[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (25)

Search Parameters:
Keywords = Micrurus

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 1842 KiB  
Article
Retrospective Evaluation of Clinical and Clinicopathologic Findings, Case Management, and Outcome for Dogs and Cats Exposed to Micrurus fulvius (Eastern Coral Snake): 92 Cases (2021–2022)
by Jordan M. Sullivan, Taelor L. Aasen, Corey J. Fisher and Michael Schaer
Toxins 2024, 16(6), 246; https://doi.org/10.3390/toxins16060246 - 27 May 2024
Viewed by 1196
Abstract
This retrospective, observational study describes the clinical findings, case management trends, and outcomes of 83 dogs and nine cats exposed to eastern coral snakes in a university teaching hospital setting. The medical records of dogs and cats that received antivenom following coral snake [...] Read more.
This retrospective, observational study describes the clinical findings, case management trends, and outcomes of 83 dogs and nine cats exposed to eastern coral snakes in a university teaching hospital setting. The medical records of dogs and cats that received antivenom following coral snake exposure were reviewed. Data collected included signalment, time to antivenom administration, physical and laboratory characteristics at presentation, clinical course during hospitalization, length of hospitalization, and survival to discharge. The mean time from presentation to coral snake antivenom administration was 2.26 ± 1.46 h. Excluding cases where the owner declined in-hospital care, the mean hospitalization time for dogs and cats was 50.8 h and 34 h, respectively. The mean number of antivenom vials was 1.29 (1–4). Gastrointestinal signs (vomiting and ptyalism) occurred in 42.2% (35/83) of dogs and 33.3% (3/9) of cats. Peripheral neurologic system deficits (ataxia, paresis to plegia, absent reflexes, and hypoventilation) were noted in 19.6% (18/92) of dogs and cats. Hemolysis was also common in 37.9% (25/66) of dogs but was not observed in cats. Mechanical ventilation (MV) was indicated in 12% (10/83) of dogs but no cats. Acute kidney injury (AKI), while rare, was a common cause of euthanasia at 20% (2/5) and was the most common complication during MV at 44.4% (4/9). Pigmenturia/hemolysis occurred in 88.9% (8/9) of MV cases and in all cases with AKI. Despite delays in antivenom administration by several hours, dogs and cats with coral snake exposure have low mortality rates (6% of dogs (5/83) and 0% of cats). Gastrointestinal signs were common but were not predictive of progression to neurological signs. Thus, differentiating between coral snake exposure and envenomation before the onset of neurological signs remains challenging. Full article
(This article belongs to the Special Issue Pre-clinical and Clinical Management of Snakebite Envenomation)
Show Figures

Figure 1

Figure 1
<p>Three examples of the pinpoint tissue reaction left by the small fangs (proteroglyphous) of the eastern coral snake in dogs. (<b>A</b>) Small bleb on the oral mucosa of the upper lip, (<b>B</b>) slit-like wound on the outer lower lip, and (<b>C</b>) focal erythematous area on gingiva. (Courtesy, M. Schaer, University of Florida).</p>
Full article ">Figure 2
<p>An example of hemolysis (in the capillary tube on the <b>left</b>) and pigmenturia (in the urine collection cup on the <b>right</b>) from a dog undergoing hemolysis caused by eastern coral snake envenomation. This dog also had lower motor neuron dysfunction. Courtesy: M. Schaer, University of Florida.</p>
Full article ">Figure 3
<p>A mixed-breed dog that was envenomated by an eastern coral snake and had progressive lower motor neuron dysfunction with eventual respiratory paralysis that required ventilatory support for three days. Courtesy: Dr. M. Schaer, University of Florida.</p>
Full article ">
16 pages, 3236 KiB  
Article
Unveiling Novel Kunitz- and Waprin-Type Toxins in the Micrurus mipartitus Coral Snake Venom Gland: An In Silico Transcriptome Analysis
by Mónica Saldarriaga-Córdoba, Claudia Clavero-León, Paola Rey-Suarez, Vitelbina Nuñez-Rangel, Ruben Avendaño-Herrera, Stefany Solano-González and Juan F. Alzate
Toxins 2024, 16(5), 224; https://doi.org/10.3390/toxins16050224 - 11 May 2024
Cited by 2 | Viewed by 1517
Abstract
Kunitz-type peptide expression has been described in the venom of snakes of the Viperidae, Elapidae and Colubridae families. This work aimed to identify these peptides in the venom gland transcriptome of the coral snake Micrurus mipartitus. Transcriptomic analysis revealed a high diversity [...] Read more.
Kunitz-type peptide expression has been described in the venom of snakes of the Viperidae, Elapidae and Colubridae families. This work aimed to identify these peptides in the venom gland transcriptome of the coral snake Micrurus mipartitus. Transcriptomic analysis revealed a high diversity of venom-associated Kunitz serine protease inhibitor proteins (KSPIs). A total of eight copies of KSPIs were predicted and grouped into four distinctive types, including short KSPI, long KSPI, Kunitz–Waprin (Ku-WAP) proteins, and a multi-domain Kunitz-type protein. From these, one short KSPI showed high identity with Micrurus tener and Austrelaps superbus. The long KSPI group exhibited similarity within the Micrurus genus and showed homology with various elapid snakes and even with the colubrid Pantherophis guttatus. A third group suggested the presence of Kunitz domains in addition to a whey-acidic-protein-type four-disulfide core domain. Finally, the fourth group corresponded to a transcript copy with a putative 511 amino acid protein, formerly annotated as KSPI, which UniProt classified as SPINT1. In conclusion, this study showed the diversity of Kunitz-type proteins expressed in the venom gland transcriptome of M. mipartitus. Full article
(This article belongs to the Special Issue Transcriptomic and Proteomic Study on Animal Venom: Looking Forward)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Analysis of the putative short KSPIs identified in the <span class="html-italic">M. mipartitus</span> transcriptome. (<b>A</b>) Deduced amino acid sequence alignment of short Kunitz <span class="html-italic">M.m</span>-C482594 (accession number: PP439990) found in the <span class="html-italic">M. mipartitus</span> transcriptome with related sequences from public databases (<span class="html-italic">Micrurus tener (M.t)</span> and <span class="html-italic">Austrelaps superbus (A.s))</span>. Colored backgrounds indicate the signal peptide and Kunitz domains. Active site residues are indicated in red. Cysteines and the block of hydrophilic aa residues are indicated in gray. (<b>B</b>) Pairwise sequence similarity is highlighted according to the percent identity shown in the left-hand column.</p>
Full article ">Figure 2
<p>Analysis of the putative long KSPIs identified in the <span class="html-italic">M. mipartitus</span> transcriptome. (<b>A</b>) Deduced amino acid sequence alignment of long Kunitz <span class="html-italic">M.m</span>-NODE-3517 (accession number PP439991) found in <span class="html-italic">M. mipartitus</span> transcriptome with related sequences from public databases (M. tener (<span class="html-italic">M.t</span>), <span class="html-italic">M. fulvius</span> (<span class="html-italic">M.f</span>), <span class="html-italic">Notechis scutatus</span> (<span class="html-italic">N.s</span>), <span class="html-italic">Pseudonaja textilis</span> (<span class="html-italic">P.t</span>), <span class="html-italic">Austrelaps labialis</span> (<span class="html-italic">A.l</span>)<span class="html-italic">, Laticauda laticadus</span> (<span class="html-italic">L.l</span>), and <span class="html-italic">Pantherophis guttatus</span> (<span class="html-italic">P.g</span>)). The signal peptide and double Kunitz domains are highlighted by colored backgrounds. Active site residues are indicated in red, and N-linked glycosylation in position 46 is indicated in green. Cysteines and the block of hydrophilic aa residues are indicated in gray. (<b>B</b>) Pairwise sequence similarity is highlighted according to percent identity shown in the left-hand column.</p>
Full article ">Figure 3
<p>Analysis of the putative Ku-WAP fusin sequences identified in <span class="html-italic">M. mipartitus</span> transcriptome (<span class="html-italic">M.m</span>-NODE-13319, PP439993; <span class="html-italic">M.m</span>-NODE-19949, PP439992; <span class="html-italic">M.m</span>-NODE-19547, PP439994; <span class="html-italic">M.m</span>-NODE-21634, PP439996; and <span class="html-italic">M.m</span>-TRINITY-DN744, PP439995). (<b>A</b>) Deduced amino acid sequence alignment of the Ku-WAP fusin sequences found in <span class="html-italic">M. mipartitus</span> with related sequences from public databases (<span class="html-italic">Pantherophis guttatus</span> (<span class="html-italic">P.g</span>), <span class="html-italic">Pseudonaja textilis</span> (<span class="html-italic">P.t</span>), <span class="html-italic">Sistrurus catenatus edwardsii</span> (<span class="html-italic">S.c</span>), and <span class="html-italic">Notechis scutatus</span> (<span class="html-italic">N.s</span>)). * Ku-WAP fusin identified in viperids with loss of second cystein residue of the WAP domain. The different domains are highlighted by colored backgrounds. The modified residue in position 25 (Q- Pyrrolidone carboxylic acid) is depicted in yellow, the active site at positions 41-42 is depicted in red, and the red box indicates the potential cleavage site between Kunitz and WAP domains in Ku-WAP-fusin. Cysteines and the block of hydrophilic aa residues are indicated in gray. (<b>B</b>) Pairwise sequence similarity is highlighted according to percent identity shown in the left-hand column.</p>
Full article ">Figure 4
<p>Analysis of the putative Waprin sequences identified in the <span class="html-italic">M. mipartitus</span> transcriptome. Deduced amino acid sequence alignment of (<b>A</b>) single (<span class="html-italic">M.m</span>-C493867; PP439999) and (<b>C</b>) double (<span class="html-italic">Mm</span>-NODE_13492; PP439998) Waprin domains with related sequences published in the database. Signal peptides are indicated in gray, and the WAP domain is shown with a yellow background. Cysteines are indicated in gray. Pairwise sequence similarity is highlighted according to percent identity shown in the left-hand column. (<b>B</b>) Single Waprin domain and (<b>D</b>) double Waprin domain. <span class="html-italic">Micrurus mipartitus</span> (<span class="html-italic">M.m</span>), <span class="html-italic">Naja naja</span> (<span class="html-italic">N.n</span>), <span class="html-italic">Notechis scutatus</span> (<span class="html-italic">N.s</span>) and <span class="html-italic">Philodryas olfersii</span> (<span class="html-italic">P.o</span>).</p>
Full article ">Figure 5
<p>Deduced amino acid sequence alignment of multidomain putative protein (<span class="html-italic">M.m</span>-scaffold14755, PP439997) annotated as Kunitz identified in the <span class="html-italic">M. mipartitus</span> transcriptome with related sequences published in the database (<b>A</b>). Pairwise sequence similarity is highlighted according to percent identity shown in the left-hand column (<b>B</b>). The signal peptide is indicated in red. Active sites for each Kunitz domain are indicated as RG and EE residues. <span class="html-italic">Micrurus mipartitus</span> (<span class="html-italic">M.m</span>), <span class="html-italic">Pseudonaja textilis</span> (<span class="html-italic">P.t</span>), <span class="html-italic">Naja naja</span> (<span class="html-italic">N.n</span>), <span class="html-italic">Pantherophis guttatus</span> (<span class="html-italic">P.g</span>), and <span class="html-italic">Thamnophis sirtalis</span> (<span class="html-italic">T.s</span>).</p>
Full article ">
21 pages, 4265 KiB  
Article
Standard Quality Characteristics and Efficacy of a New Third-Generation Antivenom Developed in Colombia Covering Micrurus spp. Venoms
by Santiago Tabares Vélez, Lina María Preciado, Leidy Johana Vargas Muñoz, Carlos Alberto Madrid Bracamonte, Angelica Zuluaga, Jeisson Gómez Robles, Camila Renjifo-Ibañez and Sebastián Estrada-Gómez
Toxins 2024, 16(4), 183; https://doi.org/10.3390/toxins16040183 - 9 Apr 2024
Viewed by 2011
Abstract
In Colombia, Micrurus snakebites are classified as severe according to the national clinical care guidelines and must be treated with specific antivenoms. Unfortunately, these types of antivenoms are scarce in certain areas of the country and are currently reported as an unavailable vital [...] Read more.
In Colombia, Micrurus snakebites are classified as severe according to the national clinical care guidelines and must be treated with specific antivenoms. Unfortunately, these types of antivenoms are scarce in certain areas of the country and are currently reported as an unavailable vital medicine. To address this issue, La Universidad de Antioquia, through its spin-off Tech Life Saving, is leading a project to develop third-generation polyvalent freeze-dried antivenom. The goal is to ensure access to this therapy, especially in rural and dispersed areas. This project aims to evaluate the physicochemical and preclinical parameters (standard quality characteristics) of a lab-scale anti-elapid antivenom batch. The antivenom is challenged against the venoms of several Micrurus species, including M. mipartitus, M. dumerilii, M. ancoralis, M. dissoleucus, M. lemniscatus, M. medemi, M. spixii, M. surinamensis, and M. isozonus, following the standard quality characteristics set by the World Health Organization (WHO). The antivenom demonstrates an appearance consistent with standards, 100% solubility within 4 min and 25 s, an extractable volume of 10.39 mL, a pH of 6.04, an albumin concentration of 0.377 mg/mL (equivalent to 1.22% of total protein), and a protein concentration of 30.97 mg/mL. Importantly, it maintains full integrity of its F(ab′)2 fragments and exhibits purity over 98.5%. Furthermore, in mice toxicity evaluations, doses up to 15 mg/mouse show no toxic effects. The antivenom also demonstrates a significant recognition pattern against Micrurus venoms rich in phospholipase A2 (PLA2) content, as observed in M. dumerilii, M. dissoleucus, and M. isozonus. The effective dose 50 (ED50) indicates that a single vial (10 mL) can neutralize 2.33 mg of M. mipartitus venom and 3.99 mg of M. dumerilii venom. This new anti-elapid third-generation polyvalent and freeze-dried antivenom meets the physicochemical parameters set by the WHO and the regulators in Colombia. It demonstrates significant efficacy in neutralizing the venom of the most epidemiologically important Micrurus species in Colombia. Additionally, it recognizes seven other species of Micrurus venom with a higher affinity for venoms exhibiting PLA2 toxins. Fulfilling these parameters represents the first step toward proposing a new pharmacological alternative for treating snakebites in Colombia, particularly in dispersed rural areas, given that this antivenom is formulated as a freeze-dried product. Full article
(This article belongs to the Special Issue Pre-clinical and Clinical Management of Snakebite Envenomation)
Show Figures

Figure 1

Figure 1
<p>HPLC chromatographic profiles where each peak is numbered and corresponds to a lane in its respective 15% sodium dodecyl sulfate–polyacrylamide gels (SDS-PAGE) of the crude venom of (<b>A</b>) <span class="html-italic">M. mipartitus</span>, (<b>B</b>) <span class="html-italic">M. dumerilii</span>, (<b>C</b>) <span class="html-italic">M. spixii</span>, (<b>D</b>) <span class="html-italic">M. surinamensis</span>, (<b>E</b>) <span class="html-italic">M. lemniscatus</span>, (<b>F</b>) <span class="html-italic">M. medemi</span>, (<b>G</b>) <span class="html-italic">M. dissoleucus</span>, (<b>H</b>) <span class="html-italic">M. isozonus</span>, and (<b>I</b>) <span class="html-italic">M. ancoralis</span> using a C18 column (250 mm–4.6 mm), an elution gradient used: 0–70% of acetonitrile (99% in TFA 0.1%). The run was monitored at 215 nm, and the assignation of the regions was made using (<b>A</b>) Rey-Suárez et al., 2011 [<a href="#B5-toxins-16-00183" class="html-bibr">5</a>], (<b>B</b>) Rey-Suárez et al. (2016) [<a href="#B6-toxins-16-00183" class="html-bibr">6</a>], and (<b>C</b>–<b>E</b>) Sanz et al., 2019 [<a href="#B20-toxins-16-00183" class="html-bibr">20</a>,<a href="#B21-toxins-16-00183" class="html-bibr">21</a>].</p>
Full article ">Figure 1 Cont.
<p>HPLC chromatographic profiles where each peak is numbered and corresponds to a lane in its respective 15% sodium dodecyl sulfate–polyacrylamide gels (SDS-PAGE) of the crude venom of (<b>A</b>) <span class="html-italic">M. mipartitus</span>, (<b>B</b>) <span class="html-italic">M. dumerilii</span>, (<b>C</b>) <span class="html-italic">M. spixii</span>, (<b>D</b>) <span class="html-italic">M. surinamensis</span>, (<b>E</b>) <span class="html-italic">M. lemniscatus</span>, (<b>F</b>) <span class="html-italic">M. medemi</span>, (<b>G</b>) <span class="html-italic">M. dissoleucus</span>, (<b>H</b>) <span class="html-italic">M. isozonus</span>, and (<b>I</b>) <span class="html-italic">M. ancoralis</span> using a C18 column (250 mm–4.6 mm), an elution gradient used: 0–70% of acetonitrile (99% in TFA 0.1%). The run was monitored at 215 nm, and the assignation of the regions was made using (<b>A</b>) Rey-Suárez et al., 2011 [<a href="#B5-toxins-16-00183" class="html-bibr">5</a>], (<b>B</b>) Rey-Suárez et al. (2016) [<a href="#B6-toxins-16-00183" class="html-bibr">6</a>], and (<b>C</b>–<b>E</b>) Sanz et al., 2019 [<a href="#B20-toxins-16-00183" class="html-bibr">20</a>,<a href="#B21-toxins-16-00183" class="html-bibr">21</a>].</p>
Full article ">Figure 1 Cont.
<p>HPLC chromatographic profiles where each peak is numbered and corresponds to a lane in its respective 15% sodium dodecyl sulfate–polyacrylamide gels (SDS-PAGE) of the crude venom of (<b>A</b>) <span class="html-italic">M. mipartitus</span>, (<b>B</b>) <span class="html-italic">M. dumerilii</span>, (<b>C</b>) <span class="html-italic">M. spixii</span>, (<b>D</b>) <span class="html-italic">M. surinamensis</span>, (<b>E</b>) <span class="html-italic">M. lemniscatus</span>, (<b>F</b>) <span class="html-italic">M. medemi</span>, (<b>G</b>) <span class="html-italic">M. dissoleucus</span>, (<b>H</b>) <span class="html-italic">M. isozonus</span>, and (<b>I</b>) <span class="html-italic">M. ancoralis</span> using a C18 column (250 mm–4.6 mm), an elution gradient used: 0–70% of acetonitrile (99% in TFA 0.1%). The run was monitored at 215 nm, and the assignation of the regions was made using (<b>A</b>) Rey-Suárez et al., 2011 [<a href="#B5-toxins-16-00183" class="html-bibr">5</a>], (<b>B</b>) Rey-Suárez et al. (2016) [<a href="#B6-toxins-16-00183" class="html-bibr">6</a>], and (<b>C</b>–<b>E</b>) Sanz et al., 2019 [<a href="#B20-toxins-16-00183" class="html-bibr">20</a>,<a href="#B21-toxins-16-00183" class="html-bibr">21</a>].</p>
Full article ">Figure 1 Cont.
<p>HPLC chromatographic profiles where each peak is numbered and corresponds to a lane in its respective 15% sodium dodecyl sulfate–polyacrylamide gels (SDS-PAGE) of the crude venom of (<b>A</b>) <span class="html-italic">M. mipartitus</span>, (<b>B</b>) <span class="html-italic">M. dumerilii</span>, (<b>C</b>) <span class="html-italic">M. spixii</span>, (<b>D</b>) <span class="html-italic">M. surinamensis</span>, (<b>E</b>) <span class="html-italic">M. lemniscatus</span>, (<b>F</b>) <span class="html-italic">M. medemi</span>, (<b>G</b>) <span class="html-italic">M. dissoleucus</span>, (<b>H</b>) <span class="html-italic">M. isozonus</span>, and (<b>I</b>) <span class="html-italic">M. ancoralis</span> using a C18 column (250 mm–4.6 mm), an elution gradient used: 0–70% of acetonitrile (99% in TFA 0.1%). The run was monitored at 215 nm, and the assignation of the regions was made using (<b>A</b>) Rey-Suárez et al., 2011 [<a href="#B5-toxins-16-00183" class="html-bibr">5</a>], (<b>B</b>) Rey-Suárez et al. (2016) [<a href="#B6-toxins-16-00183" class="html-bibr">6</a>], and (<b>C</b>–<b>E</b>) Sanz et al., 2019 [<a href="#B20-toxins-16-00183" class="html-bibr">20</a>,<a href="#B21-toxins-16-00183" class="html-bibr">21</a>].</p>
Full article ">Figure 2
<p>Appearance test of the antivenom product freeze-dried on a white background (<b>A</b>) and black background (<b>B</b>) and resuspended on a white background (<b>C</b>) and black background (<b>D</b>).</p>
Full article ">Figure 3
<p>(<b>A</b>). SEC-HPLC of one freeze-dried polyvalent antivenom vial using a column Sec 3000 Phenomenex, mobile phase buffer Na<sub>2</sub>HPO<sub>4</sub> 15 mM NaH<sub>2</sub>PO<sub>4</sub> 30 mM NaCl 200 mM (pH 7.0), flux rate 0.5 mL/min, and detection of UV 280 nm. The red dashed line indicates the beginning and end peak elution time to measure the area under the curve (<b>B</b>). Electrophoresis of the freeze-dried polyvalent antivenom (Lines 1 and 2) and INS antivenom (Line 3) under non-reducing conditions using 12% sodium dodecyl sulfate–polyacrylamide gels (SDS-PAGE). Lowercase letters a–e indicate the recognized bands with the respective calculated MW using Gel Analyzer 19.1 [<a href="#B23-toxins-16-00183" class="html-bibr">23</a>].</p>
Full article ">Figure 3 Cont.
<p>(<b>A</b>). SEC-HPLC of one freeze-dried polyvalent antivenom vial using a column Sec 3000 Phenomenex, mobile phase buffer Na<sub>2</sub>HPO<sub>4</sub> 15 mM NaH<sub>2</sub>PO<sub>4</sub> 30 mM NaCl 200 mM (pH 7.0), flux rate 0.5 mL/min, and detection of UV 280 nm. The red dashed line indicates the beginning and end peak elution time to measure the area under the curve (<b>B</b>). Electrophoresis of the freeze-dried polyvalent antivenom (Lines 1 and 2) and INS antivenom (Line 3) under non-reducing conditions using 12% sodium dodecyl sulfate–polyacrylamide gels (SDS-PAGE). Lowercase letters a–e indicate the recognized bands with the respective calculated MW using Gel Analyzer 19.1 [<a href="#B23-toxins-16-00183" class="html-bibr">23</a>].</p>
Full article ">Figure 4
<p>Double immunodiffusion test in agarose gel at 1% of (<b>A</b>) <span class="html-italic">M. mipartitus</span>, (<b>B</b>) <span class="html-italic">M. dumerilii</span>, (<b>C</b>) <span class="html-italic">M. isozonus</span>, (<b>D</b>) <span class="html-italic">M. dissoleucus</span>, (<b>E</b>) <span class="html-italic">M. ancoralis</span>, (<b>F</b>) <span class="html-italic">M. lemniscatus</span>, (<b>G</b>) <span class="html-italic">M. spixii</span>, (<b>H</b>) <span class="html-italic">M. medemi</span>, and (<b>I</b>) <span class="html-italic">M. surinamensis</span> venoms (V), 30 µL at a concentration of 1 mg/mL against antivenom of Universidad de Antioquia (AV), 30 µL at a concentration of 30 mg/mL, and water for injection as negative control (C−).</p>
Full article ">Figure 5
<p>Antivenom recognition in ELISA test for the complete <span class="html-italic">Micrurus</span> venoms. The lines with circular (●) marks refer to species with PLA<sub>2</sub>s-predominant venoms and (<b>A</b>,<b>B</b>) squares (■) 3FTxs-predominant (<b>C</b>). The cut-off point was three times the capacity of recognition of the plasm of a non-immunized horse. Each dilution was tested in quadruplicate using two separate vials (8 repetitions in total).</p>
Full article ">
25 pages, 6998 KiB  
Article
The Cloning and Characterization of a Three-Finger Toxin Homolog (NXH8) from the Coralsnake Micrurus corallinus That Interacts with Skeletal Muscle Nicotinic Acetylcholine Receptors
by Henrique Roman-Ramos, Álvaro R. B. Prieto-da-Silva, Humberto Dellê, Rafael S. Floriano, Lourdes Dias, Stephen Hyslop, Raphael Schezaro-Ramos, Denis Servent, Gilles Mourier, Jéssica Lopes de Oliveira, Douglas Edgard Lemes, Letícia V. Costa-Lotufo, Jane S. Oliveira, Milene Cristina Menezes, Regina P. Markus and Paulo Lee Ho
Toxins 2024, 16(4), 164; https://doi.org/10.3390/toxins16040164 - 22 Mar 2024
Cited by 1 | Viewed by 1821
Abstract
Coralsnakes (Micrurus spp.) are the only elapids found throughout the Americas. They are recognized for their highly neurotoxic venom, which is comprised of a wide variety of toxins, including the stable, low-mass toxins known as three-finger toxins (3FTx). Due to difficulties in [...] Read more.
Coralsnakes (Micrurus spp.) are the only elapids found throughout the Americas. They are recognized for their highly neurotoxic venom, which is comprised of a wide variety of toxins, including the stable, low-mass toxins known as three-finger toxins (3FTx). Due to difficulties in venom extraction and availability, research on coralsnake venoms is still very limited when compared to that of other Elapidae snakes like cobras, kraits, and mambas. In this study, two previously described 3FTx from the venom of M. corallinus, NXH1 (3SOC1_MICCO), and NXH8 (3NO48_MICCO) were characterized. Using in silico, in vitro, and ex vivo experiments, the biological activities of these toxins were predicted and evaluated. The results showed that only NXH8 was capable of binding to skeletal muscle cells and modulating the activity of nAChRs in nerve–diaphragm preparations. These effects were antagonized by anti-rNXH8 or antielapidic sera. Sequence analysis revealed that the NXH1 toxin possesses eight cysteine residues and four disulfide bonds, while the NXH8 toxin has a primary structure similar to that of non-conventional 3FTx, with an additional disulfide bond on the first loop. These findings add more information related to the structural diversity present within the 3FTx class, while expanding our understanding of the mechanisms of the toxicity of this coralsnake venom and opening new perspectives for developing more effective therapeutic interventions. Full article
(This article belongs to the Section Animal Venoms)
Show Figures

Figure 1

Figure 1
<p>The nucleotide and deduced protein sequence of the nxh8 cDNA clone. The cDNA clone structure of the nxh8 gene from the venom gland of <span class="html-italic">Micrurus corallinus</span> (EMBL data bank accession number AJ344067) comprises the following regions: a 5′ untranslated region (UTR) (1–5 bp), a signal peptide coding sequence (6–68 bp), a mature peptide coding sequence (69–263 bp, highlighted in a box), and a 3′UTR (264–470 bp). The polyadenylation signal is indicated in bold. The deduced mature peptide sequence (NXH8) starts at the Leu residue and is indicated by an arrow. The sequence utilized for the 5′ primer design for PCR amplification and subcloning into the pRSET C expression vector is emphasized with an underline. The blue numbers on the left indicate amino acid residue positions, while the black numbers on the right indicate nucleotide positions.</p>
Full article ">Figure 2
<p>Alignment of NXH8 (3NO48_MICCO) with different Elapidae 3FTx highlighting functionally invariant residues related to skeletal muscle nAChR binding. Functionally invariant residues in three-finger α-neurotoxins that interact with muscular nicotinic acetylcholine receptors (nAChRs) are shaded in gray, based on experimental mutational data for Erabutoxin-a (3S1EA_LATSE) [<a href="#B78-toxins-16-00164" class="html-bibr">78</a>,<a href="#B79-toxins-16-00164" class="html-bibr">79</a>], NMM I (3S11_NAJMO) [<a href="#B80-toxins-16-00164" class="html-bibr">80</a>,<a href="#B81-toxins-16-00164" class="html-bibr">81</a>], and α-Cobratoxin (3L21_NAJKA) [<a href="#B82-toxins-16-00164" class="html-bibr">82</a>,<a href="#B83-toxins-16-00164" class="html-bibr">83</a>], as well as structural data for α-Bungarotoxin (3L21A_BUNMU) [<a href="#B84-toxins-16-00164" class="html-bibr">84</a>,<a href="#B85-toxins-16-00164" class="html-bibr">85</a>]. Residues critical for binding to skeletal muscle (α1)2β1γδ nAChRs, which are shared between short- and long-chain α-neurotoxins, are indicated in bold. Putative functional conserved residues in toxins such as WTX (3NO2_NAJKA), Candoxin (3NO4_BUNCA), and NXH8 (3NO48_MICCO), with analogous binding functions, are also presented. The cysteine residues and structurally invariant residues are enclosed in boxes, and the disulfide bridges are highlighted. Furthermore, the segments contributing to the three loops and the C-terminus are labeled. The fifth disulfide bridge in long-chain α-neurotoxins, between Cys<sup>32</sup> and Cys<sup>36</sup>, is represented by a dotted line, while the corresponding bridge in non-conventional three-finger toxins, between Cys<sup>6</sup> and Cys<sup>11</sup>, is shown as a dashed line.</p>
Full article ">Figure 3
<p>Alignment of NXH1 (3SOC1_MICCO) with different Elapidae 3FTx highlighting functionally invariant residues related to skeletal muscle nAChR binding. Functionally invariant residues in three-finger α-neurotoxins that interact with muscular nicotinic acetylcholine receptors (nAChRs) are shaded in gray, based on experimental mutational data for Erabutoxin-a (3S1EA_LATSE). Residues critical for binding to skeletal muscle (α)2βγδ nAChRs, which are shared between short- and long-chain α-neurotoxins, are indicated in bold. Putative functional conserved residues in toxins such as NXH1 from <span class="html-italic">M. corallinus</span> (3SOC1_MICCO), short neurotoxin 1 from <span class="html-italic">Hydrophis cyanocinctus</span> (3S11_HYDCY), short neurotoxins 1 from <span class="html-italic">Hydrophis schitosus</span> (3S11_HYDSC), Toxin 5, from <span class="html-italic">H. schitosus</span> (3S15_HYDSC), short neurotoxin A, from <span class="html-italic">Aipysurus laevis</span> (3S11_AIPLA), short neurotoxin B, from <span class="html-italic">A. laevis</span> (3S12_AIPLA), short neurotoxin c, from <span class="html-italic">A. laevis</span> (3S13_AIPLA), short neurotoxin D, from <span class="html-italic">A. laevis</span> (3S14_AIPLA), Cobrotoxin-b, from <span class="html-italic">Naja atra</span> (3S1CC_NAJAT), short neurotoxin 1, from <span class="html-italic">Hydrophis lapemoides</span> (3S11_HYDLA), short neurotoxins 1, from <span class="html-italic">Naja oxiana</span> (3S11_NAJOX), and three-finger toxin Mnn I, from <span class="html-italic">Micrurus nigrocinctus</span> (3S11_MICNI). The cysteine residues and structurally invariant residues are enclosed in boxes, and the disulfide bridges are highlighted. Furthermore, the segments contributing to the three loops and the C-terminus are labeled.</p>
Full article ">Figure 4
<p>Alignment of NXH1 (3SOC1_MICCO) with two fasciculins 3FTx from <span class="html-italic">Dendroaspis</span> spp. Putative functional conserved residues in toxins such as NXH1 from <span class="html-italic">M. corallinus</span> (3SOC1_MICCO), Fasciculin-1, from <span class="html-italic">D. angusticeps</span> (3SE1_DENAN), Acetylcholinesterase toxin C, from <span class="html-italic">D. polylepis polylepis</span> (3SEC_DENPO). The cysteine residues and structurally invariant residues are enclosed in boxes, and the disulfide bridges are highlighted. Furthermore, the segments contributing to the three loops and the C-terminus are labeled.</p>
Full article ">Figure 5
<p>Dendrogram of three-finger toxins family. A blue arrow (<span style="color:blue">➔</span>) indicates the NXH8 ally group. 3FTx with a fifth disulfide bridge at loop I are indicated by a star (*). The dendrogram shows that 3FTx with a disulfide bridge at loop I is a non-homogeneous group, probably with distinct functions. The <span class="html-italic">M. corallinus</span> three-finger toxins NXH1 and NXH8 (shaded) are unrelated. Orphan groups’ clades are numbered in agreement with Fry, B.G. et al. [<a href="#B36-toxins-16-00164" class="html-bibr">36</a>]. The parentheses display, respectively, the identity and similarity percentages of each toxin in relation to NXH8. Bootstrap values are shown in red.</p>
Full article ">Figure 6
<p>SDS-PAGE analysis of recombinant expression and purification of rNXH8 in <span class="html-italic">Escherichia coli</span> cells transformed with pRSETC-<span class="html-italic">nxh8</span> plasmid, with each lane loaded with a 10 μL sample volume. (<b>a</b>) <span class="html-italic">E. coli</span> BL21 (DE3)-pRSETC-nxh8 cell extracts: 1. Molecular-mass marker; 2. Cell extract before IPTG induction; 3. Cell extract after IPTG induction; (<b>b</b>) Metal ion affinity chromatography, under denaturing conditions (8 M urea/10 mM 2-ME), of rNXH8 from solubilized inclusion bodies: 1. Molecular-mass marker; 2. <span class="html-italic">E. coli</span> BL21 (DE3)-pRSETC-<span class="html-italic">nxh8</span> cell extract after IPTG induction; 3. Supernatant after cell lysate centrifugation; 4. Precipitate after cell lysate centrifugation; 5. Inclusion bodies after solubilization in 8 M urea/10 mM 2-ME buffer; 6. Non-adsorbed material (flow through), after sample loading; 7. Ni<sup>2+</sup>-charged resin after sample loading; 8. Non-adsorbed material after column wash; 9. Adsorbed material after elution; (<b>c</b>) Metal ion affinity chromatography of refolded rNXH8: 1. Molecular-mass marker; 2. Cell lysate extract; 3. Column wash with 5 mM imidazole; 4. Adsorbed material after elution with 1 column volume of 250 mM imidazole elution buffer; 5. Adsorbed material after elution with five column volumes of 250 mM imidazole elution buffer. 6. Adsorbed material after EDTA chelating elution; (<b>d</b>) Reducing SDS-PAGE analysis of rNXH8 aggregate formation during dialysis (2-ME was added to samples): 1. Molecular-mass marker; 2. Inclusion bodies after solubilization in 8 M urea/10 mM 2-ME buffer; 3. Adsorbed material after elution with 2 M urea/10 mM Imidazole buffer; 4. Full dialysis material; 5. Supernatant after dialysis material centrifugation; 6. Precipitate after dialysis material centrifugation. (<b>e</b>) Non-reducing SDS-PAGE analysis of rNXH8-aggregate formation during dialysis (2-ME was not added to sample): 1. Molecular-mass marker; 2. Inclusion bodies after solubilization in 8 M urea/10 mM 2-ME buffer; 3. Adsorbed material after elution with 2 M urea/10 mM Imidazole buffer; 4. Full dialysis material; 5. Supernatant after dialysis material centrifugation; 6. Precipitate after dialysis material centrifugation.</p>
Full article ">Figure 7
<p>Cross-reactivity of anti-rNXH8 in diverse snake venoms by Western-blot. (<b>a</b>) SDS-PAGE stained with Coomassie Blue R250 (gradient 10–20%), with each lane loaded with a 10 μL sample volume. (<b>b</b>) Western blot of a replica gel after electroblotting to a nitrocellulose support incubated with anti-NXH8 polyclonal serum. 1. Molecular-mass marker; 2. Venom of <span class="html-italic">M. corallinus</span>; 3. Venom of <span class="html-italic">M. ibiboboca</span>; 4. Venom of <span class="html-italic">M. lemniscatus</span>; 5. Venom of <span class="html-italic">M. spixii</span>; 6. Venom of <span class="html-italic">M. frontalis</span>; 7. Venom of <span class="html-italic">M. altirostris</span>; 8. Venom of <span class="html-italic">M. surinamensis</span>; 9. Venom of <span class="html-italic">M. carinicauda dumerilli</span>; 10. Venom of <span class="html-italic">M. hemprichii</span>; 11. Venom of <span class="html-italic">M. spixii martiusi</span>; 12. Venom of <span class="html-italic">M. decoratus</span>; 13. Molecular-mass marker; 14. Venom of <span class="html-italic">Dendroaspis angusticeps</span>; 15. Venom of <span class="html-italic">Notechis scutatus scutatus</span>; 16. Venom of <span class="html-italic">Bungarus multicinctus</span>; 17. Venom of <span class="html-italic">Crotalus durissus terrificus</span>; 18. Purified cardiotoxin IV <span class="html-italic">Naja naja kaouthia</span>; 19. Venom of <span class="html-italic">Bothrops jararaca</span>. 20. Molecular-mass marker. The polyclonal serum against recombinant NXH8 reacts with homologous <span class="html-italic">M. corallinus</span> venom and with heterologous venoms from <span class="html-italic">M. altirostris</span> (lane 07), <span class="html-italic">Dendroaspis angusticeps</span> (lane 14) and <span class="html-italic">Bungarus multicinctus</span> (lane 16).</p>
Full article ">Figure 8
<p>Acetylcholine receptor-binding assay utilizing primary skeletal muscle cell-membrane preparations from newborn rats. The assay employed [<sup>125</sup>I]-labeled α-Bungarotoxin as a tracer. Membrane preparations were incubated for one hour with either nicotine or the <span class="html-italic">M. corallinus</span> crude venom, followed by a wash in binding buffer, and then further incubated for an additional hour at room temperature with [<sup>125</sup>I]-labeled α-Bungarotoxin. Subsequently, the membranes were rewashed, and the resulting pellets, obtained by centrifugation, were analyzed for residual radioactivity. In the neutralizing venom assay, diminishing quantities of crude venom were pre-incubated for 30 min with 10 µL of polyclonal sera (anti-rNXH8, anti-rNXH1, or anti-MIX*) before the addition to the membrane preparation. Assays were performed in quadruplicate. Statistical analysis was conducted using a 2-way ANOVA with multiple comparisons in GraphPad Prism 10 (GraphPad Software, Boston, MA, USA), with a significance threshold of <span class="html-italic">p</span> &lt; 0.05. Different lowercase letters (a–e) above the bars denote statistically distinct groups. Note: * Anti-MIX represents a 1:1 mixture of anti-rNXH8 and anti-rNXH1.</p>
Full article ">Figure 9
<p>Representative recordings showing the neuromuscular activities NXH1 and NXH8 three-finger toxins in indirectly stimulated PND preparations at 37 °C. (<b>a</b>) Reversion of synthetic NXH8 (sNXH8) blockade by saline washing (<span class="html-italic">W</span>); (<b>b</b>) Reversion of sNXH8 blockade by neostigmine (29 μM); (<b>c</b>) Reversion of sNXH8 (10 μg/mL) blockade by 3,4-DAP (230 μM); (<b>d</b>) sNXH8 activity after pre-incubation (37 °C, 30 min, 1:1 <span class="html-italic">v</span>/<span class="html-italic">w</span> antivenom–toxin ratio) with antielapidic serum from <span class="html-italic">Instituto Butantan</span>; (<b>e</b>) sNXH8 activity after pre-incubation (37 °C, 30 min, 1:1 <span class="html-italic">v</span>/<span class="html-italic">w</span> antivenom–toxin ratio) with anti-<span class="html-italic">Oxyuranus scutellatus</span> (Coastal Taipan) serum; (<b>f</b>) Synthetic NXH1 (sNXH1) activity; (<b>g</b>) Recombinant NXH8 (rNXH8) activity.</p>
Full article ">
13 pages, 1864 KiB  
Article
Immunological Cross-Reactivity and Preclinical Assessment of a Colombian Anticoral Antivenom against the Venoms of Three Micrurus Species
by Ariadna Rodríguez-Vargas, Adrián Marcelo Franco-Vásquez, Miguel Triana-Cerón, Shaha Noor Alam-Rojas, Derly C. Escobar-Wilches, Gerardo Corzo, Fernando Lazcano-Pérez, Roberto Arreguín-Espinosa and Francisco Ruiz-Gómez
Toxins 2024, 16(2), 104; https://doi.org/10.3390/toxins16020104 - 15 Feb 2024
Viewed by 2568
Abstract
Snakebite accident treatment requires the administration of antivenoms that provide efficacy and effectiveness against several snake venoms of the same genus or family. The low number of immunogenic components in venom mixtures that allow the production of antivenoms consequently gives them partial neutralization [...] Read more.
Snakebite accident treatment requires the administration of antivenoms that provide efficacy and effectiveness against several snake venoms of the same genus or family. The low number of immunogenic components in venom mixtures that allow the production of antivenoms consequently gives them partial neutralization and a suboptimal pharmacological response. This study evaluates the immunorecognition and neutralizing efficacy of the polyvalent anticoral antivenom from the Instituto Nacional de Salud (INS) of Colombia against the heterologous endemic venoms of Micrurus medemi, and M. sangilensis, and M. helleri by assessing immunoreactivity through affinity chromatography, ELISA, Western blot, and neutralization capability. Immunorecognition towards the venoms of M. medemi and M. sangilensis showed values of 62% and 68% of the protein composition according to the immunoaffinity matrix, respectively. The analysis by Western blot depicted the highest recognition patterns for M. medemi, followed by M. sangilensis, and finally by M. helleri. These findings suggest that the venom compositions are closely related and exhibit similar recognition by the antivenom. According to enzyme immunoassays, M. helleri requires a higher amount of antivenom to achieve recognition than the others. Besides reinforcing the evaluation of INS antivenom capability, this work recommends the use of M. helleri in the production of Colombian antisera. Full article
(This article belongs to the Special Issue Snake Venom: Toxicology and Associated Countermeasures)
Show Figures

Figure 1

Figure 1
<p>Cross-immunorecognition through ELISA assay of INS polyvalent anticoral antivenom against the crude venoms of <span class="html-italic">Micrurus helleri</span>, <span class="html-italic">M. medemi</span>, and <span class="html-italic">M. sangilensis</span>. IgG of non-immunized horses was used as negative control {C(−)}. <span class="html-italic">p</span>-value = 0.05. Each point represents the average of three measurements.</p>
Full article ">Figure 2
<p>Electrophoretic and immunorecognition profiles. (<b>a</b>) SDS-PAGE, 15% of the three venoms under reducing conditions (MWM: Molecular Weight Marker); (<b>b</b>) Western blot of INS polyvalent anticoral antivenom against the venoms of three Colombian <span class="html-italic">Micrurus</span> snakes. <span class="html-italic">Micrurus medemi</span> (lane 1), <span class="html-italic">M. helleri</span> (lane 2), and <span class="html-italic">M. sangilensis</span> (lane 3). (<b>c</b>) Western blot densitometry analysis using GelAnalyzer software v. 23.1.1 (available at <a href="http://www.gelanalyzer.com" target="_blank">www.gelanalyzer.com</a> by Istvan Lazar Jr., PhD and Istvan Lazar Sr., PhD, CSc).</p>
Full article ">Figure 3
<p>RP-HPLC chromatograms and heat maps of the not retained (red) and retained (blue) fractions from the affinity matrix coupled to anticoral antivenom produced by the Instituto Nacional de Salud. The black solid line shows the chromatogram of the whole venom. The heat maps indicate the immunorecognition of venom fractions by the anticoral antivenom. The immunorecognition percentages for the three venom fractions by RP-HPLC from the immunoaffinity column are based on calculated relative abundances. (<b>a</b>) <span class="html-italic">M. medemi</span>, (<b>b</b>) <span class="html-italic">M. helleri</span>, and (<b>c</b>) <span class="html-italic">M. sangilensis</span>.</p>
Full article ">
18 pages, 4636 KiB  
Systematic Review
Knowledge about Snake Venoms and Toxins from Colombia: A Systematic Review
by Jaime Andrés Pereañez, Lina María Preciado and Paola Rey-Suárez
Toxins 2023, 15(11), 658; https://doi.org/10.3390/toxins15110658 - 15 Nov 2023
Cited by 1 | Viewed by 2934
Abstract
Colombia encompasses three mountain ranges that divide the country into five natural regions: Andes, Pacific, Caribbean, Amazon, and Orinoquia. These regions offer an impressive range of climates, altitudes, and landscapes, which lead to a high snake biodiversity. Of the almost 300 snake species [...] Read more.
Colombia encompasses three mountain ranges that divide the country into five natural regions: Andes, Pacific, Caribbean, Amazon, and Orinoquia. These regions offer an impressive range of climates, altitudes, and landscapes, which lead to a high snake biodiversity. Of the almost 300 snake species reported in Colombia, nearly 50 are categorized as venomous. This high diversity of species contrasts with the small number of studies to characterize their venom compositions and natural history in the different ecoregions. This work reviews the available information about the venom composition, isolated toxins, and potential applications of snake species found in Colombia. Data compilation was conducted according to the PRISMA guidelines, and the systematic literature search was carried out in Pubmed/MEDLINE. Venom proteomes from nine Viperidae and three Elapidae species have been described using quantitative analytical strategies. In addition, venoms of three Colubridae species have been studied. Bioactivities reported for some of the venoms or isolated components—such as antibacterial, cytotoxicity on tumoral cell lines, and antiplasmodial properties—may be of interest to develop potential applications. Overall, this review indicates that, despite recent progress in the characterization of venoms from several Colombian snakes, it is necessary to perform further studies on the many species whose venoms remain essentially unexplored, especially those of the poorly known genus Micrurus. Full article
(This article belongs to the Section Animal Venoms)
Show Figures

Figure 1

Figure 1
<p>PRISMA flow diagram for the literature search strategy.</p>
Full article ">Figure 2
<p>Quantitative venom proteomes from Viperidae family. Venom localities: <span class="html-italic">B. atrox</span> (Meta) [<a href="#B17-toxins-15-00658" class="html-bibr">17</a>], <span class="html-italic">B. asper</span> (Cauca) [<a href="#B15-toxins-15-00658" class="html-bibr">15</a>,<a href="#B16-toxins-15-00658" class="html-bibr">16</a>], <span class="html-italic">B. ayerbei</span> (Cauca) [<a href="#B16-toxins-15-00658" class="html-bibr">16</a>], <span class="html-italic">B. rhombeatus</span> (Cauca) [<a href="#B15-toxins-15-00658" class="html-bibr">15</a>], <span class="html-italic">B. punctatus</span> (Antioquia) [<a href="#B18-toxins-15-00658" class="html-bibr">18</a>], <span class="html-italic">P. lansbergii</span> (Caribbean) [<a href="#B20-toxins-15-00658" class="html-bibr">20</a>], <span class="html-italic">B. myersi</span> (Valle del Cauca) [<a href="#B19-toxins-15-00658" class="html-bibr">19</a>], <span class="html-italic">C. d. cumanensis</span> (pool from Meta, Tolima, Cundinamarca, and Magdalena), <span class="html-italic">L. acrochorda</span> (pool from Antioquia and Chocó). Abbreviations for protein family names: PLA<sub>2</sub>s: phospholipase A<sub>2</sub>; SVMPs: metalloproteinase; LAAO: L-amino acid oxidase; CTL: C-type lectin/lectin-like; CRISP: cysteine-rich secretory protein; Dis: Disintegrins; SVSPs: serine proteinase; Nuc: nucleotidase; PDE: phosphodiesterase; Hya: hyaluronidase; NGF: nerve growth factor; PLB: phospholipase B; PNP: peptides and/or nonproteinaceous components; BPP: bradykinin-potentiating peptide. The unknown fractions were not considered in this graph.</p>
Full article ">Figure 3
<p>Quantitative venom proteomes from the Elapidae family. Venom localities: <span class="html-italic">M. mipartitus</span> (Antioquia) [<a href="#B29-toxins-15-00658" class="html-bibr">29</a>], <span class="html-italic">M. dumerilii</span> (Antioquia) [<a href="#B30-toxins-15-00658" class="html-bibr">30</a>], and <span class="html-italic">M. lemniscatus helleri</span> (Amazonas) [<a href="#B31-toxins-15-00658" class="html-bibr">31</a>]. Abbreviations for protein family names: 3FTx: three-finger toxins; PLA<sub>2</sub>s: phospholipase A<sub>2</sub>; SVMPs: metalloproteinase; LAAO: L-amino acid oxidase; CTL: C-type lectin/lectin-like; SVSPs: serine proteinase; Nuc: nucleotidase; PDE: phosphodiesterase; Hya: hyaluronidase; Kun: Kunitz-type inhibitors; PLB: phospholipase B; PNP: peptides and/or nonproteinaceous components. The unknown fractions were not considered in this graph.</p>
Full article ">Figure 4
<p>Biological activities reported for Colombian snake venoms from the Viperidae family. <span class="html-italic">B. atrox</span> (Meta) [<a href="#B33-toxins-15-00658" class="html-bibr">33</a>]; <span class="html-italic">B. asper</span> (Cauca and Antioquia) [<a href="#B16-toxins-15-00658" class="html-bibr">16</a>,<a href="#B34-toxins-15-00658" class="html-bibr">34</a>]; <span class="html-italic">B. ayerbei</span> (Cauca) [<a href="#B16-toxins-15-00658" class="html-bibr">16</a>]; <span class="html-italic">B. rhombeatus</span> (Cauca) [<a href="#B16-toxins-15-00658" class="html-bibr">16</a>]; <span class="html-italic">B. punctatus</span> (Antioquia) [<a href="#B18-toxins-15-00658" class="html-bibr">18</a>]; <span class="html-italic">P. lansbergii</span> (Atlántico) [<a href="#B20-toxins-15-00658" class="html-bibr">20</a>]; <span class="html-italic">B. myersi</span> (Valle del Cauca) [<a href="#B19-toxins-15-00658" class="html-bibr">19</a>]; <span class="html-italic">P. nasutum</span> (Antioquia) [<a href="#B34-toxins-15-00658" class="html-bibr">34</a>]; <span class="html-italic">B. schlegelii</span> (Antioquia) [<a href="#B34-toxins-15-00658" class="html-bibr">34</a>,<a href="#B37-toxins-15-00658" class="html-bibr">37</a>]; <span class="html-italic">L. acrochorda</span> (Antioquia) [<a href="#B34-toxins-15-00658" class="html-bibr">34</a>]; <span class="html-italic">C. d. cumanensis</span> (Meta) [<a href="#B34-toxins-15-00658" class="html-bibr">34</a>].</p>
Full article ">Figure 5
<p>Geographical distribution of snake venoms of Viperidae (blue) and Elapidae (yellow) families described in this systematic review.</p>
Full article ">
22 pages, 4249 KiB  
Article
Unveiling the Venom Composition of the Colombian Coral Snakes Micrurus helleri, M. medemi, and M. sangilensis
by Ariadna Rodríguez-Vargas, Adrián Marcelo Franco-Vásquez, Janeth Alejandra Bolívar-Barbosa, Nohora Vega, Edgar Reyes-Montaño, Roberto Arreguín-Espinosa, Alejandro Carbajal-Saucedo, Teddy Angarita-Sierra and Francisco Ruiz-Gómez
Toxins 2023, 15(11), 622; https://doi.org/10.3390/toxins15110622 - 24 Oct 2023
Cited by 6 | Viewed by 3524
Abstract
Little is known of the biochemical composition and functional features of the venoms of poorly known Colombian coral snakes. Here, we provide a preliminary characterization of the venom of two Colombian endemic coral snake species, Micrurus medemi and M. sangilensis, as well [...] Read more.
Little is known of the biochemical composition and functional features of the venoms of poorly known Colombian coral snakes. Here, we provide a preliminary characterization of the venom of two Colombian endemic coral snake species, Micrurus medemi and M. sangilensis, as well as Colombian populations of M. helleri. Electrophoresis and RP-HPLC techniques were used to identify venom components, and assays were conducted to detect enzyme activities, including phospholipase A2, hyaluronidase, and protease activities. The median lethal dose was determined using murine models. Cytotoxic activities in primary cultures from hippocampal neurons and cancer cell lines were evaluated. The venom profiles revealed similarities in electrophoretic separation among proteins under 20 kDa. The differences in chromatographic profiles were significant, mainly between the fractions containing medium-/large-sized and hydrophobic proteins; this was corroborated by a proteomic analysis which showed the expected composition of neurotoxins from the PLA2 (~38%) and 3FTx (~17%) families; however, a considerable quantity of metalloproteinases (~12%) was detected. PLA2 activity and protease activity were higher in M. helleri venom according to qualitative and quantitative assays. M. medemi venom had the highest lethality. All venoms decreased cell viability when tested on tumoral cell cultures, and M. helleri venom had the highest activity in neuronal primary culture. These preliminary studies shed light on the venoms of understudied coral snakes and broaden the range of sources that could be used for subsequent investigations of components with applications to specific diseases. Our findings also have implications for the clinical manifestations of snake envenoming and improvements in its medical management. Full article
(This article belongs to the Section Animal Venoms)
Show Figures

Figure 1

Figure 1
<p>Separation of <span class="html-italic">M. helleri, M. medemi,</span> and <span class="html-italic">M. sangilensis</span> venoms obtained using 15% SDS-PAGE under reducing conditions. Each lane was seeded with 20 μg of protein. Std.: molecular weight standard.</p>
Full article ">Figure 2
<p>Chromatographic profiles of whole venoms on a C18 column (Discovery<sup>®</sup>, 5 μm particle diameter; 250 × 4.6 mm) highlighting more abundant fractions (bold and underlined) subsequently observed in 12.5% SDS-PAGE under reducing conditions. <span class="html-italic">M. helleri</span> (<b>a</b>,<b>b</b>), <span class="html-italic">M. medemi</span> (<b>c</b>,<b>d</b>), and <span class="html-italic">M. sangilensis</span> (<b>e</b>,<b>f</b>). For analytical purposes, the chromatograms were divided into three sections, the limits of which are indicated by dotted lines. See the text for more detail. Std.: molecular weight standard (kDa).</p>
Full article ">Figure 2 Cont.
<p>Chromatographic profiles of whole venoms on a C18 column (Discovery<sup>®</sup>, 5 μm particle diameter; 250 × 4.6 mm) highlighting more abundant fractions (bold and underlined) subsequently observed in 12.5% SDS-PAGE under reducing conditions. <span class="html-italic">M. helleri</span> (<b>a</b>,<b>b</b>), <span class="html-italic">M. medemi</span> (<b>c</b>,<b>d</b>), and <span class="html-italic">M. sangilensis</span> (<b>e</b>,<b>f</b>). For analytical purposes, the chromatograms were divided into three sections, the limits of which are indicated by dotted lines. See the text for more detail. Std.: molecular weight standard (kDa).</p>
Full article ">Figure 3
<p>Percentages of the protein families in the venoms of (<b>a</b>) <span class="html-italic">M. helleri</span>, (<b>b</b>) <span class="html-italic">M. medemi,</span> and (<b>c</b>) <span class="html-italic">M. sangilensis</span>. Snake photos provided by Juan Pablo Hurtado-Gómez.</p>
Full article ">Figure 4
<p>Comparison of the number of shared and unique toxin-related proteins among the venoms of <span class="html-italic">M. helleri</span> (red; <span class="html-italic">n</span> = 53), <span class="html-italic">M. medemi</span> (yellow; <span class="html-italic">n</span> = 60), and <span class="html-italic">M. sangilensis</span> (light grey; <span class="html-italic">n</span> = 58). Note the high number of shared proteins between <span class="html-italic">M. medemi</span> and <span class="html-italic">M. sangilensis</span> and the higher number of unique peptides in <span class="html-italic">M. helleri</span> compared with <span class="html-italic">M. medemi</span> and <span class="html-italic">M. sangilensis</span>.</p>
Full article ">Figure 5
<p>Relative abundances of identified proteins using the normalized spectral abundance factor (NSAF). The color represents the protein families in this study. PLA<sub>2</sub>: phospholipases A<sub>2</sub>, 3FTx: three-finger toxins, SVMP: snake venom metalloproteinases, PLB: phospholipase B, LAAO: L-amino acid oxidases, SVSP: snake venom serine proteases, CTL: C-type lectin, HYA: hyaluronidases, VNGF: venom nerve growth factor, and MIN: minority compounds.</p>
Full article ">Figure 6
<p>Enzymatic activities for <span class="html-italic">M. helleri, M. medemi,</span> and <span class="html-italic">M. sangilensis</span> venoms. (<b>a</b>) Phospholipase A<sub>2</sub> assay in agarose and 10% egg yolk solution using 5 μg of each venom. The translucent halos formed around each well can be observed. Top: phosphate buffered saline (pH 7.4) used as the negative control (C−), and <span class="html-italic">Crotalus durissus cumanensis</span> venom used as the positive control (C+) (5 μg). Bottom: <span class="html-italic">M. helleri</span>, <span class="html-italic">M. medemi,</span> and <span class="html-italic">M. sangilensis</span>. (<b>b</b>) Determination of phospholipase A<sub>2</sub> activity by a colorimetric assay in a medium containing lecithin as the substrate (triplicate). Bars denote ± standard error. Panel (<b>c</b>) shows hyaluronidase activity. Non-stained areas depict positive activity. Std.: molecular weight standard. (<b>d</b>) Protease activity of the venoms obtained using an EnzChek<sup>®</sup> Protease Assay Kit. Trypsin was used as a positive control. Values represent the mean of three replicates. Bars denote standard error.</p>
Full article ">Figure 7
<p>Cell viability of three cell lines due to the effect of <span class="html-italic">Micrurus</span> venoms. (<b>a</b>) Hippocampal neuronal assay. Bar plots illustrate the percentage of cell viability in neuronal primary cultures after exposure to varying concentrations (2, 3, 12, 50, and 100 μg/mL) of the venoms. (<b>b</b>) Cell viability in the HTB-132 cell line assay. Bar plots illustrate the percentage of cell viability after exposure to varying concentrations (2, 3, 25, and 100 μg/mL) of the venoms. (<b>c</b>) Cell viability of the PC3 cell line assay. Bar plots illustrate the percentage of cell viability after exposure to varying concentrations (2, 3, 6, 13, 25, 50, and 100 μg/mL) of the venoms. Bars denote ± standard error. Statistical significance compared with the control is indicated by (*) symbols. <span class="html-italic">M. helleri</span> (orange), <span class="html-italic">M. medemi</span> (green), and <span class="html-italic">M. sangilensis</span> (red) venoms.</p>
Full article ">Figure 7 Cont.
<p>Cell viability of three cell lines due to the effect of <span class="html-italic">Micrurus</span> venoms. (<b>a</b>) Hippocampal neuronal assay. Bar plots illustrate the percentage of cell viability in neuronal primary cultures after exposure to varying concentrations (2, 3, 12, 50, and 100 μg/mL) of the venoms. (<b>b</b>) Cell viability in the HTB-132 cell line assay. Bar plots illustrate the percentage of cell viability after exposure to varying concentrations (2, 3, 25, and 100 μg/mL) of the venoms. (<b>c</b>) Cell viability of the PC3 cell line assay. Bar plots illustrate the percentage of cell viability after exposure to varying concentrations (2, 3, 6, 13, 25, 50, and 100 μg/mL) of the venoms. Bars denote ± standard error. Statistical significance compared with the control is indicated by (*) symbols. <span class="html-italic">M. helleri</span> (orange), <span class="html-italic">M. medemi</span> (green), and <span class="html-italic">M. sangilensis</span> (red) venoms.</p>
Full article ">
17 pages, 5255 KiB  
Article
Monoclonal-Based Antivenomics Reveals Conserved Neutralizing Epitopes in Type I PLA2 Molecules from Coral Snakes
by Carlos Corrêa-Netto, Marcelo A. Strauch, Marcos Monteiro-Machado, Ricardo Teixeira-Araújo, Juliana Guzzo Fonseca, Moema Leitão-Araújo, Maria Lúcia Machado-Alves, Libia Sanz, Juan J. Calvete, Paulo A. Melo and Russolina Benedeta Zingali
Toxins 2023, 15(1), 15; https://doi.org/10.3390/toxins15010015 - 26 Dec 2022
Cited by 1 | Viewed by 1929
Abstract
For over a century, polyclonal antibodies have been used to treat snakebite envenoming and are still considered by the WHO as the only scientifically validated treatment for snakebites. Nevertheless, moderate innovations have been introduced to this immunotherapy. New strategies and approaches to understanding [...] Read more.
For over a century, polyclonal antibodies have been used to treat snakebite envenoming and are still considered by the WHO as the only scientifically validated treatment for snakebites. Nevertheless, moderate innovations have been introduced to this immunotherapy. New strategies and approaches to understanding how antibodies recognize and neutralize snake toxins represent a challenge for next-generation antivenoms. The neurotoxic activity of Micrurus venom is mainly due to two distinct protein families, three-finger toxins (3FTx) and phospholipases A2 (PLA2). Structural conservation among protein family members may represent an opportunity to generate neutralizing monoclonal antibodies (mAbs) against family-conserved epitopes. In this work, we sought to produce a set of monoclonal antibodies against the most toxic components of M. altirostris venom. To this end, the crude venom was fractionated, and its major toxic proteins were identified and used to generate a panel of five mAbs. The specificity of these mAbs was characterized by ELISA and antivenomics approaches. Two of the generated mAbs recognized PLA2 epitopes. They inhibited PLA2 catalytic activity and showed paraspecific neutralization against the myotoxicity from the lethal effect of Micrurus and Naja venoms’ PLA2s. Epitope conservation among venom PLA2 molecules suggests the possibility of generating pan-PLA2 neutralizing antibodies. Full article
(This article belongs to the Special Issue Biotechnological Potential of Animal Venom and Toxins)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">M. altirostris</span> venom fractionation by gel filtration. The crude venom (20 mg) was submitted on TSK<sup>®</sup>-gel filtration. On the right side, an SDS-PAGE (15%) of the same venom was carried out under reducing conditions.</p>
Full article ">Figure 2
<p>Monoclonal-based antivenomics. Five monoclonal antibodies (mAb) were analyzed in these experiments. (<b>A</b>) shows the results to mAb 4B3, (<b>B</b>) to mAb 2G2, (<b>C</b>) to mAb 2B, (<b>D</b>) to mAb 3B2, (<b>E</b>) to mAb 1E8, and Panel (<b>F</b>) to the association of mAb 1E8 and 3B2. The RP-HPLC separation of <span class="html-italic">M. altirostris</span> venom (2 mg) is demonstrated in (<b>A</b>–<b>I</b>), and this chromatogram was used for comparison with (<b>B</b>–<b>F</b>) experiments. The non-bound fractions from monoclonal antibodies’ affinity columns were analyzed in A-II, B-II, C-II, D-II, E-II, and F-II. The reverse-phase separations of the monoclonal-immunocaptured proteins are demonstrated in A-III, B-III, C-III, D-III, and E-III. Red boxes indicate peaks retained by the antibodies. Green box indicate other PLA2 peaks not retained by 3B2 mAB.</p>
Full article ">Figure 2 Cont.
<p>Monoclonal-based antivenomics. Five monoclonal antibodies (mAb) were analyzed in these experiments. (<b>A</b>) shows the results to mAb 4B3, (<b>B</b>) to mAb 2G2, (<b>C</b>) to mAb 2B, (<b>D</b>) to mAb 3B2, (<b>E</b>) to mAb 1E8, and Panel (<b>F</b>) to the association of mAb 1E8 and 3B2. The RP-HPLC separation of <span class="html-italic">M. altirostris</span> venom (2 mg) is demonstrated in (<b>A</b>–<b>I</b>), and this chromatogram was used for comparison with (<b>B</b>–<b>F</b>) experiments. The non-bound fractions from monoclonal antibodies’ affinity columns were analyzed in A-II, B-II, C-II, D-II, E-II, and F-II. The reverse-phase separations of the monoclonal-immunocaptured proteins are demonstrated in A-III, B-III, C-III, D-III, and E-III. Red boxes indicate peaks retained by the antibodies. Green box indicate other PLA2 peaks not retained by 3B2 mAB.</p>
Full article ">Figure 3
<p>Effect of mAbs on the PLA<sub>2</sub> enzymatic activity. Effect of immunoglobulins from clones 1E8 and 3B2 toward phospholipase activity. (<b>A</b>) shows the absorbance variation promoted by <span class="html-italic">M. corallinus</span>, <span class="html-italic">M. altirostris</span>, and <span class="html-italic">Naja naja</span> snakes (1–10 µg/mL) in an egg yolk solution (<span class="html-italic">n</span> = 8). (<b>B</b>,<b>C</b>) display respectively, the inhibition of <span class="html-italic">M. altirostris</span> (3 µg/mL) and <span class="html-italic">N. naja</span> venom (3 µg/mL) phospholipase activity by two immunoglobulins from clones 1E8 and 3B2 alone or as a mixture. (<b>D</b>), the change in the absorbance of egg yolk solution promoted by pooled PLA2 (P4) is shown. (<b>E</b>) displays the inhibition of pooled PLA<sub>2</sub> (P4—0.3 µg/mL) by the immunoglobulins from clones 1E8 and 3B2 alone or combined. Data express mean ± SEM. (<span class="html-italic">n</span> = 8) * <span class="html-italic">p</span> &lt; 0.05 vs. venom (<b>B</b>,<b>C</b>) or pooled PLA<sub>2</sub> (<b>E</b>). Filled squares represented on (<b>B</b>,<b>C</b>,<b>E</b>) indicate that each mAb was combined according to the concentration indicated in the abscissa axis.</p>
Full article ">Figure 4
<p>Dot blot of the <span class="html-italic">M. altirostris</span> venom against the monoclonal antibodies 1E8 (<b>A</b>,<b>B</b>) and 3B2 (<b>C</b>,<b>D</b>). In (<b>A</b>,<b>C</b>), native venom, (<b>B</b>,<b>D</b>) denatured venom: boiled for 5 min with 5% 2-mercaptoethanol.</p>
Full article ">Figure 5
<p>Effect of mAbs on myotoxicity. (<b>A</b>) shows the myotoxic effect of <span class="html-italic">M. altirostris</span> and pooled PLA<sub>2</sub> (P4) in mice (<span class="html-italic">n</span> = 4–8). (<b>B</b>) the antagonism of the higher dose of immunoglobulins from clones 1E8 and 3B2 against pooled PLA<sub>2</sub> (<span class="html-italic">n</span> = 4). (<b>C</b>) displays the antagonism of this same dose against <span class="html-italic">N. naja</span> and <span class="html-italic">M. altirostris</span> crude venoms. (<span class="html-italic">n</span> = 4). Data express mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05 vs. PSS (<b>A</b>) and * <span class="html-italic">p</span> &lt; 0.05 vs. related venom in (<b>B</b>,<b>C</b>).</p>
Full article ">Figure 6
<p>Survival curve. Mice were injected with pooled PLA<sub>2</sub> alone or premixed for 15 min with immunoglobulins from clones 1E8 and 3B2 (i.p.) 12 mg/kg (<b>A</b>) or 24 mg/kg (<b>B</b>), and survival was observed for 24 h (<span class="html-italic">n</span> = 6–7). Data express the percentage of live mice, and the <span class="html-italic">p</span>-value represents the Mantel–Cox test analysis.</p>
Full article ">
15 pages, 2603 KiB  
Article
First Insights into the Venom Composition of Two Ecuadorian Coral Snakes
by Josselin A. Hernández-Altamirano, David Salazar-Valenzuela, Evencio J. Medina-Villamizar, Diego R. Quirola, Ketan Patel, Sakthivel Vaiyapuri, Bruno Lomonte and José R. Almeida
Int. J. Mol. Sci. 2022, 23(23), 14686; https://doi.org/10.3390/ijms232314686 - 24 Nov 2022
Cited by 3 | Viewed by 3225
Abstract
Micrurus is a medically relevant genus of venomous snakes composed of 85 species. Bites caused by coral snakes are rare, but they are usually associated with very severe and life-threatening clinical manifestations. Ecuador is a highly biodiverse country with a complex natural environment, [...] Read more.
Micrurus is a medically relevant genus of venomous snakes composed of 85 species. Bites caused by coral snakes are rare, but they are usually associated with very severe and life-threatening clinical manifestations. Ecuador is a highly biodiverse country with a complex natural environment, which is home to approximately 20% of identified Micrurus species. Additionally, it is on the list of Latin American countries with the highest number of snakebites. However, there is no local antivenom available against the Ecuadorian snake venoms, and the biochemistry of these venoms has been poorly explored. Only a limited number of samples collected in the country from the Viperidae family were recently characterised. Therefore, this study addressed the compositional patterns of two coral snake venoms from Ecuador, M. helleri and M. mipartitus, using venomics strategies, integrating sample fractionation, gel electrophoresis, and mass spectrometry. Chromatographic and electrophoretic profiles of these snake venoms revealed interspecific variability, which was ascertained by mass spectrometry. The two venoms followed the recently recognised dichotomic toxin expression trends displayed by Micrurus species: M. helleri venom contains a high proportion (72%) of phospholipase A2, whereas M. mipartitus venom is dominated by three-finger toxins (63%). A few additional protein families were also detected in these venoms. Overall, these results provide the first comprehensive views on the composition of two Ecuadorian coral snake venoms and expand the knowledge of Micrurus venom phenotypes. These findings open novel perspectives to further research the functional aspects of these biological cocktails of PLA2s and 3FTxs and stress the need for the preclinical evaluation of the currently used antivenoms for therapeutic purposes in Ecuador. Full article
(This article belongs to the Special Issue Pharmacological Insights of Venoms)
Show Figures

Figure 1

Figure 1
<p>The current knowledge about <span class="html-italic">Micrurus</span> venom in the literature. (<b>A</b>) <span class="html-italic">Micrurus</span> specimens collected from different countries such as Argentina, Brazil, Colombia, Costa Rica, Mexico, and United States were proteomically characterised (dots represent geographical sites where analysed venoms/specimens were collected). However, the composition of Ecuadorian coral venom is still unknown. The question mark (yellow) points out this gap in the understanding of coral snake venoms in a biodiverse region. Red dots represent specimens with a PLA<sub>2</sub>-predominant profile, while blue dots highlight venoms with a higher abundance of 3FTXs. Dual-colour dots illustrate regions where both compositional patterns have been identified. (<b>B</b>) Detailed profile of key venom toxins in Elapid snake venoms. PLA<sub>2</sub>-dominant or 3FTx-dominant profiles dictate an intriguing general trend in the distribution of proteins in venoms along the American continent.</p>
Full article ">Figure 2
<p>Venom collection and species occurrences of Ecuadorian coral snakes. (<b>A</b>) Sample from <span class="html-italic">M. helleri</span> (red) was extracted from a specimen found in Nangaritza (Zamora Chinchipe), while venom from <span class="html-italic">M. mipartitus</span> (blue) was obtained from a snake found in Quinindé, Jevon Forest Biological Station (Esmeraldas). (<b>B</b>) Spatial distribution of these two <span class="html-italic">Micrurus</span> species in Ecuador. This map was constructed using geospatial data from (GBIF, <a href="http://www.gbif.org" target="_blank">http://www.gbif.org</a>) and VertNet (<a href="http://www.vertnet.org" target="_blank">http://www.vertnet.org</a>). Both biodiversity-related repositories were accessed on 18 September 2022. (<b>C</b>) <span class="html-italic">M. helleri</span> (photo by María José Quiroz; reproduced here from <a href="https://bioweb.bio" target="_blank">https://bioweb.bio</a>, accessed on 1 November 2022, under a CC BY-NC-ND 4.0 License) and (<b>D</b>) <span class="html-italic">M. mipartitus</span> specimens (photo by Diego R. Quirola).</p>
Full article ">Figure 3
<p>Fractionation of the Ecuadorian coral venoms of (<b>A</b>) <span class="html-italic">M. helleri</span> and (<b>B</b>) <span class="html-italic">M. mipartitus</span> by RP-HPLC. The peaks were identified according to their elution through a C18 column for further characterisation by electrophoresis and mass spectrometry analyses. The separation of venom toxins was performed in the gradient (red-dashed lines) elution mode described in the methodology section.</p>
Full article ">Figure 4
<p>Electrophoretic profiles of Ecuadorian coral snake venom fractions of (<b>A</b>) <span class="html-italic">Micrurus helleri</span> and (<b>B</b>) <span class="html-italic">Micrurus mipartitus</span>. Low-molecular mass proteins (5–17 KDa) are highly abundant in these two samples of Ecuadorian elapids. The last chromatographic fractions contain higher molecular mass toxins, around 60 KDa.</p>
Full article ">Figure 5
<p>Proteomic insights into two Ecuadorian coral snake venoms. The pie charts represent the relative abundance of protein families found in the venom proteome of (<b>A</b>) <span class="html-italic">M. helleri</span> and (<b>B</b>) <span class="html-italic">M. mipartitus</span>. The relative abundance of protein families is expressed as % of the total venom toxin content, estimated based on chromatographic peak areas. UNK: unknown/unidentified.</p>
Full article ">
19 pages, 3256 KiB  
Article
Heterologous Expression and Immunogenic Potential of the Most Abundant Phospholipase A2 from Coral Snake Micrurus dumerilii to Develop Antivenoms
by Luz E. Romero-Giraldo, Sergio Pulido, Mario A. Berrío, María F. Flórez, Paola Rey-Suárez, Vitelbina Nuñez and Jaime A. Pereañez
Toxins 2022, 14(12), 825; https://doi.org/10.3390/toxins14120825 - 24 Nov 2022
Cited by 3 | Viewed by 2019
Abstract
Micrurus dumerilii is a coral snake of clinic interest in Colombia. Its venom is mainly composed of phospholipases A2 being MdumPLA2 the most abundant protein. Nevertheless, Micrurus species produce a low quantity of venom, which makes it difficult to produce anticoral [...] Read more.
Micrurus dumerilii is a coral snake of clinic interest in Colombia. Its venom is mainly composed of phospholipases A2 being MdumPLA2 the most abundant protein. Nevertheless, Micrurus species produce a low quantity of venom, which makes it difficult to produce anticoral antivenoms. Therefore, in this work, we present the recombinant expression of MdumPLA2 to evaluate its biological activities and its immunogenic potential to produce antivenoms. For this, a genetic construct rMdumPLA2 was cloned into the pET28a vector and expressed heterologously in bacteria. His-rMdumPLA2 was extracted from inclusion bodies, refolded in vitro, and isolated using affinity and RP-HPLC chromatography. His-rMdumPLA2 was shown to have phospholipase A2 activity, a weak anticoagulant effect, and induced myonecrosis and edema. The anti-His-rMdumPLA2 antibodies produced in rabbits recognized native PLA2, the complete venom of M. dumerilii, and a phospholipase from another species of the Micrurus genus. Antibodies neutralized 100% of the in vitro phospholipase activity of the recombinant toxin and a moderate percentage of the myotoxic activity of M. dumerilii venom in mice. These results indicate that His-rMdumPLA2 could be used as an immunogen to improve anticoral antivenoms development. This work is the first report of an M. dumerilii functional recombinant PLA2. Full article
(This article belongs to the Section Animal Venoms)
Show Figures

Figure 1

Figure 1
<p>Plasmid carrying His-rMdumPLA<sub>2</sub>. (<b>A</b>) Genetic construction His-rMdumPLA<sub>2</sub>. (<b>B</b>) pET28a digestion test with XhoI and EcoRV in 1% agarose gel-stained ethidium bromide (Sigma, Saint Louis, MO, USA). MW: molecular weight marker (1 kb Plus DNA Ladder) (NEB); (1): His-rMdumPLA<sub>2</sub> clone 1; (2): His-rMdumPLA<sub>2</sub> clone 2; (3) His-rMdumPLA<sub>2</sub> clone 3. The arrow indicates the fragment of 1749 bp with the expected size: target insert (451 bp) plus a fragment of 1411 bp (EcoRV-XhoI, coordinates 3797–5207) without 113 bp released in the cut. EcoRV cuts after position 158, and XhoI cuts after position 1573 on pET28a.</p>
Full article ">Figure 2
<p>Expression of His-rMdumPLA<sub>2</sub> in <span class="html-italic">E. coli</span> BL21 (DE3). A total of 10 µg of each sample were loaded in 14% Tris-Tricine SDS-PAGE gel and Coomassie G-250 stain. His-rMdumPLA<sub>2</sub> molecular mass is about 17 kDa (arrow), according to molecular weight marker (11–250 kDa) (New England Biolabs, Ipswich, MA, USA). (<b>A</b>) His-rMdumPLA<sub>2</sub> expression at three different times. MW: molecular weight markers (11–250 kDa), 0, 2 and 8 h. MW: molecular weight marker; T: total protein; S: soluble; I: insoluble. (<b>B</b>) Refolding and isolation by affinity chromatography. MW: molecular weight marker; IBs: solubilized inclusion bodies; FP: folded protein; MW: molecular weight marker; FT: Flowthrough; W1: Wash 1; W2: Wash 2; E: Elution. (<b>C</b>) TEV cut in 14% Tris-Tricine SDS-PAGE gel and Coomassie G-250 stain. A total of 10 µg of rMdumPLA<sub>2</sub> was loaded and detected as a band of approximately 14 kDa. (<b>D</b>) Western blot using the anti-MdumPLA<sub>2</sub> native antibody [1:100] coupled to peroxidase, obtained from rabbit inoculated with MdumPLA<sub>2</sub> isolated from <span class="html-italic">M. dumerilii</span> venom by RP-HPLC. MdumPLA<sub>2</sub> molecular mass is about 13 kDa (arrow in the left panel) and His-rMdumPLA<sub>2</sub> molecular mass is about 17 kDa (arrow in the right panel). Precision Plus Protein Kaleidoscope Standard (10–250 kDa) (Sigma, Saint Louis, MO, USA) was used as molecular weight marker.</p>
Full article ">Figure 3
<p>His-rMdumPLA<sub>2</sub> purification by RP-HPLC chromatography. His-rMdumPLA<sub>2</sub> purification by RP-HPLC chromatography on a C18 column (250 × 10 mm) eluted at 1 mL/min with an acetonitrile linear gradient. His-rMdumPLA<sub>2</sub> was collected in the peak that eluted at 18.2 min.</p>
Full article ">Figure 4
<p>Biological activities of His-rMdumPLA<sub>2</sub>. (<b>A</b>) PLA<sub>2</sub> activity by hydrolysis of substrate 4-NOBA. (<b>B</b>) PLA<sub>2</sub> activity by indirect hemolysis. The bars show the diameter of the hemolytic halo in millimeters. The diameter of the hemolytic halo of negative control (PBS) was zero. (<b>C</b>) Myotoxic activity. The CK activity was determined from mouse plasma by kinetic assay, and the absorbance was recorded at 340 nm. (<b>D</b>) Edema-inducing activity. Edema was estimated by the percentage increment in the weight of the footpad with respect to the negative control (SS). * Indicates statistically significant differences with corresponding negative control (<span class="html-italic">p</span> &lt; 0.05). The data show the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>The titration curve of average antibodies in serum from four bleedings by ELISA against His-rMdumPLA<sub>2</sub>. A 96-well plate was coated with His-rMdumPLA<sub>2,</sub> and serum from four bleedings was used in dilutions from 1:10 to 1:10,000. A peroxidase-labeled anti-rabbit IgG conjugate detected bound antibodies. The absorbance of antibodies anti-His-rMdumPLA<sub>2</sub> is shown in the function of the dilution. * Indicates statistically significant differences with the pre-immune serum. The data show the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 6
<p>Immunoreactivity of serum and IgG anti-His-rMdumPLA<sub>2</sub> against <span class="html-italic">M. dumerilii</span> venom (V-Mdum), His-rMdumPLA<sub>2</sub>, MdumPLA<sub>2</sub>, the native 24 and 25 fractions (F-24, F-25) close to MdumPLA<sub>2</sub> in the PLA<sub>2</sub> region from <span class="html-italic">M. dumerilii</span> proteome, and MmipPLA<sub>2</sub> by ELISA. * Indicates statistically significant differences with the pre-immune serum (<span class="html-italic">p</span> &lt; 0.0001). The data show the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 7
<p>Neutralization of the myotoxic activity of <span class="html-italic">M. dumerilii</span> venom by anti-His-rMdumPLA<sub>2</sub>-IgG. 3 mg IgG anti-His-rMdumPLA<sub>2</sub> mixed with 5 µg <span class="html-italic">M. dumerilii</span> venom was injected by the intramuscular route in mice, and CK activity was determined in serum. * Indicates statistically significant differences with anti-His-rMdumPLA<sub>2</sub>-IgG (<span class="html-italic">p</span> = 0.0005). The data show the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">
13 pages, 1379 KiB  
Article
Recovery from the Neuroparalysis Caused by the Micrurus nigrocinctus Venom Is Accelerated by an Agonist of the CXCR4 Receptor
by Marco Stazi, Federico Fabris, Julián Fernández, Giorgia D'Este, Michela Rigoni, Aram Megighian, José María Gutiérrez, Bruno Lomonte and Cesare Montecucco
Toxins 2022, 14(8), 531; https://doi.org/10.3390/toxins14080531 - 2 Aug 2022
Cited by 4 | Viewed by 1936
Abstract
Snake envenoming is a major but neglected human disease in tropical and subtropical regions. Among venomous snakes in the Americas, coral snakes of the genus Micrurus are particularly dangerous because they cause a peripheral neuroparalysis that can persist for many days or, in [...] Read more.
Snake envenoming is a major but neglected human disease in tropical and subtropical regions. Among venomous snakes in the Americas, coral snakes of the genus Micrurus are particularly dangerous because they cause a peripheral neuroparalysis that can persist for many days or, in severe cases, progress to death. Ventilatory support and the use of snake species-specific antivenoms may prevent death from respiratory paralysis in most cases. However, there is a general consensus that additional and non-expensive treatments that can be delivered even long after the snake bite are needed. Neurotoxic degeneration of peripheral motor neurons activates pro-regenerative intercellular signaling programs, the greatest of which consist of the chemokine CXCL12α, produced by perisynaptic Schwann cells, which act on the CXCR4 receptor expressed on damaged neuronal axons. We recently found that the CXCR4 agonist NUCC-390 promotes axonal growth. Here, we show that the venom of the highly neurotoxic snake Micrurus nigrocinctus causes a complete degeneration of motor axon terminals of the soleus muscle, followed by functional regeneration whose time course is greatly accelerated by NUCC-390. These results suggest that NUCC-390 is a potential candidate for treating human patients envenomed by Micrurus nigrocinctus as well as other neurotoxic Micrurus spp. in order to improve the recovery of normal neuromuscular physiology, thus reducing the mortality and hospital costs of envenoming. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">M. nigrocinctus</span> venom injection in the soleus muscle induced an acute and reversible degeneration of MAT at the NMJ. The degeneration induced by the injection of <span class="html-italic">M. nigrocintus</span> total venom (6 µg/kg) was monitored by immunostaining. The motor axon terminal (MAT) is identified by VAMP-1 immunostaining (<span class="html-italic">green</span>), post-synaptic nAChRs is identified by fluorescent α-BTx (<span class="html-italic">red</span>), and the axon terminal is identified by neurofilament (NF) staining (<span class="html-italic">white</span>). Scale bars: 50 µm. The pie chart shows the extent of neurodegeneration (<span class="html-italic">red</span> part) compared with intact NMJ (<span class="html-italic">green</span>), followed by their respective quantification and statistics (paired <span class="html-italic">t</span> test, <span class="html-italic">n</span>= 4). *** <span class="html-italic">p</span> &lt; 0.001. **** <span class="html-italic">p</span> &lt; 0.0001, ns (not significant).</p>
Full article ">Figure 2
<p>CXCR4 receptor is expressed in neuronal axons after <span class="html-italic">M. nigrocinctus</span> venom injection. CXCR4 staining (<span class="html-italic">red</span>) at soleus NMJs in controls (upper panels) and after <span class="html-italic">M. nigrocintus</span> venom injection (lower panels). Schwann cells (SCs) were GFP-positive (<span class="html-italic">cyan</span>), and the axon terminal was identified by NF staining (<span class="html-italic">green</span>). Scale bars: 10 µm.</p>
Full article ">Figure 3
<p>NUCC-390 promoted NMJ recovery of function after <span class="html-italic">M. nigrocinctus</span> venom intoxication. (<b>A</b>) Schematic representation of the technique of measurement of the evoked junctional potentials (EJPs). (<b>B</b>) Temporal scheme of the administration of <span class="html-italic">M. nigrocinctus</span> venom and NUCC-390 administration. (<b>C</b>) EJPs of soleus muscles 96 h post injection of <span class="html-italic">M. nigrocinctus</span> venom in the hind limb with or without daily NUCC-390 administrations. Each bar represents the mean of the EJP amplitude ± SEM from <span class="html-italic">n</span> = 4 (number of analyzed fibers = 12, one-way ANOVA with Tukey’s multiple comparison test; ns = not significant). * <span class="html-italic">p</span> &lt; 0.05. (<b>D</b>) Representative immunostaining of NMJs performed on the same muscles used for EJP measurements. (<b>E</b>) Quantification of NMJs from (<b>D</b>) (paired <span class="html-italic">t</span> test, <span class="html-italic">n</span> = 4). ** <span class="html-italic">p</span> &lt; 0.01. MAT is identified by VAMP-1 immunostaining (<span class="html-italic">green</span>), post-synaptic nAChRs are identified by fluorescent α-BTx (<span class="html-italic">red</span>), and the axon terminal is identified by NF staining (<span class="html-italic">white</span>). White asterisks indicate still-degenerated NMJs. Scale bars: 50 µm.</p>
Full article ">Figure 4
<p>NUCC-390 increased the recovery of compound muscle action potential from paralysis induced by <span class="html-italic">M. nigrocinctus</span> venom. (<b>A</b>) Temporal scheme of the intoxication of the soleus muscle induced by local injection of <span class="html-italic">M. nigrocinctus</span> venom with or without NUCC-390’s daily administration. (<b>B</b>) CMAP values recorded for gastrocnemius muscles 96 h after envenomation with <span class="html-italic">M. nigrocinctus</span> venom with or without NUCC-390’s daily local administration. Data are expressed as a ratio between envenomed and control animals CMAP amplitude ± SEM (<span class="html-italic">n</span> = 5, one-way ANOVA with Tukey’s multiple comparison test; ns = not significant) * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) Representative traces of CMAP analysis of control animals (<b>left</b>) and 96 h after envenomation (<b>right</b>). (<b>D</b>) Representative immunostaining of intoxicated NMJs performed on the same muscles used for CMAP analysis. Motor neuron axon terminals are identified by VAMP-1 immunostaining in <span class="html-italic">green</span>, post-synaptic nAChRs are identified by fluorescent α-BTx (<span class="html-italic">red</span>), and the axon is identified by NF staining in <span class="html-italic">white</span>. Regenerated NMJs are indicated by white asterisks. Scale bars: 50 μm.</p>
Full article ">Figure 5
<p>NUCC-390 increased lung ventilation recovery in mice intoxicated with <span class="html-italic">M. nigrocinctus</span> venom. (<b>A</b>) Representative traces of mediastinum pressure variations in untreated mice (t0) or 24 h, 96 h, or 168 h after <span class="html-italic">M. nigrocinctus</span> administration (6 μg/kg) with NUCC-390 or vehicle i.p. daily injection (3.2 mg/kg in 40 μL physiological solution containing 0.2% gelatin). (<b>B</b>) The inferred ventilation index (IVI) was estimated as measurement of the peak area of 20 consecutive events of the traces (±SEM) obtained from a group of mice envenomed with <span class="html-italic">M. nigrocinctus</span> venom with (orange trace) or without (black trace) NUCC-390 daily administration (one-way ANOVA with Tukey’s multiple comparison test, <span class="html-italic">n</span> = 4). ** <span class="html-italic">p</span> &lt; 0.01. **** <span class="html-italic">p</span> &lt; 0.0001. (<b>C</b>) Representative immunostaining of envenomed NMJ 96 h after envenoming at the level of the diaphragm muscle daily injected i.p. with either vehicle (upper panels) or NUCC-390 (lower panels). Motor axon terminals are identified by VAMP-1 immunostaining (<span class="html-italic">green</span>), post-synaptic muscle membrane is identified by nAChR staining with fluorescent α-BTx (<span class="html-italic">red</span>), and axon is identified by NF staining (<span class="html-italic">white</span>). Scale bars: 50 μm. (<b>D</b>) Quantification of recovered NMJ stained in panel C (paired <span class="html-italic">t</span> test, <span class="html-italic">n</span> = 4). *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
15 pages, 1071 KiB  
Article
Divergent Specialization of Simple Venom Gene Profiles among Rear-Fanged Snake Genera (Helicops and Leptodeira, Dipsadinae, Colubridae)
by Peter A. Cerda, Jenna M. Crowe-Riddell, Deise J. P. Gonçalves, Drew A. Larson, Thomas F. Duda and Alison R. Davis Rabosky
Toxins 2022, 14(7), 489; https://doi.org/10.3390/toxins14070489 - 15 Jul 2022
Cited by 6 | Viewed by 2733
Abstract
Many venomous animals express toxins that show extraordinary levels of variation both within and among species. In snakes, most studies of venom variation focus on front-fanged species in the families Viperidae and Elapidae, even though rear-fanged snakes in other families vary along the [...] Read more.
Many venomous animals express toxins that show extraordinary levels of variation both within and among species. In snakes, most studies of venom variation focus on front-fanged species in the families Viperidae and Elapidae, even though rear-fanged snakes in other families vary along the same ecological axes important to venom evolution. Here we characterized venom gland transcriptomes from 19 snakes across two dipsadine rear-fanged genera (Leptodeira and Helicops, Colubridae) and two front-fanged genera (Bothrops, Viperidae; Micrurus, Elapidae). We compared patterns of composition, variation, and diversity in venom transcripts within and among all four genera. Venom gland transcriptomes of rear-fanged Helicops and Leptodeira and front-fanged Micrurus are each dominated by expression of single toxin families (C-type lectins, snake venom metalloproteinase, and phospholipase A2, respectively), unlike highly diverse front-fanged Bothrops venoms. In addition, expression patterns of congeners are much more similar to each other than they are to species from other genera. These results illustrate the repeatability of simple venom profiles in rear-fanged snakes and the potential for relatively constrained venom composition within genera. Full article
(This article belongs to the Section Animal Venoms)
Show Figures

Figure 1

Figure 1
<p>Representative venom delivery system for each genus and venom transcript counts of individuals. (<b>A</b>) Soft-tissue scan of the venom delivery system for each genus mapped onto a simplified phylogeny developed in this study. (<b>B</b>) Isolated maxilla bone from each representative genus showing position of fangs (arrow). (<b>C</b>) Unique toxin transcripts recovered from each <span class="html-italic">Helicops</span> individual. (<b>D</b>) Unique toxin transcripts recovered from each <span class="html-italic">Leptodeira</span> individual. (<b>E</b>) Unique toxin transcripts recovered from each <span class="html-italic">Micrurus</span> individual. (<b>F</b>) Unique toxin transcripts recovered from each <span class="html-italic">Bothrops</span> individual. MUSM = Museo de Historia Natural, Universidad Nacional Mayor de San Marco, UMMZ = University of Michigan Museum of Zoology, SVMP = snake venom metalloproteinase, PLA2 = phospholipase A2, 3FTx = three-finger toxin, CTL = C-type lectin, SVSP = snake venom serine proteinase, BPP = bradykinin-potentiating peptides, LAAO = L-amino acid oxidase, CRiSP = cystine rich secretory protein, Kunitz = Kunitz-type serine protease. Micro-CT scans of specimens vouchered in the University of Michigan Museum of Zoology (UMMZ) and Museo de Historia Natural de la Universidad Nacional Major de San Marcos (MUSM). We deposited these micro-CT scans used for venom and fang morphology for public access in the Morphosource ‘Scan All Snakes’ Project ID 00000C374 (<a href="https://www.morphosource.org/Detail/ProjectDetail/Show/project_id/374" target="_blank">https://www.morphosource.org/Detail/ProjectDetail/Show/project_id/374</a>).</p>
Full article ">Figure 2
<p>Expression of toxin gene families from venom gland transcriptomes (<b>center</b>) and overall venom transcriptome diversity (<b>right</b>) mapped to phylogeny inferred from mitochondrial gene sequences (<b>left</b>). Note that these data suggest there is generally higher similarity in venom profiles among individuals within genera than among genera for both metrics. <span class="html-italic">Bothrops</span> and <span class="html-italic">Micrurus</span> species are front-fanged, while <span class="html-italic">Leptodeira</span> and <span class="html-italic">Helicops</span> species are rear-fanged. SVMP = snake venom metalloproteinase, PLA2 = phospholipase A2, 3FTx = three-finger toxin, CTL = C-type lectin, SVSP = snake venom serine proteinase, BPP = bradykinin-potentiating peptides, LAAO = L-amino acid oxidase, CRiSP = cystine-rich secretory protein, Kunitz = Kunitz-type serine protease.</p>
Full article ">
18 pages, 3398 KiB  
Article
Pharmacological Screening of Venoms from Five Brazilian Micrurus Species on Different Ion Channels
by Jessica Matos Kleiz-Ferreira, Hans Bernaerts, Ernesto Lopes Pinheiro-Junior, Steve Peigneur, Russolina Benedeta Zingali and Jan Tytgat
Int. J. Mol. Sci. 2022, 23(14), 7714; https://doi.org/10.3390/ijms23147714 - 13 Jul 2022
Cited by 4 | Viewed by 2394
Abstract
Coral snake venoms from the Micrurus genus are a natural library of components with multiple targets, yet are poorly explored. In Brazil, 34 Micrurus species are currently described, and just a few have been investigated for their venom activities. Micrurus venoms are composed [...] Read more.
Coral snake venoms from the Micrurus genus are a natural library of components with multiple targets, yet are poorly explored. In Brazil, 34 Micrurus species are currently described, and just a few have been investigated for their venom activities. Micrurus venoms are composed mainly of phospholipases A2 and three-finger toxins, which are responsible for neuromuscular blockade—the main envenomation outcome in humans. Beyond these two major toxin families, minor components are also important for the global venom activity, including Kunitz-peptides, serine proteases, 5′ nucleotidases, among others. In the present study, we used the two-microelectrode voltage clamp technique to explore the crude venom activities of five different Micrurus species from the south and southeast of Brazil: M. altirostris, M. corallinus, M. frontalis, M. carvalhoi and M. decoratus. All five venoms induced full inhibition of the muscle-type α1β1δε nAChR with different levels of reversibility. We found M. altirostris and M. frontalis venoms acting as partial inhibitors of the neuronal-type α7 nAChR with an interesting subsequent potentiation after one washout. We discovered that M. altirostris and M. corallinus venoms modulate the α1β2 GABAAR. Interestingly, the screening on KV1.3 showed that all five Micrurus venoms act as inhibitors, being totally reversible after the washout. Since this activity seems to be conserved among different species, we hypothesized that the Micrurus venoms may rely on potassium channel inhibitory activity as an important feature of their envenomation strategy. Finally, tests on NaV1.2 and NaV1.4 showed that these channels do not seem to be targeted by Micrurus venoms. In summary, the venoms tested are multifunctional, each of them acting on at least two different types of targets. Full article
(This article belongs to the Special Issue Venoms and Ion Channels 2.0)
Show Figures

Figure 1

Figure 1
<p>Geographic distribution of the five <span class="html-italic">Micrurus</span> species. The colored areas represent the current consolidated geographic distribution of the <span class="html-italic">Micrurus</span> species, and the dotted colored lines represent the distribution areas under scientific revision. The areas where the specimens used in this work were collected are indicated in the map with the names of the species.</p>
Full article ">Figure 2
<p>Crude venom chromatographic profiles. (<b>A</b>) <span class="html-italic">Micrurus altirostris</span>, (<b>B</b>) <span class="html-italic">Micrurus corallinus</span>, (<b>C</b>) <span class="html-italic">Micrurus frontalis</span>, (<b>D</b>) <span class="html-italic">Micrurus carvalhoi</span> and (<b>E</b>) <span class="html-italic">Micrurus decoratus</span> venoms. A total of 500 µg of crude venom was used to obtain the RP-HPLC profile. Based on the published proteome information of <span class="html-italic">Micrurus</span> species venoms [<a href="#B8-ijms-23-07714" class="html-bibr">8</a>,<a href="#B9-ijms-23-07714" class="html-bibr">9</a>,<a href="#B10-ijms-23-07714" class="html-bibr">10</a>], we show in the figure the approximate elution regions of 3FTxs and PLA<sub>2</sub>s, marked by dotted lines in red and green, respectively. The right superior panel in each figure shows the full area of the chromatogram, where the blue box represents the zoomed area and the purple line represents the acetonitrile (ACN) gradient.</p>
Full article ">Figure 3
<p>Cytotoxic evaluation of crude venoms on <span class="html-italic">X. laevis</span> non-injected oocytes. (<b>A</b>) <span class="html-italic">Micrurus altirostris</span>, <span class="html-italic">Micrurus corallinus</span>, <span class="html-italic">Micrurus carvalhoi</span> and <span class="html-italic">Micrurus frontalis</span> venom activity. (<b>B</b>) Positive control-<span class="html-italic">Apis mellifera</span> venom activity. The black lines show the control, and the red lines show the venom activity. In panel (<b>A</b>), the dotted dark gray and light gray lines show 100 s and 200 s of past time from the venom activity, respectively. The blue box shows a zoom of the graph on panel (<b>B</b>). The graphs show the trace of one experiment from a series of independent experiments. (<b>C</b>) Inward and outward current graph. (<b>D</b>) Reversal potential graph.</p>
Full article ">Figure 4
<p><span class="html-italic">Micrurus frontalis</span> venom causes a Cl<sup>−</sup> outward current by acting directly or indirectly on Ca<sup>2+</sup>-activated Cl<sup>−</sup> channels in <span class="html-italic">X. laevis</span> oocytes. The black line shows the control (the oocyte perfused with ND96); the red line shows the activity of 50 µg of <span class="html-italic">M. frontalis</span> venom on oocytes; the blue line shows the oocyte current in the presence of 100 µM of NA; and the green line shows the activity of 50 µg of <span class="html-italic">M. frontalis</span> venom after cell incubation with 100 µM of NA, where the outward current evoked by the venom is no longer observed. The blue box on the right of the panel shows a zoom of the graph. The graphs show the trace of one experiment from a series of independent experiments.</p>
Full article ">Figure 5
<p>Screening of crude venoms from different <span class="html-italic">Micrurus</span> species in different ligand-gated ion channels. (<b>A</b>) Muscle-type α1β1δε nAChR, (<b>B</b>) Neuronal-type α7 nAChR and (<b>C</b>) α1β2 GABA<sub>A</sub>R. The black lines represent the control pulses, the red lines represent the venom incubation, the blue lines represent the receptor activity after the incubation with 10 µg of crude venom, and the gray lines represent the agonist pulse after the cell washout. The triangles mark the point where the application of the agonist was started. The <span class="html-italic">n</span> from each experiment is shown as well as the current/time reference bar. The graphs show the trace of one experiment from a series of independent experiments, with the exception of tests with <span class="html-italic">n</span> = 1.</p>
Full article ">Figure 6
<p>Screening of crude venoms from different <span class="html-italic">Micrurus</span> species in different voltage-gated ion channels. (<b>A</b>) K<sub>V</sub>1.3. The inhibition percentage and the SEM are shown in blue for each graph. (<b>B</b>) Na<sub>V</sub>1.2. The black lines represent the control and the blue lines represent the channel response after incubation with 10 µg of crude venom. The dotted lines represent the zero current level. The <span class="html-italic">n</span> from each experiment is shown as well as the current/time reference bar. The graphs show the trace of one experiment from a series of independent experiments, with the exception of tests with <span class="html-italic">n</span> = 1.</p>
Full article ">
10 pages, 3379 KiB  
Article
Anti-Neurotoxins from Micrurus mipartitus in the Development of Coral Snake Antivenoms
by Ana Cardona-Ruda, Paola Rey-Suárez and Vitelbina Núñez
Toxins 2022, 14(4), 265; https://doi.org/10.3390/toxins14040265 - 9 Apr 2022
Cited by 7 | Viewed by 3459
Abstract
In Colombia, the genus Micrurus includes 30 species, of which M. mipartitus and M. dumerilii are the most widely distributed. Micrurus causes less than 3% of the approximately 5000 cases of snakebite per year. The elapid envenomation caused by the snakes from the [...] Read more.
In Colombia, the genus Micrurus includes 30 species, of which M. mipartitus and M. dumerilii are the most widely distributed. Micrurus causes less than 3% of the approximately 5000 cases of snakebite per year. The elapid envenomation caused by the snakes from the Micrurus genus, are characterized by the severity of their clinical manifestations, due to the venom neurotoxic components such as three-finger toxins (3FTx) and phospholipases (PLA2). The treatment for snakebites is the administration of specific antivenoms, however, some of them have limitations in their neutralizing ability. A strategy proposed to improve antivenoms is to produce antibodies against the main components of the venom. The aim of this work was to produce an antivenom, using an immunization protocol including the main 3FTx and PLA2 responsible for M. mipartitus lethality. The antibody titers were determined by ELISA in rabbits’ serum. The immunized animals elicited a response against toxins and whole venom. The Immunoglobulin G (IgGs) obtained were able to neutralize the lethal effect of their homologous toxins. A combination of antivenom from M. mipartitus with antitoxins improved their neutralizing ability. In the same way, a mixture of anti 3FTx and PLA2 protected the mice from a 1.5 median lethal dose (LD50) of M. mipartitus venom. The results showed that this might be a way to improve antibody titers specificity against the relevant toxins in M. mipartitus venom and indicated that there is a possibility to develop and use recombinant 3FTx and PLA2 toxins as immunogens to produce antivenoms. Additionally, this represents an alternative to reduce the amount of venom used in anti-coral antivenom production. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Micrurus mipartitus</span> snake. Orange nuchal band and red terminal rings are evident. Source: Serpentarium University of Antioquia.</p>
Full article ">Figure 2
<p>(<b>A</b>): Elution profile of <span class="html-italic">Micrurus mipartitus</span> venom proteins by RP-HPLC. Two mg of venom was fractionated on a C18 column, as described in the materials and methods. The numbers show the fractions selected for evaluation of the lethal effect. Numbers in blue indicate the fractions that they showed a lethal effect in mice and numbers in red indicate the most lethal and abundant fractions and those used as immunogens. (<b>B</b>): Lethal fractions were analyzed by 15% gradient SDS-PAGE under non-reducing conditions. Mm7: fraction 7, Mm8: fraction 8, Mm20: fraction 20 and Mm21: fraction 21, molecular mass markers indicated in the left (kDa). (<b>C</b>): Mm8 (50 µg) and (<b>D</b>): Mm20 (100 µg) toxin purity profile by RP-HPLC, as described in the materials and methods.</p>
Full article ">Figure 3
<p>ELISA reactivity sera from each bleed against the <span class="html-italic">M. mipartitus</span> whole venom, Mm8 or Mm20. A 96-well plate was coated with each immunogen (Mm8, Mm20 and whole <span class="html-italic">M. mipartitus</span> venom). Serum from each bleeding was used at dilution of 1:100. * Indicates statistically significant differences with the preimmune serum. + indicates statistically significant differences against the Mm8 toxin (<span class="html-italic">p</span> &lt; 0.05). Each bar represents the mean ± SD (<span class="html-italic">n</span> = 2).</p>
Full article ">Figure 4
<p>(<b>A</b>): Titration curve of antibodies in serum from bleeding five by ELISA against <span class="html-italic">M. mipartitus</span> whole venom and its fractions. A 96-well plate was coated with complete <span class="html-italic">M. mipartitus</span> venom and serum from bleeding five was used in dilutions from 1:100 to 1:3200. (<b>B</b>): Titration curve of each IgG by ELISA. A 96-well plate was coated with each immunogen (<span class="html-italic">M. mipartitus</span> whole venom, Mm8 and Mm20), and dilutions of each homologous IgG were added at 1:100 to 1:3200 dilutions. Bound antibodies were detected by a peroxidase-labeled anti-rabbit IgG conjugate. Each point represents the mean + SD (<span class="html-italic">n</span> = 2).</p>
Full article ">Figure 5
<p>Hyperimmune sera and caprylic acid-extracted IgG were evaluated on 10% SDS-PAGE under non-reduced conditions and stained with Coomassie Blue R-250. MW: broad range molecular mass marker (kDa). (<b>a</b>): Albumin standard, (<b>b</b>): anti-Mm serum, (<b>c</b>): anti-Mm20 serum, (<b>d</b>): anti-Mm8 serum, (<b>e</b>): IgG obtained from anti-Mm serum, (<b>f</b>): IgG obtained from anti-Mm20 serum, (<b>g</b>): IgG obtained from anti-Mm8 serum. Red arrows indicate IgG band and blue arrows indicate albumin band.</p>
Full article ">Figure 6
<p>Cross-recognition of anti-toxin IgG against fractions obtained from <span class="html-italic">M. mipartitus</span> venom by RP-HPLC. Anti-Mm IgG recognition against protein families present in the whole venom according to the proteome described by [<a href="#B10-toxins-14-00265" class="html-bibr">10</a>]. Each bar represents the mean + SD (<span class="html-italic">n</span> = 2). * Indicates statistically significant differences with the preimmune serum.</p>
Full article ">Figure 7
<p>(<b>A</b>): Immunorecognition of anti-Mm8 and anti-Mm IgGs against fractions of the 3FTx family; * indicates statistically significant differences with respect to anti-Mm IgG. (<b>B</b>): Immunorecognition of anti-Mm20 and anti-Mm IgGs against fractions of the PLA2 family. Each bar represents the mean + SD (<span class="html-italic">n</span> = 2). * Indicates statistically significant differences with the preimmune serum.</p>
Full article ">
19 pages, 5267 KiB  
Article
Three-Finger Toxins from Brazilian Coral Snakes: From Molecular Framework to Insights in Biological Function
by Jessica Matos Kleiz-Ferreira, Nuria Cirauqui, Edson Araujo Trajano, Marcius da Silva Almeida and Russolina Benedeta Zingali
Toxins 2021, 13(5), 328; https://doi.org/10.3390/toxins13050328 - 30 Apr 2021
Cited by 3 | Viewed by 3808
Abstract
Studies on 3FTxs around the world are showing the amazing diversity in these proteins both in structure and function. In Brazil, we have not realized the broad variety of their amino acid sequences and probable diversified structures and targets. In this context, this [...] Read more.
Studies on 3FTxs around the world are showing the amazing diversity in these proteins both in structure and function. In Brazil, we have not realized the broad variety of their amino acid sequences and probable diversified structures and targets. In this context, this work aims to conduct an in silico systematic study on available 3FTxs found in Micrurus species from Brazil. We elaborated a specific guideline for this toxin family. First, we grouped them according to their structural homologue predicted by HHPred server and further curated manually. For each group, we selected one sequence and constructed a representative structural model. By looking at conserved features and comparing with the information available in the literature for this toxin family, we managed to point to potential biological functions. In parallel, the phylogenetic relationship was estimated for our database by maximum likelihood analyses and a phylogenetic tree was constructed including the homologous 3FTx previously characterized. Our results highlighted an astonishing diversity inside this family of toxins, showing some groups with expected functional similarities to known 3FTxs, and pointing out others with potential novel roles and perhaps structures. Moreover, this classification guideline may be useful to aid future studies on these abundant toxins. Full article
(This article belongs to the Section Animal Venoms)
Show Figures

Figure 1

Figure 1
<p>Sequence alignment of 3FTx groups. Groups are presented with their consensus amino acids and conservation score on a scale from 0 to 10 (10 being indicated by an asterisk as completely conserved, and + for similar amino acids). The cysteine residues forming the disulfide bridges are linked by black lines on the top of the alignment. The extra disulfide bridges are marked with an asterisk. The figures were prepared with the Jalview software [<a href="#B34-toxins-13-00328" class="html-bibr">34</a>] and colored using the Clustalx scheme and conservation.</p>
Full article ">Figure 2
<p>Representative 3D structural models of each 3FTx group. Each group and subgroup are represented by a 3D structure modeled from a representative sequence and identified by its database accession code. The best model was chosen according to the lowest DOPE score (<span class="html-italic">Group 1-A</span> -3469.3; <span class="html-italic">Group 1-B</span> -4223.5; <span class="html-italic">Group 1-C</span> -3663.6; <span class="html-italic">Group 2-A</span> -3945.3; <span class="html-italic">Group 2-B</span> -3770.5; <span class="html-italic">Group 3</span> -4401.7; <span class="html-italic">Group 4</span> -3720.2; <span class="html-italic">Group 5</span> -3608.5; <span class="html-italic">Group 6</span> -5266.9; <span class="html-italic">Group 7</span> -5035.04; <span class="html-italic">Group 8-A</span> -3868.4; <span class="html-italic">Group 8-B</span> -4417.3; <span class="html-italic">Group 9</span> -4235.3). The cysteine residues that form disulfide bridges are marked yellow. The structures are presented from the N-terminus to the C-terminus, and the fingers I, II and III are positioned from the left to the right. Colored boxes represent each group of toxins.</p>
Full article ">Figure 3
<p>Phylogenetic tree of the 3FTxs sequence groups. The evolutionary analysis was generated using the maximum likelihood method and the Whelan and Goldman model. The tree with the highest log likelihood (−2630.25) is shown. The 3FTxs groups are indicated with colored boxes. 3FTxs already characterized were included in this analysis. The sequences found in the tree inside a different group from the one suggested in <a href="#toxins-13-00328-f001" class="html-fig">Figure 1</a> are indicated in blue text, with a grey arrow. The five big branches are shown here in dark blue.</p>
Full article ">Figure 4
<p>Functional amino acid residues and important regions in the representative 3FTxs of some groups. Some important/key amino acids discussed along the text are highlighted here in their corresponding groups. The coloring scheme is: red for negative amino acids, blue for positive, pale orange for aromatic resides, pale pink for glycine, yellow for cysteines, and green for the others. The grey arrows are indicating the orientation of the N and C-terminus. The fingers I, II, and III are positioned from the left to the right.</p>
Full article ">Figure 4 Cont.
<p>Functional amino acid residues and important regions in the representative 3FTxs of some groups. Some important/key amino acids discussed along the text are highlighted here in their corresponding groups. The coloring scheme is: red for negative amino acids, blue for positive, pale orange for aromatic resides, pale pink for glycine, yellow for cysteines, and green for the others. The grey arrows are indicating the orientation of the N and C-terminus. The fingers I, II, and III are positioned from the left to the right.</p>
Full article ">
Back to TopTop