ACTA BIOLOGICA SLOVENICA LJUBLJANA 2013 Vol. 56, [t. 1: 11–25 ACTA BIOLOGICA SLOVENICA LJUBLJANA 2013 Vol. 56, [t. 1: 11–25 The synergy of xenobiotics in honey bee Apis mellifera: mechanisms and effects Sinergizem ksenobiotikov v medonosni cebeli Apis mellifera: mehanizmi in ucinki. Gordana Glavan* and Janko Božic University of Ljubljana, Biotehnical faculty, Department of Biology, Vecna pot 111, Ljubljana SI-1000, Slovenia *correspondence: gordana.glavan@bf.uni-lj.si Abstract:Duringforagingactivitieshoneybeesarefrequentlyexposedtodifferent xenobiotics, most of them are agrochemical pesticides and beehive chemicals. Many pesticides are applied together and synergism is likely to occur in different organisms. The risk of synergisms is neglected and relatively few studies were performed concerning the effects and synergy mechanism of different xenobiotic combinations in honeybees. The understanding of synergy mechanisms between xenobiotics is very important for the control of defined mixtures use and also for the prediction of potential toxicity of newly developed substances in agriculture and apiculture. This review is focused on the effects, mechanisms and molecular targets of xenobiotics in honeybees and possible complex mechanisms of their synergisms. The main threat for honeybees are insecticides which primary molecular targets are few neuronal molecules therefore causing the impairment of neuronal system that have a profound effect on honeybee behavior, cognitive functions and physiology. However, the majority of synergistic effects observed in honeybees were ascribed to the inhibition of detoxifying midgut enzymes P450 involved in xenobiotic metabolism since most of studies were done with the mixtures xenobiotic/P450 inhibitor. The main inhibitors of P450 enzymes are specific compounds used to prolong the effects of pesticides as well as some fungicides. Some insecticides can also interact with these enzymes and influence the xenobiotis. Although the primary mechanisms of action of individual xenobiotics especially insecticides are well known and there are possible interactions in honeybees at their primary target sites, this issue is underestimated and it warrants further investigation. Keywords: synergism, xenobiotic, Apis mellifera, mechanism, pesticide, P450 Izvlecek: Medonosne cebele so med iskanjem hrane pogosto izpostavljene razlicnim ksenobiotikom, vecinomaso to fitofarmacevtska sredstva in panjske kemikalije. Številna fitofarmacevtska sredstva se uporablja skupaj in znano je, da lahko pride do sinergisticnih interakcij v organizmih. Tveganje za nastanek sinergizmov je podcenjeno in narejenih je relativno malo študij na cebelah o ucinkih in mehanizmih sinergizmov razlicnih kombinacij ksenobiotikov. Razumevanje mehanizmov sinergizmov ksenobiotikov je zelo pomembno za nadzor nad uporabo definiranih Acta Biologica Slovenica, 56 (1), 2013 mešanic in napovedovanje potencialne toksicnosti novih ksenobiotikov v kmetijstvu in cebelarstvu. Pregledni clanek se osredotoca na ucinke, mehanizme in molekulske tarce ksenobiotikov v medonosnih cebelah in osvetljuje morebitne primere ter mehanizmenastankasinergizmov. Najboljnevarnizacebelesoinsekticidi,katerihprimarne tarce so nekatere molekule živcnih celic,zato le-ti motijo delovanje živcnega sistema. Insekticidi zato lahko mocno vplivajo na vedenje, kognitivne funkcije in fiziologijo cebel. Kljub temu raziskovalci vecino sinergijskih ucinkov v cebelah razlagajo z inhibicijo crevesnih detoksifikacijskih encimov P450, ki presnavljajo ksenobiotike, saj je bila vecina študij narejena z mešanicami ksenobiotik/zaviralec encimov P450. Glavni zaviralci encimov P450 so specificni inhibitorji za podaljšanje ucinka fitofarmacevtskih sredstev ter nekateri fungicidi. Tudi nekateri insekticidi lahko vplivajo na delovanje encimovP450 in tako vplivajo na interakcije med ksenobiotiki. Ceprav so primarni mehanizmi delovanja posameznih ksenobiotikov, še posebej insekticidov, precej znani in so sinergizmina ciljnih tarcah pri cebelah možni, je to podrocje pod- cenjeno in neraziskano. Kljucne besede: sinergizem, ksenobiotik, Apis mellifera, mehanizem, pesticidi, P450 Introduction In addition to gathering nectar to produce honey, honey bees carry out another crucial function: pollination of agricultural crops, home gardens, orchards and wildlife habitat. A substantial decline of honey bee populations so called colony collapse disorder was observed in the last 15 years in many countries in Europe and in North America (vanEngelsdorp and Meixner 2010).Colony numbers in Europe for example decreased from over 21 million in 1970 to about 15.5millionin2007(FAO,2009),aseveredecline occurredafter1990.Manyfactorssuchasdiseases, parasites, xenobiotics (pesticides and veterinary products), the environment, and socio-economic factorsprobablyinfluencemanagedbeepopulation, workingaloneorincombinations(vanEngelsdorp and Meixner 2010). Honey bees may frequently become exposed to xenobiotics, environmental chemicals as a consequence of their foraging activities. Most of them are agrochemical pesticides and beehive or veterinary products, many of them of insecticide action, used against parasitic honey bee mites: Acarapis sp., Varroa destructor and Aethina tumida (Thompson 2012). The use of pesticides to control weeds, fungi and arthropod pests seems inevitable in modern agriculture which seeks for thehighestyieldsoftheproduces.Nectarforaging bees are likely to experienced highest exposure to both sprayed and systemic seed and soil treatments compounds followed by nurse and brood- attending bees. The residues of pesticides were found in pollen, wax and nectar within colonies, pollen and nectar residues from plants, in pollen loads on bees returning to the hives and in adult workers (Thompson 2012). Pesticide regulations sofarfocusedmainlyonprotectionofbeesagainst direct poisoning (Thompson and Wilkins 2003, Desneuxetal.,2007).Thedirectpoisoningisnow regulatedandpreventedbytheimplementationof EuropeanCouncilDirective91/414inEurope,and theFederalInsecticide,FungicideandRodenticide ActintheUS(Desneuxetal.2007,vanEngelsdorp and Meixner 2010). The standard approaches for determination of acute pesticide toxicity in bees are the calculation of the LD50 (median lethal dose) or LC50 (median lethal concentration) of a given substance with respect to adult bees or larvae.Inspiteofmoreorlesscontrolledprotection against direct poisoning, massive dying of honey bees is still present. For this reason many studies are focusing to the chronic sub-lethal exposure of xenobiotics causing a variety of sub-lethal effects on bees (reviewed by Desneux 2007) which are physiological and behavioral, affecting the honeybee colony as a whole, resulting in the perturbations of learning and communication ability.Evenmore,asmanypesticidesareapplied Glavan et al.: The synergy of xenobiotics in honey bee Apis mellifera: mechanisms and effects together,scientistsarearguingforyearsthattoxic exposures to pesticides should be measured as they would normally occur, in combination with one another. The most intriguing or concerning aspectofpesticidemixturesistheopportunityfor complex interactions such as a synergy when the administration of one chemical increases the toxic- ityofanother.Therearerelativelyfewexperimental data regarding synergistic effects of pesticides on honeybees, but in some cases pesticide mixtures, particularlywithinsecticides,havebeenshownto besynergistic,withreportedincreases intoxicity of up to100-fold (Thomson 1996). However, the effects of pesticide exposure on colony health is not systematicallymonitored,andtheEnvironmental Protection Agency (EPA) does not require data on sub-lethal or synergistic effects for pesticide registration (NAS, 2009) therefore this specific issue warrants special attention. This review focuses on the mechanisms and molecular targets of xenobiotics in honeybees which could be the basis of their synergism, especially insecticides which are the most potentially dangerousforhoneybees.Sinceofgreaterpotential to cause underestimation of the risk posed to the honeybeecoloniesdeclinethemodesofsynergisms of xenobiotics known so far are emphasized and summarizedinthissectionofourreview.Theaim istoexemplifythepossiblecomplexmechanisms of their interaction. Mechanisms and effectsof xenobiotics Agrochemical pesticides Theagrochemicalpesticidesaffectinghoneybee colonies arefungicides, herbicides and insecticides, applied to crops (Johnson et al. 2010). The large numberofcommercialpesticidesusedworldwide, whetherbasedonnaturalproductsorbeingentirely ofsyntheticorigin,actonrelativelyfew,perhaps 95biochemicaltargetsinpestinsects,weeds,and destructive fungi (Casida 2009). Herbicides in general are blocking photosynthesis, carotenoid synthesis,oraromaticandbranchedchainamino acidsynthesisessentialinplants.Manyfungicides inhibit ergosterol (the fungal sterol) or tubulin biosynthesis or cytochrome c reductase. Others disturb basic cellular functions (Casida 2009). The pesticides that represent a main threat to the honeybees are insecticides. Many of the most effective insecticides in current use act on the insect nervous system (Narahashi 1992, Bloomquist 1996).Others are insect growth regulators(Tasei2001,Thomsonetal.2005).The growth regulating insecticides are functioning as juvenile hormone analogues (fenoxycarb), chitin synthesis inhibitors (diflubenzuron), ecdysteroid synthesis inhibitors (azadirachtin) and ecdysteroid analogues (tebufenozide) (Tasei 2001, Thomson et al. 2005).The main nerve targets of current insecticides are voltage-gated sodium channels, an enzyme acetylcholinesterase (AChE) and re- ceptorsforneurotransmitters:L-glutamate-gated chloridechannelsworkingasglutamatereceptors (GluRs),ionotropic.-aminobutyricacid(GABA) receptors(GABARs) binding GABAand nicotinic cholinergic receptors (nAChRs) stimululated by acetylcholine (ACh) (Coats 1990, Fukuto 1990, Zlotkin1999,Bloomquist2003,Raymond-Delpech etal.2005,WolstenholmeandRogers2005,Davies etal.2007,JeschkeandNauen2008).Voltage-gated sodium channels are molecular targets for three big groups of insecticides, pyrethroid,DDT-type and organ chlorine insecticides (Coats 1990). These channels mediate the transient wave of sodium entry spreading along the nerve axons and dendrites, carrying the action potential along thesestructuresandareubiquitousinthehoneybee nervous system (Sattelle and Yamamoto 1988, Narahashi 1992). All three groups of insecticides cause death due to hyperexcitation of the nerves, butinslightlydifferentway.Pyrethroidpesticides by binding to voltage-gated sodium channels in- ducehyperexcitationthatresultsfromprolongation of the open phase of sodium gate function results inneurotoxic effectssuch astremorsandconvulsions. TheDDT-typeinsecticides,DDT(dichloro diphenyl trichloroethane) and DDT analogues (N-alkylamides,dihydropyrazoles),actprimarily on the peripheral nervous system (Coats 1990). The mechanism of DDTis the prevention of the deactivation or closing of that gate after activation and membrane depolarization. The result is a persistent leakage of Na+ ions through the nervemembrane,creatingadestabilizingnegative afterpotential.Thehyperexcitabilityofthenerveis theconsequenceoftrainsofrepetitivedischarges in the neuron after a single stimulus and/or occur Acta Biologica Slovenica, 56 (1), 2013 spontaneously (Coats 1990).The acute toxic effectsinanimalsoforganchlorineinsecticidesare alsoduetohyperexcitationinthenervoussystem and death is frequently recognized as respiratory failure after the disruption of nervous system function (Coats 1990). Organophosphate and carbamate insecticides are inhibiting the action of AChE (Fukuto 1990). AChEis an enzyme that terminates the syn- apticactionsofACh,theimportantneurotransmitter of sensory neurons and interneurons of insect brain which is necessary for sensory-input processing and learning in honey bee (Massoulie et al. 1993, Homberg1994,Weinberger2006).AChEiswidely distributed in the insect brain, the thoracic and abdominal segments and the abdominal ganglia (Kreissl in Bicker 1989, Thany et al. 2010).The potential target sites for organophosphate and carbamate insecticides in the honeybee brain are the optic lobes, antennal afferents projecting into the dorsal lobe, fibers connecting the two brain hemispheres, and within the protocerebrum and the mushroom bodies where AChE is highly expressed (Kreissl in Bicker 1989). AChE was found also in the compound eye and ocelli (Kral 1980, Kral and Schneider 1981).The inhibition of AChE by organophosphate and carbamate insecticides causes irreversible blockage leading to accumulation of the enzyme which results in overstimulation of cholinergic receptors (Fukuto 1990).AsAChisamajorneurotransmitterofinsect nervoussystem(Homberg1994)theinhibitionof AChEcouldcauseasystemicfailureintheinsect body. Widely used organophosphate as hive var- roacides is coumaphos. Insecticides that act selectively on insect nAChRs as potent agonists are neonicotinoids (Jeschke and Nauen 2008). Among ionotropic receptors affected by insecticides, nAChRs are the most abundant excitatory postsynaptic receptors (Sattelle 1980).The central nervous system of insects is rich in nAChRs more so than any other organism (Jones and Sattelle 2010).They are located postysinaptically and directly activated by ACh, released from presynaptic cholinergic neurons facilitating fast excitatory synaptic transmission (Thany etal. 2010).In thehoneybee brain the highest binding site densities for nAChR arelocalizedinthesuboesophagealganglion,the optic tubercles, optic lobes medulla and lobula, antennal lobes, dorsal lobes and the a-lobes of the mushroom bodies (Scheidler et al. 1990). Neonicotinoids cause excitation of the neurons and because of a high concentration of nACh receptors in honeybees the eventual paralysis could be very profound occurring at low concentration of neonicotinoids, leading to death. TheinsecticidesthatinterferewithGABARs are pyrethroidsandphenylpyrazoleinsecticides (Raymond-Delpech et al. 2005, Davies et al. 2007). In insects GABARs are associated with neurotransmitter GABA mediating inhibitory synaptic transmission in the nervous system and at nerve-muscle junctions (Homberg 1994). In the central nervous system of honeybees the neurotransmitter GABA is generally present in neuropil, especially in structures that are associated with learning and memory, such as antennal lobe and the mushroom body and the optic lobe (SchaferandBicker1986,ElHassanietal.2009). The presence of the neurotransmitter GABA in the honeybee brain was shown mainly for local interneuronsandlessintheprojectionneurons.In thebrainandsubesophagealgangliononlyminority of neurons contained GABA(Bicker et al. 1985, Meyer et al. 1986, Schafer and Bicker 1986). By targeting the GABARs which are chloride channels pyrethroids and phenylpyrazole insecticides disrupt normal neuronal influx (e.g., passage of chloride ions) and, at sufficient doses, causing excessiveneuralexcitation,severeparalysis,and death (Cole et al. 1993, Gunasekara et al. 2007). Most known representative of phenylpyrazole insecticidesisfipronilandwidelyusedpyrethroid ashivevarroacidesistau-fluvalinate.Pyrethroids are very complex group regarding the molecular mechanisms of their functioning, because they don’tbindonlytoGABA-gatedchloridechannel, but they can also interfere with other molecules such as calcium regulation. They could inhibit bothCa-ATPaseandCa-MgATPase(Coats1990). In this respect direct effects on neurotransmitter releasehavebeenobserved,aswellastheinhibition of Ca2+ uptake. However, they have also various secondary targets such as signal transduction pathwaysbyalteringtheproteinphosphorylation cascade that may result, among other things, in programmed cell death (Ray and Fry 2006). In mammals a variety of different effects of pyre- throidswerediscoveredlikemodulationofprotein Glavan et al.: The synergy of xenobiotics in honey bee Apis mellifera: mechanisms and effects phosphorylation,voltage-gatedsodiumchannels, voltage-gated chloride channels, noradrenaline release, membrane depolarization, GABA-gated chloride channels, nicotinic receptors, mitochondrial complex I, apopotosis induction, voltage- gatedcalciumchannels,lymphocyteproliferation, volume-sensitiveanionchannels,calciumATP-ase, intercellular gap junctions and chromosomal damage, but many of these effects were not shown for insects (Ray and Fry 2006). The insecticides that activate GluRs which bind neurotransmitter L-glutamate are avermectin and milbemycin(Raymond-Delpech et al. 2005, WolstenholmeandRogers2005).Thedistribution of GluRs in the nervous system of honeybees is not known but they probably modulate excitability in the nervous system and muscle cells as neurotransmitterL-glutamateisenrichedinthese tissues (Cully et al. 1996). Studies performed by Maleszka et al. (2000) and Locatelli et al. (2005) suggested that glutamatergic neurons in the honeybee brain, in particular those found in themushroombodies,maybepartofthecircuitry involvedinprocessingofolfactorymemory.Inthe honeybee, a high level of a glutamate transporter is present in the optic lobes and in restricted areas of the mushroom bodies corresponding to the Kenyon cells of the calyces (Kucharski et al. 2000). GluRs are permeable to chloride ions and theactivationofthesereceptorswithinsecticides avermectin and milbemycin causes a very long- lasting hyperpolarization or depolarization of the neuron or muscle cell and therefore blocking further function leading to paralysis and death (Wolstenholme and Rogers 2005). Insecticides have various neural effects in honeybees that were in details reviewed by Belzunces et al. (2012). They impair cognitive functions, including learning and memory, ha- bituation,olfactionandgustation,navigationand orientation. They affect also behavior, including foraging and physiological functions, including thermoregulation and muscle activity. Acaricides Commonly used in hive varroacides are amitraz, coumaphosandtau-fluvalinate(Johnsonetal. 2010). Amitraz is a formamidine pesticide. The mode of action of formamidine pesticides such in insects is believed to be the toxic effects on a Gprotein-coupledreceptorforaneuromodulator octopamine, working as octopaminergic agonists (Evans and Gee 1980, Dudai et al. 1987). High levels of octopamine in the honey bee brain are associatedwithincreasedforagingbehavior(Schulz and Robinson 2001). Forager honey bees treated withoctopamineincreasedthereportedresource valuewhencommunicatingviathedancelanguage (Barronetal.2007).However,theeffectsofamitraz on foraging activity of honeybees were not investigated, but the acute toxicity of this compound was showninlarvaewhereitincreasesapoptoticcell death in the midgut (Gregorc and Bowen 2000). Another popular in hive varroacide is tau-fluvalinate which was initially very effective at controlling Varroa mites by blocking voltage-gated sodium channels (Davies et al. 2007). Tau-fluvalinate was quite promising since it is tolerated by bees inhighconcentrationsduetorapiddetoxification bycytochromeP450monooxygenases,butmany Varroa populations are now resistant (Lodesaniet et al. 1995, Johnson et al. 2009). However, tau- fluvalinateisnotcompletelyharmless,highdoses couldaffectqueenstogrowsmalleranddronesto dieuntil reaching sexual maturity (Rindereret et al. 1999, Haarmann et al. 2002).As the efficacy of tau-fluvalinate against Varroa was beginning to decrease, coumaphos, an organophosphate pesticide, was starting to be used (Elzen and Westervelt 2002). Although honey bees can tolerate similar to tau-fluvalinate therapeutic doses of coumaphos, probably as a result of detoxicative P450 activity (Johnsonet et al. 2009), negative effects from coumaphos exposure were observed. Queens exposed to coumaphos were smaller, sufferedhighermortalityandweremorelikelytobe rejected when brought into a colony (Haarmann etal.2002,Collinsetal.2004,Pettisetal.2004). Dronespermviabilitywaslowerinstoredsperm collected from drones treated with coumaphos (Burleyetal.2008).Coumaphosalsoaffectsfood transferbetweenworkersofhoneybee(Bevketal. 2011).Fenpyroximateisapyrazoleacaricidethat presumably kills mites through the inhibition of electrontransportinthemitochondriaatcomplex I, thereby interfering with energy metabolism (Mo- toba et al.1992). It wasfound that chronic exposure tofenpyroximate causesthe increasedgeneration of reactive oxygenspecies (Sherer et al. 2007). Acta Biologica Slovenica, 56 (1), 2013 Two monoterpenoid components of plant- derived essential oils, thymoland menthol, are used for control of Varroa and tracheal mites. Theywerefoundtobeamongthemosttoxicofall terpenoidstestedwhenappliedtohoneybeesasa fumigant(EllisandBaxendale1997).Thethymol molecular targets include binding to octopamine receptors (Enan 2001) and AChE (Priestley et al. 2003), but also insect tyramine and GABAreceptors( Blenauetal.2011).Receptoractivationleads to changes in the concentration of intracellular second messengers such as cAMPor InsP3/Ca2+. Thymol could affect honeybees inducing brood removal (Marchetti and Barbattini 1984, Floris et al. 2004) and the increase of queen mortality (Whittington et al. 2000). Exposure to thymol was shown to decrease phototactic behavior in the honeybee (Bergougnoux et al. 2013). Among organic acids, formic acid and oxalicacidareusedasvarroacides.Formicacidis inhibiting electrontransport in the mitochondria bindingofcytochromecoxidaseinmitesandmay produce a neuroexcitatory effect on arthropod neurons (Keyhani and Keyhani 1980, Song and Scharf 2008). Formic acid can reduce worker longevity (Underwood and Currie 2003) and harmingbroodsurvival(Fries1991).Themodeof action of oxalic acid against Varroa is unknown, but in mammals it interferes with mitochondrial electron transport leading to increased production of reactive oxygen species and to kidney toxicity (Caoetal.2004,Meimaridouetal.2005).Repeated treatment of colonies with oxalic acid can result in higher queen mortality and a reduction in the amount of sealed brood (Higes et al. 1999). The midguts of honey bees fed oxalic acid in sugar water exhibited an elevated level of cell death (Gregorc and Smodisskerl 2007).Recent studies arefocusingonmolecularmechanismsunderlying thesub-lethaleffectsofin-hiveacaricidesonhoney bees. Using a gene expression profiling Boncristiani et al. (2012) found that thymol,coumaphos and formic acid are able to alter detoxification gene expression pathways, components of the immune system responsible for cellular response anddevelopmentalgenes.Thisstudyindicatesthat these acaricides could significantly influence the health of individual honey bees and entire colonies (Boncristiani et al. 2012). Mechanisms and factors influencing the synergy of xenobiotics applied to honeybees Understanding the toxicity and synergy of chemicals in organisms requires considering the molecular mechanisms involved as well as the relationshipsbetweenexposureconcentrationand toxic effects with time (Tennekes and Sánchez- Bayo 2013).In addition, the relevance of synergy of xenobiotics is a subject to understanding the routesofapplication,thewayoftransportationto targetmolecules in thetissueand themetabolism of pesticides in the target organism, all having a profound influence on the concentration and chemical structure of active substances at target sites. The analysis of the studies when monitoring the residuesinhoneybeesfollowingin-hivetreatments or pesticide applications revealed that the highest exposure routes were sprayed and systemic seed and soil compound treatmentsto which preferentially foraging bees are exposed during collecting contaminated nectar and the direct exposure to acaricidesusedinbeehives(Thomson2012).This is probably due to the availability of relatively high concentration of agricultural pesticides and in-hive compounds, but also the time between pesticide application to crops and bee exposure is very important as many pesticides degrade or dilute in the environment. The importance of other routes of exposure such as dusts produced during sowing of treated seeds, water from puddles or guttation droplets and beeswax might be relevantbutdataaboutthesearelimited.Thefinal actions of xenobiotics are greatly dependent on the mode of exposure, acute, sub-chronic and chronic, defining the nature and the intensity of their effects. Metabolism of xenobiotics elicited by intrinsic enzymes is remarkably important as it could result in the elevation or decrease of theirtoxicityoritcouldproducedifferenteffects. Chemicalinteractionsbetweenxenobioticsinthe mixture are also possible, causing the changes in chemicalstructuresofparticularsubstance.There are also other factors such as physiological states of the organisms including age, the season and the capacity of immune system that have impact on synergism (Thomson 2012). For example, the immunesystemofhoneybeescouldbeprofoundly affected by various pathogens, bacterial, fungal Glavan et al.: The synergy of xenobiotics in honey bee Apis mellifera: mechanisms and effects and viral pathogens as well as ecto- and endo- parasites that in many cases elevate the toxicity of xenobiotics. Most of the studies in honeybees have focused on the synergisms at the level of midgut enzymes when certain xenobiotic inhibit the detoxifying ability of these enzymes and potentiate the toxicity of another substance, but the synergism at target site is poorly investigated. The synergism at the level of midgut detoxifying enzymes Probablythemostfrequentwayofthetransfer of xenobiotics into honeybee tissue is the consummation of contaminated nectar and absorption in the midgut thought the midgut wall into the hemolymph,butalsopassagethoughtcuticleand sometimes inhalation of vaporous compounds is possible. In the midgut of the honeybee xenobiotics are metabolized by enzymes glutathione-Stransferases (GSTs), cytochrome P450 monooxygenases (P450s) and carboxyl/cholinesterases (CCEs)(ScottandWen2001,Enayatietal.2005, Wheelocketal.2005).Theseenzymesmetabolize pesticidesbydifferentmechanisms,butP450sare probably the most important for honeybees as they play a significant role in the detoxification of phytochemicals present in the nectar, honey and pollen that bees consume (Mao et al. 2009). They catalyze a range of reactions including oxidation and demethylation which decrease pesticide activity or produce active metabolites (Scott and Wen 2001).For example, they convert the thion to oxon forms of organophosphorus pesticides or change neonicotinoid thiamethoxam to clothianidin. P450s can also oxidaze aromatic rings of tau-fluvalinate and flumethrin used in varroa control (Ortiz de Motellano and De Voss 2005). GSTs in insects can metabolize insecticides by facilitating their reductive dehydrochlorination or byconjugationreactionswithreducedglutathione, toproducewater-solublemetabolitesthataremore readily excreted. In addition, they contribute to the removal of toxic oxygen free radical species produced through the action of pesticides (Enayati et al. 2005).Carboxylesterases (CaEs) are hydrolazes and catalyze the hydrolysis of carboxyl estersofthreedifferentclassesofagrochemicals, pyrethroids, organophosphates and carbamates via the addition of water (Wheelocket al. 2005). The selective toxicity of xenobiotics is affected by the ratio and the levels of metabolizing enzymes which fluctuate in different insect species and alsoinindividualorganism.Thelevelofenzymes could be affected also by the season, the study on winterhoneybeesdemonstratedreducedlevelsof P450-mediateddetoxificationsincethesynergism between pyrethroid deltamethrin and the P450inhibitingfungicideprochlorazwasmuchreduced during winter periods (Meled et al. 1998). By far the majority of the studies of pesticide synergism in honeybees have focused to P450 enzymes that are inhibited by specific pesticides mostly by monitoring the toxicity calculation of the LD50 or LC50 (Table 1.).The developers of insecticide synergists have often exploited inhibition of P450s activity to prolong the efficacy of pesticideswhichareotherwiserapidlydetoxified. ItwasshownthatP450-inhibitorselevatedtoxicity of pyrethroids (cyfluthrin, permethrinand tau-fluvalinate), neonikotinoidinsecticides(imidacloprid, acetadimiprid, thiacloprid), and carbamate insecticide carbaryl (Georghiou and Atkins Jr. 1964, Yu et al. 1984, Hagler et al. 1989, Iwasa et al. 2004, Johnson et al. 2006). It was also found that the classic P450 inhibitors PBO synergize with varroacides tau-fluvalinate and coumaphos at high levels but other inhibitors have minor effect (Johnson et al., 2009, Johnson et al., 2013). Many examples of synergy have been reported between EBI (ergosterol biosynthesis inhibitor) fungicides such as prochloraz, propiconazole, epoxiconazole, carbendazimand insecticides due to the fungicide inhibitory action on P450s. This was the case with neonicotinoids (acetamiprid, thiacloprid, imidacloprid) and pyrethroid insecticides (deltamethrin, lambda-cyhalothrin, alphacypermethrin) (Pilling 1992, Meled et al. 1998, Vandame and Belzunces 1998a, Vandame and Belzunces 1998b, Papaefthimiou and Theo- philidis 2001, Thompson and Wilkins 2003, Schmucketal.2003,Iwasaetal.2004,Thompson 2013).TheeffectsofEBIfungicidesonthecontact toxicity of the active ingredients of the pyrethroid varroacides flumethrin and tau-fluvalinate are synergized by the fungicides with relatively high increases in toxicities (Thompson and Wilkins 2003). Another EBI fungicide prochloraz which is also a P450s inhibitor elevated the toxicity of the acaricides coumaphos and fenpyroximate Acta Biologica Slovenica, 56 (1), 2013 (Johnson et al. 2013). The studies on synergism between insecticidesin honeybees were rarely conducted, mostof thembetween in-hiveacaricides. Johnson et al. (2009) observed a large increase in the toxicity of tau-fluvalinate to bees that had been treated previously with coumaphos, and a moderate increase in the toxicity of coumpahos in bees treated previously with tau-fluvalinate. These compounds were chosen due to their low toxicity to honey bees which were attributed to rapid detoxification mediated by P450s. The synergisms occurred also between in-hive miticides coumaphos, thymol, amitraz, fenpyroximate and oxalic acid (Johnson et al. 2013). The observed synergism was explained as a result of competition between miticides for access to detoxicative P450s (Johnson et al. 2009).See the Table 1. for the list of synergisms of xenobiotics observed in honeybee. Thesynergismswerefoundalsoforcarbamate insecticides (carbaryl, carbofuran) and herbicide atrazine but the mechanism of this synergy is unknown (Sonnet et al. 1978). The synergy be- tweenmonoterpenoidthymolandtau-fluvalinate orcoumaphoswasobservedandwasexplainedto betheconsequenceoftheP450sinhibitoryactivity of thymol, but thymol inhibitory property was shown only in human liver microsomes but not for honeybee midgut (Johnson et al. 2013). In other organisms, the synergisms were studied between insecticides and insecticide/ herbicideatthelevelofdetoxifyingenzymes.The interactionssuchasacompetitionwithmetabolic enzymes esterases are possible that are maybe not very significant for honeybees since it was shown that the role of these enzymes participating in the detoxification of xenobiotics is minor. It was also shown that certain organophosphate insecticides could bind to the active site associated with esterase enzymes responsible for detoxification of pyrethroid-based insecticides and so organophosphate insecticides may beconsidered usefulsynergistsforpyrethroids(Cloyd2011).The synergisms at the level of detoxifying enzymes was described also for organophosphates and pyretroids, P450 activated by organophosphates decrease the organism’s ability to detoxify pyretroids due to esterases inhibition, so greater than additive toxicity is often observed (Hernández et al. 2013). Recent studies have demonstrated the potentiatingeffectsoftriazineherbicides,suchas atrazinetothetoxicityoforganophosphateswhen these herbicides stimulate P450 activity by increas ing the rate of bioactivation of organophosphates resulting in the potentiation of the cholinesterase inhibiting property of organophosphates (Hernández et al. 2013). It seems that the regulation of the P450s in honeybeesisunique.Contrarytootherinsects,in the honey bee these enzymes are rarely induced by a substrate itself. The honeybee genome has substantially fewer protein coding genesfor xenobioticdetoxifying enzymes than Drosophila melanogaster and Anopheles gambiae (Claudianosetal. 2006)andmanyresearchersfailedtodemonstrate anincreaseofmidgutdetoxifyingenzymesinduced byxenobiotics(Yuetal.1984).Evenexposureto phenobarbitalwhichisaninducerofP450sshowed no alterations in the expression of many of P450 genestestedinhoneybees(Maoetal.2011).Only twostudiesindicatedtheincreaseinP450activity in honeybees. Application of tau-fluvalinate and coumaphos elevated specific detoxifying P450 enzymes CYP9Q1, CYP9Q2, and CYP9Q3 and benzo(a)pyrenemonooxidaseactivityinhoneybee guts was induced by exposure to benzo(a)pyrene itselfandbythein-hiveacaricidestau-fluvalinate and cymiazole hydrochloride (Kezic et al. 1992, Mao et al. 2011). As it has been at first suggested that reduced diversity of detoxification enzymes may contribute to the sensitivity of honey bees to certain pesticides (Claudianos et al. 2006) the importanceofmidgutdetoxifyingenzymesP450 in honeybees was highlighted by the studies with specificP450-inhibitors.Twostudiesindicatethat GSTsandCaEsareactivedetoxifyingenzymesin honeybeesbuttheyplayarelativelyminorrolein detoxification as compared to P450s (Johnson et al. 2009, Iwasa et al. 2004). The CaEs inhibitor DEF(S,S,S-tributylphosphorotrithioate)andGSTs inhibitor DEM (diethyl maleate) were shown to increase the toxicity of certain pyrethroids and neonicotinoids but this effect was significantly smallerthatfortheP450inhibitorPBO(piperonyl butoxide) (Iwasa et al. 2004). Recently the interesting study performed by Johnson et al. (2012) suggested that regulation of honey bee P450s is affected by chemicals occurring naturally in the hive environment in the nectar, pollen and propolis since only quercetin, a common pollen Glavan et al.: The synergy of xenobiotics in honey bee Apis mellifera: mechanisms and effects Table 1: The list of synergisms of xenobiotics in honeybee Apis mellifera and proposed mechanisms. Tabela 1:Seznam sinergizmov med ksenobiotiki v medonosni cebeli Apis mellifera in predlagani mehanizmi. Xenobiotic Xenobiotic (P450 inhibitor) Reference Mechanism of synergy: inhibition of P450 detoxifying enzymes pyrethroid insecticides classical P450-inhibitor cyfluthrin piperonyl butoxide (Johnson et al. 2006) permethrin piperonyl butoxide (Hagler et al. 1989) lambda-cyhalothrin piperonyl butoxide (Johnson et al. 2006) tau-fluvalinate piperonyl butoxide neonicotinoid insecticides classical P450-inhibitor (Johnson et al. 2006; Johnson et al. 2013) imidacloprid piperonyl butoxide (Iwasa et al., 2004, Johnson et al. 2012) acetadimiprid piperonyl butoxide (Iwasa et al. 2004) thiacloprid carbamate insecticide piperonyl butoxide classical P450-inhibitor (Iwasa et al. 2004) carbaryl piperonyl butoxide (Georghiou and Atkins Jr. 1964) hive varroacides classical P450-inhibitor tau-fluvalinate piperonyl butoxide (Johnson et al. 2009, Johnson et al. 2013) coumaphos piperonyl butoxide (Johnson et al. 2009, Johnson et al. 2013) fenpyroximate piperonyl butoxide (Johnson et al., 2013) neonicotinoid insecticides EBI (ergosterol biosynthesis inhibitor) fungicides acetamiprid epoxiconazole, propiconazole, triadimefon, triflumizole, uniconazole-P (Iwasa et al. 2004) thiacloprid prochloraz, propiconazole, tebuconazole, triflumizole (Schmuck et al. 2003, Iwasa et al. 2004) imidacloprid propiconazole, triflumizole (Iwasa et al. 2004) pyrethroid insecticides EBI (ergosterol biosynthesis inhibitor) fungicides deltamethrin difenoconazole+carbendazim, prochloraz, prochloraz+ difenoconazole 850 (Belzunces and Colin 1993, Colin and Belzunces 1992, Papaefthimiou and Theophilidis 2001, Vandame and Belzunces1998b, Vandame and Belzunces 1998a) lambda-cyhalothrin difenconazole, flusilazole, prochloraz, propiconazole, tebuconazole, thiophanate-methyl (Thompson and Wilkins 2003) alphacypermethrin difenconazole, flusilazole, prochloraz, propiconazole, tebuconazole (Thompson and Wilkins 2003) hive varroacides EBI (ergosterol biosynthesis inhibitor) fungicides coumaphos prochloraz (Johnson et al. 2013) flumethrin carbendazim, difenconazole, (Thompson and Wilkins 2003) flusilazole, prochloraz, propiconazole, tebuconazole, thiophanate-methyl 20 Acta Biologica Slovenica, 56 (1), 2013 Xenobiotic tau-fluvalinate fenpyroximate Xenobiotic (P450 inhibitor) carbendazim, difenconazole, flusilazole, prochloraz, propiconazole, tebuconazole, thiophanate-methyl, myclobutanil, metconazole, fenbuconazole, prochloraz Reference (Thompson and Wilkins 2003, Johnson et al. 2013 ) (Johnson et al. 2013) hive varroacides hive varroacides coumaphos tau-fluvalinate (Johnson et al. 2009, 2013) thymol tau-fluvalinate, coumaphos (Johnson et al. 2013) amitraz tau-fluvalinate, coumaphos, (Johnson et al. 2013) fenpyroximate fenpyroximate tau-fluvalinate, coumaphos (Johnson et al. 2013) Mechanism of synergy: increased oxidative stress hive varroacides Fungicides (mitochondrial inhibitors) tau-fluvalinate pyraclostrobin, boscalid (Johnson et al. 2013) fenpyroximate pyraclostrobin (Johnson et al. 2013) Unknown mechanism of synergy oxalic acid tau-fluvalinate, fenpyroximate, (Johnson et al. 2013) amitraz, thymol herbicide atrazine carbamate insecticides (carbaryl, (Sonnet et al. 1978) carbofuran) thio and coumaphos varroacide (Lienau 1990) dithiophosphoric ester pesticides – ethyl parathion, dimethoate, dialifos thiacloprid fungicides cyprodinil, tolyfluanid (Schmuck et al. 2003) (neonicotionoid) alphacypermethrin, fungicide chlorothalonil (Thompson and Wilkins lambda-cyhalothrin fungicide chlorothalonil 2003) and honey constituent, reduced tau-fluvalinate toxicity. Bees fed with extracts of honey, pollen andpropolisshowedelevatedexpressionofthree CYP6AS P450 genes. Non-naturally occurring inducers of cytochrome P450 enzymes did not alter the toxicity of certain xenobiotics and it seems that a wide range of synthetic pesticides do not induce in bees. It is now clearthat certain substancesfoundinbeeproductssuchasquercetin, p-coumaric acid, pinocembrin, and pinobanksin 5-methyl ether naturally elevate the levels of bee detoxifying enzymes P450 and probably helping bees to resist the toxicity of certain xenobiotics (Johnson et al. 2012, Mao et al. 2013). The synergism of xenobiotics working at the same targets Although the basic molecular mechanisms of most xenobiotics are more or less known, the possible mechanisms of their synergy at primary target sites in honeybees are unexplored. One of plausible mechanism of this synergy is that effects at the site of toxic action include increased responseofthesite(suchasareceptor)following initial pesticide exposure and according to this direct synergistic effect could be predicted for sub- stancesthathavesimilartargets(Thomson1996). In this respect only one study in honeybees was performed,onsemi-isolatedheart(Papaefthimiou and Theophilidis 2001).In this study the synergistic effect was observed between EBI fungicide Glavan et al.: The synergy of xenobiotics in honey bee Apis mellifera: mechanisms and effects prochlorazandpyrethroidinsecticidedeltamethrin which rapidly decreased the frequency and the force of the cardiac contractions with marked effects at 0.01 µM, equivalent to internal doses of 4–5 µg kg-1 body weight. Prochloraz showed to be more cardiotoxic than deltamethrin, what seemed surprising since deltamethrin is a neurotoxicsubstancewhereasprochlorazisaninhibitor andaninducerofdetoxifyingenzymesP450.So, authors concluded that there must be the neural basis of the deltamethrin prochloraz synergy. Belzunces et al. (2012) suggested that the basis of their synergy is the interaction of these two pesticides with shared molecular targets, such as ATPases, potassium and calcium channels. The existence of the synergy of insecticides at primary target sites was demonstrated also in the cockroach P. americana,inthecercal-afferent giant-interneuron synapses of the terminal abdominal ganglion (Corbel et al. 2006). Authors demonstrated that pyrethroid permethrin and carbamatepropoxurinsecticidesappliedtogether increaseddrasticallytheAChconcentrationwithin the synaptic cleft, which thereby stimulated a negative feedback of ACh release mediated by presynaptic muscarinic receptors causing the synergism. Johnsonetal.(2013)demonstrated5-fold increaseinthetoxicityoftau-fluvalinatebyamitraz pretreatment in honeybees. Interactions between formamidinesandpyrethroidsareknowninother insects and may be due to synergism at the target site through cooperative binding (Liu and Plapp 1992).It was shown that formamidine pesticides working as octopaminergic agonists change the binding properties of pyrethroid insecticides to nervemembranesodiumchannels.Thismechanism could be also the cause for the synergism of tau- fluvalinate and amitraz in honeybees. Thomson (1996) reported another possible mechanism of pesticide interaction, when esterases can act as irreversible binding sites for organophosphate and carbamate insecticides, reducing the levels available to bind to AChE within the brain. Thus, prior exposure to an organophosphate may result in a reduction in the number of available binding sites and an increase in the blood levels of free pesticide.Allthesemechanismsmentionedabove are possible for honeybees, but they are not explored. The lack of similar studies suggests that the mechanisms of synergisms of xenobiotics at primary target sites in honeybees are very much ignored and underestimated and a need for additional studies is unavoidable. Conclusions Bees are very often exposed to mixtures of products applied to plants on which they forage suchasfungicides,herbicidesandinsecticidesand inadditionveryhighlevelsofvarroacidesmaybe present within colonies. The very potential risk from most mixtures of these substances is the developmentofsynergismsthatcanprofoundlyaffect honeybeecoloniesandmaysignificantlycontribute to honeybee colony loss observed in the last 15 years. This risk is underestimated and relatively few relevant studies were performed concerning the effects and mechanisms of synergy of different xenobiotic combinations. The understanding the mechanisms of synergy betweenxenobiotics is very important for the restriction of the use of defined mixtures and also for the prediction of potential toxicity of newly developed substances in agriculture and apiculture. Many observable effects are induced by xenobiotics such as alternation of cognitive functions, behavior or integrity of physiological functions, many of them unambiguouslyexplainedbythemechanismsofxenobiotic actions at primary target sites. In spite of these physiologicalmechanismsofactionofindividual xenobioticsaremoreorlessidentified,especially forinsecticides,themajorityofsynergisticeffects observed in honeybees is ascribed to the inhibition of detoxifyingmidgutenzymes P450 involvedin xenobioticmetabolism.Evenmore,asmostofthe studies focused on synergistic effect of mixure xenobiotic/ P450inhibitor,onlyfewwereperformed oninsecticide/insecticideinteractions.Johnsonet al. (2013) proposed that the synergistic interactions occur when the compounds work through different modes of action, but few experiments in insects studying the synergism of insecticides at target sites suggest that the synergism is also possibleforsubstancesworkingthroughthesame mode of action, at least when they are working on the same system such as cholinergic synapse. Therefore,theaspectofmechanismsofsynergism at the similar targets is underestimated since only one study was performed in honeybee and Acta Biologica Slovenica, 56 (1), 2013 therefore this issue demands extra investigation. The improved knowledge of the mechanisms of pesticide and bee-hive compound interactions would prevent the negative impact on beneficial organisms like honeybees. Povzetek Medonosne cebele so med iskanjem hrane pogosto izpostavljene razlicnim ksenobiotikom, vecinoma so to fitofarmacevtska sredstva in panjske kemikalije. Cebele v zadnjih 15 letih množicno umirajo, vzrok za to naj bi bila tudi uporabaksenobiotikov.Številnafitofarmacevtska sredstva se uporablja hkrati in znano je, da lahko pride do sinergisticnih interakcij v organizmih. Tveganjezanastaneksinergizmovjepodcenjeno in narejenih je relativno malo študij na cebelah o ucinkih in mehanizmih sinergizmov razlicnih kombinacij ksenobiotikov. Razumevanje mehanizmovsinergizmovksenobiotikovjezelopomembno za nadzor nad uporabo definiranih mešanic in napovedovanje potencialne toksicnosti novih ksenobiotikovvkmetijstvuincebelarstvu. Pregledni clanek se osredotoca na ucinke, mehanizme in molekulske tarce ksenobiotikov v medonosnih cebelahinosvetljujeprimeremehanizmovnastanka sinergizmov. Opisani so tudi drugi dejavniki,ki vplivajonanjihovnastanek,okoljskiinfiziološki, poudarek je na detoksifikacijskih encimih me- donosne cebele. Najbolj nevarni za cebele so insekticidi,kidelujejopredvsemnanekajrazlicnih živcnihmolekularnihtarcintakomotijodelovanje References živcnega sistema, kar vpliva na vedenje, kognitivne funkcije in fiziologijo cebel. Glavne živcne tarceinsekticidovsonapetostnoodvisninatrijevi kanalcki, encim acetilholinesteraza, glutamatni receptorji, receptorji za gama-aminomasleno kislino in nikotinski receptorji. Znane skupine insekticidov so piretroidi, DDT, DDT podobni insekticidi, organofosfati, karbamati, fenilpirazolni pesticidi ter neonikotinoidi. Kljub temu, da so tarce delovanja posameznih ksenobiotikov,še posebej insekticidov, precej znani, raziskovalci vecino sinergijskih ucinkov v cebelah razlagajo zinhibicijocrevesnihdetoksifikacijskihencimov P450, ki presnavljajo ksenobiotike. Vecina študij je bila namrec narejena z mešanicami ksenobiotik/ zaviralecencimovP450.Glavnizaviralciencimov P450 sospecificni inhibitorji,ki jihdodajajofitofarmacevtskimsredstvomzapodaljšanjeucinkater nekaterifungicidi.Študijnacebelah,skaterimibi raziskovalisinergizemmedinsekticidi,skorajni. Ceprav so sinergizmi ksenobiotikov, še posebej insekticidov,naprimarnihciljnihtarcahpricebelah možni,sajsobiliprikazanipridrugihorganizmih, je ta vidik podcenjen. Narejena je bila samo ena raziskava mehanizmov na tarcnem mestu pri medonosni cebeli, pa še to med insekticidom in fungicidom. Acknowledgements The work was financially supported by the Slovenian Research Agency (Grant No. V1-1129). Barron, A.B., Maleszka, R., Vander Meer, R.K., Robinson, G.E., 2007. Octopamine modulates honey bee dance behavior. PNAS, 104 (5), 1703–1707. Belzunces, L.P., Tchamitchian, S., Brunet, J.-L., 2012. Neural effects of insecticides in the honey bee. Apidologie, 43 (3), 348–370. Bergougnoux, M., Treilho, U.M., Armengaud, C., 2013. Exposure to thymol decreased phototactic behavior in the honeybee (Apis mellifera) in laboratory conditions. Apidologie, 44(1), 88–89. Bevk, D., Kralj, J., Cokl, A., 2012. Coumaphos affects food transfer between workers of honeybee Apis mellifera. Apidologie 43, 465–470. Bicker, G., Schafer, S., Kingan, T.G., 1985. Mushroom body feedback interneurones in the honey bee show GABA-like immunoreactivity.Brain. Res., 360 (1-2), 394–397. Blenau,W.,Rademacher,E.,Baumann,A.,2011.Plantessentialoilsandformamidinesasinsecticides/ acaricides: what are the molecular targets? Apidologie, 43 (3), 334–347. Glavan et al.: The synergy of xenobiotics in honey bee Apis mellifera: mechanisms and effects Bloomquist, J.R., 2003. Chloride Channels as Tools for Developing Selective Insecticides. Archives of Insect Biochemistry and Physiology, 54 (4), 145–156. Boncristiani, H., Underwood, R., Schwarz, R., Evans, J.D., Pettis, J., vanEngelsdorp, D., 2012. Direct effect of acaricides on pathogen loads and gene expression levels in honey bees Apis mellifera. Journal of Insect Physiology, 58 (5), 613–620. Burley, L., Fell, R., Saacke, R., 2008. Survival of honey bee (Hymenoptera: Apidae) spermatozoa incubated at room temperature from drones exposed to miticides. J. Econ. Entomol., 101 (4), 1081–1087. Cao, L.C., Honeyman, T.W., Cooney, R., Kennington, L., Scheid, C.R., Jonassen, J.A., 2004. Mitochondrial dysfunction is a primary event in renal cell oxalate toxicity. Kidney Int., 66 (5), 1890–1900. Casida, J.E., 2009. Pest toxicology: the primary mechanisms of pesticide action. Chem Res Toxicol., 22 (4), 609–619. Claudianos, C., Ranson, H., Johnson, R.M., Biswas, S., Schuler, M.A., Berenbaum, M.R., Feyereisen, R., and Oakeshott, J.G., 2006. Adeficit of detoxification enzymes: pesticide sensitivity and environmental response in the honeybee. Insect Mol. Biol., 15 (5), 615–636. Cloyd, R.A., 2011. Pesticide Mixtures. In: Stoytcheva, M. (ed.): Pesticides-Formulations, Effects, Fate,InTech, pp. 69–80. Coats, J.R., 1990. Mechanisms of toxic action and structure-activity relationships for organochlorine and synthetic pyrethroid insecticides. Environ. Health Perspect., 87, 255–262. Cole,L.M.,Roush,R.T.,Casida,J.E.,1993.ActionofphenylpyrazoleinsecticidesattheGABA-gated chloride channel. Pestic. Biochem. Physiol., 46 (1), 47–54. Collins, A.M., Pettis, J.S., Wilbanks, R., Feldlaufer, M.F., 2004. Performance of honey bee (Apis mellifera) queens reared in beeswax cells impregnated with coumaphos. J. Apic. Res., 43 (3), 128–134. Corbel, V., Stankiewicz, M., Bonnet, J., Grolleau, F., Hougard, J.M., Lapied, B., 2006. Synergism betweeninsecticidespermethrinandpropoxuroccursthroughactivationofpresynapticmuscarinic negative feedback of acetylcholine releasein theinsectcentralnervous system. Neurotoxicology., 27 (4), 508–519. Cully, D.F., Wilkinson, H., Vassilatis, D.K., Etter, A., Arena, J.P., 1996. Molecular biology and electrophysiology of glutamate-gated chloride channels of invertebrates. Parasitol., 113 (S1), S191–S200. Davies,T.G.E.,Field,L.M.,Usherwood,P.N.R.,Williamson,M.S.,2007.DDT,pyrethrins,pyrethroids and insect sodium channels. IUBMB-Life, 59 (3), 151–162. Desneux, N., Decourtye, A., Delpuech, J.-M., 2007. The sublethal effects of pesticides on beneficial arthropods. Annu. Rev. Entomol., 52, 81–106. Dudai, Y., Buxbaum, J., Corfas, G., Ofarim M., 1987. Formamidines interact with Drosophila octopamine receptors, alter the flies’behaviour and reduce their learning ability. J. Comp. Physiol., 161 (5), 739–746. El Hassani, A.K., Dupuis, J.P., Gauthier, M., Armengaud, C., 2009. Glutamatergic and GABAergic effects of fipronil on olfactory learning and memory in the honeybee. Invertebrate Neuroscience, 9 (2), 91–100. Ellis, M.D., Baxendale, F.P., 1997. Toxicity of seven monoterpenoids to tracheal mites(Acari: Tarsonemidae) and their honey bee (Hymenoptera: Apidae) hosts when applied as Fumigants. J. Econ. Entomol., 90 (5), 1087–1091. Elzen, P.J., Westervelt, D., 2002. Detection of coumaphos resistance in Varroa destructor in Florida. Am. Bee J., 142 (4), 291–292. Enan, E., 2001. Insecticidal activity of essential oils: octopaminergic sites of action. Comparative Biochemistry and Physiology Part C, Toxicol. Pharmacol., 130 (3), 325–337. Acta Biologica Slovenica, 56 (1), 2013 Enayati,A.A., Ranson, H., Hemingway, J., 2005. Insect glutathione transferases and insecticide resistance. Insect Molecular Biology, 14 (1), 3–8. Evans, P.D., Gee, J.D., 1980. Action of formamidine pesticides on octopamine receptors. Nature, 287 (5777), 60–62. Floris, I., Satta, A., Cabras, P., Garau, V.L., Angioni, A., 2004. Comparison between two thymol formulations in the control of Varroa destructor: effectiveness, persistence, and residues. J. Econ. Entomol., 97 (2), 187–191. Food and Agriculture Organization of the United Nations (FAO), 2009. FAOSTAT.http://faostat.fao.org Fries, I., 1991. Treatment of sealed honey bee brood with formic acid for control of Varroa jacobsoni. Am. Bee J., 131 (5), 313–314. Fukuto,T.R., 1990. Mechanism of action of organophosphorus and carbamate insecticides. Environ. Health Perspect., 87, 245–254. Georghiou, G.P., Atkins Jr. E.L., 1964. Temperature coefficients of toxicity of certain n-methyl carbamates against honeybees, and the effect of the synergist piperonyl butoxide. Journal of Apicultural Research, 3, 31–35. Gregorc, A., Bowen, I., 2000. Histochemical characterization of cell death in honeybee larvae midgut after treatment with Paenibacillus larvae, amitraz and oxytetracycline. Cell. Biol. Int., 24 (5), 319–324. Gregorc, A., Smodisskerl, M.I., 2007. Toxicological and immunohistochemical testing of honeybees after oxalic acid and rotenone treatments.Apidologie, 38 (3), 296–305. Gunasekara, A.S., Truong, T., Goh, K.S., Suprlock, F., Tjeerdema, R.S., 2007. Environmental fate and toxicology of fipronil. J. Pestic. Sci., 32 (3), 189–199. Haarmann, T., Spivak, M.,Weaver, D.,Weaver, B., Glenn, T., 2002. Effects of fluvalinate and coumaphosonqueenhoneybees( Hymenoptera:Apidae)intwocommercialqueenrearingoperations. J. Econ. Entomol., 95 (1), 28–35. Hagler, J.R., Waller, G.D., Lewis, B.E., 1989. Mortality of honeybees Hymenoptera Apidae exposed to permethrin and combinations of permethrin with piperonyl butoxide. Journal of Apicultural Research, 28 (4), 208–211. Hernández, A.F., Parrón, T., Tsatsakis, A.M., Requena, M., Alarcón, R., López-Guarnido, O., 2013. Toxic effects of pesticide mixtures at a molecular level: Their relevance to human health. Toxicology, 307, 136–145. Higes, M., Meana, A., Suarez, M., Llorente, J., 1999. Negative long-term effects on bee colonies treated with oxalic acid against Varroa jacobsoni Oud.. Apidologie, 30 (4), 289–292. Homberg, U. 1994: Distribution of Neurotransmitters in the Insect Brain. Progress in Zoology, Vol. 40, Fischer, Stuttgart, 90 pp. Iwasa, T., Motoyama, N., Ambrose, J.T., Roe, R.M., 2004. Mechanism for the differential toxicity of neonicotinoid insecticides in the honey bee, Apis mellifera. Crop Protection, 23 (5), 371–378. Jeschke, P., Nauen,R., 2008. Neonicotinoids – from zero to hero in insecticide chemistry.Pest. Manag. Sci., 64 (11), 1084–1098. Johnson,R.M.,Wen,Z.,Schuler,M.A.,Berenbaum,M.R.,2006.MediationofPyrethroidInsecticide Toxicity to Honey Bees (Hymenoptera: Apidae) by Cytochrome P450 Monooxygenases. Journal of Economic Entomology, 99(4), 1046–1050. Johnson, R.M., Pollock, H.S., Berenbaum, M.R., 2009. Synergistic interactions between in-hive miticides in Apis mellifera. J. Econ. Entomol., 102 (2), 474–479. Johnson, R.M., Ellis, M.D., Mullin, C.A., Frazier, M., 2010. Pesticides and honey bee toxicity – USA. Apidologie, 41 (3), 312–331. Johnson,R.M., Mao,W., Pollock,H.S., Niu,G., Schuler,M. A., Berenbaum,M. R., 2012. Ecologically Appropriate Xenobiotics Induce Cytochrome P450s in Apis mellifera. PLoS ONE, 7 (2), p.e31051. Glavan et al.: The synergy of xenobiotics in honey bee Apis mellifera: mechanisms and effects Johnson, R.M., Dahlgren, L., Siegfried, B.D., Ellis, M.D., 2013. Acaricide, Fungicide and Drug Interactions in Honey Bees (Apis mellifera). PLoS ONE, 8 (1), e54092. Jones, A.K., Sattelle,D.B., 2010. Diversity of Insect Nicotinic Acetylcholine Receptor Subunits: in Insect Nicotinic Acetylcholine Receptors. Advances in Experimental Medicine and Biology, 683, 25–43. Keyhani, J., Keyhani, E., 1980. EPR study of the effect of formate on cytochrome oxidase. Biochem. Biophys. Res. Comm., 92 (1), 327–333. Kezic, N., Lucic, D., Sulimanovic, D., 1992. Induction of mixed function oxidase activity in honey bee as a bioassay for detection of environmental xenobiotics. Apidologie, 23 (3), 217–223. Kral, K., 1980. Acetylcholinesterase in the ocellus of Apis mellifica. J. Insect Physiol. 26 (12), 807–809. Kral, K., Schneider, L., 1981. Fine structural localization of acetylcholinesterase activity in the compound eye of the honeybee (Apis mellifica L.). Cell Tissue Res., 221 (2), 351–359. Kreissl, S., Bicker, G., 1989. Histochemistry of acetylcholinesterase and immunocytochemistry of an acetylcholinereceptor-like antigen in thebrain of the honeybee. J. Comp. Neurol., 286 (1), 71–84. Kucharski, R., Ball, E.E., Hayward, D.C., Maleszka, R., 2000. Molecular cloning and expression analysis of a cDNA encoding a glutamate transporter in the honeybee brain. Gene, 244 (1–2), 399–405. Lienau , F.W., 1990. Effect of varroacide and pesticide treatment on honeybees. Apidologie, 21 (4), 375–377. Liu, M.Y., Plapp, F.W. Jr, 1992. Mechanism of formamidine synergism of pyrethroids. Pestic. Bio- chem. Physiol., 43(2), 134–140. Lodesani, M., Colombo, M., Spreafico, M., 1995. Ineffectiveness of Apistan® treatment against the mite Varroa jacobsoni Oud. in several districts of Lombardy (Italy). Apidologie, 26 (1), 67–72. Locatelli, F., Bundrock, G., Müller, U., 2005. Focal and temporal release of glutamate in the mushroom bodies improves olfactory memory in Apis mellifera. Journal of Neuroscience, 25 (50), 11614–11618. Maleszka,R.,Helliwell,P.,Kucharski,R.,2000.Pharmacologicalinterferencewithglutamatere-uptake impairs long-term memory in the honeybee Apis mellifera. Behav. Brain. Res., 115 (1), 49–53. Mao, W., Rupasinghe, S.G., Johnson, R.M., Zangerl, A.R., Schuler, M.A., Berenbaum, M.R., 2009. Quercetin-metabolizingCYP6ASenzymesofthepollinatorApismellifera(Hymenoptera:Apidae). Comp. Biochem. Physiol. B Biochem. Mo.l Biol., 154 (4), 427–434. Mao, W., Schuler, M., Berenbaum, M., 2011. CYP9Q-mediated detoxification of acaricides in the honey bee (Apis mellifera). PNAS, 108 (31), 12657–12662. Mao,W., Schuler,M.A., Berenbaum,M.R., 2013. Honey constituents up-regulate detoxification and immunity genes in the western honey bee Apis mellifera. PNAS, doi/10.1073/pnas.1303884110. Marchetti, S., Barbattini, R., 1984. Comparative effectiveness of treatments used to control Varroa jacobsoni Oud.. Apidologie, 15 (4), 363–378. Massoulie, L., Pezzementi, L., Bon, S., Krejci, E., Vallette, F.M.,1993. Molecular and cellular biology of cholinesterases. Prog. Neurobiol., 41 (1), 31–91. Meimaridou, E., Jacobson, J., Seddon, A.M., Noronha-Dutra, A.A., Robertson, W.G., Hothersal, J.S., 2005. Crystal and microparticle effects on MDCK cell superoxide production: oxalate- specific mitochondrial membrane potential changes. Free Radical Biology and Medicine, 38 (12), 1553–1564. Meled,M., Thrasyvoulou,A., Belzunces,L.P., 1998. Seasonal variations in susceptibility of Apis mellifera to the synergistic action of prochloraz and deltamethrin. Environmental Toxicology and Chemistry, 17 (12), 2517–2520. Meyer, E.P., Matute, C., Streit, P., Nässel, D.R., 1986. Insect optic lobe neurons identifiable with monoclonal antibodies to GABA. Histochemistry, 84 (3), 207–216. Acta Biologica Slovenica, 56 (1), 2013 Motoba, K., Suzuki, T., Uchida, M., 1992. Effect of a new acaricide, fenpyroximate, on energy- metabolism and mitochondrial morphology in adult female Tetranychus urticae (two-spotted spider mite). Pestic. Biochem. Physiol, 43 (1), 37–44. Narahashi, T., 1992. Nerve membrane Na+ channels as targets of insecticides. Trends. Pharmacol. Sci., 13 (6), 236–241. Ortiz de Motellano, P.R., De Voss, J.J., 2005: Substrate oxidation by cytochrome P450 enzymes. In: OrtizdeMotellano,P.R.(ed.):CytochromeP450,structure,mechanism,andbiochemistry,Kluwer Academic/Plenum Publishers, New York. pp. 183–245. Papaefthimiou, C., Theophilidis, G., 2001. The cardiotoxic action of the pyrethroid insecticide delta- methrin, the azole fungicide prochloraz and their synergy on the semi-isolated heart of the bee Apis mellifera macedonica. Pesticide Biochemistry and Physiology, 69 (2), 77–91. Pettis,J.S.,Collins,A.M.,Wilbanks,R.,Feldlaufer,M.F.,2004.Effectsofcoumaphosonqueenrearing in the honey bee, Apis mellifera. Apidologie, 35 (6), 605–610. Pilling, E.D., 1992. Evidence for pesticide synergism is the honeybee (Apis mellifera). Aspects of Applied Biology, 31 (1), 43–47. Priestley,C.M.,Williamson,E.M.,Wafford,K.A.,Sattelle,D.B.,2003.Thymol,aconstituentofthyme essentialoil,isapositiveallostericmodulatorofhumanGABA(A)receptorsandahomooligomeric GABA receptor from Drosophila melanogaster. British J. Pharmacol., 140 (8), 1363–1372. Ray, D.E., Fry, J.R., 2006. Areassessment of the neurotoxicity of pyrethroid insecticides. Pharmacol. Ther., 111 (1), 174–93. Raymond-Delpech, V., Matsuda, K., Sattelle, B.M., Rauh, J.J., Sattelle, D.B., 2005. Ion channels: molecular targets of neuroactive insecticides. Invertebr. Neurosci., 5 (3–4), 119–133. Rinderer, T.E., De Guzman, L.I., Lancaster, V.A., Delatte, G.T., Stelzer, J.A., 1999. Varroa in the mating yard: I. The effects of Varroa jacobsoni and Apistan® on drone honey bees. Am. Bee J., 139 (3), 134–139. Sattelle, D.B., 1980. Acetylcholine receptors of insects. Adv. Insect Physiol., 15, 215–315. Sattelle,D.B.,Yamamoto,D.,1988.Moleculartargetsofpyrethroidinsecticides.Adv.InsectPhysiol., 20, 147–213. Schafer, S., Bicker, G., 1986. Distribution of GABA-like immunoreactivity in the brain and suboesophageal ganglion of the honeybee. J. Comp. Neurol., 246 (3), 287–300. Scheidler, A., Kaulen, P., Bruning, G.,Erber, J.,1990. Quantitative autoradiographic localization of [125I]alpha-bungarotoxin binding sites in the honeybee brain. Brain. Res., 534 (1–2), 332–335. Schulz, D.J., Robinson, G.E., 2001. Octopamine influences division of labor in honey bee colonies. J. Compar. Physiol. A, 187 (1), 53–61. Scott, J.G., Wen Z.,2001. Cytochromes P450 of insects: the tip of the iceberg. Pest. Manag. Sci., 57 (10), 958–967. Sherer,T.B.,Richardson,J.R.,Testa,C.M.,Seo,B.B.,Panov,A.V.,Yagi,T.,Matsuno-Magi,A.,Miller, G.W., Greenamyre, J.T., 2007. Mechanism of toxicity of pesticides acting at complex I: relevance to environmental etiologies of Parkinson’s disease. J. Neurochem., 100 (6), 1469–1479. Schmuck, R., Stadler, T., Schmidt, H.W., 2003. Field relevance of a synergistic effect observed in the laboratory between an EBI fungicide and a chloronicotinyl insecticide in the honeybee (Apis mellifera L, Hymenoptera). Pest Management Science, 59 (3), 279–286. Sonnet, P.E., Lye, T.L., Sackett, R.R., 1978. Effects of selected herbicides on the toxicity of several insecticides to honey bees. Environ. Entomol., 7 (2), 254–256. Song, C., Scharf, M.E., 2008. Formic acid: Aneurologically active, hydrolyzed metabolite of insecticidal formate esters. Pestic. Biochem. Physiol., 92 (2), 77–82. Tasei, J.N., 2001. Effects of insect growth regulators on honey bees and non-apis bees. Apidologie, 32 (6), 527–545. Tennekes,H.A.,Sánchez-Bayo,F.,2013.Themolecularbasisofsimplerelationshipsbetweenexposure concentration andtoxic effects with time. Toxicology, 309, 39–51. Glavan et al.: The synergy of xenobiotics in honey bee Apis mellifera: mechanisms and effects Thany,S.H.,Tricoire-Leignel,H.,Lapied,B.,2010.Identificationofcholinergicsynaptictransmission in the insect nervous system. Adv. Exp. Med. Biol., 683, 1–10. Thompson, H.M., 1996. Interactions between pesticides; a review of reported effects and their implications for wildlife risk assessment. Ecotoxicology, 5 (2), 59–81. Thompson,H.M.,Wilkins,S.,2003.Assessmentofthesynergyandrepellencyofpyrethroid/fungicide mixtures. Bulletin of Insectology, 56 (1), 131–134. Thompson, H.M., Wilkins, S., Battersby, A.H., Waite, R.J., Wilkinson, D., 2005. The effects of four insect growth regulating (IGR) insecticides on honeybee (Apis melliferaL.) colony development queen rearing and drone sperm production. Ecotoxicology, 14 (7), 757–769. Thompson, H.M., 2012. Interaction between pesticides and other factors in effects on bees. http:// www.efsa.europa.eu/en/supporting/doc/340e.pdf Underwood, R., Currie, R., 2003. The effects of temperature and dose of formic acid on treatment efficacy against Varroa destructor (Acari: Varroidae), a parasite of Apis mellifera (Hymenoptera: Apidae). Exp. Appl. Acarol., 29 (3–4), 303–313. Yu,S.,Robinson,F.,Nation,J.,1984.DetoxificationcapacityinthehoneybeeApismelliferaL.Pestic Biochem Physiol, 22 (3), 360–368. Vandame, R., Belzunces, L.P., 1998a. Erratum to “Joint actions of deltamethrin and azole fungicides on honey bee thermoregulation”.Neurosci Lett., 255 (1), 61. Vandame, R., Belzunces, L.P., 1998b. Joint actions of deltamethrin and azole fungicides on honey bee thermoregulation. Neurosci Lett., 251(1), 57–60. vanEngelsdorp, D., Meixner, M.D., 2010. A historical review of managed honey bee populations in Europe and the United States and the factors that may affect them. J. Invertebr. Pathol., 103 (Suppl 1), S80–95. Weinberger, N.M., 2006. Food for Thought: Honeybee Foraging, Memory, and Acetylcholine. Sci. STKE, 336, p. pe23. Wheelock,C.E., Shan, G., Ottea, J., 2005. Overview of carboxylesterases and their role in the metabolism of insecticides. J. Pestic. Sci., 30 (2), 75–83. Whittington, R., Winston, M.L., Melathopoulos, A.P., Higo, H.A., 2000. Evaluation of the botanical oils neem, thymol, and canola sprayed to control Varroa jacobsoni Oud. (Acari: Varroidae) and Acarapsis woodi (Acari: Tarsonemidae) in colonies of honey bees (Apis mellifera L.,Hymenoptera: Apidae). Am. Bee J., 140 (7), 565–572. Wolstenholme, A.J., Rogers, A.T., 2005. Glutamate-gated chloride channels and the mode of action of the avermectin/milbemycin anthelmintics. Parasitology, 131 (Suppl S1), S85–S95. Zlotkin, E., 1999. The insect voltage-gated sodium channel as target of insecticides. Annu. Rev. Entomol., 44, 429–455.