EVOLUTIONARY AND FUNCTIONAL CHRACTERIZATION OF OXIDATIVE STRESS PROTEINS IN DENDROCTONUS PONDEROSAE HOPKINS (CURCULIONIDAE: SCOLYTINAE) by Luke Spooner B.Sc., University of Northern British Columbia, 2012 B. Ed, Simon Fraser University, 2014 THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF MASTER OF SCIENCE IN NATURAL RESOURCES AND ENVIRONMENTAL STUDIES (BIOLOGY) UNIVERSITY OF NORTHERN BRITISH COLUMBIA December 2020 © Luke Spooner, 2020 I Abstract Understanding the mountain pine beetle detoxification systems is vital for predicting its continued spread into the novel jack pine host. Phylogenetic analyses were conducted for mountain pine beetle catalase, glutathione peroxidase, superoxide dismutase, and peroxiredoxin. These proteins were generally conserved, but there were differences in some key functional motifs. Specifically, a peroxiredoxin (DPPrx1) contained a unique combination of hyperoxidation motifs. DPPrx1 and a superoxide dismutase (DPSOD1) were selected for further functional analyses and demonstrated higher reactivity when compared to other SOD and Prx proteins. Also, DPPrx1 experiences hyperoxidation at a lower H2O2 concentration (~0.06 mM) than human peroxiredoxin (~0.12 mM). In other systems, hyperoxidized peroxiredoxin does act as a signal molecule for the expression of other oxidative stress proteins. Therefore, due to its relatively high reactivity and potential role as a cellular signal, DPPrx1 could serve as a future pest management target. II Table of Contents Abstract II Table of Contents III List of Tables IV List of Figures V Chapter 1: Introduction 1.1. Mountain pine beetle ecology and outbreak 1.2 Pine Chemical Defenses 1.3 Mountain Pine Beetle detoxification 1.4 Mountain pine beetle Oxidative stress proteins 1.5 Investigation of recombinant oxidative stress proteins 1.6 Research question 1.7 Objectives 1.8 Hypothesis 8 8 9 11 13 18 18 18 19 Chapter 2 Phylogenetics and Sequence Analyses 2.1 Introduction 2.2 Research Question 2.3 Objectives 2.4 Hypothesis 2.5 Materials and methods 2.6 Results 2.7 Discussion 2.8 Conclusions and Future Directions 20 20 27 27 28 28 31 45 52 Chapter 3: Functional Characterization of DPPRx1 and DPSOD1 3.1 Introduction 3.2 Research Question 3.3 Objectives 3.4 Hypothesis 3.5 Materials and Methods 3.6 Results 3.7 Discussion 3.8 Conclusions and Future Directions 54 54 62 63 63 63 76 82 89 Chapter 4: Conclusions 92 Works Cited 97 III List of Tables Table 2.1 - Major primary oxidative stress proteins (superoxide dismutase, peroxiredoxin, glutathione peroxidase, and catalase) for Tribolium castaneum (TC), Anoplophora glabripennis (AG), Bombyx mori (BM), Bicyclus anynana (BA), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Drosophila melanogaster (DM), and Anopheles gambiae (AGa). ......................................................................................................... 29 Table 2.2 - Primary oxidative stress proteins of Dendroctonus ponderosae and their key features....................................................................................................................... 32 Table 3.1 - Cloning primers for DPPrx1 and DPSOD2, ATTB sequences are bolded .... 64 Table 3.2 - The proportion and type of reagent in each SOD sample and blank in the Sigma-Aldrich WST-1 plate assay ............................................................................ 76 IV List of Figures Figure 1.1 -Physical, chemical, and protein-based defense mechanisms employed by pine trees, mountain pine beetles, and their fungal and bacterial symbionts. ................... 12 Figure 1.2 - Key oxidative stress proteins (Superoxide dismutase, Peroxiredoxin, Catalase, and Glutathione peroxidase) used by the mountain pine beetle ................. 16 Figure 2.1 - Alignment for catalase sequences from Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (Aga), and Anoplophora glabripennis (AG). ............................ 34 Figure 2.2 - Neighbour-joining tree of insect catalase sequences .................................... 35 Figure 2.3 - Alignments for glutathione peroxidase sequences ........................................ 36 Figure 2.4 - Neighbour-joining tree of insect glutathione peroxidase sequences ............. 37 Figure 2.5 - Alignments for superoxide dismutase sequences .......................................... 38 Figure 2.6 - Neighbour-joining tree of insect superoxide dismutase sequences ............... 39 Figure 2.7 - Alignments for peroxiredoxin 1 and peroxiredoxin 6 sequences................. 41 Figure 2.8 - Neighbour-joining tree of insect peroxiredoxin ............................................ 43 Figure 2.9 - Alignments for peroxiredoxin 5 sequences ................................................... 44 Figure 2.10 - Neighbour-joining tree of insect peroxiredoxin 5 sequences ...................... 45 Figure 3.1 - The reaction mechanism for a typical 2-cysteine peroxiredoxin .................. 59 Figure 3.2 - Primary protein structure of DPPrx1. ............................................................ 60 Figure 3.3 - Overview of the LR and BP steps ................................................................. 65 Figure 3.4 – A standard curve generated from BSA standards of differing concentrations and their respective absorbance readings at 595 nm. ................................................ 71 Figure 3.5 - Antioxidant activity of DPPrx1. .................................................................... 77 Figure 3.6 - Formation of HRP compound 1 in varying concentrations of DPPrx1 ........ 78 Figure 3.7 - Determination of second-order rate constant (9.8 x 107 M-1s-1) for the reaction of reduced DPPrx1 with H2O2 through competition with horseradish peroxidase. ................................................................................................................. 79 Figure 3.8. - Determination of pKa (5.4213) for reduced DPPrx1 with H2O2 ................. 80 Figure 3.9. Nonreducing SDS-PAGE analysis of purified, reduced DPPrx1. .................. 81 V Figure 3.10 - (a) Inhibition percentages of WST-1 reduction for differing concentrations of DPSOD1 ................................................................................................................ 82 VI Acknowledgements I would first of all like to thank my supervisor Dr. Dezene Huber for providing me with this opportunity and trusting me to conduct this research. In addition, my wonderful committee members, Dr. Philip Batista and Dr. Brent Murray. My first introduction to research was as an undergraduate student in both Dezene and Brent’s labs, so getting to continue my education with both of them has been very special. As well, Phil spent many hours helping me troubleshoot my PCR experiments and refine my phylogenetic tree analyses. I will always be grateful for his time, insights, and supportive nature. I would also like to express my gratitude to Dr. Nicole Sukdeo who has helped me multiple times in this project, particularly during the protein expression portion where without her guidance and support I am not sure whether I would have ever managed to generate sufficient protein yields. As well, Caitlin Pitt at the UNBC genetics lab could always be counted on to provide quick and reliable sequencing results and Victoria Rezendes, Alexander Douglas, and Kian Draper spent many hours assisting me in the lab. I am thankful for my friends and colleagues at the Academic Success Center and Prince George Public Library, both of which provided me with meaningful and reliable employment throughout my masters. Finally, to all of my family and friends who have always supported me in my academic pursuits, to quote Dr. Andrew H. Knoll, “In both nature and in nurture, I am lucky.” VII Chapter 1: Introduction 1.1. Mountain pine beetle ecology and outbreak Bark beetles (Curculionidae: Scolytinae) are a group of phloeophagous insects that have a major impact on conifer health in forest ecosystems and contribute to overall tree mortality. This is particularly true in the case of the mountain pine beetle, Dendroctonus ponderosae Hopkins, which is endemic to western North America. Historically, the range of this species extends from northern Mexico (latitude 31° north) to central British Columbia (BC) (latitude 56° north). It can feed on and reproduce in all North American species of pine, with the exception of Pinus jeffreyi (Jeffrey pine), but its primary host species is lodgepole pine, Pinus contorta Douglas (Safranyik et al. 2010). Populations of the mountain pine beetle fluctuate in four phases: endemic (low population density), incipient-endemic (population density increases), epidemic (populations reach outbreak level), and post-epidemic (populations decline) (Carroll et al. 2006). Typically, an epidemic phase lasts five years, however in 1994 the largest outbreak in recorded history began, and to date, has killed 750 million cubic meters of lodgepole pine stands in British Columbia and now in Alberta (Janes et al. 2014; Aukema et al. 2006). Despite the severity of this outbreak, at normal low population levels mountain pine beetle are generally restricted to trees in a weakened condition (Saab et al. 2014). With the massive spike in population that occurred from this most recent outbreak, a major expansion in the range of the mountain pine beetle also took place. The mountain pine beetle travelled farther north than previously observed and also crossed the Canadian Rocky Mountains east into Alberta (Janes et al. 2014; Cullingham et al. 2011). There the beetle was able to proliferate within a hybrid zone comprised of lodgepole pine 8 and jack pine, Pinus banksiana (Lusebrink et al. 2013; Cullingham et al. 2011). Eventually the mountain pine beetle was able to progress into forests comprised primarily of jack pines, which extend as far as to eastern Canada and the United States. The mountain pine beetle alongside its fungal and other microbial symbionts have been coevolving with the lodgepole pine for eons, but this is not the case for jack pines (Clark et al. 2014). Jack pines have different resin chemistry compared to lodgepole pine, and continued success of the mountain pine beetle within this novel host species is contingent upon many factors, but particularly on how it will be able to overcome the trees’ defenses (Clark et al. 2014; Janes et al. 2014). Therefore, developing a better understanding of mechanisms that mountain pine beetle use to overcome the host tree’s defense systems could be critical towards predicting the continued spread of infestation across novel jack pine host extents of range. 1.2 Pine Chemical Defenses The two main groups of secondary metabolites produced by pines are the phenols and terpenes. Phenols are a key part of the pine defense system, and they are important in disrupting the progress of affecting fungal or bacterial infection (Vornam et al. 2019; Tholl 2015). Therefore, in terms of bark beetle defenses, terpenes are usually considered the much more noteworthy group. They are the largest group of secondary metabolites, with 40,000 unique compounds found in nature (Tholl 2015). They serve a number of functions, ranging from growth and development to facilitating specific interactions between conifers and their environments (Tholl 2015). The relationship and impact pine terpenes have on the mountain pine beetle is a complicated one. Seybold et al. (2006) discussed how the mountain pine beetle uses 9 monoterpenes in the biosynthesis of its aggregation pheromones. Aggregation pheromones are essential to the success of the mountain pine beetle, as they help to recruit more individuals to the host tree, ultimately overwhelming the host tree’s defenses (Chiu et al. 2019). That said, terpenes are also harmful to the mountain pine beetle, but their exact impact is still not fully understood. Many are toxic and cause death, while others appear to disrupt ovary and egg development, serve as food deterrents, or even mimic the juvenile hormone of mountain pine beetle and hinder maturation (Chiu et al. 2019; Seybold et al. 2006). Additionally, it has been found that the monoterpenes, αpinene and β-pinene, serve as kairomones for parasitoids and predaceous insects like the wrinkled bark beetle, Lasconotus complex (Seybold et al. 2006). The role of terpenes in this tri-trophic interaction is an area that requires more in-depth study and is another intriguing potential function of these chemicals (Shikano et al. 2017). In addition to the initial toxicity of these compounds, the breakdown of terpenes by invading bark beetles produces highly degradative reactive oxygen species (ROS) that the invading beetles then have to contend with (Mason et al. 2018; Birben et al. 2012). Conifer trees are believed to have evolved over 200 million years ago, and shortly after (~65 million years later) the first wood boring beetles evolved (Seybold et al. 2006). Therefore, as the primary host of mountain pine beetle, lodgepole pine has had millions of years to adapt and refine the aforementioned defenses to specifically contend with the mountain pine beetle. However, the mountain pine beetle has in turn been able to optimize its own set of strategies to overcome these obstacles (Seybold et al. 2006). 10 1.3 Mountain Pine Beetle detoxification Mountain pine beetles have a number of pathways dedicated to isolation and degradation - a multigenic process that has been the focus of many studies (Robert et al. 2016; Pitt et al. 2014; Keeling et al. 2013; Robert et al. 2013; Bonnett et al. 2012). With the successful sequencing of the mountain pine beetle genome, it is now possible to identify specific genes and gene families involved in these different pathways (Keeling et al. 2013). Particularly, studies investigating the global mountain pine beetle production of RNA transcripts (transcriptome) and proteins (proteome) (Pitt et al. 2014; Robert et al. 2013), that occur before and after host colonization, have particularly garnered a considerable amount of attention. 11 Physical/Chemical%defenses% Phenolic"lignin"layer" DehydraEon% J" Pine%Tree% Host% (Pinus&Spp.)" Chemical%Defenses% MonoterpenesP"*" DiterpenesP"*" Sesquiterpenes V" Phenols S" J*" Bark%Beetle% Predators% monoterpenes% OxidaEve%Stress% Super"oxide"dismutase"(EIM)U" Ferri n U" Peroxiredoxin" Glutathione"peroxidase" CatalaseU" Free%% Radicals% Physical%Breakdown% T" FV" Mountain%Pine%Beetle% (Dendroctonus&ponderosae)% Proteins% Chi nase" Gluconase" Porins" Lec ns" Chemicall%BreakdownU" Cytochrome"P4 0" Glutathione"S5transferase" Glucosyl"transferase" Esterase" ABC"Transporter" Alcohol"degydrogenase" Mechanical%Defenses% Stone"Cells" Calcium"oxalate"crystals" Resin"Ducts"(Trauma c/axial*)" Fiber"lignifica on" Trichomes" JE*" 12 Figure 1.1 -Physical, chemical, and protein-based defense mechanisms employed by pine trees, mountain pine beetles, and their fungal and bacterial symbionts. Red arrows represent predators of the mountain pine beetle that are attracted to monoterpene kairomones. Thick green arrows represent induced defenses, thinner arrows represent constitutive defenses, green curved arrows indicate the plant growth regulators (J=methyl jasmonate and E=ethylene). The curved orange arrows represent the aggregation pheromones (T=transverbenone) and the anitaggregation pheromones (V=verbenone and F=frontalin) derived from the pine monterpenes. Chemical defenses are paired with a superscript indicating which pathway, MVA (V), MEP (P), or shikimic acid (S), produces the specific compound. For the mountain pine beetle and their fungal symbionts, a superscript U is used to indicate genes that were unregulated in previous transcriptomic or proteomic work. Fungal%Symbionts% (Grossmania&clavigera)" Free%% Radicals% J" J" Chemical%Defenses% MonoterpenesP" DiterpenesP" Sesquiterpene V" Phenols S" Physical/Chemical%defenses% Phenolic"lignin"layer" OxidaEve%Stress% Cu/Zn5"Superoxide"dismutase U" PeroxidaseU" Thioredoxin U" Thioredoxin"reducataseU" Chemical%%Breakdown% Oxidoreductase U" Flavoprotein"monoxygenaseU" Dehydrogenase/reductaseU" 35hydroxyacyl5CoA"dehydrogenases" Aldehyde"dehydrogenases" Oxidoreducatse" Enoyl5CoA"hydratase" Chemical%BreakdownM" Bacterial%Symbionts:% (Pseudomonas,&Rahnella,& Serra:a,&and&Burkholderia% Proteins% Chi nase" Gluconase" Porins" Lec ns" When comparing the results of the studies analyzing the mountain pine beetle transcriptome and proteome, it was found that gene products involved in detoxification were upregulated. Specifically, several gene families displayed upregulation: cytochrome p450s, ABC transporters, and oxidative stress proteins (Pitt et al. 2014;Robert et al. 2013). Despite that, the number of products upregulated from each gene family varied between the transcriptome and the proteome, it is likely that these genes play a considerable role in detoxification (Pitt et al. 2014, Robert et al. 2013). The p450 family of proteins has been investigated for their suspected role in trans-verbenol production, but their suspected primary role is breaking down the secondary metabolites of plants. ABC transporters function to actively pump toxins out of the cell, somewhat in opposition to the porin proteins utilized by the trees. The other major family of proteins that was upregulated was the oxidative stress proteins and will therefore be the focus of this project (Figure 1.1) (Pitt et al. 2014; Robert et al. 2013). 1.4 Mountain pine beetle Oxidative stress proteins As discussed, the factors influencing mountain pine beetle colonization of pine trees are many and complicated. This study’s primary focus was on the oxidative stress system. Specifically, it was conducted with the intention of functionally and phylogenetically characterizing the mountain pine beetle’s primary oxidative stress proteins with a focus on those that displayed upregulation in previous transcriptomic and proteomic analyses. Oxidative stress proteins moderate the harmful effects of reactive oxygen species. All aerobic organisms continually produce reactive oxygen species, also known as free 13 radicals. However, when exposed to toxic chemicals, like the secondary metabolites of pine trees, exposure to reactive oxygen species increases considerably (Birben et al. 2012). This is believed to be due to the breakdown of the lignin polymer, and the number of phenolic compounds generated as a result of this process. In other species these compounds have been shown to lead to the formation of the free radicals quinone and aglycone (Mason and Bowers 2017). Free radicals can be extremely harmful to the mountain pine beetle because they deplete fat stores, degrade DNA and protein, and damage gut tissue (Mason 2016; Dmochowska-Ślęzak et al. 2016). Therefore, being able to properly metabolize such compounds is important to the mountain pine beetle’s survival during host colonization. This is further supported by the upregulation of these proteins in blue stain fungus, Grossmania clavigera, that was exposed to detoxification conditions (DiGuistini et al. 2011). The oxidative stress proteins of mountain pine beetle have not been studied in terms of functional characteristics. However, within other organisms, there is a general understanding of what substrates are degraded by each enzyme (Ajuwon et al. 2015). Primary oxidative stress proteins are those that directly break down free radicals and include superoxide dismutase (SOD), peroxiredoxin (Prx), glutathione peroxidase (GPx), and catalase. Many of these proteins displayed upregulation in the mountain pine beetle transcriptome and proteome during host colonization (Pitt et al. 2014; Robert et al. 2013). In addition to these four are thioredoxin and thioredoxin reductase (TR) which are essential oxidative stress proteins but are designated as ‘peripheral’ to the main pathways because they do not directly interact with the free radicals, so were not examined in this study (Sheng et al. 2014). They were also not upregulated during host colonization in the 14 transcriptome and proteome of the mountain pine beetle (Pitt et al. 2014; Robert et al. 2013). One of the most notable of these upregulated oxidative stress proteins is superoxide dismutase, which processes the harmful superoxide anion (O2-) into freefloating oxygen and H2O2. There are three main families of animal SOD – extracellular, intracellular, and mitochondrial (Broxton and Culotta 2019). All three families are present within the mountain pine beetle, but only the intracellular variant demonstrated upregulation during host colonization, when the insect is likely to be detoxifying host secondary metabolites (Pitt et al. 2014; Robert et al. 2013). Mitochondrial SOD differs from the other SOD types in terms of structure and that its metal cofactor is a manganese ion rather than copper or zinc. The primary function of mitochondrial SOD is to breakdown the free radicals produced from aerobic respiration, and therefore it would make sense that its expression would not be heavily impacted during host colonization (Zelko et al. 2002). It is curious that the extracellular variant did not show any upregulation (Pitt et al. 2014; Robert et al. 2013), considering that Colinet et al. (2011) found that these proteins are found in the venom and digestive juices of some parasitic wasps and flies as they have antivirulence properties that protect against the host’s immune response. SOD breaks down the superoxide anion into H2O2, which if left to build up within the cells could prove lethal, but instead it is readily processed by peroxiredoxin, catalase, and GPx (Ajuwon et al. 2015). In contrast to the multiple proteins for degrading H2O2, the superoxide anion is only broken down by SOD (Figure 1.1; Figure 1.2). Therefore, 15 SOD is likely critical to successful mitigation of free radicals in the mountain pine beetle (Dmochowska-Ślęzak et al. 2016). Figure 1.2 - Key oxidative stress proteins (Superoxide dismutase, Peroxiredoxin, Catalase, and Glutathione peroxidase) used by the mountain pine beetle, their free radical substrates, and primary electron donors, derived from Ajuwon et al. (2015). The peroxiredoxins are a family of thiol-specific proteins containing no metal cofactors. In addition to their antioxidant activities, they are involved with cell signaling cascades, cell proliferation, and immune responses (Shi et al. 2014). Currently there are six identified isoforms of peroxiredoxin (Shi et al. 2014). The most notable way of distinguishing the different peroxiredoxin types is by their number of cysteine groups: Prx1–4 are classified as the typical 2-cysteine type, Prx5 is the atypical 2 cysteine type, and Prx6 is the one cysteine type. Prx1,2, and 6 are found in the cystosol; Prx4 contains 16 an extracellular N-terminal signal; Prx3 is located within the mitochondria; Prx5 can be found in the cytosol, mitochondria, and peroxisomes; and Prx 6 is cytosolic but primarily localized to the outer membranes of the peroxisomes, mitochondria, and nucleus (Rhee et al. 2016; Shi et al. 2014). All six isoforms are present in the mountain pine beetle and five were upregulated during host colonization: Prx1, Prx2, Prx4, Prx5 and Prx6 (Pitt et al. 2014; Robert et al. 2013). Despite being a key component of the antioxidant systems of mammals (with there being six isoforms in humans), glutathione peroxidases were not originally considered a part of insect systems. However, isoforms were initially identified in Drosophilia melanogaster, and have since been discovered in many other insect species (Corona and Robinson 2006). GPx's are a family of antioxidant protein that use glutathione or thioredoxins as electron donors, to reduce H2O2. Which electron donor they use is dependent on whether they contain a cysteine or a selenocysteine as the electron acceptor (Dias et al. 2016). As will be discussed in Chapter 2, mountain pine beetles only have two GPx isoforms that are both localized within the cytosol and both use cysteine as their main electron acceptor. Neither protein demonstrated upregulation in the transcriptome or proteome during host colonization (Pitt et al. 2014; Robert et al. 2013), but since GPx is a primary protein in the antioxidant pathway, they were included among the proteins deemed suitable for further analyses. In contrast to these other proteins, there is only one isoform of catalase that has been identified within other vertebrate and invertebrate systems. The one known exception is Caenorhabditis elegans where the catalase gene is believed to have duplicated (Corona and Robinson 2006). Consistent with these other systems, mountain 17 pine beetle has a single catalase that is intracellular and was not upregulated in the transcriptome and proteome during host colonization (Pitt et al. 2014; Robert et al. 2013). However, it is also a key part of the oxidative stress pathway, so it was also deemed suitable for further analyses. 1.5 Investigation of recombinant oxidative stress proteins The information garnered from this process, and the resulting phylogenetic and functional assay data could be useful for other researchers working in metabolomics and looking to functionally characterize proteins. As well, the findings from this study could help to confirm the role of these proteins within mountain pine beetle. This research is also valuable because the findings could be used in designing outbreak models that could be predictive of the mountain pine beetle’s continued range expansion. 1.6 Research question 1) Do the mountain pine beetle’s primary oxidative stress proteins (catalase, glutathione peroxidase, superoxide dismutase, and peroxiredoxin) have unique evolutionary histories or functional features compared to those found in other organisms? 1.7 Objectives 1) Construct amino acid alignments and neighbor joining trees for mountain pine beetle catalase, glutathione peroxidase, peroxiredoxin, and superoxide dismutase and their homologs found in other insect species with well-characterized genomes (Chapter 2). 18 2) Design and optimize protocols allowing for the expression and purification of mountain pine beetle catalase, glutathione peroxidase, peroxiredoxin, and superoxide dismutase (Chapter 3). 3) Design and optimize protocols allowing for the functional characterization of mountain pine beetle superoxide dismutase, peroxiredoxin, glutathione peroxidase, and catalase (Chapter 3). 1.8 Hypothesis 1) The mountain beetle’s primary oxidative stress proteins will have evolved unique functional features enabling them to mitigate the high oxidative stress environments produced by their host trees. 19 Chapter 2 Phylogenetics and Sequence Analyses 2.1 Introduction Mitigating the damaging effects of ROS is essential to an organism’s metabolic functioning. Therefore, the important role of oxidative stress proteins is ubiquitous across all life forms. However, across different taxonomic groups, there are differences in abundance and reactivity for each protein type. For instance, glutathione reductase is common in vertebrates but mainly absent in many insect groups; instead analogous genes like thioredoxin and thioredoxin peroxidase have evolved to act as functional equivalents (Couto et al. 2016). Catalase, glutathione peroxidase, superoxide dismutase, and peroxiredoxin play a primary role in reducing oxidative stress, but taxonomic variations are still present. Corona and Robinson (2006) conducted a phylogenetic analysis of the primary and secondary antioxidant system of Apis mellifera (western honeybee). They found that although the basic components of the system remained consistent (mainly the specific antioxidant proteins present) there were considerably more paralogs present for each protein in the western honeybee. The authors attributed this to the differences between social and solitary insects, and the different selective pressures that each type of insect may be exposed to (Corona and Robinson 2006). With the completion of the mountain pine beetle genome (Keeling et al. 2013), a similar analysis can now be done on the mountain pine beetle antioxidant system. With bark beetles living the majority of their lives exposed to their host tree’s defensive compounds, it is likely that their primary oxidative stress proteins had to undergo unique adaptions in order to survive those harsh conditions Proposed Evolution of the Primary antioxidant proteins 20 The evolution of oxidative stress proteins is believed to be directly linked to the Earth’s Great Oxidation Event (Sheng et al. 2014). As cyanobacteria began to introduce oxygen gas to the environment, the production of the superoxide anion soon followed. Superoxide is a free radical that can be highly degradative to DNA and other biologically essential molecules, however, it displays varying reactivity within the different regions of the cell. When compartmentalized within organelles like mitochondria, that express their own superoxide dismutase, superoxide levels are maintained at minimal and nondamaging levels. Without this antioxidant system, the ROS generated would be extremely damaging when in contact with other cellular regions, ultimately resulting in cell death (Turrens 2003). The superoxide radical had a drastic impact on the metabolic processes used by the organisms of the time, resulting in the evolution of the enzymes in the antioxidant pathways. In order to catalyze the reduction and oxidation the of the superoxide anion it is believed that superoxide dismutase (SOD) was the first of the oxidative stress proteins to develop in the common ancestor of modern Archaeans ~2.42.0 billion years ago (Sheng et al. 2014). The evolution of oxidative stress pathways was not only key for withstanding the degradative impacts from ROS like superoxide and H2O2, but likely also for the development of some of the first cellular signaling pathways (Inupakutika et al. 2016). Once these ROS could be maintained at distinct REDOX levels, they could then be coopted by the cells for signal transduction. These molecules could be useful for indicating when the cell is experiencing stress conditions, and because different ROS have different properties (e.g. reactivity, diffusion rate, etc.), each ROS could trigger a unique response in the cell. It is likely that ROS could have been some of the first signal 21 molecules to be incorporated for cellular signaling – and oxidative stress proteins, like SOD, would have been crucial for their regulation (Inupakutika et al. 2016). Furthermore, there is support that the oxidative stress proteins themselves can serve as signal molecules within the cell, particularly the hyperoxidized form of peroxiredoxin. This peroxiredoxin conformation occurs when oxidative stress is high and is believed to act as some sort of signal, resulting in the expression of additional oxidative stress proteins (Neumann et al. 2009). Superoxide dismutase The evolution of SOD is believed to have occurred not only in response to increased oxygen in the environment but also to increased accessibility to REDOX-active transition metals which are transition metals that enable the SOD proteins to catalyze their redox cycling reactions (Jomova et al. 2012). The evolution of SOD enzymes has resulted in a classification of these enzymes based on their associated cofactors. Originally, iron (Fe2+) occurred in abundance in the world’s oceans and was used as the first metal cofactor of these original SODs, as well as the similar superoxide reductases (SOR), that evolved separately in prokaryotes (Sheng et al. 2014). However, as the atmosphere became increasingly oxygenated, the available Fe2+ was converted into insoluble ferric oxyhydroxides. Manganese (Mn2+) and Nickel (Ni2+) cofactors were then employed as alternatives to iron (Inupakutika et al. 2016). As the atmospheric concentration of oxygen increased, copper (Cu) and zinc (Zn) were gradually oxidized into the soluble Cu2+ and Zn2+ ions and became the primary metal cofactors used by most SOD and SOR variants (Sheng et al. 2014). These are the metal cofactors still used by cytosolic and perixosomic SODs, with cytosolic SODs 22 evolving much earlier than peroxisomic SODs. Fe-SOD is found within the chloroplasts and Mn-SOD within the mitochondria, a characteristic that can be attributed to the bacterial origins of these organelles. There is also a Ni-SOD variant that is restricted to gram-negative bacteria and cyanobacteria (Inupakutika et al. 2016). Mammalian SODs are among the most thoroughly characterized enzymes of this type because of their essential role in human health. For instance, a mutation in one of the subunits of the human Cu/Zn-SOD1 gene has been linked to an increase amyotrophic lateral sclerosis (ALS), or Lou Gehrig’s disease (Sheng et al. 2014). However, in insect systems, such as mountain pine beetle, these enzymes are not as well characterized. That said, although it is not specific to insects, generally it is understood that mitochondrial Mn-SODs primarily break down the superoxide anion surges referred to as “superoxide flashes” that occur as a by-product from cellular respiration. Some cytosolic Cu/Zn SODs are located near the mitochondria and are believed to further assist with this, but others are located in other regions of the cytosol and could break down the superoxide anions produced by peroxisomes and other degradative structures (Sheng et al. 2014). The exact role of extracellular Cu/Zn SOD in insects is even more tentative. Their function could be particularly important for an organism like the mountain pine beetle that is actively breaking down high concentrations of toxic compounds (e.g. terpenes and phenols) and consistently needing to mitigate the damaging effects of the resulting free radicals that are produced (Parker et al. 2017). Furthermore, the fungal and microbial symbionts are also actively breaking down these compounds, so it is likely that there would be a large quantity of superoxide radicals produced in the extracellular matrix that the beetles would need to contend with (Gretscher et al. 2016). 23 When SOD breaks down the superoxide anion, it produces oxygen and hydrogen peroxide (H2O2). Although less damaging than the superoxide anion, H2O2 is still a reactive oxygen species and required the evolution of new oxidative stress proteins in order to mitigate its damaging effects. There is evidence to support that shortly after its evolution, the original gene encoding SOD underwent a duplication event in the initial ancestral organism. Subsequent mutations in each of the duplicates led to one becoming the original Cu/Zn-SOD gene, and the other becoming the first catalase protein (Klotz & Loewen 2003). Catalase Originally discovered and named in 1900 (Loew 1900), catalases have become the most commonly studied antioxidant proteins due to their presence across all life forms and relatively straightforward reaction mechanisms. Catalase breaks down two H2O2 molecules into two water molecules and one oxygen molecule, and as such is a key pH regulator. Disrupted catalase activity has been linked to a number of diseases (e.g. Alzheimer's disease) (Nandi et al. 2019). With the exception of C. elegans, which has two catalase genes, most organisms are known to only contain one catalase gene (Petriv and Rachubinski 2004). Typically, catalase proteins lack any kind of signal peptide, restricting them to the cytosol, and are large tetramers typically comprised of four polypeptide chains that are ~500 amino acids long (Ashokan et al. 2011). Three families of catalase have been defined: typical heme containing catalases, atypical heme containing catalases, and manganese catalases (Zamocky et al. 2009). Manganese catalases represent only a small portion of catalase proteins (30 proteins) found within bacteria. Phylogenetic studies show that this family diverged very 24 early on in the evolution of catalases (Mhamdi et al. 2010; Zamocky et. al. 2009). The atypical heme containing catalases are absent in Animalia and Plantae. The most abundant type of catalase – the typical heme containing catalases – are found across all Archaea, Bacteria, Eukarya (including Plantae and Animalia). The spread of this family of catalase across these disparate life forms is believed to be due to multiple instances of lateral gene transfer events between eukaryotes and their microbial symbionts (Zamocky et al. 2009). Glutathione peroxidase A study conducted by Margis et al. (2008), supported that glutathione peroxidase does not have a linear evolutionary history. The evidence indicates that several lateral gene transfers resulted in the formation of the different GPx genes. Margis et al. (2008) were able to organize the different glutathione peroxidases into three polyphyletic groups: Group 1 (metazoans), Group 2 (fungi, bacteria, cyanobacteria), Group 3 (plants). GPx catalyzes the breakdown of H2O2 using either the glutathione or thioredoxin (TPX) molecule as an electron donor. In mammals, glutathione is the primary electron reductant, and initially it was believed that genes encoding enzymes with GPx activity were absent from insect genomes. However, with the completion of the Drosophila melanogaster genome, two GPx homologs were discovered that use TPX as the primary reductant. As a result, they were labeled as GPx homologs with TPX activity. Two GPx homologs were also identified within the Apis mellifera and Anopheles gambiae genomes (Corona and Robinson 2006). Most mammals have six distinct GPx isoforms, with GPx4 being the most similar to those found in insects. Mammalian GPx4 shows more similarity to the GPx4 isoforms found in other organisms (insects, plants, and bacteria), indicating 25 that the other mammalian GPx proteins share a distinct phylogenetic origin from the GPx4 proteins which are generally regarded as being phylogenetically ancient and having first evolved prior to the divergence of these different organismal groups (Margis et al. 2008). Peroxiredoxin Peroxiredoxins (Prx's), also known as thioredoxin peroxidases, are a family of proteins that are ubiquitous across all three domains of life (Knoops et al. 2007). In the last couple decades, the Prx family has become one of the most widely studied families of oxidative stress proteins, due to their prevalence and their many different functions (e.g. protein chaperones, oxidative stress mitigation, cellular signaling, etc.) (McGonigle et al. 1998). In addition, this group is unique among the previously mentioned oxidative stress proteins in that they lack metal cofactors, relying solely on the cysteines in their active site to facilitate catalysis. Although similar to GPx, they exclusively use thioredoxin as an electron donor instead of the glutathione molecule. Despite having a similar threedimensional structure and having functional similarities to other oxidative stress proteins, there is little sequence similarity between the peroxiredoxins and the other primary oxidative stress proteins (Chae et al. 1994). The Prx family of proteins is large and diverse in terms of their functional role and sequence similarity, making it challenging to categorize individual proteins. Nelson et al. (2010) were able to classify Prx’s based on the similarity of the Prx active site into six subfamilies. When constructing a tree with both insect and human homologs, it was found that the different Prx types formed distinct clades regardless of species. As a result, Nelson et al. (2010) came to the conclusion that different Prx types were present in the 26 common ancestor of both of these groups. This was further supported when it was determined that each member of these particular clades had the same subcellular location. There are more Prx proteins in eukaryotes than in prokaryotes, and they appear to have other functional roles in addition to reducing oxidative stress. When in their hyperoxidized stated, Prx proteins are believed to play a crucial role in cellular signaling. Bacterial Prx (Ahpc) shows limited susceptibility to hyperoxidation, indicating that it is probably not involved in cellular signaling. In contrast, many of the Prx proteins found in eukaryotes do demonstrate this susceptibility. Therefore, the expansion in the number of unique Prx proteins in eukaryotes is likely a reflection of the expansion in function that is required in more complex eukaryotic organisms (Bolduc et al. 2018). Despite their importance in other systems, all of the previously discussed proteins have yet to be characterized in mountain pine beetle. 2.2 Research Question 1) Are the primary oxidative stress proteins (catalase, glutathione peroxidase, superoxide dismutase, and peroxiredoxin) of the mountain pine beetle phylogenetically similar to the primary oxidative stress proteins of other insects with well characterized genomes? 2.3 Objectives 1) Conduct phylogenetic and bioinformatic analyses comparing the antioxidant genes of mountain pine beetle to two other Coleopterans and two species from Lepidoptera, Diptera, and Hymenoptera. 27 2.4 Hypothesis 1) The Oxidative stress proteins of the mountain pine beetle will cluster most closely with species that have a similar life history, such as Asian longhorn beetle. 2.5 Materials and methods Annotation of Dendroctonus ponderosae genes In order to conduct phylogenetic analyses on the mountain pine beetle oxidative stress proteins, homologous protein sequences were identified from eight other insect species. The following species used were selected based on availability of their respective annotated genomes: Tribolium castaneum (Herbst) (red flour beetle), Anoplophora glabripennis (Motschulsky) (Asian longhorn beetle), Bombyx mori (L.) (silk moth), Bicyclus anynana (Butler) (squinting bush brown), Apis mellifera (L.) (honey bee), Pseudomyrmex gracilis (Fabricius) (elongate twig ant), Drosophila melanogaster (fruit fly), and Anopheles gambiae (Giles) (African malaria mosquito). Each homologous sequence was obtained by inputting the accession number of the mountain pine beetle protein of interest into the BLASTp program at NCBI (Altschul et al. 1990). To further support that the generated protein was actually that particular homolog, the proteins that were generated from BLASTp were cross referenced with previous studies that worked with them to confirm that those authors also identified the protein as that particular homolog. The accession number for each protein is listed in Table 2 28 XP_624806 Prx5 - NP_001171540 GPx2 CAT KMQ89785 XP_020288373 XP_020280704 XP_020300222 XP_395319 NP_001171492 XP_020282120 XP_020289705 XP_020300009 NP_001164444 GPx1 Prx6b Prx6a XP_003249289 Prx1 NP_536731 - NP_728870 NP_523463 ABV32570 NP_0010271 91 NP_477510 NP_0011631 08 XP_016766983 SOD CCS XP_020295228 NP_476735 SOD3 NP_001037084 XP_0202291397 DM CAA35210 PG XP_020282033 SOD1 NP_001171498 AM BA - - NP_001037084 XP_023935668 BM NP_001037083 XP_023942413 ABV32570 - XP_023941442 - ABL09376 - NP_001036912 XP_023940851 - XP_003436755 NP_001036999 XP_023941251 XP_320690 XP_308753 XP_001238545 NP_001040386 XP_023942413 XP_308081 XP_018574477 XP_004922164 XP_023955178 AAS17758 XP_311594 AGa XP_018575508 XP_018566914 XP_018566915 XP_018580062 XP_018569307 XP_018575310 XP_018578437 XP_018574477 XP_0185694 XP_018579184 AG NP_001164309 NP_001164309 XP_008198961 XP_970660 XP_968419 XP_969254 XP_018578437 XP_975577 XP_975577 XP_968284 TC Table 2.1 - Major primary oxidative stress proteins (superoxide dismutase, peroxiredoxin, glutathione peroxidase, and catalase) for Tribolium castaneum (TC), Anoplophora glabripennis (AG), Bombyx mori (BM), Bicyclus anynana (BA), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Drosophila melanogaster (DM), and Anopheles gambiae (AGa). 29 Determination of protein characteristics For the mountain pine beetle oxidative stress proteins, the number of exons was obtained from the NCBI summary for each gene and the open reading frame size was determined using the NCBI ORF finder (NCBI Resource Coordinators 2018). Subcellular localization of the proteins was determined using WoLF PSORT (Horton et al. 2007) and confirmed with SignalP-5.0 (Almagro et al. 2019). Protein weight was determined using BioSuite Sequence Manipulation Suite (Stothard 2000) and confirmed in SWISS-MODEL (Guex et al. 2009). It was difficult to determine the specific subfamily of some of the typical one-cysteine Prx proteins, so the Deacon Active Site (Genbank 9nr, Nov 2010 edition) was used to determine the particular subfamily for these Prx proteins. The N-terminal mitochondrial sequences in the Prx5 sequences were identified using TargetP-2.0 (Almagro et al. 2019). Phylogenetic analysis The different types of each protein were grouped together for the different phylogenetic investigations. Following the example of Wang et al. (2008) and Wang et al. (2016) it was deemed prudent to separate the atypical peroxiredoxins (the oxidizing and resolving cysteines are on the same subunit) and typical peroxiredoxins (the oxidizing and resolving cysteines are on different subunits) as these proteins are dissimilar enough that grouping them together makes it difficult for either family to resolve properly. Multiple sequence alignments were created using CLUSTALW 2.1 in Geneious Prime 2020.0, under the following conditions: gap opening=15.0, gap extension= 0.3 and using a protein weight matrix of the Gonnet series. Any gaps in the alignments were 30 removed as missing information in a pairwise manner. Amino acid substitutions were identified across the key protein domains, and deemed conserved, semiconserved, or nonconserved using the NCBI BLOSUM62 database. Based on the generated amino acid alignments, ProtTest 2.4 (Oakley et al. 2014) was used to determine the most suitable best-fit model to use for each protein family’s phylogenetic tree. A Whelan And Goldman model (WAG) was determined to be the most suitable for Prx, SOD, and GP. LG was deemed to be the most suitable model for catalase. Using these matrices derived from the best fit model, each neighbour joining tree was constructed using the PAUP 4.0 b10 (Swofford 2003) program implemented in Geneious Prime 9. In order to assess the significance of the branch distribution, a bootstrapping value based on 1000 replications was generated for each alignment. 2.6 Results Ten genes of interest were identified in the mountain pine beetle genome (Table 2.2). Both mountain pine beetle GPx proteins were cytosolic and roughly the same size, differing at only a few amino acids. There was no information on NCBI about the number of exons present for either of the mountain pine beetle’s GPx proteins, but all of the other proteins contained multiple exons. Two of the peroxiredoxin genes were assigned to the Prx6 subfamily, so one was labeled Prx6a (extracellular) and the other Prx6b (cytosolic). One mountain pine beetle SOD (AEE61971) was found to contain a copper chaperone domain (CCS) resulting in the relabeling of this protein from a SOD to a SOD-CCS to highlight the differing function of this enzyme (Table 2.2). Due to this reclassification, the mountain pine beetle SOD-CCS was not included in any phylogenetic analyses with the mountain pine beetle SOD proteins. 31 Table 2.2 - Primary oxidative stress proteins of Dendroctonus ponderosae and their key features. Information includes accession numbers, number of base pairs in their associated gene, number of base pairs in the open reading frame of their associated gene, number of exons, number of amino acids, weight in kiloDaltons, and subcellular localization. Nd is used to indicate unknown information. C is used to indicate protein found in the cytoplasm, E is used to indicate a protein found in the extracellular matrix, and m is used to indicate a protein found in the mitochondria. Gene Accession number Gene length (bp) ORF length (ORF) Number of Exons AA Protein size (kDa) Cellular localization SOD1 AEE63298 887 263 2 152 15.57 C SOD3 AEE63612 2107 285 5 204 21.05 E SODCCS AEE61971 2917 324 4 231 24.63 C Prx1 AEE61832 1157 571 3 195 21.57 C Prx5 AEE62552 1059 369 4 122 13.42 M Prx6a AEE61766 1141 813 4 270 26.48 E Prx6b AEE62426 1280 195 4 64 7.58 C GPx1 AEE61713 723 504 nd 167 19.09 C GPx2 AEE61775 930 504 nd 167 18.89 C CAT AEE61710 1649 1536 8 511 58.6 C Figure 2.1 is an alignment of the different insect catalase sequences. The alignment shows a high degree of conservation among the three catalytic amino acids, with no variation occurring. Although the heme ligand binding (F*RERI*ERVVHAKGAG) and active site motifs (R*F*YAD*H) also showed a high degree of conservation, there was variation in the heme ligand-binding motif at the second amino acid in the motif (the 65th site in the alignment). At this position there appears to be a conservative replacement between the hymenopterans (both species), Asian longhorn beetle, and mountain pine beetle, which contained an asparagine residue, 32 with the remaining sequences that contained aspartic acid. Within the active site motif there was more variation between the insect catalases. However, the mountain pine beetle sequences did not have any substitutions within this region. 33 H N Y Figure 2.1 - Alignment for catalase sequences from Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (Aga), and Anoplophora glabripennis (AG). The heme-ligand binding motif (F*RERI*ERVVHAKGAGA) and the active site (R*F*YADT) are highlighted in a black box. The three catalytic amino acids are identified with red letters above their locations (Qin et al. 2016). 34 The phylogenetic analysis of catalase showed the proteins separated based on order (Figure 2.2). Mountain pine beetle catalase was most similar to that of Asian long horn beetle (Figure 2.2). As the lowest bootstrapping value was 73, these groupings have relatively high statistical support. Figure 2.2 - Neighbour-joining tree of insect catalase sequences: Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (AGa), and Anoplophora glabripennis (AG). The tree was constructed using the distance matrix based on the LG model derived from catalase amino acid sequences obtained from NCBI. The tree was prepared using Geneious Prime. The GenBank accession number is displayed for each sequence and the bolded values above each node represent the bootstrap support, and the values below each branch indicate branch length. Comparison of insect glutathione peroxidase amino acid sequences showed that there were some amino acid differences within the three conserved glutathione peroxidase domains (Figure 2.3). The most notable of these was in the domain closest to 35 the N-terminus where both mountain pine beetle GPx’s had a lysine substitution. Many of the other species also had a lysine at this site, but the four other coleopteran GPx proteins had a glutamine (Figure 2.3). Figure 2.3 - Alignments for glutathione peroxidase sequences from Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (AGa), and Anoplophora glabripennis (AG). The conserved glutathione peroxidase domains (A, B, and C respectively) are highlighted in hollow black boxes (Jiu et al. 2015). Phylogenetic analysis of insect glutathione peroxidases showed that when both were present, the GPx1 and GPx2 sequences formed a species-specific cluster (Figure 2.4). An exception was the Asian longhorn beetle where a species-specific clade did not form. Asian longhorn beetle did still group with the rest of the coleopterans, forming a distinct clade, the rest of the insects sequences also grouped together based on order. The mountain pine beetle clade was basal to the other coleopterans. That said, the bootstrap value associated with the divergence of the coleopteran clade is low (45). 36 Figure 2.4 - Neighbour-joining tree of insect glutathione peroxidase sequences Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (AGa), and Anoplophora glabripennis (AG). The tree was constructed using the distance matrix based on the Whelan and Goldman (WAG) model from glutathione peroxidase amino acid sequences obtained from NCBI. The tree was prepared using Geneious Prime. The tree was prepared using Geneious Prime. The GenBank accession number is displayed for each sequence and the bolded values above each node represent the bootstrap support, and the values below each branch indicate branch length. The alignment of insect SODs did showcase some variability within the conserved SOD domains (Figure 2.5). DP SOD1 had an amino that differed from the consensus in each of those domains, but overall, DPSOD3 contained more substitutions that deviated from the consensus within these domains. 37 C C SS Z SS SS Z ZZ C SS Figure 2.5 - Alignments for superoxide dismutase sequences from Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (AGa), and Anoplophora glabripennis (AG). Conserved superoxide dismutase domains are highlighted in a black box. Amino acids involved in zinc and copper binding are identified with red letters above their locations, labeled with a Z or a C, respectively. The amino acids identified with a red S are involved in disulphide formation (Parker et al. 2004). The phylogeny of insect SOD’s showed a distinction between the SOD1 and SOD3 groups, with each forming a separate monophyletic clade (Figure 2.6). Generally, the proteins separated based on order. The main exception being African malarial mosquito SOD1, which was basal to the rest of the proteins. The mountain pine beetle 38 isoforms of each protein showed the most similarity with the respective Asian longhorn beetle isoforms. There was a high degree of bootstrap support associated with the branch order of the SOD3 clade (97) but not with the SOD1 clade (32). The bootstrap support within each clade was low. Figure 2.6 - Neighbour-joining tree of insect superoxide dismutase sequences Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (AGa), and Anoplophora glabripennis (AG). The tree was constructed using the distance matrix based on the Whelan and Goldman (WAG) model derived from superoxide dismutase amino acid sequences obtained from NCBI. The tree was prepared using Geneious Prime. The tree was prepared using Geneious Prime. The GenBank accession number is displayed for each sequence and the bolded values above each node represent the bootstrap support, and the values below each branch indicate branch length. Comparison of the Prx amino acid sequences showed that the GGLG and YF hyperoxidation motifs were only present in the Prx1 sequences (Figure 2.7). However, the mountain pine beetle Prx1 sequence was the only Prx sequence that did not contain 39 the traditional YF motif, but instead underwent an amino acid substitution resulting in a FF motif. Hypersensitivity motif A was present also present in only the Prx1 sequences, excluding the fruit fly Prx sequence. Hyperoxidation motif B was not present in any of the sequences. 40 Motif A Motif A Motif B GGLG YF Figure 2.7 - Alignments for peroxiredoxin 1 and peroxiredoxin 6 sequences from Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (AGa), and Anoplophora glabripennis (AG). The YF and GGLG hypersensitivity motifs are identified with red letters above their locations. The locations of hypersensitivity motifs A (D-X8-N/G-X10-H-X27-S/G) and B (T-X3-S/T) are highlighted in black boxes and labelled with their respective names (Bolduc et al. 2018). 41 The different Prx1 and Prx6 proteins form distinct phylogenetic groups based on protein type rather than species (Figure 2.8). For each protein type, the proteins formed species-specific clades, with the exception of the Prx1 group where the Asian longhorn beetle’s sequence is basal to the rest of the Prx1 sequences. The branch order of each of the Prx6b, Prx6a, and Prx1 clades indicated a high degree of bootstrap support at 96, 78, and 100 respectively. 42 Prx6b Prx6a Prx1 Figure 2.8 - Neighbour-joining tree of insect peroxiredoxin sequences Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (AGa), and Anoplophora glabripennis (AG). The tree was constructed using the distance matrix based on the Whelan and Goldman (WAG) model derived from peroxiredoxin amino acid sequences obtained from NCBI. The tree was prepared using Geneious Prime. The GenBank accession number is displayed for each sequence and the bolded values above each node represent the bootstrap support, and the values below each branch indicate branch length. Among the Prx5 proteins, there was some variability in the composition and length of the mitochondrial leader sequences, but none of the proteins had a peroxisomal 43 target sequence at their C-terminal (Figure 2.9). The conserved active site residues showed no variation among the different sequences. V T R Figure 2.9 - Alignments for peroxiredoxin 5 sequences from Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (AGa), and Anoplophora glabripennis (AG). Predicted mitochondrial presequences are highlighted in the hollow boxes and conserved active site residues are marked with a red letter (Cox et al. 2010). The Prx5 proteins separated based on species, except for both of the Prx5 proteins from the Dipterans, which displayed more of a ladderlike structure, but formed a general monophyletic clade with the Lepidopteran proteins. The coleopteran and hymenopteran clades demonstrated a high degree of bootstrap support with 100 and 91 respectively, but the dipteran and lepidopteran clade only had an associated bootstrap value of 50. 44 Figure 2.10 - Neighbour-joining tree of insect peroxiredoxin 5 sequences Dendroctonus ponderosae (DP), Drosophila melanogaster (DM), Bombyx mori (BM), Bicyclus anynana (BA), Tribolium castaneum (TC), Apis mellifera (AM), Pseudomyrmex gracilis (PG), Anopheles gambiae (AGa), and Anoplophora glabripennis (AG). The tree was constructed using the distance matrix based on the Whelan and Goldman (WAG) model derived from peroxiredoxin 5 amino acid sequences obtained from NCBI. The tree was prepared using Geneious Prime. The GenBank accession number is displayed for each sequence and the bolded values above each node represent the bootstrap support, and the values below each branch indicate branch length. 2.7 Discussion In this study, I conducted a comparative and phylogenetic analysis of the mountain pine beetle oxidative stress proteins. Some of the trees that I generated resolved based on insect order, with Dipterans and Lepidopterans typically grouping together and separately from the Coleopterans and Hymenopterans. These findings suggest that the evolution of these antioxidant proteins parallel the species evolution and that these proteins have an important conserved function (Misof et al. 2014). Misof et al. (2014) conducted phylogenetic analyses on 1478 protein-coding insect genes, looking at both nucleotide and amino acid sequences. From their generated trees, they proposed that Hymenopterans first evolved during the Permian period with the Coleopterans appearing shortly after. The Lepidopterans and Dipterans did not appear until the Triassic; therefore, these two orders are more closely related to each other than to Coleoptera or 45 Hymenoptera. A configuration that supports this hypothesis was observed in the trees for catalase, SOD, Prx5, and Prx6b. Mountain pine beetle catalase, SOD, and Prx6b also demonstrated the greatest similarity to their respective Asian longhorn beetle counterparts. The similarity between the mountain pine beetle and Asian longhorn isoforms for each of these proteins was supported by both high bootstrap values and similar branch lengths. The exception was SOD1 which had a lower bootstrap value of 32. Catalase and SOD are two of the primary proteins involved in free radical breakdown; therefore, the fact that these proteins show the most similarity to the Asian longhorn beetle homolog is not surprising in that the two insects do share a common life history, with the majority of their development occurring under the bark of trees (Faccoli et al. 2014). As a result, they would both require key antioxidant proteins like SOD and catalase to function similarly to mitigate the defensive chemicals of their respective host trees. DPPrx1, DPPrx5, and DPPrx6a did not group with their counterparts from Asian longhorn beetle, and with the exception of Prx6a, these groupings had high bootstrap support. This could potentially be attributed to the diversity in function that has been described within peroxiredoxin proteins (Bolduc et al. 2018). Peroxiredoxins have been found to play many roles in addition to reducing oxidative stress. Therefore, they may serve different physiological roles in mountain pine beetles and Asian longhorn beetles, despite the similar environments these organisms share. In particular, this could be the case for mountain pine beetle Prx1 as it did contain a variation of the YF hypersensitivity motif, which will be further discussed in Chapter 3. As previously discussed, hyperoxidized peroxiredoxin has been found to act as a cellular signal. Therefore, the 46 presence of this motif in mountain pine beetle and not in Asian longhorn beetle supports that Prx1 may play a more multifaceted role in the mountain pine beetle or the Asian longhorn beetle. Regarding catalase, there is only one isoform of this protein present in most organisms, and it is regarded as one of the primary proteins involved in the mitigation of H2O2 (Nandi et al. 2019). Consequently, it is especially true that the respective mountain pine beetle and Asian longhorn beetle catalases would need to function similarly to manage the similar environments both organisms inhabit. This was reflected in the amino acid alignment of these two proteins, where among key amino acid residues, variation was only seen for two residues (Figure 2.1). Within the Heme ligand-binding motif, mountain pine beetles and Asian longhorns showed variation when compared to the other sequences, but even then they both contained identical conserved aspartic acid to asparagine substitutions. Massoumi-Alamouti et al. (2014) looked at specific genes coding for detoxification proteins in several pine pathogen species found within the genus Grossmania. Their findings supported that species adapted to a particular pine tree shared certain gene variants, which could be reflective of the specific pine chemistries they are exposed to. As supported by the work done by Massoumi-Alamouti et al. (2014), the presence of these identical substitutions in the catalase sequences of both mountain pine beetle and Asian longhorn beetle could be due to the shared selective pressures each of these beetles’ experiences within their associated host trees. This sequence similarity supports that the proteins would function similarly under oxidative stress conditions. Mountain pine beetle catalase is the largest of its primary oxidative stress proteins. A reason for the dramatic size difference between catalase and the other 47 oxidative stress proteins that perform a similar role, is that the catalase gene is believed to have initially evolved from a duplication mutation within the Cu/Zn-SOD gene, resulting in a larger gene and gene product (Corona and Robinson 2006)(Table 2.1). Originally, catalase was considered to be the sole scavenger of H2O2 in insects. This was attributed to the reduced number of insect glutathione peroxidases that had been identified (Yamamoto et al. 2005). However, despite that the number of identified glutathione peroxidase isoforms in insects is reduced, these proteins in addition to peroxiredoxins have been identified within most insect orders, so catalase no longer has the distinction of being regarded as the sole H2O2 scavenger in insects (Yamamoto et al. 2005). All of the trees containing different protein isoforms demonstrated the formation of distinct clades based on the specific protein isoform, with the exception of the GPx tree. Instead, the GPx tree demonstrated distinct clustering by species when both GPx1 and GPx2 sequences were present (Figure 2.4). This species grouping perhaps is indicative of functional diversification of the different GPx proteins that would have occurred after the species diverged from one another. Wang et al. (2001) studied the antioxidant systems of insect species and found that GPx activity was generally highly reduced. This could potentially account for our inability to generate so many GPx2 orthologs. Aside from the elongated twig ant, the coleopterans were the only species to have both GPx1 and GPx2. Potentially, the duplication event that resulted in both orthologs may have occurred after the divergence of coleoptera from the other insect groups (Wang et al. 2001). In particular, the lepidopterans that I examined were found to have a smaller suite of available antioxidant 48 proteins as neither species had extracellular SOD or Prx homologs (Table 2.2). This suggests that this insect order may have evolved a different oxidative stress strategy. From the mountain pine beetle genome (Keeling et al. 2013), a putative SOD protein (AEE61971) was queried against a sequence database to find homologous proteins in other organisms. Interestingly all of the matches were identified as copper chaperone proteins (CCS) and when these sequences were aligned, this AEE61971 gene and its homologous proteins did have some of the conserved SOD domains, but others were missing, most notably the zinc binding domains. Based on these findings, it is likely that AEE61971 is not actually a SOD protein but actually a copper chaperone for SOD (CCS). CCS proteins have some conserved domains associated with SOD proteins, amino acid residues involved in copper binding, and an N-terminal heavy metal associated domain (Corona & Robinson 2006). The latter feature is unique to these proteins and present in the AEE61971 protein. The function of the CCS protein is to bring and incorporate the copper cofactor to the Cu/Zn SOD as it is folding. The structure of the CCS protein generally mirrors that of a Cu/Zn SOD protein – a heterodimer with copper binding residues (Figure 2.5). This similarity in structure is believed to facilitate proper recognition, binding, and activation between the CCS and SOD proteins (Boyd et al. 2018). As a result, the CCS genes share similar sequence similarities with that of SOD proteins. Based on this similarity, when the mountain pine beetle draft genome was initially characterized, this gene may have been misidentified as a SOD protein. Zhang et al. (2017) investigated the SOD proteins of Dastarcus helophoroides (Coleoptera: Bothrideridae) and found two distinct groups consistent with the 49 intracellular SOD1 and extracellular SOD3 groups. This distribution is consistent with the SOD phylogenetic tree that I constructed for this study. As discussed by Corona and Robinson (2006), this indicates that these proteins evolved independently of one another. The general separation between the extracellular SOD3 group and intracellular SOD1 group is consistent with other studies that have been conducted looking at the phylogeny of SOD proteins (Zhang et al. 2017; Corona and Robinson 2006) Based on the branch lengths associated with each tree, the extracellular SOD3 family showed less diversification in their amino acid sequences than the SOD 1 family. This was surprising, as Laukkanen et al. (2015) found that the SOD3 family is directly involved in mediating oxidative stress from the environment as well as in cellular signalling, so I was expecting a greater degree of variation to be observed within this group, based on the differing conditions found within the environments of each organism. Due to the general reduction in measurable GPx protein activity in insect cells (Wang et al. 2001), the overall evolutionary trend is to see a greater number of peroxiredoxin isoforms in insects compared to other organisms. This appears to be the case in mountain pine beetle as there are two Prx6 isoforms (an extracellular and intracellular variant) in its genome with homologous Prx6 proteins being found in the other insect genomes as well. In addition, these proteins often display a higher reactivity, carrying out their associated reactions much faster (Wang et al. 2016). Consistent with the work done by Wang et al. (2016) looking at silk moth peroxiredoxin proteins, the typical one-cysteine peroxiredoxins (Prx6), and the typical two-cysteine peroxiredoxins (Prx1) formed distinct groups, with the latter group containing the conserved FVCP and EVCP regions characteristic of these families (Figure 2.7; Figure 2.8). Corona and 50 Robinson (2006) phylogenetically compared honeybee peroxiredoxins to other invertebrates and vertebrates and also found that the proteins separated based on the specific protein homolog rather than species and suggest that these peroxiredoxin families may have diverged prior to the speciation of invertebrates and vertebrates. This conclusion is further supported by the distribution of the peroxiredoxin tree in Figure 2.8. Some Prx5 proteins have a peroxisomal SQL sequence near their C-terminus, such as PMP20 from Candida boidinii (a species of yeast). The presence of this sequence can result in the peroxiredoxin being localized to the mitochondria or the peroxisomes. All the Prx5 sequences investigated in this study had a mitochondrial leader sequence and lacked an SQL peroxisomal sequence, supporting that these proteins are exclusively mitochondrial (Knoops et al. 2011). With Prx5 located within the mitochondria, its origins are likely bacterial having been incorporated into the eukaryote genome around the same time that the mitochondrion was also incorporated into the eukaryote cell (Djuika et al. 2015, Wang et al. 2016). In addition to mediating oxidative stress from H2O2, the hyperoxidized form of some peroxiredoxins have been found to play an important role in cellular signalling. In general, the typical two-cysteine peroxiredoxins are more susceptible to hyperoxidation than the atypical or typical one-cysteine peroxiredoxins. The YF and GGLG motifs have long been supported to increase susceptibility of peroxiredoxins to hypersensitivity in Prx1 proteins (Wood et al. 2003). Consistent with the current knowledge on these motifs, none of the typical or atypical peroxiredoxins from mountain pine beetle contained these motifs. All of the insect Prx1 proteins did have the GGLG motif. They all also had the YF motif except for the Prx1 from mountain pine beetle. In this protein the tyrosine in the 51 motif was substituted with a phenylalanine, making it FF instead of YF. A previous study by Schneider et al. (2008) found that substituting a phenylalanine with a tyrosine in the active site of a fructose-6-phosphate aldolase protein resulted in a >70 fold increase in protein activity when compared to the wild type. This substitution` in such a key regulatory domain of the DPPrx1 protein, could have a considerable impact on the susceptibility of that protein to hyperoxidation (Figure 2.7). Recently Bolduc et al. (2018) identified two novel motifs that drastically impact the overall sensitivity of 2-cysteine Prx proteins to hypersensitivity: Motif A and Motif B. It was found that in human and bacterial Prx proteins the different combinations of the four motifs could have considerable impacts on the overall hypersensitivity of the protein. Despite being present in bacterial and human Prx proteins, motif B has not been identified in any of the other insect sequences. Bolduc et al. (2018) compared the relative hypersensitivity of 2-cysteine Prx proteins with varying combinations of these hypersensitivity motifs. However, they did not look at the reactivity of a peroxiredoxin with just GGLG and Motif A. Therefore, mountain pine beetle Prx1 could offer a unique opportunity to further expand on how these domains impact overall reactivity, particularly when the modified YF motif is also taken into account. 2.8 Conclusions and Future Directions In this study I examined the sequence structure and phylogenetic relationships of the primary oxidative stress proteins of the mountain pine beetle, which to my knowledge has previously not been done for a species of bark beetle. Overall, I found that the key components of the antioxidant system in insects are also conserved in the mountain pine beetle. By examining the sequence structure of these genes, it allowed for any unique 52 features in the mountain pine beetle oxidative stress proteins to be identified. One of the most notable differences was that the mountain pine beetle was the only insect to have its Prx1 containing a FF instead of the conserved YF. A mutation in such a key regulatory motif could have critical impacts on the functionality of this protein and could be indicative of the unique life history of the mountain pine beetle compared to the other insects in this study. 53 Chapter 3: Functional Characterization of DPPRx1 and DPSOD1 3.1 Introduction Reactive oxygen species (ROS), including the superoxide anion (O2-) and hydrogen peroxide (H2O2) are considered primary lines of chemical defense in pine trees (Li et al. 2016). ROS are produced by trees as defensive compounds and are by-products of lignin breakdown during mountain pine beetle colonization (Birben et al. 2012). These molecules can be incredibly harmful as they can damage lipids, protein, and DNA. Therefore, mitigating ROS damage is essential to mountain pine beetle survival during colonization of pine trees. To my knowledge, no studies have been performed on the function of oxidative stress proteins in bark beetles. The mountain pine beetle has a number of oxidative stress proteins, including catalase, glutathione peroxidase, peroxiredoxin and superoxide dismutase, many of which showed upregulation during host colonization (Pitt et al. 2014; Robert et al. 2013). I was able to generate sufficient quantities of heterologously expressed mountain pine beetle peroxiredoxin-1 (DPPrx1) and superoxide dismutase-1 (DPSOD1), for functional characterization. SOD is essential as it is one of the only proteins able to breakdown the superoxide free radical (O2-) and peroxiredoxin performs a variety of vital antioxidant and signal transduction roles (Tasaki et al. 2018; Winterbourn 2008). One recent study by Li et al. (2016) investigated the role of a typical 2-cysteine peroxiredoxin (BxPrx) in the pine wood nematode, Bursaphelenchus xylophilus. Like the mountain pine beetle, pine wood nematodes feed on the parenchymal cells of pine species, and their population has reached epidemic levels in some regions, drastically impacting pine forests across Asia and Europe. In addition, under infestation conditions, Li et al. (2016) found that BxPrx expression was elevated, similar to the increased 54 expression of DPPrx1. The pine wood nematode is responsible for the pine wilt disease, which has had considerable ecological and economic impacts on Asian and European pine stands. A previous study found that 2-cysteine peroxiredoxin proteins are overexpressed in C. elegans (parasitic nematode), as they play a critical role in the breakdown of free radicals produced both metabolically within the nematode and as an immune response by the host organism (McGonigle et al. 1998). Although Li et al. (2016) did not determine a rate constant for BxPrx, they were able to assess its activity using a metal catalyzed oxidation (MCO) system, which assays for the ability of peroxiredoxins to prevent DNA degradation by H 2O2, and via disc assays, which compare the survival rate of bacterial cells grown in high concentrations of H2O2 when BxPrx is present and when it is absent. They found that BxPrx has a high antioxidant activity, indicating that it likely plays a role in mitigating the damage from free radicals that build up during host colonization. To date there has been no attempt to characterize the activity of any bark beetle peroxiredoxins. That said, Zhang and Lu (2015) looked at the antioxidant functions of an atypical peroxiredoxin (BMPrx5) from Bombyx mori (silk moth). Using quantitative real time PCR, they found that BMPrx5 was upregulated in the haemocytes of the silk moth under conditions triggering oxidative stress (i.e. bacterial infection and H2O2 injection). This response is consistent with the work done by Pitt et al. (2014) and Robert et al. (2013) who observed the upregulation of these genes in the transcriptome and proteome, respectively, of the mountain pine beetle under detoxification conditions. Using a combination of approaches that include an MCO system, disc diffusion, and viability assays (which also evaluates the bacterial survival when grown in high H2O2 55 concentrations), it was confirmed that silk moth peroxiredoxin displays considerable antioxidant functions (Zhang and Lu 2015). However, the authors did not characterize BMPrx5 by determining rate constants or pKa values. Based on these findings it is proposed that the physiological purpose of DPPx1 are to mitigate the damaging impacts of H2O2 through the cysteine residues within the active site of the protein having a high oxidizing tendency. However, the threedimensional structure of DPPx1 and its active site can greatly influence its overall activity and oxidizing potential. Similar to the previously described studies, the general affinity of DPPx1 for H2O2 can also be confirmed using a MCO system, but in order to gain a more in depth understand of the physiological role of DPPrx1 and its affinity for H2O2, a rate constant must be determined for its reaction with H2O2 (Nelson et al. 2008). Rate constant determination A competition-based assay with horseradish peroxidase is a relatively straightforward and affordable approach to generate a second order rate constant for a peroxiredoxin with H2O2 (Nelson and Parsonage 2011). Horseradish peroxidase (HRP) reacts with hydrogen peroxide (H2O2) resulting in an oxidized form of HRP called HRP compound 1, which absorbs less light at 403 nm compared to unreacted HRP. Peroxiredoxin has a higher affinity for H2O2 than horseradish peroxidase (HRP), as a result there is a reduction in the net decrease in absorbance at 403 nm (from HRP compound 1 formation) when a functional peroxiredoxin is present in a solution containing horseradish peroxidase and H 2O2. This allows for the fractional HRP inhibition to be calculated because it is proportional to the concentrations of peroxiredoxin in the reaction mix (Ogusucu et al. 2007). 56 Originally, rate constants for this reaction were acquired using steady state kinetics (kcat/kM). Using this approach, the rate constants for the typical 2-cysteine peroxiredoxins, Tsa1 and Tsa2 from Saccharomyces cerevisiae, were determined to be 104 to 105 M−1s−1, respectively (Park et al. 2000). This was surprising because for a protein that appears to be as essential as peroxiredoxin, it is expected that it would react more rapidly with its substrate and have a second order rate constant that is larger than 105 M−1s−1. However, using the horseradish competition assay, Ogusucu et al. (2007) measured rate constants that were two orders of magnitude larger than originally reported. These findings were further supported by kinetic approaches performed in a stopped-flow instrument, indicating that the horseradish competition assay garnered a more accurate rate constant than the steady state kinetic approach previously employed. Therefore, I used a horseradish competition assay to determine rate constants for DPPrx1 because of its accuracy and its ease of implementation. Due to DPPrx1’s seemingly important role during detoxification, based on transcriptomic and proteomic data, it was suspected that the generated rate constant would be close to the value of 107 M−1s−1 (Ogusucu et al. 2007). pKa determination The second order rate constant of horseradish peroxidase with H 2O2 is pH independent from pH 4.5 to 7.0, which makes the horseradish peroxidase competition assay a useful tool for determining the pKa of the active site of a peroxiredoxin protein, as it allows for a range of pH values to be tested (Dunford 1999; Job et al. 1978). This pK a value should be indicative of the pKa value of the oxidizing cysteine (CP); whereas, the 57 pKa of the resolving cysteine (CR) should be more comparable to that of a free cysteine group (pKa=8.4) (Ogusucu et al. 2007). Sensitivity to Hyperoxidation A physiologically important aspect of DPPx1 could be its potential role as a molecular signal. As discussed in Chapter 2, DPPx1 is a cytosolic homodimer, with each monomer containing a CP and a CR. It is critical to the catalytic function of 2-cysteine peroxiredoxin that the Cp is quickly deprotonated at the neutral pH of the cytosol, becoming a thiolate (CPS-). Therefore, the pKa of Cp is relatively low compared to that of free cysteines (Nelson et al. 2008). This thiolate has a high affinity for reacting with H2O2, to become a sulfenic acid (CPSOH). Condensation then occurs between the sulfenic acid and the resolving cysteine (SR), forming a disulfide linkage. Therefore, each peroxiredoxin protein is able to form two disulfide linkages and reduce four H2O2 molecules. The homodimer will eventually be reduced via disulfide exchange facilitated by the thioredoxin-thioredoxin reductase system (Ledgerwood et al. 2017). However, when free radical concentrations are high the sulfenic acid can be further oxidized to form a sulfinic acid (CPSO2H) and sulfonic acid (CpSO3H). These hyperoxidized states can no longer be reduced by thioredoxin, but sulfiredoxin, an ATPdependent enzyme that is much slower, can reduce sulfinic acid. The conversion of the thiolate to sulfenic acid is believed to be an irreversible reaction. As hyperoxidized peroxiredoxin is catalytically inactive, this eliminates any antioxidant capacity of the enzyme (Ledgerwood et al. 2017). 58 !CpS% %%HSCR!% !CRSH% %SCP!% (Reduced*State)% H2O2 H2O ThioredoxinThioredoxin Reductase System Sulfiredoxin !CPSOH%%HSCR!% H !CRSH%%HOSCp!% %%%%%%%(Sulfenic*Acid)* Low H2O2 Concentration !CpS%!SCR!% High H2O2 Concentration !CPSO2H% %HSCR!% !CRSH%%HOS2Cp!% !CRS%!SCp!% (Disulfide*Forma0on)* (Hyperoxidized*State)% Figure 3.1 - The reaction mechanism for a typical 2-cysteine peroxiredoxin, adapted from Ledgerwood et al. (2017). The homodimer is represented by the yellow and green blocks with their respective cysteine residues. The fork in the diagram indicates the two pathways that lead to either the disulfide or hyperoxidized state. The reduction pathways for each of the oxidized forms and the main enzymes involved are represented by the curved arrows. Initially, the propensity for peroxiredoxin to undergo hyperoxidation under extreme oxidative stress seems counterintuitive to the role of an oxidative stress protein. However, there is evidence that hyperoxidized peroxiredoxin serves a valuable role as a messenger within the greater antioxidant response systems of a cell. In S. cerevisiae, thiol-based peroxidases have been found to act as ‘peroxide sensors’, with their hyperoxidized forms oxidizing the transcription factor Yap1 (Delaunay 2002). This transcription factor regulates the expression of a glutathione peroxidase protein called Orp1. This protein has a much higher affinity for H2O2 than peroxiredoxin and will 59 therefore be more effective at reducing damage from free radicals (Delaunay 2002). A similar mechanism happens with Pap1, a homolog of Yap 1, found in Schizosaccharomyces pombe (Delaunay 2002). It is regulated by a hyperoxidized peroxiredoxin called Tpx1 (Vivancos et al. 2005). Similar mechanisms are also found in plants, such as Arabidopsis thaliana, where a 2-cysteine peroxiredoxin was shown to regulate the transcriptional activity of Rap24, a transcription factor that was seen to upregulate the expression of the plant’s antioxidant defense (Rudnik et al. 2017; Rhee et al. 2012). As discussed in Chapter 2, there are unique motifs that provide a peroxiredoxin molecule with varying degrees of sensitivity to hyperoxidation. Two key motifs that convey sensitivity to hyperoxidation are the primary sequences Glycine-GlycineLeucine-Glycine (GGLG) located near the middle of the peroxiredoxin active site, and Tyrosine-Phenylalanine (YF), located near the C-terminus of the proteins (Figure 3.1). These motifs restrict bonding between the sulfenic acid and the resolving cysteine, thus leaving the cysteine residues susceptible to the formation of sulfinic and sulfonic acids. It is common for the conserved YF sequence to be missing from typical 2-cysteine peroxiredoxins, which is the case for DPPrx1 (Kumagai et al. 2009). M S P A L QK P A P S F K GT A V V D GQ F K D I S L ED Y K GQ Y V V L F F Y P L D F T F V C P T E I I A F SD R I E D" E F K K I K T A V I GA ST D SH F SH L AW I N T P R K QGG L G SMN I P L L A D K N L E I A R S YGV L D E ST G *" H" *" I A F RG L F I I D P K G I L RQV T I N D L P V GR SV D E T L R L V Q A FQ F T D EH G E V C P A GWT PGK K T I K PQV D A SK E F F S ST N Figure 3.2 - Primary protein structure of DPPrx1. Functional cysteines, flanked by a valine and a proline (VCP), are boxed in blue. The conserved sensitivity motif GGLG is boxed in purple. Hyperoxidation motif A is underlined in red. Sites where a lysine was found in place of conserved glycine are highlighted with a *. 60 Because DPPrx1 is missing the YF motif that is associated with sensitivity to hyperoxidation (see Chapter 2) it is likely that the protein may not show susceptibility to hyperoxidation. However, Bolduc et al. (2018) did an intensive sequence analysis of bacterial and human peroxiredoxins. In addition to the YF and GGLG motifs, they identified two novel protein motifs associated with sensitivity to hyperoxidation – Motif A (See Figure 3.2) and Motif B (T-X3-S/T). In addition to sequence analysis, Bolduc et al. (2018) performed functional analysis on different peroxiredoxins and found it difficult to predict how these particular motifs impact catalytic function. Bolduc et al. (2018) found that the peroxiredoxins that displayed the highest susceptibility to hyperoxidation were those that contained the GGLG and YF motifs, with only either Motif A or Motif B. Consequently, the peroxiredoxins containing both Motif A and Motif B showed the greatest resistance to hyperoxidation, regardless of whether the GGLG and YF motifs were present. We can infer based on the presence of motif A, that the sensitivity of DPPrx1 to hyperoxidation is likely high. However, the presence of motif A with GGLG in DPPrx1, is a combination that was not seen by Boluc et al. (2018), so the exact effect it will have on hyperoxidation is unknown. Another unique aspect of the specific motif A present in DPPrx1 is that in the sites where bacterial and human peroxiredoxins have glycines, DPPrx1 has lysines. In other proteins that have experienced a mutation resulting in the conversion of a glycine to lysine, there is almost always an impact on the overall function, except in cases where the amino acids are not located near the active site (Tsuzuki et al. 2003). However, Motif A completely overlaps with the active site of DPPrx1 and includes one of its functional cysteines (Figure 3.2). Therefore, these mutations could have a considerable impact on 61 both the antioxidant and/or signal transduction functions of DPPrx1 (Haijema et al. 1996). Functional analysis of DPSOD1 Similar to peroxiredoxin, there is no prior research literature describing the activity of Cu/Zn-SOD in bark beetles. However, there have been studies looking at a Tribolium castaneum (red flour beetle) SODs (Ferro et al. 2017; Behrens et al. 2014). The detoxification system of the red flour beetle is an interesting one in that it the insect has experienced extensive selective pressures from insecticide usage. This is similar to how the mountain pine beetle has existed alongside the selection pressure context of host tree defensive metabolites. Using transcriptomic data collected from red flour beetle populations infected by Bacillus thuringiensis, a common bioinsecticide used in agriculture, Behrens et al. (2014) determined that oxidative stress proteins experienced significant upregulation. Based on this upregulation, red flour beetle SOD, as one of the primary oxidative stress proteins, was deemed a suitable selection for functional analysis. Ferro et al. (2017) observed that the red flour beetle SOD, when ‘primed’ via exposure to Bacillus thuringiensis infection, displayed a distinct increase in SOD activity which was determined based on its ability to inhibit nitrolium blue tetrazolium. 3.2 Research Question 1) Do the primary oxidative stress proteins (catalase, glutathione peroxidase, superoxide dismutase, and peroxiredoxin) of the mountain pine beetle have unique functional features (i.e. increased reactivity)? 62 3.3 Objectives 1) Design and optimize a protocol allowing for the expression and purification of mountain pine beetle superoxide dismutase 1 and peroxiredoxin 1. 2) Functionally characterize mountain pine beetle superoxide dismutase 1 by performing a water-soluble tetrazolium (WST-1) salt competition assay to determine how much is required to achieve 50% inhibition of xanthine oxidase (IC50). 3) Determine the pKa and second order rate constant values for mountain pine beetle peroxiredoxin 1 by performing a competition-based horseradish peroxidase assay. In addition, perform a nonreducing SDS-PAGE to determine the susceptibility of mountain pine beetle peroxiredoxin 1 to hyperoxidation. 3.4 Hypothesis 1) Based on the unique life history of the mountain pine beetle, their oxidative stress proteins will demonstrate unique functional features compared to the oxidative stress proteins of other organisms. 3.5 Materials and Methods Entry Clone Construction ElectroMAX DH10B T1 Phage-Resistant electro-competent cells containing the genes of interest were retrieved from a mountain pine beetle cDNA library glycerol stock and were grown at 37°C on lysogeny broth (LB) agar plates containing chloramphenicol (30 μg/ml). The following day, four colonies were selected and grown overnight at 37°C in 5 mL of LB containing chloramphenicol (30μg/ml). Plasmids were extracted using the QIAprep ® Spin Miniprep kit (Qiagen). 63 The Gateway® Cloning Procedure TM (Invitrogen) was selected to introduce genes of interest into One Shot® BL21-AI™ E. coli (Invitrogen) competent cells via the plasmid pDEST-17 (Invitrogen) as it employs a single recombination reaction. To insert the genes of interest, specific recombination sites (attB sites) that are 25 base pairs long are needed to flank the genes. These were incorporated into the primers (Table 3.1) prior to polymerase chain reaction (PCR) amplification. Primers for DPPrx1 and DPSOD1 were designed and purchased from Integrated DNA technologies (Table 3.1). Table 3.1 - Cloning primers for DPPrx1 and DPSOD2, ATTB sequences are bolded Primer DPPrx1F DPPrx1R DPSOD1F DPSOD1R Sequence (5’–3’) GGGGACAAGTTTGTACAAAAAAGCAGGCTTAATGTCTCCTGCACTTCAGAA GGGGACCACTTTGTACAAGAAAGCTGGGTACTATTAGTTAGTTGAGCTGAAAA GGGGACAAGTTTGTACAAAAAAGCAGGCTTCATGGTTAAAGCCGTTGCCGTGTTG GGGGACCACTTTGTACAAGAAAGCTGGGTCCTACTAGTTGGGCTGTGCAAGG PCR reaction conditions were: 3 μL of extracted plasmid, 1X reaction buffer, 0.5 μL of dNTP’s, 2.5 μM of paired forward and reverse primers (containing attB sites), (Table 3.1), 0.25 μL (1.25 units) of Platinum Taq high fidelity DNA polymerase (Invitrogen), and 17.75 μL of milliQ-grade water. Reactions (25 μL) were run at 94°C for 1 minute; followed by 34 cycles of 94°C for 30 seconds, 66°C (DPPrx1) or 69°C (DPSOD1) for 30 seconds, and 72°C for 90 seconds; 72°C for 10 minutes. PCR product gene sequences were confirmed through restriction digestion analysis and dye-terminator Sanger sequencing at the UNBC NALS Genetics Laboratory. For the restriction digest, the purified plasmids (1 µg) were mixed with HindIII (1 unit) and digested at 37°C for 4 hours, and then run on a 1% agarose gel at 1.5 volts for 1.5 hours. Upon confirming that proper amplification and cloning had occurred, the 64 Gateway® Cloning Procedure was employed to transfer the plasmid to the One Shot® BL21-AI E. coli expression system (Figure 3.3). Gene*of*interest* +" pDONRTM221* 1" pDESTTM17* pDonr* +" 2" pDONRTM221* 1)"BP"reac on" *****Crea?ng*an** *****entry*clone* +" pDESTTM17* 2)"LR"reac on" *****Crea?ng*an** *****expression* *****clone* 3" BL213A1*E.#coli*cells* 3)"Plasmid"Transforma on* Figure 3.3 - Overview of the LR and BP steps adapted from the Gateway® Technology Cloning protocol (2003). The first step of the Gateway® Cloning Procedure, referred to as the BP reaction, is designed to incorporate the gene of interest to a transfer vector (pDONR™) (Figure 3.3). Depending on the amplicon concentration, which was measured using a Nanodrop spectrophotometer, 1–10 μL of the PCR product was mixed with 2 μL of pDONR™ (150 ng/μL) and 1–9 μL of TE buffer to make a final volume of 10 μL. Lastly, 4 μL of BP Clonase™ II (thawed on ice for 2 minutes and vortexed twice for 2 seconds) was added to the mixture. The reaction was then incubated at room temperature (~21°C) for 1 hour. To stop the reaction Proteinase K (2 mg/mL) was added and the reaction was incubated for ten minutes at 37°C. The transfer vectors generated during the BP reaction were transformed into competent cells. Frozen 50 μL aliquots of TOP10 library efficiency cells (stored at 80°C) were kept on ice until thawed, and then 1 μL of BP reaction mix was added, mixed 65 by lightly flicking the closed tube, and left on ice for 30 minutes. The cells were then heat shocked in a 42°C water bath for 30 seconds, and the tube was immediately transferred back to the ice and left to sit for one minute. SOC medium (450 μL) was added and the tube was shaken horizontally at 200 rpm for 1 hour at 37°C. Selection of transformed cells relies on the kanamycin resistance gene found in the pDONR™ plasmid. Therefore, 20 μL and 100 μL of the above cell mix were streaked on LB agar plates containing kanamycin (30μg/ml). Plates were incubated overnight at 37°C. From each plate four colonies were transferred to 5 mL of LB treated with kanamycin (30μg/ml) and incubated at 37°C overnight in a shaker (300 rpm). To confirm that the BP reaction had occurred successfully, pDONR™ plasmids were purified using the Qiagen QIAprepâ Spin Miniprep Kit and submitted to the UNBC NALS Genetics Laboratory for sequencing. The second part of the Gateway® Cloning Procedure is called the LR reaction and involves the transfer of the gene of interest from pDONR to the final expression vector (pDEST-17) by LR Clonase™ II. The steps of the LR reaction are similar to those of the BP reaction but using purified pDONR containing either DPPrx1 or DPSOD in place of PCR product (Figure 3.3). For cell transformation steps the BL21-AI™ strain of E. coli was used in place of library efficiency cells with ampicillin (30μg/ml) or carbenicillin (30μg/ml) instead of kanamycin for selection, as this is the antibiotic pDEST™-17 confers resistance to. Of the potential expression systems that could have been employed, BL21-AI™ cells were selected for their affordability, ease of use, and because they are deficient in two key proteases (lon and OmpT) that would otherwise reduce the overall yield of protein product (Amid et al. 2011). DPPrx1 induction 66 In order to optimize the amount of soluble protein produced, different induction parameters (e.g. temperature, rotary shaker rpm setting, induction time, and heavy metal supplementation) were compared. For DPPRx1, 50 μL of BL21-AI™ cell stock was grown overnight and added to a 250 mL Erlenmeyer containing 50 mL of LB growth media, and 500 μL of ampicillin (30μg/ml). Cultures were shaken at 300 rpm in a rotary shaker-incubator (Shel Lab) at a temperature of 37°C. The tops of the flasks were covered with tin foil to allow for air-exchange. After two hours, the optical density (OD) at 600 nm was measured using a SmartSpec™ Plus Spectrophotometer (Bio-rad). Following the protocol described by Wang et al. (2016), if they reached an OD of 0.4–0.6 the cultures were induced with 0.4 mM of arabinose. If the desired optical density had not yet been reached then the cultures were left to incubate further, and the OD was checked every 30 minutes and induced after the desired value was reached. After arabinose induction, the cultures were then left to shake and incubate at 300 rpm at 37°C for six hours. After the induction was finished, the cultures were aliquoted into both 15 mL conical and 1.5 mL microcentrifuge tubes and centrifuged at 13,000 rpm for three minutes using a Sorvall Legend Micro 17R centrifuge (Thermo Fisher) and a Sorvall Legend XIR centrifuge (Thermo Fisher) respectively. The supernatant was poured off, and the cell pellets were stored at -20°C until they were used for further analysis. DPSOD1 Induction Inductions were performed on the BL21-AI™ cell cultures containing DPSOD1 in LB growth media but did not result in the production of any soluble protein. Therefore, the minimal media induction conditions employed by Asmail (2008) were employed. The minimal media was made of 5 g/L glucose, 1 mg/L of biotin, 2 mg/L of thiamin, 1g/L of 67 ammonium sulphate, 0.125% Casamino acids, 25 mM phosphate buffer, 12.5 ng/mL of ampicillin, and 0.004% of heavy metal stock solution. Heavy metal stock solution was prepared with the following components: 100 mg MoNa2O4•2H2O, 90.4 mg CoCl2• 6H2O, 35 mg CuSO4• 5H2O, 200 mg MnSO4•H2O, 854.6 mg MgSO4• 7H2O, 250 mg ZnSO4• 7H2O, 500 mg CaCl2• 2H2O, 200 mg H3BO3, and FeCl2• 4H2O all mixed in 1 liter of 1M HCl. The mixtures were stirred overnight with a magnetic stir bar and filtered using WhatmanTM Qualitative filter paper: Grade 1 Circles (GE Healthcare). 50 mL of overnight cell culture, grown in LB media, was mixed with 50 mL of minimal media in a 300 mL Erlenmeyer flask covered with tinfoil and shaken at 37°C at 300 rpm until an OD of 0.6–0.8 was reached. At this point, the flasks were put on ice for five minutes and then arabinose was added to a final concentration of 0.5 mM. The cultures were then incubated at room temperature (~20°C), and mixed with a magnetic stir bar for 18 hours. After that, the cultures were both 15 mL falcon and 1.5 mL microcentrifuge ,tubes and centrifuged at 13,000 rpm for three minutes. The cultures collected in the 1.5 mL microcentrifuge tubes were run on an SDS-PAGE, using NuPAGE™ 4-12% Bis/Trispolyacrylamide gels (Invitrogen), to confirm protein presence and size. In addition, to assess protein solubility, cultures were centrifuged and SDS-PAGE was performed on both the pellet and supernatant. The subsequent purifications and assays were performed on the cultures stored in the 15 mL falcon tubes. The resulting supernatant was poured away, and the cell pellets were stored at -20°C for further analysis. Protein Purification 68 Pellets from the 15 mL aliquots were resuspended in 1 mL of GE His SpinTrap Binding Buffer and 50 μL of 200 mM phenylmethylsulfonyl fluoride (PMSF) and left to sit on ice for 30 minutes with regular intervals of stirring using a glass rod. Resuspended cell cultures were transferred to precooled 15 mL conical tubes and, while still on ice, sonicated using an Model 150T ultrasonic dismembranator (Fisher Scientific). The cell suspension was sonicated on ice using intervals of five pulses of 30 second duration on output five with a 30 second break in between pulsing cycles. The supernatant containing soluble SOD was recovered after sonication by centrifugation at 13,000 x g for 10 minutes. SDS-PAGE was performed on sub-samples of the pellets and lysates to compare levels of soluble and insoluble recombinant SOD produced during induction. The rest of the lysate was transferred to a new microcentrifuge tube and used for chromatographic purification of SOD after SDS-PAGE confirmation of solubility. To assess the size and quantity of the expressed protein, SDS-PAGE was performed with NuPAGE™ 4-12% Bis/Trispolyacrylamide gels (Invitrogen) and run for 1.5 hours at 125 V, under reducing conditions, and run with a 1 Kb protein ladder (Thermo Scientific). The protein purification protocol was adapted from a general protocols for purification with GE Healthcare His SpinTrap column. Before finalizing the purification protocol, five different imidazole concentrations were screened for the binding buffer conditions: 0 mM, 20 mM, 40 mM, 60 mM, 100 mM, and 300 mM. 20 mM resulted in the highest yield, which was also the concentration recommended in the GE purification protocol. His SpinTrap columns were inverted and shaken to disperse the medium and then centrifuged for 30 seconds at 70 x g. To prime the column, 600 µl of 20 mM binding 69 buffer was added to the column and spun for 30 seconds at 70 x g. 600 µl of lysate was applied to the column and incubated at room temperature for 30 minutes to promote maximum protein binding before centrifugation at 70 x g for 30 seconds. The column was washed with 600 µl of 20 mM binding buffer to remove protein contaminants, and then the protein was eluted with 200 µl of 500 mM elution buffer pH 7. This step was repeated in a separate tube to collect any remaining protein attached to the column. Protein quantification I used the Bradford protein assay (Bio-Rad) to quantify the amount of protein collected at different steps of the purification. Concentrated bovine serum albumen (BSA) was diluted with MilliQ-grade water to prepare BSA standards in the following concentrations: 0.0 mg/ml, 0.05 mg/ml, 0.1 mg/ml, 0.2 mg/ml, 0.4 mg/ml, 0.6 mg/ml, 0.8 mg/ml and 1.0 mg/ml. In a 96 well plate, 250 μL of Quick Start™ Bradford 1X reagent (Bio-Rad) was added to each well followed by 10 μL of the appropriate BSA standard or purified protein. The contents of each well were mixed by pipetting up and down, with care being taken to ensure that no bubbles formed, and then then the absorbance was read at 595 nm using a FilterMax F5 Multi-Mode Microplate Reader (Molecular Devices). Each Bradford standard and protein sample were run in triplicate. Using the absorbance reading from the different BSA standards, a standard curve was generated and could be used to determine the protein concentration of the unknown samples (Figure 3.4). 70 1.6 y = 1.0009x + 0.3694 R² = 0.978 Absorbance at 595 nm 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 0.2 0.4 0.6 0.8 Concentration (mg/ml) 1 1.2 Figure 3.4 – A standard curve generated from BSA standards of differing concentrations and their respective absorbance readings at 595 nm. Metal-catalyzed Oxidation System In order to confirm the antioxidant function of DPPrx1, an MCO (metal-catalyzed oxidation) system was employed (Latifi et al. 2007). The following protocol was adapted from Lim et al. (1993). 10 mM of Dithiothreitol (DTT) and 40 μM of FeCl3 were mixed to produce free oxidant radicals (i.e. hydroxyl radical and hydroperoxyl radical species). In addition, 20 mM of EDTA was added with varying amounts of DPPrx1 (0–8 μM). The final volume for each reaction was 50 μL. Differing volumes of MilliQ-grade water were added to each reaction in order to bring the total volume of each reaction to 60 μL. These reactions were then incubated at 37°C for one hour in a shaking incubator (300 rpm). Plasmid DNA (500ng pUC19) was then added to the reactions and incubated for 4 hours at 20°C The impact of the MCO on the structure of the DNA was then analyzed by loading 20 μL of each reaction on to a 1% agarose gel which was then electrophoresed at 100 volts for 30 minutes and 50 volts for one hour (Zhang and Lu 2015). Horseradish peroxidase competition assay for DPPrx1 71 A 0.5 mM HRP stock solution was prepared by dissolving HRP in 25 mM standard Prx buffer, pH 7.0 (25 mM potassium phosphate and 1 mM, EDTA [pH 7.0])(Nelson and Parsonage 2011). The final concentration of HRP working stock was determined by reading the absorbance using a SmartSpec™ Plus Spectrophotometer FilterMax F5 Multi-Mode Microplate Reader (Bio-Rad) at 403 nm. As well, a cuvette with a 1 cm pathlength and a molar extinction coefficient of 1.02 x 105 M-1cm-1 were used. Aliquots were then stored at -80°C. To reduce purified DPPrx1 (active conformation), it was mixed with 10mM of dithiothreitol (DTT) and incubated at room temperature for 1 hour. From the reduced peroxiredoxin, five different peroxiredoxin working stocks were prepared with concentrations of 2 μg/ml, 4 μg/mL, 6 μg/ml, 8 μg/mL, and 12 μg/mL. 2.25 μL of HRP stock solution was added to each well of the plate and DPPrx1 and MilliQ-grade water were then added to achieve the desired final DPPrx1 concentration and a total reaction volume of 150 μL. The reactions for each concentration was run in triplicate. To initiate the reaction between HRP and H2O2 , 10 μL of 45 μM H2O2 was added to each well of the plate. Due to the instability of compound 1, the absorbance at 403 nm was then measured within 30 seconds of adding H2O2 in order to determine the extent to which HRP compound 1 had formed. The mixture was pipetted up and down and care was taken to prevent the formation of bubbles and the absorbance at 403 nm was recorded (Nelson and Parsonage 2011). DPPrx1 rate constant calculation The proportion of H2O2 reacting with horseradish peroxidase was determined from the initial observed absorbance (ΔAobvs) reading at 403 nm. The concentration of 72 DPPrx1 is proportional to the difference between that observed absorbance (ΔAobvs) of each sample and the maximum change in absorbance (ΔAmax) at 403 nm (ΔAmax -ΔAobvs) after adding DPPrx1. The ΔAmax can be calculated from averaging all the absorbance readings from the wells containing 0 mg/mL of peroxiredoxin (i.e. no inhibition of the HRP H2O2 reaction). Fractional HRP inhibition can then be calculated by the ratio of H2O2 reacting with horseradish peroxidase to H2O2 reacting with peroxiredoxin: ([ΔAmaxΔAobvs]/ ΔAobvs). Therefore, based on this formula and the rate equations for peroxiredoxin and HRP’s reactions with H2O2 we can generate equation 1. [ΔAmax − ΔAobvs] [(Px3 Px3)] = 012 4567 ) [(HRP HRP)] ΔAobvs This equation was reorganized to solve for kPx3, allowing for a k-value to be calculated for each concentration of DPPrx1. 9 ΔAmax − ΔAobvs : [(HRP HRP)] = [(Px3 Px3)] 012 4567 ) ΔAobvs The kPx3 values were plotted against the DPPrx1 concentrations and a linear regression was performed. The generated slope is equivalent to the second order rate constant for DPPrx1. Based on values documented by Nelson et al. (2008), the kHRP used in these calculations was 1.8 × 107 M-1 s-.1 pKa analysis of DPPrx1 In order to determine the functional pKa of DPPrx1, the HRP competition assay was conducted at seven different pH values (4.5, 5, 5.5, 6, 6.5, 7, 7.5) with each concentration of DPPrx1 again being run in triplicate. The previously described DPPrx1 expression protocol was conducted again with purified DPPrx1 eluted and stored in elution buffers at each of the desired pH values. To achieve the desired pH values for the 73 elution buffer, pH was measured using a Cyberscan pH 1100/2100 pH calibrator (Eutech Instruments) and adjusted using 1M hydrochloric acid (pH 2.1). The only difference in the protocol for the horseradish competition assay used for pKa determination was that 1.8 × 107 M-1 s-1 was only used as the kHRP value for pH values of 5 and higher. For pH values less than 5, the k HRP was calculated using equation 3. KHRP = 1.78 x 107 M-1 s-1 012 4567 3) [H+] [H+] + K +K K 2 1 2 The K-values used for these calculations in Equation 3, as recommended by Nelson et al. (2008), were K1 = 5.64 X 10-4 and K2 = 1.29 x 10-4. The resulting second order rate constants were then plotted against their associated pH values and fitted directly to Equation 4 using CurveExpert Professional Software v.2.6.5 (Hyams 2018). Y= [ A x 10 pH)+ B x 10 pKa)] 10pKa + 10pH 012 4567 ) In Equation 4, A indicates the upper plateau from the plotted data associated with the higher pH values, and B is the lower plateau from the plotted data associated with the lower pH values. Y corresponds to the ε 240 and was calculated from the amount of protein in solution as described using the Bradford assay (Nelson et al. 2008). The HRP competition assay was conducted three times at each of the pH values being investigated, with the exception of pH 4.5 (run once), pH 5 (run in duplicate), and pH 7.5 (run in duplicate) where it was difficult to generate enough protein yield to run the assays in triplicate. DPPrx1 Hyperoxidation analyses To determine whether the conformation of reduced DPPrx1 changed when exposed to higher concentrations of H2O2, a SDS-PAGE was performed under 74 nonreducing conditions. First DPPrx1 was reduced with 10mM of dithiothreitol (DTT) and incubated at room temperature for 1 hour. Nine 5 μM aliquots of DPPrx1 were prepared and each was mixed with a different concentration of H2O2 ranging from 0–0.96 mM. These reactions were then mixed with 20-40 μL of SDS loading buffer (Laemmeli’s solution) and boiled for ten minutes and then run on 4-12% bis/trispolyacrylamide gels for 1.5 hours at 125 V (Peskin et al. 2013). Xanthine Oxidase SOD Assay In order to functionally characterize DPSOD1, a water-soluble tetrazolium (WST-1) salt competition assay with xanthine oxidase and tetrazolium was conducted using the WST-1 SOD assay kit (Sigma-Aldrich), scaled for microplate readers. Bovine SOD standards were prepared using superoxide dismutase from bovine erythrocytes () in the following concentrations: 200 U/mL, 100 U/ml, 50 U/mL, 20 U/mL, 10 U/mL, 5 U/mL, 1 U/mL, 0.1 U/mL, 0.05 U/mL, 0.01 U/mL, and 0.001 U/mL. An enzyme working solution containing xanthine oxidase, was prepared by diluting 15 µL of enzyme stock solution with 2.5 mL of dilution buffer. Both solutions were included with the Sigma-Aldrich kit. WST working solution was also prepared by diluting 1 mL working stock solution in 19 mL of dilution buffer. Both the enzyme working solution and working solution were stored for long-term use (up to one month) at -20°C and 4°C respectively. The DPSOD1 assay and the subsequent inhibition rate (%) calculations were conducted using the protocol provided with the WST-1 plate SOD assay kit (SigmaAldrich). The components of the sample, standard and blank solutions were made as described in Table 2 using a multichannel pipette and all were run in triplicate. 75 Absorbance readings at 560 nm were obtained using a FilterMax F5 Multi-Mode Microplate Reader. Table 3.2 - The proportion and type of reagent in each SOD sample and blank in the Sigma-Aldrich WST-1 plate assay Sample solution ddH2O WST working solution Enzyme working solution Dilution buffer Sample 20 µL – 200 µL 20 µL – Blank 1 – 20 µL 200 µL 20 µL – Blank 2 20 µL – 200 µL – 20 µL Blank 3 – 20 µL 200 µL – 20 µL The assay was performed with 1ng, 2ng, 4 ng, 6 ng, and 8 ng of DPSOD1. Each concentration was run in triplicate, and the mean percent inhibition for each DPSOD1 concentration was determined. These mean percent inhibition values were plotted against the mass of SOD used, and an exponential curve was fitted to the data using CurveExpert Professional Software (Equation 5). Y = a – e-bx) (Equation 5) From these plots the amount of DPSOD1 and bovine SOD that produced 50% inhibition of xanthine oxidase (IC50) or one SOD unit was determined. The activity of DPSOD1 can then be expressed in terms of its IC50 value or in terms of bovine SOD equivalents where the final value is expressed as a ratio of the DPSOD1 IC50 value to the Bovine SOD IC50 value (Peskin and Winterbourn 2017). 3.6 Results DNA cleavage by the metal-catalyzed oxidation system (MCO) A MCO system was used to establish the antioxidant-based ability of DPRX1 to protect DNA from cleavage, in an investigation of the likely physiological role of this 76 protein in the cellular context. The mixture of DTT and Fe3+ produces DNA damaging free radicals, which degrade the supercoiled form (SF) of DNA, producing the more linear nicked form (NF) (Figure 3.5). The SF form travels further down the gel compared to the NF form. If DPPrx3 has antioxidant activity, a brighter SF band should be observed in the mixtures with a higher concentration of DPPrx3 compared to those with a lower concentration (Zhang and Lu 2015). In the wells with 0 μg/mL of peroxiredoxin, there was no distinct lower band visible, only a bright upper band. However, as the concentration of DPPrx1 increased the intensity of the upper NF band decreased and the intensity of the lower SF band increased as a result of more of the SF form remaining intact (Figure 3.5). NF! SF! !!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!0!!!!!!!!!!!!!25!!!!!!!!!!!!50!!!!!!!!!!!75!!!!!!!!!!100!!!!!!!!200!!!!!!!!!!!–!!!!!!!!!!!!!!!–!!!!!!!!!!!!!–!!!!!!!!!!!!!!!–!!!!!!!!!!!DPPrx1(μg/mL)! !!!!!!!!!!!!!!!!!!!!!!!!!!!+!!!!!!!!!!!!!!+!!!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!!!+!!!!!!!!!!!DTT! !!!!!!!!!!!!!!!!!!!!!!!!!!!+!!!!!!!!!!!!!!+!!!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!–!!!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!!!–!!!!!!!!!!!FeCl3! !!!!!!!!!!!!!!!!!!!!!!!!!!!+!!!!!!!!!!!!!!+!!!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!+!!!!!!!!!!!!!!!+!!!!!!!!!!!!!–!!!!!!!!!!!!!!!+!!!!!!!!!!!pUC19! Figure 3.5 - Antioxidant activity of DPPrx1. DTT and Fe3+ were mixed to generate free radicals and incubated for 1 hour for 37°C. Afterwards pUC19 plasmid DNA (500 ng) and DPPrx1 in concentrations ranging 0-200 µg/mL were added to the reaction and incubated for another 4 hours at 37°C. The resulting reaction mixes were run (100 volts for 0.5 hours, followed by 50 volts for 1 hour) on a 1.0% agarose gel stained with ethidium bromide. NF and SF indicate the nicked and supercoiled plasmid DNA respectively. Second order rate constant for DPPrx1 To further characterize the enzymatic activity of DPPrx1, a competition assay assessing the ability of DPPrx1 to outcompete horseradish peroxidase for H2O2 was employed. HRP has its maximum absorbance at a wavelength of 403 nm. Therefore, the 77 degree to which HRP compound 1 production was inhibited was first determined by measuring the absorbance at 403nm with increasing concentrations of DPPrx1. The difference between these values and the end point absorbance at 403 nm for reactions containing only HRP and H2O2 was then determined (Figure 3.6; Figure 3.7). Decrease in Absorbance at 403 nm ±1 SE 0.12 0.1 0.08 0.06 0.04 0.02 0 0 2 4 DPPrx1 [μM] 6 8 12 Figure 3.6 - Formation of HRP compound 1 in varying concentrations of DPPrx1, as determined by the difference obtained by subtracting absorbance at 403nm of H2O2 reacting with just HRP (∆AObs) and that of H2O2 reacting with HRP in the presence of varying concentrations of DPPrx1 (∆Amax-∆AObs). Error bars indicate standard error for samples that were run in triplicate with protein from three separate DPPrx1 inductions. 78 1600 {[F/(1-F)]}kHRP(HRP)s-1 ±1 SE 1400 y = 98.479x + 66.083 R² = 0.9705 1200 1000 800 600 400 200 0 0 2 4 6 DPPrx1 [μM] 8 10 12 14 Figure 3.7 - Determination of second-order rate constant (9.8 x 107 M-1s-1) for the reaction of reduced DPPrx1 with H2O2 through competition with horseradish peroxidase. The second order rate constants were calculated using Equation 1 and were plotted against their associated DPPrx1 concentration and fit by linear regression. Samples were run in triplicate with protein from three separate DPPrx1 inductions, and the calculated second order rate constants were averaged. The error bars represent standard error. As the concentration of DPPrx1 increased, the absorbance at 403 nm decreased (Figure 3.6). The difference in absorbance before and after DPPrx1 was added and was used to calculate the fractional inhibition of each reaction mixture, and a linear regression was performed on the plot of fractional inhibition versus DPPrx1 concentration. The resulting slope provided a second order rate constant of 9.8 x 107 M-1s-1 for DPPrx1 (Figure 3.7). pKa analysis of DPPrx1 To assess the pH dependence of DPPrx1 with H2O2, the HRP competition assay was again employed but across a pH range. The resulting second order rate constants were plotted against their corresponding pH values (Figure 3.8). 79 Second order rate constants for DPPrx1 120 100 80 60 40 20 0 -20 4.3 4.8 5.3 5.8 pH 6.3 6.8 7.3 Figure 3.8. - Determination of pKa (5.4213) for reduced DPPrx1 with H2O2. Second order rate constants for the reaction of H2O2 with DPPrx1 were plotted against the pH in which they were generated; data was fit to Equation 2 as described in the methods. The number of points at each pH is the number of replicates at that particular pH. As the pH for the assay increased, the resulting k values also increased. After a pH of 6 the rate constants leveled out and were relatively similar. Overall, a functional pKa value of 5.4 was determined from fitting the data to Equation 2 (Figure 3.8). Hyperoxidation analyses for DPPrx1 In order to determine the structure of DPPrx1 under differing oxidative stress conditions, reduced DPPrx1 was mixed with varying concentrations of H2O2 and the resulting band patterns were compared via SDS-PAGE (Figure 3.9). At H2O2 concentrations of 0.02 mM and 0.04 mM, three bands were generated: a major band at ~40 kDa, a smaller band just underneath it, and another minor band at ~22 kDa. At H2O2 concentration of 0 mM and above 0.06 mM, only two bands formed, a major band at ~40 kDa and a minor one at ~22 kDa (Figure 3.9). 80 Figure 3.9. Nonreducing SDS-PAGE analysis of purified, reduced DPPrx1. 5 µM of DPPrx1 was mixed with varying concentration of H2O2 (0–0.96 mM). 1 kb ladder is in lane 1. WST-1 assay on DPSOD1 In order to assess the relative activity of DPSOD1, an inhibition curve was generated form varying DPSOD1 concentrations using the WST-1 SOD assay (Figure 3.10). From the inhibition curves, the amount of DPSOD1 that results in 50% inhibition of WST-1 (IC50) reduction was determined to be 0.95 ng (Figure 3.10). This value was less than the amount of Bovine SOD, which was 1.62 ng. As the activity of bovine SOD is often used a common reference point when characterizing SOD activity, the activity of DPSOD1 can be expressed as a ratio of its IC50 to that of the bovine standard – 1.62 ng per 0.95 ng of protein. 81 Inhibition of WST-1 reduction (%) ±1 SE 120.00 100.00 80.00 60.00 40.00 20.00 0.00 -0.50 -20.00 1.50 3.50 5.50 7.50 9.50 DPSOD1 (ng) 100.00 80.00 60.00 ±1 SE Inhibition of WST-1 reduction (%) 120.00 40.00 20.00 0.00 -0.50 0.50 1.50 -20.00 2.50 3.50 4.50 5.50 6.50 7.50 Bovine SOD (ng) Figure 3.10 - (a) Inhibition percentages of WST-1 reduction for differing concentrations of DPSOD1, fitted to an exponential curve to generate an inhibition curve. (b) Inhibition percentages of WST-1 reduction for differing concentrations of bovine Cu, Zn-SOD fitted to an exponential curve to generate an inhibition curve. Each point is the mean value for the three wells containing that specific standard, and the error bars represent standard error. 3.7 Discussion MCO DNA protection assay One of the key physiological roles of peroxiredoxin is to mitigate the damaging effects of H2O2, as it is particularly damaging to proteins, lipids, and most notably DNA 82 (Mason 2016; Dmochowska-Ślęzak et al. 2016). As demonstrated in the MCO system, when DPPrx1 was present in higher concentrations, the accumulation of nicked DNA was minimized, supporting that DPPrx1 possesses antioxidant activity (Figure 3.5). This is consistent with the previously mentioned work done by Li et al. (2016) looking at a typical 2-cysteine peroxiredoxin from the pine wood nematode, which lives in a similar host environment. The results from this MCO experiment support the important antioxidant role of DPPrx1 for the mountain pine beetle. Preventing this kind of molecular damage would be essential to the success of the mountain pine beetle when first entering the phloem and surviving the flood of defensive compounds produced by the tree that can initiate the generation of free radicals (Mason 2016). Peroxiredoxin functional assay The second order rate constant generated for DPPrx1 with H2O2 was 9.8 x 107 Ms (Figure 3.6). This value is two orders of magnitude higher than the rate constant 1 -1 reported for human typical 2-cysteine peroxiredoxin, 4.9 x 105 M-1s-1 (Carvalho et al. 2017). However, it is more comparable to another human peroxiredoxin, 4 x 107 M-1s-1 (Ogusucu et al. 2007), and to those found in Salmonella typhimurium, AhpC - 2.3 x 107 M-1s-1 (Parsonage et al. 2008), and S. cerevisae, Tsa1 - 2.2 × 107 M−1 s −1 and Tsa2 - 1.3 × 107 M−1 s −1 (Nelson et al. 2008). However, of these five peroxiredoxins, DPPrx1 has the highest second order rate constant, indicating that DPPrx1 is more reactive with H 2O2 than these other peroxiredoxins. The reactivity of the protein likely plays a key part in mountain pine beetle survival. This high reactivity would better enable the protein to protect the cell against oxidative damage, but also more readily participate in any hydrogen peroxide-mediated cellular signaling. 83 The horseradish competition assay is a useful tool to elucidate the reactivity of DPPrx1 with H2O2 but cannot be used to study the relationship between DPPrx1 and organic peroxides. There are several proposed signal transduction pathways that use peroxiredoxin as the initial messenger that convey the message through both covalent and noncovalent interactions (Ogusucu et al. 2007). Therefore, DPPrx1 is likely involved in other pathways, but future research would be required to determine the particular pathways, protein interaction partners, and DPPrx1’s reactivity with the associated target molecules. One such study would involve a variant of the horseradish peroxidase competition assay that could be used to assess the role of DPPrx1 in other pathways as described by Peschenko and Shichi (2001). In this approach, another competition assay called the selenium glutathione peroxidase/glutathione reductase/NADPH coupled assay is incorporated into the protocol. Initially, peroxynitrite and organic peroxides are mixed with the peroxiredoxin of interest. If the peroxiredoxin has a high affinity for the organic peroxides, then there should be a reduction in free organic peroxides in the mixture. Those free organic peroxides can be by quantified with the glutathione peroxidase/glutathione reductase/NADPH coupled assay. The difference in organic peroxides can then be used to generate a second order rate constant between peroxiredoxin and the organic peroxides (Ogusucu et al. 2007; Peshenko and Schichi 2001). Potentially, this approach could be employed with DPPrx1 to better assess the role of this protein in other detoxification pathways used by the mountain pine beetle, with more ecologically relevant substrate contexts. Peroxiredoxin pKa analysis 84 DPPrx1 has a functional pKa of 5.4, which is consistent with the pKa value for Tsa1 from S. cerevisae, which also has a pKa of 5.4. This pKa is higher than that of other proteins that have thiolate containing active sites, such as E. coli DsbA at 3.5. However, the pK a for DPPx1 is lower than that of free cysteine (6.7–7.1). A pKa value lower than that of free cysteine is important because at a pKa of 6 there would be 93% deprotonation of the cysteine at a pH of 7, but at a pH of 8.5 only 3% of the cysteine would show deprotonation (Nelson et al. 2008). As the deprotonated thiolate form is much more reactive with H2O2, having a pKa below 6 highly increases the reactivity of the peroxiredoxin protein. However, the impact this would have on the overall function of the protein is somewhat balanced by the fact that once the thiolate is formed, the nucleophilicity diminishes as the pK a value decreases (Nelson et al. 2008). Physiologically this becomes relevant, as DPPrx1 is a cytosolic protein. The cytoplasm of the cells is approximately neutral, so the combination of the lower pKa value and the decreased nucleophilicity after thiolate formation, allows for it to be able to carry out antioxidant activities in the cytosol (Nelson et al. 2008). Hyperoxidation of DPPrx1 In the absence of H2O2, DPPrx1 ran as a major and minor band, likely corresponding to the ~22 kDa monomer and the ~43kDa dimer, respectively. Initially as the H 2O2 concentration increased, the homodimer (both the single disulfide and the double disulfide forms) became the major bands, as this is the functionally relevant conformation of the protein. At 40 µM, the monomer once again becomes the dominant band (Figure 3.9), corresponding to the hyperoxidized form (Figure 3.1) (Peskin et al. 2013). 85 The hyperoxidation of typical 2-cysteine peroxiredoxins in insects has not been characterized. It was observed that DPPrx1 underwent hyperoxidation at a lower concentration of H2O2 (~0.06 mM) than human peroxiredoxin (~0.12 mM). This can potentially be attributed to the higher rate constant of DPPrx1 (Figure 3.7) and the presence of hyperoxidation sensitivity motifs (Figure 3.2). This hyperoxidized peroxiredoxin has been found to serve many signalling purposes in other organisms suggesting that hyperoxidized DPPrx1 could play a similar role in the mountain pine beetle during host colonization (Delaunay 2002). This signal transduction could be essential during host colonization when mountain beetles are undergoing considerable oxidative stress which would require more proteins to mitigate the damaging impact of free radicals like catalase and glutathione peroxidase which could more effectively breakdown H2O2. α-pinene is a terpene produced by pine trees and is used as a defensive toxin against mountain pine beetles during colonization (Seybold et al. 2006). However, the mountain pine beetle has evolved a mechanism to incorporate the α-pinene molecule into the structure of its primary aggregation pheromone, trans-verbenol (Chiu et al. 2019). Therefore, a compound that was originally produced by trees to withstand mountain pine beetle attacks has been coopted by the mountain pine beetle to serve as a key component for overwhelming the tree defenses. Potentially, a similar situation could be occurring with H2O2, where a compound originally produced to defend against mountain pine beetle attack has been incorporated by the beetle’s defense systems as a means for better overcoming the host tree’s defenses. 86 H2O2 is produced in high quantities during the breakdown of phenols and terpenes that occurs during tree colonization and can freely cross the plasma membrane of cells. The results from the hyperoxidation experiments support that when exposed to high concentration of H2O2, DPPrx1 undergoes a hyperoxidized change in oligomeric state. As discussed by Delaunay (2002), hyperoxidized peroxiredoxin acts as a signaling molecule in other organisms. One of the most likely signaling targets for hyperoxidized DPPrx1 would be proteins that lead to the expression of oxidative stress proteins like glutathione peroxidase and catalase that typically have a higher affinity for H2O2. That said, it is likely that DPPrx1 is involved in many different signaling pathways because there are 196 unique protein interactors for human peroxiredoxin 1 (Ledgerwood et al. 2017) Ledgerwood et al. (2017) have proposed three different signal transduction pathways caused by hyperoxidized peroxiredoxin. The first proposed mechanism is known as the ‘disulfide relay’, where hyperoxidized peroxiredoxin triggers disulfide formation in a targeted protein or proteins, which then proceed to pass on this signal through other proteins (Okazaki et al. 2007; Ichijo et al. 1997). This signal transduction was original studied in S. cerevisiae but there is evidence that supports its presence in mammalian cells as well. As reviewed in Ledgerwood et al. (2017), this mechanism is carried out between thiol containing proteins. The second mechanism that has been proposed for peroxiredoxin signaling is noncovalent interactions of the hyperoxidized peroxiredoxin with its target proteins. These mechanisms would allow the peroxiredoxin to pass the signal on to proteins without thiol groups. The exact mechanism in which this signal is passed on is not fully known, but it appears peroxiredoxins are able to impact non-redox proteins like phosphatase, a tensin 87 homologue (PTEN), and epidermal growth factor (EGF) without changing their redox states (Kwon et al. 2004; Lee et al. 2002). These proteins have been referred to as peroxiredoxin interacting partners and identifying them within the mountain pine beetle is key to elucidating the exact signaling roles of DPPrx1 (Ledgerwood et al. 2017). The third type of peroxiredoxin signal transduction is where the hyperoxidized peroxiredoxin is not actually acting as the signal, but because it is catalytically inactive, the subsequent build-up of H 2O2 is acting as the initial signal. This mechanism is referred to as the floodgate model and has been proposed as playing a key part during the cell cycle (Wood et al. 2003). In this specific scenario, it is not the build-up of H 2O2 that causes the inactivation of the peroxiredoxin but rather inhibitory phosphorylation by another protein (Ledgerwood et al. 2017; Wood et al. 2003). This allows for the build-up of H2O2 and the oxidation and inhibition of a centrosome-bound phosphatase. This would prevent the early degradation of mitotic activators and allows for the cell cycle to continue (Lim et al. 2015). However, the exact mechanism for how H2O2 channels and accumulates at the centrosome is still not understood (Woo et al. 2010). Each of the three previously mentioned pathways is dependent on involvement of a hyperoxidized peroxiredoxin. Therefore, the fact that DPPrx1 shows sensitivity to hyperoxidation (Figure 3.9) indicates that it could potentially be a part of many cellsignaling systems. Superoxide dismutase functional assay Upon conducting a WST-1 assay on DPSOD1, an IC50 value of 0.95 ng was generated. This is lower than that of the SOD obtained from bovine erythrocytes -1.62 ng (Figure 3.10). In addition, this IC50 is much lower than that of a Cu, Zn SOD from S 88 cerivisiae, that has an IC50 value of 1.5 μg (Peskin and Winterbourn 2017). Therefore, this suggests that DPSOD1 may be a relatively more reactive SOD. This is similar to the higher rate constant generated for DPPrx1 where the increased reactivity could be attributed to the unique life history of the mountain pine beetle, immersing itself in toxic compounds during host tree colonization. Other insects that are found in environments with a high concentration of free radicals, such as Tribolium castaneum and Reticulitermes speratus, also appear to have more reactive SOD (Tasaki et al. 2018; Ferro et al. 2017). Therefore, it is likely that the life history of mountain pine beetle requires a highly reactive suite of oxidative stress proteins to successfully survive the host tree’s defenses 3.8 Conclusions and Future Directions The findings of this study support the likely physiological role of peroxiredoxin within the mountain pine beetle, as a key part in preventing biomacromolecule damage, such as DNA nicking, due to the presence of reactive oxygen species. As well, the functional profile of DPPrx1 was further expanded through characterizing the reaction constant and pKa value for DPPrx1 with a horseradish peroxidase competition assay. A potential aspect of DPPrx1’s functionality that was not tested for in this study is its ability to mitigate H 2O2 levels independent of its functional cysteines. This catalytic activity is similar to that of catalase and associated with a GVN motif located in the active site (Sun et al. 2017). This particular motif is present in DPPrx1 making this protein a suitable candidate for further tests determining whether DPPrx1 also has this catalytic function (Sun et al. 2017). 89 Furthermore, my findings support that DPPrx1 is sensitive to hyperoxidation; this susceptibility to hyperoxidation makes sense with the relatively high second order rate constant that we determined for the enzyme. This potential role in signal transduction could be essential when the mountain beetle cells are undergoing considerable oxidative stress, as it could trigger the expression of other oxidative stress proteins as discussed previously. However, there are many other signaling pathways hyperoxidized peroxiredoxin is believed to be a part of (i.e. the H2O2 floodgate model, noncovalent signaling interactions, and the disulfide relay) (Ledgerwood et al. 2017). Therefore, it is likely that cellular signaling is the primary role of DPPrx1 during host colonization, even more so than its role as an antioxidant. At this point, the signaling roles of peroxiredoxin in the mountain pine beetle are mainly speculative, making it an area for future research. However, characterizing some of the other pathways and molecules that may be interacting with peroxiredoxin may be challenging as currently most approaches have only been conducted with high throughput mass spectrometric-based methods and overexpression of tagged peroxiredoxins (Portillo-Ledesma 2018). The former approach is highly technical and the second is problematic as the tag often impacts the activity of the peroxiredoxin. Conducting a study of this type could require designing a novel experimental approach. In addition to the different pathways that hyperoxidized DPPrx1 may be involved in, the unique active site of this enzyme is another potential focal point for future investigation. As discussed in the introduction, the glycine residues that are present in the motif A’s of human and bacterial Prx1s are not present in DPPrx1 (Bolduc et al. 2018). There is a potential link between the presence of these lysines and the high reactivity of 90 DPPrx1, particularly when considering why DPPrx1 shows such a high sensitivity to hyperoxidation, despite missing the YR motif typically associated with 2-cysteine peroxiredoxin being prone to hyperoxidation (Figure 3.2; Figure 3.9). An interesting line of investigation would be to compare the DPSOD1 IC50 value to some of the other mountain pine beetle SODs. Most notably DPSOD3, which is an extracellular protein, meaning it is secreted into the periplasmic space. Other studies support that extracellular SODs are often secreted to break down superoxide before it can enter the cell and cause any damage either on its own or from its breakdown and subsequent formation of H 2O2. In addition, because there is evidence that extracellular SOD may also be involved in redox signaling pathways (Fukai & Fukai 2011), it would be interesting to see how the difference in cellular localization and function might impact the overall activity of these two SOD proteins. 91 Chapter 4: Conclusions Mitigating the destructive impacts of reactive oxygen species (ROS) is essential to the mountain pine beetle’s success during host colonization. A better understanding of these oxidative stress proteins and their unique functionality will lead to an overall increase in the understanding of bark beetles and the molecular mechanisms underlying host detoxification. Therefore, findings from this study could potentially be applied to other bark beetles like, Dendroctonus rufipennis, the spruce beetle or other economically important Dendroctonus spp. Spruce beetle currently of growing concern in British Columbia as its population is increasing in some areas, considerably impacting the local forest ecosystems and economies (Campbell et al. 2019) and various other bark beetles are also important in BC and elsewhere in the mountain west. In this study I selected five key oxidative stress proteins from the mountain pine beetle – superoxide dismutase (SOD), catalase, glutathione peroxidase (GPx), and peroxiredoxin (Prx) – for functional and phylogenetic analyses. Including the sequence and phylogenetic data in this study enabled me to compare the active sites of these proteins to those of other insect species, thus providing a molecular context for the functional data. From the generated sequence alignments, I found a variation in the YF hyperoxidation motif of DPPRx1 (Chapter 2). A non-conserved amino acid substitution in such a key reactivity motif could have a considerable impact on that protein’s reactivity. In fact, this impact on reactivity was potentially observed in this study as the generated second-order rate constant was 9.8 x 107 M-1s-1, which was two orders of magnitude higher than the human homolog (Ogusucu et al. 2007). As well, DPPrx1 92 displayed a higher susceptibility to hyperoxidation compared to human Prx proteins. This is notable as hyperoxidized Prx is known to act as a signal for the expression of other oxidative stress proteins (Delaunay 2002). This increase in activity and susceptibility to hyperoxidation indicates that DPPrx1 could be a key component of the mountain pine beetle primary antioxidant system. Future studies could investigate DPPrx1 homologs in other bark beetles, such as spruce beetle or Ips spp., to see if they have the same unique FF motif and function similarly. Additionally, one could do site-specific mutagenesis, changing the motif to YF, and then conduct comparative kinetic assays. These DPPrx1 features could be unique to bark beetles and could have evolved to enable them to better overcome the host tree’s defenses. If this is the case, disrupting these proteins and the molecular pathways that they are part of could be a potential target for future pest management strategies. To gather this functional data, I obtained the coding region for each protein from a mountain pine beetle cDNA library and inserted into an expression vector using the Gateway Cloning ProcedureTM. This approach generated enough DPPrx1 and DPSOD1 to proceed with functional analyses. Competition assays were used to functionally characterize these proteins, a xanthine oxidase assay for DPSOD1 and a horseradish peroxidase for DPPrx1. Competition-based assays do not directly measure the activity of a protein but do so indirectly through the disruption of a competing protein’s reaction, and as a result, a reaction constant can be successfully generated. In order to acquire a sufficient protein yield for enzymological characterization, a systematic stepwise approach had to be employed. This was effective, but very time consuming. Therefore, these finished protocols could be very useful to other researchers 93 undertaking functional analyses on similar proteins. At the very least, they would provide a general framework. These protocols could be used in functional analyses related to a variety of research fields (agriculture, medicine, ecology, etc.). Despite generating purified recombinant enzyme for DPPRx1 and DPSOD1 characterization, I was unable to do so for the other oxidative stress proteins being analyzed. Most of these were expressed as insoluble inclusion bodies. There are a number of solutions that future researchers could employ to remedy this problem. Wingfield (2015) discussed denaturing misfolded proteins and refolding them under the conditions optimal for that specific protein. However, the yield is usually reduced (5%-20%) as a result of this procedure. Another approach would be to use an expression system other than E. coli. As these are eukaryotic proteins, potentially a eukaryotic expression system could be preferable. Although complex molecular processes vary considerably amongst eukaryotic organisms, protein modifications like glycosylation, phosphorylation, and fatty acid addition can be better facilitated in eukaryotic systems than prokaryotic ones (Possee 1997; Giesse et al. 1996). Cereghino and Clegg (1999) found success with yeast when trying to express proteins that had been clustering into inclusion bodies. As mentioned in the introductory chapter of this thesis, insect cell lines are also an option and would best replicate the cellular conditions of the mountain pine beetle, and insect cell lines can also be quickly scaled up, increasing the total protein yield (Grennan et al. 1996). Another method to promote proper protein folding, particularly for any of the extracellular proteins (DPSOD2 and DPPrx6) and membrane bound proteins (DPPrx5), would be to fuse a signal sequence to the recombinant proteins. These signal sequences 94 are key for directing proteins to their respective organelles and cellular regions. Zhang et al. (2018) compared the effect of a number of signal peptides on a thermostable lipase, lipBJ10, from Pseudomonas fluorescens. Although there was variation based on the specific signal peptide that was being used, in general they found that attaching these signal peptides to the N-terminus increased protein solubility. That said, the success of this approach is highly dependent, on the particular proteins being expressed and their associated secretory pathways. Singh et al. (2013) compared the expression yields of a thioredoxin protein that had two different attached leader sequences to wild types, and found that in general, those with leader sequences demonstrated decreased thermostability and increased likelihood for protein aggregation. Another potential solution to this problem would be to express the associated chaperone with each protein. For instance, in chapter 2 I found that one of the proteins that was designated as ‘SOD-like’ was most likely a copper chaperone for SOD (CCS); a protein which functions to ensure the proper insertion of the copper cofactor into the fully folded SOD. I supplemented the DPSOD1 inductions with copper and obtained a sufficient yield of soluble DPSOD1 (Williams et al. 2016). However, the majority of the protein was still in the insoluble state, and all of the DPSOD2 was insoluble. Potentially, if the expression vector contained both the DPSOD gene of interest and the CCS protein then there would be a greater yield of properly folded protein. A study by Eiamphungporn et al. (2018) compared the co-expression of a Cu/Zn SOD with its corresponding CCS protein and found that not only was their an increase in the yield of soluble protein, but that there was also a three-fold increase in specific activity. Therefore, now that I have identified a potential CCS protein, co-expressing it with 95 mountain pine beetle SOD proteins may not only change the protein yield but also the functional characterization of the proteins. The mountain pine beetle is an organism with a unique life history, spending the majority of its existence in a host tree that defends itself with resin-based toxins. Therefore, continuing to examine the proteins it uses for detoxification, like the oxidative stress proteins, will provide further intriguing insights as to what molecular modifications had to evolve in order for it to not only survive, but flourish in these conditions. As these functional and bioinformatics protocols become more refined, the more in-depth our understanding of these systems becomes. This is noteworthy because the better we understand these symptoms, the better we will be able to predict future outbreak patterns and create targeted pest management strategies. 96 Works Cited Ajuwon, O. R., Marnewick, J., & Davids, L. (2015). Rooibos (Aspalathus linearis) and its Major Flavonoids — Potential Against Oxidative Stress-Induced Conditions. INTEC Open Science, 171–218. http://dx.doi.org/10.5772/61614 Almagro Armenteros, J. J., Tsirigos, K. D., Sønderby, C. K., Petersen, T. N., Winther, O., Brunak, S., von Heijne, G., & Nielsen, H. (2019). SignalP 5.0 improves signal peptide predictions using deep neural networks. Nature biotechnology, 37(4), 420–423. https://doi.org/10.1038/s41587-019-0036-z Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. (1990) Basic local alignment search tool. J. Mol. Biol. 215:403–410. PubMed. Amid, A., Ismail, N. A., Yusof, F., Salleh, H. M. (2011). Expression, purification, and characterization of a recombinant stem bromelain from Ananas comosus. Process Biochemistry, 46(12), 2232–2239. Asmail, N. (2008). Cloning and characterization of superoxide dismutase from the ectoparasite Sarcoptes scabiei. Uppsala University. Ashokan, S. K. V., Mundagunar, D. S., & Mundagunar, Y. D. (2010). Catalase: Phylogenetic Characterization to explore protein cluster. Journal of Research in Bioinformatics, 1(1), 001-008. Aukema, B., Carroll, A., Zhu, J., Raffa, K., Sickley, T. A., & Taylor, S. W. (2006). Landscape level analysis of mountain pine beetle in British Columbia, Canada: spatiotemporal development and spatial synchrony within the present outbreak. Ecography, 29, 427–441. Behrens, S., Peuß, R., Milutinović, B., Eggert, H., Esser, D., Rosenstiel, P., Schulenburg, H., Bornberg-Bauer, E., & Kurtz, J. (2014). Infection routes matter in population-specific responses of the red flour beetle to the entomopathogen Bacillus thuringiensis. BMC genomics, 15(1), 445. Birben, E., Sahiner, U. M., Sackesen, C., Erzurum, S., & Kalayci, O. (2012). Oxidative stress and antioxidant defense. The World Allergy Organization journal, 5(1), 9–19. https://doi.org/10.1097/WOX.0b013e3182439613 Bolduc, J. A., Nelson, K. J., Haynes, A. C., Lee, J., Reisz, J. A., Graff, A. H., Clodfelter, J. E., Parsonage, D., Poole, L. B., Furdui, C. M., & Lowther, W. T. (2018). Novel hyperoxidation resistance motifs in 2-Cys peroxiredoxins. The Journal of biological chemistry, 293(30), 11901–11912. https://doi.org/10.1074/jbc.RA117.001690 Bonnett, T. R., Robert J. A., Pitt C., Fraser J. D., Keeling C. I., Bohlmann J., & Huber D. P. (2012). Global and comparative proteomic profiling of overwintering and developing 97 mountain pine beetle, Dendroctonus ponderosae (Coleoptera: Curculionidae), larvae. Insect Biochem. Mol Biol., 42(12), 890–901. https://doi.org/10.1016/j.ibmb.2012.08.003 Boyd, S. D., Liu, L., Bulla, L., & Winkler, D. D. (2018). Quantifying the Interaction between Copper-Zinc Superoxide Dismutase (Sod1) and its Copper Chaperone (Ccs1). Journal of proteomics & bioinformatics, 11(4), 473. https://doi.org/10.4172/jpb.1000473 Broxton, C. N., & Culotta, V. C. (2016). SOD Enzymes and Microbial Pathogens: Surviving the Oxidative Storm of Infection. PLoS pathogens, 12(1), e1005295. https://doi.org/10.1371/journal.ppat.1005295 Campbell, E. M., Antos, J. A., vanAkker, L. (2019). Resilience of southern Yukon boreal forests to spruce beetle outbreaks. Forest Ecology and Management, 433, 52–63. https://doi.org/10.1016/j.foreco.2018.10.037 Carroll, A.L., Aukema, B.H., Raffa, K.F., Linton, D.A., Smith, G.D., Lindgren, B.S. (2006). Mountain pine beetle outbreak development: the endemic-incipient transition. Mountain Pine Beetle Initiative Project #1.03 Working Paper. Canada Forest Services, Pacific Forestry Centre, Victoria, BC, Canada. Carvalho, L., Truzzi, D. R., Fallani, T. S., Alves, S. V., Toledo, J. C., Jr, Augusto, O., Netto, L., & Meotti, F. C. (2017). Urate hydroperoxide oxidizes human peroxiredoxin 1 and peroxiredoxin 2. The Journal of biological chemistry, 292(21), 8705–8715. https://doi.org/10.1074/jbc.M116.767657 Cereghino, G. P., & Cregg, J. M. (1999). Applications of yeast in biotechnology: protein production and genetic analysis. Current opinion in biotechnology, 10(5), 422–427. https://doi.org/10.1016/s0958-1669(99)00004-x Chae, H. Z., Chung, S. J., & Rhee, S. G. (1994). Thioredoxin-dependent peroxide reductase from yeast. The Journal of biological chemistry, 269(44), 27670–27678. Chiu, C.C., Keeling, C. I. & Bohlmann, J. (2019) The cytochrome P450 CYP6DE1 catalyzes the conversion of α-pinene into the mountain pine beetle aggregation pheromone transverbenol. Sci. Rep. 9(1477). https://doi.org/10.1038/s41598-018-38047-8 Clark, E. L., Pitt, C., Carroll, A., Lindgren, B., & Huber, D. P. (2014). Comparison of lodgepole and jack pine resin chemistry: implications for range expansion by the mountain pine beetle, Dendroctonus ponderosae (Coleoptera: Curculionidae). PeerJ, 2. Colinet, D., Cazes, D., Belghazi, M., Gatti, J. L., & Poirié, M. (2011). Extracellular superoxide dismutase in insects: characterization, function, and interspecific variation in parasitoid wasp venom. The Journal of biological chemistry, 286(46), 40110–40121. https://doi.org/10.1074/jbc.M111.288845 98 Corona, M., & Robinson, G. E. (2006). Genes of the antioxidant system of the honey bee: annotation and phylogeny. Insect molecular biology, 15(5), 687–701. https://doi.org/10.1111/j.1365-2583.2006.00695.x Couto, N., Wood, J. L., & Barber, J. (2016). The role of glutathione reductase and related enzymes on cellular redox homoeostasis network. Free radical biology & medicine, 95, 27–42. Cox, A. G., Winterbourn, C. C., & Hampton, M. B. (2009). Mitochondrial peroxiredoxin involvement in antioxidant defence and redox signalling. The Biochemical journal, 425(2), 313–325. https://doi.org/10.1042/BJ20091541 Cullingham, C. I., Cooke, J. E., Dang, S., Davis, C. S., Cooke, B. J., & Coltman, D. W. (2011). Mountain pine beetle host-range expansion threatens the boreal forest. Molecular ecology, 20(10), 2157–2171. https://doi.org/10.1111/j.1365-294X.2011.05086.x Delaunay, A., Pflieger, D., Barrault, M. B., Vinh, J., & Toledano, M. B. (2002). A thiol peroxidase is an H2O2 receptor and redox-transducer in gene activation. Cell, 111(4), 471–481. https://doi.org/10.1016/s0092-8674(02)01048-6 Dias, F. A., Gandara, A. C., Perdomo, H. D., Gonçalves, R. S., Oliveira, C. R., Oliveira, R. L., Citelli, M., Polycarpo, C. R., Santesmasses, D., Mariotti, M., Guigó, R., Braz, G. R., Missirlis, F., & Oliveira, P. L. (2016). Identification of a selenium-dependent glutathione peroxidase in the blood-sucking insect Rhodnius prolixus. Insect biochemistry and molecular biology, 69, 105–114. https://doi.org/10.1016/j.ibmb.2015.08.007 DiGuistini, S., Wang, Y., Liao, N. Y., Taylor, G., Tanguay, P., Feau, N., Henrissat, B., Chan, S. K., Hesse-Orce, U., Alamouti, S. M., Tsui, C. K., Docking, R. T., Levasseur, A., Haridas, S., Robertson, G., Birol, I., Holt, R. A., Marra, M. A., Hamelin, R. C., Hirst, M., … Breuil, C. (2011). Genome and transcriptome analyses of the mountain pine beetle-fungal symbiont Grosmannia clavigera, a lodgepole pine pathogen. Proceedings of the National Academy of Sciences of the United States of America, 108(6), 2504–2509. https://doi.org/10.1073/pnas.1011289108 Djuika, C. F., Huerta-Cepas, J., Przyborski, J. M., Deil, S., Sanchez, C. P., Doerks, T., Bork, P., Lanzer, M., & Deponte, M. (2015). Prokaryotic ancestry and gene fusion of a dual localized peroxiredoxin in malaria parasites. Microbial cell (Graz, Austria), 2(1), 5–13. https://doi.org/10.15698/mic2015.01.182 Dmochowska-Ślęzak, K. , Dmitryjuk, M. , Zaobidna, E. , & Żółtowska, K. (2016). The antioxidant defense system of Varroa destructor mites facilitates the infestation of Apis mellifera. Journal of Apicultural Science, 60(1). https://doi.org/10.1515/jas-2016-0010 Dunford. H. B. (1999). Heme Peroxidases. Wiley-VCH, New York. 507 pp. 99 Eiamphungporn, W., Yainoy, S., & Prachayasittikul, V. (2016). Enhancement of Solubility and Specific Activity of a Cu/Zn Superoxide Dismutase by Co-expression with a Copper Chaperone in Escherichia coli. Iranian journal of biotechnology, 14(4), 243–249. https://doi.org/10.15171/ijb.1465 Faccoli, M., Favaro, R., Concheri, G., Squartini, A., & Battisti, A. (2016). Tree colonization by the Asian longhorn beetle, Anoplophora glabripennis (Coleoptera: Cerambycidae): effect of habitat and tree suitability. Insect science, 23(2), 288–296. https://doi.org/10.1111/1744-7917.12192 Ferro, K., Ferro, D., Corrà, F., Bakiu, R., Santovito, G., & Kurtz, J. (2017). Cu,Zn Superoxide Dismutase Genes in Tribolium castaneum: Evolution, Molecular Characterisation, and Gene Expression during Immune Priming. Frontiers in immunology, 8, 1811. https://doi.org/10.3389/fimmu.2017.01811 Fukai, T., & Ushio-Fukai, M. (2011). Superoxide dismutases: role in redox signaling, vascular function, and diseases. Antioxidants & redox signaling, 15(6), 1583–1606. https://doi.org/10.1089/ars.2011.3999 Giesse, S., Gram, H., Kleuser, B., & Kocher, H. P. (1996). Eukaryotic Expression Systems: A Comparison. Protein Expression and Purification, 8(3), 271–282. https://doi.org/10.1006/prep.1996.0101 Grennan Jones, F., Wolstenholme, A., Fowler, S., Smith, S., Ziemnicka, K., Bradbury, J., Furmaniak, J., & Rees Smith, B. (1996). High-level expression of recombinant immunoreactive thyroid peroxidase in the High Five insect cell line. Journal of molecular endocrinology, 17(2), 165–174. https://doi.org/10.1677/jme.0.0170165 Gretscher, R., Streicher, P., Strauß, A.S., Wielsch, N., Stock, M., Wang, D., Boland, W., Burse, A. (2016). A common theme in extracellular fluids of beetles: extracellular superoxide dismutases crucial for balancing ROS in response to microbial challenge. Scientific Reports, 6, 24082. https://doi.org/10.1038/srep24082 Guex, N., Peitsch, M. C., & Schwede, T. (2009). Automated comparative protein structure modeling with SWISS-MODEL and Swiss-PdbViewer: a historical perspective. Electrophoresis, 30(1), S162–S173. https://doi.org/10.1002/elps.200900140 Gui-Qin, S., Ze, Z., Kun-Lun, J., Kun, Z., Dong-Xu, A., Gang, W., Bao-Long, Z., & He-Nan, Y. (2014) Characterization and expression analysis of peroxiredoxin family genes from the silkworm Bombyx mori in response to phoxim and chlorpyrifos. Pesticide Biochemistry and Physiology, 114, 24–31. https://doi.org/10.1016/j.pestbp.2014.07.007 Haijema, B. J., Noback, M., Hesseling, A., Kooistra, J., Venema, G., & Meima, R. (1996). Replacement of the lysine residue in the consensus ATP-binding sequence of the AddA subunit of AddAB drastically affects chromosomal recombination in transformation and 100 transduction of Bacillus subtilis. Molecular microbiology, 21(5), 989–999. https://doi.org/10.1046/j.1365-2958.1996.601424.x Ichijo, H., Nishida, E., Irie, K., ten Dijke, P., Saitoh, M., Moriguchi, T., Takagi, M., Matsumoto, K., Miyazono, K. and Gotoh, Y. (1997). Induction of apoptosis by ASK1, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science, 275, 90–94. Horton, P., Park, K. J., Obayashi, T., Fujita, N., Harada, H., Adams-Collier, C. J., & Nakai, K. (2007). WoLF PSORT: protein localization predictor. Nucleic acids research, 35, W585– W587. https://doi.org/10.1093/nar/gkm259 Hyams, D. G., CurveExpert software, http://www.curveexpert.net, 2010. Inupakutika, M. A., Sengupta, S., Devireddy, A. R., Azad, R. K., & Mittler, R. (2016). The evolution of reactive oxygen species metabolism. Journal of experimental botany, 67(21), 5933–5943. https://doi.org/10.1093/jxb/erw382 Janes, J.K., Li, Y., Keeling, C.I., Yuen, M.M.S., Boone, C.K., Cooke, J.E.K., Bohlmann, J., Huber, D.P.W., Murray, B.W., Coltman, D.W., & Sperling, F.A.H. (2014). How the mountain pine beetle (Dendroctonus ponderosae) breached the Canadian Rocky Mountains. Molecular Biology and Evolution, 31, 1803–1815. Jiu, M., Li, J., Gao, X., Wang, L., Wang, X., & Liu, S. (2015). Identification and characterization of two phospholipid hydroperoxide glutathione peroxidase genes from the Mediterranean species of the whitefly Bemisia tabaci complex. Archives of insect biochemistry and physiology, 89 1, 54–67. Jomova, K., Baros, S. & Valko, M. (2012). Redox active metal-induced oxidative stress in biological systems. Transition Metal Chemistry, 37, 127–134. https://doi.org/10.1007/s11243-012-9583-6 Job, D., & Dunford, H. (1978). Horseradish peroxidase. XXVIII. Formation and reactivity of the alkaline form. Evidence for an enzyme–substrate complex in compound 1 formation. Canadian Journal of Chemistry, 56 (10) Keeling, C. I., Yuen, M. M., Liao, N. Y., Roderick Docking, T., Chan, S. K., Taylor, G. A., Palmquist, D. L., Jackman, S. D., Nguyen, A., Li, M., Henderson, H., Janes, J. K., Zhao, Y., Pandoh, P., Moore, R., Sperling, F. A., W Huber, D. P., Birol, I., Jones, S. J., & Bohlmann, J. (2013). Draft genome of the mountain pine beetle, Dendroctonus ponderosae Hopkins, a major forest pest. Genome Biology, 14(3). https://doi.org/10.1186/gb-2013-14-3-r27 Klotz, M. G., & Loewen, P. C. (2003). The molecular evolution of catalatic hydroperoxidases: evidence for multiple lateral transfer of genes between prokaryota and from bacteria into eukaryota. Molecular biology and evolution, 20(7), 1098–1112. https://doi.org/10.1093/molbev/msg129 101 Knoops, B., Loumaye, E., & Van Der Eecken, V. (2007). Evolution of the Peroxiredoxins. Peroxiredoxin Systems. Knoops, B., Goemaere, J., Van der Eecken, V., & Declercq, J. P. (2011). Peroxiredoxin 5: structure, mechanism, and function of the mammalian atypical 2-Cys peroxiredoxin. Antioxidants & redox signaling, 15(3), 817–829. https://doi.org/10.1089/ars.2010.3584 Kobayashi, Y., Nojima, Y., Sakamoto, T., Iwabuchi, K., Nakazato, T., Bono, H., Toyoda, A., Fujiyama, A., Kanost, M., & Tabunoki, H. (2019). Comparative analysis of seven types of superoxide dismutases for their ability to respond to oxidative stress in Bombyx mori. Scientific Reports, 9(1):2170. Kumagai, T., Osada, Y., Ohta, N., & Kanazawa, T. (2009). Peroxiredoxin-1 from Schistosoma japonicum functions as a scavenger against hydrogen peroxide but not nitric oxide. Molecular and biochemical parasitology, 164(1), 26–31. https://doi.org/10.1016/j.molbiopara.2008.11.002 Kwon J, Lee SR, Yang KS, Ahn Y, Kim YJ, Stadtman ER, Rhee SG. Reversible oxidation and inactivation of the tumor suppressor PTEN in cells stimulated with peptide growth factors. Proc Natl Acad Sci U S A. 2004 Nov 23;101(47):16419-24. doi: 10.1073/pnas.0407396101. Epub 2004 Nov 8. PMID: 15534200; PMCID: PMC534546. Latifi, A., Ruiz, M., Jean, R., & Zhang, C. C. (2007). PrxQ-A, a member of the peroxiredoxin Q family, plays a major role in defense against oxidative stress in the cyanobacterium Anabaena sp. strain PCC7120. Free radical biology & medicine, 42(3), 424–431. Laukkanen M. O., Cammarota F., Esposito T., Salvatore M., & Castellone M. D. (2015) Extracellular Superoxide Dismutase Regulates the Expression of Small GTPase Regulatory Proteins GEFs, GAPs, and GDI. PLoS ONE 10(3). https://doi.org/10.1371/journal.pone.0121441 Ledgerwood E. C. , Marshall J. W. , & Weijman, J. F. (2017). The role of peroxiredoxin 1 in redox sensing and transducing. Archives of Biochemistry and Biophysics, 617, 60–67. https://doi.org/10.1016/j.abb.2016.10.009 Lee, S. R., Yang, K. S., Kwon, J., Lee, C., Jeong, W., & Rhee, S. G. (2002). Reversible inactivation of the tumor suppressor PTEN by H 2O2. The Journal of biological chemistry, 277(23), 20336–20342. https://doi.org/10.1074/jbc.M111899200 Li, Z., Zhang, Q., & Zhou, X. (2016). A 2-Cys peroxiredoxin in response to oxidative stress in the pine wood nematode, Bursaphelenchus xylophilus. Scientific Reports, 6. Lim, Y. S., Cha, M. K., Kim, H. K., Uhm, T. B., Park, J. W., Kim, K., & Kim, I. H. (1993). Removals of hydrogen peroxide and hydroxyl radical by thiol-specific antioxidant protein 102 as a possible role in vivo. Biochemical and biophysical research communications, 192(1), 273–280. Lim, J. M., Lee, K. S., Woo, H. A., Kang, D., & Rhee, S. G. (2015). Control of the pericentrosomal H2O2 level by peroxiredoxin I is critical for mitotic progression. The Journal of cell biology, 210(1), 23–33. https://doi.org/10.1083/jcb.201412068 Loew O. (1900). A New Enzyme OF General Occurence in Organismis. Science, 11(279), 701– 702. https://doi.org/10.1126/science.11.279.701 Lusebrink, I., Erbilgin, N., & Evenden, M. L. (2013). The lodgepole x jack pine hybrid zone in Alberta, Canada: a stepping stone for the mountain pine beetle on its journey east across the boreal forest. J. Chem. Ecol. 39, 1209–1220. Margis, R., Dunand, C., Teixeira, F. K., & Margis-Pinheiro, M. (2008). Glutathione peroxidase family - an evolutionary overview. The FEBS journal, 275(15), 3959–3970. https://doi.org/10.1111/j.1742-4658.2008.06542.x Mason R. P. (2016). Imaging free radicals in organelles, cells, tissue, and in vivo with immunospin trapping. Redox biology, 8, 422–429. https://doi.org/10.1016/j.redox.2016.04.0 Mason, P. A., & Bowers, M. D. (2017). Localization of Defensive Chemicals in Two Congeneric Butterflies (Euphydryas, Nymphalidae). Journal of Chemical Ecology, 43, 480–486. Mason, C. J., Keefover-Ring, K., Villari, C., Klutsch, J., Cook, S., Bonello, P., Erbilgin, N., Raffa, K., & Townsend, P. (2018). Anatomical defences against bark beetles relate to degree of historical exposure between species and are allocated independently of chemical defences within trees. Plant, cell & environment, 42(2), 633–646. Massoumi-Alamouti, S., Haridas, S., Feau, N., Robertson, G., Bohlmann, J., & Breuil, C. (2014). Comparative genomics of the pine pathogens and beetle symbionts in the genus Grosmannia. Molecular biology and evolution, 31(6), 1454–1474. https://doi.org/10.1093/molbev/msu102 McGonigle, S., Dalton, J.P., & James, E. R. 1998. Peroxiredoxins: A New Antioxidant Family. Parasitology Today, 14(4), 139–145. Mhamdi, A., Queval, G., Chaouch, S., Vanderauwera, S., Van Breusegem, F., & Noctor, G. (2010). Catalase function in plants: a focus on Arabidopsis mutants as stress-mimic models. Journal of experimental botany, 61(15), 4197–4220. https://doi.org/10.1093/jxb/erq282 Misof, B., Liu, S., Meusemann, K., Peters, R. S., Donath, A., Mayer, C., Frandsen, P. B., Ware, J., Flouri, T., Beutel, R. G., Niehuis, O., Petersen, M., Izquierdo-Carrasco, F., Wappler, T., Rust, J., Aberer, A. J., Aspöck, U., Aspöck, H., Bartel, D., Blanke, A., … Zhou, X. 103 (2014). Phylogenomics resolves the timing and pattern of insect evolution. Science, 346(6210), 763–767. https://doi.org/10.1126/science.1257570 Nandi, A., Yan, L. J., Jana, C. K., & Das, N. (2019). Role of Catalase in Oxidative Stress- and Age-Associated Degenerative Diseases. Oxidative medicine and cellular longevity, 2019, 9613090. https://doi.org/10.1155/2019/9613090 National Center for Biotechnology Information (NCBI)[Internet]. Bethesda (MD): National Library of Medicine (US), National Center for Biotechnology Information; [1988] – [cited 2020 Dec 18]. Available from: https://www.ncbi.nlm.nih.gov/ NCBI Resource Coordinators (2018). Database resources of the National Center for Biotechnology Information. Nucleic acids research, 46(D1), D8–D13. https://doi.org/10.1093/nar/gkx1095 Nelson, K. J., & Parsonage, D. (2011). Measurement of peroxiredoxin activity. Current protocols in toxicology, Chapter 7, Unit7.10. https://doi.org/10.1002/0471140856.tx0710s49 Nelson, K. J., Parsonage, D., Hall, A., Karplus, P. A., & Poole, L. B. (2008). Cysteine pK(a) values for the bacterial peroxiredoxin AhpC. Biochemistry, 47(48), 12860–12868. https://doi.org/10.1021/bi801718d Nelson, K. J., Knutson, S. T., Soito, L., Klomsiri, C., Poole, L. B., & Fetrow, J. S. (2011). Analysis of the peroxiredoxin family: using active-site structure and sequence information for global classification and residue analysis. Proteins, 79(3), 947–964. https://doi.org/10.1002/prot.22936 Neumann, C. A., Cao, J., & Manevich, Y. (2009). Peroxiredoxin 1 and its role in cell signaling. Cell cycle (Georgetown, Tex.), 8(24), 4072–4078. https://doi.org/10.4161/cc.8.24.10242 Oakley, T. H., Alexandrou, M. A., Ngo, R., Pankey, M. S., Churchill, C. C. K., & Lopker, K. B. (2014). Osiris: Accessible and reproducible phylogenetic and phylogenomic analyses within the Galaxy workflow management system. BMC Bioinformatics, 15, 230. Ogusucu, R., Rettori, D., Munhoz, D. C., Netto, L. E., & Augusto, O. (2007). Reactions of yeast thioredoxin peroxidases I and II with hydrogen peroxide and peroxynitrite: rate constants by competitive kinetics. Free radical biology & medicine, 42(3), 326–334. https://doi.org/10.1016/j.freeradbiomed.2006.10.042 Okazaki, S., Tachibana, T., Naganuma, A., Mano, N., & Kuge, S. (2007). Multistep disulfide bond formation in Yap1 is required for sensing and transduction of H2O2 stress signal. Molecular cell, 27(4), 675–688. https://doi.org/10.1016/j.molcel.2007.06.035 104 Park, S. G., Cha, M. K., Jeong, W., & Kim, I. H. (2000). Distinct physiological functions of thiol peroxidase isoenzymes in Saccharomyces cerevisiae. The Journal of biological chemistry, 275(8), 5723–5732. https://doi.org/10.1074/jbc.275.8.5723 Parker, J., Parker, K.M., & Keller, L. (2004). Molecular phylogenetic evidence for an extracellular Cu Zn superoxide dismutase gene in insects. Insect Molecular Biology, 13. Parsonage, D., Nelson, K. J., Ferrer-Sueta, G., Alley, S., Karplus, P. A., Furdui, C. M., & Poole, L. B. (2015). Dissecting peroxiredoxin catalysis: separating binding, peroxidation, and resolution for a bacterial AhpC. Biochemistry, 54(7), 1567–1575. https://doi.org/10.1021/bi501515w Peshenko, I. V., & Shichi, H. (2001). Oxidation of active center cysteine of bovine 1-Cys peroxiredoxin to the cysteine sulfenic acid form by peroxide and peroxynitrite. Free radical biology & medicine, 31(3), 292–303. https://doi.org/10.1016/s08915849(01)00579-2 Peskin, A. V., Dickerhof, N., Poynton, R. A., Paton, L. N., Pace, P. E., Hampton, M. B., & Winterbourn, C. C. (2013). Hyperoxidation of peroxiredoxins 2 and 3: rate constants for the reactions of the sulfenic acid of the peroxidatic cysteine. The Journal of biological chemistry, 288(20), 14170–14177. https://doi.org/10.1074/jbc.M113.460881 Peskin, A. V., & Winterbourn, C. C. (2017). Assay of superoxide dismutase activity in a plate assay using WST-1. Free radical biology & medicine, 103, 188–191. https://doi.org/10.1016/j.freeradbiomed.2016.12.033 Petriv, O. I., & Rachubinski, R. (2004). Lack of Peroxisomal Catalase Causes a Progeric Phenotype in Caenorhabditis elegans*. Journal of Biological Chemistry, 279, 19996– 20001. Pitt, C., Robert, J. A., Bonnett, T. R., Keeling, C. I., Bohlmann, J., & Huber, D. P. (2014). Proteomics indicators of the rapidly shifting physiology from whole mountain pine beetle, Dendroctonus ponderosae (Coleoptera: Curculionidae), adults during early host colonization. PloS one, 9(10), e110673. https://doi.org/10.1371/journal.pone.0110673 Portillo-Ledesma, S., Randall, L. M., Parsonage, D., Dalla Rizza, J., Karplus, P. A., Poole, L. B., Denicola, A., & Ferrer-Sueta, G. (2018). Differential Kinetics of Two-Cysteine Peroxiredoxin Disulfide Formation Reveal a Novel Model for Peroxide Sensing. Biochemistry, 57(24), 3416–3424. https://doi.org/10.1021/acs.biochem.8b00188 Possee R. D. (1997). Baculoviruses as expression vectors. Current opinion in biotechnology, 8(5), 569–572. https://doi.org/10.1016/s0958-1669(97)80030-4 Qin, J., Lu, M., Zheng, Y., & Du, Y. (2016). Molecular Cloning, Characterization, and Functional Analysis of Catalase in Frankliniella occidentalis (Thysanoptera: Thripidae). Annals of the Entomological Society of America, 110, 212–220. 105 Rhee S. G. (2016). Overview on Peroxiredoxin. Molecules and cells, 39(1), 1–5. https://doi.org/10.14348/molcells.2016.2368 Rhee, S. G., Woo, H. A., Kil, I. S., & Bae, S. H (2012). Peroxiredoxin functions as a peroxidase and a regulator and sensor of local peroxides. J. Biol. Chem., 287, 4403–4410. Robert, J. A., Bonnett, T., Pitt, C., Spooner, L. J., Fraser, J., Yuen, M. M., Keeling, C. I., Bohlmann, J., & Huber, D. P. (2016). Gene expression analysis of overwintering mountain pine beetle larvae suggests multiple systems involved in overwintering stress, cold hardiness, and preparation for spring development. PeerJ, 4, e2109. https://doi.org/10.7717/peerj.2109 Robert, J. A., Pitt, C., Bonnett, T. R., Yuen, M. M., Keeling, C. I., Bohlmann, J., & Huber, D. P. (2013). Disentangling detoxification: gene expression analysis of feeding mountain pine beetle illuminates molecular-level host chemical defense detoxification mechanisms. PloS one, 8(11), e77777. Rudnik, R., Bulcha, J. T., Reifschneider, E., Ellersiek, U., & Baier, M. (2017). Specificity versus redundancy in the RAP2.4 transcription factor family of Arabidopsis thaliana: transcriptional regulation of genes for chloroplast peroxidases. BMC plant biology, 17(1), 144. https://doi.org/10.1186/s12870-017-1092-5 Saab, V. A., Latif, Q. S., Rowland, M.M., Johnson, T. N., Chalfoun, A. D., Buskirk, S. W., Heyward, J. E., Dresser, M. A. (2014). Ecological Consequences of Mountain Pine Beetle Outbreaks for Wildlife in Western North American Forests. Forest Science, 60(3)3, 539–559. https://doi.org/10.5849/forsci.13-022 Seybold, S., Huber, D.P., Lee, J., Graves, A.D., & Bohlmann, J. (2006). Pine monoterpenes and pine bark beetles: a marriage of convenience for defense and chemical communication. Phytochemistry Reviews, 5, 143–178. Sheng, Y., Abreu, I. A., Cabelli, D. E., Maroney, M. J., Miller, A. F., Teixeira, M., & Valentine, J. S. (2014). Superoxide dismutases and superoxide reductases. Chemical reviews, 114(7), 3854–3918. https://doi.org/10.1021/cr4005296 Safranyik, L., Carroll, A. L., Régnière, J., Langor, D. W., Riel, W. G., Shore, T. L., Peter, B., Cooke, B. J., Nealis, V. J., & Taylor, S. W. 2010. Potential for Range Expansion of Mountain Pine Beetle into the Boreal Forest of North America. The Canadian Entomologist, 142(5), 415–442. https://doi.org/10.4039/n08-CPA01 Schneider, S., Sandalova, T., Schneider, G., Sprenger, G. A., & Samland, A. K. (2008). Replacement of a phenylalanine by a tyrosine in the active site confers fructose-6phosphate aldolase activity to the transaldolase of Escherichia coli and human origin. The Journal of biological chemistry, 283(44), 30064–30072. https://doi.org/10.1074/jbc.M803184200 106 Shikano, I., Rosa, C., Tan, C. W., & Felton, G. W. (2017). Tritrophic Interactions: MicrobeMediated Plant Effects on Insect Herbivores. Annual review of phytopathology, 55, 313– 331. https://doi.org/10.1146/annurev-phyto-080516-035319 Singh, P., Sharma, L., Kulothungan, S. R., Adkar, B. V., Prajapati, R. S., Ali, P. S., Krishnan, B., & Varadarajan, R. (2013). Effect of signal peptide on stability and folding of Escherichia coli thioredoxin. PloS one, 8(5), e63442. https://doi.org/10.1371/journal.pone.0063442 Stadtman, E. R., & Oliver, C. N. (1991). Metal-catalyzed oxidation of proteins. Physiological consequences. The Journal of biological chemistry, 266(4), 2005–2008. Stothard, P. (2000) The Sequence Manipulation Suite: JavaScript programs for analyzing and formatting protein and DNA sequences. Biotechniques, 28, 1102–1104. Sun, C. C., Dong, W. R., Shao, T., Li, J. Y., Zhao, J., Nie, L., Xiang, L. X., Zhu, G., & Shao, J. Z. (2017). Peroxiredoxin 1 (Prx1) is a dual-function enzyme by possessing Cysindependent catalase-like activity. The Biochemical journal, 474(8), 1373–1394. https://doi.org/10.1042/BCJ20160851 Swofford, D.L. (2003) PAUP*. Phylogenetic analysis using parsimony (* and other methods). Version 4. Sinauer Associates, Sunderland. Tasaki, E., Kobayashi, K., Matsuura, K., & Luchi, Y. (2018). Long-Lived Termite Queens Exhibit High Cu/Zn-Superoxide Dismutase Activity. Oxidative medicine and cellular longevity, 2018, 5127251. https://doi.org/10.1155/2018/5127251 Tholl D. (2015). Biosynthesis and biological functions of terpenoids in plants. Advances in biochemical engineering/biotechnology, 148, 63–106. https://doi.org/10.1007/10_2014_295 Tsuzuki, D., Hichiya, H., Okuda, Y., Yamamoto, S., Tamagake, K., Shinoda, S., & Narimatsu, S. (2003). Alteration in catalytic properties of human CYP2D6 caused by substitution of glycine-42 with arginine, lysine and glutamic acid. Drug metabolism and pharmacokinetics, 18(1), 79–85. Turrens, J. F. (2003). Mitochondrial formation of reactive oxygen species. The Journal of physiology, 552(Pt 2), 335–344. https://doi.org/10.1113/jphysiol.2003.049478 Vivancos, A. P., Castillo, E. A., Biteau, B., Nicot, C., Ayté, J., Toledano, M. B., & Hidalgo, E. (2005). A cysteine-sulfinic acid in peroxiredoxin regulates H2O2-sensing by the antioxidant Pap1 pathway. Proceedings of the National Academy of Sciences of the United States of America, 102(25), 8875–8880. Vornam, B., Leinemann, L., Peters, F.S., Wolff, A., Leha, A., Salina-Riester, G., Shumacher, J., & Gaiing, O. (2019). Response of Scots pine (Pinus sylvestris) seedlings subjected to 107 artificial infection with the fungus Sphaeropsis sapinea. Plant Mol. Biol. Rep., 37, 214– 223. https://doi.org/10.1007/s11105-019-01149-2 Wang, Q., Chen, K., Yao, Q., Zhao, Y., Li, Y., Shen, H., & Mu, R. (2008). Identification and characterization of a novel 1-Cys peroxiredoxin from silkworm, Bombyx mori. Comparative Biochemistry and physiology. Biochemistry & Molecular Biology, 149(1), 176–182. https://doi.org/10.1016/j.cbpb.2007.09.005. Wang, Q., Chen, K., Yao, Q., Zhao, Y., Li, Y., Shen, H., & Mu, R. (2008). Identification and characterization of a novel 1-Cys peroxiredoxin from silkworm, Bombyx mori. Comparative Biochemistry and physiology. Biochemistry & Molecular Biology, 149(1), 176–182. https://doi.org/10.1016/j.cbpb.2007.09.005. Wang, Q., Zhou, Y., Chen, K., & Ju, X. (2016). Identification and characterization of an atypical 2‐cys peroxiredoxin from the silkworm, Bombyx mori. Insect Molecular Biology, 25(4), 347–354. Wang, Y., Oberley, L. W., & Murhammer, D. W. (2001). Evidence of oxidative stress following the viral infection of two lepidopteran insect cell lines. Free radical biology & medicine, 31(11), 1448–1455. https://doi.org/10.1016/s0891-5849(01)00728-6 Williams, J. R., Trias, E., Beilby, P. R., Lopez, N. I., Labut, E. M., Bradford, C. S., Roberts, B. R., McAllum, E. J., Crouch, P. J., Rhoads, T. W., Pereira, C., Son, M., Elliott, J. L., Franco, M. C., Estévez, A. G., Barbeito, L., & Beckman, J. S. (2016). Copper delivery to the CNS by CuATSM effectively treats motor neuron disease in SOD(G93A) mice coexpressing the Copper-Chaperone-for-SOD. Neurobiology of disease, 89, 1–9. https://doi.org/10.1016/j.nbd.2016.01.020 Wingfield P. T. (2015). Overview of the purification of recombinant proteins. Current protocols in protein science, 80, 6.1.1–6.1.35. https://doi.org/10.1002/0471140864.ps0601s80 Winterbourn, C. C. (2008). Reconciling the chemistry and biology of reactive oxygen species. Nature chemical biology, 4(5), 278–286. https://doi.org/10.1038/nchembio.85 Winterbourn, C., & Peskin, A. (2016). Kinetic Approaches to Measuring Peroxiredoxin Reactivity. Molecules and Cells, 39, 26–30. Woo, H. A., Yim, S. H., Shin, D. H., Kang, D., Yu, D. Y., & Rhee, S. G. (2010). Inactivation of peroxiredoxin I by phosphorylation allows localized H 2O2 accumulation for cell signaling. Cell, 140(4), 517–528. https://doi.org/10.1016/j.cell.2010.01.009 Wood, Z. A., Poole, L. B., & Karplus, P. A. (2003). Peroxiredoxin evolution and the regulation of hydrogen peroxide signaling. Science, 300(5619), 650–653. https://doi.org/10.1126/science.1080405 108 Xikeranmu, Z., Abdunasir, M., Ma, J., Tusong, K., & Liu, X. (2019). Characterization of two copper/zinc superoxide dismutases (Cu/Zn-SODs) from the desert beetle Microdera punctipennis and their activities in protecting E. coli cells against cold. Cryobiology, 87, 15–27. https://doi.org/10.1016/j.cryobiol.2019.03.006 Yamamoto, K., Banno, Y., Fujii, H., Miake, F., Kashige, N., & Aso, Y. (2005). Catalase from the silkworm, Bombyx mori: Gene sequence, distribution, and overexpression. Insect Biochemistry and Molecular Biology, 35(4), 277–283. https://doi.org/10.1016/j.ibmb.2005.01.00 Zamocky, M., Furtmüller, P. G., Bellei, M., Battistuzzi, G., Stadlmann, J., Vlasits, J., & Obinger, C. (2009). Intracellular catalase/peroxidase from the phytopathogenic rice blast fungus Magnaporthe grisea: expression analysis and biochemical characterization of the recombinant protein. The Biochemical journal, 418(2), 443–451. https://doi.org/10.1042/BJ20081478 Zelko, I. N., Mariani, T. J., & Folz, R. J. (2002). Superoxide dismutase multigene family: a comparison of the CuZn-SOD (SOD1), Mn-SOD (SOD2), and EC-SOD (SOD3) genhang, L., & Lu, Z. (2015). Expression, purification and characterization of an atypical 2-Cys peroxiredoxin from the silkworm, Bombyx mori. Insect molecular biology, 24(2), 203–212. https://doi.org/10.1111/imb.12149 Zhang, L, & Lu, Z. (2015). Expression, purification and characterization of an atypical 2-Cys peroxiredoxin from the silkworm, Bombyx mori. Insect Molecular Biology, 24(2), 203– 212. Zhang, Z., Guo, R., & Li, M. (2017). Identification, characterization and expression analysis of superoxide dismutase genes in Dastarcus helophoroides (Coleoptera: Bothrideridae). Gene, 606, 25–34. https://doi.org/10.1016/j.gene.2017.01.003 109