University of Southern Denmark

GiTx1(β/κ-theraphotoxin-Gi1a), a novel toxin from the venom of Brazilian Grammostola iheringi (, Theraphosidae) Isolation, structural assessments and activity on voltage-gated ion channels Montandon, Gabriela Gontijo; Cassoli, Juliana Silva; Peigneur, Steve; Braga, Thiago Verano; Moreira Dos Santos, Daniel; Bittencourt Paiva, Ana Luiza; Ricardo de Moraes, Éder; Kushmerick, Christopher; Borges, Márcia Helena; Richardson, Michael; Monteiro de Castro Pimenta, Adriano; Kjeldsen, Frank; Vasconcelos Diniz, Marcelo Ribeiro; Tytgat, Jan; Elena de Lima, Maria Published in: Biochimie

DOI: 10.1016/j.biochi.2020.07.008

Publication date: 2020

Document version: Accepted manuscript

Document license: CC BY-NC-ND

Citation for pulished version (APA): Montandon, G. G., Cassoli, J. S., Peigneur, S., Braga, T. V., Moreira Dos Santos, D., Bittencourt Paiva, A. L., Ricardo de Moraes, É., Kushmerick, C., Borges, M. H., Richardson, M., Monteiro de Castro Pimenta, A., Kjeldsen, F., Vasconcelos Diniz, M. R., Tytgat, J., & Elena de Lima, M. (2020). GiTx1(β/κ-theraphotoxin-Gi1a), a novel toxin from the venom of Brazilian tarantula Grammostola iheringi (Mygalomorphae, Theraphosidae): Isolation, structural assessments and activity on voltage-gated ion channels. Biochimie, 176, 138-149. https://doi.org/10.1016/j.biochi.2020.07.008

Go to publication entry in University of Southern Denmark's Research Portal

Terms of use This work is brought to you by the University of Southern Denmark. Unless otherwise specified it has been shared according to the terms for self-archiving. If no other license is stated, these terms apply:

• You may download this work for personal use only. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying this open access version If you believe that this document breaches copyright please contact us providing details and we will investigate your claim. Please direct all enquiries to [email protected]

Download date: 05. Oct. 2021 1 GiTx1(β/κ-theraphotoxin-Gi1a), a novel toxin from the venom of Brazilian 2 tarantula Grammostola iheringi (Mygalomorphae, Theraphosidae): 3 isolation, structural assessments and activity on voltage-gated ion channels 4 5 Gabriela Gontijo Montandona; Juliana Silva Cassolia,b; Steve Peigneurb; Thiago Verano 6 Bragac,d; Daniel Moreira dos Santosa; Ana Luiza Bittencourt Paivaf; Éder Ricardo de 7 Moraesc; Christopher Kuschmerickc; Márcia Helena Borgesf; Michael Richardsonf; Adriano 8 Monteiro de Castro Pimentaa; Frank Kjeldsend; Marcelo Ribeiro Vasconcelos Dinizf; Jan 9 Tytgatb; Maria Elena de Limaa,h* 10 11 Affiliations 12 13 aDepartamento de Bioquímica e Imunologia, Universidade Federal de Minas Gerais - UFMG, 14 Av. Antônio Carlos 6627, 31270-901 Belo Horizonte, MG, Brazil. 15 16 bPharmacology and Toxicology, University of Leuven (KULeuven) - Campus Gasthuisberg, 17 PO Box 922, Herestraat 49, 3000 Leuven, Belgium. 18 19 cDepartamento de Fisiologia e Biofísica, Universidade Federal de Minas Gerais - UFMG, Av. 20 Antônio Carlos 6627, 31270-901 Belo Horizonte, MG, Brazil. 21 22 dProtein Research Group, Department of Biochemistry and Molecular Biology, University of 23 Southern Denmark, Campusvej 55, 5230 Odense M, Denmark. 24 25 fDiretoria de Pesquisa e Desenvolvimento, Fundação Ezequiel Dias, Rua Conde Pereira 26 Carneiro 80, 30510-010 Belo Horizonte, MG, Brazil. 27 28 hSanta Casa de Belo Horizonte: Ensino e Pesquisa, Rua Domingos Vieira, 590, 30.150-240 29 Belo Horizonte, MG, Brazil. 30 31 32 * Corresponding author: Maria Elena de Lima - Santa Casa de Belo Horizonte: Ensino e 33 Pesquisa, Rua Domingos Vieira, 590, 30.150-240 Belo Horizonte, MG, Brazil. 34 E-mail: [email protected]; [email protected] (M.E. de Lima). 35

36 Keywords 37 Grammostola, tarantula venom, GiTx1, β/κ-TRTX-Gi1a, toxin 38

39

40

41

42

43

44

1

45 Abstract

46 Spider venoms, despite of their toxicity, represent rich sources of active compounds with

47 biotechnological potential. However, in view of the diversity of the spider , the full

48 potential of their venom molecules is still far from being known. In this work, we report the

49 purification, the structural and functional characterization of GiTx1 (β/κ-TRTX-Gi1a), the first

50 toxin purified from the venom of the Brazilian tarantula spider Grammostola iheringi. This

51 toxin was isolated by chromatographic procedures, completely sequenced through automated

52 Edman’s degradation and tandem mass spectrometry and its spatial structure was predicted by

53 molecular modeling. It has a MW of 3.585 Da and the following sequence of amino acids:

54 SCQKWMWTCDQRPCCEDMVCKLWCKIIK. The pharmacological activities of GiTx1

55 were characterized by electrophysiology using two different approaches: (i) whole-cell patch

56 clamp on dorsal root ganglia neurons (DRG) and (ii) two-electrode voltage-clamp on voltage-

57 gated sodium and potassium channels subtypes expressed in Xenopus laevis oocytes. GiTx1,

58 at 2 μM, caused a partial block of inward ( 40 %) and outward ( 20 %) currents in DRG

59 cells, blocked rNav1.2, rNav1.4 and mNav1.6 and had a noticeable effect on VdNav, an

60 sodium channel isoform, among others. IC50 values of 156.39 ± 14.90 nM for Nav1.6 and

61 124.05 ± 12.99 nM for VdNav, were obtained. In addition, this toxin was active on rKv4.3 and

62 hERG isoforms. However, it was not effective on Shaker IR and rKv2.1 potassium channel

63 subtypes. These multiple effects on ion channels characterize GiTx1 as a promiscuous toxin.

64

65

66

67 68

69

70

71

2

72 1. Introduction

73 Spider venoms are complex mixtures of hundreds of compounds that act on different

74 targets when introduced into other organisms due to predatory or defense senses [1]. The

75 chemical nature of these molecules is very diverse, including salts, nucleotides, free amino

76 acids, glucose, neurotransmitters, acylpolyamines, and mainly peptide toxins and proteins

77 [2,3]. Because of the diverse mixture of components, spider venoms venom represent an

78 important source of novel biologically active compounds potentially useful as tools for

79 pharmacological and agricultural researches – e.g. development of drugs and insecticides (for

80 review see: [4,5]).

81 from the Theraphosidae family (infraorder Mygalomorphae), also called

82 and “caranguejeiras”, are a sizable group of large, hairy and diverse spiders species,

83 distribute nearly all over globe. Example are species belonging the genus Theraphosa,

84 Lasiodora and Grammostola, the last mainly found in the Southern region of Brazil, Argentina,

85 Chile, Paraguay Bolivia and Uruguay.

86 Tarantula toxins comprise a group of peptides with 31 - 41 amino acids residues and 3

87 - 4 disulfide bridges, displaying DDH or ICK structural motifs [3,6,7]. Several studies have

88 demonstrated the wide molecular repertoire of toxins from tarantula venoms [3,8–12]. To date,

89 information on 450 toxins from the Theraphosidae family have been deposited in the

90 Arachnoserver database [13], however, few of these toxins have been functionally

91 characterized.

92 The spider Grammostola iheringi is one of best-known spider species from the

93 Theraphosidae family in Brazil. With a broad characterization approach based on

94 Multidimensional Protein Identification Technology (MudPIT) and bioinformatic analysis

95 using PepExplorer tool, our group has identified more than three hundred proteins in the venom

96 of Grammostola iheringi [14]. Through alignments searches in databases, we were able to

97 identify about 30% of these sequences, which include proteins (metalloproteinases, cysteine

3

98 proteinases, aspartic proteinases, carboxypeptidases and cysteine-rich secretory enzymes –

99 CRISP) and toxins. We have found in the G. iheringi venomous extract 63 toxins with 20-86

100 amino acid residues (MW 2-10 kDa) and 6-12 cysteines. Through sequence homology, it was

101 possible to infer functional activity for some of these toxins, including analgesic,

102 antiarrhythmic and insecticidal activity, likely through modulation of ion channels since

103 tarantula’s toxins have been described mainly as ion channel modulators. However, several

104 neurotoxins are yet to be characterized, such as the 24 neurotoxins in the venom of G. iheringi

105 identified by Borges et al, 2016, with remaining unknown targets.

106 In this work, a new short ICK toxin from G. iheringi venom, GiTx1 (β/κ-TRTX-Gi1a),

107 was purified and structurally characterized. Furthermore, its actions on voltage-gated sodium

108 and potassium channels were characterized using whole-cell patch clamp and two electrode

109 voltage-clamp techniques. GiTx1 is the first toxin purified from the venom of G. iheringi and

110 its structural and pharmacological characterization provide a relevant contribution on the

111 venom study of this spider.

112 2. Material and Methods

113 2.1. collection and Venom Extraction

114 Specimens of Grammostola iheringi (Sisgen #A9BCC40) were collected at the border

115 of Santa Catarina and Rio Grande do Sul states (Coordinates: 27°31'31"S, 51°47'7" W), in the

116 south region of Brazil. The species were identified by Dr. Sylvia Lucas from the

117 Laboratory of Butantan Institute (Brazil). The venom was obtained from eighteen adult

118 individuals by electrical stimulation (5V) of the chelicerae. Soluble venom was pooled,

119 lyophilized and stored at -20 oC until use.

120 2.2. Purification procedures

121 G. iheringi crude venom was fractionated by reverse phase chromatography on HPLC

122 system, according Richardson et al., 2006 [15]. A preparative column Vydac C4 (22 x 250 mm)

123 (Grace – Deerfield, USA) previously equilibrated with 0.1 % aqueous trifluoroacetic acid

4

124 (TFA) (Eluent A), was used in this procedure. For each run, 30 mg of the lyophilized venom

125 were dissolved in 2 mL of Eluent A. Samples were centrifuged and supernatants loaded into

126 column. Components were eluted with 0.1% trifluoroacetic acid in acetonitrile (Eluent B) using

127 the following gradient system: 0 to 20 min, 100% A; 20 to 30 min, 0-20% B; 30 to 110 min;

128 20-40% B; 110 to 130 min, 40-50% B; 130 to 150 min, 50-70% B, being Eluent B a solution

129 of 0.1% trifluoroacetic acid in acetonitrile (ACN). Flow rate was 5 mL/min and elution was

130 monitored by absorbance at 214 nm. Eluted fractions were manually collected and lyophilized.

131 The fraction of interest was purified by another reverse phase chromatography on HPLC, using

132 an analytical column Sephasil Peptide C18 5µ ST (4.6 x 250 mm) (GE Healthcare - Uppsala,

133 Sweden) equilibrated with Eluent A. The peptide was eluted by a linear gradient of 20 – 50 %

134 of eluent B during 40 min. Flow rate was 1 mL/min and chromatographic runs monitored by

135 absorbance at 214 nm. Samples were collected and lyophilized for structural and functional

136 experiments.

137 2.3. Peptide sequencing by Edman’s degradation

138 The purified peptide was previously S-reduced and alkylated with vinyl pyridine as

139 described in Yuen et al, 1988 [16]. The reduced and alkylated sample was desalted by RPC-

140 HPLC using a Vydac C4 column (22 x 250 mm) (Grace - Deerfield, USA) and one half was

141 digested in solution with trypsin. Digested peptides were purified by RPC-HPLC using a Vydac

142 C18 column (4.6 x 250 mm, Small Pore). Both reduced/alkylated and digested peptides were

143 sequenced by Edman degradation using the automated PPSQ-21A protein sequencer

144 (Shimadzu - Kyoto, Japan) coupled to reverse phase separation of PTH-amino acids on a

145 WAKOSIL-PTH (4.6 x 250 mm) column (Wako, Osaka, Japan), according to the

146 manufacturer’s instructions.

147 2.4. Mass Spectrometry (MS) Analyses

148 2.4.1. Sample preparation and sample desalting.

5

149 Eighty micrograms of the purified peptide were resuspended in 40 µL of 20 mM

150 triethylammonium bicarbonate (Sigma). Disulfide-bridged cysteine residues were reduced

151 with 10 mM TCEP (Pierce) for 20 min at 57 oC, followed by 40 min at room temperature. Thiol

152 groups were blocked with 30 mM iodoacetamide (Sigma) for 45 min at room temperature and

153 protected from light. Acetylation of the primary amines was achieved by incubating at room

154 temperature the peptide with acetic anhydride/dichloromethane (1:1) for 1 h.

155 The native sample and its chemically derived forms (i.e., reduced, alkylated and

156 acetylated) were desalted prior mass spectrometry analyses using self-made microcolumns

157 packed with Poros R2 resin (PerSeptive Biosystems) with a small plug of C18 extraction disk

158 (3M Bioanalytical Technologies) to close the tip. Samples were acidified with TFA (0.1% final

159 concentration) and loaded onto microcolumns previously equilibrated with 0.1% TFA and

160 washed 3 times with this solution. Peptides were eluted from the column with 70 % ACN/ 0.1

161 % TFA, lyophilized and stored at -20 ºC until required.

162 2.4.2. MALDI-TOF MS

163 MALDI-TOF mass spectrometry analyses were performed using an UltrafleXtreme

164 mass spectrometer (Bruker Daltonics), in positive ion mode and controlled by the software

165 FlexControl 3.4 (Bruker Daltonics). Samples were mixed with α-ciano-4-hydroxycynamic acid

166 (1:1, v/v) directly on the MALDI target and dried at room temperature. Analyses were

167 performed using reflected mode with external calibration using a peptide calibration standard

168 (Bruker Daltonics). Results were visualized using the software FlexAnalysis 3.4 (Bruker

169 Daltonics).

170 2.4.3. Reverse-phase nanoLC-MS/MS

171 Samples were resuspended in 0.1% formic acid (FA) and loaded onto a reverse-phase

172 in-house packed Reprosil-Pur C18-AQ column (3 μm; Dr. Maisch GmbH) with 15 cm length

173 and 75 μm inner diameter using an Easy-LC nanoHPLC system (Thermo Fisher Scientific).

174 Peptides were eluted from the column with a chromatography gradient from 0−34% solvent B

6

175 (90% ACN, 0.1% FA) for 23 at a flow rate of 250 nL/min directly into the mass spectrometer.

176 The Orbitrap XL (Thermo Fisher Scientific) was operated in a data-dependent mode (DDA).

177 After a survey scan (3 µscan; mass range m/z 300 – 1800; max injection time = 500 ms / AGC

178 target = 1x106; 30,000 FWHM at 400 m/z), the 2 most intense ion species. Were selected for

179 ETD fragmentation (1 µscan; max injection time = 100 ms / AGC target = 2x105; 30,000

180 FWHM at 400 m/z).

181 2.5. Similarity searches

182 Similarity searches were performed using BLAST tool against the UniProtKB/Swiss-

183 Prot database [17] and Clustal Omega [18] was used to align similar sequences.

184 2.6. Molecular modeling

185 Modeller v9.22 software [19] was used to determine the structure by molecular

186 modeling. The three-dimensional homology models of toxin targets were constructed using the

187 NMR structures of the templates (GsmTx2 or κ-TRTX-Gr2a, PDB: 1LUP, PaTx1 or κ-TRTX-

188 Ps1a, PDB: 1V7F and ProTx-II or β/ω-TRTX-Tp2a, PDB: 2N9T) assuming alignment between

189 targets and template. The Python script was written, and 50 best models were generated. The

190 best predicted model with least DOPE (discrete optimized protein energy) score was selected

191 for further investigation. Models were evaluated by the lowest DOPE by PROCHECK [20]

192 statistics and subsequently checked by VERIFY- 3D [21], ProSA [22] and Ramachandran plot

193 [23]. Tertiary structure models were visualized using Discovery Studio Visualizer 17.2.

194 2.7. Toxicity assays

195 Toxicity assays on mammals were performed on male Swiss mice (Mus musculus) of

196 about 20 g. Samples with different concentrations of crude venom or toxin, diluted in saline

197 with 0.1% BSA were injected intraperitoneal (i.p) or intracerebral (i.c). Toxicity to insects

198 were evaluated by intrathoracic injections into adult house flies (Musca domestica), as

199 described (Figueiredo et al., 1995). Previously to injections, flies were cold anesthetized at 4

200 oC during 2-3 min. Injections (1 L/fly) were carried out with a syringe (Hamilton, Reno, USA)

7

201 coupled to an automatic volume dispenser and a capillary glass needle. Control groups (mouse

202 and fly) received saline solution containing 0.1% of bovine serum albumin (BSA). Observation

203 of the intoxication symptoms were carried out immediately after injection and 2 and 24 h later.

204 The median lethal dose 50 % (LD50) value was estimated by the probit method (Finney, 1964).

205 The experimental procedures involving mice were approved by the Animal Experimentation

206 Ethics Committee (CETEA) of the Federal University of Minas Gerais (protocol 43/2005).

207 2.8. Culture of rat dorsal root ganglia (DRG) neurons cells

208 Culture of DRG neurons was done as previously described in Moraes et al., 2011 [24].

209 Briefly, male adults Wistar rats were killed by decapitation. Ganglia from all spinal segments

210 were removed and placed in cold Ca2+ -free Ringer solution. Ganglia were sectioned with

211 iridectomy scissors and incubated at 37°C for 20 minutes in Ca2 + -free Ringer plus papain (1

212 mg/ml) activated by cystein (0.03 mg/ml), followed by 20 minutes in Ca2 + -free Ringer plus

213 collagenase (2.5 mg/ml), also at 37°C. The enzyme incubation was stopped with F-12 media

214 supplemented with 10% fetal bovine serum (FBS) and 100 U/ml penicillin/streptomycin. The

215 cells were released by gentle trituration through Pasteur pipettes with fire-polished tips. Cells

216 were platted on glass discs pre-coated with poly-D-lysine and laminin inside Petri dishes. After

217 2 hours in an incubator at 37°C to allow the cells to settle, the media was exchanged with

218 Leibovitz’s L-15 medium supplemented with 10% FBS, 5 mM glucose, 5 mM NaHEPES and

219 100 U/ml penicillin/ streptomycin. Cultures were stored in a normal atmosphere at room

220 temperature and used within 72 h.

221 2.9. Heterologous expression of voltage-gated ion channels

222 For the expression of the Voltage Gated Potassium Channels (Shaker IR, rKv2.1, rKv4.3

223 and hERG) and the Voltage Gated Sodium Channels (rNav1.2, rNav1.3, rNav1.4, hNav1.5,

224 mNav1.6, rNav1.8, DmNav1, BgNav1, VdNav1) in Xenopus laevis oocytes, the linearized

225 plasmids were transcribed using the T7 or SP6 mMESSAGEmMACHINE transcription kit

226 (Ambion, Life TechnologiesTM - Carlsbad, USA). The harvesting of stage V-VI oocytes from

8

227 anaesthetized female X. laevis frog was performed according to Liman et al. (1992). Oocytes

228 were injected with 50 nL of cRNA at a concentration of 1 ng/nL using a micro-injector

229 (Drummond Scientific, USA). The oocytes were incubated in a solution containing (in mM):

230 NaCl, 96; KCl, 2; CaCl2, 1.8; MgCl2, 2 and HEPES, 5 (pH 7.4), supplemented with 50 mg/L

231 gentamycin sulfate. Whole-cell currents from oocytes were recorded 2–5 days after injection.

232 2.10. Electrophysiological recordings

233 2.10.1. DRG neurons

234 Currents from DRG neurons were accessed using whole cell patch-clamp technique and

235 were recorded at room temperature using a patch clamp amplifier (Axopatch 200B - Axon

236 Instruments, USA), controlled by pClamp7 software (Axon Instruments, USA) (Hamill et al.,

237 1981; Moraes et al., 2011). Pipettes open-tip resistance were < 1.5 MΩ. Pipettes were covered

238 with dental wax to approximately 0.1 mm from tips to reduce their electric capacitance and

239 thus facilitate cancellation of capacitive currents. Pipette solution was constituted by (in mM):

240 KCl, 30, K-Gluconate, 110; NaCl, 4; HEPES, 10; EGTA, 10; MgCl2, 5; MgATP, 2; NaGTP,

241 0.3; pH 7.0. The composition of bath solution was (in mM): NaCl, 140; KCl, 2.5; HEPES,

242 10;MES, 10; CaCl2, 2; MgCl2, 1; pH 7.4. Cell capacitance and series resistance were properly

243 compensated. Leak subtraction was performed using a -P/4 protocol. Elicited currents were

244 filtered at 1 kHz and sampled at 0.5 kHz (for outward currents) or at 20 kHz (for inward

245 currents) using a four-pole low-pass Bessel filter. Cells were held at -80 mV. Inward currents

246 were elicited by 100 ms depolarization pulses to 0 mV preceded by a 50-ms pre-pulse to -120

247 mV. Outward currents were evoked by a 200 ms depolarization pulses to 0 mV also from a

248 holding potential of -80 mV.

249 2.10.3. Xenopus laevis oocytes

250 For electrophysiological measurements in oocytes, two-electrode voltage clamp

251 recordings were performed at room temperature (18 - 22 oC) using a Geneclamp 500 amplifier

252 (Axon Instruments - Sunnyvale, California, USA) controlled by a pClamp data acquisition

9

253 system (Axon Instruments - Sunnyvale, California, USA). Bath solution composition was

254 ND96 (in mM): NaCl, 96; KCl, 2; CaCl2, 1.8; MgCl2, 2 and HEPES, 5 (pH 7.4). Voltage and

255 current electrodes were filled with 3M KCl. Resistances of both electrodes were kept between

256 0.7 and 1.5 MΩ. Elicited currents were sampled at 1 kHz and filtered at 0.5 kHz (for potassium

257 currents) or sampled at 20 kHz and filtered at 2 kHz (for sodium currents) using a four-pole

258 low-pass Bessel filter. Leak subtraction was performed using a-P/4 protocol. Shaker currents

259 were evoked by 500 ms depolarizations to 0 mV followed by a 500 ms pulse to -50 mV, from

260 a holding potential of -90 mV. Current traces of hERG channels were elicited by applying a

261 +40 mV pre-pulse for 2 s followed by a step to -120 mV for 2 s. Kv2.1 and Kv4.3 currents were

262 elicited by 500 ms pulses to +20 mV from a holding potential of -90 mV. Sodium current traces

263 were evoked by 100 ms depolarization to 0 mV from a holding potential of -90 mV. To assess

264 the concentration dependency of the peptide induced inhibitory effects, a concentration-

265 response curve was constructed, in which the percentage of current inhibition was plotted as a

266 function of toxin concentration. Data were fitted with the Hill equation: y = 100/[1 +

h 267 (IC50/[toxin]) ], were y is the amplitude of the toxin-induced effect, IC50 is the toxin

268 concentration at half-maximal efficacy, [toxin] is the toxin concentration, and h is the Hill

269 coefficient.

270 3. Results and Discussion

271 3.1. Toxicity, fractionation and mass fingerprinting of G. iheringi venom and isolation of

272 GiTx1

273 Toxicity of the total venom was checked in mouse by i.p. (n=2), i.c. (n=2) routes and in flies

274 (n=10) by intrathoracic injections. The first effects caused in mice by the i.p. injection of the

275 venom (800, 1,600, 3,200, 6,400 and 12,800 ng/mouse), occurred on average 30 minutes after

276 its application. These effects included paralysis and limited mobility at the lowest doses (800

277 ng and 1600 ng/animal), complete paralysis and death occurred in the highest doses (3,200 ng;

278 6,400 ng and 12,800 ng/animal). For all doses, when non-lethal, the effects were reversible and

10

279 after 24 hours, the mice had no visible symptoms. Higher toxicity in mice was observed when

280 the venom was injected by i.c via (100, 200, 400, 800 and 1,600 ng/mouse) being lethal at

281 1,600 ng/mouse. In insects that receive 15, 31, 62, 125, 250, 500 and 1000 ng/fly, toxicity was

282 visible from doses of 31 ng/fly, being lethal from 500 ng/fly. The LD50 calculated was 205.45

283 ng/fly (or 10.3 μg/g of fly). As most of the values reported from the literature were obtained

284 in tests using crickets the comparison of our results with these data is not adequate. The LD50

285 was not calculated to mouse, due to limitation on animal number and also because our first

286 objective was the characterization of novel insecticidal toxins. However, neurotoxic effects on

287 mouse were observed in all i.c tested doses (100, 200, 400, 800, 1400 ng), such as rotating

288 movements, tail erection, disorientation, piloerection, difficulty in locomotion followed by

289 paralysis. Grunts, screw movements and cyanosis were also noted in higher doses (800 and

290 1,600 ng/mouse). Few studies show comparative toxicity data from the venoms of tarantulas.

291 Lethal doses described vary between 210-950 μg of dry venom per mouse (20g). Despite our

292 values are far below that, Bücherl et al. 1971 [25] demonstrated that the toxicity and effects of

293 the venom vary widely with the injected route. Thus, the strict comparison of these data remains

294 complicated by the divergence of doses and routes used in the scarce studies with these venoms.

295 Crude venom fractionation and toxin purification were performed in two

296 chromatographic steps. Initially, 30 mg from G. iheringi venom were applied to a reversed-

297 phase C4 column, resulting in 33 fractions (Fig 1A). Each obtained fraction was submitted to

298 mass spectrometry analyses. The mass spectra results were summarized in Fig. 1B, which

299 represents the correspondence between molecular masses vs % ACN elution of venom

300 components.

11

301

302 Figure 1: G. iheringi venom fractionation and venom molecular mass profile obtained by

303 MALDI-TOF MS analysis. (A) Preparative RPC profile of G. iheringi venom on HPLC

304 system. The column (Vydac C4 250/22 mm) was equilibrated with 0.1% aqueous

305 trifluoroacetic acid, and 33 fractions eluted by a gradient of 0.1% trifluoroacetic acid in

306 acetonitrile (5 mL/min flow rate) as described in item 2.2. The elution was monitored at 214

307 nm. (B) Molecular mass vs % ACN elution plots of G. iheringi venom.

12

308

309 The results showed that the peptide fractions of G. iheringi venom comprised molecular

310 masses ranging from m/z 3,585 Da to 43,202 Da. Most of the eluted components seem to depict

311 a linear correlation between molecular masses and hydrophobicity. As other tarantula venoms

312 [3], G. iheringi components also displayed such bimodal behavior.

313 Following MS analysis, insect toxicity assays were performed in all eluted fractions

314 from the first chromatographic step, in exception of the fraction 20 that seemed to be pure and

315 was selected for further characterization. Table 1 shows a summary of the results with the

316 fractions obtained. Fractions 12 to 21 displayed high toxicity to adult flies including prolonged

317 paralyses and inability to fly. LD50 ranged from 3 to 171 ng/fly, hereby confirming the presence

318 of insecticidal toxins in G. iheringi venomous extract, as reported by Borges et al., 2016 [14].

319 Table 1. Insecticidal fractions from Grammostola iheringi venom and their respective 320 molecular masses. 321 Fractions Molecular weight of [M+H]+ in Da 12 nd* 13 6683; 6859; 6194; 6033; 7063 14 3867; 5182 15 5822; 5198; 3867; 3899 16 3978; 3884; 3633 17 6855; 3980; 3897; 4044 18 3999; 4165; 7998; 7999; 8909 19 4165; 4000; 6855 20 3588 21 3581; 3681 nd (not determinated) 322 323 The assessment of molecular masses was performed using reverse phase chromatography (RPC) on 324 HPLC followed by matrix assisted laser desorption/ionization mass spectrometry in offline mode. 325 Attempts to reach a reasonable amount of insecticidal pure peptide were carried out,

326 and, due to the highest protein concentration among the insecticidal fractions and its high state

327 of purity, the fraction 20 was further investigated. This sample was submitted to analytical RPC

328 using a linear gradient of ACN and a pure peptide - toxin GiTx1- was eluted with 36% of ACN

329 (Fig 2A).

13

330 Its toxicity was evaluated in flies and in mouse. Toxicity in flies ranged from 100 to

331 400 ng (causing paralyses) and death when 900 ng or above. Comparing the toxicities of GiTx1

332 (LD50 = 20.9 µg/g and LD100 = 45 µg/g ) to that of HwTx-II (LD50 = 127 + 54 µg/g and LD100

333 = 200 µg/g) from the Chinese spider Selenocosmia huwena [26], it shows that GiTx1 is six

334 times more toxic and also more lethal to the insects, than the HwTx-II. The toxicity trials in

335 mouse showed different effects: within the testing doses (i.p = 5 µg; i.c = 0.5 µg and 1

336 µg/animal) the toxin showed very low toxicity – causing reversible paralyses at the first hour

337 when injected i.c, at dose of 1µg/animal.

338 3.2. Peptide sequencing and sequence analyses

339 The molecular mass of GiTx1 was determined by MALDI-TOF/MS as 3,585.2 Da

340 (upper panel of Figure 2B). Because of a mass increase of 6 Da following the reduction of the

341 disulfide bonds, it is possible to assume that the native toxin contains 3 disulfide-bridges

342 (Figure 2B middle panel), and 6 cysteine residues, confirmed by the mass increase of 342 Da

343 (6 x 57 Da) after alkylation of thiol groups (Figure 2B, lower panel).

14

344

345

15

346 Figure 2: Purification of GiTx1 and MALDI-TOF MS analysis. (A) Analytical RPC of

347 subfraction 20 obtained from the preparative RPC step. The column (Sephasil Peptide C18 5µ

348 ST - 4.6 x 250 mm) was equilibrated with 0.1% aqueous trifluoroacetic acid, and GiTx1 was

349 eluted by a linear gradient of 0.1 % trifluoroacetic acid in acetonitrile according to item 2.2.

350 The elution was monitored at 214 nm. (B) Spectra of the native toxin and its chemically derived

351 forms. Each spectrum contains an insert representing the isotope distribution of GiTx1. In the

352 upper panel the monoisotopic mass of the native toxin is indicated. Mass increase of 6 Da

353 following the reduction of the disulfide bonds is depicted in middle panel. The 3 dissulfide-

354 bridged cysteines (6 cysteine residues) were confirmed by the mass increase of 342 Da (6x57

355 Da) following alkylation of the thiol groups in the lower panel.

356

357 In addition, GiTx1 was submitted to Edman’s sequencing in order to assess its primary

358 amino acids sequence. The sequence obtained of the reduced, alkylated and digested sample

359 and of the native sample comprises 30 amino acid residues, including 6 cysteine residues, that

360 were previously detected by MALDI-MS. Sequencing by Edman's degradation had

361 confidentially identified 29 amino acid residues out of 30 cycles, exception to the last cycle.

362 Comparing the theoretical mass of the obtained sequence and the experimental mass resulted

363 in nominal 128 Da, suggesting that a residue of lysine, glutamine or even an amidated glutamic

364 acid, could be the candidates for last residue of GiTx1 sequence. In order to solve the toxin

365 entire amino acid sequence, the reduced and alkylated toxin was subjected to ETD LTQ-

366 Orbitrap MS/MS analysis. Its primary structure was de novo sequenced, assisted by Edman

367 degradation’s data (Fig. 3A). Acetylation of primary amines was also carried out to confirm

368 the number of lysine residues that could be present in toxin sequence. The observed mass

369 increase of 252 Da is consistent with six primary amines, five in lysine residues and the free

370 amine at the N-terminus (Fig. 3B). The amino acids sequence data reported for GiTx1 toxin

371 was deposited in UniProt database under the accession number C0HJJ7.

16

372

373 Figure 3: High-resolution ETD Orbitrap MS/MS analysis of GiTx1 and acetylation of its

374 primary amines. (A) The reduced and alkylated GiTx1was subjected to ETD fragmentation

375 and its primary structure was de novo sequenced manually and assisted with the Edman

376 degradation result. The figure represents a deconvoluted MS/MS spectrum of the ion 656.80

377 m/z (z = 6). (B) The reduced and alkylated toxin was acetylated in order to confirm the number

378 of lysine residues present in the toxin. The observed mass increase is consistent with 6 primary

379 amines, which mean that the toxin indeed contains 5 lysine residues and a free N-terminal.

380 Acetylation = 42 Da.

381

17

382 3.3. Sequence alignments and molecular modeling

383 Searches on protein databases revealed a high similarity index of GiTx1 to other

384 tarantula peptides. The multiple alignment with such peptides is shown in Fig. 4.

385

386 Figure 4: Sequence and alignment of GiTx1 with other tarantula’s toxins. Full sequence

387 and alignment of GiTx1 with other similar proteins from spiders found at UniProtKB database.

388 Cysteine residues are highlighted in red. Identical and similar residues are boxed in black and

389 grey, respectively. The percentages of similarity (Sim.) of each sequence with GiTx1 are

390 indicated next to each sequence. The peptides κ-TRTX-Ps1b and κ-TRTX-Ps1a are from

391 Paraphysa scrofa venom. β-TRTX-Gr1a, β-TRTX-Gr1b, κ-TRTX-Gr2a and κ-TRTX-Gr2b

392 are from Grammostola rosea venom. Finally, β/ω-TRTX-Tp2a is from pruriens

393 venom. Information about these sequences can be accessed in ArachnoServer [13].

394

395 GiTx1 displays cysteine residues framework similar to the aligned toxins, which

396 disulfide bridge pattern is CysI-CysIV, CysII-CysV and CysIII-CysIV. Following the remarkable

397 similarity concerning the cysteine positions, it is proposed that GiTx1 has a disulfide bridge

398 pattern similar to this group of spider toxins, adopting the ICK (Inhibitory Cystine Knot)

399 conformation [6], as a short loop ICK type toxin [3]. Apart from cysteine residues, it is possible

400 to notice also in the Fig.4, similarities and differences among amino acid residues that may

401 interfere in their pharmacological activities. The first residue of the GiTx1 sequence is differing

402 from the other peptides. It distinguishes itself from the others for being the only N-terminal

403 residue that is a serine and not a tyrosine. At position 11th, a glutamine is a distinct residue

18

404 facing the glutamic acid and serine residues. The same situation can be seen for the 14th

405 residue, the presence of a proline residue in GiTx1 contrasts the conserved lysine residues in

406 other known sequences. Other non-conservative amino acid residues from GiTx1 in

407 comparison to aligned sequences are also present at positions 27th and 29th. Despite the

408 observed differences, it is also valid to emphasize some similarities among the aligned

409 sequences. Overall, the toxin sequences from figure 5 displays a positive net charge (+1 to +4).

410 Other resemblance is in terms of positioning of hydrophobic residues. They are mainly located

411 between the first and the second cysteines residues, just before the fifth and the sixth cysteine

412 residues.

413 Among the aligned toxin sequences shown in fig.4, three of them had their solution

414 structures determined by two-dimensional NMR spectroscopy [27,28]. Solution structures of

415 κTrTx-Gr2a (GsmTx2), β/ω-TRTX-Tp2a (ProTx-II) and κTrTx-Ps1a (PaTx1), are available at

416 Protein Data Bank (PDB) under accession numbers 1LUP, 2N9T and 1V7F, respectively.

417 Using these three structures, it was achievable to predict the solution structure of GiTx1 by

418 molecular modeling tools (see item 2.6). The GiTx1 predicted structure is presented in Fig. 5.

19

419

420 Figure 5: Molecular model of GiTx1. (A) Stereo view of the peptide and (B) its rotated view

421 180 degrees to the left of the orientation in panel A (counterclockwise). The β-strands are in

422 light blue, loop regions in green and the disulfide bridges in yellow. Blue colors represent

423 positive charges, as red colors represent negative charges. (C) Electrostatic potential of the

424 molecular surface of GiTx1. (D) Electrostatic potential of GiTx1, rotated 180 degrees to the

425 left of the orientation in panel C (counterclockwise).

426

427 The molecular model for GiTx1 presents two distinct loops, and two short stranded

428 antiparallel β sheets. The model shows an overall charge distribution that reveals an

429 electrostatic anisotropy represented by the dipole moment shown in Fig. 5C and 5D, in which

430 electrostatic potential of the molecular surface of GiTx1 is shown. Charged residues are visibly

431 located in one side of the structure, while the opposite side is composed by hydrophobic amino

432 acids (TRP5, MET6, TRP7, THR8, LEU23 and TRP24). These characteristics are similar to

433 β/ω-TRTX-Tp2a [29,30]. Henriques et al., 2016 demonstrated the membrane binding of the

20

434 hydrophobic face of β/ω-TRTX-Tp2a and its importance for the interaction with the sodium

435 channel voltage sensor domain, proposing a correlation between membrane binding affinity

436 and its inhibitory potential. β/ω-TRTX-Tp2a insertion in the membrane favors its correct

437 orientation with the channel, facilitating the electrostatic interactions between the polar amino

438 acids of the toxin with the channel [30].

439 3.4. Electrophysiological recordings

440 Considering the structural similarity with other tarantula toxins and their biological

441 activities, electrophysiological experiments were conducted to determine if GiTx1 modulates

442 inward or outward ionic currents in DRG neurons. These experiments showed that the toxin,

443 at a concentration of 2M, GiTx1 reversibly inhibited 40% of inward current and 20% of

444 outward currents, suggesting its action upon voltage-gated sodium and potassium channels

445 (Fig. 6).

446

447

448 Figure 6: Normalized current plots from whole-cell patch clamp recordings. Effects of

449 GiTx1 (2M) upon sodium and potassium (inward and outward) currents from DRG cells

450 (n=8).

451 In addition, the action of GiTx1 was investigated upon a broad variety of sodium and

452 potassium channels subtypes expressed in X. laevis oocytes. The mammalian (rNav1.2,

453 rNav1.3, rNav1.4, hNav1.5, mNav1.6, hNav1.8), insect (DmNav1, BgNav1) and arachnid

21

454 (VdNav1) voltage gated sodium channel (VGSC) subtypes, as well as the voltage gated

455 potassium (VGPC) insect channel Shaker (Shaker IR) and mammalian channels Shab (rKv2.1),

456 Shal (rKv4.3) and erg1 (hERG) - were tested by two electrode voltage-clamp technique using

457 2µM of GiTx1. The electrophysiological recordings of the action of GiTx1 on these channels

458 are shown in Fig. 7A. The results show that at 2 µM, GiTx1 was able to inhibit the ionic currents

459 through all tested VGSCs isoforms (Fig. 7A). Among the VGPCs, the toxin showed complete

460 and partial inhibition on Kv4.3 and hERG channels respectively, but did not have any effect

461 on Shaker IR and Kv2.1 (Fig.7A). Fig. 7B summarizes GiTx1 effects in terms of percentage of

462 inhibition on the tested channels. The toxin effectiveness upon the inhibition of the ionic

463 currents through the sodium channels Nav1.6 and VdNav and the potassium channel hERG

464 was determined by dose-response curves, as shown in Fig. 7C. Measured IC50 values were

465 156.39 ± 14.90 nM for Nav1.6 and 124.05 ± 12.99 nM for VdNav, indicating that GiTx1a has

466 a similar affinity for these channels. In contrast, the IC50 value obtained for hERG channel

467 was 3695.17  127.22 nM. This concentration is 25 times higher than for the sodium channel

468 subtypes, suggesting that GiTx1 has a much higher affinity for sodium channels.

469

22

470 23

471 Figure 7. Effects of GiTx1 upon voltage-gated ion channels subtypes, expressed in 472 Xenopus laevis oocytes. (A) Screening of GiTx1 at 2µM. The tested mammalian isoforms were 473 raised from rat (r), mouse (m) and human (h). The invertrebrate subtypes were originated from 474 fruit-fly (Dm, Drosophyla melanogaster), cockroach (Bg, Blatella germanica) and mite (Vd, 475 Varroa destructor). Representative whole-cell current traces in the absence and in the presence 476 of 2 µM GiTx1 are shown for each channel. Dotted line indicates the zero-current level. The * 477 marks steady state current traces after application of GiTx1 (2µM) (n  3). (B) GiTx1 (2µM) 478 blocking effect on tested sodium and potassium channels isoforms (n  3). (C) Dose-response 479 curves of GiTx1 on mNav1.6, VdNav and hERG channels. Curves were obtained by plotting 480 the percentage of blocked current as a function of increasing toxin concentrations. All data are 481 presented as mean ± standard error (n ≥ 3). 482 483 Despite the differences in affinities presented for sodium and potassium voltage-gated

484 sodium channels, GiTx1 toxin can be considered a promiscuous toxin. The action of

485 promiscuous toxins is not specific for an ion channel family and they can act on ion channels

486 with different selectivity [3,31]. This promiscuous activity upon ion channels is largely found

487 in ICK toxins from tarantula venoms. For instance, Hainan toxins showed high affinity for Kv

488 channels and low affinity for Cav channels, whereas the opposite was observed for ω-TRTX-

489 Gr1a (GsTxSIA) [32]. Likewise, several works have been reporting the promiscuous activity

490 of tarantula’s toxins [33–36]. As consequence of the promiscuity displayed by these toxins, its

491 functional prediction becomes a difficult task. For example, κ-TRTX-Ps1a and κ-TRTX-Gr2a,

492 that were taken as templates for the prediction of the GiTx1 structure, are toxins that share 90%

493 of similarity in terms of sequence, however, their pharmacological actions have been reported

494 as modulators of Kv4.2 and Kv4.3 channels [37] or as blockers of mechano-sensitive channel

495 [27]. Unlike, μ-TRTX-Hh2a (HwTx4) and β-TRTX-Ps1a (PaurTx3) toxins share only 40% of

496 structural similarity; nevertheless they modulate neuronal sodium channels with equivalent

497 affinities [38]. Besides that, it was reported that the activity of some toxins upon specific ion

498 channels can be dose-dependent, which suggests that spider toxins have a number of putative

499 binding sites able to alter the function of a voltage-gated ion channel in different modes, at the

24

500 same time [34]. These findings emphasize the importance of the functional characterization of

501 spider toxins activity upon different ion channels and at different toxin concentrations. Thus,

502 additional studies may further illustrate other possible pharmacological actions of GiTx1.

503 Nevertheless, considering the ion channel activities reported for GiTx1 in this study and its

504 high homology to tarantula toxins with β action on sodium channels (Fig. 4), this toxin should

505 be renamed as β/κ-theraphotoxin-Gi1a according to the rational nomenclature proposed by

506 King et al, 2008 [39].

507 4. Final remarks

508 In this report, we have described the isolation, sequencing and the structural and

509 pharmacological characterization of GiTx1, the first peptide purified from the venom of the

510 spider G. iheringi. Additionally, the peptide mass diversity of G. iheringi venom was obtained

511 by mass spectrometry analysis.

512 A current database investigation revealed a similarity of GiTx1 with theraphosid toxins

513 bearing the short loop ICK toxin group. This was confirmed by its predicted 3D structure,

514 which showed that GiTx1 has two loops, two short stranded antiparallel β and an electrostatic

515 anisotropy, that are some features of short loop ICK toxin group. The electrophysiological

516 experiments performed on DRG neurons and X. laevis oocytes show that GiTx1 was effective

517 upon voltage-gated sodium and potassium channels. The action on invertebrate channels was

518 one of the identified targets of the toxin underlying the toxic effects observed in flies injected

519 with the venom as well as with this toxin. In summary, the present work provided a relevant

520 contribution for the description of the toxicity and the molecular diversity of the spider G.

521 iheringi venom and for the structural and functional characterization of a novel toxin belonging

522 to the short loop ICK family.

523

524 Acknowledgments

25

525 The authors are thankful to Dr. Sylvia Lucas (Laboratory of from Butantan Institute

526 – São Paulo, SP, Brazil) by careful identification of specimens Grammostola iheringi.

527 We are grateful D. J. Snyders for sharing rKv2.1 and rKv4.3. The Shaker IR clone was kindly

528 provided by G. Yellen. We thank M. Keating for sharing hERG. We are grateful to G. Mandel

529 (Stony Brook University, Stony Brook, NY) for providing the rNav1.4 clone, R.G. Kallen

530 (University of Pennsylvania, PA) for providing hNav1.5, A.L. Goldin (University of

531 California, Irvine, CA) for providing mNav1.6, J.N. Wood (University College, London, UK)

532 for providing rNav1.8, S.H. Heinemann (Friedrich-Schiller-Universita, Jena, Germany) for

533 sharing rβ1, and M.S. Williamson (Institute of Arable Crops Research Rothamsted, Harpenden,

534 UK) for providing DmNav1 and tipE clone. J.T. was funded by GOC2319N and GOA4919N

535 (FWO-Vlaanderen) and CELSA/17/047 (BOF - KU Leuven). KU Leuven funding

536 (PDM/19/164) supports S.P. We thank the following Brazilian Agencies for the financial

537 support: Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - CAPES, Fundação

538 de Amparo à Pesquisa do Estado de Minas Gerais - FAPEMIG and Conselho Nacional de

539 Desenvolvimento Científico e Tecnológico – CNPq.

540

541 References

542 [1] L. Kuhn-Nentwig, R. Stöcklin, W. Nentwig, Venom composition and strategies in

543 spiders. is everything possible?, 2011. https://doi.org/10.1016/B978-0-12-387668-

544 3.00001-5.

545 [2] E. Grishin, Polypeptide neurotoxins from spider venoms, Eur. J. Biochem. (1999).

546 https://doi.org/10.1046/j.1432-1327.1999.00622.x.

547 [3] P. Escoubas, L. Rash, Tarantulas: Eight-legged pharmacists and combinatorial

548 chemists, Toxicon. (2004). https://doi.org/10.1016/j.toxicon.2004.02.007.

549 [4] G.F. King, M.C. Hardy, Spider-Venom Peptides: Structure, Pharmacology, and

550 Potential for Control of Insect Pests, Annu. Rev. Entomol. (2013).

26

551 https://doi.org/10.1146/annurev-ento-120811-153650.

552 [5] N.J. Saez, V. Herzig, Versatile spider venom peptides and their medical and

553 agricultural applications, Toxicon. (2019).

554 https://doi.org/10.1016/j.toxicon.2018.11.298.

555 [6] R.S. Norton, P.K. Pallaghy, The cystine knot structure of ion channel toxins and

556 related polypeptides, in: Toxicon, 1998. https://doi.org/10.1016/S0041-

557 0101(98)00149-4.

558 [7] X.H. Wang, M. Connor, R. Smith, M.W. Maciejewski, M.E.H. Howden, G.M.

559 Nicholson, M.J. Christie, G.F. King, Discovery and characterization of a family of

560 insecticidal neurotoxins with a rare vicinal disulfide bridge, Nat. Struct. Biol. (2000).

561 https://doi.org/10.1038/75921.

562 [8] Z. Liao, J. Cao, S. Li, X. Yan, W. Hu, Q. He, J. Chen, J. Tang, J. Xie, S. Liang,

563 Proteomic and peptidomic analysis of the venom from Chinese tarantula Chilobrachys

564 jingzhao, Proteomics. (2007). https://doi.org/10.1002/pmic.200600785.

565 [9] J. Chen, L. Zhao, L. Jiang, E. Meng, Y. Zhang, X. Xiong, S. Liang, Transcriptome

566 analysis revealed novel possible venom components and cellular processes of the

567 tarantula Chilobrachys jingzhao venom gland, Toxicon. (2008).

568 https://doi.org/10.1016/j.toxicon.2008.08.003.

569 [10] E. Diego-García, S. Peigneur, E. Waelkens, S. Debaveye, J. Tytgat, Venom

570 components from Citharischius crawshayi spider (Family Theraphosidae): Exploring

571 transcriptome, venomics, and function, Cell. Mol. Life Sci. 67 (2010) 2799–2813.

572 https://doi.org/10.1007/s00018-010-0359-x.

573 [11] X. Tang, Y. Zhang, W. Hu, D. Xu, H. Tao, X. Yang, Y. Li, L. Jiang, S. Liang,

574 Molecular diversification of peptide toxins from the tarantula haplopelma hainanum

575 (Ornithoctonus hainana) venom based on transcriptomic, peptidomic, and genomic

576 analyses, J. Proteome Res. (2010). https://doi.org/10.1021/pr1000016.

27

577 [12] C.B.F. Mourão, F.N. Oliveira, A.C. E Carvalho, C.J. Arenas, H.M. Duque, J.C.

578 Gonçalves, J.K.A. Macêdo, P. Galante, C.A. Schwartz, M.R. Mortari, M. de F.M.

579 Almeida Santos, E.F. Schwartz, Venomic and pharmacological activity of

580 Acanthoscurria paulensis (Theraphosidae) spider venom, Toxicon. (2013).

581 https://doi.org/10.1016/j.toxicon.2012.11.008.

582 [13] V. Herzig, D.L.A. Wood, F. Newell, P.A. Chaumeil, Q. Kaas, G.J. Binford, G.M.

583 Nicholson, D. Gorse, G.F. King, ArachnoServer 2.0, an updated online resource for

584 spider toxin sequences and structures, Nucleic Acids Res. (2011).

585 https://doi.org/10.1093/nar/gkq1058.

586 [14] M.H. Borges, S.G. Figueiredo, F. V. Leprevost, M.E. De Lima, M. do N. Cordeiro,

587 M.R.V. Diniz, J. Moresco, P.C. Carvalho, J.R. Yates, Venomous extract protein profile

588 of Brazilian tarantula Grammostola iheringi: Searching for potential biotechnological

589 applications, J. Proteomics. (2016). https://doi.org/10.1016/j.jprot.2016.01.013.

590 [15] M. Richardson, A.M.C. Pimenta, M.P. Bemquerer, M.M. Santoro, P.S.L. Beirao, M.E.

591 Lima, S.G. Figueiredo, C. Bloch, E.A.R. Vasconcelos, F.A.P. Campos, P.C. Gomes,

592 M.N. Cordeiro, Comparison of the partial proteomes of the venoms of Brazilian

593 spiders of the genus Phoneutria, Comp. Biochem. Physiol. - C Toxicol. Pharmacol.

594 142 (2006) 173–187. https://doi.org/10.1016/j.cbpc.2005.09.010.

595 [16] S. Yuen, M.W. Hunkapiller, J. Wilson, P.A.U.M. Yuan, Sylvia yuen, michael w.

596 hunkapiller, 15 (1988) 5–15.

597 [17] A. Bateman et al. UniProt: The universal protein knowledgebase, Nucleic Acids Res.

598 (2017). https://doi.org/10.1093/nar/gkw1099.

599 [18] F. Sievers, A. Wilm, D. Dineen, T.J. Gibson, K. Karplus, W. Li, R. Lopez, H.

600 McWilliam, M. Remmert, J. Söding, J.D. Thompson, D.G. Higgins, Fast, scalable

601 generation of high-quality protein multiple sequence alignments using Clustal Omega,

602 Mol. Syst. Biol. (2011). https://doi.org/10.1038/msb.2011.75.

28

603 [19] A. Šali, T.L. Blundell, Comparative protein modelling by satisfaction of spatial

604 restraints, J. Mol. Biol. (1993). https://doi.org/10.1006/jmbi.1993.1626.

605 [20] R.A. Laskowski, M.W. MacArthur, D.S. Moss, J.M. Thornton, PROCHECK: a

606 program to check the stereochemical quality of protein structures, J. Appl. Crystallogr.

607 (1993). https://doi.org/10.1107/s0021889892009944.

608 [21] D. Eisenberg, R. Lüthy, J.U. Bowie, VERIFY3D: Assessment of protein models with

609 three-dimensional profiles, Methods Enzymol. (1997). https://doi.org/10.1016/S0076-

610 6879(97)77022-8.

611 [22] M. Wiederstein, M.J. Sippl, ProSA-web: Interactive web service for the recognition of

612 errors in three-dimensional structures of proteins, Nucleic Acids Res. (2007).

613 https://doi.org/10.1093/nar/gkm290.

614 [23] S.C. Lovell, I.W. Davis, W.B.A. Iii, P.I.W. de Bakker, J.M. Word, M.G. Prisant, J.S.

615 Richardson, D.C. Richardson, Structure validation by Calpha geometry: Phi,psi and

616 Cbeta deviation, Proteins. (2003).

617 [24] E.R. Moraes, E. Kalapothakis, L.A. Naves, C. Kushmerick, Differential effects of

618 tityus bahiensis scorpion venom on tetrodotoxin-sensitive and tetrodotoxin-resistant

619 sodium currents, Neurotox. Res. (2011). https://doi.org/10.1007/s12640-009-9144-8.

620 [25] W. BÜCHERL, Classification, Biology, and Venom Extraction of Scorpions, in:

621 Venom. Anim. Their Venoms, 1971. https://doi.org/10.1016/b978-0-12-138903-

622 1.50019-3.

623 [26] Q. Shu, S.P. Liang, Purification and characterization of huwentoxin-II, a neurotoxic

624 peptide from the venom of the Chinese bird spider Selenocosmia huwena, J. Pept. Res.

625 (1999). https://doi.org/10.1034/j.1399-3011.1999.00039.x.

626 [27] R.E. Oswald, T.M. Suchyna, R. McFeeters, P. Gottlieb, F. Sachs, Solution structure of

627 peptide toxins that block mechanosensitive ion channels, J. Biol. Chem. (2002).

628 https://doi.org/10.1074/jbc.M202715200.

29

629 [28] B. Chagot, P. Escoubas, E. Villegas, C. Bernard, G. Ferrat, G. Corzo, M. Lazdunski,

630 H. Darbon, Solution structure of Phrixotoxin 1, a specific peptide inhibitor of Kv4

631 potassium channels from the venom of the theraphosid spider Phrixotrichus auratus ,

632 Protein Sci. (2004). https://doi.org/10.1110/ps.03584304.

633 [29] S.T. Henriques, E. Deplazes, N. Lawrence, O. Cheneval, S. Chaousis, M. Inserra, P.

634 Thongyoo, G.F. King, A.E. Mark, I. Vetter, D.J. Craik, C.I. Schroeder, Interaction of

635 tarantula venom peptide ProTx-II with lipid membranes is a prerequisite for its

636 inhibition of human voltage-gated sodium channel NaV 1.7, J. Biol. Chem. (2016).

637 https://doi.org/10.1074/jbc.M116.729095.

638 [30] H. Xu, T. Li, A. Rohou, C.P. Arthur, F. Tzakoniati, E. Wong, A. Estevez, C. Kugel, Y.

639 Franke, J. Chen, C. Ciferri, D.H. Hackos, C.M. Koth, J. Payandeh, Structural Basis of

640 Nav1.7 Inhibition by a Gating-Modifier Spider Toxin, Cell. (2019).

641 https://doi.org/10.1016/j.cell.2018.12.018.

642 [31] A.J. Agwa, S. Peigneur, C.Y. Chow, N. Lawrence, D.J. Craik, J. Tytgat, G.F. King,

643 S.T. Henriques, C.I. Schroeder, Gating modifier toxins isolated from spider venom:

644 Modulation of voltage-gated sodium channels and the role of lipid membranes, J. Biol.

645 Chem. (2018). https://doi.org/10.1074/jbc.RA118.002553.

646 [32] Y. Li-Smerin, K.J. Swartz, Gating modifier toxins reveal a conserved structural motif

647 in voltage-gated Ca2+ and K+ channels, Proc. Natl. Acad. Sci. U. S. A. (1998).

648 https://doi.org/10.1073/pnas.95.15.8585.

649 [33] S. Sokolov, R.L. Kraus, T. Scheuer, W. a Catterall, Inhibition of sodium channel

650 gating by trapping the domain II voltage sensor with protoxin II., Mol. Pharmacol. 73

651 (2008) 1020–1028. https://doi.org/mol.107.041046 [pii]\r10.1124/mol.107.041046.

652 [34] E. Redaelli, R.R. Cassulini, D.F. Silva, H. Clement, E. Schiavon, F.Z. Zamudio, G.

653 Odell, A. Arcangeli, J.J. Clare, A. Alago, R.C. Rodríguez De La Vega, L.D. Possani,

654 E. Wanke, Target promiscuity and heterogeneous effects of tarantula venom peptides

30

655 affecting Na+ and K+ ion channels, J. Biol. Chem. (2010).

656 https://doi.org/10.1074/jbc.M109.054718.

657 [35] J.K. Klint, S. Senff, D.B. Rupasinghe, S.Y. Er, V. Herzig, G.M. Nicholson, G.F. King,

658 Spider-venom peptides that target voltage-gated sodium channels: Pharmacological

659 tools and potential therapeutic leads, Toxicon. 60 (2012) 478–491.

660 https://doi.org/10.1016/j.toxicon.2012.04.337.

661 [36] J.J. Smith, V. Herzig, M.P. Ikonomopoulou, S. Dziemborowicz, F. Bosmans, G.M.

662 Nicholson, G.F. King, Insect-active toxins with promiscuous pharmacology from the

663 African theraphosid spider Monocentropus balfouri, Toxins (Basel). (2017).

664 https://doi.org/10.3390/toxins9050155.

665 [37] S. Diochot, M.D. Drici, D. Moinier, M. Fink, M. Lazdunski, Effects of phrixotoxins on

666 the Kv4 family of potassium channels and implications for the role of Itoj in cardiac

667 electrogenesis, Br. J. Pharmacol. (1999). https://doi.org/10.1038/sj.bjp.0702283.

668 [38] F. Bosmans, L. Rash, S. Zhu, S. Diochot, M. Lazdunski, P. Escoubas, J. Tytgat, Four

669 novel tarantula toxins as selective modulators of voltage-gated sodium channel

670 subtypes., Mol. Pharmacol. 69 (2006) 419–429.

671 https://doi.org/10.1124/mol.105.015941.

672 [39] G.F. King, M.C. Gentz, P. Escoubas, G.M. Nicholson, A rational nomenclature for

673 naming peptide toxins from spiders and other venomous , Toxicon. 52 (2008)

674 264–276. https://doi.org/10.1016/j.toxicon.2008.05.020.

675

31