Ecological Venomics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Proteomics 135 (2016) 62–72

Contents lists available at ScienceDirect

Journal of Proteomics

journal homepage: www.elsevier.com/locate/jprot

Ecological venomics: How genomics, transcriptomics and proteomics can


shed new light on the ecology and evolution of venom
Kartik Sunagar a, David Morgenstern b,⁎, Adam M. Reitzel c, Yehu Moran a,⁎
a
Department of Ecology, Evolution and Behavior, Alexander Silberman Institute of Life Sciences, Hebrew University of Jerusalem, Jerusalem 91904, Israel
b
Proteomics Resource Center, Langone Medical Center, New York University, New York, USA
c
Department of Biological Sciences, University of North Carolina at Charlotte, Charlotte, NC 28223, USA

a r t i c l e i n f o a b s t r a c t

Article history: Animal venom is a complex cocktail of bioactive chemicals that traditionally drew interest mostly from biochem-
Received 15 July 2015 ists and pharmacologists. However, in recent years the evolutionary and ecological importance of venom is real-
Received in revised form 2 September 2015 ized as this trait has direct and strong influence on interactions between species. Moreover, venom content can
Accepted 9 September 2015
be modulated by environmental factors. Like many other fields of biology, venom research has been revolution-
Available online 15 September 2015
ized in recent years by the introduction of systems biology approaches, i.e., genomics, transcriptomics and prote-
Keywords:
omics. The employment of these methods in venom research is known as ‘venomics’. In this review we describe
Venom the history and recent advancements of venomics and discuss how they are employed in studying venom in
Toxin general and in particular in the context of evolutionary ecology. We also discuss the pitfalls and challenges of
Genomics venomics and what the future may hold for this emerging scientific field.
Transcriptomics © 2015 Elsevier B.V. All rights reserved.
Proteomics
Venomics

1. Introduction venoms is of growing interest for the pharmacological and biotechno-


logical communities as they are increasingly recognized as a rich source
Venom is defined as a secretion, produced in a specialized gland or for lead compounds that can drive forward the development of insecti-
cell of one animal that is actively delivered to the target animal through cides and pharmaceutical drugs [10–12]. However, venom can also
the infliction of a wound. It is a complex mixture of toxin peptides, pro- serve as a model system where the relationship between genetic vari-
teins, salts and other chemicals employed mostly for prey capture and/ ability, protein biochemistry and interspecific interactions is more ame-
or for defense from predators and aggressors [1,2]. Further, venoms are nable for elucidation and definition than in most other systems. After all,
also employed by some animals such as leeches, ticks and vampire bats an increase in the potency of venom can directly increase the fitness of
to facilitate their specialized blood feeding habits [3,4]. As venoms are the predator secreting the venom and decrease the fitness of the prey or
characterized by an unusual diversity of components, and may contain vice versa, depending on which species is venomous. Thus, venom po-
hundreds of toxin peptides they generate considerable interest among tency can directly affect the strength of antagonistic interaction be-
evolutionary biologists and biochemists alike. The structural variability tween the prey and the predator. Hence, evolutionary ecology can
of animal toxins is remarkable [5] and they exhibit a large array of bio- greatly benefit from the study of venom. Unfortunately, in many cases
logical activities [1]. Genes coding these peptides are hypothesized to where much is known about venom composition we lack knowledge
evolve under positive Darwinian selection due to their participation in regarding the ecology of the venomous animal and in many cases
an evolutionary “arms race”, where the evolution of venom resistance where the ecology is well-understood little is known about venom com-
in prey and the invention of potent venom components in the secreting position. There are indications from snakes that the diet of a venomous
animal exert reciprocal selection pressures [6–8]. On the other hand, it is species might be closely-tied to its venom composition [13–15]. More-
clear now that several protein families with non-venomous functions over, it seems that cone snails employ different venom arsenals for
are recurrently and independently recruited into the venoms of differ- prey capture and defense [16] and that scorpions can control the com-
ent animal lineages [3], and that some venoms evolve mostly under pu- position of the mixture they inject via their sting [17]. It is plausible
rifying selection with only episodic positive selection [9]. The study of that such tight relationships between interspecific interactions and
venom composition might be due to the high metabolic cost of venom
⁎ Corresponding authors.
production, which makes potent venom advantageous, as less venom
E-mail addresses: venomologist@gmail.com (D. Morgenstern), is required for neutralizing prey or foe [18]. However, to better under-
yehu.moran@mail.huji.ac.il (Y. Moran). stand such selective pressures we need a much better picture of the

http://dx.doi.org/10.1016/j.jprot.2015.09.015
1874-3919/© 2015 Elsevier B.V. All rights reserved.
K. Sunagar et al. / Journal of Proteomics 135 (2016) 62–72 63

complete venom composition, its temporal dynamics and venom varia- venomous animal to be sequenced [28] (Table 1). However, it has an ex-
tion between individuals and populations as well as spatiotemporal var- tremely streamlined venom with only a handful of components [29].
iations in diet and interspecific interactions. Filling these gaps with the The next animal to have its genome sequenced was the starlet anemone
help of “omic” tools could help us understand the ecological factor and Nematostella vectensis that serves as an important model organism in
evolutionary pressures that shaped the venom and also what role the evolutionary developmental biology studies [30] (Table 1). This species
venom plays in the interspecific interactions of the venomous animal is a representative of the venomous phylum Cnidaria (sea anemones,
and its ecological niche. corals, jellyfish and hydroids), but the contents of its venom were un-
The term “venomics” was first used as a description for proteomic known before its genome was sequenced [31]. Thus, this species provid-
study of snake venom composition [19,20], but in 2006 as a description ed the first example for a genome-guided venom discovery [32,33].
of an ambitious project aimed to provide a full picture of venom-related Following Nematostella, the genomes of other cnidarians such as Hydra
biology by sequencing the full genomes, transcriptomes of venom magnipapillata and the reef-building coral Acropora digitifera were
glands and venom protein contents of several venomous animal species sequenced as well [34,35] (Table 1). Another genome of a venomous an-
[21]. Since then it has also been used for describing studies of much imal to be sequenced by the Sanger method is that of platypus
smaller scales that use several systems biology approaches for the com- (Ornithorhynchus anatinus), which uses its venom for intraspecific ag-
prehensive investigation of venom components, e.g., studies combining gression among males during the mating season [36]. Lastly, the full ge-
shotgun mass spectrometry analysis of venom components with tran- nome of the parasitic wasp Nasonia vitripennis was sequenced by the
scriptome sequencing of a venom gland for charting an extensive list same method as well as partial genomic sequence of two other Nasonia
of venom contents and getting a more complete picture on the proteo- species [37]. The females of these tiny wasp species inject venom not
mic landscape of venoms and their biology [22,23]. The field of only for paralyzing host fly larvae, but also for manipulating the host's
venomics is growing fast but even with the current availability of new gene regulation, possibly for making it more suitable for the needs of
high-throughput methods, the de novo sequencing of genomes, their progeny [38]. The sequencing of the Nasonia genome provided
transcriptomes and proteomes is far from an easy feat. In this perspec- an important tool in the study of their venom that can be obtained
tive paper we will discuss the current state of knowledge, the pitfalls only in miniscule amounts, and hopefully provides an important step
and challenges awaiting for studies in this field and what are the future in the direction of understanding how these complex genetic manipula-
directions that could help it grow in the context of evolutionary ecology. tions are achieved [39].
The introduction of high throughput sequencing techniques, also
2. Genomics and transcriptomics in venom research known as Next Generation Sequencing (NGS), in the last decade revolu-
tionized the field of genomics. Initially, pyrosequencing technology
2.1. Genomics of venomous animals platforms, such as those by 454 and then reversible terminator technol-
ogy platforms such as the ones from Illumina, made sequencing of mil-
The DNA sequencing methods developed in parallel by Sanger and lions of bps by a single run possible [40]. In recent years NGS outputs are
Gilbert brought a true revolution to the field of biology [24,25]. Roughly booming, for example, at the time of writing this manuscript (July
25 years later in a remarkable project by a large consortium and an in- 2015), the Illumina HiSeq 4000 sequencing system offered the ability
vestment of more than 3 billion US dollars and 10 years of work, the to sequence 1500 gigabases per run, i.e., to sequence about 12 human
full human genome was sequenced by applying the method developed genomes at a ×30 coverage in one run (Table 2). This second genomic
by Sanger [26]. As this method is costly and relatively slow, using it to revolution has had a profound effect on all fields of biology as sequenc-
fully sequence a genome of several hundreds of millions of base pairs ing costs per bp are constantly decreasing. The NGS technologies en-
(bp) would cost several millions of US dollars and would take several abled the sequencing of the genomes of the scorpion Mesobuthus
years. However, a full-genome sequence of an organism is priceless for martensii[41] and the king cobra (Ophiophagus hannah) [42] (Table 1).
biologists as it can give the full repertoire of genes, not only those that Sequencing of the full genomes of these two animals coupled to RNA-
are actively transcribed, but also the many additional genetic features Seq (the sequencing of expressed protein-coding genes via complemen-
such as introns, intergenic regions and cis- and trans-transcriptional tary DNA) revealed the nearly-full landscape of venom-encoding genes
regulatory elements that have pivotal roles in the control of gene ex- in these animals. In addition, the genome sequences provided novel
pression and in the evolution and physiology of an organism [27]. Still, insights into several evolutionary and zoological aspects related to
the astounding costs have put a severe limitation on the ability to venom: from the ability of scorpions to resist their own venom to a pos-
sequence full animal genomes and very few genomes of venomous an- sible answer to the enigmatic origin of venom glands in snakes [41,42].
imals have been sequenced to date. As an organism of pivotal agricultur- NGS also enabled the sequencing of the first two spider genomes
al importance the European honeybee Apis mellifera was the first [43] (Table 1). The vast majority of known spiders are venomous, and

Table 1
Sequenced genomes of venomous animals.

Organism Approximate genome size Sequencing platform Sequencing depth Scaffold N50 References for
(million base pairs) (see Table 2 for details) (thousand base pairs) genome and venom

Arthropods
Acanthoscurria geniculata (tarantula) 6500 Illumina 40× 47 [43]
Apis mellifera (honey bee) 262 Originally Sanger, later 6× (Sanger); 20× (SOLiD); Originally 359, [28,29,73]
454 and SOLiD 4× (454) later 997
Mesobuthus martensii (scorpion) 1323 Illumina 248× 223 [41,171]
Nasonia vitripennis (parasitoid wasp) 295 Sanger 6× 709 [37,39]
Solenopsis invicta (fire ant) 352 Illumina and 454 ? 720 [46,47]
Stegodyphus mimosarum (spider) 2550 Illumina 91× 480 [43]
Cnidaria
Hydra magnipapillata 1050 Sanger 8× 92.5 [34,172,173]
Nematostella vectensis (sea anemone) 357 Sanger 6.5× 470 [30,32,174]
Vertebrates
Ophiophagus hannah (king cobra) 1590 Illumina ? 226 [42]
Ornithorhynchusanatinus(platypus) 1840 Sanger 6× 967 [36,175]
64 K. Sunagar et al. / Journal of Proteomics 135 (2016) 62–72

Table 2
Major Next Generation Sequencing (NGS) platforms.

Platform Systems Sequencing Detection Run time Read length Reads per run Output
(in base pairs)

Illumina (previously Solexa) HiSeq 3000/HiSeq 4000 Synthesis Fluorescence b1–3.5 days ≥75% of bases Above Q30 5 billion 1.5 Tb
at 2 × 150 bp
Miseq Synthesis Fluorescence 56 h ≥70% of bases above Q30 25 million 13.2–15 Gb
at 2 × 300 bp
Life Technologies Ion torrent Synthesis Proton release 2h 200–400 4 million 1.5–2 Gb
(previously Invitrogen) Abi/solid Ligation Fluorescence 10 days 2 × 50 2.4 billion 240 Gb
Proton Synthesis Proton release 4h 125 60–80 million 8–10 Gb
Roche (previously 454) GS FLX+ Synthesis Pyrophosphate 23 h Up to 1000 1 million 700 Mb
GS Junior System Synthesis Pyrophosphate 10 h 700 0.1 million 40 Mb
Pacific Biosciences PacBio RS II Synthesis: one Fluorescence up to 240 min 10,000–15,000 0.8 million ~500 Mb–1 Gb
molecule at a time
Helicos Heliscope Synthesis: one Fluorescence 30 days 25,000–35,000 500 million 35 Gb
molecule at a time

Legend: bp: base pair; Mb: megabase pairs; Gb: gigabase pairs; Q: quality score.

the sequencing of the genomes of the African social velvet spider, composition and evolution of venom in various poorly studied venom-
Stegodyphus mimosarum, and the Brazilian white-knee tarantula, ous animals, such as sea anemones, polychaete worms, remipede crus-
Acanthoscurria geniculata provided the first in-depth data regarding taceans, ticks and platypus [53–57]. The transcriptomic approach is
venom genomics in spiders and the proteases that play an important especially important when proteomics are not feasible due to limited
role in the maturation process of the small cysteine-rich peptides that venom accessibility. As reviewed by von Reumont et al. in 2014 [58],
form the major groups of toxins in their venom. there are multiple phyla whose venom is inaccessible for proteomic in-
NGS was recently applied in an unprecedented scale for the se- vestigation due to the very limited size or number of venom glands or
quencing of the genomes of more than a dozen hymenopteran species the inability to extract venom from them. In such cases the fact that
due to the great interest in their agricultural role, phenotypic plasticity transcriptomics incorporate PCR-based amplification steps is invalu-
and in the evolution of eusociality [44,45]. It is likely that many of able, even if it comes at the price of less accurate quantification due to
these species are venomous, yet practically nothing is known about amplification biases.
their venom, with the notable exception of the fire ant, Solenopsis Venom-gland transcriptomics involve the sequencing of mRNAs ex-
invicta[46,47]. tracted from the venom gland to reveal a snapshot of the gland's tran-
While genome sequences provide the full genetic data of a venom- scriptome at a given time point, which reflects the recent expression
ous organism, they are not sufficient for determining the exact gene of genes encoding venom components. This strategy not only permits
boundaries and the cell and tissue expression dynamics and specificity the identification of novel bioactive components in venom but also in
of genes. For addressing these questions transcriptomic and proteomic understanding how each toxin family varies between species or within
approaches are required. various populations of the same species. It can further shed light into
shifts in toxin expression profiles across developmental stages. Indeed,
2.2. Venom transcriptomics ontogenic or developmental stage associated variation in venom has
also been studied by sequencing the venom gland transcriptomes of
Transcriptomics show which transcripts are currently present in a various developmental-stages in snakes [59–61]. Application of integra-
sample and hence enable the identification of genes that are transcrip- tive omics technologies in the central American rattlesnake (Crotalus
tionally active (or at least been recently active) in a given animal, tissue simus simus) revealed an ontogenic shift in venom expression [59]. Pro-
or cell. Further, transcriptomics also reveal untranslated regions (UTRs), teomic examinations (approaches discussed below) in this study
which can carry important regulatory roles and are many times difficult identified ontogenetic shifts in the venom that involved a gradual re-
to annotate accurately from genomic data. Early studies of venom duction in the expression of certain toxins, such as serine proteinases
transcriptomes employed the preparation of complementary DNA and PLA2s. When these snakes reached an age of 9 to 18 months, an in-
(cDNA) libraries by generating a double stranded cDNA and its cloning creased secretion of PI and PIII metalloproteinases was observed. Deep
into a plasmid vector, followed by sequencing by the Sanger method. sequencing of small RNAs in the venom glands of various developmen-
The resulting sequences are known as Expressed Sequence Tags tal stages identified several microRNAs (miRNAs) that were found to be
(ESTs). Despite its limited sequencing depth and relative cost inefficien- complementary to the transcripts of various venom protein encoding
cy compared with NGS approaches, this method is still occasionally mRNAs. Hence, the authors speculated that miRNAs may participate in
employed as it manages to give an initial insight into the venom content regulating the venom gland expression profiles of C. s. simus, and thus
of animals, as well as sheds light on the venom's physiological and bio- result in ontogenetic-shifts in venom-composition. However, since
chemical roles [48–51]. However, RNA-Seq methods are taking over the miRNAs are mainly known to fine-tune expression profiles in most
whole field of transcriptomics, including venom transcriptomics, as other animals and they affect gene expression and protein output by
they are much more cost-effective due to their high-throughput nature less than 2-fold [62,63], the actual contribution of miRNAs in introduc-
[52]. Transcriptomics are also used many times instead of genomics as ing variation in snake venom composition remains to be experimentally
they cost only a small fraction of the latter due to the high intronic validated.
and intergenic content of most eukaryotic genomes. Moreover, the Transcriptomics can also permit the understanding of how various
assembly of a transcriptome is a much simpler bioinformatic task factors influence the composition of venoms, and ultimately the ecology
compared to genome assembly. of venomous animals. Specifically, there are strong indications from
Transcriptomics enables the identification of various domains of the several studies in venomics that various ecological and environmental
transcripts including signal peptide, pro-peptide and the pre-pro factors, such as diet [15,64], gender [64,65], geographical distribution
peptide domains that are rarely caught at the proteomic level due to [15,66], and temperature and UV radiation [67] may influence venom
the fact that they are cleaved very quickly. Moreover, venom gland composition in animals. Variability in venom is not only observed across
transcriptomic investigations have provided novel insights into the species, but within species as well, where different populations utilize
K. Sunagar et al. / Journal of Proteomics 135 (2016) 62–72 65

different venom proteins for prey incapacitation [14,68]. An intriguing and assemble them into distinct transcripts. However, setting the
question is what part of the variability in venom measured in individ- threshold in a way that will ideally balance between real variation and
uals from the same species is derived from genetic variation and how erroneous reads remains a challenge.
much is due to plasticity of gene expression in response to environmen- A possible solution for assembling longer genomes is the combina-
tal experience. Like any quantitative character, the particular phenotype tion of Sanger sequencing and NGS: the Sanger sequencing provides
(here, venom composition) potentially reflects both: certain popula- longer reads for bridging, but at a much higher cost per bp, whereas
tions may have locally adapted venom composition resulting from ge- the NGS provides numerous short reads at a much cheaper price. The
netic adaptation over numerous generations (i.e., genotypic variation) honeybee genome was recently re-assembled from a combination of
and individuals within each population may have the molecular mech- reads from various sequencing methods, resulting in much longer and
anism to modulate venom composition based on within lifetime experi- informative genomic scaffolds [73] (Table 1). Another solution is the
ence (i.e., environmental variation). Moreover, individuals within a production of several distinct NGS libraries with vastly different insert
species may differ in the relative extent of modulations to their venom. sizes. This approach produced a well-assembled spider genome that
Venom gland transcriptomics of various populations of a species can was sequenced exclusively by the Illumina HiSeq technology [43]
help in revealing such differences in venom compositions. For example, (Table 2). Lastly, some of the NGS platforms like Heliscope, PacBio
a combined proteomic and transcriptomic investigation of the venom (Table 2) or MinION [74] can provide extremely long reads that can
glands of various populations of Crotalus oreganus helleri (the Southern prove very useful when combined with short read data from other plat-
Pacific Rattlesnake), one of the most medically significant snakes in all forms [75,76]. These platforms also allow to sequence single molecules,
of North America, revealed remarkably different venom compositions, enabling sequencing of full transcripts up to 10 kb in length [77], a qual-
despite relatively short geographical distribution of the examined pop- ity that may prove extremely useful for documenting the true variability
ulations [69]. While hemorrhagic and tissue destroying toxins, such as of toxin-encoding transcripts, including specific alleles with all linked
the snake venom metalloproteinases (SVMPs) were abundantly secret- polymorphisms.
ed by the Catalina Island and Phelan populations, the Loma Linda and Another major challenge in genomics of eukaryotes is annotating full
the Idyllwild populations secreted SVMPs in moderate and trace genes, from their transcription start site (TSS) to the last nucleotide,
amounts, respectively. Moreover, the presynaptic neurotoxic PLA2 com- while encompassing all exons, including the alternative ones in splice
plex, previously only identified in the Mohave rattlesnake (Crotalus variants. In the past a major part of gene annotation was based on pre-
scutulatus) and the neotropical rattlesnake (Crotalus durissus terrificus) diction algorithms such as Genscan that used mostly common patterns
were identified in the Idyllwild population of the San Jacinto Mountains. in genomic sequences that typify exons-introns boundaries [78]. The
Thus, transcriptomic and proteomic examinations revealed why combination of transcriptomics and genomics for gene prediction and
treating envenomations by this species has been a clinician's nightmare. annotation leads to a significant improvement in accuracy, and there
Venom transcriptome sequencing and evolutionary assessments also is no doubt that for this task NGS data is priceless, as a single sequencing
identified variations in the molecular evolutionary rates of various run can cover even the rarest splice variants. However, how to combine
toxin families in distinct populations [69]. all these data together in order to map full genes and to avoid artifacts is
To fully discern the effects of genotype and environment in the com- not a simple task and various softwares that employ distinct approaches
position of venom, the most straightforward approach would be to to reach this aim are available [79]. Beyond the annotation of physical
compare the venom of genetically identical individuals exposed to a gene structure, annotating function is of paramount importance.
range of environmental conditions. The use of clonal individuals like Genomes of model organisms such as yeast, flies, nematodes and mice
those available for cnidarians or male hymenopterans removes genetic provide the gold-standard for functional annotation, as genes in these
variation from any measured transcriptomic and/or proteomic response organisms were studied at extreme depth, the function of most of
in the phenotype. Then, by comparing this response in multiple clonal their genes (but not all) is presently known. Further, the large research
lines from a single species, one could quantify the genotype vs. environ- communities using these organisms make the functional annotations
ment contribution to the phenotypic variation. To our knowledge, this available in well-maintained public databases [80–82]. Functional an-
experimental design has yet to be employed in the study of venomics. notation of genes in non-model organisms, such as venomous animals,
is based mostly on the homology to annotated genes of model
2.3. Challenges for venom genomics and transcriptomics: the devil is in the organisms. However, the problem is that toxic/venomous function is
assembly and annotation non-existent in classic model organisms and as many toxin genes
were recruited from genes of non-venomous function [3] they might
While NGS methods provide extraordinary amounts of data they be annotated with the wrong function. The use of manually annotated
also provide new challenges: most of these methods provide reads databases that specialize in toxins can help in the process of annotating
length of about paired-end 150–300 bp, whereas the Sanger method novel toxin genes according to sequence homology [83,84]. However,
can provide reads of about 700–1000 bp (Table 2). This has dramatic another problem arises due to the homology of toxins and the original
effect on the ability to assemble transcripts and genes, as demonstrated physiological proteins they were recruited from. Differentiating the
by the considerably shorter scaffolds obtained in many of the two apart can be difficult. An example for such difficulties can be
genome projects that employed NGS techniques rather than Sanger found in the work of Whittington et al. [57] where 454 as well as
(Table 1). Moreover, toxin genes are especially prone to mis-assembly Illumina NGS platforms were used for sequencing the venom gland of
as a single genome may contain numerous copies that may differ from platypus, and the data was compared to older Sanger data for tran-
one another by only a handful of nucleotides [33]. This is extremely scripts of this animal's bill, brain, liver, spleen, and testis. The idea was
problematic as native single amino variations in a toxin can result in to use the data from the non-venom-related tissues to exclude physio-
dramatically different pharmacological activities [70]. Assembly of logical genes that are not toxins. The problem is that venom genes, as
venom gland transcriptomes was shown to be especially prone to this least in early evolutionary stages after recruitment might be expressed
typical sequence variation as most common transcriptome assemblers in both venom glands and other tissues in parallel and setting an appro-
such as Trinity [71] are designed to collapse such “minor” differences priate cutoff is challenging. Moreover, expression quantification in the
into a single transcript as slight variation in most other systems is con- platypus study was especially problematic as the Sanger data was limit-
sidered to be the result of either technical sequencing artifacts or allelic ed for the non-venomous tissues. The authors arbitrarily decided that
variation that is unimportant in the context of generating a reference only genes which are expressed in three or more non-venomous tissues
transcriptome. Hence, it is important to develop assembly tools such are not toxin-encoding genes. The problematic aspects of this study
as VTBuilder [72], which is specifically built to capture those differences were criticized by Terrat and Ducancel [85] and in this correspondence
66 K. Sunagar et al. / Journal of Proteomics 135 (2016) 62–72

the authors discussed what possible criteria can or cannot be used to an- approach provided a unique and exciting opportunity for comparative
notate a toxin based on nucleotide data and homology. Another study analyses of whole venoms, rather than concentrating on specific toxin
[86], which encountered similar problems, attempted to test the com- families. Indeed, this approach continues to be popular, and with the
mon origin of venom in Toxicofera — a hypothetical clade that includes improved resolution of modern chromatography and 2D-
venomous snakes and lizards [87,88]. While the data generated in this electrophoresis a greater degree of detail can be achieved [108–111].
study was interesting and highlighted the importance of quantitative Additionally, these two analytical methods could be combined with
comparison of venom proteins in the venom gland and physiological zymography or functional assays of chromatographic fractions to
tissues, it contained questionable interpretations and conclusions. The allow the functional characterization of unique gel spots or fractions.
authors arbitrarily classified a protein as highly secreted only if it repre- These approaches enabled the onset of fascinating studies tying var-
sents the top 25% of all transcripts in at least two out of four examined ious factors to variability in venom composition — the canonical work
venom gland tissues. Any protein that did not represent the top 25% of by Daltry et al. [15], tied dietary differences in geographic populations
the transcript was considered to be playing a homeostatic function. of a pitviper species with geographic variations, sex-related venom
Thus, the authors likely ignored venom components that may be secret- variability [109,112], or a basic population proteomics analysis [113].
ed in lower amounts but could be potent enough none-the-less. More- However, this approach provides a lower resolution overview of
over, the potency of venom could be increased when components venom compared to the level achieved by sequence-based approaches
function in synergy, as was shown for several scorpion toxins [89]. (genomic or proteomic); chromatographic peaks or gel spots often
Since venom is considered as energetically expensive to secrete [18], contain multiple proteins, even under orthogonal separations, while
an extremely toxic component could be secreted in limited amounts multiple peaks and gel spots may contain various proteoforms [114]
in the venom gland. Hence, it becomes essential to employ toxicity and user-induced artifacts (oxidations, carbamylations and degrada-
assays that can reveal whether a particular component secreted in tion). Moreover, successful experiments require large quantities
what seems to be an insignificant amount actually participates in the of starting material, biasing early studies to organisms producing
envenoming process or not. such quantities (i.e., snakes). Thus, qualitative and quantitative analyses
Another challenge with annotating transcripts and genes as toxin- of sequence variants and their phylogenetic relationships require
encoding is the fact that the sequenced animal may have novel toxin higher sensitivity and considerably higher resolution, such as mass
types that are non-homologous to previously studied toxins. Sequenc- spectrometry.
ing by low coverage (36,864 ESTs) the transcriptome of nematocytes
(the cnidarian stinging cells) of Hydra revealed that at least 15% of the 3.2. Venomics and mass spectrometry
clusters did not have any significant homology to a previously known
protein [90]. Along the same lines, out of 20 proteins which are released The pivotal work of Tanaka et al. demonstrated the ability to identify
from the nematocyst (stinging capsule) of Nematostella during dis- proteins using mass spectrometry [115]. It provided the unique ability
charge (stinging), only six had clear homologs in other organisms to improve the resolution of protein analysis down to unique
[32]. Hence, homology-based annotation of toxins from transcriptomic proteoforms (see definition at [116]), with mass spectrometry analyses
and genomic data can be insufficient for giving a full picture of the of venoms coming soon afterwards [117,118]. Mass spectrometry
venom arsenal of a venomous animal and proteomics are many times showed superiority in sensitivity and resolution compared to chroma-
essential in order to determine what proteins are actually secreted tography and electrophoresis, while providing identifiable sequence in-
from a venom gland or a venom producing cell and can be considered formation for distinct precursors. In combination with fractionation, the
as reliable toxin candidates. separation and identification of individual venom components became
feasible by mass spectrometry as it allows the observation of individual
3. Venom proteomics components in venom. Unfortunately, limitations of instrument resolu-
tion favored the study of venoms rich in small peptides such as those
The availability of genetic information allows for the identification of from scorpions and spiders [65,119,120]. In contrast, the sequencing
the footprint of evolutionary processes through the history of the of the toxins in question remained a very low throughput process, with-
species in question, either in comparison to other species or within out the availability of a genomic database. To this day, most software an-
the species population. However, genetic data provides only a part of alyzing mass spectral data require a sequence database to compare the
the picture; genomic and transcriptomic data cannot always provide spectra against (see [121] for review on interpretation of data mass
accurate quantitative data [91–93] nor can it provide insight into post- spectrometry). Lack of such databases due to high costs of genomic
translational modifications [81,94,95], which may be important for the sequencing in the early days of venom mass spectrometry made it ex-
toxin's functionality and therefore, for understanding its evolution and tremely difficult to accurately identify individual protein toxins. The
ecological role. availability of genomic databases provided the means to rapidly identify
protein groups. In most cases, these analyses are performed “Bottom-
3.1. Venom proteomics in the pre-genomic era Up” (Fig. 1): the proteome is either digested unfractionated (“shotgun”)
or following fractionation by chromatographic or electrophoretic
Modern proteomics rely heavily on the application of mass spec- means. The digests are then analyzed by the mass spectrometer and
trometry to analyze complex proteinaceous mixtures. The ability to con- the data generated are then matched against the genomic database.
vert the acquired spectra into protein identification relies almost This approach rarely provides the identification of the full toxin se-
exclusively on the availability of a genomic database [96,97]. As a result, quences, even with a fully accurate genomic database. Still, it is suffi-
venom proteomics flourished in recent years due to the affordability of cient to match MS data to associate chromatographic peaks or gel
NGS. However, evolutionary venom proteomics have relied on the spots to identify a family of toxins, providing the opportunity to give a
availability of sequence information as early as the 1960s [98] and the functional context to the variability observed in earlier studies. As a re-
main challenge in this era resulted from the difficulty of obtaining sult a flood of studies looking at intra- [122,123] and inter- [19,124–
sequence information en-masse (through de-novo sequencing of indi- 127] specific variations in venom composition from various phyla, as
vidually isolated proteins). These difficulties resulted in two distinct ap- well as ontogenic [122,128], sex [129], geography [122,130,131] and
proaches of early venom proteomics — either through sequence based diet [132] based variations, were conducted.
phylogenetic analyses [99] or through a venom-centric description of Later studies have demonstrated an additional utility of combining
their composition by low-resolution profiling through electrophoretic proteomic and genomic analyses: while traditionally proteomic analy-
[98,100–103] and chromatographic [104–107] patterns. The latter ses followed the process of genomic data generation and library
K. Sunagar et al. / Journal of Proteomics 135 (2016) 62–72 67

assembly, combining proteomics and genomics in the process of ge- not provide an alternative identification. It remains to the user to iden-
nome/transcriptome assembly, or proteogenomics, allows the identifi- tify the correct sequence of the protein in question. This requirement
cation and annotation of novel unidentified proteins [133,134], as well slows down the analysis considerably, yet it is immensely quicker
as improved genomic sequence assembly [135]. than manual de-novo sequencing.
What about automated de-novo sequencing? There are currently
3.3. Current achievements and ongoing challenges in venom proteomics: several de-novo solutions in the market, either free [143–145] or com-
top-down proteomics mercial [146]. These methods may well be effective for simple spectra,
derived from short peptides, yet they are still inadequate for the analysis
High throughput proteomics have flourished in recent years as a re- of spectra derived from larger peptides and proteins. This is not to say
sult of the availability of genomic databases and search software, yet the that de-novo sequencing has no future in Top-Down proteomics;
determination of individual intact proteoforms remains a challenge. Ad- though still in development, multiple publications have presented
vances in instrumentation and data analysis in the last decade has slow- some promise for Top-Down proteomics, either through a combination
ly brought forth the ability to characterize individual intact proteins of Bottom-Up and Top-Down data [147,148] or through a constrained
(i.e., Top-Down; Fig. 1) at a considerably faster pace than ever before, approach [149] that requires a sequence “anchor” from which the de-
and at a higher throughput (see for review [136]). This capability is novo algorithm works.
especially appealing for venom research, as even very slight variations Despite its limitations Top-Down proteomics, either in its more
in sequence or post-translational modifications (PTMs) may result in “primitive form” or the modern one, provided an immensely important
significant changes in activity (see example for the effect of point muta- finding about venom biology: the understanding that venom produc-
tions in synthetic [137] and native [70] toxins), making their identifica- tion and secretion is not monolithic and uniform, but rather a complex
tion paramount. Indeed, Top-Down venom proteomic analyses were mixture whose composition is susceptible to biochemical and behavior-
performed recently on peptidic venoms [138,139] as well as small- al modulation. Studies have hinted of this phenomenon in scorpions
protein venom [140]. Unfortunately, these experimental achievements [17,103], snake [150] and cone snails [16,123,151]. It is still unclear
highlighted the inadequacy of the downstream data analysis; proteo- what evolutionary processes drive venom modulation, but it provides
mics search engines require utmost sequence accuracy, as sequence the crucial context to evolutionary signals identified from the nucleo-
variants and unknown PTMs (compared to the database) will result in tide data.
false negative identification of the protein in question. The current solu-
tion in most cases is improving the database through the production of 3.4. Current achievements and ongoing challenges in venom proteomics: in
extensive variant libraries or individual genomic libraries. This solution situ venom proteomics
is typically not feasible for non-model organisms in general and venom-
ous animals in particular, due to the associated cost and effort required Another significant challenge facing the field of venomics is venom
in production and assembly of genomic and transcriptomic libraries for accessibility. Venom studies have concentrated so far on organisms
every organism in question. The current viable option is the use of toler- whose venom is easily accessible — first snakes and scorpions, followed
ant database search engines, such as Byonic [141] and ProSight [142], by spiders, cone snails and then a whole plethora of venomous organ-
that allow the identification of proteins that are very similar to se- isms, from vampire bats [4] to Arctic cephalopods [152]. Unfortunately,
quences in the database, yet not identical. This type of analysis provides in many of these cases the “venom” sample is really a whole tissue ly-
identification (and maybe localization) of the similar protein, yet it does sate containing the venom gland, leading to a contamination of the

Fig. 1. Typical workflows for “Bottom up” and “Top down” proteomic experiments. Typical proteomics analysis begins with protein extract, which is in our case milked or extracted venom.
The sample undergoes the reduction of disulfide bonds and cysteine alkylation to prevent the reformation of disulfide bonds to enable fragmentation in the mass spectrometer. Under a
Bottom-up approach the venom is enzymatically or chemically digested, generating short sequence fragments which are then acquired by the mass spectrometer. These short reads are
then searched against a known protein database. Identified peptides are used to reconstruct. While this method is more suitable for most mass spectrometers, as well as search engines, it
suffers from drawbacks such as incomplete coverage (illustrated by the missing green lines) and inability to reconstruct the correct proteoform state of protein (colored balls). In contrast,
in top-down proteomics, the proteins are analyzed intact, skipping sample processing and retaining any information related to the proteoform state of the sequence. However, top-down
acquisition of data requires specialized instruments, software and higher skill in experimental design. It is also less adequate for analysis of large proteins or proteins of low solubility.
68 K. Sunagar et al. / Journal of Proteomics 135 (2016) 62–72

venom with cellular proteins. Still, one may consider such contamina- quantitation using mass spectrometry to peptidic venoms. Limited
tion as the lesser evil, as most venoms are completely inaccessible Top-Down quantitation of MS1 spectra of venoms has been published
[58]. Even in the case of the extensively studies spider venoms, one previously using label free quantitation [16,123]. Quantitation of
can find many clades of spiders too small to analyze their venom venoms composed of large proteins is feasible using a veteran method
successfully — enough to prevent gaining a full picture of the Araneae — 2D gel electrophoresis, using a DIGE (Differential Gel Electrophore-
venomic landscape. Quantitatively, those inaccessible species include sis). While this method may not match the resolution of mass spectrom-
N40,000 species of spiders, ~ 10,000 species of pseudoscorpions, etry, the degree of separation for most venoms is probably sufficient.
N40,000 species of Acari and thousands of other species across the Protein identification can be performed either from a reference gel or
tree of life. from the DIGE gel itself using Bottom-up approach afterwards. Curious-
In-situ mass spectrometry, in the form of Imaging Mass Spectrome- ly, despite the obvious advantages of DIGE for venom research (espe-
try (IMS), may circumvent this problem; IMS provides a unique oppor- cially snake venom research), DIGE was mainly used to identify
tunity to provide information about the arrangement of proteins and qualitative differences between distant venoms, rather than being uti-
small molecules in the tissue, highlighting spatial specificity of mole- lized for fulfilling its quantitative potential [169,170]. Follow-up analy-
cules that will be lost otherwise in tissue extraction (for reviews see sis with Bottom-Up proteomics of individual gel spots provides the
[153,154]). However, the basic capability of this technique in analyzing protein identification to match to the quantitative information.
proteins straight off tissue provides in immensely important tool for While relative quantitative venomics is gradually taking off, this is
venom researchers to study. Both Bottom-Up [155] and Top-Down not the case for absolute quantitation. The difficulties associated with
[156] proteomics have been demonstrated in the context of imaging absolute quantitation, requiring known amounts of non-labeled or
mass spectrometry, and while it is not a mature technology as of yet, (preferentially) isotopically labeled peptide for each precursor in ques-
current developments in instrumentation and data analysis suggest tion, make this technique largely inaccessible as comparative analyses
that we may be able to perform routine off-tissue proteomics within between different toxins within a single venom or between venoms
the next decade. will require large monetary investment in the production of labeled
While in-situ sequencing may be still in the future, several exciting peptides or large-scale purification of the individual toxins in question.
studies have been published utilizing IMS to provide information of As the quantity of material tends to be a limiting factor in many
the arrangement of toxins in the secretory tissue. Undheim et al. have venomics studies, it is difficult to envision the incorporation of absolute
presented convincing evidence for spatial differentiation of toxin pro- quantitation as a staple method for venom research. Instead, lower
duction in the venom glands of centipedes due to ecological pressures accuracy methods such as densitometry, may be the most realistic
[127,157], and such differentiation was also detected in cone snail option for future quantitative analyses.
venoms due to the differentiation of the venom into scenario specific
venoms that are stored in distinct and separated compartment [16].
4. Concluding remarks
3.5. Current achievements and ongoing challenges in venom proteomics:
quantitative venom proteomics Venomous animals are usually highly dependent on their venoms
for survival, presumably placing this trait under extreme levels of selec-
One of the biggest advantages of protein-based analyses is their abil- tion. Further, venom plays an important part in the interspecific interac-
ity to convey quantitative information of various toxins in the venom tions of the animals producing it and in shaping the ecological role of a
and between individual venoms. Early studies have used indirect ap- species. There is little doubt that omic techniques are providing power-
proaches, particularly ELISA [158–160], enzymatic studies [158–162], ful instruments for understanding the evolutionary ecology of venom,
chromatography [163] and electrophoresis [164,165], providing quanti- yet like any other scientific method they should be utilized properly
tative information without the need for sequence data. However, the and their results should be interpreted with caution.
quantitative data available with these methods is of low resolution, The field of venomics is undoubtedly benefiting from the current
and cannot distinguish between individual proteins with similar activi- advances in proteomics, transcriptomics and genomics. The unique bio-
ties (although 2D zymography may come very close to it). More impor- chemical nature of venom; i.e., the fact that it contains a large array of
tantly, these methods disconnect the data derived from proteins, from sequence variants whose differences are often significant for their activ-
any evolutionary information derived from genomic or transcriptomic ity, presents a special challenge for omics analyses. Still, it appears that
data. To achieve this connection, mass spectrometry-based proteomics the field is yet to make full use of the newest technologies available.
is a key technology. Quantitative mass spectrometry has developed im- We expect that the promiscuity of high resolution and high throughput
mensely in the recent decades (see reviews [166,167]), dividing into instrumentation, along with improvements in data-analysis technolo-
two main experimental approaches — relative and absolute quantita- gies will lead to improved accessibility of these methods to venom
tion. Relative quantification is the most common approach as it is the researchers. Further, we believe that this improved availability of the
easiest to perform. It compares the abundance of the same peptide or newest methods will enable research in the near future that will provide
protein based on the spectral data of the precursor itself (MS1), its new dimensions into the evolution and ecology of venom.
fragments (MS2) or a reporter ion (chemical labeling). In most cases,
protein quantitation is performed on a digest of the sample. Unfortu- Transparency document
nately, Bottom-Up quantitative proteomics suffer from all of the prob-
lems listed in the above section dealing with Top-Down proteomics as The Transparency document associated with this article can be
many individual toxins carry a single peptide that is unique to them found in the online version.
and distinguishes them from other variants. The reliance on a single
peptide for quantitation is highly unfavorable in quantitative proteo-
mics and should be avoided. As a result, neither label-free [43] nor Acknowledgments
chemical labeling [128,168] took the same importance in venomics as
in general proteomics. KS was supported by a Marie Skłodowska-Curie Individual Fellow-
A possible solution would be to adopt a Top-Down approach for ship (654294). The work in the research group of YM is supported by
quantitation, absolving the need to identify a specific digested peptide the Israel Science Foundation grant no. 691/14. Work on cnidarian
for the quantitation of a specific toxin. Unfortunately, the technical dif- venom ecology in the labs of YM and AMR is supported by the Binational
ficulties in the analysis of proteins (N20 kDa) limit Top-Down Science Foundation grant no. 2013119.
K. Sunagar et al. / Journal of Proteomics 135 (2016) 62–72 69

References [32] Y. Moran, D. Praher, A. Schlesinger, A. Ayalon, Y. Tal, U. Technau, Analysis of soluble
protein contents from the nematocysts of a model sea anemone sheds light on
[1] N.R. Casewell, W. Wuster, F.J. Vonk, R.A. Harrison, B.G. Fry, Complex cocktails: the venom evolution, Mar. Biotechnol. (NY) 15 (2013) 329–339.
evolutionary novelty of venoms, Trends Ecol. Evol. 28 (2013) 219–229. [33] Y. Moran, H. Weinberger, J.C. Sullivan, A.M. Reitzel, J.R. Finnerty, M. Gurevitz, Con-
[2] D. Mebs, Venomous and poisonous animals: ahandbook for biologists, toxicologists certed evolution of sea anemone neurotoxin genes is revealed through analysis of
and toxinologists, physicians and pharmacists, Medpharm, 2002. the Nematostella vectensis genome, Mol. Biol. Evol. 25 (2008) 737–747.
[3] B.G. Fry, K. Roelants, D.E. Champagne, H. Scheib, J.D. Tyndall, G.F. King, et al., The [34] J.A. Chapman, E.F. Kirkness, O. Simakov, S.E. Hampson, T. Mitros, T. Weinmaier,
toxicogenomic multiverse: convergent recruitment of proteins into animal et al., The dynamic genome of hydra, Nature 464 (2010) 592–596.
venoms, Annu. Rev. Genomics Hum. Genet. 10 (2009) 483–511. [35] C. Shinzato, E. Shoguchi, T. Kawashima, M. Hamada, K. Hisata, M. Tanaka, et al.,
[4] D.H.W. Low, K. Sunagar, E.A.B. Undheim, S.A. Ali, A.C. Alagon, T. Ruder, et al., Using the Acropora digitifera genome to understand coral responses to environ-
Dracula's children: molecular evolution of vampire bat venom, J. Proteome 89 mental change, Nature 476 (2011) 320–323.
(2013) 95–111. [36] W.C. Warren, L.W. Hillier, J.A. Marshall Graves, E. Birney, C.P. Ponting, F. Grutzner,
[5] S. Mouhat, B. Jouirou, A. Mosbah, M. De Waard, J.M. Sabatier, Diversity of folds in et al., Genome analysis of the platypus reveals unique signatures of evolution, Na-
animal toxins acting on ion channels, Biochem. J. 378 (2004) 717–726. ture 453 (2008) 175–183.
[6] T.F. Duda Jr., S.R. Palumbi, Molecular genetics of ecological diversification: duplica- [37] J.H. Werren, S. Richards, C.A. Desjardins, O. Niehuis, J. Gadau, J.K. Colbourne, et al.,
tion and rapid evolution of toxin genes of the venomous gastropod conus, Proc. Functional and evolutionary insights from the genomes of three parasitoid Nasonia
Natl. Acad. Sci. U. S. A. 96 (1999) 6820–6823. species, Science 327 (2010) 343–348.
[7] K. Nakashima, T. Ogawa, N. Oda, M. Hattori, Y. Sakaki, H. Kihara, et al., Accelerated [38] E.O. Martinson, D. Wheeler, J. Wright, Mrinalini, A.L. Siebert, J.H. Werren, Nasonia
evolution of Trimeresurus flavoviridis venom gland phospholipase A2 isozymes, vitripennis venom causes targeted gene expression changes in its fly host, Mol.
Proc. Natl. Acad. Sci. U. S. A. 90 (1993) 5964–5968. Ecol. 23 (2014) 5918–5930.
[8] K. Sunagar, W.E. Johnson, S.J. O'Brien, V. Vasconcelos, A. Antunes, Evolution of [39] D.C. de Graaf, M. Aerts, M. Brunain, C.A. Desjardins, F.J. Jacobs, J.H. Werren, et al.,
CRISPs associated with toxicoferan-reptilian venom and mammalian reproduction, Insights into the venom composition of the ectoparasitoid wasp Nasonia vitripennis
Mol. Biol. Evol. 29 (2012) 1807–1822. from bioinformatic and proteomic studies, Insect Mol. Biol. 19 (Suppl. 1) (2010)
[9] M. Jouiaei, K. Sunagar, A. Federman Gross, H. Scheib, P.F. Alewood, Y. Moran, et al., 11–26.
Evolution of an ancient venom: recognition of a novel family of cnidarian toxins [40] M. Kircher, J. Kelso, High-throughput DNA sequencing—concepts and limitations,
and the common evolutionary origin of sodium and potassium neurotoxins in BioEssays 32 (2010) 524–536.
sea anemone, Mol. Biol. Evol. 32 (2015) 1598–1610. [41] Z. Cao, Y. Yu, Y. Wu, P. Hao, Z. Di, Y. He, et al., The genome of Mesobuthus
[10] J.K. Klint, S. Senff, D.B. Rupasinghe, S.Y. Er, V. Herzig, G.M. Nicholson, et al., Spider- martensii reveals a unique adaptation model of arthropods, Nat. Commun. 4
venom peptides that target voltage-gated sodium channels: pharmacological tools (2013) 2602.
and potential therapeutic leads, Toxicon 60 (2012) 478–491. [42] F.J. Vonk, N.R. Casewell, C.V. Henkel, A.M. Heimberg, H.J. Jansen, R.J. McCleary, et al.,
[11] J.J. Smith, V. Herzig, G.F. King, P.F. Alewood, The Insecticidal Potential of Venom The king cobra genome reveals dynamic gene evolution and adaptation in the
Peptides, Cell. Mol. Life Sci. (2013). snake venom system, Proc. Natl. Acad. Sci. U. S. A. 110 (2013) 20651–20656.
[12] I. Vetter, R.J. Lewis, Therapeutic potential of cone snail venom peptides [43] K.W. Sanggaard, J.S. Bechsgaard, X. Fang, J. Duan, T.F. Dyrlund, V. Gupta, et al., Spi-
(conopeptides), Curr. Top. Med. Chem. 12 (2012) 1546–1552. der genomes provide insight into composition and evolution of venom and silk,
[13] A. Barlow, C.E. Pook, R.A. Harrison, W. Wuster, Coevolution of diet and prey- Nat. Commun. 5 (2014).
specific venom activity supports the role of selection in snake venom evolution, [44] R. Bonasio, G. Zhang, C. Ye, N.S. Mutti, X. Fang, N. Qin, et al., Genomic comparison of
Proc. Biol. Sci./R. Soc. 276 (2009) 2443–2449. the ants Camponotus floridanus and Harpegnathos saltator, Science 329 (2010)
[14] T. Chijiwa, M. Deshimaru, I. Nobuhisa, M. Nakai, T. Ogawa, N. Oda, et al., Regional 1068–1071.
evolution of venom-gland phospholipase A2 isoenzymes of Trimeresurus [45] K.M. Kapheim, H. Pan, C. Li, S.L. Salzberg, D. Puiu, T. Magoc, et al., Social evolution.
flavoviridis snakes in the southwestern islands of Japan, Biochem. J. 347 (2000) Genomic signatures of evolutionary transitions from solitary to group living, Sci-
491–499. ence 348 (2015) 1139–1143.
[15] J.C. Daltry, W. Wuster, R.S. Thorpe, Diet and snake venom evolution, Nature 379 [46] D.R. Hoffman, Ant venoms, Curr. Opin. Allergy Clin. Immunol. 10 (2010) 342–346.
(1996) 537–540. [47] Y. Wurm, J. Wang, O. Riba-Grognuz, M. Corona, S. Nygaard, B.G. Hunt, et al., The ge-
[16] S. Dutertre, A.H. Jin, I. Vetter, B. Hamilton, K. Sunagar, V. Lavergne, et al., Evolution nome of the fire ant Solenopsis invicta, Proc. Natl. Acad. Sci. U. S. A. 108 (2011)
of separate predation- and defence-evoked venoms in carnivorous cone snails, Nat. 5679–5684.
Commun. 5 (2014) 3521. [48] S. Kozlov, A. Malyavka, B. McCutchen, A. Lu, E. Schepers, R. Herrmann, et al., A novel
[17] B. Inceoglu, J. Lango, J. Jing, L. Chen, F. Doymaz, I.N. Pessah, et al., One scorpion, two strategy for the identification of toxin-like structures in spider venom, Proteins 59
venoms: prevenom of Parabuthus transvaalicus acts as an alternative type of (2005) 131–140.
venom with distinct mechanism of action, Proc. Natl. Acad. Sci. U. S. A. 100 [49] B.G. Mille, S. Peigneur, R. Predel, J. Tytgat, Transcriptomic approach reveals the mo-
(2003) 922–927. lecular diversity of Hottentotta conspersus (Buthidae) venom, Toxicon 99 (2015)
[18] Z. Nisani, D.S. Boskovic, S.G. Dunbar, W. Kelln, W.K. Hayes, Investigating the chem- 73–79.
ical profile of regenerated scorpion (Parabuthus transvaalicus) venom in relation to [50] D. Morgenstern, B.H. Rohde, G.F. King, T. Tal, D. Sher, E. Zlotkin, The tale of a resting
metabolic cost and toxicity, Toxicon 60 (2012) 315–323. gland: transcriptome of a replete venom gland from the scorpion Hottentotta
[19] A. Bazaa, N. Marrakchi, M. El Ayeb, L. Sanz, J.J. Calvete, Snake venomics: com- judaicus, Toxicon 57 (2011) 695–703.
parative analysis of the venom proteomes of the Tunisian snakes Cerastes [51] E.F. Schwartz, R.C. Rodriguez de La Vega, L.D. Possani, E. Diego-Garcia, Tran-
cerastes, Cerastes vipera and Macrovipera lebetina, Proteomics 5 (2005) scriptome analysis of the venom gland of the Mexican scorpion Hadrurus gertschi
4223–4235. (Arachnida: Scorpiones), BMC Genomics 8 (2007) 119.
[20] P. Juarez, L. Sanz, J.J. Calvete, Snake venomics: characterization of protein families [52] Z. Wang, M. Gerstein, M. Snyder, RNA-Seq: a revolutionary tool for transcriptomics,
in Sistrurus barbouri venom by cysteine mapping, N-terminal sequencing, and tan- Nat. Rev. Genet. 10 (2009) 57–63.
dem mass spectrometry analysis, Proteomics 4 (2004) 327–338. [53] I.M. Francischetti, J.G. Valenzuela, J.F. Andersen, T.N. Mather, J.M. Ribeiro,
[21] A. Menez, R. Stocklin, D. Mebs, ‘Venomics’ or: the venomous systems genome pro- Ixolaris, a novel recombinant tissue factor pathway inhibitor (TFPI) from the
ject, Toxicon: Off. J. Int. Soc. Toxinol. 47 (2006) 255–259. salivary gland of the tick, Ixodes scapularis: identification of factor X and factor
[22] D. Biass, A. Violette, N. Hulo, F. Lisacek, P. Favreau, R. Stocklin, Uncovering intense Xa as scaffolds for the inhibition of factor VIIa/tissue factor complex, Blood 99
protein diversification in a cone snail venom gland using an integrative venomics (2002) 3602–3612.
approach, J. Proteome Res. 14 (2015) 628–638. [54] J. Macrander, M.R. Brugler, M. Daly, A RNA-seq approach to identify putative toxins
[23] V.L. Viala, D. Hildebrand, M. Trusch, T.M. Fucase, J.M. Sciani, D.C. Pimenta, et al., from acrorhagi in aggressive and non-aggressive Anthopleura elegantissima polyps,
Venomics of the Australian eastern brown snake (Pseudonaja textilis): detection BMC Genomics 16 (2015) 221.
of new venom proteins and splicing variants, Toxicon (2015). [55] B.M. von Reumont, A. Blanke, S. Richter, F. Alvarez, C. Bleidorn, R.A. Jenner, The first
[24] A.M. Maxam, W. Gilbert, A new method for sequencing DNA, Proc. Natl. Acad. Sci. venomous crustacean revealed by transcriptomics and functional morphology:
U. S. A. 74 (1977) 560–564. remipede venom glands express a unique toxin cocktail dominated by enzymes
[25] F. Sanger, S. Nicklen, A.R. Coulson, DNA sequencing with chain-terminating inhib- and a neurotoxin, Mol. Biol. Evol. 31 (2014) 48–58.
itors, Proc. Natl. Acad. Sci. 74 (1977) 5463–5467. [56] B.M. von Reumont, L.I. Campbell, S. Richter, L. Hering, D. Sykes, J. Hetmank,
[26] International Human Genome Sequencing C, Finishing the euchromatic sequence et al., A polychaete's powerful punch: venom gland transcriptomics of Glycera
of the human genome, Nature 431 (2004) 931–945. reveals a complex cocktail of toxin homologs, Genome Biol. Evol. 6 (2014)
[27] E.P. Consortium, The ENCODE (ENCyclopedia Of DNA Elements) project, Science 2406–2423.
306 (2004) 636–640. [57] C.M. Whittington, A.T. Papenfuss, D.P. Locke, E.R. Mardis, R.K. Wilson, S. Abubucker,
[28] Insights into social insects from the genome of the honeybee Apis mellifera, Nature et al., Novel venom gene discovery in the platypus, Genome Biol. 11 (2010) R95.
443 (2006) 931–949. [58] B. von Reumont, L. Campbell, R. Jenner, Quo vadis venomics? A roadmap to
[29] J. Gauldie, J. Hanson, F. Rumjanek, R. Shipolini, C. Vernon, The peptide components neglected venomous invertebrates, Toxins 6 (2014) 3488–3551.
of bee venom, Eur. J. Biochem. 61 (1976) 369–376. [59] J. Durban, A. Perez, L. Sanz, A. Gomez, F. Bonilla, S. Rodriguez, et al., Integrated
[30] N.H. Putnam, M. Srivastava, U. Hellsten, B. Dirks, J. Chapman, A. Salamov, et al., Sea “omics” profiling indicates that miRNAs are modulators of the ontogenetic
anemone genome reveals ancestral eumetazoan gene repertoire and genomic or- venom composition shift in the central American rattlesnake, Crotalus simus
ganization, Science 317 (2007) 86–94. simus, BMC Genomics 14 (2013) 234.
[31] Y. Moran, M. Gurevitz, When positive selection of neurotoxin genes is missing. [60] J. Durban, P. Juarez, Y. Angulo, B. Lomonte, M. Flores-Diaz, A. Alape-Giron, et al.,
The riddle of the sea anemone Nematostella vectensis, FEBS J. 273 (2006) Profiling the venom gland transcriptomes of Costa Rican snakes by 454 pyrose-
3886–3892. quencing, BMC Genomics 12 (2011) 259.
70 K. Sunagar et al. / Journal of Proteomics 135 (2016) 62–72

[61] H.L. Gibbs, L. Sanz, M.G. Sovic, J.J. Calvete, Phylogeny-based comparative analysis of [90] S. Milde, G. Hemmrich, F. Anton-Erxleben, K. Khalturin, J. Wittlieb, T.C. Bosch,
venom proteome variation in a clade of rattlesnakes (Sistrurus sp.), PLoS ONE 8 Characterization of taxonomically restricted genes in a phylum-restricted cell
(2013), e67220. type, Genome Biol. 10 (2009) R8.
[62] D. Baek, J. Villen, C. Shin, F.D. Camargo, S.P. Gygi, D.P. Bartel, The impact of [91] N.R. Casewell, S.C. Wagstaff, W. Wüster, D.A.N. Cook, F.M.S. Bolton, S.I. King, et al.,
microRNAs on protein output, Nature 455 (2008) 64–71. Medically important differences in snake venom composition are dictated by
[63] M. Selbach, B. Schwanhausser, N. Thierfelder, Z. Fang, R. Khanin, N. Rajewsky, distinct postgenomic mechanisms, Proc. Natl. Acad. Sci. U. S. A. 111 (2014)
Widespread changes in protein synthesis induced by microRNAs, Nature 455 9205–9210.
(2008) 58–63. [92] T. Maier, A. Schmidt, M. Güell, S. Kühner, A.-C. Gavin, R. Aebersold, et al., Quantifi-
[64] H.L. Gibbs, L. Sanz, J.E. Chiucchi, T.M. Farrell, J.J. Calvete, Proteomic analysis of onto- cation of mRNA and protein and integration with protein turnover in a bacterium,
genetic and diet-related changes in venom composition of juvenile and adult Mol. Syst. Biol. 7 (2011) 511.
dusky pigmy rattlesnakes (Sistrurus miliarius barbouri), J. Proteome 74 (2011) [93] R.C. Taylor, B.-J.M. Webb Robertson, L.M. Markillie, M.H. Serres, B.E. Linggi, J.T.
2169–2179. Aldrich, et al., Changes in translational efficiency is a dominant regulatory mecha-
[65] P. Escoubas, G. Corzo, B.J. Whiteley, M.-L. Célérier, T. Nakajima, Matrix-assisted laser nism in the environmental response of bacteria, Integr. Biol. (Camb) 5 (2013)
desorption/ionization time-of-flight mass spectrometry and high-performance 1393–1406.
liquid chromatography study of quantitative and qualitative variation in tarantula [94] Z.L. Bergeron, J.B. Chun, M.R. Baker, D.W. Sandall, S. Peigneur, P.Y.C. Yu, et al., A
spider venoms, Rapid Commun. Mass Spectrom. 16 (2002) 403–413. ‘conovenomic’ analysis of the milked venom from the mollusk-hunting cone
[66] D. Chang, A.M. Olenzek, T.F. Duda Jr., Effects of geographical heterogeneity in snail conus textile—the pharmacological importance of post-translational modifi-
species interactions on the evolution of venom genes, Proc. Biol. Sci. 282 (2015). cations, Peptides 49 (2013) 145–158.
[67] S. Richier, M. Rodriguez-Lanetty, C.E. Schnitzler, V.M. Weis, Response of the [95] M. Rong, Z. Duan, J. Chen, J. Li, Y. Xiao, S. Liang, Native pyroglutamation of
symbiotic cnidarian Anthopleura elegantissima transcriptome to temperature huwentoxin-IV: a post-translational modification that increases the trapping
and UV increase, Comp. Biochem. Physiol. Part D Genomics Proteomics 3 ability to the sodium channel, PLoS ONE 8 (2013), e65984.
(2008) 283–289. [96] E. Kapp, F. Schütz, Overview of tandem mass spectrometry (MS/MS) database
[68] D.J. Orts, S. Peigneur, B. Madio, J.S. Cassoli, G.G. Montandon, A.M. Pimenta, et al., search algorithms, Curr. Protoc. Protein Sci. (2007) Chapter 25:Unit25.2.
Biochemical and electrophysiological characterization of two sea anemone type 1 [97] A.I. Nesvizhskii, Protein identification by tandem mass spectrometry and sequence
potassium toxins from a geographically distant population of Bunodosoma database searching, Methods Mol. Biol. 367 (2007) 87–119.
caissarum, Mar. Drugs 11 (2013) 655–679. [98] A.T. Tu, B.L. Adams, Phylogenetic relationships among venomous snakes of the
[69] K. Sunagar, E.A. Undheim, H. Scheib, E.C. Gren, C. Cochran, C.E. Person, et al., Intra- genus Agkistrodon from Asia and the North American continent, Nature 217
specific venom variation in the medically significant southern Pacific rattlesnake (1968) 760–762.
(Crotalus oreganus helleri): biodiscovery, clinical and evolutionary implications, J. [99] C.C. Yang, Chemistry and evolution of toxins in snake venoms, Toxicon: Off. J. Int.
Proteome 99 (2014) 68–83. Soc. Toxinol. 12 (1974) 1–43.
[70] S. Peigneur, L. Beress, C. Moller, F. Mari, W.G. Forssmann, J. Tytgat, A natural point [100] E.M. Bettke, D.D. Watt, T. Tu, Electrophoretic patterns of venoms from species
mutation changes both target selectivity and mechanism of action of sea anemone of Crotalidae and Elapidae snakes, Toxicon: Off. J. Int. Soc. Toxinol. 4 (1966)
toxins, FASEB J.: Off. Publ. Fed. Am. Soc. Exp. Biol. 26 (2012) 5141–5151. 73–76.
[71] M.G. Grabherr, B.J. Haas, M. Yassour, J.Z. Levin, D.A. Thompson, I. Amit, et al., Full- [101] C.A. Bonilla, N.V. Horner, Comparative electrophoresis of Crotalus and Agkistrodon
length transcriptome assembly from RNA-Seq data without a reference genome, venoms from North American snakes, Toxicon: Off. J. Int. Soc. Toxinol. 7 (1969)
Nat. Biotechnol. 29 (2011) 644–652. 327–329.
[72] J. Archer, G. Whiteley, N.R. Casewell, R.A. Harrison, S.C. Wagstaff, VTBuilder: a tool [102] R. Foote, J.A. MacMahon, Electrophoretic studies of rattlesnake (Crotalus &
for the assembly of multi isoform transcriptomes, BMC Bioinf. 15 (2014) 389. Sistrurus) venom: taxonomic implications, Comp. Biochem. Physiol. B 57 (1977)
[73] C.G. Elsik, K.C. Worley, A.K. Bennett, M. Beye, F. Camara, C.P. Childers, et al., Finding 235–241.
the missing honey bee genes: lessons learned from a genome upgrade, BMC [103] A. Yahel-Niv, E. Zlotkin, Comparative studies on venom obtained from individ-
Genomics 15 (2014) 86. ual scorpions by natural stings, Toxicon: Off. J. Int. Soc. Toxinol. 17 (1979)
[74] T. Laver, J. Harrison, P.A. O'Neill, K. Moore, A. Farbos, K. Paszkiewicz, et al., Assessing 435–446.
the performance of the Oxford nanopore technologies MinION, Biomol. Detect. [104] N.J. da Silva Júnior, P.R. Griffin, S.D. Aird, Comparative chromatography of
Quantif. 3 (2015) 1–8. Brazilian coral snake (Micrurus) venoms, Comp. Biochem. Physiol. B 100
[75] A.C. English, S. Richards, Y. Han, M. Wang, V. Vee, J. Qu, et al., Mind the gap: (1991) 117–126.
upgrading genomes with Pacific biosciences RS long-read sequencing technology, [105] J.L. Glenn, R.C. Straight, Intergradation of two different venom populations of the
PLoS ONE 7 (2012), e47768. Mojave rattlesnake (Crotalus scutulatus scutulatus) in Arizona, Toxicon: Off. J. Int.
[76] M.A. Madoui, S. Engelen, C. Cruaud, C. Belser, L. Bertrand, A. Alberti, et al., Genome Soc. Toxinol. 27 (1989) 411–418.
assembly using nanopore-guided long and error-free DNA reads, BMC Genomics [106] M.F. Martin, H. Rochat, P. Marchot, P.E. Bougis, Use of high performance liquid
16 (2015) 327. chromatography to demonstrate quantitative variation in components of venom
[77] H. Tilgner, F. Grubert, D. Sharon, M.P. Snyder, Defining a personal, allele-specific, from the scorpion Androctonus australisHector, Toxicon: Off. J. Int. Soc. Toxinol.
and single-molecule long-read transcriptome, Proc. Natl. Acad. Sci. U. S. A. 111 25 (1987) 569–573.
(2014) 9869–9874. [107] J.G. Soto, J.C. Perez, M.M. Lopez, M. Martinez, T.B. Quintanilla-Hernandez, M.S.
[78] C.B. Burge, S. Karlin, Finding the genes in genomic DNA, Curr. Opin. Struct. Biol. 8 Santa-Hernandez, et al., Comparative enzymatic study of HPLC-fractionated
(1998) 346–354. Crotalus venoms, Comp. Biochem. Physiol. B 93 (1989) 847–855.
[79] M. Yandell, D. Ence, A beginner's guide to eukaryotic genome annotation, Nat. Rev. [108] L.F. Machado, S. Laugesen, E.D. Botelho, C.A.O. Ricart, W. Fontes, K.C. Barbaro, et al.,
Genet. 13 (2012) 329–342. Proteome analysis of brown spider venom: identification of loxnecrogin isoforms
[80] J.M. Cherry, E.L. Hong, C. Amundsen, R. Balakrishnan, G. Binkley, E.T. Chan, et al., in Loxosceles gaucho venom, Proteomics 5 (2005) 2167–2176.
Saccharomycesgenome database: the genomics resource of budding yeast, Nucleic [109] M.C. Menezes, M.F. Furtado, S.R. Travaglia-Cardoso, A.C.M. Camargo, S.M.T. Serrano,
Acids Res. 40 (2012) D700–D705. Sex-based individual variation of snake venom proteome among eighteen Bothrops
[81] G. dos Santos, A.J. Schroeder, J.L. Goodman, V.B. Strelets, M.A. Crosby, J. Thurmond, jararaca siblings, Toxicon: Off. J. Int. Soc. Toxinol. 47 (2006) 304–312.
et al., FlyBase: introduction of the Drosophila melanogasterrelease 6 reference ge- [110] S.M.T. Serrano, J.D. Shannon, D. Wang, A.C.M. Camargo, J.W. Fox, A multiface-
nome assembly and large-scale migration of genome annotations, Nucleic Acids ted analysis of viperid snake venoms by two-dimensional gel electrophoresis:
Res. 43 (2015) D690–D697. an approach to understanding venom proteomics, Proteomics 5 (2005)
[82] T.W. Harris, J. Baran, T. Bieri, A. Cabunoc, J. Chan, W.J. Chen, et al., WormBase 2014: 501–510.
new views of curated biology, Nucleic Acids Res. 42 (2014) D789–D793. [111] J. Vejayan, T.L. Khoon, H. Ibrahim, Comparative analysis of the venom proteome of
[83] V. Herzig, D.L. Wood, F. Newell, P.A. Chaumeil, Q. Kaas, G.J. Binford, et al., four important Malaysian snake species, J. Venom. Anim. Toxins. Incl. Trop. Dis. 20
ArachnoServer 2.0, an updated online resource for spider toxin sequences and (2014) 6.
structures, Nucleic Acids Res. 39 (2011) D653–D657. [112] J.C. Daltry, G. Ponnudurai, C.K. Shin, N.H. Tan, R.S. Thorpe, W. Wüster, Electropho-
[84] F. Jungo, L. Bougueleret, I. Xenarios, S. Poux, The UniProtKB/Swiss-Prot Tox-Prot retic profiles and biological activities: intraspecific variation in the venom of the
program: a central hub of integrated venom protein data, Toxicon 60 (2012) Malayan pit viper (Calloselasma rhodostoma), Toxicon: Off. J. Int. Soc. Toxinol. 34
551–557. (1996) 67–79.
[85] Y. Terrat, F. Ducancel, Are there unequivocal criteria to label a given protein as a [113] H.L. Gibbs, L. Sanz, J.J. Calvete, Snake population venomics: proteomics-based anal-
toxin? Permissive versus conservative annotation processes, Genome Biol. 14 yses of individual variation reveals significant gene regulation effects on venom
(2013) 406. protein expression in Sistrurus rattlesnakes, J. Mol. Evol. 68 (2009) 113–125.
[86] A.D. Hargreaves, M.T. Swain, D.W. Logan, J.F. Mulley, Testing the toxicofera: com- [114] P. Kleinert, T. Kuster, D. Arnold, J. Jaeken, C.W. Heizmann, H. Troxler, Effect of
parative transcriptomics casts doubt on the single, early evolution of the reptile glycosylation on the protein pattern in 2-D-gel electrophoresis, Proteomics 7
venom system, Toxicon 92 (2014) 140–156. (2007) 15–22.
[87] B.G. Fry, N. Vidal, J.A. Norman, F.J. Vonk, H. Scheib, S.F. Ramjan, et al., Early evolution [115] K. Tanaka, H. Waki, Y. Ido, S. Akita, Y. Yoshida, T. Yoshida, et al., Protein and poly-
of the venom system in lizards and snakes, Nature 439 (2006) 584–588. mer analyses up to m/z 100 000 by laser ionization time-of-flight mass spectrom-
[88] N. Vidal, S.B. Hedges, Higher-level relationships of caenophidian snakes inferred etry, Rapid Commun. Mass Spectrom. 2 (1988) 151–153.
from four nuclear and mitochondrial genes, C. R. Biol. 325 (2002) 987–995. [116] L.M. Smith, N.L. Kelleher, Consortium for top down P. Proteoform: a single term de-
[89] L. Cohen, N. Lipstein, D. Gordon, Allosteric interactions between scorpion toxin re- scribing protein complexity, Nat. Methods 10 (2013) 186–187.
ceptor sites on voltage-gated Na channels imply a novel role for weakly active [117] J.R. Perkins, C.E. Parker, K.B. Tomer, The characterization of snake venoms using
components in arthropod venom, FASEB J.: Off. Publ. Fed. Am. Soc. Exp. Biol. 20 capillary electrophoresis in conjunction with electrospray mass spectrometry:
(2006) 1933–1935. black mambas, Electrophoresis 14 (1993) 458–468.
K. Sunagar et al. / Journal of Proteomics 135 (2016) 62–72 71

[118] J.R. Perkins, B. Smith, R.T. Gallagher, D.S. Jones, S.C. Davis, A.D. Hoffman, et al., [143] H. Chi, H. Chen, K. He, L. Wu, B. Yang, R.-X. Sun, et al., pNovo+: de novo peptide
Application of electrospray mass spectrometry and matrix-assisted laser de- sequencing using complementary HCD and ETD tandem mass spectra, J. Proteome
sorption ionization time-of-flight mass spectrometry for molecular weight as- Res. 12 (2013) 615–625.
signment of peptides in complex mixtures, J. Am. Soc. Mass Spectrom. 4 (1993) [144] K. Jeong, S. Kim, P.A. Pevzner, UniNovo: a universal tool for de novo peptide
670–684. sequencing, Bioinformatics (Oxford, England) 29 (2013) 1953–1962.
[119] P. Favreau, L. Menin, S. Michalet, F. Perret, O. Cheneval, M. Stöcklin, et al., Mass [145] Y. Yan, A. Kusalik, F.-X. Wu, A framework of de novo peptide sequencing for
spectrometry strategies for venom mapping and peptide sequencing from crude multiple tandem mass spectra, IEEE Trans. Nanobioscience. (2015).
venoms: case applications with single arthropod specimen, Toxicon: Off. J. Int. [146] B. Ma, K. Zhang, C. Hendrie, C. Liang, M. Li, A. Doherty-Kirby, et al., PEAKS: powerful
Soc. Toxinol. 47 (2006) 676–687. software for peptide de novo sequencing by tandem mass spectrometry, Rapid
[120] D.G. Nascimento, B. Rates, D.M. Santos, T. Verano-Braga, A. Barbosa-Silva, A.A.A. Commun. Mass Spectrom.: RCM 17 (2003) 2337–2342.
Dutra, et al., Moving pieces in a taxonomic puzzle: venom 2D-LC/MS and data [147] A. Guthals, K.R. Clauser, N. Bandeira, Shotgun protein sequencing with meta-contig
clustering analyses to infer phylogenetic relationships in some scorpions assembly, Mol. Cell. Proteomics 11 (2012) 1084–1096.
from the Buthidae family (Scorpiones), Toxicon: Off. J. Int. Soc. Toxinol. 47 [148] X. Liu, L.J.M. Dekker, S. Wu, M.M. Vanduijn, T.M. Luider, N. Tolić, et al., De novo pro-
(2006) 628–639. tein sequencing by combining top-down and bottom-up tandem mass spectra, J.
[121] H. Steen, M. Mann, The ABC's (and XYZ's) of peptide sequencing, Nat. Rev. Mol. Cell Proteome Res. 13 (2014) 3241–3248.
Biol. 5 (2004) 699–711. [149] R. Datta, M. Bern, Spectrum fusion: using multiple mass spectra for de novo pep-
[122] A. Alape-Girón, L. Sanz, J. Escolano, M. Flores-Díaz, M. Madrigal, M. Sasa, et al., tide sequencing, J. Comput. Biol. 16 (2009) 1169–1182.
Snake venomics of the lancehead pitviper Bothrops asper: geographic, individual, [150] J. Cascardi, B.A. Young, H.D. Husic, J. Sherma, Protein variation in the venom spat by
and ontogenetic variations, J. Proteome Res. 7 (2008) 3556–3571. the red spitting cobra, Naja pallida (Reptilia: Serpentes), Toxicon: Off. J. Int. Soc.
[123] A.M. Rodriguez, S. Dutertre, R.J. Lewis, F. Marí, Intraspecific variations in Conus Toxinol. 37 (1999) 1271–1279.
purpurascens injected venom using LC/MALDI-TOF-MS and LC-ESI-triple TOF-MS, [151] C.A. Prator, K.M. Murayama, J.R. Schulz, Venom variation during prey capture by
Anal. Bioanal. Chem. (2015). the cone snail, Conus textile, PLoS ONE 9 (2014), e98991.
[124] J.J. Calvete, P. Ghezellou, O. Paiva, T. Matainaho, A. Ghassempour, H. Goudarzi, et al., [152] E.B. Undheim, D.N. Georgieva, H.H. Thoen, J.A. Norman, J. Mork, C. Betzel, et al.,
Snake venomics of two poorly known Hydrophiinae: comparative proteomics of Venom on ice: first insights into Antarctic octopus venoms, Toxicon: Off. J. Int.
the venoms of terrestrial Toxicocalamus longissimus and marine Hydrophis Soc. Toxinol. 56 (2010) 897–913.
cyanocinctus, J. Proteome 75 (2012) 4091–4101. [153] B. Chatterji, A. Pich, MALDI imaging mass spectrometry and analysis of endogenous
[125] B. Nguyen, J. Molgó, H. Lamthanh, E. Benoit, T.A. Khuc, D.N. Ngo, et al., High accu- peptides, Expert Rev. Proteomics 10 (2013) 381–388.
racy mass spectrometry comparison of Conus bandanus and Conus marmoreus [154] M.M. Gessel, J.L. Norris, R.M. Caprioli, MALDI imaging mass spectrometry: spatial
venoms from the South Central coast of Vietnam, Toxicon: Off. J. Int. Soc. Toxinol. molecular analysis to enable a new age of discovery, J. Proteome 107 (2014)
75 (2013) 148–159. 71–82.
[126] A.K. Tashima, L. Sanz, A.C.M. Camargo, S.M.T. Serrano, J.J. Calvete, Snake venomics [155] J. Stauber, L. MacAleese, J. Franck, E. Claude, M. Snel, B.K. Kaletas, et al., On-tissue
of the Brazilian pitvipers Bothrops cotiara and Bothrops fonsecai. Identification of protein identification and imaging by MALDI-ion mobility mass spectrometry, J.
taxonomy markers, J. Proteome 71 (2008) 473–485. Am. Soc. Mass Spectrom. 21 (2010) 338–347.
[127] E.A.B. Undheim, K. Sunagar, B.R. Hamilton, A. Jones, D.J. Venter, B.G. Fry, et al., Mul- [156] A. Kiss, D.F. Smith, B.R. Reschke, M.J. Powell, R.M.A. Heeren, Top-down mass spec-
tifunctional warheads: diversification of the toxin arsenal of centipedes via novel trometry imaging of intact proteins by laser ablation ESI FT-ICR MS, Proteomics 14
multidomain transcripts, J. Proteome 102 (2014) 1–10. (2014) 1283–1289.
[128] A. Zelanis, A.K. Tashima, A.F.M. Pinto, A.F. Paes Leme, D.R. Stuginski, M.F. Furtado, [157] E.A.B. Undheim, B.R. Hamilton, N.D. Kurniawan, G. Bowlay, B.W. Cribb, D.J. Merritt,
et al., Bothrops jararaca venom proteome rearrangement upon neonate to adult et al., Production and packaging of a biological arsenal: evolution of centipede
transition, Proteomics 11 (2011) 4218–4228. venoms under morphological constraint, Proc. Natl. Acad. Sci. U. S. A. 112 (2015)
[129] D.C. Pimenta, B.C. Prezoto, K. Konno, R.L. Melo, M.F. Furtado, A.C.M. Camargo, 4026–4031.
et al., Mass spectrometric analysis of the individual variability of Bothrops [158] G. Brunda, B.S. Rao, R.K. Sarin, Quantitation of Indian krait (Bungarus caeruleus)
jararaca venom peptide fraction. Evidence for sex-based variation among the venom in human specimens of forensic origin by indirect competitive inhibition
bradykinin-potentiating peptides, Rapid Commun. Mass Spectrom. 21 (2007) enzyme-linked immunosorbent assay, J. AOAC Int. 89 (2006) 1360–1366.
1034–1042. [159] Z.E. Selvanayagam, S.G. Gnanavendhan, K.A. Ganesh, D. Rajagopal, P.V. Rao, ELISA
[130] C.T. Cologna, J.S. Cardoso, E. Jourdan, M. Degueldre, G. Upert, N. Gilles, et al., for the detection of venoms from four medically important snakes of India,
Peptidomic comparison and characterization of the major components of the Toxicon: Off. J. Int. Soc. Toxinol. 37 (1999) 757–770.
venom of the giant ant Dinoponera quadriceps collected in four different areas of [160] N.H. Tan, K.H. Yeo, M.I. Jaafar, The use of enzyme-linked immunosorbent
Brazil, J. Proteome 94 (2013) 413–422. assay for the quantitation of Calloselasma rhodostoma (Malayan pit viper)
[131] K.Y. Tan, C.H. Tan, S.Y. Fung, N.H. Tan, Venomics, lethality and neutralization of venom and venom antibodies, Toxicon: Off. J. Int. Soc. Toxinol. 30 (1992)
Naja kaouthia (monocled cobra) venoms from three different geographical regions 1609–1620.
of Southeast Asia, J. Proteome 120 (2015) 105–125. [161] S.D. Aird, N.J. da Silva, Comparative enzymatic composition of Brazilian coral snake
[132] L. Sanz, H.L. Gibbs, S.P. Mackessy, J.J. Calvete, Venom proteomes of closely relat- (Micrurus) venoms, Comp. Biochem. Physiol. B 99 (1991) 287–294.
ed Sistrurus rattlesnakes with divergent diets, J. Proteome Res. 5 (2006) [162] L.M. Mulfinger, A.W. Benton, M.W. Guralnick, R.A. Wilson, A qualitative and quan-
2098–2112. titative analysis of proteins found in vespid venoms, J. Allergy Clin. Immunol. 77
[133] G.D. Findlay, M.J. MacCoss, W.J. Swanson, Proteomic discovery of previously unan- (1986) 681–686.
notated, rapidly evolving seminal fluid genes in Drosophila, Genome Res. 19 (2009) [163] S. Namiranian, R.C. Hider, Use of HPLC to demonstrate variation of venom toxin
886–896. composition in the Thailand cobra venoms Naja naja kaouthia and Naja naja
[134] S.D. Robinson, H. Safavi-Hemami, S. Raghuraman, J.S. Imperial, A.T. siamensis, Toxicon: Off. J. Int. Soc. Toxinol. 30 (1992) 47–61.
Papenfuss, R.W. Teichert, et al., Discovery by proteogenomics and character- [164] H. Arikan, N. Alpagut Keskin, I.E. Cevik, C. Ilgaz, Age-dependent variations in the
ization of an RF-amide neuropeptide from cone snail venom, J. Proteome 114 venom proteins of Vipera xanthina (Gray, 1849) (Ophidia: Viperidae), Türk.
(2015) 38–47. Parazitol. Derg. 30 (2006) 163–165.
[135] K. Krug, S. Popic, A. Carpy, C. Taumer, B. Macek, Construction and assessment of in- [165] M. Lang Balija, A. Vrdoljak, L. Habjanec, B. Dojnović, B. Halassy, B. Vranesić, et al., The
dividualized proteogenomic databases for large-scale analysis of nonsynonymous variability of Vipera ammodytes ammodytes venoms from Croatia—biochemical
single nucleotide variants, Proteomics 14 (2014) 2699–2708. properties and biological activity, Comp. Biochem. Physiol. C Toxicol. Pharmacol. 140
[136] A.D. Catherman, O.S. Skinner, N.L. Kelleher, Top down proteomics: facts and per- (2005) 257–263.
spectives, Biochem. Biophys. Res. Commun. 445 (2014) 683–693. [166] M. Bantscheff, S. Lemeer, M.M. Savitski, B. Kuster, Quantitative mass spectrometry
[137] I. Karbat, M. Turkov, L. Cohen, R. Kahn, D. Gordon, M. Gurevitz, et al., X-ray struc- in proteomics: critical review update from 2007 to the present, Anal. Bioanal.
ture and mutagenesis of the scorpion depressant toxin LqhIT2 reveals key determi- Chem. 404 (2012) 939–965.
nants crucial for activity and anti-insect selectivity, J. Mol. Biol. 366 (2007) [167] S.W. Holman, P.F.G. Sims, C.E. Eyers, The use of selected reaction monitoring in
586–601. quantitative proteomics, Bioanalysis 4 (2012) 1763–1786.
[138] P. Anand, A. Grigoryan, M.H. Bhuiyan, B. Ueberheide, V. Russell, J. Quinoñez, et al., [168] J.-F. Gao, Y.-F. Qu, X.-Q. Zhang, Y. He, X. Ji, Neonate-to-adult transition of snake
Sample limited characterization of a novel disulfide-rich venom peptide toxin from venomics in the short-tailed pit viper, Gloydius brevicaudus, J. Proteome 84
terebrid marine snail Terebra variegata, PLoS ONE 9 (2014), e94122. (2013) 148–157.
[139] B.M. Ueberheide, D. Fenyö, P.F. Alewood, B.T. Chait, Rapid sensitive analysis of cys- [169] S.A. Ali, K. Baumann, T.N.W. Jackson, K. Wood, S. Mason, E.A.B. Undheim, et al.,
teine rich peptide venom components, Proc. Natl. Acad. Sci. U. S. A. 106 (2009) Proteomic comparison of Hypnale hypnale (hump-nosed pit-viper) and
6910–6915. Calloselasma rhodostoma (Malayan pit-viper) venoms, J. Proteome 91 (2013)
[140] D. Petras, P. Heiss, R.D. Süssmuth, J.J. Calvete, Venom proteomics of Indonesian king 338–343.
cobra, Ophiophagus hannah: integrating top-down and bottom-up approaches, J. [170] H. Safavi-Hemami, H. Hu, D.G. Gorasia, P.K. Bandyopadhyay, P.D. Veith, N.D. Young,
Proteome Res. 14 (2015) 2539–2556. et al., Combined proteomic and transcriptomic interrogation of the venom gland of
[141] M. Bern, Y. Cai, D. Goldberg, Lookup peaks: a hybrid of de novo sequencing and da- Conus geographus uncovers novel components and functional compartmentaliza-
tabase search for protein identification by tandem mass spectrometry, Anal. Chem. tion, Mol. Cell. Proteomics 13 (2014) 938–953.
79 (2007) 1393–1400. [171] X. Xu, Z. Duan, Z. Di, Y. He, J. Li, Z. Li, et al., Proteomic analysis of the venom from
[142] L. Zamdborg, R.D. LeDuc, K.J. Glowacz, Y.-B. Kim, V. Viswanathan, I.T. the scorpion Mesobuthus martensii, J. Proteome 106 (2014) 162–180.
Spaulding, et al., ProSight PTM 2.0: improved protein identification and char- [172] T. Rachamim, D. Morgenstern, D. Aharonovich, V. Brekhman, T. Lotan, D. Sher, The
acterization for top down mass spectrometry, Nucleic Acids Res. 35 (2007) dynamically evolving nematocyst content of an anthozoan, a scyphozoan, and a
W701–W706. hydrozoan, Mol. Biol. Evol. 32 (2015) 740–753.
72 K. Sunagar et al. / Journal of Proteomics 135 (2016) 62–72

[173] D. Sher, Y. Fishman, M. Zhang, M. Lebendiker, A. Gaathon, J.M. Mancheno, et al., [175] E.S. Wong, D. Morgenstern, E. Mofiz, S. Gombert, K.M. Morris, P. Temple-Smith,
Hydralysins, a new category of beta-pore-forming toxins in Cnidaria, J. Biol. et al., Proteomics and deep sequencing comparison of seasonally active venom
Chem. 280 (2005) 22847–22855. glands in the platypus reveals novel venom peptides and distinct expression pro-
[174] Y. Moran, H. Weinberger, A.M. Reitzel, J.C. Sullivan, R. Kahn, D. Gordon, et al., Intron files, Mol. Cell. Proteomics 11 (2012) 1354–1364.
retention as a posttranscriptional regulatory mechanism of neurotoxin expression
at early life stages of the starlet anemone Nematostella vectensis, J. Mol. Biol. 380
(2008) 437–443.

You might also like