Subscriber access provided by NEW YORK MED COLL
Article
Structural and antimicrobial features of peptides related to myticin C, a special defense molecule from the Mediterranean mussel Mytilus galloprovincialis Stefania Domeneghetti, Marco Franzoi, Nunzio Damiano, Rosa Norante, Nancy M. El Halfawy, Stefano Mammi, Oriano Marin, Massimo Bellanda, and Paola Venier J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.5b03491 • Publication Date (Web): 07 Oct 2015 Downloaded from http://pubs.acs.org on October 7, 2015
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
Journal of Agricultural and Food Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 32
Journal of Agricultural and Food Chemistry
1
Structural and antimicrobial features of peptides related to myticin C, a special defense
2
molecule from the Mediterranean mussel Mytilus galloprovincialis
3 4
Stefania Domeneghetti1§, Marco Franzoi1§, Nunzio Damiano2§, Rosa Norante3, Nancy M. El
5
Halfawy4, Stefano Mammi5, Oriano Marin2,3, Massimo Bellanda5, Paola Venier1*
6 7
1
Dept. of Biology, University of Padova, Via Ugo Bassi 58/B, 35131, Padova, Italy
8
2
CRIBI Biotechnology Centre, Via Ugo Bassi 58/B, 35131 Padua, Italy
9
3
Dept. of Biomedical Sciences, University of Padua, Via Ugo Bassi 58/B, 35131, Padova, Italy
10
4
Dept. of Botany and Microbiology, Alexandria University, Moharam Bey 21511, 21526
11
Alexandria, Egypt
12
5
Dept. of Chemical Sciences, University of Padova, Via Marzolo 1, 35131, Padova, Italy
13 14
§These authors equally contributed to the work
15
*Corresponding author (e-mail:
[email protected]; tel: 0039-0498276284)
16
1 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 2 of 32
17
ABSTRACT
18
Mussels (Mytilus spp.) have a large repertoire of cysteine-stabilized alphabeta peptides and myticin
19
C (MytC) was identified in some hundreds of transcript variants after in vivo immunostimulation.
20
Using a sequence expressed in Italian mussels, we computed the MytC structure and synthesized
21
the mature MytC and related peptide fragments (some of them also prepared in oxidized form) to
22
accurately assess their antibacterial and antifungal activity. Only when tested at pH 5, the reduced
23
MytC as well as reduced and oxidized fragments including structural β-elements were able to
24
inhibit Gram positive and negative bacteria (MIC range: 4-32 and 8-32 µM, respectively). Such
25
fragments caused selective E. coli killing (MBC: 8-32 µM) but scarcely inhibited two fungal
26
strains. In detail, the antimicrobial β-hairpin MytC[19-40]SOX, caused membrane-disrupting effects
27
in E. coli despite its partially ordered conformation in membrane-mimetic environments. In
28
perspective, MytC-derived peptides could be employed to protect acidic mucosal tissues, cosmetic
29
and food products and, possibly, used as adjuvants in aquaculture.
30
KEY WORDS
31
Mytilus galloprovincialis, mussel, antimicrobial peptides, myticin C
32
INTRODUCTION
33
Cationic and cysteine-rich antimicrobial peptides (AMPs) play a central role in the innate immunity
34
of marine bivalves of the genus Mytilus. Arthropod-like defensins, mytilins, mytimycins and
35
myticins were originally isolated from acidified plasma fractions displaying antimicrobial activity
36
by high performance liquid chromatography (HPLC)1-3. The AA sequences of native myticin A
37
(MytA) and myticin B (MytB) were firstly obtained by mass spectrometry and Edman degradation
38
and then confirmed by cDNA cloning and sequencing: their deduced molecular mass was consistent
39
with that of the purified peptides assuming the formation of four intramolecular disulfide bridges3.
40
Though differing in 23 out of 96 amino acids, the MytA and MytB precursors are similarly
41
organized in a highly conserved 20 AA signal peptide, a C-terminal extension of 36 AA, and a 2 ACS Paragon Plus Environment
Page 3 of 32
Journal of Agricultural and Food Chemistry
42
central 40 AA mature peptide which includes 8 invariant cysteines likely forming a canonical
43
cysteine-stabilized alpha beta (CSαβ) motif, as in mussel mytilins and defensins. Actually, these
44
AMP families diverge in primary sequence from one another and are unique to the Mytilus species4-
45
5
46
Database of NCBI (pfam10690; http://www.ncbi.nlm.nih.gov/Structure/cdd/cdd.shtml).
47
The abundance of AMPs in hemocytes of adult mussels indicated these cells as the main production
48
and storage site for both precursor and mature peptides3. While the anionic character of the C-
49
terminal extension of mussel myticins may stabilize the positively charged central region within
50
intracellular storage granules, the mature peptides are expected to exert their microbicidal role in
51
pathogen-containing phagolysosomes, when released in the hemolymph and intercellular space or
52
after massive cell degranulation4. The occurrence of MytA and MytB transcripts in tissues other
53
than hemolymph reasonably depends on the circulating and infiltrating ability of hemocytes albeit
54
other cells may play a role, such as mussel enterocytes in the case of mytilin B3,5.
55
Myticin C (MytC) transcripts were later recognized in hemocytes of immunostimulated mussels as
56
multiple sequence variants clustering apart from those of the already known MytA and MytB6. The
57
deduced MytC pre-pro peptide showed similar organization, except for a slightly longer C-terminal
58
extension of 40 AA (coding sequence: 100 AA). The virtual translation of 300 transcript variants
59
denoting MytC in GenBank (e.g., AM498001.1, AM498018.1, AM498038.1, AM497990.1)
60
suggested a molecular weight of 4.4 kDa, an isoelectric point at pH 8.8, and a 35% hydrophobicity
61
ratio.
62
High throughput amplicon sequencing revealed synonymous and non-synonymous single
63
nucleotide changes along the whole MytC precursor transcript of farmed mussels from three
64
European countries7. Such molecular diversity is consistent with the rapid and adaptive evolution of
65
peptides actively recruited to target infectious agents evolving more quickly8-10. Actually, myticin
66
precursor transcripts were the most abundant (20.2%) in hemocytes of adult mussels injected with
. The molecular hallmarks of myticin pre-pro peptides are described in the Conserved Domain
3 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 4 of 32
67
heat-killed bacteria or poly I:C compared with mytilins (14.3%), defensins (1.2%) and very rare
68
mytimycins6,11.
69
Other cysteine-rich AMPs were subsequently identified in mussel: big defensins and mytimacins12,
70
and an unusual myticusin-1 identified in M. coruscus was traced in M. galloprovincialis13. Seven
71
sequences expressed in Mytilus coruscus and highly similar to the M. galloprovincialis myticins are
72
present in GenBank (GU324724.1-GU324731.1, unpublished). The variety of mussel cysteine-rich
73
peptides is currently under study and their functional roles have still to be disentangled14.
74
Consistent with the onset of immune competence during development, MytC transcripts were
75
detected in mussel oocytes, possibly a maternal effect, at levels raising beyond the settling transition
76
at 15-18 days post-fertilization and peaking in the adult hemocytes6,15. Using specific riboprobes
77
and antisera, granular hemocytes were confirmed as source of MytC transcripts and peptides15.
78
The functional roles suggested for the mussel myticins span from antibacterial and antiviral effects
79
to chemotactic and immunomodulatory behaviour. Native MytA and MytB preferentially killed
80
Gram positive (G+; 2.25-4.5 µM MBC on Micrococcus luteus) than Gram negative (G-) bacteria
81
and MytB also inhibited the fungus Fusarium oxysporum (MIC: 5-10 µM)3. The infectivity of
82
VHSV, an enveloped virus causing hemorrhagic septicemia in fish, was significantly inhibited in
83
Chinook salmon embryo cells transiently transfected with different GFP-MytC plasmids16. In the
84
same study, raw lysates of the MytC-transfected cells induced hemocyte chemotaxis at variable
85
degree and some immune-related hemocyte genes appeared up-regulated at 48-72 h after in vivo
86
injection of one MytC construct. As proof-of-concept, one chemically synthesized variant of the
87
mature MytC partially decreased the VHSV infectivity in cultured fish cells at 62.5-125 µM range
88
and significantly induced mussel hemocyte chemotaxis in filtered sea water at 0.025-0.25 µM
89
range17. The same peptide inhibited the E. coli growth in the range 25-250 µM (70% reduction at
90
250 µM) only when prepared at pH 3, a condition suggested to increase the α-helix component in
4 ACS Paragon Plus Environment
Page 5 of 32
Journal of Agricultural and Food Chemistry
91
the peptide structure, to cause aggregation of anionic phospholipids and destabilization of
92
negatively charged membranes17.
93
MATERIALS AND METHODS
94
Modelling of the mature MytC. Precursor transcript variants of MytC were downloaded as
95
nucleotide sequences from GenBank and Sequence Read Archive (http://www.ncbi.nlm.nih.gov).
96
The AA conservation profile of the mature peptide only was achieved with the freely available tool
97
WebLogo v3.4 using the sample SRR286640 (hemocytes sequence reads from mussels farmed in
98
the Venice lagoon area)7. The 3D model of the mature peptide was computed with I-tasser18 and
99
then refined with GROMACS v. 4.6.519 (www.gromacs.org).
100
Based on Cα distances, disulfide bonds fully compatible with the cysteine positions (Cys 5-24, 10-
101
33, 14-35, 19-38) were imposed. For all simulations, the Charmm forcefield was used in a
102
5.188×5.188×3.630 nm cell filled with SPC/E water model, in isothermal-isobaric ensemble
103
(NPT)20.
104
A 1.2 nm distance cut-off was imposed both for the Lennard-Jones vdW potential and for the
105
electrostatic interactions, using the particle mesh Ewald (PME) method with Fourier spacing 0.12
106
and PME order 4. Berendsen and Parrinello-Rahman’s temperature-pressure coupling was chosen
107
for solvent equilibration and simulated annealing, respectively, with isotropic pressure coupling.
108
After a minimization of 2000 steps, the Linear Constraint Solver (LINCS) method for bond
109
constraints was chosen. For solvent equilibration, a 2 ps molecular dynamics simulation was
110
performed at 300 K, with constrained protein atom positions. Finally, simulated annealing of 10 ns
111
was performed with annealing time of 0, 2 and 8 ns, at temperatures of 300, 365 and 280 K for
112
protein and 300, 320 and 280 K for water, respectively. For the 100 ns simulation, the same
113
experimental protocol was used, except that the simulated annealing was replaced with free
114
dynamics at 300 K.
5 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 6 of 32
115
The superpose function of PyMOL (www.pymol.org) was used to overlay ribbon peptide structures;
116
the homolog search in the PDB was performed via DALI server21.
117
Peptide Synthesis. Synthetic peptides were prepared by the solid phase peptide synthesis method
118
using a multiple peptide synthesizer (Syro II, MultiSynTech GmbH) on an LL Rink Amide resin
119
(Novabiochem, Bad Soden, Germany) (Table 1). The fluoren-9-ylmethoxycarbonyl (Fmoc)
120
strategy22 was used, utilizing 2-(7-aza-1H-benzotriazole-1-yl)-1,1,3,3-tetramethyluronium
121
hexafluorophosphate (HATU) as coupling reagent23. The side-chain protected amino acid building
122
blocks used were: N-α-Fmoc-Nω-(2,2,4,6,7-pentamethyldihydrobenzofuran-5-sulfonyl)-l-arginine,
123
N-α-Fmoc-O-tert-butyl-l-tyrosine, N-α-Fmoc-O-tert-butyl-l-serine, N-α-Fmoc-Nε-(tert-
124
butyloxycarbonyl)-l-lysine, N-α-Fmoc-N(im)-trityl-l-histidine, N-α-Fmoc-S-trityl-cystine, Fmoc-
125
Gly-Ser(Psi(Me, Me)pro)-OH, Fmoc-Thr(tBu)-Ser(Psi(Me, Me)pro)-OH. Cleavage of the peptides
126
was performed by reacting the peptidyl-resins with a mixture containing TFA/ethanedithiol
127
95%/5% for 2.5 h, and then precipitated by addition of cold diethyl ether. Crude peptides were
128
purified by preparative HPLC with Prep. Nova-Pak Cartridge HRC18 (Waters Corporation,
129
Milford, MA USA). Analytical HPLC analyses were performed on an Onyx Monolithic C18
130
column (Phenomenex, Torrance, CA, USA) with a flow rate of 2 mL/min, and detection at 220 nm.
131
The molecular peptide mass was confirmed by mass spectroscopy on a MALDI TOF–TOF mass
132
spectrometer (model 4800, Applied Biosystems, Carlsbad, CA, USA). The peptide purity was in the
133
range 95-98%, as evaluated by analytical HPLC.
134
Oxidative Refolding. Following purification, the reduced peptides MytC[23-34] and MytC[19-
135
40]S were dissolved at 1 mg/mL concentration in MQ water and allowed to spontaneously oxidize
136
for 24 h. Disulphide bond formation was followed by HPLC and confirmed by mass spectroscopy
137
using a MALDI TOF-TOF mass spectrometer for 48 h at room temperature. The resulting product
138
was lyophilized. HPLC analysis was performed under the same conditions described above.
6 ACS Paragon Plus Environment
Page 7 of 32
Journal of Agricultural and Food Chemistry
139
Antibacterial Activity. Minimal Inhibitory Concentration (MIC) was determined on selected
140
bacterial strains (Escherichia coli ATCC 25922, Pseudomonas aeruginosa ATCC 27853,
141
Staphylococcus aureus ATCC 29213, Enterococcus faecalis ATCC 29212) using the broth
142
microdilution susceptibility test24. Serial two-fold dilutions of each peptide (0.5-64 µM, final
143
concentration) were prepared in 50 µL volume, using 96-well microtiter plates (Sarstedt, Germany)
144
and 50% Muller-Hinton broth (MH; Difco Laboratories, USA) buffered at pH 5 or 7. Each dilution
145
series included a control well without peptide. Fifty µL of bacterial culture (~5×105 CFU/mL in
146
50% MH broth) were then added to each well. The MIC value was defined as the lowest peptide
147
concentration that prevented visible bacterial growth after incubation for 24 h at 37 °C. The
148
minimal bactericidal concentration (MBC) corresponded to the lowest peptide concentration that
149
prevented any CFU after plating 50 µL from the wells onto nutrient agar plates. The reported MIC
150
and MBC data result from three independent assays.
151
Selected MytC fragments were tested on Micrococcus lysodeikticus ATCC 4698. Vibrio splendidus
152
LGP32 and Vibrio aestuarianus, using suitable media: Poor Broth (5 g/L NaCl, 10 g/L
153
bactotryptone, pH 7.5) for the standard G+ bacterium and Zobel medium modified for marine
154
strains (artificial sea water at a third of the strength added with 4 g/L bactopeptone and 1 g/L yeast
155
extract)25.
156
The inhibition kinetics of MytC[19-40]SOX was comparatively evaluated on E. coli and E. faecalis
157
at pH 5, using a greater bacterial inoculum (107 CFU/mL). The optical density was monitored every
158
30 min at 620 nm for 24 h with the microplate reader Infinite 200 Pro (Tecan, Grödig, Austria).
159
Antifungal Activity. The antifungal activity was evaluated on the sporogenous Aspergillus
160
brasiliensis ATCC 16404 and the yeast Candida albicans ATCC 10231 strains using the broth
161
microdilution assay26. Each peptide was prepared as indicated above (0.5-64 µM) and tested in
162
Sabouraud broth (pH 5.8) inoculated with 50 µL of the yeast or spore suspension (no peptide in
163
control wells). Yeast cells were cultivated overnight in liquid medium and their concentration 7 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 8 of 32
164
adjusted to 2.5×105 CFU/mL, while the sporogenous strain was prepared by harvesting fungal
165
spores and adjusting their concentration to 0.4–5×104 CFU/mL using a Bright Line™
166
haemocytometer (Sigma, USA). The inoculated 96-well microplates were incubated at 35°C and
167
MIC values were determined visually after 48 h. Minimal fungicidal concentration (MFC)
168
corresponded to the lowest peptide concentration that prevented any CFU after plating 50 µL from
169
the wells onto the corresponding nutrient agar plates.
170
Permeability Assays. E. coli and E. faecalis bacteria were grown at 37°C in MH broth (pH 7) up to
171
mid-logarithmic phase, washed and resuspended at 107 CFU/mL in 10 mM sodium phosphate
172
buffer, pH 7.4. The cells were subsequently incubated for 15 min in the dark with 2 µM SYTOX ®
173
Green Nucleic Acid Stain (Life Technologies), a dye that selectively permeates membrane-damaged
174
bacteria. Then, the peptide to be tested was added at the MIC value and the fluorescence increase
175
due to the binding of cationic dye to intracellular DNA was monitored every minute for 1 h
176
(excitation and emission wavelengths: 485 nm and 520 nm, respectively). Melittin was used as
177
positive control.
178
Circular Dichroism Spectroscopy of MytC [19-40]SOX. Samples were prepared at room
179
temperature by adding increasing amounts of DPC or SDS (0.5-20 mM final detergent
180
concentration) to a 30 µM peptide solution in 20 mM phosphate buffer, pH 7.4. The peptide
181
concentration was determined by UV-absorbance at 290 nm. All CD measurements were performed
182
with a Jasco J-715 spectropolarimeter (Tokyo, Japan) using a HELLMA quartz cuvette with optical
183
path length of 0.2 cm. The CD spectra were acquired from 192 nm to 250 nm, with 2 nm
184
bandwidth, 2 s time constant, and 50 nm/min scan speed; they were then processed with the Jasco J-
185
700 software (version 1.50.01). The signal/noise ratio was improved by accumulating ten scans per
186
experimental condition.
187
NMR Analysis of MytC[19-40]SOX. The lyophilized peptide was dissolved in MQ H2O to final
188
concentration of 1 mM, and incubated for 240 h at 277 K in order to achieve complete solvation. To 8 ACS Paragon Plus Environment
Page 9 of 32
Journal of Agricultural and Food Chemistry
189
the peptide samples, DPC or SDS were added to a final concentration of 300 mM; 10% 2H2O was
190
added and the pH was adjusted to 7.0. NMR experiments were performed in 5 mm tubes at 295 K
191
using an Avance DMX600 spectrometer (Bruker, Germany) operating at 599.90 MHz for 1H and
192
equipped with a 5 mm TXI xyz-triple gradient probe. COSY spectra were recorded with 1 s
193
recovery delay whereas the TOCSY spectra were acquired using a DIPSI-2 spin-lock, 70 ms
194
mixing-time and 1 s recovery delay. For the NOESY experiments, 150 ms mixing time and 1.4 s
195
recovery delay were used. Data were processed using TopspinTM (Bruker, Germany) and analyzed
196
with CARA 1.9.0 Beta 7 software (cara.nmr-software.org).
197 198 199
RESULTS AND DISCUSSION
200
Structural features of MytC. We chose the consensus sequence of a MytC variant previously
201
described11 and predominantly expressed in farmed Italian mussels7 to compute the model of
202
mature MytC using the I-tasser algorithm for protein folding prediction18. The five resulting
203
structures showed unequivocal pairing of eight conserved cysteines, though the disulfide bridges
204
were not entirely defined. Connectivity between cysteines 5-24, 10-33, 14-35 and 19-38 were set as
205
molecular constraints for model refinement by simulated annealing with GROMACS19, a flexible
206
toolkit for molecular dynamics simulation (Figure 1A). Following a further simulation at 300 K for
207
100 ns, we selected three convergent MytC structures which clearly reproduced a compact CSαβ
208
scaffold and a γ-core motif common to various AMPs and to venom components27. These models
209
showed the most variable AA residues clustered mainly on the surface in the α-helical region
210
(Figure 1B). Both the invariant cysteine array and tolerated sequence changes are essential for the
211
biological function while minimizing the risk of pathogen resistance8, hence useful for nature-
212
guided drug design28.
9 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 10 of 32
213
The γ-core motif of MytC does not present the typical GxC element and is likely stabilized by
214
aromatic stacking interactions between two highly conserved tyrosines, as previously observed in
215
tachyplesin I, a β-hairpin AMP of 17 AA isolated from horseshoe crab29. Once superimposed, the
216
structural sub-domains of MytC appeared closely similar to those of other CSαβ peptides, such as
217
defensin 1 (MGD1)30 and mytilin B31 from M. galloprovincialis and butantoxin, a scorpion venom
218
peptide stabilized by three disulfide bridges32(Figure 2). MytC (PF10690) and PDB accession codes
219
of the other peptides reported in the figure are available in public protein databases33,34.
220
While maintaining the invariant cysteine array and stacking tyrosines, changes of the variable
221
residues could be considered for the development of MytC-derived antimicrobial modules of
222
minimal size.
223
Chemical synthesis of peptides mapping the mature MytC. We synthesized and purified several
224
fragments and analogues, designed all along the 40 AA MytC sequence, aiming to explore their
225
antimicrobial and antifungal activity (Table 1). Specifically, the β-turn suspected to play a role in
226
the antimicrobial activity was covered with fragments of 12 and 22 residues: MytC[23-34],
227
MytC[19-40], and MytC[19-40]S. In the latter, three Cys were replaced with Ser, to obtain a single
228
intramolecular disulfide bridge after spontaneous oxidation for 24 h, yielding peptide MytC[19-
229
40]SOX. According to MS analysis, the empirical mass values of the reduced and oxidized
230
fragments are consistent with the expected values reported in Table 1 (Figure 3: A, B). We also
231
synthesized and purified the whole 40 AA peptide for comparative purposes (Figure 3C ). The
232
sequence of MytC[1-40] makes its synthesis particularly difficult. Partial hydrophobicity and the
233
tendency to adopt a beta structure lead to peptide aggregation on the resin. To overcome this issue,
234
a strategy based on the coupling of Gly-Ser in positions 24-25 and Thr-Ser in positions 33-34 as
235
pseudo-prolines (dimethyloxazolidine analogs) was followed. These building blocks have a
236
preference for cis-amide bonds resulting in a twisted conformation of the peptide backbone35. This
10 ACS Paragon Plus Environment
Page 11 of 32
Journal of Agricultural and Food Chemistry
237
strategy allowed us to greatly increase the global yield up to 20% after peptide purification (a
238
promising result for the synthesis of other similar peptides).
239
MytC-derived peptide fragments show pH-dependent antimicrobial and antifungal activity.
240
AMPs can exhibit differential antimicrobial activity at different pH36,17.When dissolved and tested
241
at pH 7 on G+ and G- bacteria, the nine synthetic peptides of MytC were inactive or just weakly
242
active, with minimal inhibitory concentration (MIC) values above 64 µM, except for the G+
243
Enterococcus faecalis which was inhibited by 32 µM MytC[1-40], MytC[29-40], MytC[19-40],
244
MytC[19-40]S and by 16 µM MytC[19-40]SOX (Suppl. Table S1).
245
The assays were repeated at pH 5 with the same peptides, this time evaluating both inhibitory and
246
killing activity against bacterial, fungal and yeast strains (Table 2). At acidic pH, a greater number
247
of positive charges could improve the peptide interaction with bacteria37; moreover, the ionic
248
interactions are enhanced at the lower salt content of the medium. MytC[1-40] exhibited a better
249
MIC on G+ (4 µM, E. faecalis) than G- bacteria (32 µM, E. coli), similar to folded defensins from
250
M. galloprovincialis, Crassostrea gigas and Venerupis philippinarum5,25,38. Though partially
251
including α and β elements, MytC[9-28] inhibited only E. faecalis at 16 µM.
252
The three MytC fragments of 22 AA, covering the β elements, inhibited E. coli and E. faecalis at 8-
253
16 µM and Staphylococcus aureus at 32 µM. MytC[19-40] was more effective than its substituted
254
and oxidised forms in killing E. coli (8 µM MBC). The tested peptides did not inhibit or kill
255
Candida albicans except MytC[19-40]SOX (16 µM) and MytC[19-40]S (32 µM). Overall, the
256
activating effect of the acidic vs. neutral pH supports previous findings obtained in E. coli with a
257
synthetic MytC variant17 and the antimicrobial effectiveness of MytC in the endophagosome39. In
258
our experimental conditions, E. faecalis (G+) and Pseudomonas aeruginosa (G-) were the most and
259
least sensitive strains, respectively.
260
In essence, the three 22 AA peptide fragments including two antiparallel β-strands and the putative
261
β-turn were bacteriostatic, bactericidal and somewhat fungistatic. Following a similar mapping 11 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 12 of 32
262
approach, the antimicrobial activity of tenecin 1, an insect defensin, was also located in the β-sheet
263
region40. Analogous peptide fragments from MGD1 inhibited Micrococcus lysodeikticus at 10-16
264
µM and E. coli at 22-62 µM when tested in poor broth5. Moreover, β-hairpins from insect defensins
265
killed G+ bacteria in the 1.1-14.9 µM range, with no evidence of antifungal activity41. Related to
266
these [19-40] fragments, the natively folded MytC with its putative helical element could present
267
increased stability as well as more efficient interaction/permeation of microbial membranes5,41.
268
Besides, we tested the 22 AA peptides on M. lysodeikticus, Vibrio splendidus LGP32 and Vibrio
269
aestuarianus, the latter as opportunistic components of marine bivalve microbiomes42. Only
270
MytC[19-40]S and MytC[19-40]SOX inhibited M. lysodeikticus at 32 µM (Suppl. Table S2). Earlier,
271
a β-hairpin peptide designed on MytB and stabilized by two disulphide bonds inhibited M.
272
lysodeikticus (1 mM) more than V. splendidus LGP32 (125 µM)31 and defensin-like peptides were
273
found scarcely active on Vibrio spp. even in poor broth5,43,38.
274
At the MIC value of 8 and 16 µM, respectively, MytC[19-40]SOX caused a CFU reduction of 3 logs
275
on E. coli and 4 logs on E. faecalis (MBC: 32 µM and 64 µM, respectively) (Suppl. Table S3).
276
Additional kinetic absorbance measurements performed at pH 5 with the same G- and G+ strains
277
(107 CFU/mL, 0-64 µM range) highlighted a different bacteriostatic effect. MytC[19-40]SOX
278
partially inhibited E. coli at 8 µM (MIC) and likely caused immediate killing at 16-64 µM.
279
Conversely, the E. faecalis growth was delayed but not blocked in the 8-64 µM dose range. For
280
comparison, the pore-forming helical melittin tested at the MIC value on the same strains (2 µM on
281
E. coli; 16 µM on E. faecalis) produced immediate and total growth inhibition44,45 (Figure 4).
282
The pH dependency of the observed antimicrobial effects and the low micromolar MIC of the
283
unfolded 40 AA MytC support the hypothesis that MytC can act in multiple ways and persist in vivo
284
like human defensin 5 in the paneth cells (thioredoxin-dependent reduction and zinc-mediated
285
resistance to hydrolytic breakdown)46.
12 ACS Paragon Plus Environment
Page 13 of 32
Journal of Agricultural and Food Chemistry
286
We further assessed the inhibition mechanism of MytC[19-40]SOX on E. coli and E. faecalis at the
287
MIC value by monitoring the uptake of SYTOX green, a nucleic acid binding molecule (Figure 5).
288
The positive control melittin (red line) induced immediate membrane damage on both bacterial
289
strains whereas the fluorescence increase starting five min after the addition of MytC[19-40]SOX
290
(black line) indicated a progressive damage of the E. coli membranes. Conversely, the fluorescence
291
values obtained with E. faecalis remained close to the baseline, suggesting a different mechanism of
292
action. The β-hairpin region is critical for the antimicrobial activity, even if the cysteine scaffold
293
with the α-helix is missing, and the selective membrane-disrupting action of MytC[19-40]SOX on G-
294
(E. coli) related to G+ bacteria (E. faecalis) likely depends on cell wall differences. Whether the
295
killing effect is due to direct membrane damage or consequent to events occurring in the reducing
296
cytosolic environment47 requires further study.
297
Direct accessibility to Lipid II in G+ bacteria (only reachable by perturbing the outer membrane in
298
G- bacteria) and subsequent inhibition of cell wall biosynthesis have been hypothesized to explain
299
the more pronounced antimicrobial effects of folded invertebrate defensins on G+48,39. Supposing a
300
similar behavior for MytC, our data suggest that the β-hairpin alone is able to damage the G-
301
membranes with only partial affinity to Lipid II. This could explain the lower MBC showed by the
302
MytC [19-40] fragments on E. coli compared to E. faecalis.
303
MytC[19-40]SOX adopts an unpredicted, atypical conformation in membrane-mimetic
304
environments. MytC[19-40]SOX was preliminary subjected to conformational analysis using far-
305
UV circular dichroism (CD) spectroscopy in different experimental conditions. In spite of one
306
disulfide bridge linking its N- and C-termini, the peptide did not show a defined secondary structure
307
in aqueous solution, as expected from the MytC model structure.
308
In the presence of dodecylphosphocholine (DPC) or sodium dodecyl sulfate (SDS) above the
309
critical micelle concentration, the CD spectra changed significantly, indicating at least a partially
310
structured peptide in the micellar environments. Specifically, the CD spectra displayed a positive 13 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 14 of 32
311
peak below 200 nm and a negative one at around 217 nm in the presence of either detergent. Other
312
disulfide-stabilized β-hairpins displayed a CD spectrum with two positive bands at around 200 and
313
230 nm and a negative one between 205 and 210 nm49. Five aromatic residues and the disulfide
314
bridge might explain the atypical CD spectra of MytC[19-40]SOX (Suppl. Figure S1).
315
To get more insights on the structure adopted in micelle environments, we acquired sets of two-
316
dimensional homonuclear NMR spectra of MytC[19-40]SOX in the presence of SDS or DPC. The
317
dispersion of the amide protons was larger in both detergent solutions than in aqueous buffer,
318
confirming that the peptide is at least partially structured in the micellar environments (Figure 6A).
319
In both cases however, an excessive number of spin systems suggested structural heterogeneity. In
320
the presence of SDS, severe broadening of the peaks occurred during the acquisition of the NMR
321
dataset, most likely because of aggregation events. Nevertheless, the assignment was completed at
322
least for the most populated conformer, both in SDS and DPC. Due to extensive resonance overlap,
323
line broadening and lack of long-range NOEs, it was not possible to proceed with structure
324
calculation. Nevertheless, the small negative deviation of the Hα chemical shifts from typical
325
random coil values indicates a weak propensity to form helical segments. Moreover, the comparison
326
of Hα secondary chemical shifts in the two media suggests similar peptide conformation50 (Figure
327
6B). A consecutive stretch of α(i)-NH(i+2) NOE correlations in DPC indicates the presence of a 310
328
helix between Ser-2 and Leu-9. The signal between His-10 Hα and Lys-13 HN suggests a loop
329
between these two residues. Peptide flexibility from Ser-15 to Arg-22 could explain a localized
330
absence of medium range correlations and the general lack of long range peaks in the NOESY
331
spectrum (Figure 6C). In our experimental conditions, there is no evidence of a stable,
332
independently folded β-hairpin and, consistent with the CSαβ motif predicted for the mature MytC,
333
its formation would likely require two additional disulfide bonds with the preceding α-helix. Amino
334
acids substitutions designed to provide an additional disulfide could stabilize the hairpin
335
conformation. On the other hand, the conformation adopted by MytC[19-40]SOX in the presence of 14 ACS Paragon Plus Environment
Page 15 of 32
Journal of Agricultural and Food Chemistry
336
bacterial membranes could differ from that observed in micellar environments, thus explaining its
337
antimicrobial activity.
338
In agreement to the nature-guided development of bioactive molecules, rigorous structural and
339
biological data are an essential step to understand the biological roles of MytC. In perspective,
340
MytC-derived peptides acting at low or neutral pH could be validated to protect acidic mucosal
341
tissues, cosmetic and food products and possibly used as adjuvants in aquaculture.
342
Abbreviations
343
AA, amino acid
344
AMPs, antimicrobial peptides
345
CD, circular dicroism
346
COSY, correlation spectroscopy
347
CSαβ, cysteine-stabilized alpha beta
348
DPC, dodecylphosphocholine
349
EST, expressed sequence tag
350
GFP, green fluorescence protein
351
G+/G-, Gram positive/Gram negative
352
HPLC, high performance liquid chromatography
353
MIC, minimal inhibitory concentration
354
MBC, minimal bactericidal concentration
355
MS, mass spectrometry
356
MQ, MilliQ [water]
357
Myt, myticin [MytC, myticin C] 15 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 16 of 32
358
NMR, nuclear magnetic resonance
359
NOE/NOESY, Nuclear Overhauser Effect/ Nuclear Overhauser Effect Spectroscopy
360
SDS, sodium dodecyl sulfate
361
TOCSY, total correlation spectroscopy
362
vdW, van der Waals
363
ACKNOWLEDGMENTS
364
We thank Umberto Rosani (INNOV-H20 network for technological innovation in aquaculture) for
365
updating the MytC ESTs analysis.
366
ASSOCIATED CONTENT
367
S Supporting Information
368
Tables S1, S2, S3 and Figure S1. This material is available free of charge via the Internet at
369
http://pubs.acs.org.
370
AUTHOR INFORMATION
371
Corresponding Author
372
PV. E-mail:
[email protected]. Address: Department of Biology, Via U. Bassi 58/B, 35131
373
Padova, Italy
374
SD, MF and ND did the experimental work and critically contributed to its development. RN and
375
NMEH contributed to the synthesis and antifungal assays, respectively. MB and SM, OM, and PV
376
supervised selected parts of work. PV designed the work and drafted the manuscript. All Authors
377
contributed to the manuscript and approved its final version.
378
Funding
16 ACS Paragon Plus Environment
Page 17 of 32
Journal of Agricultural and Food Chemistry
379
Work primarily funded by PRAT 2012 (CPDA128951) to PV and, in part, by PRIN 2008-2010
380
(20109XZEPR) to PV and SM. According to the objectives of the mentioned projects, SD and RN
381
were contracted as post-doc and graduated fellows, respectively. We thank the Ministry of Higher
382
Education (MoHE) and Missions Sector-Egypt for granting a visiting research scholarship to
383
NMEH.
384 385
Notes
386
The authors declare no competing financial interest.
387
17 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
388 389
REFERENCES 1. Hubert F.; Noel T.; Roch P. A member of the arthropod defensin family from edible
390
Mediterranean mussels (Mytilus galloprovincialis) Eur J Biochem. 1996, 240(1), 302-6
391
[Erratum in: Eur. J. Biochem. 240, 815].
392
Page 18 of 32
2. Charlet M.; Chernysh S.; Philippe H.; Hetru C.; Hoffmann J.A.; Bulet P. Innate immunity.
393
Isolation of several cysteine-rich antimicrobial peptides from the blood of a mollusc, Mytilus
394
edulis. J. Biol. Chem. 1996, 271, 21808-13.
395
3. Mitta G.; Hubert F.; Noël T.; Roch P. Myticin, a novel cysteine-rich antimicrobial peptide
396
isolated from haemocytes and plasma of the mussel Mytilus galloprovincialis. Eur. J.
397
Biochem. 1999, 265, 71-8.
398 399
4. Mitta G.; Vandenbulcke F.; Roch P. Original involvement of antimicrobial peptides in mussel innate immunity. FEBS Lett. 2000a, 486, 185-90.
400
5. Mitta, G.; Vandenbulcke, F.; Noël, T.; Romestand, B.; Beauvillain, J.C.; Salzet, M.; Roch,
401
P. Differential distribution and defence involvement of antimicrobial peptides in mussel. J.
402
Cell Sci. 2000b, 113, 2759-69.
403
6. Pallavicini, A.; Costa, M. del M.; Gestal, C.; Dreos, R.; Figueras, A.; Venier, P.; Novoa, B.
404
(2008) High sequence variability of myticin transcripts in hemocytes of immune-stimulated
405
mussels suggests ancient host-pathogen interactions. Dev. Comp. Immunol. 32, 213-26.
406
7. Rosani, U.; Varotto, L.; Rossi, A.; Roch, P.; Novoa, B.; Figueras, A.; Pallavicini, A.;
407
Venier, P. Massively parallel amplicon sequencing reveals isotype-specific variability of
408
antimicrobial peptide transcripts in Mytilus galloprovincialis. PLoS One 2011, 6, e26680.
409
8. Padhi, A.; Verghese, B. Molecular diversity and evolution of myticin-C antimicrobial
410
peptide variants in the Mediterranean mussel, Mytilus galloprovincialis. Peptides 2008, 29,
411
1094-101. 18 ACS Paragon Plus Environment
Page 19 of 32
Journal of Agricultural and Food Chemistry
412
9. Vera, M.; Martínez, P.; Poisa-Beiro, L.; Figueras, A.; Novoa, B. Genomic organization,
413
molecular diversification, and evolution of antimicrobial peptide myticin-C genes in the
414
mussel (Mytilus galloprovincialis). PLoS One 2011, 6, e24041.
415 416 417
10. Tassanakajon, A.; Somboonwiwat, K.; Amparyup, P. Sequence diversity and evolution of antimicrobial peptides in invertebrates. Dev. Comp. Immunol. 2014, 48, 324-41. 11. Venier, P.; De Pittà, C.; Bernante, F.; Varotto, L.; De Nardi, B.; Bovo, G.; Roch, P.; Novoa,
418
B.; Figueras, A.; Pallavicini, A.; Lanfranchi, G. MytiBase: a knowledgebase of mussel (M.
419
galloprovincialis) transcribed sequences. BMC Genomics 2009, 10, 72.
420
12. Gerdol, M.; De Moro, G.; Manfrin, C.; Venier, P.; Pallavicini, A. Big defensins and
421
mytimacins, new AMP families of the Mediterranean mussel Mytilus galloprovincialis. Dev.
422
Comp. Immunol. 2012, 36, 390-9.
423
13. Liao Z.; Wang X.C.; Liu H.H.; Fan M.H.; Sun J.J.; Shen W. Molecular characterization of a
424
novel antimicrobial peptide from Mytilus coruscus. Fish Shellfish Immunol. 2013, 34, 610-6.
425
14. Gerdol M.; Puillandre N.; De Moro G.;, Guarnaccia C.; Lucafo M.; Benincasa M.; Zlatev
426
V.; Manfrin C.; Torboli V.; Giulianini P.G.; Sava G.; Venier P.; Pallavicini A. Identification
427
and characterization of a novel family of cysteine-rich peptides(MgCRP-I) from Mytilus
428
galloprovincialis. Genome Biol. Evol. 2015, 7(8):2203-19.
429
15. Balseiro, P.; Moreira, R.; Chamorro, R.; Figueras, A.; Novoa, B. Immune responses during
430
the larval stages of Mytilus galloprovincialis: metamorphosis alters immunocompetence,
431
body shape and behavior. Fish Shellfish Immunol. 2013, 35, 438-47.
432
16. Balseiro, P.; Falcó, A.; Romero, A.; Dios, S.; Martínez-López, A.; Figueras, A.; Estepa, A.;
433
Novoa, B. Mytilus galloprovincialis myticin C: a chemotactic molecule with antiviral
434
activity and immunoregulatory properties. PLoS One 2011, 6, e23140.
19 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
435
17. Martinez-Lopez, A.; Encinar, J.A.; Medina-Gali, R.M.; Balseiro, P.; Garcia-Valtanen, P.;
436
Figueras, A.; Novoa, B.; Estepa, A. pH-dependent solution structure and activity of a
437
reduced form of the host-defense peptide myticin C (Myt C) from the mussel Mytilus
438
galloprovincialis. Mar. Drugs 2013, 11, 2328-46.
439 440 441
18. Yang, J.; Yan, R.; Roy, A.; Xu, D.; Poisson, J.; Zhang, Y. The I-TASSER Suite: Protein structure and function prediction. Nat. Methods 2015, 12, 7-8. 19. Pronk, S.; Páll, S.; Schulz, R.; Larsson, P.; Bjelkmar, P.; Apostolov, R.; Shirts, M.R.; Smith,
442
J.C.; Kasson, P.M.; van der Spoel, D.; Hess, B.; Lindahl, E. GROMACS 4.5: a high-
443
throughput and highly parallel open source molecular simulation toolkit. Bioinformatics
444
2013, 29, 845-54.
445 446 447 448 449 450 451 452
Page 20 of 32
20. van der Spoel, D.; Lindahl, E.; Hess, B. and the GROMACS development team GROMACS User Manual version 4.6.7. 2014, (accessed at www.gromacs.org). 21. Holm, L.; Rosenström, P. Dali server: conservation mapping in 3D. Nucleic Acids Res. 38(Web Server issue), 2010, W545-9. 22. Fields, G.B.; Noble, R.L. Solid phase peptide synthesis utilizing 9fluorenylmethoxycarbonyl amino acids. Int. J. Pept. Protein Res. 1990, 35, 161–214. 23. Carpino, L.A.; Ferrer, F.J. The 5,6- and 4,5-benzo derivatives of 1-hydroxy-7azabenzotriazole. Org Lett. 2001, 3, 2793-5.
453
24. Wiegand, I.; Hilpert, K.; Hancock, R.E.W. Agar and broth dilution methods to determine the
454
minimal inhibitory concentration (MIC) of antimicrobial substances. Nat. Protoc. 2008, 3,
455
163-75.
456
25. Schmitt, P.; Wilmes, M.; Pugnière, M.; Aumelas, A.; Bachère, E.; Sahl, H.G.; Schneider, T.;
457
Destoumieux-Garzón, D. Insight into invertebrate defensin mechanism of action: oyster
20 ACS Paragon Plus Environment
Page 21 of 32
Journal of Agricultural and Food Chemistry
458
defensins inhibit peptidoglycan biosynthesis by binding to lipid II. J. Biol. Chem. 2010, 285,
459
29208-16.
460
26. CLSI Reference method for broth dilution antifungal susceptibility testing of filamentous
461
fungi; approved standard CLSI document M38-A2. Clinical and Laboratory Standards
462
Institute, Wayne. 2008 (accessed at www.springer.com).
463 464 465 466 467
27. Yeaman, M.R.; Yount, N.Y. Unifying themes in host defence effector polypeptides. Nat. Rev. Microbiol. 2007, 5, 727-40. 28. Zhu, S.; Gao, B.; Tytgat, J. Phylogenetic distribution, functional epitopes and evolution of the CSalphabeta superfamily. Cell Mol. Life Sci. 2005, 62, 2257-69. 29. Laederach, A.; Andreotti, A.H.; Fulton, D.B. Solution and micelle-bound structures of
468
tachyplesin I and its active aromatic linear derivatives. Biochemistry 2002, 41, 12359-68.
469
30. Gueguen, Y.; Herpin, A.; Aumelas, A.,; Garnier, J.; Fievet, J.; Escoubas, J.M.; Bulet, P.;
470
Gonzalez, M.; Lelong, C.; Favrel ,P.; Bachère, E. Characterization of a defensin from the
471
oyster Crassostrea gigas. Recombinant production, folding, solution structure, antimicrobial
472
activities, and gene expression. J. Biol. Chem. 2006, 281, 313-23.
473
31. Roch, P.; Yang, Y.; Toubiana, M.; Aumelas, A. NMR structure of mussel mytilin, and
474
antiviral-antibacterial activities of derived synthetic peptides. Dev. Comp. Immunol. 2008,
475
32, 227-38.
476 477 478 479
32. Holaday, S.K. Jr.; Martin, B.M.; Fletcher, P.L. Jr.; Krishna, N.R. NMR solution structure of butantoxin. Arch. Biochem. Biophys. 2000, 379, 18-27. 33. Berman H.M.; Westbrook J.; Feng Z., Gilliland G.; Bhat T.N., Weissig H., Shindyalov I.N., Bourne P.E. The Protein Data Bank Nucleic Acids Res. 2000, 28, 235-42.
21 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
480
34. Finn R.D.; Bateman A.; Clements J.; Coggill P.; Eberhardt R.Y.; Eddy S.R.; Heger A.;
481
Hetherington K.; Holm L.; Mistry J.; Sonnhammer E.L.L.; Tate J.; Punta M. Pfam: the
482
protein families database. Nucleic Acids Res. 2014, 42(Database issue), D222-30.
483
35. Cremer, G.A.; Tariq, H.; Delmas, A.F. Combining a polar resin and a pseudo-proline to
484
optimize the solid-phase synthesis of a 'difficult sequence'. J Pept Sci. 2006, 12, 437-42.
485
Page 22 of 32
36. Park, S.C.; Kim, J.Y.; Lee, J.K.; Hahm, K.S.; Park, Y. Antibacterial action of new
486
antibacterial peptides, Nod1 and Nod2, isolated from Nordotis discus discus. J. Agric. Food
487
Chem. 2012, 60, 6875-81.
488 489 490
37. Huang, X.X.; Gao, C.Y.; Zhao, Q.J., Li, C.L. Antimicrobial characterization of site-directed mutagenesis of porcine beta defensin 2. PLoS One 2015, 10, e0118170. 38. Zhang L.; Yang D.; Wang Q.; Yuan Z.; Wu H.; Pei D.; Cong M.; Li F.; Ji C.; Zhao J. A
491
defensin from clam Venerupis philippinarum: Molecular characterization, localization,
492
antibacterial activity, and mechanism of action. Dev. Comp. Immunol. 2015, 51, 29-38.
493
39. Bachère, E.; Rosa R.D.; Schmitt, P.; Poirier, A.C.; Merou, N.; Charrière, G.M.;
494
Destoumieux-Garzón, D. The new insights into the oyster antimicrobial defense: Cellular,
495
molecular and genetic view. Fish Shellfish Immunol. 2015, 46, 50-64.
496
40. Lee, K.H.; Hong, S.Y.; Oh, J.E.; Kwon, M.; Yoon, J.H.; Lee, J.; Lee, B.L.; Moon, H.M.
497
Identification and characterization of the antimicrobial peptide corresponding to C-terminal
498
beta-sheet domain of tenecin 1, an antibacterial protein of larvae of Tenebrio molitor.
499
Biochem J. 1998, 334, 99-105.
500 501 502 503
41. Gao, B.; Zhu, S. An insect defensin-derived β-hairpin peptide with enhanced antibacterial activity. ACS Chem. Biol. 2014, 9, 405-13. 42. Domeneghetti S.; Varotto L.; Civettini M.; Rosani U.; Stauder M.; Pretto T.; Pezzati E.; Arcangeli G.; Turolla E.; Pallavicini A.; Venier P. Mortality occurrence and pathogen 22 ACS Paragon Plus Environment
Page 23 of 32
Journal of Agricultural and Food Chemistry
504
detection in Crassostrea gigas and Mytilus galloprovincialis close-growing in shallow
505
waters (Goro lagoon, Italy). Fish Shellfish Immunol. 2014, 41, 37-44.
506
43. Schmitt, P.; Rosa, R.D.; Duperthuy, M.; de Lorgeril, J.; Bachère, E.; Destoumieux-Garzón,
507
D. The Antimicrobial defense of the Pacific oyster, Crassostrea gigas. How diversity may
508
compensate for scarcity in the regulation of resident/pathogenic microflora. Front.
509
Microbiol. 2012, 3, 160.
510
44. Chen, C.H.; Wiedman, G.; Khan, A.; Ulmschneider, M.B. Absorption and folding of
511
melittin onto lipid bilayer membranes via unbiased atomic detail microsecond molecular
512
dynamics simulation. Biochim. Biophys. Acta 2014, 1838, 2243-9.
513
45. Upadhyay, S.K.; Wang, Y.; Zhao, T.; Ulmschneider, J.P. Insights from Micro-second
514
Atomistic Simulations of Melittin in Thin Lipid Bilayers. J Membr Biol. 2015, 248, 497-
515
503.
516
46. Zhang, Y.; Cougnon, F.B.; Wanniarachchi, Y.A.; Hayden, J.A.; Nolan, E.M. Reduction of
517
human defensin 5 affords a high-affinity zinc-chelating peptide. ACS Chem. Biol. 2013, 8,
518
1907-11.
519
47. Chileveru, H.R.; Lim,S.A.; Chairatana, P.; Wommack, A.J.; Chiang, I.L.; Nolan, E.M.
520
Visualizing attack of Escherichia coli by the antimicrobial peptide human defensin 5.
521
Biochemistry. 2015, 54, 1767-77.
522 523
48. Dias, R.O.; Franco, O.L. Cysteine-stabilized defensins: From a common fold to antibacterial activity. Peptides 2015, doi: 10.1016/j.peptides.2015.04.017 [Epub ahead of print]
524
49. Powers, J.P.; Rozek, A.; Hancock, R.E. Structure-activity relationships for the beta-hairpin
525
cationic antimicrobial peptide polyphemusin I. Biochim. Biophys. Acta 2004, 1698, 239-50.
23 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
526
Page 24 of 32
50. Ulrich, E.L.; Akutsu, H.; Doreleijers, J.F.; Harano, Y.; Ioannidis, Y.E.; Lin, J.; Livny, M.;
527
Mading, S.; Maziuk, D.; Miller, Z.; Nakatani, E.; Schulte, C.F.; Tolmie, D.E.; Kent Wenger,
528
R.; Yao, H.; Markley, J.L. BioMagResBank. Nucleic Acids Res. 2008, 36, D402-8.
529
CAPTIONS FOR FIGURES AND TABLES
530
Figure 1. (A) Superimposed ribbon structures of three converging models of the mature MytC
531
obtained by I-tasser18 and refined with GROMACS19. Arrows indicate two stacking tyrosines in the
532
two antiparallel β-strands; N- and C- termini are also indicated. (B) Mesh representation of MytC
533
highlighting cysteine (blue) and variable (red) residues. Most of the latter are clustered in the helix
534
region.
535
Figure 2. Left. Multiple alignments illustrating the primary sequence (top) and secondary structure
536
elements (bottom) of MytC and selected CSαβ peptides. Right. Superimposed structures of MytC
537
(green) with mytilin B (2eemA, top), MGD1 (1fjnA, center), and butantoxin (1c56A, bottom).
538
Despite general similarities in the molecular organization and cysteine array, the superimposition
539
reveals topological differences in terms of loop length and β-sheet twist (the γ-core region).
540
Figure 3. Analytical RP-HPLC and MS analysis of MytC[19-40]S (A), MytC[19-40]SOX (B) and
541
MytC[1-40] (C). The chromatographic profiles, indicating the high peptide purity, were obtained on
542
a monolithic 100x4.6 mm, C18 column (OnyxTM, Phenomenex, USA) with appropriate acetonitrile
543
gradients (dashed lines). Though the peptides MytC[19-40]S and MytC[19-40]SOX differ for an
544
intramolecular disulfide bridge, no significant difference in their retention time was observed. Insets
545
show the related MS spectra. The experimental m/z values are consistent with the sequences of the
546
peptides, based on their theoretical molecular masses in Table 1.
547
Figure 4. Effect of 0-64 µM MytC[19-40]SOX on the growth of the E. coli (A) and E. faecalis (B) at
548
pH 5 (inoculum: 107 CFU/mL; absorbance at 620 nm). Black lines: negative control. Red lines:
549
melittin tested at MIC (2 µM on E. coli; 16 µM on E. faecalis).
24 ACS Paragon Plus Environment
Page 25 of 32
Journal of Agricultural and Food Chemistry
550
Figure 5. Evaluation of membrane permeabilization by SYTOX green intracellular uptake in E.coli
551
(A) and E. faecalis (B). Lines indicate melittin (red), MytC[19-40]SOX (black) or no peptide (dashed
552
line). Following 15 min incubation of bacterial cells at pH 7 with 2 µM SYTOX green, each peptide
553
was added at 1x MIC and the fluorescence values were recorded every minute for 1 h (Ex. 485 nm
554
and Em. 520 nm).
555
Figure 6. (A) Amide region of the 1H-NMR spectrum of MytC[19-40]SOX in SDS (blue), DPC
556
(red) and water (green), Peak dispersion and width suggest partial structuring of the peptide
557
fragment in both membrane-mimetic environments. (B) Difference between random and
558
experimental Hα chemical shifts of assigned amino acids in SDS and DPC50. In both cases, the data
559
suggest two helical regions separated by a loop. (C) NOE signals of MytC[19-40]SOX in DPC:
560
strong peaks (black), weak signals (grey) and inferred correlations partially masked by other signals
561
(empty bars).
562
Graphic for Table of Contents. The mature myticin C peptide is stabilized by 4 disulfide bonds
563
(in yellow) within the Csαβ motif (in green and blue) and is characterized by two stacking tyrosines
564
(in background a mussel hemocyte activated by heat-killed bacteria).The sequence logo with
565
residue conservation probability is also reported. Synthetic MytC and related fragments were tested
566
on reference Gram positive/negative bacteria and fungal strains, with only weak activity on these
567
latter ones.
568
Table 1. Name, sequence, length and monoisotopic mass of the mature myticin C and related
569
peptide fragments obtained by chemical synthesis.
570
Table 2. Minimal Inhibitory Concentration (MIC) and Minimal Bactericidal or Fungicidal
571
Concentration (MBC or MFC) of the MytC peptide fragments at acidic pH.
25 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 26 of 32
26 ACS Paragon Plus Environment
Page 27 of 32
Journal of Agricultural and Food Chemistry
FIGURE GRAPHICS
Figure 1
27 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 28 of 32
Figure 2
28 ACS Paragon Plus Environment
Page 29 of 32
Journal of Agricultural and Food Chemistry
Figure 3
29 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 30 of 32
Figure 4
30 ACS Paragon Plus Environment
Page 31 of 32
Journal of Agricultural and Food Chemistry
Figure 5
Figure 6
31 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 32 of 32
Graphic for Table of Contents
32 ACS Paragon Plus Environment