococ-
cus epidermidis (clinical isolate), Enterococcus faecalis ATCC29212,
and Listeria monocytogenes (food isolate).
Polymyxin B and 1-N-phenylnaphylamine (NPN)
1
were purchased
from Sigma. 3,3-Dipropylthiacarbocyanine (DiS-C
3
-(5)) was from Mo-
lecular Probes (Eugene, Oregon). Dansyl-polymyxin B was synthesized
as described previously (18). The lipids 1-pamitoyl-2-oleoyl-sn-glycero-
3-phosphocholine (POPC) and 1-palmitoyl-2-oleoyl-sn-glycero-3-phos-
phoglycerol (POPG) were purchased from Northern Lipids Inc. (Van-
couver, British Columbia, Canada).
Synthesis and Refolding of Bactenecin—Bactenecin and variants
bac2S were synthesized by Fmoc (N-(9-fluorenyl)methoxycarbonyl)
chemistry by the Nucleic Acid/Protein Service unit at the University of
British Columbia using an Applied Biosystems, Inc. (Foster City, CA)
model 431 peptide synthesizer. The purchased bactenecin variants were
in their fully reduced forms. After a series of trials to determine the
optimal strategy, the disulfide bond was formed by air-oxidation in 0.01
M Tris buffer at room temperature for 24 h. The concentration of
* This work was supported by the Canadian Bacterial Diseases Net-
work and Micrologix Biotech Inc. The costs of publication of this article
were defrayed in part by the payment of page charges. This article must
therefore be hereby marked advertisement in accordance with 18
U.S.C. Section 1734 solely to indicate this fact.
‡ Recipient of a British Columbia Science Council Graduate Research
Engineering and Technology studentship.
§ Medical Research Council Distinguished Scientist Award. To whom
correspondence should be addressed. Tel.: 604-822-2682; Fax: 604-822-
6041; E-mail: bob@cmdr.ubc.ca.
1
The abbreviations used are: NPN, 1-N-phenylnaphylamine; DiS-C3-
(5), 3,3-dipropylthiacarbocyanine; dansyl, 5-dimethylaminonaphtha-
lene-1-sulfonyl; POPC, 1-pamitoyl-2-oleoyl-sn-glycero-3-phosphocho-
line; POPG, 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoglycerol; MALDI,
matrix-assisted laser desorption/ionization; MIC(s), minimal inhibitory
concentration(s); LPS, lipopolysaccharide.
THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 274, No. 1, Issue of January 1, pp. 29–35, 1999
© 1999 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.
This paper is available on line at http://www.jbc.org 29
bactenecins was kept below 100
m
g/ml in the oxidation buffer to mini-
mize the formation of multimers. A reversed phase column Pep RPC
HR5/5 (Amersham Pharmacia Biotech; Quebec, Canada) was used to
purify the disulfide-bonded bactenecins from their multimer by-prod-
ucts. The column was equilibrated with 0.3% (v/v) aqueous trifluoro-
acetic acid and eluted with a gradient of acetonitrile in 0.3% trifluoro-
acetic acid at a flow rate of 0.7 ml/min. Peptide concentration was
determined by amino acid analysis. Matrix-assisted laser desorption/
ionization (MALDI)-mass spectrometry (for native bactenecin only) and
acid-urea polyacrylamide gel electrophoresis (19) were used to confirm
that the disulfide bond was properly formed and a pure product
obtained.
Circular Dichroism—A Jasco (Japan) J-720 spectropolarimeter was
used to measure the circular dichroism spectra (20). The data were
collected and analyzed by Jasco software. Liposomes POPC/POPG (7:3)
were prepared by the freeze-thaw method to produce multilamellar
vesicles as described previously (21), followed by extrusion through
0.1-
m
m double-stacked Nuclepore filters using an extruder device (Li-
pex Biomembranes, Vancouver, British Columbia, Canada), resulting
in unilamellar liposomes. Peptide at a final concentration of 50
m
M was
added to 100
m
M liposomes and incubated at room temperature for 10
min before the CD measurement.
Antimicrobial Activity—The minimal inhibitory concentration of
peptides was determined by a modified 2-fold microtiter broth dilution
method modified from that of Steinberg et al. (22). Using the classical
method (23), higher concentrations of peptides tend to precipitate in the
LB broth, thus the concentrations of peptides in the sequential wells are
not accurate. Also the peptides stick to the most readily available
(tissue-culture treated polystyrene) 96-well microtiter plates. Therefore
the 23 series of dilutions was performed in Eppendoff tubes (polypro-
pylene) before mixing with LB broth. Serial of 2-fold dilutions of pep-
tides ranging from 640 to 1.25
m
g/ml were made in 0.2% bovine serum
albumin, 0.01% acetic acid buffer in the Eppendoff tubes. Ten
m
l of each
concentration was added to each corresponding well of a 96-well micro-
titer plate (polypropylene cluster; Costar Corp., Cambridge, MA). Bac-
teria were grown overnight and diluted 10
25
into fresh LB broth or
Todd Hewitt broth for Streptococcus. LB medium contained 10 g/liter
tryptone and 5 g/liter yeast extract, with no salt. Todd Hewitt contained
500 g/liter beef heart infusion, 20 g/liter bacto-neopeptone, 2 g/liter
bacto-dextrose, 2 g/liter sodium chloride, 0.4 g/liter disodium phos-
phate, 2.5 g/liter sodium carbonate. One-hundred
m
l of broth containing
about 10
4
–10
5
colony-forming units/ml of tested bacteria was added to
each well. The plate was incubated at 37 °C overnight. The MIC was
taken as the concentration at which greater than 90% of growth inhi-
bition was observed.
Dansyl-Polymyxin B Displacement Assay—E. coli UB1005 LPS was
prepared according to the phenol-chloroform-petroleum ether extrac-
tion method (24). The dansyl-polymyxin B displacement assay (25) was
used to determine the relative binding affinity of peptides for LPS.
Membrane Permeabilization Assays—The ability of peptides to per-
meabilize the outer membrane was determined by the NPN assay of
Loh et al. (26). Cytoplasmic membrane permeabilization was deter-
mined by using the membrane potential sensitive cyanine dye DiS-C
3
-
(5) (27). The mutant E. coli DC2 with increased outer membrane per-
meability was used so that DiS-C
3
-(5) could reach the cytoplasmic
membrane. Fresh LB medium was inoculated with an overnight cul-
ture, grown at 37 °C, and mid-logarithmic phase cells (A
600
5 0.5–0.6)
were collected. The cells were washed with buffer (5 mM HEPES, pH
7.2, 5 mM glucose) once, then resuspended in the same buffer to an A
600
of 0.05. The cell suspension was incubated with 0.4
m
M DiS-C
3
-(5) until
DiS-C
3
-(5) uptake was maximal (as indicated by a stable reduction in
fluorescence due to fluorescence quenching as the dye became concen-
trated in the cell by the membrane potential), and 100 mM KCl was
added to equilibrate the cytoplasmic and external K
1
concentration.
One ml of cell culture was placed in a 1-cm cuvette, and the desired
concentration of tested peptide was added. The fluorescence reading
was monitored by using a Perkin-Elmer model 650–10S fluorescence
spectrophotometer (Perkin-Elmer Corp.), with an excitation wave-
length of 622 nm and an emission wavelength of 670 nm. The maximal
increase of fluorescence due to the disruption of the cytoplasmic mem-
brane by certain concentration of cationic peptide was recorded. A blank
with only cells and the dye was used to subtract the background.
Control experiments
2
titrating with valinomycin and K
1
showed that
the increase in fluorescence was directly proportional to the membrane
potential and that a buffer concentration of 100 mM KCl prevented any
effects of the high internal K
1
concentration and corresponding oppos-
ing chemical gradient.
RESULTS
Bactenecin and Its Linear Derivative—The amino acid se-
quence of bactenecin and its linear derivative are shown in
Table I. The linear derivative (Lin-Bac2S) with two cysteine
residues replaced by two serine residues, was made to deter-
mine the importance of the disulfide bond in bactenecin’s an-
timicrobial activity. The reduced form of bactenecin was also
included in this study as a linear version of bactenecin. The
identity of these peptides was confirmed by MALDI mass spec-
trometry. The MALDI data showed the molecular mass of the
reduced bactenecin as 1486 6 1 dalton and oxidized bactenecin
as 1484 6 1 dalton, in agreement with formation of one disul-
fide bond in the latter. Linear reduced bactenecin did not
reform its disulfide bonds spontaneously within the lifetime of
these experiments, as confirmed by its gel electrophoretic mo-
bility (which was altered by disulfide bond formation).
Circular Dichroism—CD spectrometry (Fig. 1A) showed that
linear, reduced bactenecin and linear Bac2S were present in 10
m
M sodium phosphate buffer as unordered structures, which
had a strong negative ellipticity near 200 nm. The CD spectrum
of native bactenecin (Fig. 1A) demonstrated a negative elliptic-
ity near 205 nm, typical of that seen for a type I
b
-turn struc-
ture (29) and resembling oxyribonuclease and nuclease, which
are short polypeptides with a disulfide bond (28). In 60% TFE
buffer, in the presence of liposomes and 10 m
M SDS, the native
bactenecin retained a similar structure (Fig. 1, BD). However,
the reduced form and Lin-Bac2S exhibited clearly distinct
structures from those observed in the aqueous solution. In 60%
trifluoroethanol (considered a helix-inducing solvent), these
two peptides tended to form an
a
-helical structure (Fig. 1B),
whereas and in the presence of liposomes or 10 m
M SDS (a
membrane-mimicking detergent), a
b
-sheet structure was evi-
dent (Fig. 1, C and D).
Antimicrobial Activity—The MIC of bactenecin and its deriv-
atives against a range of bacteria was determined (Table II) by
using a modified broth dilution method. Bactenecin was active
against all Gram-negative bacteria tested. It was relatively
inactive (MIC 5 64
m
g/ml) against the Gram-positive bacte-
2
M. Wu, E. Maier, R. Benz, and R. E. W. Hancock, submitted for
publication.
T
ABLE I
Amino acid sequences of bactenecin and its derivatives
Peptide classification Name Amino acid sequence
a
Size Net charge
Native Cyclic bactenecin RLCRIVVIRVCR 12 14
Linear Lin Bac
RLCRIVVIRVCR 12 14
Lin Bac2S
RLSRIVVIRVSR 12 14
Lin Bac2S-NH
2
RLSRIVVIRVSR-NH
2
12 15
Lin BacR
RRLCRIVVIRVCRR 14 16
Cyclic BacR
RRLCRIVVIRVCRR 14 16
BacP
RRCPIVVIRVCR 12 14
BacP-NH
2
RRCPIVVIRVCR-NH
2
12 15
BacP3K
KKCPIVVIRVCK 12 14
a
One-letter amino acid code. Cysteine residues linked together with disulphide bonds are underlined
Antimicrobial Mechanism of the Cyclic Cationic Peptide Bactenecin30
rium S. aureus, in contrast to a previous report (12). The linear
variant Lin-Bac2S and reduced bactenecin (Lin-Bac) were in-
active against wild type Gram-negative bacteria. P. aeruginosa
Z61 and E. coli DC2 are outer membrane barrier-defective
mutants, that have more permeable outer membranes than
their parent strains, allowing potentially easier access of the
peptides to the cytoplasmic membrane. All three bactenecins
exhibited equivalent activity against these two mutants. Lin-
earization by reduction or changing cysteine to serines dramat-
ically changed the antimicrobial activity for two Gram-positive
species S. epidermidis and Enterococcus facaelis. For other
antimicrobial peptides with disulfide bonds, reduction of these
disulfides generally results in complete loss of antimicrobial
activity (30–32). In contrast, accompanying the linearization
was a shift in spectrum of activity from Gram-negative selec-
tive to Gram-positive selective, which corresponded to the sub-
FIG.1.CD spectra of bactenecin, its linear (reduced) form Lin Bac and a linear variant Lin bac2S in media of various lipophilicity.
The concentrations of peptides and liposomes were 50 and 100
m
M, respectively. CD measurements were taken in 10 mM sodium phosphate buffer
(pH 7.0) in the absence (A) and the presence (B) of POPC/POPG. C shows the spectra in the presence of 60% (v/v) TFE, and D shows the spectra
in the presence of 10 m
M SDS. Open circles, bactenecin; solid line, reduced bactenecin; dashed line, Lin-Bac2S.
T
ABLE II
Differential activity of native cyclic and linear bactenecins against Gram-negative bacteria, outer membrane-altered, antibiotic supersusceptible
mutants DC2 and Z61, and selected Gram-positive bacteria
Species and strains Relevant phenotype
MIC
Bactenecin (oxidized) Lin Bac Lin Bac2S
m
g/ml
E. coli UB1005 Parent of DC2 8 .64 32
E. coli DC2 Antibiotic-sensitive 2 2 2
P. aeruginosa K799 Parent of Z61 4 .64 .64
P. aeruginosa Z61 Antibiotic-sensitive 0.5 0.5 0.5
S. typhimurium Wild type 8 .64 .64
S. epidermidis Clinical isolate .64 8 8
E. faecalis Wild type .64 8 8
Antimicrobial Mechanism of the Cyclic Cationic Peptide Bactenecin 31
stantially different structures adopted in liposomes.
The Binding of Bactenecins to Purified E. coli UB1005—The
MIC results indicated that the interaction with the outer mem-
brane might be critical in the explaining the difference in
antimicrobial activity against Gram-negative bacteria among
three bactenecin forms. The first step of cationic peptide anti-
microbial action has been shown to involve the binding of the
cationic peptide to the negatively charged surface of the target
cells (1). In Gram-negative bacteria, this initial interaction
occurs between the cationic peptides and the negatively
charged LPS in the outer membrane (20, 33, 34). Such binding
can be quantified using the dansyl-polymyxin B displacement
assay. Dansyl-polymyxin B is a fluorescently tagged cationic
lipopeptide, which is nonfluorescent in free solution, but fluo-
resces strongly when it binds to LPS. When the peptides bind to
LPS, they displace dansyl-polymyxin B, resulting in decreased
fluorescence, which can be assessed as a function of peptide
concentration (Fig. 2). Bactenecin was a relatively weak LPS
binder compared with polymyxin B and similar to the peptide
indolicidin (13 amino acids with a net charge of 12; Ref. 20),
but it was still better than Mg
21
, the native divalent cation
associated with LPS. Most importantly, it seemed that native
cyclic bactenecin bound to LPS far better than its linear deriv-
atives, which partially explained the difference in activities
against Gram-negative bacteria.
Effect on Outer Membrane Permeability—Antimicrobial pep-
tides bind to LPS, displacing the native divalent cations. Due to
their bulky nature they disrupt the outer membrane and self-
promote their own uptake across the outer membrane (33, 34).
In order to determine whether better binding ability resulted in
better outer membrane permeabilization, a NPN assay was
performed. NPN is a neutral hydrophobic probe that is ex-
cluded by an intact outer membrane, but is taken up into the
membrane interior of an outer membrane that is disrupted by
antimicrobial peptide action. NPN fluoresces weakly in free
solution but strongly when it enters the membrane. Fig. 3
showed that polymyxin B permeabilized the outer membrane
to a 50% of maximal increase in fluorescence arbitrary units at
0.4
m
g/ml, while bactenecin, Lin-Bac2S, and linear bactenecin
caused half-maximal permeabilization at 0.8, 2, and 4.5
m
g/ml,
respectively. Bactenecin was thus better than the linearized
derivatives at permeabilizing the outer membrane of E. coli
UB1005.
Effect on the Inner Membrane Potential Gradient—It has
been proposed that the antibacterial target of cationic peptides
is at the cytoplasmic membrane. Cationic peptides are gener-
ally able to interact electrostatically with the negatively
charged headgroups of bacterial phospholipids and then insert
into the cytoplasmic membrane, forming channels or pores that
are proposed to lead to the leakage of cell contents and cell
death. However there is very little data for peptides pertaining
to measurement of the disruption of the cytoplasmic membrane
permeability barrier, despite ample evidence that membrane
disruption can occur in model membrane systems (35). Al-
though, some authors have utilized measurements of the ac-
cessibility of a normally membrane-impermeable substrate to
cytoplasmic
b
-galactosidase, this assay suffers from using a
bulky substrate (ortho-nitrophenyl galactoside) (36, 37). To
circumvent this, we have developed an assay involving the
membrane potential-sensitive dye diS-C
3
-(5) to measure the
disruption of electrical potential gradients in intact bacteria.
The use of the E. coli mutant DC2 permitted us to perform this
assay in the absence of EDTA (required by previous workers
who have used similar assays in E. coli (38, 39)). The fluores-
cent probe diS-C
3
-(5), which is a caged cation, distributes be-
tween cells and medium depending on the cytoplasmic mem-
brane potential. Once it is inside the cells, it becomes
concentrated and self-quenches its own fluorescence. If pep-
tides form channels or otherwise disrupt the membrane, the
membrane potential will be dissipated, and the DiS-C
3
-(5) will
be released into the medium causing the fluorescence to in-
crease, as can be detected by fluorescent spectrometry. In these
assays, 0.1
M KCl was added to the buffer to balance the
chemical potential of K
1
inside and outside the cells. Therefore
the MICs of bactenecin, reduced bactenecin, and bac2S in the
FIG.2.Binding of peptides to LPS as assessed by their ability
to displace dansyl-polymyxin B from E. coli UB1005 LPS. Dansyl-
polymyxin B was added to 1 ml of 3
m
g/ml LPS to a final concentration
of 1
m
M, which saturated the binding sites on LPS, and the fluorescence
sensitivity was adjusted to 90%. The peptides and Mg
21
were titrated
in, resulting in a decrease in fluorescence due to the competitive dis-
placement of dansyl-polymyxin from the LPS, resulting in a reduction
in fluorescence. Symbols: triangles, cyclic bactenecin; squares, Lin-
Bac2S; closed circles, linear (reduced) bactenecin (Lin Bac), dashed line,
polymyxin B; dotted line, indolicidin (from Ref. 20); diamonds, MgCl
2
.
FIG.3.Peptide-induced outer membrane permeabilization as-
sessed by the NPN uptake in E. coli UB1005. Mid-log phase E. coli
cells were collected and incubated with NPN in the presence of various
concentrations of native cyclic bactenecin (oxidized), linear reduced
bactenecin (Lin Bac), and Lin Bac2S. NPN was taken up into cells when
the outer membrane was disrupted by the peptides. The uptake of NPN
was measured by the increase of fluorescence. Symbols: triangles, cyclic
bactenecin; squares, Lin-Bac2S; closed circles, linear (reduced) bactene-
cin; open circles, polymyxin B.
Antimicrobial Mechanism of the Cyclic Cationic Peptide Bactenecin32
presence of 0.1 M KCl were determined and shown to be 8–16
m
g/ml (i.e. 48-fold higher than in low salt). Despite these
similar MICs for the three peptides versus E. coli DC2, the
influence of these peptides on the membrane potential was
quite different (Fig. 4). The linear bactenecins at around their
MIC (8
m
g/ml) caused a rapid increase in fluorescence that was
similar to that seen for a control
a
-helical peptide CEMA at its
MIC of 1
m
g/ml. However despite a similar 30-s delay prior to
initiation, the kinetics were somewhat slower with CEMA
causing a maximal depolarization of the cytoplasmic mem-
brane (increase in fluorescence) within 2 min, whereas the
linear bactenecins caused only 50% maximal depolarization of
the cytoplasmic membrane in this period of time. In stark
contrast to both the linear bactenecins and CEMA, native cyclic
bactenecin at 8
m
g/ml caused a very modest depolarization
within the first 5 min (14% of that observed with reduced
bactenecin, reaching a maximum of 30% in 1 h).
Structure-Activity Relationships—A series of peptides re-
lated to bactenecin were made (Table I) in an attempt to deci-
pher important features of these peptides contributing to anti-
microbial activity. Included in this series were peptides that
differed in charge due either to amidation of the carboxyl ter-
minus (Lin-Bac2S-NH
2
and BacP-NH
2
) or addition of arginines
to the NH
2
and COOH terminus (Lin-BacR and cyclic BacR),
contained an added proline residue in the bactenecin ring to
promote cyclization (BacP), or contained a substitution of three
lysines for arginines (BacP3K). In total, four linear peptides
(denoted Lin-Bac for clarity) and five cyclic peptides were in-
vestigated. Antimicrobial activity was assessed for the bacteria
studied above in addition to two Gram-positive pathogens, S.
aureus ATCC25923 and a food isolate of L. monocytogenes
(Table III). The latter Gram-positive bacterium was reasonably
susceptible to cyclic bactenecin; however, the linear bactenecin
and Lin-Bac2S were 8-fold more active (but not particularly
active against S. aureus).
Among the linear peptides, an increase in positive charge
tended to result in increased activity against Gram-negative
bacteria for both Lin-Bac2S-NH
2
(14) and Lin-BacR (15).
However neither of these peptides had activities (except
against E. coli) equivalent to that of cyclic bactenecin. The
increase in positive charge of the peptides also tended to result
in an increase in activity against the Gram-positive bacteria
(cf. BacR versus Bac, Lin-Bac2S-NH
2
versus Lin-Bac2S, BacP-
NH
2
versus BacP). Clearly amidation of the carboxyl terminus
was very favorable to antimicrobial activity against both Gram-
negative and Gram-positive bacteria. Overall, good Gram-pos-
itive activity tended to require the peptides to be linear, al-
though the cyclic peptide BacP-NH
2
had reasonable activities
against the Gram-positive bacteria S. aureus and L.
monocytogenes.
The substitution of an arginine with a proline residue in the
ring structure in BacP (and moving the arginine in place of the
leucine residue at position 2) resulted in a loss of activity
against all bacteria except E. coli. This indicated that the
three-dimensional structure of the peptide was important,
since the net charge was identical to that of bactenecin, the
overall hydrophobicity very similar, and the substitution of Arg
for Leu in position 2 was not detrimental in BacR. Interest-
ingly, the further substitution of three arginines for three ly-
sine residues in BacP3K resulted in a very weakly active pep-
tide, suggesting that these two basic residues may not be
equally effective in promoting bactenecin activity.
DISCUSSION
The
a
-helical and
b
-structured classes are two groups of
antimicrobial polycationic peptides that have been well stud-
ied. Although their precise antimicrobial mechanism is some-
what unclear, it has been proposed that the outer and the
cytoplasmic membranes of Gram-negative bacteria are their
primary and final targets respectively (1). They have been
proposed to kill bacteria by first electrostatically interacting
with the surface of the bacterial cytoplasmic membrane (after
self-promoted uptake across the outer membrane for Gram-
negative bacteria). Then under the influence of a membrane
potential, they are proposed to insert into the membrane and
form channels to leak internal constituents. However much of
this mechanism is based on data from model membrane
studies.
Bactenecin belongs to a group of cationic peptides with only
one disulfide bond. In this study, it was shown that bactenecin
was active against the wild type Gram-negative bacteria E.
coli, P. aeruginosa, and S. typhimurium, whereas the linear
derivative and reduced form were virtually inactive against
these bacteria but had gained activity against certain Gram-
positive bacteria. For other disulfide-bonded peptides such as
the
b
-sheet defensins (30), the protegrins (31), and the tachy-
plesins (32), the loss of ability to form a disulfide bond results
in a complete loss of structure and activity. Thus the observa-
tion that bactenecin, when linearized, undergoes a dramatic
shift in activity spectrum (Table II) and in structure (Fig. 2) is
unprecedented and surprising.
Furthermore, cyclic bactenecin behaved in a fundamentally
different fashion to the linear bactenecins and the
a
-helical
28-amino acid peptide CEMA (34), with respect to cytoplasmic
membrane permeabilization (depolarization of the membrane
potential gradient). Previous studies used artificial liposomes
to study the interaction of antimicrobial peptides with mem-
branes. In this study, live cells of E. coli DC2, an outer mem-
brane hyperpermeable mutant, were used in conjunction with a
fluorescent dye, diS-C
3
-(5), which was released from cells when
the membrane potential is disrupted, leading to fluorescence
dequenching. Despite their equivalent MIC value against E.
coli DC2, the pattern of interaction of bactenecin and its linear
variants with the cytoplasmic membrane was quite different.
Whereas CEMA and linear bactenecin and Lin-Bac2S caused
rapid depolarization of the membrane, cyclic bactenecin caused
only a slow and minor change in membrane potential. Thus we
FIG.4.Peptide-induced inner membrane permeabilization as-
sessed by the diS-C
3
-(5) assay. Mid-log phase cells were collected and
resuspended in buffer (5 m
M HEPES, 5 mM glucose) to an A
600
of 0.05.
A 0.4
m
M final concentration of diS-C
3
-(5) was incubated with cell
suspensions until no more quenching was detected, then 0.1
M KCl was
added. The desired peptide concentration (8
m
g/ml for the bactenecins
and 1
m
g/ml for CEMA) was added to a 1-cm cuvette containing 1 ml of
cell suspension. The fluorescence change (in arbitrary units) was ob-
served as a function of time. Symbols: triangles, cyclic bactenecin;
squares, Lin-Bac2S; closed circles, linear (reduced) bactenecin; open
circles, CEMA.
Antimicrobial Mechanism of the Cyclic Cationic Peptide Bactenecin 33
conclude that cyclic bactenecin kills cells in a completely dif-
ferent way to the other antimicrobial peptides, which have
been proposed to act on the cytoplasmic membrane of bacteria.
Although the actual mechanism of killing was not investigated
in this study, we propose that bactenecin is able to cross the
cytoplasmic membrane of Gram-negative bacteria and act on a
target inside cells (e.g. negatively charged nucleic acids).
In contrast, it would appear that the linear bactenecins are
acting in the same way on the cytoplasmic membrane of bac-
teria as do other larger peptides. The linear variants dissipated
the cytoplasmic membrane potential at the MIC and showed
partial activity on membranes (data not shown), even at con-
centrations as low as 0.125
m
g/ml (less than 1% of the MIC).
However, while it is straightforward to imagine how 28-mer
peptides like CEMA might be able to span a biological mem-
brane to form a channel by a barrel-stave mechanism (1, 5, 6),
it is not so simple to understand how a 12-mer peptide contain-
ing 50% polar residues could span such a membrane.
Both the cyclic and linear versions of bactenecin, as well as
Bac2S, were equally active against outer membrane permeabil-
ity defective mutants of E. coli and P. aeruginosa. This obser-
vation indicated that the disulfide bond was important for
interaction with the outer membrane as confirmed here. Bacte-
necin had a better binding ability for LPS and also permeabi-
lized the outer membrane better, explaining its better activity
versus wild type Gram-negative bacteria. Computer modeling
of bactenecin with InsightII software (Biosym Technologies
Inc., San Diego, CA) indicated that bactenecin was a loop
molecule with a hydrophobic ring and a positively charged face
constructed from the COOH- and NH
2
-terminal portions of the
molecule (2). Such a conformation, which was consistent with
the CD spectral studies (Fig. 2) which indicated that bactenecin
existed as a rigid
b
-turn loop molecule regardless of its envi-
ronment, may make bactenecin a more amphipathic molecule
than the unstructured linear and reduced forms, which exist in
solution as random structures. This could explain why bacte-
necin interacted better with the negatively charged LPS than
its linear and reduced form. It is also worth mentioning that
bactenecin would also be too small to span the membrane and
form pores or channels unless a multimer is involved.
The original report of the isolation of bactenecin suggested
it was active against both E. coli and S. aureus (12), whereas
we demonstrated that cyclic bactenecin has very little activ-
ity against the latter bacterium. Therefore, we are tempted to
speculate that Romeo et al. (12) were working with a mixture
of linear and cyclic bactenecin or that their preparations were
partly or completely amidated (since amidation of two of our
peptides improved activity against S. aureus by 48-fold).
Unfortunately, despite two attempts to synthesize amidated
bactenecin, we were unable to obtain a preparation suffi-
ciently pure enough to permit identification of the desired
product.
Our studies of structure activity relationships revealed cer-
tain factors that were important in the activity of the linear
and cyclic bactenecins against bacteria. The most obvious cor-
relations observed were the improvement in activity against
Gram-negative bacteria with cyclization (due to disulfide bond
formation) and with increased positive charge. In addition
while cyclization tended to decrease activity against Gram-
positive bacteria, while increasing the positive charge by addi-
tion of two arginines or by amidation of the COOH-terminal
carboxyl, led to an improvement in activity against Gram-
positive bacteria. Despite the small size of these peptides, we
observed MICs against important bacterial pathogens that are
equal to or better than much larger peptides, and we suggest
that these peptides offer a potentially fruitful basis for isolation
of antibiotic peptides for clinical use.
REFERENCES
1. Hancock, R. E. W., Falla, T., and Brown, M. H. (1995) Adv. Microb. Physiol. 37,
135–175
2. Hancock, R. E. W. (1997) Lancet 349, 418422
3. Boman, H. G., and Hultmark, D. (1987) Annu. Rev. Microbiol. 41, 103–126
4. Westerhoff, H. V., Juretic, D., Hendler, R. W., and Zasloff, M. (1989) Proc. Natl.
Acad. Sci. U. S. A. 86, 6597–6601
5. Christensen, B., Fink, J., Merrifield, R. B., and Mauzerall, D. (1988) Proc.
Natl. Acad. Sci. U. S. A. 85, 5072–5076
6. Hill, C. P., Yee, J., Selsted, M. E., and Eisenberg, D. (1991) Science 251,
1481–1485
7. Morikawa, N., Hagiwara, K., and Nakajima, T. (1992) Biochim. Biophys. Res.
Commun. 189, 184–190
8. Simmaco, M., Mignogna, G., Barra, D., and Bossa, F. (1993) FEBS Lett. 324,
159–161
9. Clark, D. P., Durell, S., Maloy, W. L., and Zasloff, M. (1994) J. Biol. Chem. 269,
10849–10855
10. Suzuki, S., Ohe, Y., Okubo, T., Kakegawa, T., and Tatemoto, K. (1995)
Biochim. Biophys. Res. Commun. 212, 249–254
11. Fehlbaum, P., Bulet, P., Chernysh, S., Briand, J. P., Roussel J. P., Letellier, L.,
Hetru, C., and Hoffmann, J. A. (1996) Proc. Natl. Acad. Sci. U. S. A. 93,
1221–1225
12. Romeo, D., Skerlavaj, B., Bolognesi, M., and Gennaro, R. (1988) J. Biol. Chem.
263, 9573–9575
13. Radermacher, S. W., Schoop, V. M., and Schluesener, H. J. (1993) J. Neurosci.
Res. 36, 657–662
14. Richmond, M. G., Clarke, D. C., and Wotton, S. (1976) Antimicrob. Agents
Chemother. 10, 215–218
15. Angus, B. L., Carey, A. M., Caron, D. A., Kropinski, A. M. B., and Hancock,
R. E. W. (1982) Antimicrob. Agents Chemother. 21, 299–309
16. Fields, P. I., Groisman, E. A., and Heffron, F. (1989) Science 243, 1059–1062
17. Kreiswirth, B. N., Lofdahl, S., Bently, M. J., O’Reilly, M., Schlievert, P. M.,
Bergdoll, M. S., and Novick, R. P. (1983) Nature 305, 709–712
18. Schindler, P. R. G., and Teuber, M. (1975) Antimicrob. Agents Chemother. 8,
94–104
19. Spiker, S. (1980) Anal. Biochem. 108, 263–265
20. Falla, T. J., Karunaratne, D. N., and Hancock, R. E. W. (1996) J. Biol. Chem.
271, 19298–19303
21. Mayer, L. D., Hope, M. J., Cullis, P. R., and Janoff, A. S. (1985) Biochim.
Biophys. Acta 817, 193–196
22. Steinberg, D. A., Hurst, M. A., Fujii, C. A., Kung, A. H., Ho, J. F., Cheng, F. C.,
Loury, D. J., and Fiddes, J. C. (1997) Antimicrob. Agents Chemother. 41,
1738–1742
23. Amsterdam, D. (1991) in Antibiotics in Laboratory Medicine (Lorian, V., ed)
pp. 72–78, Williams and Wilkins, Baltimore
24. Moore, R. A., Bates, N. C., and Hancock, R. E. W. (1986) Antimicrob. Agents
chemother. 29, 496–500
25. Sprott, G. D., Koval, S. F., and Schnaitman, C. A. (1994) in Methods for
General Molecular Bacteriology (Gerhardt, P., ed) pp. 72–103, American
Society for Microbiology, Washington, D. C.
26. Loh, B., Grant, C., and Hancock, R. E. W. (1984) Antimicrob. Agents
Chemother. 19, 777–785
27. Sims, P. J., Waggoner, A. S., Wang, C. H., and Hoffman, J. F. (1974) Biochem-
istry 13, 3315–3329
28. Venyaminov, Y. S., and Yang, J. T. (1996) in Circular Dichroism and the
TABLE III
Structure-activity relationships amongst cyclic and linear bactenecins
Bacteria
MIC
Bac Lin-Bac Lin-Bac2S Lin-Bac2S-NH
2
Lin-BacR BacR BacP BacP-NH
2
BacP3K
m
g/ml
E. coli 8 .64 32 2 4 2 4 2 16
P. aeruginosa 4 .64 .64 16 16 4 .32 16 .64
S. typhimurium 8 .64 .64 32 32 4 16 16 32
S. aureus 64 .64 16 4 .64 64 32 4 .64
S. epidermidis .64 8 8 1 4 8 .32 32 .64
E. facaelis .64 8 8 4 4 32 .32 .64 .64
L. monocytogenes 8 1 1 0.25 0.5 0.125 16 1 16
Antimicrobial Mechanism of the Cyclic Cationic Peptide Bactenecin34
Conformational Analysis of Biomolecules (Fasman, G. D., ed) pp. 65–107,
Plenum Press, New York
29. Perczel, A., and Hollo´si, M. (1996) Circular Dichroism and the Conformational
Analysis of Biomolecules (Fasman, G. D., ed) pp. 285–380, Plenum Press,
New York
30. Ganz, T., Selsted, M. E., and Lehrer, R. I. (1987) in Bacteria-Host Cell
Interaction (Horowitz, M., and Lovett, M., eds) pp. 3–14, Alan B. Liss Inc.,
New York
31. Qu, X. D., Harwis, S. S., Schafer, W. M., and Lehrer, R. I. (1997) Infect. Immun.
65, 636639
32. Tamamura, H., Ikoma, R., Niwa, M., Funakoshi, S., and Fujii, N. (1993) Chem.
Pharm. Bull. 41, 975–980
33. Sawyer, J. G., Martin, N. L., and Hancock, R. E. W. (1988) Infect. Immun. 56,
693–698
34. Piers, K. L., and Hancock, R. E. W. (1994) Mol. Microbiol. 12, 951–958
35. Silvestro, L., Gupta K., Weiser, J. N., and Axelsen, P. H. (1997) Biochemistry
36, 11452–11460
36. Lehrer, R. I., Barton, A., Daher, K. A., Harwig, S. S. L., Ganz, T., and Selsted,
M. E. (1989) J. Clin. Invest. 84, 553–561
37. Skerlavaj, B., Romeo, D., and Gennaro, R. (1990) Infect. Immun. 58,
3724–3730
38. Letellier, L., and Shechter, E. (1979) Eur. J. Biochem. 102, 441– 447
39. Ghazi, A., Schechter E., Letellier L., and Labedan, B. (1981) FEBS Lett. 125,
197–199
Antimicrobial Mechanism of the Cyclic Cationic Peptide Bactenecin 35
... To understand the action mechanisms of a-MG against Staphylococcus species, FE-SEM was used to determine the morphological changes of bacteria treated with a-MG. Interestingly, ultrastructural alterations of bacteria treated with a-MG such as surface craters, surface indentation, cell disruption, and cellular debris were comparable to the results from S. aureus treated with antimicrobial peptides (Wu and Hancock, 1999;Tan et al., 2019;Deshmukh et al., 2021) or membrane acting antimicrobials such as daptomycin (Silverman et al., 2003), suggesting that a-MG targets the cytoplasmic membrane of Staphylococcus species. a-MG causes considerable damage to membrane integrity, whereas membrane acting antimicrobials produce ion channels or transmembrane pores (Wu and Hancock, 1999;Silverman et al., 2003;Cotroneo et al., 2008). ...
... Interestingly, ultrastructural alterations of bacteria treated with a-MG such as surface craters, surface indentation, cell disruption, and cellular debris were comparable to the results from S. aureus treated with antimicrobial peptides (Wu and Hancock, 1999;Tan et al., 2019;Deshmukh et al., 2021) or membrane acting antimicrobials such as daptomycin (Silverman et al., 2003), suggesting that a-MG targets the cytoplasmic membrane of Staphylococcus species. a-MG causes considerable damage to membrane integrity, whereas membrane acting antimicrobials produce ion channels or transmembrane pores (Wu and Hancock, 1999;Silverman et al., 2003;Cotroneo et al., 2008). Koh et al. (2013) also demonstrated that a-MG directly interacts with the cytoplasmic membrane of S. aureus via non-electrostatic interactions, causing bacterial lysis. ...
Article
Full-text available
Antimicrobial resistance in Staphylococcus species from companion animals is becoming increasingly prevalent worldwide. S. pseudintermedius is a leading cause of skin infections in companion animals. α-mangostin (α-MG) exhibits various pharmacological activities, including antimicrobial activity against G (+) bacteria. This study investigated the antimicrobial activity of α-MG against clinical isolates of Staphylococcus species from companion animals and assessed the therapeutic potential of α-MG in skin diseases induced by S. pseudintermedius in a murine model. Furthermore, the action mechanisms of α-MG against S. pseudintermedius were investigated. α-MG exhibited antimicrobial activity against clinical isolates of five different Staphylococcus species from skin diseases of companion animals in vitro, but not G (-) bacteria. α-MG specifically interacted with the major histocompatibility complex II analogous protein (MAP) domain-containing protein located in the cytoplasmic membrane of S. pseudintermedius via hydroxyl groups at C-3 and C-6. Pretreatment of S. pseudintermedius with anti-MAP domain-containing protein polyclonal serum significantly reduced the antimicrobial activity of α-MG. The sub-minimum inhibitory concentration of α-MG differentially regulated 194 genes, especially metabolic pathway and virulence determinants, in S. pseudintermedius. α-MG in pluronic lecithin organogel significantly reduced the bacterial number, partially restored the epidermal barrier, and suppressed the expression of cytokine genes associated with pro-inflammatory, Th1, Th2, and Th17 in skin lesions induced by S. pseudintermedius in a murine model. Thus, α-MG is a potential therapeutic candidate for treating skin diseases caused by Staphylococcus species in companion animals.
... F. solani (10 4 conidia) were left untreated or incubated with S100A12 (25 μM) for different time points, washed with 1× PBS, and incubated with 250 nM of 3,3 0 -dipropylthiadicarbocyanine iodide (DiSc 3 -5; Sigma-Aldrich) for 30 min in dark followed by addition of 100 mM KCl as described earlier (31). The membrane potential was determined by flow cytometry (Beckman Coulter). ...
Article
Full-text available
Fungal keratitis is the foremost cause of corneal infections worldwide, of which Fusariumspp. is the common etiological agent that causes loss of vision and warrants surgical intervention. An increase in resistance to the available drugs along with severe side effects of the existing antifungals demands for new effective antimycotics. Here, we demonstrate that antimicrobial peptide S100A12 directly binds to the phospholipids of the fungal membrane, disrupts the structural integrity, and induces generation of reactive oxygen species in fungus. In addition, it inhibits biofilm formation by Fusariumspp. and exhibits antifungal property against Fusariumspp. both in vitro and in vivo. Taken together, our results delve into specific effect of S100A12 against Fusariumspp. with an aim to investigate new antifungal compounds to combat fungal keratitis.
... The membrane depolarization assessment utilized the cationic dye, 3,3 ′ -dipropylthiadi carbocyanine iodide (DiSC 3 (5)), following a modified version of our previously established method [24,55,56,65,91]. In brief, AB5075 bacterial colonies were cultured overnight on Tryptic Soy agar, and opaque colonies were dispersed in Dulbecco's Phosphate-Buffered Saline (DPBS, Gibco) until reaching an optical density equivalent to 0.5 on the McFarland standard (~1 × 10 8 CFU/mL). ...
Article
Full-text available
Acinetobacter baumannii is a gram-negative bacterium that causes hospital-acquired and opportunistic infections, resulting in pneumonia, sepsis, and severe wound infections that can be difficult to treat due to antimicrobial resistance and the formation of biofilms. There is an urgent need to develop novel antimicrobials to tackle the rapid increase in antimicrobial resistance, and antimicrobial peptides (AMPs) represent an additional class of potential agents with direct antimicrobial and/or host-defense activating activities. In this study, we present GATR-3, a synthetic, designed AMP that was modified from a cryptic peptide discovered in American alligator, as our lead peptide to target multidrug-resistant (MDR) A. baumannii. Antimicrobial susceptibility testing and antibiofilm assays were performed to assess GATR-3 against a panel of 8 MDR A. baumannii strains, including AB5075 and some clinical strains. The GATR-3 mechanism of action was determined to be via loss of membrane integrity as measured by DiSC3(5) and ethidium bromide assays. GATR-3 exhibited potent antimicrobial activity against all tested multidrug-resistant A. baumannii strains with rapid killing. Biofilms are difficult to treat and eradicate. Excitingly, GATR-3 inhibited biofilm formation and, more importantly, eradicated preformed biofilms of MDR A. baumannii AB5075, as evidenced by MBEC assays and scanning electron micrographs. GATR3 did not induce resistance in MDR A. baumannii, unlike colistin. Additionally, the toxicity of GATR-3 was evaluated using human red blood cells, HepG2 cells, and waxworms using hemolysis and MTT assays. GATR-3 demonstrated little to no cytotoxicity against HepG2 and red blood cells, even at 100 μg/mL. GATR-3 injection showed little toxicity in the waxworm model, resulting in a 90% survival rate. The therapeutic index of GATR-3 was estimated (based on the HC50/MIC against human RBCs) to be 1250. Overall, GATR-3 is a promising candidate to advance to preclinical testing to potentially treat MDR A. baumannii infections.
... The same is true for L5K5W, so it is assumed that the modified version L5K5W-W4I6 acts identically to the original peptide (Kang et al., 2009;Kim et al., 2013). Bactenecin, which is linear due to reducing conditions of the medium, also has a membranolytic effect, and barrel stave is assumed in this context (Wu & Hancock, 1999). All tested peptides showed a significant loss of activity regarding biofilm inhibition compared to their effectiveness against planktonic cells. ...
Article
Full-text available
In medical, environmental, and industrial processes, the accumulation of bacteria in biofilms can disrupt many processes. Antimicrobial peptides (AMPs) are receiving increasing attention in the development of new substances to avoid or reduce biofilm formation. There is a lack of parallel testing of the effect against biofilms in this area, as well as in the testing of other antibiofilm agents. In this paper, a high‐throughput screening was developed for the analysis of the antibiofilm activity of AMPs, differentiated into inhibition and removal of a biofilm. The sulfate‐reducing bacterium Desulfovibrio vulgaris was used as a model organism. D. vulgaris represents an undesirable bacterium, which is considered one of the major triggers of microbiologically influenced corrosion. The application of a 96‐well plate and steel rivets as a growth surface realizes real‐life conditions and at the same time establishes a flexible, simple, fast, and cost‐effective assay. All peptides tested in this study demonstrated antibiofilm activity, although these peptides should be individually selected depending on the addressed aim. For biofilm inhibition, the peptide DASamP1 is the most suitable, with a sustained effect for up to 21 days. The preferred peptides for biofilm removal are S6L3‐33, in regard to bacteria reduction, and Bactenecin, regarding total biomass reduction.
... not naturally existing) designed from an ARE isoform (Yang et al., 2017b): the synthetic fragment containing disulfide bridges were not affected by pH variations (4.0 to 10.0) and more thermostable (range 20-80°C) than their respective linear fragments (devoid of disulfide bridges) which also exhibited lowest antibacterial activities (Yang et al., 2017a). By ensuring the β-hairpin structure, the disulfide bridges confer both a higher stability and a much more efficient bactericidal activity to ALV and ARE (Andrä et al., 2009;Kang et al., 2007;Lai et al., 2002;Nan et al., 2012;Wu et al., 1999). ...
Article
Antimicrobial peptides (AMPs) play a key role in the external immunity of animals, offering an interesting model for studying the influence of the environment on the diversification and evolution of immune effectors. Alvinellacin (ALV), arenicin (ARE) and polaricin (POL, a novel AMP identified here), characterized from three marine worms inhabiting contrasted habitats (‘hot’ vents, temperate and polar respectively), possess a well conserved BRICHOS domain in their precursor molecule despite a profound amino acid and structural diversification of the C-terminal part containing the core peptide. Data not only showed that ARE, ALV and POL display an optimal bactericidal activity against the bacteria typical of the habitat where each worm species lives but also that this killing efficacy is optimal under the thermochemical conditions encountered by their producers in their environment. Moreover, the correlation between species habitat and the cysteine contents of POL, ARE and ALV led us to investigate the importance of disulfide bridges in their biological efficacy as a function of abiotic pressures (pH and temperature). The construction of variants using non-proteinogenic residues instead of cysteines (α-aminobutyric acid variants) leading to AMPs devoid of disulfide bridges, provided evidence that the disulfide pattern of the three AMPs allows for a better bactericidal activity and suggests an adaptive way to sustain the fluctuations of the worm’s environment. This work shows that the external immune effectors exemplified here by BRICHOS AMPs are evolving under strong diversifying environmental pressures to be structurally shaped and more efficient/specific under the ecological niche of their producer.
Preprint
Antimicrobial peptides (AMPs) are attractive candidates to combat antibiotic resistance for their capability to target bio-membranes and restrict a wide range of pathogens. It is a daunting challenge to discover novel AMPs due to their sparse distributions in a vast peptide universe, especially for peptides that demonstrate potencies for both bacterial membranes and viral envelopes. Here we establish a de novo AMP design framework by bridging a deep generative module and a graph-encoding activity regressor. The generative module learns hidden ‘grammars’ of AMP features and produces candidates sequentially pass antimicrobial predictor and antiviral classifiers. We discover three bifunctional AMPs and experimentally validated their abilities to inhibit a spectrum of pathogens in vitro and in animal models. Notably, P076 is a highly potent bactericide with the minimal inhibitory concentration of 0.21 μM against multidrug-resistant A. baumannii , while P002 broadly inhibits five enveloped viruses. Our study provides feasible means to uncover sequences that simultaneously encode antimicrobial and antiviral activities, thus bolstering the function spectra of AMPs to combat a wide range of drug-resistant infections.
Article
Full-text available
The outer membrane in Gram-negative bacteria consists of an asymmetric phospholipid—lipopolysaccharide bilayer that is densely packed with outer-membrane β-barrel proteins (OMPs) and lipoproteins¹. The architecture and composition of this bilayer is closely monitored and is essential to cell integrity and survival2–4. Here we find that SlyB, a lipoprotein in the PhoPQ stress regulon, forms stable stress-induced complexes with the outer-membrane proteome. SlyB comprises a 10 kDa periplasmic β-sandwich domain and a glycine zipper domain that forms a transmembrane α-helical hairpin with discrete phospholipid- and lipopolysaccharide-binding sites. After loss in lipid asymmetry, SlyB oligomerizes into ring-shaped transmembrane complexes that encapsulate β-barrel proteins into lipid nanodomains of variable size. We find that the formation of SlyB nanodomains is essential during lipopolysaccharide destabilization by antimicrobial peptides or acute cation shortage, conditions that result in a loss of OMPs and compromised outer-membrane barrier function in the absence of a functional SlyB. Our data reveal that SlyB is a compartmentalizing transmembrane guard protein that is involved in cell-envelope proteostasis and integrity, and suggest that SlyB represents a larger family of broadly conserved lipoproteins with 2TM glycine zipper domains with the ability to form lipid nanodomains.
Article
Full-text available
The rise in antibiotic-resistant strains of clinically important pathogens is a major threat to global health. The World Health Organization (WHO) has recognized the urgent need to develop alternative treatments to address the growing list of priority pathogens. Antimicrobial peptides (AMPs) rank among the suggested options with proven activity and high potential to be developed into effective drugs. Many AMPs are naturally produced by living organisms protecting the host against pathogens as a part of their innate immunity. Mechanisms associated with AMP actions include cell membrane disruption, cell wall weakening, protein synthesis inhibition, and interference in nucleic acid dynamics, inducing apoptosis and necrosis. Acinetobacter baumannii is a critical pathogen, as severe clinical implications have developed from isolates resistant to current antibiotic treatments and conventional control procedures, such as UV light, disinfectants, and drying. Here, we review the natural AMPs representing primary candidates for new anti-A. baumannii drugs in post-antibiotic-era and present computational tools to develop the next generation of AMPs with greater microbicidal activity and reduced toxicity.
Article
Full-text available
Antimicrobial peptides (AMPs) are essential elements of thehost defense system. Characterized by heterogenous structures and broad‐spectrumaction, they are promising candidates for combating multidrug resistance. Thecombined use of AMPs with other antimicrobial agents provides a new arsenal ofdrugs with synergistic action, thereby overcoming the drawback of monotherapiesduring infections. AMPs kill microbes via pore formation, thus inhibitingintracellular functions. This mechanism of action by AMPs is an advantage overantibiotics as it hinders the development of drug resistance. The synergisticeffect of AMPs will allow the repurposing of conventional antimicrobials andenhance their clinical outcomes, reduce toxicity, and, most significantly,prevent the development of resistance. In this review, various synergies ofAMPs with antimicrobials and miscellaneous agents are discussed. The effect ofstructural diversity and chemical modification on AMP properties is firstaddressed and then different combinations that can lead to synergistic action,whether this combination is between AMPs and antimicrobials, or AMPs andmiscellaneous compounds, are attended. This review can serve as guidance whenredesigning and repurposing the use of AMPs in combination with other antimicrobialagents for enhanced clinical outcomes.
Article
Full-text available
Antimicrobial peptides comprise a diverse class of molecules used in host defense by plants, insects, and animals. In this study we have isolated a novel antimicrobial peptide from the skin of the bullfrog, Rana catesbeiana. This 20 amino acid peptide, which we have termed Ranalexin, has the amino acid sequence: NH2-Phe-Leu-Gly-Gly-Leu-Ile-Lys-Ile-Val-Pro-Ala-Met-Ile-Cys-Ala-Val-Thr- Lys-Lys - Cys-COOH, and it contains a single intramolecular disulfide bond which forms a heptapeptide ring within the molecule. Structurally, Ranalexin resembles the bacterial antibiotic, polymyxin, which contains a similar heptapeptide ring. We have also cloned the cDNA for Ranalexin from a metamorphic R. catesbeiana tadpole cDNA library. Based on the cDNA sequence, it appears that Ranalexin is initially synthesized as a propeptide with a putative signal sequence and an acidic amino acid-rich region at its amino-terminal end. Interestingly, the putative signal sequence of the Ranalexin cDNA is strikingly similar to the signal sequence of opioid peptide precursors isolated from the skin of the South American frogs Phyllomedusa sauvagei and Phyllomedusa bicolor. Northern blot analysis and in situ hybridization experiments demonstrated that Ranalexin mRNA is first expressed in R. catesbeiana skin at metamorphosis and continues to be expressed into adulthood.
Article
Full-text available
Cytoplasmic granules of neutrophils store a variety of cationic polypeptides, which exert in vitro a potent antibacterial action and are potentially involved in host defense mechanisms. From an acid extract of bovine neutrophil granules we have purified over 2,000-fold a dodecapeptide exhibiting bactericidal activity against both Escherichia coli and Staphylococcus aureus at 10(-7)-10(-5) M concentration. The purification procedure involved only two steps of ion-exchange and reversed-phase chromatography. The peptide, named bactenecin, has the amino acid sequence, Arg-Leu-Cys-Arg-Ile-Val-Val-Ile-Arg-Val-Cys-Arg, maintained in a cyclic structure by a disulfide bond between the two cysteine residues. Computer modeling of the dodecapeptide resulted in a conformation in which the chain adopts an antiparallel extended structure forming a gamma turn at residue 7.
Article
Full-text available
Bactenecins are a class of arginine-rich antibacterial peptides of bovine neutrophil granules. Two bactenecins with approximate molecular weights of 5,000 and 7,000 designated Bac5 and Bac7, respectively, exert in vitro a potent bactericidal activity toward several gram-negative bacteria (R. Gennaro, B. Skerlavaj, and D. Romeo, Infect. Immun. 57:3142-3146, 1989). We have now found that this activity shows an inverse relationship to the ionic strength of the medium and is inhibited by divalent cations and greatly potentiated by lactoferrin. Under conditions supporting marked bactericidal activity, the two peptides cause a rapid increase in the permeability of both the outer and inner membranes of Escherichia coli, as shown by unmasking of periplasmic beta-lactamase and of cytoplasmic beta-galactosidase. In addition, the two bactenecins inhibit the respiration of E. coli and Klebsiella pneumoniae but not of Bac5- and Bac7-resistant Staphylococcus aureus. Furthermore, they induce a drop in ATP content in E. coli, K. pneumoniae, and Salmonella typhimurium and a marked decrease in the rates of transport and incorporation of [3H]leucine and [3H]uridine into E. coli protein and RNA, respectively. In general, all these effects become evident within 1 to 2 min and reach their maximal expression within about 5 min. Overall, these data strongly suggest that the decrease in bacterial viability is causally related to the increase in membrane permeability and the subsequent fall in respiration-linked proton motive force, with the attendant loss of cellular metabolites and macromolecular biosynthesis ability.
Chapter
β and γ turns have been recognized as forming an important group of regular or ordered secondary structures of proteins. β turns are sites where the polypeptide chain reverses its overall direction. Using a hard-sphere model-building technique, Venkatachalam (1968) explored favorable H-bonded conformations of the three consecutive amide units of β turns (see Chou and Fasman, 1977, and references therein). The x-ray structural analysis of an increasing number of proteins (Richardson, 1981; Kabsch and Sander, 1983) has revealed that β turns are common in proteins, accounting for 25–30% of the residues of the total molecule. Lewis et al. (1973) found that about one-fourth of β turns do not possess the H-bond stipulated by Venkatachalam.
Article
The fluorescence response of a positively charged cyanine dye: 3,3'-dimethylindodicarbocyanine iodide can be specifically related to the generation in Escherichia coli cells and E. coli membrane vesicles of an electrical membrane potential induced either by substrate oxidation or by an artificially imposed potassium diffusion gradient. The energy-dependent quenching of the dye fluorescence correlates well with the known effect on delta phi of: oxidation of various energy sources, external pH and solute accumulation. Thus, in the vesicles, the fluorescence quenching of the dye increases from succinate to D-lactate, to ascorbate/phenazine methosulfate and parallels the increasing ability of these electron donors to generate a delta phi. In the vesicles, delta phi is only weakly dependent on external pH, whereas in the cells, delta phi increases with increasing external pH. Lactose accumulation in the vesicles results in the partial utilization of delta phi. A calibration of the dye fluorescence in terms of delta phi has been determined using valinomycin-induced potassium diffusion potential.
Article
Though the primary action of the cationic antibiotic polymyxin B is against the membrane of susceptible bacteria, severe morphological changes are detected in the cytoplasm. Using fluorescence microscopy and a mono-N-dansyl-polymyxin B derivative, we could demonstrate aggregations of the antibiotic with cellular material, possibly nucleic acids and/or ribosomes. These aggregations were only produced by minimum inhibitory or higher concentrations of the antibiotic as shown with Salmonella and Escherichia strains differing in their polymyxin susceptibility. The outer membrane of Salmonella typhimurium revealed characteristic blebs when treated with polymyxin B. This was investigated by the gentle methods of spray-freezing and freeze-etching. The obtained electron micrographs suggest that the polymyxin-induced blebs are projections of the outer monolayer of the outer membrane. A possible mechanism of penetration of polymyxin B through the cell envelope of gram-negative bacteria is presented.
Article
Escherichia coli UB1005 and two mutants of this strain (DC2 and DC3) have been used to assess indirectly the relative ability of various β-lactam antibiotics to penetrate the outer layers of E. coli. Benzylpenicillin, ampicillin, methicillin, cloxacillin, cephaloridine, cephalothin, and cephalexin have been examined. The results confirm those obtained with other methods and show that, among the compounds studied here, cephalosporins seem to penetrate more readily than penicillins.
Article
Two unique antimicrobial peptides named brevinin-1 and -2 were isolated from the skin of the frog, Rana brevipoda porsa. Both of the peptides did not have any structural homology with bombinin nor magainin; the frog skin derived-antimicrobial peptides isolated from Bombina and Xenopus, nor even with other known antimicrobial peptides of non-amphibian origin. The minimum inhibitory concentration of brevinin-1 against the growth of St. aureus and E. coli was determined to be 8 micrograms/ml and 34 micrograms/ml while that of brevinin-2 was 8 micrograms/ml and 4 micrograms/ml, respectively, indicating the difference of the two peptides in the antimicrobial selectively on Gram-positive and Gram-negative bacteria.
Article
Defensins (molecular weight 3500 to 4000) act in the mammalian immune response by permeabilizing the plasma membranes of a broad spectrum of target organisms, including bacteria, fungi, and enveloped viruses. The high-resolution crystal structure of defensin HNP-3 (1.9 angstrom resolution, R factor 0.19) reveals a dimeric beta sheet that has an architecture very different from other lytic peptides. The dimeric assembly suggests mechanisms by which defensins might bind to and permeabilize the lipid bilayer.