Toxins 您所在的位置:网站首页 toxins free full Toxins

Toxins

2023-08-05 07:40| 来源: 网络整理| 查看: 265

Next Article in Journal The Deciphered Genome of Mesobuthus martensii Uncovers the Resistance Mysteries of Scorpion to Its Own Venom and Toxins at the Ion Channel Level Next Article in Special Issue Evolution Stings: The Origin and Diversification of Scorpion Toxin Peptide Scaffolds Previous Article in Journal Potential Degradation of Swainsonine by Intracellular Enzymes of Arthrobacter sp. HW08 Previous Article in Special Issue Atractaspis aterrima Toxins: The First Insight into the Molecular Evolution of Venom in Side-Stabbers Journals Active Journals Find a Journal Proceedings Series Topics Information For Authors For Reviewers For Editors For Librarians For Publishers For Societies For Conference Organizers Open Access Policy Institutional Open Access Program Special Issues Guidelines Editorial Process Research and Publication Ethics Article Processing Charges Awards Testimonials Author Services Initiatives Sciforum MDPI Books Preprints.org Scilit SciProfiles Encyclopedia JAMS Proceedings Series About Overview Contact Careers News Press Blog Sign In / Sign Up Notice clear Notice

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

Continue Cancel clear

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess.

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Journals Active Journals Find a Journal Proceedings Series Topics Information For Authors For Reviewers For Editors For Librarians For Publishers For Societies For Conference Organizers Open Access Policy Institutional Open Access Program Special Issues Guidelines Editorial Process Research and Publication Ethics Article Processing Charges Awards Testimonials Author Services Initiatives Sciforum MDPI Books Preprints.org Scilit SciProfiles Encyclopedia JAMS Proceedings Series About Overview Contact Careers News Press Blog Sign In / Sign Up Submit     Journals Toxins Volume 5 Issue 11 10.3390/toxins5112172 toxins-logo Submit to this Journal Review for this Journal Edit a Special Issue ► ▼ Article Menu Article Menu Subscribe SciFeed Recommended Articles Related Info Links PubMed/Medline Google Scholar More by Authors Links on DOAJ Sunagar, K. Jackson, T. N. W. Undheim, E. A. B. Ali, S. A. Antunes, A. Fry, B. G. on Google Scholar Sunagar, K. Jackson, T. N. W. Undheim, E. A. B. Ali, S. A. Antunes, A. Fry, B. G. on PubMed Sunagar, K. Jackson, T. N. W. Undheim, E. A. B. Ali, S. A. Antunes, A. Fry, B. G. /ajax/scifeed/subscribe Article Views Citations - Table of Contents Altmetric share Share announcement Help format_quote Cite question_answer Discuss in SciProfiles thumb_up ... Endorse textsms ... Comment Need Help? Support

Find support for a specific problem in the support section of our website.

Get Support Feedback

Please let us know what you think of our products and services.

Give Feedback Information

Visit our dedicated information section to learn more about MDPI.

Get Information clear JSmol Viewer clear first_page settings Order Article Reprints Font Type: Arial Georgia Verdana Font Size: Aa Aa Aa Line Spacing:    Column Width:    Background: Open AccessArticle Three-Fingered RAVERs: Rapid Accumulation of Variations in Exposed Residues of Snake Venom Toxins by Kartik SunagarKartik Sunagar Scilit Preprints.org Google Scholar 1,2, Timothy N. W. JacksonTimothy N. W. Jackson Scilit Preprints.org Google Scholar 3,4, Eivind A. B. UndheimEivind A. B. Undheim Scilit Preprints.org Google Scholar 3,4, Syed. A. AliSyed. A. Ali Scilit Preprints.org Google Scholar 3,4,5, Agostinho AntunesAgostinho Antunes Scilit Preprints.org Google Scholar 1,2 and Bryan G. FryBryan G. Fry Scilit Preprints.org Google Scholar 3,4,* 1 CIMAR/CIIMAR, Centro Interdisciplinar de Investigação Marinha e Ambiental, Universidade do Porto, Rua dos Bragas, 177, Porto 4050-123, Portugal 2 Departamento de Biologia, Faculdade de Ciências, Universidade do Porto, Rua do Campo Alegre, Porto 4169-007, Portugal 3 Venom Evolution Lab, School of Biological Sciences, The University of Queensland, St. Lucia, Queensland 4072, Australia 4 Institute for Molecular Bioscience, The University of Queensland, St. Lucia, Queensland 4072, Australia 5 HEJ Research Institute of Chemistry, International Center for Chemical and Biological Sciences (ICCBS), University of Karachi, Karachi-75270, Pakistan * Author to whom correspondence should be addressed. Toxins 2013, 5(11), 2172-2208; https://doi.org/10.3390/toxins5112172 Received: 2 October 2013 / Revised: 8 November 2013 / Accepted: 11 November 2013 / Published: 18 November 2013 (This article belongs to the Collection Evolution of Venom Systems) Download Download PDF Download PDF with Cover Download XML Download Epub Download Supplementary Material Browse Figures Versions Notes

Abstract: Three-finger toxins (3FTx) represent one of the most abundantly secreted and potently toxic components of colubrid (Colubridae), elapid (Elapidae) and psammophid (Psammophiinae subfamily of the Lamprophidae) snake venom arsenal. Despite their conserved structural similarity, they perform a diversity of biological functions. Although they are theorised to undergo adaptive evolution, the underlying diversification mechanisms remain elusive. Here, we report the molecular evolution of different 3FTx functional forms and show that positively selected point mutations have driven the rapid evolution and diversification of 3FTx. These diversification events not only correlate with the evolution of advanced venom delivery systems (VDS) in Caenophidia, but in particular the explosive diversification of the clade subsequent to the evolution of a high pressure, hollow-fanged VDS in elapids, highlighting the significant role of these toxins in the evolution of advanced snakes. We show that Type I, II and III α-neurotoxins have evolved with extreme rapidity under the influence of positive selection. We also show that novel Oxyuranus/Pseudonaja Type II forms lacking the apotypic loop-2 stabilising cysteine doublet characteristic of Type II forms are not phylogenetically basal in relation to other Type IIs as previously thought, but are the result of secondary loss of these apotypic cysteines on at least three separate occasions. Not all 3FTxs have evolved rapidly: κ-neurotoxins, which form non-covalently associated heterodimers, have experienced a relatively weaker influence of diversifying selection; while cytotoxic 3FTx, with their functional sites, dispersed over 40% of the molecular surface, have been extremely constrained by negative selection. We show that the a previous theory of 3FTx molecular evolution (termed ASSET) is evolutionarily implausible and cannot account for the considerable variation observed in very short segments of 3FTx. Instead, we propose a theory of Rapid Accumulation of Variations in Exposed Residues (RAVER) to illustrate the significance of point mutations, guided by focal mutagenesis and positive selection in the evolution and diversification of 3FTx. Keywords: positive selection; venom evolution; three-finger toxins; RAVER; focal mutagenesis 1. IntroductionVenoms are key evolutionary innovations in the Kingdom Animalia and are complex concoctions of biologically active proteins (from polypeptide globular enzymes to small peptides), salts, and organic molecules such as polyamines, amino acids and neurotransmitters [1]. Venom components originate via toxin recruitment events during which ordinary protein-encoding genes, typically those involved in key regulatory processes (such as hemostasis or neurotransmission) are duplicated, and the new copies are selectively expressed in the venom gland [1,2,3,4,5,6,7,8,9,10,11,12,13]. These novel paralogs can further duplicate and give rise to multigene families, following the “birth and death” mode of evolution, where the rapid evolution of these families results in extensive neofunctionalization of some copies, while the other non-functional forms are lost through degradation or get transformed into pseudogenes [14]. Research has shown that despite the extraordinary diversity of animal toxins, most belong to a limited number of enzymatic (e.g., phospholipases, serine proteases, metalloproteinases) and non-enzymatic (e.g., three-finger toxins, natriuretic peptides, Kunitz peptides, lectins) protein superfamilies which have been convergently recruited in various organisms to perform similar functions [1,3].Three-finger toxins (3FTx) are one of the most abundantly secreted non-enzymatic components of elapid (Elapidae), colubrid (Colubridae) and psammophiide (Psammophiinae subfamily withing the Lamprophidae) snake venom. They are characterised by a broad diversity of functional forms (Table 1). In the past, 3FTx were considered to be exclusive to elapid snake venoms [6]. The discovery of α-colubritoxin, however, revealed this potent toxin type to be widespread in “non-front-fanged” (NFF) Caenophidia snake lineages [15]. Subsequent studies revealed the broad taxonomic distribution of this toxin type [16,17,18,19,20,21,22], which appears to have been recruited into the snake venom arsenal near the base of the snake tree [9]. 3FTx are characterised by the presence of three β-loops that extend from the toxin’s small, hydrophobic core, giving them the “three-fingered” appearance and their name [23]. The plesiotypic (ancestral character state of a molecular scaffold) 3FTx form, such as that found in NFF advanced snakes, has ten cysteines in a distinctive pattern, reflective of its molecular origin from the recruitment of a LYNX/SLUR nicotinic receptor binding neuromodulation peptide [3,6,15,17]. All described 3FTx types from Henophidia, NFF and viperid (Viperidae) snakes contain these ten cysteines, and all of these forms that have been functionally investigated are α-neurotoxic, with a potency much greater against birds and/or reptiles than mammals [6,9,17,19,20,21,24]. This taxon specificity in action led to the plesiotypic α-neurotoxins being mistakenly referred to as “weak neurotoxins” (c.f. [25]). Similarly, this taxon-specific toxicity led to misinterpretation of prey-handling behaviour in an experiment investigating the role of venom in the feeding ecology of NFF snakes [26]. Table Table 1. Bioactivities of 3FTx types with characterised toxicities. Table 1. Bioactivities of 3FTx types with characterised toxicities. Functional ClassMode of ActionBasal-type α-neurotoxinsAntagonists of α1 nicotinic acetylcholine receptors, with a 100-fold greater potency to avians/reptiles than mammals. Produces flaccid paralysis [21]Type I α-neurotoxinsAntagonists of α1 nicotinic acetylcholine receptors. Produces flaccid paralysis [27]Type II α-neurotoxinsAntagonists of α1 and α7 nicotinic acetylcholine receptors. Produces flaccid paralysis [27]Type III α-neurotoxinsAntagonists of α1 nicotinic acetylcholine receptors. Produces flaccid paralysis [28] κ-neurotoxinsAntagonists of α3β2 neuronal nicotinic acetylcholine receptor subtype. Produces flaccid paralysis [29]Adrenergic/Muscarinic neurotoxinsAntagonists of a wide variety of adrenergic and muscarinic subtypes with extreme specificity for receptor subtypes [30,31,32,33,34,35,36,37,38,39,40,41]Type B Muscarinic toxinsAntagonists of M2 muscarinic acetylcholine receptors [42]ASIC channel blockersActs as a reversible gating modifier toxin by antagonistically binding to closed/inactivated ASIC1a-ASIC2a (ACCN2-ACCN1) channels in central neurons and ASIC1b-containing channels in nociceptors [43]Calcium channel blockersAntagonists of L-type calcium channels, thus inhibiting the transmission of the action potential [44]Acetylcholinesterase inhibitorsInhibit acetylcholinesterase through competitive binding [45]Platelet inhibitorsCompetitively bind to platelet GPIIb/IIIa receptor utilising the RGD functional motif, thus blocking platelet aggregation [46]CytotoxinsCell-damaging activity mediated by hydrophobic-patch on molecular surface that interacts non-specifically with hydrophobic aspects of the cell phospholipid bilayer [47]SynergisticAlone are non-toxic but form complexes with α-neurotoxins to dramatically enhance neurotoxicity [48] Toxins 05 02172 g001 1024 Figure 1. Bayesian molecular phylogeny of representative three-finger toxins. Uniprot [49] accession numbers are given for each. Cysteine framework variation is displayed, with ancestral cysteines in black and newly evolved cysteines in red. Figure 1. Bayesian molecular phylogeny of representative three-finger toxins. Uniprot [49] accession numbers are given for each. Cysteine framework variation is displayed, with ancestral cysteines in black and newly evolved cysteines in red. Toxins 05 02172 g001 Subsequent to elapid snakes evolving a high pressure, syringe-like delivery system including venom gland compressor musculature and hollow front fangs, apotypic (derived character state of a molecular scaffold) forms of 3FTx emerged, characterised by a loss of plesiotypic cysteines 2 and 3 (Figure 1), a change that probably resulted in the dramatic potentiation of α-neurotoxicity through the uncoupling of loop-1, with these apotypic forms becoming much more potent upon mammalian receptors than the more constrained plesiotypic forms [6]. This increased toxicity likely enhanced the role of these proteins in prey capture and resulted in a high level expression of α-neurotoxins (α-ntx) lacking the second and third plesiotypic cysteines (Type I (aka: short-chain) and Type II (aka: long-chain) α-ntx) in the venom glands of these snakes. The increased level of expression was accompanied by punctuated molecular evolution, resulting in the emergence of a myriad of structurally and functionally novel forms [6]. Structurally novel forms included α-ntx with newly evolved cysteines that stabilised the second loop (Type II α-ntx). Venoms of snakes from the Oxyuranus and Pseudonaja clade are unusual in containing significant amounts of 3FTx that display all the features of Type II α-ntx (including lacking the 2nd and 3rd plesiotypic cysteines) but lack the apotypic loop-2 stabilising cysteine pair characteristic of this type [50]. These toxins would thus be expected to be phylogenetically basal to the Type II α-ntx which contain the apotypic loop-2 stabilising cysteine pair. Another derived 3FTx structural variation is represented by the κ-neurotoxins (aka: κ-bungarotoxins), which form non-covalently associated heterodimers. A number of novel functions also emerged ([6]; Table 1; Figure 1), such as κ-ntx specifically targeting the neuronal nicotinic receptors. The most extreme neofunctionalisation is represented by the cytotoxins, which have deviated from the highly focused ion channel targeting of the α-ntx [51]. Instead, they exhibit a number of novel biological activities, including lysis of various types of cells (including erythrocytes and epithelial cells), enzyme inhibition (protein kinase C: [52]; Na+/K+ ATPase: [53]), depolarization and contraction of muscle cells and prevention of platelet aggregation. All structurally and functionally apotypic 3FTxs lack plesiotypic cysteines 2 and 3 (Figure 1), and are expressed in the venoms of elapid snakes in much higher levels than the α-ntxs that contain all ten plesiotypic cysteines.Although it has been hypothesised and demonstrated through (now obsolete) selection assessments that snake venom three-finger toxins have evolved under the influence of positive Darwinian selection [6,54,55], the underlying mechanism of evolution driving the diversification of functional forms remains unclear. The molecular evolution of several 3FTx forms, such as κ-ntx, plesiotypic 3FTxs from Henophidia, NFF and elapid snakes and Type III α-neurotoxins from Australian elapids, remain unstudied to date. Molecular mechanisms underlying the evolution of the 3FTx gene in viperid snakes, which have independently evolved a sophisticated high pressure, hollow-fanged venom delivery system (VDS), also remain elusive. Moreover, interpretations regarding the evolution of certain 3FTx types, such as the suggestion that cytotoxins evolve under the influence of positive selection (with an ω value of 19.5) [55], have been questionable. Recently, the molecular evolution of 3FTx homologues in Atractaspis sp. was evaluated and the influence of positive Darwinian selection on genes encoding them was highlighted [56].As the general organization of the 3FTx gene is highly conserved and ordered, Accelerated Segment Switch in Exons to alter Targeting (ASSET) has been postulated as the mechanism driving the molecular evolution of 3FTx [16,57]. This theory suggests that during the evolution of the 3FTx gene, segments in exonic regions have been exchanged with distinctly different ones and the resultant “switching” of segments has generated the observed sequence variation and the functional diversity of 3FTx. The authors further speculated that point mutations alone could not account for the diversity of 3FTx functional forms and that they could only be helpful in fine tuning receptor binding capabilities, which originally arise through ASSET [16,57]. However, this study attempted to classify regions in 3FTx based on simplistic “degree of identity” comparisons. Crucially, such analysis does not take into account the fact that proteins adopt regionally differential rates of evolution [58,59,60]. It has been previously demonstrated in other toxin types that structurally important residues are constrained by negative selection, while regions responsible for biological function and/or those forming the molecular surface accumulate variations under an arms race scenario [58,59,61]. This not only facilitates functional diversification, but also increases the number of active residues on the surface of venom components that can non-specifically interact with novel receptors in prey and induce a plethora of pharmacological effects.In the present study, we test the following hypotheses: (i) that 3FTx with specific sites of action are involved in a coevolutionary arms race with receptors of prey animals and thus are evolving under the influence of positive selection; (ii) that cytotoxic 3FTx, which interact non-specifically with cell membranes, do not experience an arms race and evolve under the constraints of negative selection pressure; (iii) that, in venoms in which 3FTx are a major component (Elapidae, Colubridae and Psammophiinae), 3FTx are rapidly evolving under positive selection (with the exception of cytotoxic 3FTx in the venom of elapid snakes); (iv) that in venoms in which 3FTx are a minor component (Viperidae), 3FTx evolve under a neutral selection regime and do not experience positive selection; and (v) the evolution of the advanced venom delivery system in elapid snakes led to an increase in diversifying pressure on 3FTx, resulting in the diversity of functional forms present in the venoms of snakes from this family. In order to test these hypotheses and provide further insight into the evolution and diversification of this toxin superfamily, we reconstructed the complex molecular evolutionary history of 3FTx. In particular, this study examined: the relative rate of evolution of the plesiotypic α-ntx (i) before and after the evolution of the advanced venom delivery apparatus in Caenophidia (advanced snakes), in order to test whether the evolution of a sophisticated VDS influenced the regime of 3FTx evolution; (ii) before and after the evolution of the sophisticated, high pressure and hollow-fanged venom delivery system in elapids and viperids to test whether the refining of VDS influenced selection pressures on 3FTx genes; (iii) the relative rate of 3FTx evolution subsequent to the loss of two plesiotypic cysteines and resultant potentiation of α-neurotoxicity; (iv) the relative rate of evolution subsequent to the apotyposis (or derivation) of non-covalently associated dimeric forms, in order to find out whether these new structural constraints affected natural selection pressures on this gene; and finally; (v) the relative rate of evolution subsequent to the apotyposis of cytotoxins was examined to understand whether the novel activity recruited by cytotoxins affected their rate of evolution. 2. ResultsBayesian and maximum-likelihood molecular phylogenetic analyses retrieved phylogenetic trees with similar topologies (Figure 1; Supplementary Figure 1), with different toxin sequences forming distinct phylogenetic clades as shown previously [3,6,17]. The Bayesian tree (convergence diagnostics: Average standard deviation of split frequencies: 0.009; PSRF = 1.00) was built using amino acid sequences, since some 3FTxs are only known from their amino acid sequences, while the maximum-likelihood tree (1,000 bootstrap replicates; GTR + I + G) was built using all the nucleotide sequences analysed in this study (n = 457). The overall topology of these phylogenetic trees was largely in concordance with the previously published 3FTx phylogeny [3,6,17].A particularly notable finding was that the novel Type II α-ntx sequences from Oxyuranus and Pseudonaja, which lack the Type II α-ntx characteristic cysteine doublet (−2C) between plesiotypic cysteines 5 and 6, were not phylogenetically basal to the other Type II, nor were they monophyletic, but rather were nested within the regular Type II forms (+2C) (Figure 2). This suggests that the apotypic cysteine doublet characteristic of Type II α-ntx was secondarily lost on at least three occasions. Toxins 05 02172 g002 1024 Figure 2. Bayesian molecular phylogeny and structural and functional evolution of +2C/−2C Type II (long-chain) α-neurotoxins. Pseudonaja/Oxyuranus −2C and +2C sequences are coloured purple and green, respectively. Sequences presented (uniprot): (1) R4FIT5 Pseudonaja modesta, (2) R4FIU6 Pseudonaja modesta, (3) A8HDK6 Pseudonaja textilis, (4) R4FK68 Pseudonaja modesta, (5) A8HDK8 Oxyuranus microlepidotus, (6) A7X4Q3 Oxyuranus microlepidotus, (7) A8HDK7 Oxyuranus microlepidotus, (8) A7X4R0 Oxyuranus microlepidotus, (9) A8HDK9 Oxyuranus scutellatus, (10) Q9W7J5 Pseudonaja textilis, (11) R4FIT0 Pseudonaja modesta, (12) R4G7K3 Pseudonaja modesta, (13) R4G321 Pseudonaja modesta, (14) R4G2J4 Pseudonaja modesta, (15) R4G319 Pseudonaja modesta and (16) R4FK75 Pseudonaja modesta. Figure 2. Bayesian molecular phylogeny and structural and functional evolution of +2C/−2C Type II (long-chain) α-neurotoxins. Pseudonaja/Oxyuranus −2C and +2C sequences are coloured purple and green, respectively. Sequences presented (uniprot): (1) R4FIT5 Pseudonaja modesta, (2) R4FIU6 Pseudonaja modesta, (3) A8HDK6 Pseudonaja textilis, (4) R4FK68 Pseudonaja modesta, (5) A8HDK8 Oxyuranus microlepidotus, (6) A7X4Q3 Oxyuranus microlepidotus, (7) A8HDK7 Oxyuranus microlepidotus, (8) A7X4R0 Oxyuranus microlepidotus, (9) A8HDK9 Oxyuranus scutellatus, (10) Q9W7J5 Pseudonaja textilis, (11) R4FIT0 Pseudonaja modesta, (12) R4G7K3 Pseudonaja modesta, (13) R4G321 Pseudonaja modesta, (14) R4G2J4 Pseudonaja modesta, (15) R4G319 Pseudonaja modesta and (16) R4FK75 Pseudonaja modesta. Toxins 05 02172 g002 Our analysis of the venom gland transcriptome of the unique viperid species Azemiops feae recovered a 3FTx transcript that was found in the same clade as the representative sequences from the pit-viper Sistrurus catenatus (Figure 1; Supplementary Figure 1). This is the first non-Sistrurus 3FTx to be recovered from a viperid snake. Based on an unresolved neighbour joining tree, viperid 3FTx were previously reported as polyphyletic, with some sequences nested within elapid 3FTx clade, and others in the “non-front-fanged” advanced snake 3FTx clade [24]. However, both Bayesian and maximum-likelihood phylogenetic analyses in this study place all viperid sequences in a monophyletic clade outside elapid 3FTx clade with strong node support (Figure 1; Supplementary Figure 1). This is consistent with the viperid snakes having diverged from the remaining advanced snakes nearly 54 million years ago [62]. viperid 3FTx homologs retrieved to date indicate that apotypic structural and functional forms, such as Type I, II and III α-ntxs, κ-ntxs and cytotoxins, are the result of apotyposis in elapid snakes after this split (Figure 1; Supplementary Figure 1).The one-ratio model (ORM), the simplest of the lineage-specific models, computed a wide range of ω (dN/dS) values for various types of 3FTx (Supplementary Tables 1–10). This highly conservative model can only detect positive selection when the ω ratio, averaged over all sites along the lineages in a phylogenetic tree, is significantly greater than one. Despite this, the ω value for most three-finger toxins was significantly greater than 1, indicating the strong influence of positive selection in shaping the evolution of 3FTx: plesiotypic α-ntx from viperids, NFF and elapids: 1.79, 1.29 and 1.30, respectively; Type I, Type II and Type III α-ntx: 1.92, 2.01 and 2.59, respectively; and κ-ntx: 1.64. Oxyuranus/Pseudonaja Type II α-ntx with (+2C) and without the cys doublet (−2C) were significantly different, with ω values of 0.97 and 2.69, respectively. In contrast to all other 3FTx types, ORM estimated an ω value of 0.32 for cytotoxins, indicating an unprecedented lack of variation in this unique three-finger toxin.The segregation of seromucous mixed maxillary glands into discrete protein and mucus glands at the base of the advanced snake tree and the subsequent evolution of a high pressure, hollow-fanged venom delivery apparatus independently in Atractaspis/Homoroselaps, elapids and viperids were major evolutionary advances in snake-feeding ecology [7,17]. In order to assess the effect of apomorphic (derived state of a morphological character) venom delivery systems on the evolution of 3FTx, we employed the three-ratio model (3RM) and the branch-site test A (BST) on the plesiotypic α-ntx from the advanced snakes (NFF, viperids and elapids). While 3RM—which assesses selection pressure only across lineages—estimated ω of 1.13, 1.37 and 2.21 for the plesiotypic α-ntx from NFF, elapids and viperids, respectively; BST—which assess selection pressures across the sites and along the lineages in a tree—estimated ω of 2.16 (22% sites), 4.09 (23% sites) and 6.54 (31% sites), respectively (Table 2). NFF and elapid comparisons were insignificant (p > 0.05) for 3RM, while all other comparisons were significant (significant at 0.001 even after Bonferroni corrections). 3RM and BST indicated a greater influence of positive selection on the plesiotypic α-ntxs in elapid and viperid snakes, which have independently evolved sophisticated high pressure and hollow-fanged venom delivery systems. Table Table 2. Lineage-specific analyses of plesiotypic Caenophidia three-finger toxins (3FTxs). Table 2. Lineage-specific analyses of plesiotypic Caenophidia three-finger toxins (3FTxs). Modelω aLikelihood (ι)Prop. of Sites with ω > 1 bSignificance cPlesiotypic α-neurotoxins from ‘non-front-fanged’ advanced snakesThree-ratio model1.13−5982.232579-P > 0.05 NSBranch-site model A2.16−5767.71045422.1%* P 1 (positive selection) and M7 (Beta) versus M8 (Beta and ω), and models that mirror the evolutionary constraints of M1 and M2 but assume that ω values are drawn from a beta distribution [125]. Only if the alternative models (M3, M2a and M8: allow sites with ω > 1) show a better fit in Likelihood Ratio Test (LRT) relative to their null models (M0, M1a and M7: do not allow sites ω > 1), are their results considered significant. It should be noted that although we have reported the results of all site-specific models (M2a, M3 and M8), only the results of M8, the most accurate model among these, were considered for downstream analyses (Figure 3, Figure 4, Figure 5, Figure 6, Figure 7, and Figure 8; Table 4 and Table 5). LRT is estimated as twice the difference in maximum likelihood values between nested models and compared with the χ2 distribution with the appropriate degree of freedom—the difference in the number of parameters between the two models. The Bayes empirical Bayes (BEB) approach [126] was used to identify amino acids under positive selection by calculating the posterior probabilities that a particular amino acid belongs to a given selection class (neutral, conserved or highly variable). Sites with greater posterior probability (PP ≥ 95%) of belonging to the “ω > 1 class” were inferred to be positively selected.FUBAR [127] implemented in HyPhy [128] was employed to detect codon sites evolving under the influence of pervasive diversifying and purifying selection pressures. The Mixed Effects Model of Evolution (MEME) [129] was also employed to efficiently detect sites that experience diversifying selection for a short period in the evolutionary timescale. Non-synonymous mutations that introduce extremely variant amino acids are more likely to influence the structure-function, and hence the fitness of the organism, relative to mutations that introduce amino acids with similar side-chains as the ancestral residues. Further support for the results of the nucleotide-level selection analyses was obtained and the radicalness of mutations were assessed using a complementary protein-level approach implemented in TreeSAAP [130].Direct comparison of ω values computed from different datasets can be misleading, as they can have different proportions of sites under selection. Hence, we assessed the selection pressures shaping different 3FTx clades by employing clade model analyses implemented in Codeml and simultaneously estimated ω values [131]. The significance of the analysis was tested by comparing the likelihood of this model with that of model M1a. To clearly depict the proportion of sites under different regimes of selection, an evolutionary fingerprint analysis was carried out using the evolutionary selection distance (ESD) algorithm implemented in datamonkey [132]. 5.7. Structural AnalysesTo depict the natural selection pressures influencing the evolution of various three-finger toxins, we mapped the sites under positive selection on the homology models created using the Phyre 2 webserver [133]. Pymol 1.3 [134] was used to visualise and generate the images of homology models. The Consurf webserver [135] was used for mapping the evolutionary selection pressures on the three-dimensional homology models. GETAREA [136] was used to calculate the Accessible Surface Area (ASA) or the solvent exposure of amino-acid side chains. It uses the atom co-ordinates of the PDB file and indicates if a residue is buried or exposed to the surrounding medium by comparing the ratio between side chain ASA and the “random coil” values per residue. An amino acid is considered to be buried if it has an ASA less than 20% and exposed if the ASA is more than or equal to 50%. When ASA ratio lies between 40% and 50%, it is highly likely that the residues have their side chains exposed to the surrounding medium. AcknowledgementsKS was funded by the PhD grant (SFRH/BD/61959/2009) from F.C.T (Fundação para a Ciência e a Tecnologia). TNWJ was funded by a PhD scholarship from the University of Queensland. AA was funded by the European Regional Development Fund (ERDF) through the COMPETE - Operational Competitiveness Programme and national funds through F.C.T under the projects PEst-C/MAR/LA0015/2013 and PTDC/AAC-AMB/121301/2010 (FCOMP-01-0124-FEDER-019490), and BGF was funded by the Australian Research Council (ARC) and the University of Queensland. SAA was the recipient of a postdoctoral fellowship (PDRF Phase II Batch-V) from Higher Education Commission (HEC Islamabad) Pakistan. EABU acknowledges funding from the University of Queensland (International Postgraduate Research Scholarship, UQ Centennial Scholarship, and UQ Advantage Top-Up Scholarship) and the Norwegian State Education Loans Fund.Conflicts of InterestThe authors declare no conflict of interest.ReferencesFry, B.G.; Roelants, K.; Champagne, D.E.; Scheib, H.; Tyndall, J.D.A.; King, G.F.; Nevalainen, T.J.; Norman, J.A.; Lewis, R.J.; Norton, R.S.; et al. The toxicogenomic multiverse: Convergent recruitment of proteins into animal venoms. Annu. Rev. Genomics Hum. Genet. 2009, 10, 483–511. [Google Scholar] [CrossRef]Escoubas, P.; Sollod, B.; King, G.F. Venom landscapes: Mining the complexity of spider venoms via a combined cDNA and mass spectrometric approach. Toxicon 2006, 47, 650–663. [Google Scholar] [CrossRef]Fry, B.G. From genome to “venome”: Molecular origin and evolution of the snake venom proteome inferred from phylogenetic analysis of toxin sequences and related body proteins. Genome Res. 2005, 15, 403–420. [Google Scholar] [CrossRef]Inceoglu, B.; Lango, J.; Jing, J.; Chen, L.; Doymaz, F.; Pessah, I.N.; Hammock, B.D. One scorpion, two venoms: Prevenom of Parabuthus transvaalicus acts as an alternative type of venom with distinct mechanism of action. Proc. Natl. Acad. Sci. USA 2003, 100, 922–927. [Google Scholar]Olivera, B.M. CONUS VENOM PEPTIDES: Reflections from the biology of clades and species. Annu. Rev. Ecol. Syst. 2002, 33, 25–47. [Google Scholar] [CrossRef]Fry, B.G.; Wüster, W.; Kini, R.M.; Brusic, V.; Khan, A.; Venkataraman, D.; Rooney, A.P. Molecular evolution and phylogeny of elapid snake venom three-finger toxins. J. Mol. Evol. 2003, 57, 110–129. [Google Scholar]Fry, B.G.; Casewell, N.R.; Wüster, W.; Vidal, N.; Young, B.; Jackson, T.N.W. The structural and functional diversification of the Toxicofera reptile venom system. Toxicon 2012, 60, 434–448. [Google Scholar]Chang, D.; Duda, T.F. Extensive and continuous duplication facilitates rapid evolution and diversification of gene families. Mol. Biol. Evol. 2012, 29, 2019–2029. [Google Scholar]Fry, B.G.; Undheim, E.A.B.; Ali, S.A.; Jackson, T.N.W.; Debono, J.; Scheib, H.; Ruder, T.; Morgenstern, D.; Cadwallader, L.; Whitehead, D.; et al. Squeezers and leaf-cutters: differential diversification and degeneration of the venom system in toxicoferan reptiles. Mol. Cell. Proteomics 2013, 12, 1881–1899. [Google Scholar] [CrossRef]Kozminsky-Atias, A.; Zilberberg, N. Molding the business end of neurotoxins by diversifying evolution. FASEB J. 2012, 26, 576–586. [Google Scholar] [CrossRef]Casewell, N.R.; Wagstaff, S.C.; Harrison, R.A.; Renjifo, C.; Wuster, W. Domain loss facilitates accelerated evolution and neofunctionalization of duplicate snake venom metalloproteinase toxin genes. Mol. Biol. Evol. 2011, 28, 2637–2649. [Google Scholar] [CrossRef]Weinberger, H.; Moran, Y.; Gordon, D.; Turkov, M.; Kahn, R.; Gurevitz, M. Positions under positive selection--key for selectivity and potency of scorpion alpha-toxins. Mol. Biol. Evol. 2010, 27, 1025–1034. [Google Scholar]Zhu, S.; Peigneur, S.; Gao, B.; Lu, X.; Cao, C.; Tytgat, J. Evolutionary diversification of Mesobuthus α-scorpion toxins affecting sodium channels. Mol. Cell. Proteomics 2012, 11, M111.012054. [Google Scholar]Nei, M.; Gu, X.; Sitnikova, T. Evolution by the birth-and-death process in multigene families of the vertebrate immune system. Proc. Natl. Acad. Sci. USA 1997, 94, 7799–7806. [Google Scholar]Fry, B.G.; Lumsden, N.G.; Wüster, W.; Wickramaratna, J.C.; Hodgson, W.C.; Kini, R.M. Isolation of a neurotoxin (alpha-colubritoxin) from a nonvenomous colubrid: evidence for early origin of venom in snakes. J. Mol. Evol. 2003, 57, 446–452. [Google Scholar] [CrossRef]Doley, R.; Pahari, S.; Mackessy, S.P.; Kini, R.M. Accelerated exchange of exon segments in Viperid three-finger toxin genes (Sistrurus catenatus edwardsii; Desert Massasauga). BMC Evol. Biol. 2008, 8, 196. [Google Scholar]Fry, B.G.; Scheib, H.; van der Weerd, L.; Young, B.; McNaughtan, J.; Ramjan, S.F.R.; Vidal, N.; Poelmann, R.E.; Norman, J.A. Evolution of an arsenal: Structural and functional diversification of the venom system in the advanced snakes (Caenophidia). Mol. Cell. Proteomics 2008, 7, 215–246. [Google Scholar]Fry, B.G.; Wüster, W.; Ryan Ramjan, S.F.; Jackson, T.; Martelli, P.; Kini, R.M. Analysis of Colubroidea snake venoms by liquid chromatography with mass spectrometry: Evolutionary and toxinological implications. Rapid Commun. Mass Spectrom. 2003, 17, 2047–2062. [Google Scholar]Lumsden, N.G.; Fry, B.G.; Ventura, S.; Kini, R.M.; Hodgson, W.C. The in vitro and in vivo pharmacological activity of Boiga dendrophila (mangrove catsnake) venom. Auton. Autacoid Pharmacol. 2004, 24, 107–113. [Google Scholar]Pawlak, J.; Mackessy, S.P.; Fry, B.G.; Bhatia, M.; Mourier, G.; Fruchart-Gaillard, C.; Servent, D.; Ménez, R.; Stura, E.; Ménez, A.; et al. Denmotoxin, a three-finger toxin from the colubrid snake Boiga dendrophila (Mangrove Catsnake) with bird-specific activity. J. Biol. Chem. 2006, 281, 29030–29041. [Google Scholar] [CrossRef]Pawlak, J.; Mackessy, S.P.; Sixberry, N.M.; Stura, E.A.; le Du, M.H.; Ménez, R.; Foo, C.S.; Ménez, A.; Nirthanan, S.; Kini, R.M. Irditoxin, a novel covalently linked heterodimeric three-finger toxin with high taxon-specific neurotoxicity. FASEB J. 2009, 23, 534–545. [Google Scholar]Lumsden, N.G.; Fry, B.G.; Manjunatha Kini, R.; Hodgson, W.C. In vitro neuromuscular activity of “colubrid” venoms: Clinical and evolutionary implications. Toxicon 2004, 43, 819–827. [Google Scholar]Kini, R.M.; Doley, R. Structure, function and evolution of three-finger toxins: Mini proteins with multiple targets. Toxicon 2010, 56, 855–867. [Google Scholar] [CrossRef]Pahari, S.; Mackessy, S.P.; Kini, R.M. The venom gland transcriptome of the Desert Massasauga Rattlesnake (Sistrurus catenatus edwardsii): Towards an understanding of venom composition among advanced snakes (Superfamily Colubroidea). BMC Mol. Biol. 2007, 8, 115. [Google Scholar] [CrossRef]Utkin, Y.N.; Kukhtina, V.V.; Kryukova, E.V.; Chiodini, F.; Bertrand, D.; Methfessel, C.; Tsetlin, V.I. “Weak Toxin” from naja kaouthia is a nontoxic antagonist of α7 and muscle-type nicotinic acetylcholine receptors. J. Biol. Chem. 2001, 276, 15810–15815. [Google Scholar] [CrossRef]Rochelle, M.J.; Kardong, K.V. Constriction versus envenomation in prey capture by the brown tree snake, boiga irregularis (Squamata: Colubridae). Herpetologica 1993, 49, 301–304. [Google Scholar]Barber, C.M.; Isbister, G.K.; Hodgson, W.C. Alpha neurotoxins. Toxicon 2013, 66, 47–58. [Google Scholar]Gong, N.; Armugam, A.; Jeyaseelan, K. Postsynaptic short-chain neurotoxins from Pseudonaja textilis. cDNA cloning, expression and protein characterization. Eur. J. Biochem. 1999, 265, 982–989. [Google Scholar]Chiappinelli, V.A.; Weaver, W.R.; McLane, K.E.; Conti-Fine, B.M.; Fiordalisi, J.J.; Grant, G.A. Binding of native kappa-neurotoxins and site-directed mutants to nicotinic acetylcholine receptors. Toxicon 1996, 34, 1243–1256. [Google Scholar] [CrossRef]Carsi, J.M.; Potter, L.T. m1-toxin isotoxins from the green mamba (Dendroaspis angusticeps) that selectively block m1 muscarinic receptors. Toxicon 2000, 38, 187–198. [Google Scholar] [CrossRef]Fruchart-Gaillard, C.; Mourier, G.; Marquer, C.; Stura, E.; Birdsall, N.J.M.; Servent, D. Different interactions between MT7 toxin and the human muscarinic M1 receptor in its free and N-methylscopolamine-occupied states. Mol. Pharmacol. 2008, 74, 1554–1563. [Google Scholar] [CrossRef]Jianmongkol, S.; Marable, B.R.; Berkman, C.E.; Talley, T.T.; Thompson, C.M.; Richardson, R.J. Kinetic evidence for different mechanisms of acetylcholinesterase inhibition by (1R)- and (1S)-stereoisomers of isomalathion. Toxicol. Appl. Pharmacol. 1999, 155, 43–53. [Google Scholar] [CrossRef]Jolkkonen, M.; van Giersbergen, P.L.; Hellman, U.; Wernstedt, C.; Karlsson, E. A toxin from the green mamba Dendroaspis angusticeps: Amino acid sequence and selectivity for muscarinic m4 receptors. FEBS Lett. 1994, 352, 91–94. [Google Scholar] [CrossRef]Karlsson, E.; Jolkkonen, M.; Satyapan, N.; Adem, A.; Kumlin, E.; Hellman, U.; Wernstedt, C. Protein toxins that bind to muscarinic acetylcholine receptors. Ann. N. Y. Acad. Sci. 1994, 710, 153–161. [Google Scholar] [CrossRef]Koivula, K.; Rondinelli, S.; Nasman, J. The three-finger toxin MTalpha is a selective alpha(2B)-adrenoceptor antagonist. Toxicon 2010, 56, 440–447. [Google Scholar] [CrossRef]Kornisiuk, E.; Jerusalinsky, D.; Cerveñansky, C.; Harvey, A.L. Binding of muscarinic toxins MTx1 and MTx2 from the venom of the green mamba Dendroaspis angusticeps to cloned human muscarinic cholinoceptors. Toxicon 1995, 33, 11–18. [Google Scholar] [CrossRef]Krajewski, J.L.; Dickerson, I.M.; Potter, L.T. Site-directed mutagenesis of m1-toxin1: Two amino acids responsible for stable toxin binding to M(1) muscarinic receptors. Mol. Pharmacol. 2001, 60, 725–731. [Google Scholar]Kukhtina, V.V.; Weise, C.; Muranova, T.A.; Starkov, V.G.; Franke, P.; Hucho, F.; Wnendt, S.; Gillen, C.; Tsetlin, V.I.; Utkin, Y.N. Muscarinic toxin-like proteins from cobra venom. Eur. J. Biochem. 2000, 267, 6784–6789. [Google Scholar] [CrossRef]Quinton, L.; Girard, E.; Maiga, A.; Rekik, M.; Lluel, P.; Masuyer, G.; Larregola, M.; Marquer, C.; Ciolek, J.; Magnin, T.; et al. Isolation and pharmacological characterization of AdTx1, a natural peptide displaying specific insurmountable antagonism of the alpha1A-adrenoceptor. Br. J. Pharmacol. 2010, 159, 316–325. [Google Scholar] [CrossRef]Rouget, C.; Quinton, L.; Maïga, A.; Gales, C.; Masuyer, G.; Malosse, C.; Chamot-Rooke, J.; Thai, R.; Mourier, G.; de Pauw, E.; et al. Identification of a novel snake peptide toxin displaying high affinity and antagonist behaviour for the α2-adrenoceptors. Br. J. Pharmacol. 2010, 161, 1361–1374. [Google Scholar] [CrossRef]Toomela, T.; Jolkkonen, M.; Rinken, A.; Järv, J.; Karlsson, E. Two-step binding of green mamba toxin to muscarinic acetylcholine receptor. FEBS Lett. 1994, 352, 95–97. [Google Scholar] [CrossRef]Carsi, J.M.; Valentine, H.H.; Potter, L.T. m2-toxin: A selective ligand for M2 muscarinic receptors. Mol. Pharmacol. 1999, 56, 933–937. [Google Scholar]Diochot, S.; Baron, A.; Salinas, M.; Douguet, D.; Scarzello, S.; Dabert-Gay, A.S.; Debayle, D.; Friend, V.; Alloui, A.; Lazdunski, M.; et al. Black mamba venom peptides target acid-sensing ion channels to abolish pain. Nature 2012, 490, 552–555. [Google Scholar] [CrossRef]De Weille, J.R.; Schweitz, H.; Maes, P.; Tartar, A.; Lazdunski, M. Calciseptine, a peptide isolated from black mamba venom, is a specific blocker of the L-type calcium channel. Proc. Natl. Acad. Sci. USA 1991, 88, 2437–2440. [Google Scholar] [CrossRef]Karlsson, E.; Mbugua, P.M.; Rodriguez-Ithurralde, D. Fasciculins, anticholinesterase toxins from the venom of the green mamba Dendroaspis angusticeps. J. Physiol. (Paris) 1984, 79, 232–240. [Google Scholar]McDowell, R.S.; Dennis, M.S.; Louie, A.; Shuster, M.; Mulkerrin, M.G.; Lazarus, R.A. Mambin, a potent glycoprotein IIb-IIIa antagonist and platelet aggregation inhibitor structurally related to the short neurotoxins. Biochemistry 1992, 31, 4766–4772. [Google Scholar]Konshina, A.G.; Dubovskii, P.V.; Efremov, R.G. Structure and dynamics of cardiotoxins. Curr. Protein Pept. Sci. 2012, 13, 570–584. [Google Scholar] [CrossRef]Joubert, F.J.; Viljoen, C.C. Snake venom. The amino-acid sequence of the subunits of two reduced and S-carboxymethylated proteins (C8S2 and C9S3) from Dendroaspis angusticeps venom. Hoppe. Seylers. Z. Physiol. Chem. 1979, 360, 1075–1090. [Google Scholar] [CrossRef]Uniprot. Available online: www.uniprot.org (accessed on 14 November 2013).St Pierre, L.; Fischer, H.; Adams, D.J.; Schenning, M.; Lavidis, N.; de Jersey, J.; Masci, P.P.; Lavin, M.F. Distinct activities of novel neurotoxins from Australian venomous snakes for nicotinic acetylcholine receptors. Cell. Mol. Life Sci. 2007, 64, 2829–2840. [Google Scholar] [CrossRef]Bhaskaran, R.; Huang, C.C.; Chang, D.K.; Yu, C. Cardiotoxin III from the Taiwan cobra (Naja naja atra). Determination of structure in solution and comparison with short neurotoxins. J. Mol. Biol. 1994, 235, 1291–1301. [Google Scholar] [CrossRef]Harvey, A.L.; Marshall, R.J.; Karlsson, E. Effects of purified cardiotoxins from the Thailand cobra (Naja naja siamensis) on isolated skeletal and cardiac muscle preparations. Toxicon 1982, 20, 379–396. [Google Scholar] [CrossRef]Bougis, P.E.; Khélif, A.; Rochat, H. On the inhibition of [Na+,K+]-ATPases by the components of Naja mossambica mossambica venom: Evidence for two distinct rat brain [Na+,K+]-ATPase activities. Biochemistry 1989, 28, 3037–3043. [Google Scholar] [CrossRef]Gong, N.; Armugam, A.; Jeyaseelan, K. Molecular cloning, characterization and evolution of the gene encoding a new group of short-chain alpha-neurotoxins in an Australian elapid, Pseudonaja textilis. FEBS Lett. 2000, 473, 303–310. [Google Scholar] [CrossRef]Jiang, Y.; Li, Y.; Lee, W.; Xu, X.; Zhang, Y.Y.; Zhao, R.; Wang, W. Venom gland transcriptomes of two elapid snakes (Bungarus multicinctus and Naja atra) and evolution of toxin genes. BMC Genomics 2011, 12, 1. [Google Scholar] [CrossRef]Terrat, Y.; Sunagar, K.; Fry, B.; Jackson, T.; Scheib, H.; Fourmy, R.; Verdenaud, M.; Blanchet, G.; Antunes, A.; Ducancel, F. Atractaspis aterrima Toxins: The First Insight into the Molecular Evolution of Venom in Side-Stabbers. Toxins 2013, 5, 1948–1964. [Google Scholar] [CrossRef]Doley, R.; Mackessy, S.P.; Kini, R.M. Role of accelerated segment switch in exons to alter targeting (ASSET) in the molecular evolution of snake venom proteins. BMC Evol. Biol. 2009, 9, 146. [Google Scholar] [CrossRef]Sunagar, K.; Johnson, W.E.; O’Brien, S.J.; Vasconcelos, V.; Antunes, A. Evolution of CRISPs associated with toxicoferan-reptilian venom and mammalian reproduction. Mol. Biol. Evol. 2012, 29, 1807–1822. [Google Scholar] [CrossRef]Brust, A.; Sunagar, K.; Undheim, E.A.B.; Vetter, I.; Yang, D.C.; Casewell, N.R.; Jackson, T.N.W.; Koludarov, I.; Alewood, P.F.; Hodgson, W.C.; et al. Differential evolution and neofunctionalization of snake venom metalloprotease domains. Mol. Cell. Proteomics 2013, 12, 651–663. [Google Scholar] [CrossRef]Sunagar, K.; Fry, B.G.; Jackson, T.; Casewell, N.R.; Undheim, E.A.B.; Vidal, N.; Ali, S.; King, G.F.; Vasudevan, K.; Vasconcelos, V.; et al. Molecular evolution of vertebrate neurotrophins: Co-option of the highly conserved nerve growth factor gene into the advanced snake venom arsenal. PLoS One 2013, in press. [Google Scholar]Low, D.H.W.; Sunagar, K.; Undheim, E.A.B.; Ali, S.A.; Alagon, A.C.; Ruder, T.; Jackson, T.N.W.; Pineda Gonzalez, S.; King, G.F.; Jones, A.; Antunes, A.; Fry, B.G. Dracula’s children: Molecular evolution of vampire bat venom. J. Proteomics 2013, 89, 95–111. [Google Scholar] [CrossRef]Vidal, N.; Rage, J.; Couloux, A.; Hedges, S.B. Snakes (Serpentes). In The Timetree of Life; Hedges, S.B., Kumar, S., Eds.; Oxford University Press: New York, NY, USA, 2009; pp. 390–397. [Google Scholar]Kosakovsky Pond, S.L.; Murrell, B.; Fourment, M.; Frost, S.D.W.; Delport, W.; Scheffler, K. A random effects branch-site model for detecting episodic diversifying selection. Mol. Biol. Evol. 2011, 28, 3033–3043. [Google Scholar] [CrossRef]Fry, B.G.; Vidal, N.; Norman, J.A.; Vonk, F.J.; Scheib, H.; Ramjan, S.F.R.; Kuruppu, S.; Fung, K.; Hedges, S.B.; Richardson, M.K.; et al. Early evolution of the venom system in lizards and snakes. Nature 2006, 439, 584–588. [Google Scholar] [CrossRef]Fry, B.G.; Vidal, N.; van der Weerd, L.; Kochva, E.; Renjifo, C. Evolution and diversification of the Toxicofera reptile venom system. J. Proteomics 2009, 72, 127–136. [Google Scholar] [CrossRef]Fry, B.G.; Winter, K.; Norman, J.A.; Roelants, K.; Nabuurs, R.J.A.; van Osch, M.J.P.; Teeuwisse, W.M.; van der Weerd, L.; McNaughtan, J.E.; Kwok, H.F.; et al. Functional and structural diversification of the Anguimorpha lizard venom system. Mol. Cell. Proteomics 2010, 9, 2369–2390. [Google Scholar] [CrossRef]Fry, B.G.; Scheib, H.; de L M Junqueira de Azevedo, I.; Silva, D.A.; Casewell, N.R. Novel transcripts in the maxillary venom glands of advanced snakes. Toxicon 2012, 59, 696–708. [Google Scholar] [CrossRef]Li, M.; Fry, B.G.; Kini, R.M. Putting the brakes on snake venom evolution: The unique molecular evolutionary patterns of Aipysurus eydouxii (Marbled sea snake) phospholipase A2 toxins. Mol. Biol. Evol. 2005, 22, 934–941. [Google Scholar] [CrossRef]Junqueira de Azevedo, I.L.; Farsky, S.H.; Oliveira, M.L.; Ho, P.L. Molecular cloning and expression of a functional snake venom vascular endothelium growth factor (VEGF) from the Bothrops insularis pit viper. A new member of the VEGF family of proteins. J. Biol. Chem. 2001, 276, 39836–39842. [Google Scholar]Deufel, A.; Cundall, D. Prey transport in “palatine-erecting” elapid snakes. J. Morphol. 2003, 258, 358–375. [Google Scholar] [CrossRef]Deufel, A.; Cundall, D. Feeding in Atractaspis (Serpentes: Atractaspididae): A study in conflicting functional constraints. Zoology (Jena) 2003, 106, 43–61. [Google Scholar] [CrossRef]Gopalakrishnakone, P.; Kochva, E. Venom glands and some associated muscles in sea snakes. J. Morphol. 1990, 205, 85–96. [Google Scholar] [CrossRef]McCarthy, C.J. Adaptations of sea snakes that eat fish eggs; with a note on the throat musculature of Aipysurus eydouxi (Gray, 1849). J. Nat. Hist. 1987, 21, 1119–1128. [Google Scholar] [CrossRef]Salomão, M.G.; Laporta-Ferreira, I.L. The role of secretions from the supralabial, infralabial, and Duvernoy’s glands of the slug-eating snake Sibynomorphus mikani (Colubridae: Dipsadinae) in the immobilization of molluscan prey. J. Herpetol. 1994, 28, 369–371. [Google Scholar] [CrossRef]Voris, H.K. Fish eggs as the apparent sole food item for a genus of sea snake, emydocephalus (krefft). Ecology 1966, 47, 152. [Google Scholar] [CrossRef]VORIS, H.K.; VORIS, H.H. Feeding strategies in marine snakes: An analysis of evolutionary, morphological, behavioral and ecological relationships. Integr. Comp. Biol. 1983, 23, 411–425. [Google Scholar] [CrossRef]Limpus, C.J. The venom apparatus and venom yields of sub-tropical Queensland Hydrophiidae. Toxicon Suppl. 1978, 39–70. [Google Scholar]Cundall, D.; Deufel, A. Influence of the venom delivery system on intraoral prey transport in snakes. Zool. Anzeiger - A J. Comp. Zool. 2006, 245, 193–210. [Google Scholar] [CrossRef]Deufel, A.; Cundall, D. Functional plasticity of the venom delivery system in snakes with a focus on the poststrike prey release behavior. Zool. Anz. 2006, 245, 249–267. [Google Scholar] [CrossRef]Jackson, T.N.W.; Casewell, N.R.; Fry, B.G. Response to “Replies to Fry et al. (Toxicon 2012, 60/4, 434-448). Part A. Analyses of squamate reptile oral glands and their products: A call for caution in formal assignment of terminology designating biological function”. Toxicon 2013, 64, 106–112. [Google Scholar] [CrossRef]Jackson, T.N.W.; Casewell, N.R.; Fry, B.G. Response to “Replies to Fry et al. (Toxicon 2012, 60/4, 434–448). Part B. Analyses of squamate reptile oral glands and their products: A call for caution in formal assignment of terminology designating biological function”. Toxicon 2013, 64, 113–115. [Google Scholar] [CrossRef]Prado-Franceschi, J.; Hyslop, S.; Cogo, J.C.; Andrade, A.L.; Assakura, M.; Cruz-Hofling, M.A.; Rodrigues-Simioni, L. The effects of Duvernoy’s gland secretion from the xenodontine colubrid Philodryas olfersii on striated muscle and the neuromuscular junction: Partial characterization of a neuromuscular fraction. Toxicon 1996, 34, 459–466. [Google Scholar] [CrossRef]Shine, R. Comparative ecology of three australian snake species of the genus cacophis (Serpentes: Elapidae). Copeia 1980, 1980, 831–838. [Google Scholar] [CrossRef]Shine, R. Reproduction, feeding and growth in the australian burrowing snake vermicella annulata. J. Herpetol. 1980, 14, 71–77. [Google Scholar] [CrossRef]Shine, R. Ecology of Australian elapid snakes of the genera furina and glyphodon. J. Herpetol. 1981, 15, 219. [Google Scholar] [CrossRef]Shine, R. Food habits and reproductive biology of small australian snakes of the genera unechis and suta (Elapidae). J. Herpetol. 1988, 22, 307. [Google Scholar] [CrossRef]Shine, R. Constraints, allometry, and adaptation - food-habits and reproductive-biology of Australian brownsnakes (Pseudonaja, Elapidae). Herpetologica 1989, 45, 195–207. [Google Scholar]St Pierre, L.; Birrell, G.W.; Earl, S.T.; Wallis, T.P.; Gorman, J.J.; de Jersey, J.; Masci, P.P.; Lavin, M.F. Diversity of toxic components from the venom of the evolutionarily distinct black whip snake, Demansia vestigiata. J. Proteome Res. 2007, 6, 3093–3107. [Google Scholar] [CrossRef]Condrea, E. Hemolytic disorders associated with a primary red cell membrane defect. Experientia 1976, 32, 537–542. [Google Scholar] [CrossRef]Dufton, M.J.; Hider, R.C. Structure and pharmacology of elapid cytotoxins. Pharmacol. Ther. 1988, 36, 1–40. [Google Scholar] [CrossRef]Grant, G.A.; Chiappinelli, V.A. kappa-Bungarotoxin: Complete amino acid sequence of a neuronal nicotinic receptor probe. Biochemistry 1985, 24, 1532–1537. [Google Scholar] [CrossRef]Dewan, J.C.; Grant, G.A.; Sacchettini, J.C. Crystal structure of kappa-bungarotoxin at 2.3-A resolution. Biochemistry 1994, 33, 13147–13154. [Google Scholar] [CrossRef]Chang, L.-S.S.; Chung, C.; Lin, J.; Hong, E. Organization and phylogenetic analysis of kappa-bungarotoxin genes from Bungarus multicinctus (Taiwan banded krait). Genetica 2002, 115, 213–221. [Google Scholar] [CrossRef]Chang, L.; Lin, S.; Huang, H.; Hsiao, M. Genetic organization of alpha-bungarotoxins from Bungarus multicinctus (Taiwan banded krait): Evidence showing that the production of alpha-bungarotoxin isotoxins is not derived from edited mRNAs. Nucleic Acids Res. 1999, 27, 3970–3975. [Google Scholar] [CrossRef]Chang, L.; Lin, J.; Wu, P.; Chang, C.; Hong, E. cDNA sequence analysis and expression of kappa-bungarotoxin from Taiwan banded krait. Biochem. Biophys. Res. Commun. 1997, 230, 192–195. [Google Scholar] [CrossRef]Danse, J.M.; Garnier, J.M. cDNA deduced amino-acid sequences of two novel kappa-neurotoxins from Bungarus multicinctus. Nucleic Acids Res. 1990, 18, 1050. [Google Scholar] [CrossRef]Koludarov, I.; Sunagar, K.; Undheim, E.A.B.; Jackson, T.N.W.; Ruder, T.; Whitehead, D.; Saucedo, A.C.; Mora, G.R.; Alagon, A.C.; King, G.; et al. Structural and molecular diversification of the Anguimorpha lizard mandibular venom gland system in the arboreal species Abronia graminea. J. Mol. Evol. 2012, 75, 168–183. [Google Scholar] [CrossRef]Jansa, S.A.; Voss, R.S. Adaptive evolution of the venom-targeted vWF protein in opossums that eat pitvipers. PLoS One 2011, 6, e20997. [Google Scholar] [CrossRef]Voss, R.S. Opossums (Mammalia: Didelphidae) in the diets of Neotropical pitvipers (Serpentes: Crotalinae): Evidence for alternative coevolutionary outcomes? Toxicon 2013, 66, 1–6. [Google Scholar] [CrossRef]Nisani, Z.; Boskovic, D.S.; Dunbar, S.G.; Kelln, W.; Hayes, W.K. Investigating the chemical profile of regenerated scorpion (Parabuthus transvaalicus) venom in relation to metabolic cost and toxicity. Toxicon 2012, 60, 315–323. [Google Scholar] [CrossRef]Nisani, Z.; Dunbar, S.G.; Hayes, W.K. Cost of venom regeneration in Parabuthus transvaalicus (Arachnida: Buthidae). Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 2007, 147, 509–513. [Google Scholar] [CrossRef]Morgenstern, D.; King, G.F. The venom optimization hypothesis revisited. Toxicon 2013, 63, 120–128. [Google Scholar] [CrossRef]Kini, R.M.; Chan, Y.M. Accelerated evolution and molecular surface of venom phospholipase A2 enzymes. J. Mol. Evol. 1999, 48, 125–132. [Google Scholar] [CrossRef]Delpietro, H.A.; Russo, R.G. Acquired resistance to saliva anticoagulants by prey previously fed upon by vampire bats (Desmodus rotundus): Evidence for immune response. J. Mamm. 2009, 90, 1132–1138. [Google Scholar] [CrossRef]Pillet, L.; Trémeau, O.; Ducancel, F.; Drevet, P.; Zinn-Justin, S.; Pinkasfeld, S.; Boulain, J.C.; Ménez, A. Genetic engineering of snake toxins. Role of invariant residues in the structural and functional properties of a curaremimetic toxin, as probed by site-directed mutagenesis. J. Biol. Chem. 1993, 268, 909–916. [Google Scholar]Antil, S.; Servent, D.; Menez, A. Variability among the sites by which curaremimetic toxins bind to torpedo acetylcholine receptor, as revealed by identification of the functional residues of alpha-cobratoxin. J. Biol. Chem. 1999, 274, 34851–34858. [Google Scholar] [CrossRef]Antil-Delbeke, S.; Gaillard, C.; Tamiya, T.; Corringer, P.J.; Changeux, J.P.; Servent, D. Ménez, a Molecular determinants by which a long chain toxin from snake venom interacts with the neuronal alpha 7-nicotinic acetylcholine receptor. J. Biol. Chem. 2000, 275, 29594–29601. [Google Scholar]Trémeau, O.; Lemaire, C.; Drevet, P.; Pinkasfeld, S.; Ducancel, F.; Boulain, J.C.; Ménez, A. Genetic engineering of snake toxins. The functional site of Erabutoxin a, as delineated by site-directed mutagenesis, includes variant residues. J. Biol. Chem. 1995, 270, 9362–9369. [Google Scholar]Heyborne, W.H.; Mackessy, S.P. Identification and characterization of a taxon-specific three-finger toxin from the venom of the Green Vinesnake (Oxybelis fulgidus; family Colubridae). Biochimie 2013, 95, 1923–1932. [Google Scholar] [CrossRef]Gotz, S.; Garcia-Gomez, J.M.; Terol, J.; Williams, T.D.; Nagaraj, S.H.; Nueda, M.J.; Robles, M.; Talon, M.; Dopazo, J.; Conesa, A. High-throughput functional annotation and data mining with the Blast2GO suite. Nucleic Acids Res. 2008, 36, 3420–3435. [Google Scholar] [CrossRef]NCBI. Available online: www.ncbi.nlm.nih.gov (accessed on 14 November 2013).Altschul, S.F.; Madden, T.L.; Schaffer, A.A.; Zhang, J.; Zhang, Z.; Miller, W.; Lipman, D.J. Gapped BLAST and PSI-BLAST: A new generation of protein database search programs. Nucleic Acids Res. 1997, 25, 3389–3402. [Google Scholar] [CrossRef]Edgar, R.C. MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004, 32, 1792–1797. [Google Scholar] [CrossRef]Ronquist, F.; Teslenko, M.; van der Mark, P.; Ayres, D.L.; Darling, A.; Hohna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck, J.P. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef]Whelan, S.; Goldman, N. A general empirical model of protein evolution derived from multiple protein families using a maximum-likelihood approach. Mol. Biol. Evol. 2001, 18, 691–699. [Google Scholar] [CrossRef]Guindon, S.; Dufayard, J.F.; Lefort, V.; Anisimova, M.; Hordijk, W.; Gascuel, O. New algorithms and methods to estimate maximum-likelihood phylogenies: Assessing the performance of PhyML 3.0. Syst. Biol. 2010, 59, 307–321. [Google Scholar] [CrossRef]Posada, D.; Crandall, K.A. The effect of recombination on the accuracy of phylogeny estimation. J. Mol. Evol. 2002, 54, 396–402. [Google Scholar]Delport, W.; Poon, A.F.; Frost, S.D.; Kosakovsky Pond, S.L. Datamonkey 2010: A suite of phylogenetic analysis tools for evolutionary biology. Bioinformatics 2010, 26, 2455–2457. [Google Scholar] [CrossRef]Kosakovsky Pond, S.L.; Posada, D.; Gravenor, M.B.; Woelk, C.H.; Frost, S.D.W.; Pond, S.L.K. Automated phylogenetic detection of recombination using a genetic algorithm. Mol. Biol. Evol. 2006, 23, 1891–1901. [Google Scholar] [CrossRef]Goldman, N.; Yang, Z. A codon-based model of nucleotide substitution for protein-coding DNA sequences. Mol. Biol. Evol. 1994, 11, 725–736. [Google Scholar]Yang, Z. Likelihood ratio tests for detecting positive selection and application to primate lysozyme evolution. Mol. Biol. Evol. 1998, 15, 568–573. [Google Scholar] [CrossRef]Yang, Z. PAML 4: Phylogenetic analysis by maximum likelihood. Mol. Biol. Evol. 2007, 24, 1586–1591. [Google Scholar] [CrossRef]Zhang, J.; Nielsen, R.; Yang, Z. Evaluation of an improved branch-site likelihood method for detecting positive selection at the molecular level. Mol. Biol. Evol. 2005, 22, 2472–2479. [Google Scholar] [CrossRef]Yang, Z.; Nielsen, R. Codon-substitution models for detecting molecular adaptation at individual sites along specific lineages. Mol. Biol. Evol. 2002, 19, 908–917. [Google Scholar] [CrossRef]Nielsen, R.; Yang, Z. Likelihood models for detecting positively selected amino acid sites and applications to the HIV-1 envelope gene. Genetics 1998, 148, 929–936. [Google Scholar]Yang, Z.; Wong, W.S.W.; Nielsen, R. Bayes empirical Bayes inference of amino acid sites under positive selection. Mol. Biol. Evol. 2005, 22, 1107–1118. [Google Scholar] [CrossRef]Murrell, B.; Moola, S.; Mabona, A.; Weighill, T.; Sheward, D.; Kosakovsky Pond, S.L.; Scheffler, K. FUBAR: A fast, unconstrained bayesian approximation for inferring selection. Mol. Biol. Evol. 2013, 30, 1196–1205. [Google Scholar] [CrossRef]Pond, S.L.K.; Frost, S.D.W.; Muse, S.V. HyPhy: Hypothesis testing using phylogenies. Bioinformatics 2005, 21, 676–679. [Google Scholar] [CrossRef]Murrell, B.; Wertheim, J.O.; Moola, S.; Weighill, T.; Scheffler, K.; Kosakovsky Pond, S.L. Detecting individual sites subject to episodic diversifying selection. PLoS Genet. 2012, 8, e1002764. [Google Scholar] [CrossRef]Woolley, S.; Johnson, J.; Smith, M.J.; Crandall, K.A.; McClellan, D.A. TreeSAAP: Selection on amino acid properties using phylogenetic trees. Bioinformatics 2003, 19, 671–672. [Google Scholar] [CrossRef]Bielawski, J.P.; Yang, Z. A maximum likelihood method for detecting functional divergence at individual codon sites, with application to gene family evolution. J. Mol. Evol. 2004, 59, 121–132. [Google Scholar]Pond, S.L.K.; Scheffler, K.; Gravenor, M.B.; Poon, A.F.Y.; Frost, S.D.W. Evolutionary fingerprinting of genes. Mol. Biol. Evol. 2010, 27, 520–536. [Google Scholar] [CrossRef]Kelley, L.A.; Sternberg, M.J.E. Protein structure prediction on the Web: A case study using the Phyre server. Nat. Protoc. 2009, 4, 363–371. [Google Scholar] [CrossRef]DeLano, W.L. The PyMOL Molecular Graphics System; DeLano Scientific: San Carlos, CA, USA, 2002. [Google Scholar]Armon, A.; Graur, D.; Ben-Tal, N. ConSurf: An algorithmic tool for the identification of functional regions in proteins by surface mapping of phylogenetic information. J. Mol. Biol. 2001, 307, 447–463. [Google Scholar] [CrossRef]Fraczkiewicz, R.; Braun, W. Exact and efficient analytical calculation of the accessible surface areas and their gradients for macromolecules. J. Comput. Chem. 1998, 19, 319–333. [Google Scholar] [CrossRef] Supplementary Files Supplementary File 1:

Supplementary (ZIP, 3550 KB)

Share and Cite MDPI and ACS Style

Sunagar, K.; Jackson, T.N.W.; Undheim, E.A.B.; Ali, S.A.; Antunes, A.; Fry, B.G. Three-Fingered RAVERs: Rapid Accumulation of Variations in Exposed Residues of Snake Venom Toxins. Toxins 2013, 5, 2172-2208. https://doi.org/10.3390/toxins5112172

AMA Style

Sunagar K, Jackson TNW, Undheim EAB, Ali SA, Antunes A, Fry BG. Three-Fingered RAVERs: Rapid Accumulation of Variations in Exposed Residues of Snake Venom Toxins. Toxins. 2013; 5(11):2172-2208. https://doi.org/10.3390/toxins5112172

Chicago/Turabian Style

Sunagar, Kartik, Timothy N. W. Jackson, Eivind A. B. Undheim, Syed. A. Ali, Agostinho Antunes, and Bryan G. Fry. 2013. "Three-Fingered RAVERs: Rapid Accumulation of Variations in Exposed Residues of Snake Venom Toxins" Toxins 5, no. 11: 2172-2208. https://doi.org/10.3390/toxins5112172

Find Other Styles Article Metrics No No Article Access Statistics For more information on the journal statistics, click here. Multiple requests from the same IP address are counted as one view. Supplementary Material Supplementary File 1:

Supplementary (ZIP, 3550 KiB)

clear Zoom | Orient | As Lines | As Sticks | As Cartoon | As Surface | Previous Scene | Next Scene Cite Export citation file: BibTeX | EndNote | RIS MDPI and ACS Style

Sunagar, K.; Jackson, T.N.W.; Undheim, E.A.B.; Ali, S.A.; Antunes, A.; Fry, B.G. Three-Fingered RAVERs: Rapid Accumulation of Variations in Exposed Residues of Snake Venom Toxins. Toxins 2013, 5, 2172-2208. https://doi.org/10.3390/toxins5112172

AMA Style

Sunagar K, Jackson TNW, Undheim EAB, Ali SA, Antunes A, Fry BG. Three-Fingered RAVERs: Rapid Accumulation of Variations in Exposed Residues of Snake Venom Toxins. Toxins. 2013; 5(11):2172-2208. https://doi.org/10.3390/toxins5112172

Chicago/Turabian Style

Sunagar, Kartik, Timothy N. W. Jackson, Eivind A. B. Undheim, Syed. A. Ali, Agostinho Antunes, and Bryan G. Fry. 2013. "Three-Fingered RAVERs: Rapid Accumulation of Variations in Exposed Residues of Snake Venom Toxins" Toxins 5, no. 11: 2172-2208. https://doi.org/10.3390/toxins5112172

Find Other Styles clear Toxins, EISSN 2072-6651, Published by MDPI RSS Content Alert Further Information Article Processing Charges Pay an Invoice Open Access Policy Contact MDPI Jobs at MDPI Guidelines For Authors For Reviewers For Editors For Librarians For Publishers For Societies For Conference Organizers MDPI Initiatives Sciforum MDPI Books Preprints.org Scilit SciProfiles Encyclopedia JAMS Proceedings Series Follow MDPI LinkedIn Facebook Twitter MDPI © 1996-2023 MDPI (Basel, Switzerland) unless otherwise stated Disclaimer Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. Terms and Conditions Privacy Policy We use cookies on our website to ensure you get the best experience. Read more about our cookies here. Accept Share Link Copy clear Share https://www.mdpi.com/61038 clear Back to TopTop


【本文地址】

公司简介

联系我们

今日新闻

    推荐新闻

    专题文章
      CopyRight 2018-2019 实验室设备网 版权所有