Peptides 25 (2004) 2069–2078

Structural and functional characterization of N-terminally blocked peptides isolated from the venom of the social wasp Polybia paulista Susan Pereira Ribeiroa , Maria Anita Mendesa,b,c , Lucilene Delazari dos Santosa,b,c , Bibiana Monson de Souzaa,b,c , Maur´ıcio Ribeiro Marquesa,b,c , Walter Filgueira de Azevedo Jr.c,d , Mario Sergio Palmaa,b,c,∗ a

Department of Biology, CEIS/IBRC, UNESP, Rio Claro, SP, CEP 13506-900, Brazil b Institute of Immunological Investigations/MCT, CNPq, Rio Claro, SP, Brazil c Center for Applied Toxinology CAT/CEPID, FAPESP, Rio Claro, SP, Brazil d Department of Physics, IBILCE-UNESP, S.J.Rio Preto, SP, Brazil

Received 25 June 2004; received in revised form 20 August 2004; accepted 24 August 2004 Available online 2 October 2004

Abstract Two novel peptides were isolated from the crude venom of the social wasp Polybia paulista, by using RP-HPLC under a gradient of MeCN from 5 to 60% (v/v) and named Polybine-I and -II. Further purification of these peptides under normal phase chromatography, rendered pure enough preparations to be sequenced by Edman degradation chemistry. However, both peptides did not interact with phenylisothiocyanate reagent, suggesting the existence of a chemically blocked N-terminus. Therefore, the sequences of both peptides were assigned by ESI-MS/MS under CID conditions, as follows: Polybine-I Ac-SADLVKKIWDNPAL-NH2 (Mr 1610 Da) and Polybine-II Ac-SVDMVMKGLKIWPL-NH2 (Mr 1657 Da). During the tandem mass spectrometry experiments, a loss of 43 a.m.u. was observed from the N-terminal residue of each peptide, suggesting the acetylation of the N-terminus. Subsequently, the peptides with and without acetylation were synthesized on solid phase and submitted to functional characterizations; the biological activities investigated were: hemolysis, chemotaxis of polymorphonucleated leukocytes (PMNL), mast cell degranulation and antibiosis. The results revealed that the acetylated peptides exhibited more pronounced chemotaxis of PMNL cells and mast cell degranulation than the respective non-acetylated congeners; no hemolytic and antibiotic activities were observed, irrespective to the blockage or not of the ␣-amino groups of the N-terminal residues of each peptide. Therefore, the N-terminal acetylation may be related to the increase of the inflammatory activity of both peptides. © 2004 Elsevier Inc. All rights reserved. Keywords: Polybia paulista; N-terminally blocked peptides; Tandem mass spectrometry; Inflammatory peptides; Wasp venom toxins

1. Introduction The venoms of social Hymenoptera such as hornets, paper wasps and honeybees, are used both for self-defense and the defense their larvae from predators and intruders of the nest. The stinging caused by these insects produce severe pain, inflammation, local tissue damage and occasionally death in large vertebrates, including man [9]. Hymenoptera venoms are complex mixtures of biochemical and pharmacologically ∗

Corresponding author. Tel.: +55 19 3526 4163; fax: +55 19 3534 8523. E-mail address: [email protected] (M.S. Palma).

0196-9781/$ – see front matter © 2004 Elsevier Inc. All rights reserved. doi:10.1016/j.peptides.2004.08.019

active components such as biogenic amines, peptides and proteins. The composition of neotropical vespid venoms has been subjected to few investigations, since the production of venoms by social wasps is very small and most of wasp species usually present a population of some tens of individuals; thus, there is a limited availability of venoms as raw materials [26]. Wasps venoms are composed of a complex mixture of free amino acids, biogenic amines, a series of polycationic peptides and proteins [9]. The polycationic peptides represent the major group of components of vespid venoms; the most important families of cationic peptides are represented

2070

S.P. Ribeiro et al. / Peptides 25 (2004) 2069–2078

by the mastoparans, chemotactic peptides and wasp kinins [21]. In addition, the neotropical vespid venoms still present other polycationic peptide components, not characterized up to now. Polycationic peptides have been reported to be involved with the occurrence of inflammation, including initial lysis of cell membrane or mast cell degranulation, leading to histamine release and consequent vasodilatation. Increasing chemotaxis of neutrophils and T helper cells resulting in leukocytes recruitment to the inflammation site generally is also observed for these peptides. This type of peptides in general contain from 12 to 50 amino acids residues, with a net positive charge between +2 and +7 due to the excess of basic amino acid residues and also present more than 50% of their amino acid sequences constituted of hydrophobic residues [8]. These structural features may attribute to these peptides amphipathic, ␣-helical conformation which in turn is related to their ability to interact with the anionic components of the bacterial membranes, providing their assembling in these membranes associated with pores formation [4]. Because of this, some of these peptides also can present antimicrobial actions [5,12,15]. The amphipathic, ␣-helical conformations also may permit the assembling of some of these cationic peptides with the zwitterionic membranes of mammalian cells, making some of these peptides to act as hemolysins [5,15]. The primary structural information is the basis for understanding the secondary structure and function of these peptides, which requires reliable and fast amino acid sequencing. Among existing methods for sequence determination, Edman degradation chemistry is the notably the most used one. However, some proteins and peptides cannot be directly sequenced by Edman degradation because they have a blocked N-terminal residue. N-acetyl ‘blocked’ termini are common among the eukaryotic proteins, for which more than 10 kinds of blocking groups are currently known, such as: methyl, acetyl, formyl or pyroglutamyl [11]. A series of specific chemical reagents and/or enzymes have been used to remove these blocked groups; however, some kinds of deblocking reagents present side reactions in some peptide bonds and certain side chains of the amino acids [13,27]. So, this process generally spends lots of time, requiring relatively high amounts of sample and even analyze with difficulty [10]. Mass spectrometry provides the capability for sequence determination because of the breakthrough of mass spectrometric techniques such as: electrospray ionization (ESI), tandem mass spectrometry (MS/MS) [20] and matrix-assisted laser desorption ionization (MALDI) with post-source decay (PSD) capability [26]. Peptide sequencing became one of the major fields of mass spectrometry, driven by the growing analytical demand. This type of application has some important advantages compared with classical sequencing methods such as Edman degradation: (I) the ability to perform sequence analysis of a peptide within a mixture because it can be mass selected by the instrument; (II) fragmentation of a molecular ion is possible even

in the presence of an N-terminal modification of the peptide, a considerable problem for the Edman degradation protocols; (III) the sample amount necessary for analysis is usually less than 1 pmol to obtain a high quality mass spectrum, which provides the MW of the sample as well as an indication of its purity [26]. In the present work two novel polycationic peptides were isolated from the crude venom of social wasp Polybia paulista, by using RP-HPLC under a gradient of MeCN from 5 to 60% (v/v). Further purification of these peptides, rendered pure enough preparations to be sequenced by Edman degradation chemistry. However, both peptides did not react with phenylisothiocyanate reagent, suggesting the existence of a chemically blocked N-terminus. Thus, the sequences of both peptides were assigned by ESI-MS/MS. Subsequently, the peptides with and without acetylation were synthesized on solid phase and submitted to functional characterizations.

2. Material and methods 2.1. Sample preparation The wasps were collected in Rio Claro-SP, southeast Brazil, and immediately frozen and stored at −20 ◦ C. The venom reservoirs of 500 worker wasps were removed by dissection with surgical micro scissors and washed with 1:1 acetonitrile (MeCN, Aldrich):water containing 0.1% (v/v) trifluoroacetic acid (TFA, Aldrich) to solubilize the peptides. The extract was then centrifuged at 10,000 × g for 15 min at 4 ◦ C; the supernatant was collected, dried by centrifugation under reduced pressure in a Speed-Vac system, and maintained at −20◦ C until be used. 2.2. Peptide purification The biological material from the dried extract was solubilized in 5% (v/v) MeCN at a concentration of 100 ␮g/mL and chromatographed under RP-HPLC conditions in a Shiseido Nucleosil C-18 (ODS) column (250 mm × 10.0 mm; 5 ␮m), at a flow rate of 2 mL/min, using a gradient from 5 to 60% (v/v) MeCN (containing 0.1% TFA), at 30 ◦ C for 45 min. The elution was monitored at 215 nm with a UV-DAD detector (Shimadzu, model SPD-M10A) and each peak eluted was manually collected in 5-mL glass vials. A peak of interest was submitted to chromatography under normal phase conditions using a Brownlee CN-column (250 mm × 4.6 mm, 5 ␮m), under isocratic elution with 50% (v/v) MeCN (containing 0.1% TFA) at a flow rate of 600 ␮L/min, for 25 min at 30 ◦ C. The elution was monitored at 215 nm, and fractions were manually collected in 5-mL glass vials. 2.3. Mass spectrometry analysis ESI-MS mass spectra were acquired on a triple quadrupole (Quatro II) mass spectrometer instrument (Micromass, UK),

S.P. Ribeiro et al. / Peptides 25 (2004) 2069–2078

equipped with a standard electrospray probe, adjusted to ca. 5 ␮L min−1 . During all experiments, the source temperature was maintained at 80 ◦ C and the needle voltage at 3.6 kV, applying a drying gas flow (nitrogen) of 200 L h−1 and a nebulizer gas flow (nitrogen) of 20 L h−1 . The mass spectrometer was calibrated with intact horse heart myoglobin and its typical cone-voltage induced fragments. The cone sample to skimmer lens voltage, controlling the ion transfer to the mass analyzer, was maintained at 30 V. About 50 pmol of each sample was injected into electrospray transport solvent. The ESI spectra were obtained in the continuous acquisition mode, scanning from m/z 100 to 2000 with a scan time of 5 s. The N-terminal blocked peptides were sequenced through the technique of tandem mass spectrometry by using positive electrospray ionization (ESI+). Typical conditions were: a capillary voltage of 3 kV, a cone voltage of 30 V, collision gas pressure of 3.5 × 10−3 mBar and a desolvation gas temperature of 80 ◦ C. The double charged diprotonated molecules of the precursors (presenting peak widths of 0.14 Da) were selected in Q1 and subjected to collision-induced dissociation (CID) with argon gas at 50 eV collision energy; the product ions (presenting peak widths 0.14 Da) were detected in Q3. All the MS/MS experiments were performed using the Quatro II triple-quadrupole instrument. The ESI spectra were obtained in the continuous acquisition mode, scanning from m/z 40 to the m/z value of the [M + H]+ ion of each peptide, with a scan time of 5 s. The spectral data were acquired and monitored by using the MassLynx software. Peptide sequences were assigned manually from the ESIMS/MS product ion mass spectra with the help of the PepSeq software (Micromass, UK) as described elsewhere [3,20]. 2.4. Derivatization of lysine residues In order to distinguish between the isobaric amino acid residues lysine and glutamine, the peptides were submitted to acetylation as follows: 5 ␮L of acetic anhydride was add to 5 ␮L of native peptide solution; the solution was then vortexed and incubated for 45 min at 37 ◦ C under shaking. After incubation the solutions were centrifuged and analyzed by ESI-MS. 2.5. ESI-MS analysis of chemically modified peptides After the acetylation reaction, the molecular masses of the acetylated peptides were analyzed by ESI-MS. About 1 pmole of each peptide was injected into electrospray transport solvent by using a microsyringe (50 ␮L) coupled to a micro infusion pump (KD Scientific) at a flow rate of 4 ␮L min−1 . The spectra were acquired in the continuous acquisition mode, scanning from m/z 100 to 2500 at a scan time of 5 s, using a capillary voltage of 3 kV, a cone voltage of 30 V and a desolvation gas temperature of 80 ◦ C.

2071

2.6. C-terminal sequencing The C-terminal sequencing of the peptides was performed based on derivatization with acetylisothiocyanate to yield amino acid thiohydantoins (TH-Aas; this simple chemistry was automated using a ABI 473A N-terminal sequencer. All reagents (R1, trimethylsilylisothiocyanate; R3, alkaline thiocyanate for cleavage) and solvents required for sequencing were accommodated on the sequencer, which was modified to deliver liquid R2 (acetyl chloride) to the reaction vessel. The conversion flask was used for preparing the TH-AAs for analysis by on-line HPLC, using a graphitized carbon (Hypercarb) column. This setting was permited to determine at least five residues from the C terminus. 2.7. Peptide synthesis and purification The peptides were prepared by step-wise manual solidphase synthesis using N-9-fluorophenylmethoxy-carbonyl (Fmoc) chemistry with Novasyn TGS resin (NovaBiochem). Side-chain protective groups included t-butyl for serine and t-butoxycarbonyl for lysine. To prepare peptides presenting the N-terminal residue acetylated, after the coupling of the last amino acid residue, acetic anhydride was added to the peptide-resine complexes and maintained under shaking for 1 h to promote the acetylation. Cleavage of the peptide-resin complexes was achieved by treatment with trifluoroacetic acid/1,2-ethanedithiol/ anisole/phenol/water (82.5:2.5:5:5:5 by volume), using 10 mL g−1 of complex at room temperature for 2 h. After filtration to remove the resin, ethyl ether at 4 ◦ C was added to the soluble material causing precipitation of the crude peptides, which were collected as a pellet after centrifugation at 1000 × g, for 15 min at room temperature. The crude peptides were solubilized in water and chromatographed under RPHPLC conditions using a semi-preparative column (Shiseido C18, 250 mm × 10 mm, 5 ␮m), under isocratic elution with 40% (v/v) acetonitrile in water containing 0.1% (v/v) trifluoroacetic acid at a flow rate of 2 mL/min. The elution was monitored at 215 nm with a UV-DAD detector (Shimadzu, model SPD-M10A) and each fraction eluted was manually collected in 2 mL plastic vials. The homogeneity and sequences of the synthetic peptides (both acetylated and non-acetylated ones) were assessed by analytical HPLC and ESI-MS analysis. 2.8. Biological activities Mast cell degranulation was determined by measuring the release of ␤-d-glucosaminidase, which co-localizes with histamine, as proposed by Hide et al. [6]. Mast cells were obtained by peritoneal washing of adult Wistar rats with a solution containing 0.877 g NaCl, 0.028 g KCl, 0.043 g NaH2 PO4 , 0.048 g KH2 PO4 , 0.10 g glucose, 0.10 g BSA, 90 ␮L CaCl2 (2 mol/L) solution and 50 ␮L Liquemine, in

2072

S.P. Ribeiro et al. / Peptides 25 (2004) 2069–2078

100mL water. The cells were incubated in the presence of the peptides for 15 min at 37 ◦ C. After centrifugation the supernatants were sampled for ␤-d-glucosaminidase assay. Briefly, 50 ␮L of each sample was incubated in 50 ␮L of the substrate (3 mg of p-nitrophenyl-N-acetyl-␤-d-glucosaminidine dissolved in 10 mL of 0.2 mol/L sodium citrate solution at pH 4.5) for 6 h at 37 ◦ C. The absorbance of the colored product was assessed at 405 nm and the values were expressed as the percentage of total ␤-d-glucosaminidase activity from rat mast cell suspensions, determined in lysed mast cells in presence of 0.1% (v/v) Triton X-100 (considered as 100% reference). The hemolytic activity of the peptides were determined as follows: 50 ␮L of red blood cell suspensions from Wistar rats were washed three times with physiological saline solution (NaCl, 0.85% and CaCl2 , 10−3 mol/L) and suspended in 50 mL of the same solution; this cell suspension was designated as washed rat red blood cells (WRRBC). Aliquots of WRRBC were incubated at 37 ◦ C in presence of each peptide, for 120 min with gentle mixing. The samples were then centrifuged and the absorbance of the supernatants was measured at 540 nm. The absorbance measured from lysed WRRBC in presence of 1% (v/v) Triton X-100 was considered to be 100%. The chemotaxis activities were assayed in a specific multi chamber apparatus (NEURO PROBE) by using polymorphonucleated leukocytes (PMNL), obtained from subcutaneous inflammatory induction in Wistar rats. The upper chambers were filled with 200 ␮L of a PMNL suspension (∼2.7 × 105 cell/mL in 0.9% NaCl solution—physiological saline solution) and the lower chambers were filled with 400 ␮L of physiological saline solution containing the peptides (about 500 ng per assay). A polycarbonate membrane containing pores of 10 ␮m of diameter (NEURO PROBE) was placed between both chambers. The chemotaxis chamber was incubated at 37 ◦ C for 1 h. After incubation, cells in the lower chamber were counted using a Ne¨ubauer chamber. The antimicrobial activity was assayed by using the disc diffusion method as described below. Colonies of the bateria Staphylococcus aureus (ATCC 6538), Bacilus turinghiensis, Escherichia coli (ATCC 25922) and Psudomonas aeruginosa (ATCC 15442) were used in the present assay. Transferring cells made inocula of the colonies of each bateria species from solid culture media with a wire loop to a tube containing 2 mL of M¨uller Hinton broth medium. The liquid inocula were incubated at 34 ◦ C until they achieved a turbidity of 0.5 against a McFarland standard scale; Petri dishes (10 mL) were filled with 5 mL of melted M¨uller Hinton Agar and mixed with 20 ␮L of individual bacterial inoculum. The diffusion disks (6 mm of diameter) containing different concentrations of each peptide were applied on the surface of agar media, and the Petri dishes were incubated at 34 ◦ C for 24 h. After this incubation time, the diameters of the inhibition zones were measured.

2.9. Statistical analysis Results are expressed as means ± S.D. of five experiments. Differences between two groups (acetylated and nonacetylated peptides) were evaluated with Students’ t-tests. Values of P < 0.05 were considered significant.

3. Results Edman degradation is used as the classical method to get the amino acid sequence of a protein and/or of a peptide. However, many polypeptides cannot be directly sequenced by Edman degradation because they have a blocked N-terminal residue. Several published procedures have been combined to develop a general strategy for the specific identification of a N-terminally blocked amino acid residue, its removal and amino acid sequencing of the removing chain [13,14,27]. We have found two novel peptides extracted from the venom of the wasp Polybia paulista. The chromatogram obtained by RP-HPLC (Fig. 1a) reveals 13 peaks (designated 1–13). Fractions 1–3 are constituted of free amino acids and biogenic amines; fraction 4 was characterized as serotonine; fractions 5–9 are very low abundant venom components not characterized up to now; fractions 11 and 12 are constituted of mastoparans peptides biochemically and functionally characterized in a previous publication [26]. Fraction 13 showed pronounced mast cell degranulation activities. However, the ESI/MS spectrum of this fraction showed a series of different compounds, with peptides of molecular mass 1610 and 1657 Da (not showed results) representing the major components. The purification of that fraction by using HPLC under normal-phase conditions resulted in two new fractions (Fr-13.1 and Fr-13.2) as showed in the Fig. 1b. The novel peptides were identified by their molecular masses, 1610 Da (in Fr-13.1) and 1657 Da (in Fr-13.2), respectively, as shown by the ESI mass spectra of the isolated peptides (Fig. 2a and b); these spectra contain both [M + H]+ ions at m/z 1611 and 1658 for Fr-13.1 and Fr-13.2, respectively, and the [M + 2H]2+ ions at m/z 807 and 830. At first, we tried to determine the amino acid sequence by Edman degradation chemistry. However, it was not possible, since there was no coupling reaction between the ␣-amino groups of the N-terminal residue of both peptides with the reagent phenylisothiocyanate, suggesting the existence of Nterminally modified ␣-amino groups in both peptides. Afterward, the amino acid sequence has been determined by ESI-MS/MS. 3.1. Characterization by tandem mass spectrometry of peptides In order to establish the peptides structural identities the CID mass spectra of the [M + H]+ ion at mass-to-charge (m/z 1611 and 1658) were generated. The mass spectra shown in Fig. 3a and b are an example of the structural information

S.P. Ribeiro et al. / Peptides 25 (2004) 2069–2078

2073

Fig. 1. (a) Chromatogram profile of fractionation of the crude venom from the wasp Polybia paulista under reversed-phase HPLC with a Nucleosil C-18 (ODS) SHISEIDO column (250 mm × 10 mm), under linear gradient from 5 to 60% (v/v) MeCN (containing 0.1% TFA), at a flow rate of 2.0 mL/min over 60 min by monitoring at UV 214 nm. (b) Rechromatography of fraction 13 under normal phase conditions using a Brownlee CN-column (250 mm × 4.6 mm, 5 ␮m), under isocratic elution with 50% (v/v) MeCN containing 0.1% TFA at a flow rate of 600 ␮L/min, for 25 min at 30 ◦ C.

that can be obtained from the CID mass spectrum of both peptides. Analysis of the product ion spectra shows peaks at the low mass end probably representing the imonium ions (H2 N CHR)+ (where R is a side chain group). The presence of certain amino acids can be deduced from peaks at high mass correspond to ions formed by the loss of an amino acid side chain from the precursor ion [M + H]+ . By searching for consecutive peaks of mass difference corresponding to the mass of the NH CHR CO residue of natural amino acid residues, it is possible to recognize one or more series of peaks. In the Fig. 3a and b, those peaks were assigned as a series of ‘y’ and ‘b’ types ions-fragments for both peptides. The complete identification is confirmed by the presence of the others ions series and also by the presence of the peaks, which correspond to neutral loss (NH2 or H2 O) from the main peaks.

Fig. 2. ESI mass spectra for the peptides present in the fraction 13 from the crude venom of Polybia paulista: (a) Polybine-I and (b) Polybine-II. The ions [M + H]+ and [M + 2H]2+ ions are observed for both peptides and were assigned as A1 and A2, respectively.

In Fig. 3a is represented the tandem mass spectrum of the peptide of fraction 13.1, where the first major ion in the high mass end was assigned with m/z 1611 (molecular ion as [M + H]+ ); if considered the lost of ammonia (17 a.m.u.) from the [M + H]+ ion, this will result in the ion of m/z 1594. Thus, a series of b-ions can be observed in this spectrum: m/z 1567 (b14 ), m/z 1479. (b13 ), m/z 1408 (b12 ), m/z 1313 (b11 ), m/z 1198 (b10 ), m/z 1083 (b9 ), 897.23 (b8 ), m/z 784 (b7 ), m/z 656 (b6 ), m/z 528 (b5 ), m/z 429 (b4 ), and m/z 316. (b3 ), m/z 200 (b2 ) and m/z 129 (b1 ). The subtraction of the m/z values between consecutive b-ions permitted the assignment of the sequence of the peptide with a few ambiguities in relation to the isobaric residues (I/L and K/Q), as showed in the Fig. 3a. Another interesting feature in that mass spectrum is the presence of the m/z 1567.43 Da (Fig. 3a), indicating the loss of 43 mass units from the monoprotonated molecular, which corresponds to the elimination of acetyl group from the first amino acid residue from the N-terminal position. Thus, the sequence of the peptide from fraction 13.1 is: AcS-A-D-L/IV-K/Q-K/Q-I/L-W-D-N-P-A-L/I. The tandem mass spectrum of the peptide of fraction 13.2 is showed in the Fig. 3b, which was submitted to the same analysis and the b-type ion series was also found to be the major series of ions; the first major ion in the high mass end was assigned with m/z 1658. (molecular ion as [M + H]+ ); if considered the lost of ammonia (17 a.m.u.) from the [M + H]+ ion, this will result in the ion of m/z 1641 (b14 ). Thus, a series of b-ions can be observed in this spectrum: m/z 1528.22 (b13 ), m/z 1430 (b12 ), m/z 1244 (b11 ), m/z 1131 (b10 ), m/z 1003 (b9 ), m/z 890 (b8 ), m/z 833 (b7 ), m/z 705 (b6 ), m/z 574 (b5 ), m/z 475 (b4 ), m/z 344 (b3 ), m/z 230 (b2 ) and m/z 129 (b1 ). The subtraction of the m/z values between consecutive b-ions permitted the assignment of the sequence of the peptide also with a few ambiguities in relation to the isobaric residues (I/L and K/Q). The presence of the m/z 1615.60 in-

2074

S.P. Ribeiro et al. / Peptides 25 (2004) 2069–2078

Fig. 3. The ESI-MS/MS spectra of the [M + H]+ ion, acquired in the 60–70 eV collision energy range, acquired for the molecular ions as [M + H]+ : (a) m/z 1611 and (b) m/z 1658. The mass differences between the consecutive bn ions and their correspondence to the amino acid sequence are shown.

dicates the loss of acetyl group (Fig. 3b) indicating the loss of 43 mass units from the monoprotonated molecular, which corresponds to the elimination of acetyl group from the first amino acid residue from the N-terminal position. Thus, the sequence indicated by the MS/MS spectrum of Fig. 3b is: AcS-V-D-M-V-M-K/Q-G-L/I-K/Q-I/L-W-P-L/I. These experiments allowed the determination of the amino acid sequences of both peptides as shown in Fig. 3a and b. However, the consideration of only the b- and y-types of fragment-ions still leaves unknown the choices between the isobaric residues I/L and K/Q. In order to address these questions, we looked for d-type fragment ions that permit the distinction between I and L [20]. The analysis of the ESI-MS/MS spectrum for the peptide of fraction 13.1 (Fig. 3a–b) revealed the presence of ions of m/z 359 and 841, assigned as d4 and d8 corresponding to Ile and Leu residues, respectively. The peptide of fraction 13.2 was submitted to the same process (Fig. 3b) and ions of m/z 933 and 1188 were assigned as d9 and d11 ions, corresponding to Ile, and, Leu, respectively. The d-type fragment-ions are

formed due to the differential fragmentation of the side chains of these amino acid residues; thus, the distinction between the I/L residues based on the search of d-type ions is not effective, when these residues are localized at the C-terminal position, since the fragmentation of side chain does not occur at this position [20,26]. The distinction between K and Q was achieved through the use of acetylation of the peptides, which occurs for ␣and ␧-amino groups; thus only unblocked N-termini and the side-chains of K residue react. The intact peptides of fractions 13.1 and 13.2 were submitted to acetylation reaction to determine whether a lysine residue was present; after in vitro acetylation, the molecular masses of the modified peptides were analyzed by ESI-MS. The [M + H]+ ions were observed at m/z 1695.0 and 1742.0 for peptides of fractions 13.1 and 13.2, respectively (spectra not shown). Thus, it was possible to observe that the masses for both peptides increased by 84 Da (corresponding to 42 Da × 2), indicating that there were two sites of acetylation in each peptide. Since the ␣amino group from the N-terminal residue was acetylated in

S.P. Ribeiro et al. / Peptides 25 (2004) 2069–2078

both them, the results above suggest the presence of two Lys residues in each peptide. Therefore, the putative K/Q residues in both peptides were all identified as lysine. Consequently, taken into account the results obtained up to this point the sequence of peptides of fractions 13.1 and 13.2 are: peptide 13.1 Ac-S-A-D-I-V-K-K-L-W-D-N-P-A-L/I; peptide 13.2 Ac-SV-D-M-V-M-K-G-I-K-L-W-P-L/I. These results still show ambiguities concerning to the Cterminal distinction between the residues L and I. However, the C-terminal sequencing revealed the following partial sequences: peptide 13.1: D-N-P-A-L; peptide 13.2: K-L-W-PL. In spite to have permitted the assignment of only five amino acid residues from the C-teminal side of each peptide, these results corroborated the previous sequence assigment made by tandem mass spectrometry and also solved the remaining ambiguity between the isobaric residues I/L at the C-terminus for both peptides. It must be emphasized that the molar masses of these peptides fit to the sequences above, considering the C-terminal residues in the amidated form, as already observed for the most of peptide toxins from Hymenoptera venoms [12]. The similarities and/or homologies with other known polypeptides were performed with the protein/peptide engine search tool Mascot (http://www.matrixscience.com) that revealed no similarity with any known toxic peptide or toxin. Since these peptides constitute novel toxins first time described in animal venoms, we named them by taking into consideration the wasp scientific name as Polybine-I and -II. Thus, considering the results above the complete sequences of both peptides are: Polybine-I Ac-S-A-D-I-V-K-K-L-W-D-N-P-A-L-NH2 Polybine-II Ac-S-V-D-M-V-M-K-G-I-K-L-W-P-L-NH2 3.2. Biological activities After determined that the N-terminus were blocked in both peptides, it was wondered about the role of the N-terminal acetylation of these peptides. For this reason, both peptides were synthesized, with and without acetylation at the N-terminal residue and submitted to biological characterizations. Therefore, mast cell degranulation, hemolysis, chemotaxis and antimicrobiosis activities were assayed for all peptides. Fig. 4 shows the results of rat peritonial mast cell degranulation activities for peptides Polybine-I and -II and their non-acetylated. Non-acetylated forms of the Polybine peptides showed reduced mast cell degranulation activities from 10−8 to 10−4 mol/L. However, both Polybine-I and -II promoted potent degranulation of rat peritonial mast cells; a concentration of 10−6 mol/L, which is within the physiological concentration of these peptides under natural conditions, resulted in 58% degranulation. Thus, apparently the acetylation of the ␣-amino group of the N-teminal serine residue of both Polybine peptides increased in more than 40% their power of

2075

Fig. 4. Degranulation activity in rat peritoneal mast cell. The activity was determined by measuring the release of the granule marker, ␤-dglucosaminidase, which co-localizes with histamine and the values for ␤-dglucosaminidase released in the medium were expressed in the percentage of total ␤-d-glucosaminidase. Values are mean ± S.D. (n = 5); (*) P < 0.05 vs. non-acetylated peptide.

Fig. 5. Hemolytic activity in washed rat red blood cells (WRRBC). The absorbance measured at 540 nm from lysed WRRBC in presence of 1% (v/v) Triton X-100 was considered as 100%. Values are mean ± S.D. (n = 5).

degranulation of rat peritoneal mast cells, when compared to the non-acetylated forms (significant at level of P < 0.05). Hemolytic activity was also examined. As observed in the Fig. 5, no peptide presented hemolytic activity, despite to be acetylated or not. The same way, the assays of antibiosis revealed that the Polybine peptides are not active both against Gram-positive and Gram-negative bacteria (results not shown). The assays of chemotaxis revealed that Polybine-I and -II constitute more potent chemoattractant peptides for PMNL cells, than their non-acetylated congeners (Fig. 6a and b). The acetylated forms are from 30 to 100% more potent that the non-acetylated ones, depending on each peptide concentration (significant at level of P < 0.05).

4. Discussion Among the Hymenopteram peptide toxins is very rare the natural occurrence of in vivo chemical modifications of the ␣-amino group of the N-terminal amino acid residue, while

2076

S.P. Ribeiro et al. / Peptides 25 (2004) 2069–2078

Despite to be potent mast cell degranulators and chemoatractants of PNML cells, Polybine-I and -II do not present sequence similarity neither with the mastoparans nor with the classical wasp chemotactic peptides. Thus, the Polybines constitute a novel family of inflammatory peptides from wasp venoms, not related to the previously known peptide toxins, probably presenting novel mechanisms of action. Mast cells stimulated by immunoglobulin-E, release some chemical mediators (e.g., histamine) by exocytosis. They can, however, also be triggered to undergo exocytosis by nonantigenic stimuli. Compounds such as substance P, neuropeptide Y, wasp venom mastoparans, a series of different polycationic mast cell degranulating peptides [16,22] and the compound 48/80 [25] can stimulate the release of vasoactive and inflammatory compounds from mast cells. The mechanism by which cationic peptides stimulate mast cell exocytosis remains unclear; it is generally accepted that heterotrimeric GTP binding proteins (G proteins) are involved in the stimulation of mast cell degranulation by basic secretagogues. [23] The delivering of stored compounds from mast cells granules may occur either due to the cytolytic effect of the peptides or due to exocytosis activated by the binding of the peptides to G-proteins [7]. However, there are at least two mechanisms proposed to explain mast cell degranulation based in the interaction of polycationic peptides with the G-proteins: Fig. 6. Histogram showing comparatively the chemotaxis assays of PMNL cells to the peptides: (a) non-acetylated Polybine-I and Polybine-II and (b) non-acetylated Polybine-I and Polybine-II. Bars represent the mean ± S.D. (n = 5); (*) P < 0.05 vs. non-acetylated peptide.

the large most of these toxins usually present amidated Cterminus. The sequencing of Polybine-I and -II by using ESI-MS/MS showed the power of tandem mass spectrometry for the structural assignment of the blocking group at the N-terminal position of each peptide and their subsequent sequencing. The present investigation presents the first description of this type of post-translational modification among the Hymenopteram peptide toxins. Polybine-I and -II are polycationic peptides, however, the biological activities typically associated to ␣-helical conformations, such as hemolysis and antibiosis, were not observed for these peptides, despite the condition of their N-termini to be or not blocked. Since no experimental data were accquired up to now to characterize the secondary structures these peptides, the primary sequence of Polybine-I and -II were submitted to the evaluation of the algorithm PSIPRED–secondary structure prediction method [11], by using the PSIPRED server [19]. The results revealed that Polybine-I presents a propensity of only 40% ␣-helix in the midle of the molecule, located between two coiled regions in the extremities; meanwhile, Polybine-II presented 100% propensity to be a coil. These predictions of low/no extension of ␣-helical conformations for the Polybine peptides apparently may explain the absence of hemolytic and antibiotic activity.

(i) The mastoparans have been proposed to stimulate secretion from mast cells, by directly activating GTP-binding regulatory proteins (G proteins), in a manner similar to that of G-protein coupled receptors by a mechanism remarkably similar to that used by agonist-bound receptors [8]. Mastoparans can bind to G(i)- and G(0)-subunits adopting an amphiphilic alpha-helical conformation in both cases, requiring a receptor for its binding. The lysine residues are known to be crucial for activity of these peptides and it is thus, likely that at least the polar face of the amphiphylic helix must be in contact with the Gproteins [8]. (ii) The polycationic non-mastoparan peptides are proposed to stimulate mast cells as result of their interaction with negatively charged sites (e.g., sialic acid containing glycoproteins) at the mast cell membrane [2,23]. In this mechanism there is a direct interaction of polycationic peptides with heterotrimeric G-proteins through a receptor-independent mechanism of peptide-induced mast cell degranulation [23], probably involving the interaction of the polycationic peptides with G(␣)-subunit subsequent to their translocation across the plasma membrane [18]. In contrast to the requirements for the activation of purified G(i)-proteins by mastoparan peptides, amphipathic helicity was proposed to be not obligatory for non-mastoparan peptide-induced mast cell activation [1]. Apparently, this observation could support the Polybine peptides to act through this mechanism.

S.P. Ribeiro et al. / Peptides 25 (2004) 2069–2078

Chemotaxis is the phenomenon in which bacteria, other organisms, or single cells of multicellular organisms to direct their movements according to certain chemicals in their environment. The recruitment of leukocytes to a site of tissue injury (e.g., a wasp stinging) also constitutes a leading cause for inflammatory responses. Mechanistically, it involves a cascade of cellular events precisely regulated by temporal and spatial presentation of a repertoire of molecules in the migrating leukocytes and their surroundings (microenvironments) [17]. For the most part, eukaryotic cells sense the presence of chemotactic stimuli by using stereospecific 7-transmembrane heterotrimeric G-protein coupled receptors; the activity of the chemotactic peptides generally is dependent of specific signals mediated by G-proteins located on the plasma membrane of the chemoattracted cells, making the interaction cell/peptide relatively selective [24]. Despite the mechanism of action of the Polybines is not known, these peptides clearly play a potent chemotactic effect on rat PMNL cells. Therefore, both the results of mast cell degranulation and chemotaxis of PMNL cells are pointing to the possibility that the Polybine toxins, naturally occurring as polycationic N-terminally acetylated peptides, may be related to the induction of inflammatory processes, by interacting with Gproteins. The acetylation of the ␣-amino group at the Nterminus of these peptides may play the role of biological activities enhancer; but also, it may be speculated another concomitant role, as a protective group to prevent the action of some exoproteinases over the peptides.

Acknowledgments This work was supported by a grant from the S˜ao Paulo State Research Foundation (FAPESP, SMOLBNet 01/075320). Maria Anita Mendes is Postdoctoral fellow from FAPESP (Proc. 01/05060-4), Bibiana Monson de Souza and Maur´ıcio Ribeiro Marques are Doctoral students fellows from FAPESP. Susan Pereira Ribeiro is undergraduate student; Mario Sergio Palma (Proc. 300377/2003-5) and Walter Filgueira de Azevedo Junior (Proc. 300851/98-7) are researching for the Brazilian Council for Scientific and Technological Development (CNPq).

[4] [5] [6]

[7]

[8]

[9]

[10]

[11] [12]

[13]

[14] [15]

[16]

[17]

[18]

[19] [20]

References [1] Cross MLJ, Ennis M, Krause E, Dathe M, Lorenz D, Krause G, et al. Influence of ␣-helicity, amphipathicity and d-amino acid incorporation on the peptide induced mast cell activation. Eur J Pharmacol 1995;291:291–300. [2] Emadi-Khiav B, Mousli M, Bronner C, Landry Y. Human and rat cutaneous mast cells: involvement of a G-protein in the response to peptidergic stimuli. Eur J Pharmacol 1995;272:97–102. [3] Fontana R, Mendes MA, Souza BM, Konno K, C´esar LMM, Malaspina O, et al. Jelleines: a family of antimicrobial pep-

[21]

[22]

[23]

2077

tides from the Royal Jelly of honeybees (Apis mellifera). Peptides 2004;25:919–28. Gallo RL, Huttner KM. Antimicrobial peptides: an emerging concept in cutaneous biology. J Investig Dermatol 1998;111:739–43. Hancock REW, Chapple DS. Peptide antibiotics. Antimicrob Agents Chemother 1999;43:1223–317. Hide I, Bennett JP, Pizzey A, Boonen G, Sagi DB, Gomperts BD, et al. Degranulation of individual mast-cells on response to Ca2+ and guanine-nucleotides–an all-or-none event. J Cell Biol 1993;3:585– 93. Higashijima T, Uzu S, Nakajima T, Ross EM. Mastoparan a peptide toxin from wasp venom, mimics receptors by activating GTP-binding regulatory proteins (G proteins). J Biol Chem 1988;263:6491–4. Higashijima T, Burnier J, Ross EM. Regulation of Gi and G0 by mastoparan related peptides and hydrophilic amines. J Biol Chem 1990;265:14176–86. Hirai Y, Yasuhara T, Yoshida H, Nakajima T, Fujino M, Kitada C. A new mast cell degranulating peptide ‘mastoparan’ in the venom of Vespula lewisii. Chem Pharm Bull 1979;27:1942–4. Huang RH, Wang DC. Mass spectrometric strategy for primary structure determination of N-terminally blocked peptides. J Chromatogr B 2004;803:167–72. Jones DT. Protein secondary structure prediction based on positionspecific scoring matrices. J Mol Biol 1999;292:195–202. Konno K, Hisada M, Fontana R, Lorenzi CCB, Naoki H, Itagaki Y, et al. Anoplin, a novel antimicrobial peptide from the venom of the solitary wasp Anoplius samariensis. Biochim Biophys Acta 2001;1550:70–80. Kouach M, Belalche D, Jaquinod M, Couppez M, Kmiecik D, Ricart G, et al. Application of electrospray and fast atom bombardment mass spectrometry to the identification of post-translational and other chemical modifications of proteins and peptides. Biol Mass Spectrom 1994;23:283–94. Krishna RG. Techniques in protein chemistry, vol. 111. San Diego: Academic Press; 1992. p. 77–84. Krishnakumari V, Nagaraj R. Antimicrobial and hemolytic activities of crabrolin, a 13-residue from the venom of the European hornet, Vespa crabro, and its analogs. J Pep Res 1997;50:88–93. Landry Y, Bronner C, Mousli M, Fischer T, Vall A. The activation of mast cells: molecular targets and transducing processes for antigenic and non-antigenic stimuli. Bull Inst Pasteur 1992;90:83–98. Le Y, Gong W, Li B, Dunlop NM, Shen W, Su SB, et al. Utilization of two seven-transmembrane, G-protein coupled receptors, formyl peptide receptor-like 1 and formyl peptide receptor, by the synthetic hexapeptide WKYMVm for human phagocyte activation. J Immunol 1999;163:6777–84. Lorenz D, Wiesner B, Zipper J, Winkler A, Krause E, Beyermann M, et al. Mechanism of peptide-induced mast cell degranulation translocation and Patch-Clamp Studies. J Gen Physiol 1998;112:577– 91. McGuffin LJ, Bryson K, Jones DT. The PSIPRED protein structure prediction server. Bioinformatics 2000;16:404–5. Mendes MA, Souza BM, Santos LD, Palma MS. Structural characterization of novel chemotactic and mastoparan peptides from the venom of the socialwasp Agelaia pallipes pallipes by highperformance liquid chromatography/electropray ionization tandem mass spectrometry. Rapid Commun Mass Spectrom 2004;18:636– 42. Mendes MA, Souza BM, Marques MR, Palma MS. Structural and biological characterization of two novel peptides from the venom of the neotropical social wasp Agelaia pallipes pallipes. Toxicon 2004;44:67–74. Mousli M, Bronner C, Bueb JL, Tschirhart E, Gies JP, Landry Y. Activation of rat peritoneal mast cells by substance P and mastoparan. J Pharmacol Exp Ther 1989;250:329–35. Mousli M, Bueb JW, Bronner C, Rouot B, Landry Y. G-protein activation: a receptor-independent mode of action for cationic am-

2078

S.P. Ribeiro et al. / Peptides 25 (2004) 2069–2078

phiphilic neuropeptides and venom peptides. Trends Pharmacol Sci 1990;11:358–62. [24] Nagashima K, Masamichi K, Saito K, Yasuhara T, Tsukamoto T, Mori M, et al. The role of lysine at 7th position of wasp chemotactic peptides. Biochem Biophys Res Commun 1990;168:844–9. [25] Paton WDM. Compound 48/80: a potent histamine releaser. Br J Pharmacol 1951;6:499–508.

[26] Souza BM, Marques MR, Tomazela DM, Eberlin MN, Mendes MA, Palma MS. Mass spectrometric characterization of two novel inflammatory peptides from the venom of the social wasp Polybia Paulista. Rapid Commun Mass Spectrom 2004;18(10):1095–102. [27] Vorm O, Roepstorff P. Peptide sequence information derived by partial acid hydrolysis and matrix-assisted laser desorption/ionization mass spectrometry. Biol Mass Spectrom 1994;23:734–40.

Structural and functional characterization of N ...

Oct 2, 2004 - a solution containing 0.877g NaCl, 0.028g KCl, 0.043g. NaH2PO4, 0.048g KH2PO4, 0.10g glucose, 0.10g BSA,. 90 μL CaCl2 (2mol/L) solution ...

197KB Sizes 2 Downloads 186 Views

Recommend Documents

Structural and functional characterization of two novel ...
E-mail address: [email protected] (M.S. Palma). the wasps; mellitin ... their sequences were determined by automated Edman dedra- dation chemistry.

Synthesis, structural characterization and cytotoxic activity of ... - Arkivoc
Aug 7, 2017 - N = 0.0 ppm) as an external standard. The chemical shifts are reported in ppm (δ); multiplicities are indicated by s (singlet), d (doublet), t (triplet), q (quartet), m (multiplet) and br (broad). Coupling ...... the dose-dependent inh

Functional, Structural and Non-Structural Preparedness ...
May 11, 2016 - In many devel- oping countries, health facilities and the education and ... included technical equipment and evaluated adminis- trative aspects ...

Structural and biological characterization of three ... - Semantic Scholar
Nov 11, 2004 - European hornet, Vespa crabro, and its analogs. J. Peptide Res. 50, 88–93. Manning, D.R., Gilman, A.G., 1983. The regulatory components of adenylate cyclase and transducin. A family of structurally homologous guanine nucleotide-bindi

Structural characterization of novel chemotactic and ...
cal Investigations (Millennium Institute-MCT/CNPq), Rio. Claro/SP 13506-900 ..... the side chain of the C-terminal amino acid in the fragment ion. When two ...

Synthesis and structural characterization of a stable betaine ... - Arkivoc
more than one moiety of a stable radical are called polyradicals, and they .... following typical settings: number of scans 1, centre field 3350 G, sweep field ..... 20. http://www.niehs.nih.gov/research/resources/software/tox-pharm/tools/index.cfm.

Functional and clinical characterization of KCNJ2 ...
A reduction in Kir2.1 prolonged the terminal phase of the cardiac action potential, ..... cal examination. ... taneous action potentials, DADs, and terminal-phase.

Structural and Functional Studies of Aspergillus oryzae ...
Oct 7, 2009 - reactivity, stability, and structure of Aspergillus oryzae cutinase and compare it to the ..... PocketFinder function.28 For all above analysis 1CEX pdb file was ..... of the active site in A. oryzae cutinase can also explain its abilit

Isolation and partial characterization of an N-acetylgalactosamine ...
agarose was obtained fromSigma Chemical Co. (St. Louis, MO, U.S.A.). Extraction .... 3M-mercaptoethanesulphonic acid. Cysteinewas determined as cysteic ...

An X-ray nanodiffraction technique for structural characterization of ...
Author(s) of this paper may load this reprint on their own web site provided that this cover page is retained. ... of the third-generation synchrotron radiation source and advanced high- ... discusses the application of the technique to studies of ti

Density-functional theory characterization of acid sites ...
deprotonation energy is not correlated to the O–H bond length or vibrational stretch frequency. .... be only 0.2 kJ/mol, so our use of fixed lattice parameters is ..... measured by the rate of hexane cracking, increases linearly .... the best [12].

Structural characterization of the large soluble ...
the most highly conserved domain within the members of the dynamin family. .... 27 °C; average Rh ¼ 22.37 nm; (B) 100 lM GED in 0.1 M phos- phate buffer with 1% ... free and mobile while the rest were buried in the inter- ior of the oligomers.

Characterization of the N-Acetyl-5-neuraminic Acid ...
2008 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A.. JANUARY 11 ...... mutant R127K gaveO-LOS express- ..... Hood, D. W., Cox, A. D., Gilbert, M., Makepeace, K., Walsh, S., Deadman,. M. E., Cody ...

Characterization of the N-Acetyl-5-neuraminic Acid ...
SiaP crystals were flash-frozen in mother liquor containing. 10% (v/v) glycerol. ..... Electron density map of the Neu5Ac-binding site of SiaP. The figure shows the ...

Synthesis and spectroscopic characterization of double ... - Arkivoc
Dec 4, 2016 - with the elaboration at positions 2, 3 or 6, depending on the application ..... CHaHbO), 4.32 (dd, J 5.9, 11.7 Hz, 1H, CHaHbO), 4.80 (d, J2.0 Hz, ...

Characterization and Parameterized Generation of ...
The development of new architectures for Field-Programmable Gate Arrays ..... is analogous to the conclusions of Shew 63] who studied the application of.

SYNTHESIS AND CHARACTERIZATION OF ...
1 Faculty of Chemical Technology, Hanoi University of Technology. 2 Institute of .... their different degrees of ionization depending on pH values. Actually, the ...

Characterization of the Psychological, Physiological and ... - CiteSeerX
Aug 31, 2011 - inhibitors [8], acetylcholine esterase inhibitors [9] and metabolites ...... Data was stored on a dedicated windows XP laptop PC for post.

SYNTHESIS, CHARACTERIZATION AND ANTIBACTERIAL ...
SYNTHESIS, CHARACTERIZATION AND ANTIBACTE ... T C-4 OF 7-HYDROXY-4- METHYL COUMARIN.pdf. SYNTHESIS, CHARACTERIZATION AND ...