Skip to main content

Structural determinants of the hyperalgesic activity of myotoxic Lys49-phospholipase A2

Abstract

Background

Bothropstoxin-I (BthTx-I) is a Lys49-phospholipase A2 (Lys49-PLA2) from the venom of Bothrops jararacussu, which despite of the lack of catalytic activity induces myotoxicity, inflammation and pain. The C-terminal region of the Lys49-PLA2s is important for these effects; however, the amino acid residues that determine hyperalgesia and edema are unknown. The aim of this study was to characterize the structural determinants for the Lys49-PLA2-induced nociception and inflammation.

Methods

Scanning alanine mutagenesis in the active-site and C-terminal regions of BthTx-I has been used to study the structural determinants of toxin activities. The R118A mutant was employed as this substitution decreases PLA2 myotoxicity. In addition, K115A and K116A mutants – which contribute to decrease cytotoxicity – and the K122A mutant – which decreases both myotoxicity and cytotoxicity – were also used. The H48Q mutant – which does not interfere with membrane damage or myotoxic activity – was used to evaluate if the PLA2 catalytic site is relevant for the non-catalytic PLA2-induced pain and inflammation. Wistar male rats received intraplantar injections with mutant PLA2. Subsequently, hyperalgesia and edema were evaluated by the paw pressure test and by a plethysmometer. Native and recombinant BthTx-I were used as controls.

Results

Native and recombinant BthTx-I induced hyperalgesia and edema, which peaked at 2 h. The R118A mutant did not induce nociception or edema. The mutations K115A and K116A abolished hyperalgesia without interfering with edema. Finally, the K122A mutant did not induce hyperalgesia and presented a decreased inflammatory response.

Conclusions

The results obtained with the BthTx-I mutants suggest, for the first time, that there are distinct residues responsible for the hyperalgesia and edema induced by BthTx-I. In addition, we also showed that cytolytic activity is essential for the hyperalgesic effect but not for edematogenic activity, corroborating previous data showing that edema and hyperalgesia can occur in a non-dependent manner. Understanding the structure-activity relationship in BthTx-I has opened new possibilities to discover the target for PLA2-induced pain.

Background

Phospholipases A2 (PLA2; EC 3.1.1.4) are enzymes that hydrolyze the sn-2 acyl bond of glycerophospholipids, releasing free fatty acids and lysophospholipids [1]. Secretory PLA2s are found in a wide variety of biological fluids such as inflammatory exudates, and the venoms of arthropods, mollusks and snakes [2]. These enzymes are abundant in Bothrops snake venoms and display pharmacological activities characterized by myotoxic, neurotoxic, anticoagulant, hypotensive, hemolytic, platelet aggregation inhibition, bactericidal, pro-inflammatory and nociceptive effects [24]. A subfamily of class IIA PLA2s has been purified from the venoms of several viperid snakes, in which the Asp49 residue is replaced by Lys [5, 6]. These Ly49-PLA2s conserve the basic structural fold of this family of enzymes but lack catalytic activity.

While the Lys49-PLA2s do not show catalytic activity, in vitro studies showed they are able to disrupt liposome membranes and release their contents by a Ca2+-independent mechanism that does not involve hydrolysis of membrane phospholipids [7]. Despite the lack of catalytic activity, the in vivo activities of the Lys49-PLA2s include myonecrosis, bactericidal activity, local inflammation and pain [6, 813]. Chacur et al. [11] have demonstrated that the C-terminal cationic/hydrophobic sequence corresponding to amino acids 115–129 of a Lys49-PLA2 isolated from Bothrops asper is critical for the sensation of pain. This finding is supported by the demonstration that heparin partially neutralizes hyperalgesia induced by this toxin, and the direct induction of hyperalgesia by the peptide corresponding to amino acids 115–129, although having lower activity than the native toxin. Despite this evidence, the amino acids responsible for this effect are unknown.

Scanning alanine mutagenesis is a useful strategy to study the structural determinants of the activities of Lys49-PLA2. In this regard, Chioato et al. [14] have demonstrated that amino acid residues in C-terminal region of a Lys49-PLA2 from the venom of Bothrops jararacussu (BthTx-I) determine its biological activity. It has been demonstrated that the Lys122Ala mutant does not display myotoxic activity while Arg115Ala and Arg116Ala mutants do not display membrane-damaging activities. Moreover, His48Gln substitution, which eliminates any possible catalytic activity, does not influence the biological or membrane damaging proprieties of BthTx-I. Using these well-characterized functional point mutants in the active-site and C-terminal regions of the BthTx-I, we aimed to characterize the structural determinants for the Lys49-PLA2-induced nociception and inflammation, and more specifically, the edematogenic response.

Methods

Protein purification from crude venom

Bothropstoxin-I (BthTx-I) was purified from crude lyophilized Bothrops jararacussu venom using a single step cation-exchange chromatography as previously described [15]. The BthTx-I was eluted as a single peak and then dialyzed against 5 mM Tris–HCl, pH 7.5, for 36 h with buffer changes every 12 h and concentrated 10-fold by lyophilization. Protein purity was evaluated by silver staining of SDS-PAGE gels [16].

Site directed mutagenesis

A full-length cDNA encoding BthTx-I has been previously isolated from B. jararacussu venom gland cDNA by RT-PCR (GenBank Acc. No. X78599) [17], and subcloned into the expression vector pET3-d [18]. The nucleotide sequencing has confirmed the construct in which Ser1 of the BthTx-I is preceded by a Met, and a stop codon immediately follows Cys133. After linearization of this construct with ScaI, site directed mutagenesis of the BthTx-I was performed by PCR mutagenesis [19] to introduce single mutations: Lys115 → Ala (K115A), Lys116 → Ala (K116A), Arg118 → Ala (R118A), Lys122 → Ala (K122A) and His48 → Gln (H48Q). The final PCR reactions were performed using oligonucleotides complementary to the vector sequences flanking the BthTx-I insert which contained restriction sites for XbaI (5′-extremity) and BamHI (3′-extremity). After digestion with these enzymes, the amplified fragments were subcloned into the equivalent sites in the expression vector pET3d and fully sequenced.

Recombinant protein expression and purification

A 150-mL volume of growth medium (2.5 g yeast extract; 10 mM MgSO4; 15 μg/L chloramphenicol; 150 μg/L ampicillin; pH 7.5) was inoculated with Escherichia coli strain BL21(DE3)pLysS transformed with the native or mutant constructs in pET3d, and grown at 37 °C to an A600 of 0.6. Recombinant protein expression was induced by addition of 0.6 mM isopropylthiogalactoside, and the culture was grown for an additional period of 5 h. Inclusion bodies were isolated from bacterial pellets by repeated rounds of sonication in 20 mL of lysis buffer (50 mM Tris–HCl, pH 8.0; 1 mM EDTA; 0.4 M urea; 1% Triton X-100) followed by centrifugation at 12,000 g. The protocol for the solubilization and refolding of recombinant BthTx-I in the presence of a gel filtration medium was performed as previously described [18]. The refolded protein was applied directly to the cation exchange column and eluted as described earlier for the purification of the native BthTx-I from crude venom.

Animals

Male Wistar rats, weighing between 170 and 190g were used. Rats were housed in a temperature-controlled (21 ± 2 °C) and light-controlled (12/12 h light/dark cycle) room with standard rodent rations and water ad libitum. All procedures were conducted in accordance with the guidelines of the International Association for the Study of Pain [20] and were approved by the Institutional Animal Care Committee of the Butantan Institute (CEUAIB, protocol number 118/2002).

Pharmacological treatments

For the evaluation of hyperalgesia and allodynia, animals were injected with either 0.1 mL of sterile phosphate-buffered saline (PBS) solution (control animals) or 0.1 mL PBS containing the appropriate concentration of native, recombinant or mutant BthTx-I into the subplantar surface of one hind paw. For evaluation of edema, whereas toxins were injected into a hind paw and the PBS was administered in the contralateral paw.

Evaluation of mechanical hyperalgesia (Randall and Selitto test)

An Ugo-Basile pressure apparatus [21] was used to assess pressure pain thresholds prior to, and again at different times after, intraplantar injection of native, recombinant, mutant BthTx-I or vehicle into the right hind paw. The contralateral paw was not injected. Testing was blind with respect to group designation. Briefly, a force (in g) with increasing magnitude was applied to the paw. The force needed to induce paw withdrawal was recorded as the pain threshold. To reduce stress, the rats were exposed to the testing procedure the day before the experiment, as previously described [22].

Evaluation of low threshold mechanical allodynia (von Frey)

The von Frey test [23] was used to assess low-threshold mechanical pain thresholds prior to intraplantar injection of the toxins or PBS (control) at different periods of time later on. This test was performed as previously described in detail, using the modified up-down method [24]. Briefly, a logarithmic series of ten calibrated Semmes-Weinstein monofilaments (von Frey hairs, Stoelting, Wood Dale, USA) was applied to the right hind paw to determine the stimulus intensity threshold stiffness required to elicit a paw withdrawal response.

Log stiffness of the hairs is determined by log10 (milligrams × 10) and ranged from 3.61 (407 mg) to 5.18 (15.136 mg). Basal line assessment was initiated with the 2.041 mg hair. In the event of a paw withdrawal, the same hair was again presented 30–60 s later. If the response was again elicited, the 407 mg monofilament was presented. In the absence of a paw withdrawal response to the 407 mg stimulus, the next stronger monofilament was presented (692 mg). The monofilament that elicited a clear response was recorded, and was presented once again 30–60 s later. If the animal withdrew its paw on two consecutive trials with the same stiffness value, no further von Frey hairs were tested.

However, in the absence of a response to the initial 2.041 mg monofilament, presentation of monofilaments continued in ascending order until two consecutive responses were elicited from the same monofilament. All single responses were recorded, but assessment was complete only after two consecutive responses were elicited from the same monofilament. In instances when rats failed to respond, the strongest stimulus (15.136 mg) was considered to be the cut-off value.

Responses that occurred to the weakest stimulus (407 mg) were assigned the lower cut-off value for that time point. To reduce stress, rats were habituated to the experimental environment on each of the four days before experiments. Behavioral responses were used to calculate the 50% paw withdrawal threshold (absolute threshold) by fitting a Gaussian integral psychometric function using a maximum-likelihood fitting method. This fitting method allows parametric analyses [24, 25].

Evaluation of edema

The volume increase (edema) of paws up to the tibiotarsal articulation was plethysmographically measured before toxin or PBS (control) injection and subsequently at chosen time intervals according to the method of Van Arman et al. [26]. The percentage of increase in paw volume was determined for each paw. The difference between the values obtained for both paws was used as a measure of edema.

Statistical analysis

The results are presented as the mean ± SEM. The statistical evaluation of data was conducted using a two-way analysis of variance (ANOVA) with post-hoc testing by Tukey. A value of p < 0.05 was considered significant.

Results

Characterization of hyperalgesia and edema induced by native and recombinant bothropstoxin-I

Intraplantar injection of 2.5 μg of native BthTx-I did not alter the sensitivity to pressure pain, as measured by the Randall and Sellito test. In contrast, doses of 5, 10 and 20 μg/paw decreased the pain threshold (28%, 43% and 42%, respectively) of the animals as compared to the basal values. The peak of mechanical hyperalgesia was detected at 2 h. Intraplantar PBS injection (vehicle control) did not modify the pain threshold of the animals (Fig. 1a). Native BthTx-I also induced a significant edematogenic response when injected at 10 and 20 μg/paw. The dose of 10 μg/paw caused maximal response at 2 h (44%), while a dose of 20 μg/paw showed a peak effect 1 h (47%) after toxin administration, decreasing thereafter and completely disappearing within 24 h (Fig. 1b).

Fig. 1
figure 1

Effect of native and recombinant bothropstoxin I injection on the pain threshold and paw volume. a Effect of the native-BthTx intraplantar injection on the pain threshold and b on paw volume, at different doses. c Effect of recombinant BthTx-I intraplantar injection on the pain threshold and d on paw volume, at different doses. e Comparison between native and recombinant BthTx-I hyperalgesic and f edematogenic effects. The pain threshold of the animals was determined in rat hind paw before at different times after the intraplantar injection of PBS (control group) or toxins. Sensitivity to pain was measured as the threshold response to pressure and expressed as g. The edema was determined by an increase in the volume in the right hind paw of rats compared with the control contralateral paw. The paw volume was obtained pletsmografically. Each point represents the mean ± SEM of six animals. *Significantly different from mean values before venom injection and #different from BthTx-I-recombinant (BthTx-I rec) (p < 0.05)

In order to evaluate whether the recombinant form of BthTx-I induces hyperalgesia and edema, a dose response curve was performed for recombinant BthTx-I. As is the case for the native toxin, the intraplantar injection of 2.5 μg/paw of recombinant BthTx-I did not induce significant alteration in the pain threshold of the animals. Doses of 5, 10 and 20 μg/paw significantly decreased the pain threshold as compared with baseline, 2h after toxin injection (27%, 42% and 41%, to 5, 10 and 20 μg/paw, respectively). Injection of PBS (control) did not modify the pain threshold of the animals (Fig. 1c). Moreover, the doses of 5, 10 and 20 μg/paw caused a significant edematogenic response. The peak of the edematogenic response was detected 2 h after injection of 5 μg/paw (34%) of BthTx-I, or 1 h after the administration of 10 (46%) or 20 μg/paw (50%) of the toxin. Intraplantar injection of PBS (vehicle control) did not alter the paw volume of the animals (Fig. 1d).

For comparative analysis, native and recombinant BthTx-I at 10 μg/paw induced similar intensity of hyperalgesia (Fig. 1e). Although the recombinant BthTx-I did not induce edema at the same magnitude as native at 1 h, both toxins induced similar edema 2 h after the treatment (Fig. 1f). Therefore, the dose of 10 μg/paw was selected for subsequent studies.

Effect of BthTx-I site-directed mutagenesis on rat pain threshold and paw volume

In order to investigate whether the residues involved in the determination of myotoxic activities were also critical for hyperalgesia and edema, the BthTx-I mutant R118A was tested. BthTx-I-induced hyperalgesia was blocked by R118A mutation (in which the myotoxic activity is reduced) (Fig. 2a). In addition, the R118A mutation significantly decreased the edema induced by recombinant BthTx-I (Fig. 2b).

Fig. 2
figure 2

Effect of BthTx-I R118A (Arg118 → Ala) mutagenesis on rat pain threshold and paw volume. Decrease in a threshold response and b paw volume were determined in rat hind paw before and 2, 4, 6 and 24 h after the intraplantar injection of PBS (control group) or R118A, or recombinant BthTx-I. The paw volume was obtained pletsmografically. Sensitivity to pain was measured as the threshold response to pressure and expressed as g. The paw volume was obtained pletsmografically. Each point represents the mean ± SEM of six animals. *Significantly different from mean values before venom injection and #different from BthTx-I-recombinant (BthTx-I rec) (p < 0.05)

In order to investigate whether the residues involved in the BthTx-I membrane damaging activities were also critical for hyperalgesia and edema, the K115A and K116A mutants were tested. BthTx-I induced hyperalgesia was completely abolished by both K115A and K116A (which reduce the membrane-damaging activity) (Fig 3a).

Fig. 3
figure 3

Effect of BthTx-I K115A and K116A (Lys115 → Ala and Lys116 → Ala) mutagenesis on rat pain threshold and paw volume. Decrease in a threshold response and b paw volume were determined in rat hind paw before and 2, 4, 6 and 24 h after the intraplantar injection of PBS (control group) or R115A, or R116A, or recombinant BthTx-I. Sensitivity to pain was measured as the threshold response to pressure and expressed as g. The paw volume was obtained pletsmografically. Each point represents the mean ± SEM of six animals. *Significantly different from mean values before venom injection (p < 0.05)

The K122A mutation, which significantly reduces both myotoxic and membrane damaging activities also reduced the rat hind paw hyperalgesia induced by BthTx-I (Fig. 4a). The K122A mutation also significantly decreased the edema induced by recombinant BthTx-I (Fig. 4b).

Fig. 4
figure 4

Effect of BthTx-I K122A (Lys122 → Ala) mutagenesis on rat pain threshold and paw volume. Decrease in a threshold response and b paw volume were determined in rat hind paw before and 2, 4, 6 and 24 h after the intraplantar injection of PBS (control group) or R122A, or recombinant BthTx-I. Sensitivity to pain was measured as the threshold response to pressure and expressed as g. The paw volume was obtained pletsmografically. Each point represents the mean ± SEM of six animals. *Significantly different from mean values before venom injection (p < 0.05)

The H48Q mutation eliminates catalytic activity in class II PLA2s, and even though no catalytic activity is detected in BthTx-I, this mutant was used as a control to eliminate the possibility that the observed effects are the result of phospholipid hydrolysis. The H48Q mutation did not modify the hyperalgesia or edematogenic response induced by BthTx-I (Fig. 5a and Fig 5b).

Fig. 5
figure 5

Effect of H48Q (BthTx-I His48 → Gln) mutagenesis on rat pain threshold and paw volume. Decrease in a threshold response and b paw volume were determined in rat hind paw before and 2, 4, 6 and 24 h after the intraplantar injection of PBS (control group) or H48Q, or recombinant BthTx-I. Sensitivity to pain was measured as the threshold response to pressure and expressed as g. The paw volume was obtained pletsmografically. Each point represents the mean ± SEM of six animals. *Significantly different from mean values before venom injection (p < 0.05)

Characterization of allodynia induced by native and recombinant bothropstoxin-I

Intraplantar injection (10 μg/paw) of either the native or recombinant lowered withdrawal thresholds, as measured by the von Frey test. This effect was observed 2 h after native Bthtx-I (65%) or recombinant BthTx-I (58%) injection, and completely disappeared within 24 h. The injection of PBS (control group) did not modify the pain threshold of the animals (Fig. 6a).

Fig 6
figure 6

Characterization of allodynia induced by native and recombinant BthTx-I (a) and mutants BthTx-I (b). Decrease in tactile threshold was evaluated by von Frey test, before and 2, 4, 6 and 24 h after the intraplantar injection of PBS (control group), native, recombinant BthTx-I (a) or BthTx-I mutants (b). Sensitivity to pain was measured as the threshold response to tactile and expressed as g Log (mg × 10). Each point represents the mean ± SEM of six animals. *Significantly different from mean values before venom injection (p < 0.05)

Effect of BthTx-I site-directed mutagenesis on allodynia

The intraplantar administration of H48Q (10 μg/paw) lowered withdrawal thresholds, comparing with the baseline. The allodynic effect was observed 2 h after toxin injection (58%). The other BthTx-I mutants (Arg118 → Ala, Arg115 → Ala, Arg116 → Ala, Arg122 → Ala) did not alter the threshold (Fig. 6b).

Discussion

Site-directed mutagenesis studies can identify the structural determinants for biological activities of venom PLA2s. In this study, we demonstrate for the first time the involvement of amino acids in the C-terminal region in the nociceptive activity of the BthTx-I, a non-catalytic Lys49-PLA2 from Bothrops jararacussu venom. In addition, we have demonstrated that the same residues that are determinants of myotoxicity of the BthTx-I are also involved in hyperalgesia and inflammation, whereas the residues responsible for the cytolytic activity only contribute to the nociceptive effect of the protein.

In the experimental procedures, the recombinant BthTx-I was used as control for all the behavior experiments performed. Circular dichroism spectroscopic analysis has previously confirmed that the protein secondary structures were preserved in the recombinant molecule, as well as its biological activities (myotoxicity and cytolytic effects). Here we have shown that the recombinant BthTx-I displays hyperalgesic and edematogenic responses with similar onset, intensity and time course to those observed for the native BthTx-I. The edematogenic activity of the native BthTx-1 has been previously demonstrated [27, 28]. However, to the best of our knowledge, this is the first report showing that BthTx-I induces hyperalgesia in an experimental model of pain evaluation.

It is well established that the PLA2 component contributes to the local effects induced by the Bothrops snake envenomation. Our group has previously demonstrated that both the Lys49-PLA2 that is devoid of catalytic activity, and the catalytically active Asp49-PLA2 from Bothrops asper venom cause significant local hyperalgesia in rat paws after intraplantar injection. The hyperalgesic effect induced by these PLA2s is mediated by biogenic amines, bradykinin, cytokines, prostaglandins and sympathomimetic amines that may interact and be sequentially released [11].

Because Lys49-PLA2s are unable to catalyze phospholipid hydrolysis, their toxicity has been explained by some mechanisms that differ from that of their catalytically active PLA2 counterparts. It has been suggested that the C-terminal region of Lys49-PLA2s from Bothrops venoms is critical for their biological activities [14, 2933]. The nociceptive effect of Lys49-PLA2 has also been investigated, and we have previously shown that the C-terminal region of the Lys49-PLA2 from Bothrops asper venom is important for hyperalgesia, since the intraplantar injection of a peptide corresponding to amino acids 115–119 in the C-terminal region of the protein induces hyperalgesia in rats. In contrast, a C-terminal peptide derived from the same region of the Asp49-PLA2 did not show any nociceptive effect [11].

It has been demonstrated that the C-terminal region of the Lys-PLA2 region is also responsible for cytolytic, edematogenic, and myotoxic activities of this PLA2 [31, 34, 35]. Furthermore, this C-terminal region is also endowed with bactericidal activity, and a peptide corresponding to residues 115–119 of the BthTx-I reproduces the antimicrobial effect of the role Lys49-PLA2 [36, 37]. Taken together these data indicate that the C-terminal region may have an important role in the biological effects of venom-derived Lys49-PLA2. However, crystallographic and site-directed mutagenesis studies have suggested that additional residues, other than those located at C-terminal, participate in Lys49-PLA2 toxicity, and Lys20 is also critical for the myotoxic activity of this molecule [29, 32].

Despite the importance of the C-terminal region for the different biological activities of the Lys49-PLA2s, the membrane damaging, myotoxic and bactericidal activities have distinct structural determinants. This suggestion is based on scanning alanine mutagenesis studies showing that the structural determinants of the bactericidal activity are more extensive and only partially overlap with the structural determinants of the myotoxic and cytolytic activities [14, 29]. However, the results obtained in the present study indicate a degree of similarity in the structural determinants involved in the myotoxic, cytolytic, hyperalgesic and edematogenic effects. This observation is based on data showing that:

  • The same residue responsible for the myotoxic activity (R118) also contributes to the edematogenic and hyperalgesic responses induced by BthTx-I [14].

  • Residues 115 and 116, which are determinants for the Ca2+-independent membrane damaging activity of the BthTx-I, are also critical for the hyperalgesic effect of this Lys-PLA2, but not for the edematogenic response [12, 14, 29].

  • The R122A, which contributes to both the myotoxic and cytolytic activities of BthTX-I, is also important for hyperalgesia and edema.

The results obtained in this study indicate that the BthTx-I-induced hyperalgesia depends on the main biological activities of this Lys-PLA2, since elimination of the myotoxic and cytolytic activities also abolished hyperalgesia. In contrast, the edematogenic response is less dependent on the cytolytic effects of BthTx-I, since the elimination of myotoxicity interfered with this activity. These data also suggest that hyperalgesia and edema caused by this Lys-PLA2s are not directly correlated.

Previous data from our group have demonstrated that different mediators are involved in the genesis of hyperalgesia and edema caused by Lys49 and Asp49-PLA2s from Bothrops snake venoms, reinforcing the suggestion that hyperalgesia and edema induced by BthTx-I are not directly dependent [11].

The mechanisms that contribute to myotoxicity and to hyperalgesia and edema of the Lys49-PLA2s are not yet characterized. Preliminary data have demonstrated that the R118 mutant, which is devoid of myotoxic activity, induces decreased edematogenic activity, and a substantial drop in the number of local neutrophils (Zambelli and Cury, 2004, personal communication), confirming the importance of myotoxicity to the inflammatory response caused by this molecule. Despite the evidences indicating a possible correlation between myotoxicity and inflammation, we should also consider that a reduction in myotoxicity abrogates hyperalgesia by a mechanism independent of inflammation. The Lys49-PLA2 from B. asper induces a ATP and K+ release from C2C12 myotubes in culture and from mouse muscles [38]. It has been demonstrated that these mediators can directly induce pain by activating purinergic receptors or inducing membrane depolarization of peripheral sensory nerves [3840]. Therefore, a direct link between myotoxicity and pain generation may exist.

The data showing that the residues 115 and 116 are critical for hyperalgesia, but not for edema formation, also adds evidence that nociception and edema are not directly correlated, and may have distinct structural determinants. Although there is no data available to explain how these residues contribute exclusively for hyperalgesia, these mutations are involved in the BthTx-I cytolytic activity and a direct effect of cytotoxicity on pain generation is a possibility. Further experiments are necessary to investigate this hypothesis.

To further characterize the structural determinants involved in BthTx-I-induced hyperalgesia, and to evaluate whether a putative residual catalytic activity of the Lys49-PLA2 could play a role in the hyperalgesia and the edema induced by BthTx-I, we tested the H48Q mutant in our experimental conditions. The hydrolytic mechanism of Asp49-PLA2s involves the His48 in the catalytic site that activates a conserved water molecule, thereby initiating the nucleophilic attack on the sn-2 position of the phospholipid substrate and the H48Q mutation abolishes this activity [41, 42]. Here we observe that H48Q mutation does not interfere with the hyperalgesic and edematogenic effects induced by the BthTx-I, confirming that only the C-terminal amino acids are relevant for the BthTx-I pharmacological activities. It is important to highlight that the H48Q mutation does not alter the myotoxicity or the membrane damaging effects of the BthTx-I [14].

The data presented here suggest that residues present in the C-terminal region of the BthTx-I are important for hyperalgesia and edema. Previous studies have demonstrated that the synthetic peptide 115–129 was unable to mimic some effects of the whole toxin, such as in vitro and in vivo myotoxicity, and toxicity against epithelial cells and erythrocytes [30, 37]. These findings indicate that other residues present in the whole molecule or the quaternary structure of Lys49-PLA2s are critical for the for the biological activity of these molecules [32, 43]. However, the absence of toxicity observed for the synthetic peptide 115–129 do not exclude the participation of C-terminal residues in the pharmacological effects evaluated.

Conclusions

In conclusion, the results obtained with the BthTx-I mutants suggest, for the first time, that there are distinct residues responsible for the hyperalgesia and edema induced by BthTx-I. In addition, we also showed that the cytolytic activity is essential for the hyperalgesic effect but not for edematogenic activity, reinforcing previous data showing that edema and hyperalgesia can occur independently. A better understanding about the structure-activity relationship can open new avenues of investigation to identify the target for PLA2-induced pain.

Abbreviations

ANOVA:

analysis of variance

BthTx-I:

bothropstoxin-I

cDNA:

complementary deoxyribonucleic acid

EDTA:

ethylenediaminetetraacetic acid

H48Q:

His48 → Gln

K115A:

Lys115 → Ala

K116A:

Lys116 → Ala

K122A:

Lys122 → Ala

MgSO4 :

magnesium sulfate

PLA2 :

phospholipase A2

PLA2-Lys49:

phospholipase A2 with a Lys at position 49

R118A:

Arg118 → Ala

RT-PCR:

reverse transcriptase polymerase chain reaction

Tris–HCl:

Tris hydrochloride

References

  1. Dennis EA. Diversity of group types, regulation, and function of phospholipase A2. J Biol Chem. 1994;269(18):13057–60.

    CAS  PubMed  Google Scholar 

  2. Valentin E, Lambeau G. Increasing molecular diversity of secreted phospholipases A2 and their receptors and binding proteins. Biochim Biophys Acta. 2000;1488(1–2):59–70.

    Article  CAS  PubMed  Google Scholar 

  3. Harris LK, Franson RC. [1-14C] oleate-labeled autoclaved yeast: a membranous substrate for measuring phospholipase A2 activity in vitro. Anal Biochem. 1991;193(2):191–6.

    Article  CAS  PubMed  Google Scholar 

  4. Dunn RD, Broady KW. Snake inhibitors of phospholipase A2 enzymes. Biochim Biophys Acta. 2001;1533(1):29–37.

    Article  CAS  PubMed  Google Scholar 

  5. Arni RK, Ward RJ. Phospholipase A2--a structural review. Toxicon. 1996;34(8):827–41.

    Article  CAS  PubMed  Google Scholar 

  6. Gutiérrez JM, Lomonte B. Phospholipase A2 myotoxins from Bothrops snake venoms. Toxicon. 1995;33(11):1405–24.

    Article  PubMed  Google Scholar 

  7. de Oliveira AH, Giglio JR, Andrião-Escarso SH, Ito AS, Ward RJ. A pH-induced dissociation of the dimeric form of a lysine 49-phospholipase A2 abolishes Ca2 + −independent membrane damaging activity. Biochemistry. 2001;40(23):6912–20.

    Article  PubMed  Google Scholar 

  8. Ownby CL. Selistre de Araujo HS, White SP, Fletcher JE. Lysine 49 phospholipase A2 proteins. Toxicon. 1999;37(3):411–45.

    Article  CAS  PubMed  Google Scholar 

  9. Soares AM, Guerra-Sá R, Borja-Oliveira CR, Rodrigues VM, Rodrigues-Simioni L, Rodrigues V, et al. Structural and functional characterization of BnSP-7, a Lys49 myotoxic phospholipase A2 homologue from Bothrops neuwiedi pauloensis venom. Arch Biochem Biophys. 2000;378(2):201–9.

    Article  CAS  PubMed  Google Scholar 

  10. Páramo L, Lomonte B, Pizarro-Cerdá J, Bengoechea JA, Gorvel JP, Moreno E. Bactericidal activity of Lys49 and Asp49 myotoxic phospholipases A2 from Bothrops asper snake venom--synthetic Lys49 myotoxin II-(115–129)- peptide identifies its bactericidal region. Eur J Biochem. 1998;253(2):452–61.

    Article  PubMed  Google Scholar 

  11. Chacur M, Longo I, Picolo G, Gutiérrez JM, Lomonte B, Guerra JL, et al. Hyperalgesia induced by Asp49 and Lys49 phospholipases A2 from Bothrops asper snake venom: pharmacological mediation and molecular determinants. Toxicon. 2003;41(6):667–78.

    Article  CAS  PubMed  Google Scholar 

  12. Mamede CCN, de Sousa BB, Pereira DFC, Matias MS, de Queiroz MR, de Morais NCG, et al. Comparative analysis of local effects caused by Bothrops alternatus and Bothrops moojeni snake venoms: enzymatic contributions and inflammatory modulations. Toxicon. 2016;117:37–45.

    Article  CAS  PubMed  Google Scholar 

  13. Teixeira CF, Landucci EC, Antunes E, Chacur M, Cury Y. Inflammatory effects of snake venom myotoxic phospholipases A2. Toxicon. 2003;42(8):947–62.

    Article  CAS  PubMed  Google Scholar 

  14. Chioato L, De Oliveira AH, Ruller R, Sá JM, Ward RJ. Distinct sites for myotoxic and membrane-damaging activities in the C-terminal region of a Lys49-phospholipase A2. Biochem J. 2002;366(Pt 3):971–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Ruller R, Aragão EA, Chioato L, Ferreira TL, de Oliveira AH, Sà JM, et al. A predominant role for hydrogen bonding in the stability of the homodimer of bothropstoxin-I, a lysine 49-phospholipase A2. Biochimie. 2005;87(11):993–1003.

    Article  CAS  PubMed  Google Scholar 

  16. Laemmli UK. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature. 1970;227(5259):680–5.

    Article  CAS  PubMed  Google Scholar 

  17. Ward RJ, Monesi N, Arni RK, Larson RE, Paço-Larson ML. Sequence of a cDNA encoding bothropstoxin I, a myotoxin from the venom of Bothrops jararacussu. Gene. 1995;156(2):305–6.

    Article  CAS  PubMed  Google Scholar 

  18. Ward RJ, de Oliveira AH, Bortoleto RK, Rosa JC, Faça VM, Greene LJ. Refolding and purification of Bothropstoxin-I, a Lys49-phospholipase A2 homologue, expressed as inclusion bodies in Escherichia coli. Protein Expr Purif. 2001;21(1):134–40.

    Article  CAS  PubMed  Google Scholar 

  19. Nelson RM, Long GL. A general method of site-specific mutagenesis using a modification of the Thermus aquaticus polymerase chain reaction. Anal Biochem. 1989;180(1):147–51.

    Article  CAS  PubMed  Google Scholar 

  20. Zimmermann M. Ethical guidelines for investigations of experimental pain in conscious animals. Pain. 1983;16(2):109–10.

    Article  CAS  PubMed  Google Scholar 

  21. Randallo LO, Selitto JJ. A method for measurement of analgesic activity on inflamed tissue. Arch Int Pharmacodyn Ther. 1957;111(4):409–19.

    Google Scholar 

  22. Zambelli VO, Gross ER, Chen CH, Gutierrez VP, Cury Y, Mochly-Rosen D. Aldehyde dehydrogenase-2 regulates nociception in rodent models of acute inflammatory pain. Sci Transl Med. 2014;6(251):251ra118.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Chaplan SR, Bach FW, Pogrel JW, Chung JM, Yaksh TL. Quantitative assessment of tactile allodynia in the rat paw. J Neurosci Methods. 1994;53(1):55–63.

    Article  CAS  PubMed  Google Scholar 

  24. Milligan ED, Mehmert KK, Hinde JL, Harvey LO, Martin D, Tracey KJ, et al. Thermal hyperalgesia and mechanical allodynia produced by intrathecal administration of the human immunodeficiency virus-1 (HIV-1) envelope glycoprotein, gp120. Brain Res. 2000;861(1):105–16.

    Article  CAS  PubMed  Google Scholar 

  25. Milligan ED, Twining C, Chacur M, Biedenkapp J, O’Connor K, Poole S, et al. Spinal glia and proinflammatory cytokines mediate mirror-image neuropathic pain in rats. J Neurosci. 2003;23(3):1026–40.

    CAS  PubMed  Google Scholar 

  26. Van Arman CG, Nuss GW. Plasma bradykininogen levels in adjuvant arthritis and carrageenan inflammation. J Pathol. 1969;99(3):245–50.

    Article  PubMed  Google Scholar 

  27. Melaré R, Floriano RS, Gracia M, Rodrigues-Simioni L, da Cruz-Höfling MA, Rocha T. Ultrastructural aspects of mouse nerve-muscle preparation exposed to Bothrops jararacussu and Bothrops bilineatus venoms and their toxins BthTX-I and Bbil-TX: Unknown myotoxic effects. Microsc Res Tech. 2016;79(11):1082–9.

    Article  PubMed  Google Scholar 

  28. Landucci EC, Castro RC, Pereira MF, Cintra AC, Giglio JR, Marangoni S, et al. Mast cell degranulation induced by two phospholipase A2 homologues: dissociation between enzymatic and biological activities. Eur J Pharmacol. 1998;343(2–3):257–63.

    Article  CAS  PubMed  Google Scholar 

  29. Chioato L, Aragão EA, Lopes Ferreira T, Ivo de Medeiros A. Mapping of the structural determinants of artificial and biological membrane damaging activities of a Lys49 phospholipase A2 by scanning alanine mutagenesis. Biochim Biophys Acta. 2007;176(5):1247–57.

    Article  Google Scholar 

  30. Lomonte B, Angulo Y, Santamaría C. Comparative study of synthetic peptides corresponding to region 115–129 in Lys49 myotoxic phospholipases A2 from snake venoms. Toxicon. 2003;42(3):307–12.

    Article  CAS  PubMed  Google Scholar 

  31. Núñez CE, Angulo Y, Lomonte B. Identification of the myotoxic site of the Lys49 phospholipase A(2) from Agkistrodon piscivorus piscivorus snake venom: synthetic C-terminal peptides from Lys49, but not from Asp49 myotoxins, exert membrane-damaging activities. Toxicon. 2001;39(10):1587–94.

    Article  PubMed  Google Scholar 

  32. dos Santos JI, Soares AM, Fontes MRM. Comparative structural studies on Lys49-phospholipases A(2) from Bothrops genus reveal their myotoxic site. J Struct Biol. 2009;167(2):106–16.

    Article  PubMed  Google Scholar 

  33. Lomonte B, Angulo Y, Moreno E. Synthetic peptides derived from the C-terminal region of Lys49 phospholipase A2 homologues from viperidae snake venoms: biomimetic activities and potential applications. Curr Pharm Des. 2010;16(28):3224–30.

    Article  CAS  PubMed  Google Scholar 

  34. Lomonte B, Moreno E, Tarkowski A, Hanson LA, Maccarana M. Neutralizing interaction between heparins and myotoxin II, a lysine 49 phospholipase A2 from Bothrops asper snake venom. Identification of a heparin-binding and cytolytic toxin region by the use of synthetic peptides and molecular modeling. J Biol Chem. 1994;269(47):29867–73.

    CAS  PubMed  Google Scholar 

  35. Calderón L, Lomonte B. Immunochemical characterization and role in toxic activities of region 115–129 of myotoxin II, a Lys49 phospholipase A2 from Bothrops asper snake venom. Arch Biochem Biophys. 1998;358:343–50.

    Article  PubMed  Google Scholar 

  36. Aragão EA, Chioato L, Ward RJ. Permeabilization of E. coli K12 inner and outer membranes by bothropstoxin-I, A LYS49 phospholipase A2 from Bothrops jararacussu. Toxicon. 2008;51(4):538–46.

    Article  PubMed  Google Scholar 

  37. Santos-Filho NA, Lorenzon EN, Ramos MA, Santos CT, Piccoli JP, Bauab TM, et al. Synthesis and characterization of an antibacterial and non-toxic dimeric peptide derived from the C-terminal region of Bothropstoxin-I. Toxicon. 2015;103:160–8.

    Article  CAS  PubMed  Google Scholar 

  38. Cintra-Francischinelli M, Caccin P, Chiavegato A, Pizzo P, Carmignoto G, Angulo Y, et al. Bothrops snake myotoxins induce a large efflux of ATP and potassium with spreading of cell damage and pain. Proc Natl Acad Sci. 2010;107(32):14140–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Burnstock G. Purines and sensory nerves. Handb Exp Pharmacol. 2009;194:333–92.

    Article  CAS  Google Scholar 

  40. Lim RKS. Neuropharmacology of pain and analgesia. Pharmacology of Pain: Proceedings of the First International Pharmacological edited by Lim RKS, Armstrong DEG. PardoPharmacol Pain 9. 2013. p. 169–217.

  41. Verheij HM, Volwerk JJ, Jansen EH, Puyk WC, Dijkstra BW, Drenth J, et al. Methylation of histidine-48 in pancreatic phospholipase A2. Role of histidine and calcium ion in the catalytic mechanism. Biochemistry. 1980;19(4):743–50.

    Article  CAS  PubMed  Google Scholar 

  42. Scott DL, Otwinowski Z, Gelb MH, Sigler PB. Crystal structure of bee-venom phospholipase A2 in a complex with a transition-state analogue. Science. 1990;250(4987):1563–6.

    Article  CAS  PubMed  Google Scholar 

  43. Lomonte B, Rangel J. Snake venom Lys49 myotoxins: From phospholipases A(2) to non-enzymatic membrane disruptors. Toxicon. 2012;60(4):520–30.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

Thanks are to the Center for the Study of Venoms and Venomous Animals (CEVAP) of UNESP for enabling the publication of this paper (Edital Toxinologia CAPES no. 063/2010, Process no. 230.38.006285/2011-21, AUXPE Toxinologia 1219/2011).

Funding

This work was supported by a grant from the State of São Paulo Research Foundation (FAPESP, grant no. 2002/12906-0) and by Butantan Foundation, São Paulo, Brazil.

Authors’ contributions

YC, RJW and VOZ conceived the project. VOZ, LC, VPG and YC designed and performed the experiments and carried out data analysis. VOZ wrote the manuscript with input from RJW. All authors read and approved the final manuscript.

Competing interests

The authors declare that they have no competing interests.

Consent for publication

Not applicable.

Ethics approval

All animal tests were conducted in accordance with the guidelines of the International Association for the Study of Pain [20]. The present study was approved by the Institutional Animal Care Committee of the Butantan Institute (CEUAIB, protocol number 118/2002).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yara Cury.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zambelli, V.O., Chioato, L., Gutierrez, V.P. et al. Structural determinants of the hyperalgesic activity of myotoxic Lys49-phospholipase A2 . J Venom Anim Toxins Incl Trop Dis 23, 7 (2017). https://doi.org/10.1186/s40409-017-0099-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40409-017-0099-6

Keywords