!" ! #  $  %   &' ()*'  )    % #

            +  ,    + + # -        )     +  +   #  ,           ) )     #                #          +       )+         #    +        .         ) +      )+          +      #       +           .             #  +     +   #/ ) +         )              0 )+          1          +   +         +     #2+           &   3#()+    ,   )   +   )    &. ()+          4    #       +          0               # )       +    +   +          )  +  +            #     0     +                     )            5            #              5    #2 0     +   +  0        ,                         #                          #   6 +         ) ,   &7(80! #2        ) ,     +          )          #   )                                 0   #              )    0           #        +                )           )             #

       !" # "!   

   !"  911 ## 1  : ; 99 9 9 !8!4

<-*=>"=!>>=>!4 <-*=>"=!>>=>!4?>

           ! 

    )! ?=!  

BARLEY DEFENSE GENES AGAINST

Aleksandra Losvik

Barley defense genes against aphids

Aleksandra Losvik ©Aleksandra Losvik, Stockholm University 2018

ISBN print 978-91-7797-145-0 ISBN PDF 978-91-7797-146-7

Cover photo by Aleksandra Losvik Printed in Sweden by Universitetsservice US-AB, Stockholm 2018 Distributor: Department of Ecology, Environment and Plant Sciences To my dearest Johan Thank you for the adventures of my lifetime

Sammanfattning

Bladlöss är skadeinsekter av stor betydelse över hela världen. Genom att livnära sig från växternas ledningsrör undandrar de direkt näring från växterna. Skador av angrepp syns ofta som bladinrullning, gula eller svarta bladfläckar eller andra växtdeformationer. Deras roll som skadeinsekter hänger också samman med att de överför virussjukdomar. Bladlusangrepp hanteras inom jordbruket med bekämpningsmedel, vilka kan utgöra ett hot mot miljön. Ett alternativt angreppssätt kan vara att identifiera och karakterisera genetiska faktorer som bidrar till bladlusresistens, samt medel som inducerar resistens, med avsikten att utnyttja denna kunskap i förädlingsprogram.

Syftet med denna avhandling var att identifiera sådana resistensgener och karakterisera deras roller i samspelet mellan växt och bladlus. Detta genomfördes med två angreppssätt. För det första klonades två förmodade bladlusresistensgener från korn och överfördes till Arabidopsis och korn och effekterna av transformationen undersöktes på två bladlusarter. Generna var

CI2c som kodar för en kymotrypsinhämmare och LOX2.2. som kodar för ett lipoxygenas. Bladlusarterna var havrebladlus (Rhopalosiphum padi L.) som är en skadegörare på viktiga sädesslag och generalisten persikbladlus ( persicae Sulzer) som är skadegörare på växter tillhörande mer än 40 olika familjer. Effekterna av transformation studerades också med avseende på uttrycket av andra försvarsrelaterade gener i de transformerade plantorna.

Det andra angreppssättet var att behandla plantor med luftburna ämnen och studera i vad mån detta framkallade resistens, följt av tester med havrebladlus på de behandlade växterna.

Studierna med överuttryck av CI2c visade att genprodukten i Arabidopsis gav en övergående reducering av persikbladlusens fruktsamhet, men i korn indirekt minskade samma arts undvikande av plantan på grund av att plantans försvar trycktes ned. Transformationen hade ingen effekt på havrebladlusens beteende eller fruktsamhet. Överuttryck av LOX2.2. i korn påverkade uttrycket av andra gener som regleras av jasmonsyra och minskade kortvarigt fruktsamheten både av havrebladlus och persikbladlus.

Studierna med behandling av luftburna ämnen gav stöd för idén att bladlusresistens kan framkallas med denna metod. Flera försvarsgener visade högre uttryck efter behandling med metylsalicylat, metyljasmonat och (Z)-3- hexen-1-ol. Av dessa tre undersökta ämnen, visade metyljasmonat den största potentialen som försvarsframkallande ämne, då den orsakade kortvarig reducering av fruktsamhet hos bladlusen.

Sammanfattningsvis stöder denna avhandling idéerna att korngenerna

CI2c och LOX2.2. spelar en roll i bladlusresistens och att viss bladlusresistens kan framkallas genom externa faktorer. Bladlössen kan direkt påverkas av genprodukten eller det kan finnas en indirekt effekt, som orsakas av förändringar i genuttryck av andra gener som är inblandade i växtens försvar.

De observerade negativa effekterna på bladlössen var av måttlig omfattning och det är troligt att de var för sig inte har någon stark negativ effekt, men att de kan bidra till att förmedla bladlusresistens.

List of publications

This thesis is based on the work contained in the following papers and man- uscripts, referred to in Roman numerals in the text:

I: Losvik A, Beste L, Mehrabi S, Jonsson L (2017). The protease inhibitor

CI2c gene induced by bird cherry-oat in barley inhibits green

peach aphid fecundity in transgenic Arabidopsis. International Journal of

Molecular Sciences 18: 1317.

II: Losvik A, Beste L, Stephens J, Jonsson L (2018). Overexpression of the

aphid-induced serine protease inhibitor CI2c gene in barley affects the

generalist green peach aphid, not the specialist bird cherry-oat aphid.

Accepted for publication in PLoS One.

III: Losvik A, Beste L, Glinwood R, Ivarson E, Stephens J, Zhu Li-Hua, Jons-

son L (2017). Overexpression and down-regulation of barley lipoxygen-

ase LOX2.2 affects jasmonate-regulated genes and aphid fecundity. In-

ternational Journal of Molecular Sciences 18: 2765.

IV: Losvik A, Glinwood R, Jonsson L. Effects of treatments of barley with

methyl salicylate, methyl jasmonate or (Z)-3-hexen-1-ol on barley gene

expression and fecundity of bird cherry-oat aphid (Manuscript).

Papers I, II and III are reproduced with the permission of their respective pub- lisher.

My contributions to the papers

Paper I: Was main responsible for the experimental design, conducted

the experiments, except for the cloning and transformation and

carried out the analysis of data. Wrote the first draft of the man-

uscript.

Paper II: Was main responsible for the experimental design, conducted

the experiments, except for the cloning and transformation,

and analyzed the data. Contributed to writing the manuscript.

Paper III: Designed and conducted the experiments, except for the clon-

ing and transformation, and performed the analysis of data.

Contributed to writing the manuscript.

Paper IV: Participated in the experimental design, conducted the experi-

ments and the analysis of data. Wrote the first draft of the man-

uscript.

Contents

Introduction ...... 11

Aphids as plant pests ...... 13

Host specificity ...... 13

Bird cherry-oat aphid ...... 14

Green peach aphid ...... 16

Host recognition and feeding ...... 17

Reproduction ...... 18

Damage to barley ...... 19

Plant defenses against aphids ...... 19

Direct defenses ...... 20

Indirect plant defenses ...... 23

Triggering of plant defense response ...... 24

Pattern-triggered immunity (basal defense) ...... 25

Effector-triggered immunity (active defense) ...... 27

Aphid-associated effectors ...... 28

Defense signaling pathways ...... 29

The cross-talk of SA and JA pathways ...... 34

Aphid resistance genes...... 35

Aphid management in agriculture ...... 36 Aim of the thesis ...... 41

Main approaches ...... 43

Hypotheses for the thesis work...... 43

Putative resistance factors against aphids ...... 43

Induction of resistance ...... 45 Comments on methodology ...... 47

Plant transformation ...... 47

Selection and controls ...... 49

Confirmation of transformation ...... 50

Plant phenotype ...... 52

Aphid performance ...... 53

Antixenosis ...... 53

Antibiosis ...... 54 Main results ...... 57

Effects of transformation or plant treatments with volatiles ...... 57

Effects of transformation or treatments on aphid performance ...... 60

CI2c overexpression ...... 60

LOX2.2 overexpression and down-regulation ...... 63

Volatile treatments ...... 64

Other effects of transformation or treatments ...... 67

Differences in aphid performance ...... 70

Between aphid species: BCA and GPA on barley ...... 70

Between host plants: GPA on Arabidopsis and barley ...... 72 Conclusions ...... 73

Future perspectives ...... 75

Acknowledgements ...... 79

References ...... 81

Abbreviations

AOS Allene oxide synthase

BCA Bird cherry-oat aphid

BCI Barley chemically induced

BYDV Barley yellow dwarf virus

CI2c Barley chymotrypsin inhibitor

COI1 Coronatine-insensitive 1

CRISPR Clustered regularly interspaced short palindrome repeats

DAMPs Damage-associated molecular patterns dsRNA Double-stranded RNA

Eβf (E)-β-farnesene

EPG Electrical penetration graph

ETI Effector-triggered immunity

GLV Green leaf volatile

GMO Genetically modified organism

GPA Green peach aphid h Hour

HAMPs Herbivore-associated molecular patterns

HPL Hydroperoxide lyase

Hx Hydroxamic acid

JA Jasmonic acid

JAZ Jasmonate ZIM domain

LOX Lipoxygenase

MeJA Methyl jasmonate

MeSA Methyl salicylate

NPBT New plant breeding technique

NPR1 Non-expressor of PR genes 1

ODN Oligodeoxynucleotides

PAMPs Pathogen-associated molecular patterns

PI Proteinase inhibitor

PRR Pattern-recognition receptors ps Phloem specific

PTGS Post-transcriptional gene silencing

PTI Pattern-triggered immunity

R-gene Resistance gene

ROS Reactive oxygen species qPCR Quantitative polymerase chain reaction

RNAi RNA-mediated interference

SA Salicylic acid siRNA Small interfering RNA

TILLING Targeting induced local lesions in genomes

TPPE Thionin proprotein-processing enzyme

VIGS Virus-induced gene silencing

WT Wild type

Introduction

Barley (Hordeum vulgare L.) is a self-pollinating, diploid species, a member of the grass family (Poaceae) and a major cereal grain grown in temperate re- gions of the world. Domestication of barley began about 10,000 years ago from its wild progenitor H. vulgare ssp. spontaneum (C. Koch) (Zohary and Hopf,

2000). Barley has evolved to include several forms. The two–row and six–row varieties differ in the protein content; the two-row barley has lower protein content and more fermentable sugar and is used for brewing, while six-row barley is used to feed livestock (Ullrich, 2011). Traditionally, spring malting barley varieties were grown in regions with moderate temperatures and suf- ficient rainfall during the growth season, whereas winter barley was mostly grown in the warmer and dry parts of Europe (Pržulj et al., 1998). Barley was probably initially used as human food, but due to the rise in prominence of wheat and rice, eventually evolved into a feed and later into a source of fer- mentable material for the production of beverages (Ullrich, 2011). In 2014, bar- ley ranked fourth in acreage among major cereal crops, after maize, rice and wheat, covering more than 49 million hectares worldwide, with a total aver- age yield of 2,9 tons per hectare (FAOSTAT, 2014).

Barley is a host for numerous pathogens such as Puccinia species which are causal agents of steam, leaf and stripe rust, Xantomonas campestris causing bac- terial blight, Blumeria graminis f.sp. hordei causing powdery mildew and

Fusarium fungi causing head blight infections (Paulitz and Steffenson, 2011).

Aside from bacterial and fungal pathogens, barley is also hosting a few virus species, among them barley yellow dwarf virus (BYDV). The most important pests on barley are aphids, which cause direct damage by feeding from

11 the plant’s phloem and indirect damage by transmitting plant viruses (e.g.

BYDV) (Miller and Rasochová, 1997; Plumb, 2002; Gray and Gildow, 2003).

Chemical control strategies are commonly used to constrain aphid popula- tions, however in spite of strict regulations for pesticide transport and appli- cation, the many drawbacks of insecticide usage necessitate a search for alter- native methods. Breeding for aphid resistance is a less harmful pest manage- ment strategy (Åhman et al., 2010). The pursuit to find aphid resistance factors is enhanced by the emergence of new molecular techniques in plant science.

The sequencing of the barley genome that was finished in 2012 facilitates mo- lecular studies on this plant and simplifies the production of new resistant cultivars (Mayer et al., 2012). The genome size has been reported to be about

5.1 gigabases with around 40 thousand coding genes (Mayer et al., 2012).

A barley breeding program was initiated, attempting to introduce genes for aphid resistance from barley wild ancestor H. vulgare ssp. spontaneum into cultivated barley, and those attempts resulted in moderately aphid resistant double haploid lines (Åhman et al., 2000). The original resistance bioassays showed ca. 15% lower increase in nymphal weight on those lines, when com- pared to an aphid-susceptible barley cultivar (Åhman et al., 2000). The barley wild ancestor, an aphid resistant barley double haploid line (DH28:4) and two barley susceptible cultivars were used in a comparative transcript profiling study based on a microarray, which resulted in the identification of candidate genetic factors contributing to the observed aphid resistance. Among them were gene sequences coding for a protease inhibitor (PI), a lipoxygenase

(LOX) and thionins (Delp et al., 2009; Mehrabi et al., 2014). The work in this thesis aims to explain the contribution of two of the suggested resistance fac- tors (PI and LOX) to aphid resistance by studying aphid responses to plants with overexpression or downregulation of the selected genes as well as re- sponses of the transgenic plants to aphid feeding at the molecular level. In

12 addition, the effects of volatile treatments were studied on different cultivars of barley, with regard to activation of plant defense, followed by bioassays with bird cherry-oat aphid (BCA, Rhopalosiphum padi L.) on the treated plants.

Aphids as plant pests

Host specificity

Aphids are members of the superfamily Aphidoidea of the order Hemip- tera. They are small (1-10 mm), soft-bodied insect pests which feed on the phloem sap. There are more than 4000 species of aphids distributed world- wide (Dixon 1985a; Dixon et al., 1987). Most aphid species (95%) are monoph- agous or oligophagous and show high host specificity. Polyphagous species, such as the black bean aphid (Aphis fabae Scopoli) are associated with a wide range of primary and secondary hosts (Blackman and Eastop, 2007). Consid- ering aphids’ small body and limited energy reserves, the time they can sur- vive without a suitable food source is narrow. Thus, aphids are associated mainly with abundant plants species and the greatest amount of aphid species occurs in temperate regions (Dixon, 1985b). Most aphid species are monoe- cious, completing their life cycle on one host and some of them, like the pea aphid (Acyrthosiphon pisum Harris) and grain aphid (Sitobion avenae Fabricius) are known to live only on herbaceous hosts or grasses, respectively. Only 10% of aphids are host-altering species (Figure 1) and spend the summer on dif- ferent plant species than the primary (winter) host, where sexual reproduction occurs. The secondary host, where populations of parthenogenetic females develop, is often an herbaceous crop plant species, and those aphids are im- portant from the agronomic point of view (Williams and Dixon, 2007).

13

Figure 1. Life cycle of host-altering BCA. A fundatrix (foundress) is a wing- less parthenogenetic female that produces live offspring during spring; af- ter one or two generations, winged offspring (alataes) migrate to the sec- ondary host to establish a population of parthenogenetic females. With the approach of winter, a winged female (gynopara) migrates to the primary host and produces ovipara - wingless, sexual female that mates with a winged male to produce overwintering eggs (adapted from Finlay and Luck, 2011).

Bird cherry-oat aphid

The colour of the BCA (Figure 2) varies from yellow-green to green-olive and greenish-black with a rust-coloured pigmentation around the cornicles. It is a cosmopolitan, host-alternating insect species (Figure 1) occurring in tem- perate regions. Permanent parthenogenetic development can however result from warmer weather conditions or a decrease of the geographical range of the primary host plant (Simon et al., 1996; Williams and Dixon, 2007;

Strażyński and Ruszkowska, 2015). BCA with a holocyclic life cycle uses bird cherry (Prunus padus L.) as its primary host and all major cereals and pasture

14 grasses as secondary host. The overwintering eggs on the primary host hatch in the spring and after two to three wingless generations, winged emigrant morphs appear (Dixon, 1971; Glinwood and Pettersson, 2000a). Alataes fly to the summer host where they give rise to further parthenogenetic generations

(Wiktelius et al., 1990). Winged migrants are born and migrate to new host plants when the nutrient source becomes depleted. With the approach of cold weather and longer periods of darkness, the winged gynoparae and males return to bird cherry, whose senescing leaves provide a rich food source

(Dixon and Glen, 1971). Apterous oviparae are born and mate with the winged males to produce overwintering eggs (Figure 1) (Dixon, 1971).

Figure 2. Apterous adult BCA feeding on oat leaf. Photo by A. Losvik

15 Green peach aphid

The green peach aphid (GPA, Myzus persicae Sulzer) (Figure 3) is a cosmo- politan, polyphagous aphid species. The color varies from pale greenish-yel- low and various shades of green, to pink, red, or even almost black (Blackman and Eastop, 2007). The life cycle includes primary and secondary hosts; how- ever, it can reproduce asexually during the warmer months, and throughout the year in warmer climates. The primary host comprise stone fruits like peach

(Prunus persicae L.), Canadian plum (P. nigra Aiton), black cherry (P. serotine

Ehrhart), and dwarf Russian almond (P. tenella Batsch) (Blackman and Eastop,

2000). As secondary hosts GPA uses a large variety of plants, including crops belonging to Brassicaceae, Solanaceae and Cucurbitaceae families. Altogether, the host range comprises plant species belonging to more than 40 plant fami- lies (Blackman and Eastop, 2000; Louis and Shah, 2013). Furthermore, GPA is a vector for many viral plant diseases and its resistance to a large number of insecticides has been reported (Louis and Shah, 2013; Bass et al., 2014).

Figure 3. Apterous adult GPA feeding on kohlrabi leaf (Brassica oleracea L.). Photo by A. Losvik

16 Host recognition and feeding

Aphids are poor flyers and their migration depends strongly on air move- ment. Visual and olfactory cues can initiate the landing of a winged migrant, an alatae, chemical and nutritional cues available to the aphid may lead to host acceptance or rejection (Powell et al., 2006; Nam et al., 2013; Smith and

Chuang, 2014). To establish a feeding site, aphids scan the surface of a poten- tial host to detect veins and penetrate the tissue. The stylet pathway is mostly intercellular and is facilitated by gelling saliva excreted by aphids at the time of probing (Will et al., 2012). Gel saliva is mainly composed of proteins (in- cluding phenoloxidases, peroxidases, pectinases, β-glucosidases), phospho- lipids and conjugated carbohydrates (Urbanska et al., 1998; Miles, 1999;

Cherqui and Tjallingii, 2000). While progressing toward the phloem, stylets briefly puncture many cells and after withdrawal the punctured site is quickly sealed by gelling saliva (Tjallingii and Esch, 1993; Powell et al., 2006; Tjallingii,

2006). There are indications that aphids discriminate between host and non- host plants before reaching the phloem, by uptake of small amounts of intra- cellular substance. Those studies were based on electrical penetration graph

(EPG) recordings, combined with observations of aphid behavior (Powell et al., 2006; Goggin, 2007). In the EPG technique, an aphid and plant are made parts of an electric circuit, which is completed when the aphid inserts its stylet into the plant. Voltage fluctuations in the circuit create distinct EPG wave- forms, patterns of amplitude, frequency, and voltage level, which correlate with the insect´s stylet activities (Tjallingii and Esch, 1993). During the phase of superficial piercing, as well as during initial establishment in the phloem, aphids secrete watery saliva, believed to aid in gustatory exploration and pre- vent plant defense responses (Miles, 1999; Powell et al., 2006; Will et al., 2009;

Moreno et al., 2011). Watery saliva contains a complex mixture of proteins e.g. hydrolases, oxidases and calcium-binding proteins, the latest involved in the

17 inhibition of calcium-mediated occlusion of phloem sieve elements (Urbanska et al., 1998; Miles, 1999; Carolan et al., 2009). Saliva composition differs be- tween aphid species and is dependent on aphid’s diet (Cherqui and Tjallingii,

2000; Carolan et al., 2009; Will et al., 2009). The successful recognition and host acceptance may result in a compatible or incompatible reaction, where respec- tively aphids reach reproductive success or where they are not able to exploit a resistant plant (Powell et al., 2006; Nam et al., 2013). The herbivore-specific compounds of aphid salivary secretions can have elicitor-like properties and serve as herbivore-associated molecular patterns (HAMPs) (Mithöfer and

Boland, 2008). Additionally, saliva components and stylet progress through plant tissue may potentially lead to a release of damage-associated molecular patterns (DAMPs) from plant cells. The induction of plant innate immune re- sponse by HAMPs and DAMPs is discussed in more detail in section “Trig- gering of plant defense responses”.

Aphids feed passively on phloem sap, which is rich in sugars and poor in amino acids, especially the essential ones (Dinant et al., 2010). Because of this unbalanced diet, aphids need to ingest large amounts of food and the result is the excreted honey dew (Douglas, 2006). The diet is further supplemented with essential amino acids produced by the primary obligate prokaryotic symbiont from the genus Buchnera, present in bacteriocytes – specialized fat body cells localized near the aphid’s gut (Baumann et al., 1995; Douglas, 1998).

Reproduction

With few exceptions, aphids live in colonies, with both advantages and drawbacks. Aggregation provides increased protection from natural enemies, however colonies can grow to a size at which the colony loses sustainability on the host plant (Pettersson et al., 2007). This extremely rapid growth of a colony is secured by parthenogenesis, where offspring is produced from

18 unfertilized eggs, and telescopic generations, where as a result, fully devel- oped nymphs are born. Aphid nymphs molt during growth and pass through

3-4 instars before becoming adults (Awmack and Leather, 2007). For most aphid species, parthenogenesis is cyclical, with periods of asexual reproduc- tion alternating with sexual reproduction (holocyclic life cycle, Figure 1).

Some species, such as GPA are examples where all generations can comprise of asexual (parthenogenetic) females (Dixon, 1985c; Williams and Dixon,

2007).

Damage to barley

There are a few aphid species associated with barley and they cause differ- ent injury symptoms. The Russian wheat aphid (Diuraphis noxia Kurdjumov) causes leaf rolling and discoloration, whereas chlorosis and necrosis of plant leaves are characteristic for greenbug (Schizapis graminum Rondani) attack

(Quisenberry and Ni, 2007). BCA is a pest on grasses, causing no visible inju- ries to plants, but can reduce yield directly by withdrawing nutrients from the plant phloem and by transmitting plant viruses, for example the BYDV. This virus is the causal agent of one of the most important viral disease of barley and other small grain cereal crops worldwide (Paulitz and Steffenson, 2011).

It causes leaf discolouration and dwarfing and the infection may lead to losses approaching 100% if appearing early in the season (Irwin and Thresh, 1990;

Mathre, 1997; Paulitz and Steffenson, 2011; Ordon and Perovic, 2013).

Plant defenses against aphids

Non-host resistance includes traits preventing a plant from being a host and is the basic defense used by plants. It can be defined as the immunity, which is shared by all genotypes of a plant species including susceptible plants and is able to limit non-adapted herbivore populations. This type of

19 resistance has considerable value for crop improvement, since it is typically both broad-spectrum and durable (Lee et al., 2016). Traits that deter herbivore performance on host plants contribute to host plant resistance (Goggin, 2007).

Insect responses to host resistance characteristics can be of behavioral

(antixenosis) or physiological (antibiosis) nature (Smith, 2005a; 2005b). Plants can employ direct and indirect defense mechanisms for protection against pathogens and herbivores, both can be present constitutively before insect at- tack or can be induced.

Direct defenses

Surface barriers

The main determinants of plant resistance to herbivores are direct mecha- nisms, which include physical and chemical traits. The physical barriers can be created by the presence of leaf surface wax, cutin layers and trichomes

(Hanley at al., 2007; War et al., 2012). This line of defense was shown to have variable effects on aphids. A high density of trichomes was shown to have negative effects on GPA feeding on wild species of tomato (Lycopersicon)

(Wójcicka, 2015), and the presence of waxes impaired probing behavior of S. avenae on triticale (× Triticosecale Wittmack) plants (Wójcicka, 2015), however there exist also reports showing no effects of trichomes on aphids (Åhman et al., 2000; Sato and Kudoh, 2015; Karley et al., 2016).

Lignin is a phenolic heteropolymer deposited predominantly in the walls of cells showing secondary thickening, especially abundant in vascular (xy- lem tracheids, vessel elements) and support tissues (sclereid cells). Lignin depositions provide rigidity and imperviousness to the cells (Vanholme et al.,

2010). In addition, lignin biosynthesis can be induced upon various biotic and abiotic stress conditions and plays an important role in plant defense against

20 pathogens and , by hindering pathogen entry or increasing leaf tough- ness to reduce feeding by herbivores (War et al., 2012). It was shown that the chrysanthemum aphid (Macrosiphoniella sanborni Gillette) induced CmMYB19, a transcription factor important in the regulation of lignin synthesis. Overex- pression of CmMYB19 restricted the multiplication of aphids by enhancing accumulation of lignin (Wang et al., 2017). Callose and lignin depositions were also induced by melon/cotton aphid (Aphis gossypii Glover) in the cell walls of aphid-resistant melon genotypes, carrying the Vat gene and hinder- ing aphids from reaching the vascular tissue, but not in aphid susceptible lines

(Villada et al., 2009).

Biochemical defense

Chemical barriers comprise a big group of plant secondary metabolites which are present constitutively or are induced by pathogen/herbivore attack.

These compounds belong to various chemical classes such as phenolics, ter- penoids and nitrogen-containing organic compounds (Mithöfer and Boland,

2012). Some attention has been directed toward understanding the involve- ment of cereal secondary metabolites in resistance to aphids (Leszczynski and

Dixon, 1990; Forslund et al., 1998; Åhman et al., 2000; Larsson et al., 2006;

Goławska et al., 2008; Larsson et al., 2011).

Gramine is an indole alkaloid which occurs in leaves of certain barley cul- tivars (Forslund et al., 1998; Larsson et al., 2006). It’s involvement in defense against aphids has been suggested, however results from different studies do not provide clear evidence to support this idea. It has been shown that gra- mine is present in epidermis and mesophyll parenchyma cells of barley seed- lings and can modify S. graminum and BCA feeding behavior (Zúñiga et al.,

1988). It has been correlated with barley seedling resistance to the two above- mentioned aphid species (Zúñiga et al., 1985; Zúñiga and Corcuera, 1986).

21 More recent studies however show no negative effect of high gramine content on BCA settling (Forslund et al., 1998) or growth (Åhman et al., 2000).

Another group of secondary metabolites present in barley is the cyano- genic glycosides, which release hydrogen cyanide upon tissue damage. Simi- larly to gramine, no conclusive evidence has been provided with regard to their involvement in defense against aphids. In one of the studies, the cyano- genic glycoside dhurrin occurring in sorghum (Sorghum bicolor Moench) was shown to be deterrent to S. graminum (Dreyer et al., 1981) in experiments where it was supplemented to aphids’ artificial diet. This effect was however not observed for the corn leaf aphid (Rhopalosiphum maidis Fitch) feeding on different sorghum varieties with varying content of dhurrin in the field

(Woodhead et al., 1980). It was further demonstrated that BCA settling did not differ on cultivars with different levels of cyanogenic glucosides (Forslund et al., 1998).

Cyclic hydroxamic acids (Hx) are a group of secondary metabolites char- acteristic for the family Poaceae and occur in wild grasses, as well as in culti- vated crops such as maize (Zea mais L.), wheat (Triticum aestivum L.), rye (Se- cale cereale L.), and triticale, but are absent from barley, oats (Avena sativa L.) and rice (Oryza sativa L.) (Zúñiga et al., 1983; Niemeyer, 1988; Slesak et al.,

2001). Hx seem to be defensive factors against insects (Ahmad et al., 2011;

Meihls et al., 2013). It was shown by using genetically-defined maize lines, that the growth and survival of BCA, as well as reproduction of R. maidis were negatively affected on plants accumulating 2,4-dihydroxy-7-methoxy-2H-1,4- benzoxazin-3(4H)-one and its glucoside form (Ahmad et al., 2011; Meihls et al., 2013).

Another example of plant defense compounds with suggested protective functions are the thionins. They are small (~5kDa), cysteine-rich peptides and can be divided into the families of α/β-thionins and γ-thionins (plant

22 defensins) (Westermann and Craik, 2010). The barley leaf thionins are en- coded by a large gene family, with more than 50 members per haploid genome

(Bohlmann et al., 1988; Bunge et al., 1992). Thionins are involved in protection against pathogenic bacteria and fungi by targeting the pathogen’s membrane

(Stec, 2006). Their involvement in protection against insects has also been sug- gested (Jones et al., 1985; Bohlmann et al., 1988; Charity et al., 2005; Pelegrini and Franco, 2005). Four barley thionin sequences were found to be constitu- tively expressed in moderately aphid resistant barley genotypes and all were upregulated by aphid feeding (Delp et al., 2009). It was shown that GPA prog- eny production on Nicotiana benthamiana (Domin) leaves transiently overex- pressing barley thionin was reduced nearly 50% when compared to the vector control, the mechanism of this reduction remains however unknown

(Escudero-Martinez et al., 2017). Barley leaf thionins have been localized both in cell walls and intracellularly (Charity et al., 2005), which creates the possi- bility for aphids to encounter thionins during penetration or probing.

It has long been suggested that secondary metabolites might protect against generalists, but not against specialists (van Emden, 1972). Glucosin- olates constitute a much-studied example of this idea. Upon tissue damage, these compounds are broken down by the enzyme myrosinase to toxic isothi- ocyanates, nitriles and other compounds with documented anti-herbivore ef- fects (Kelly et al., 1998; Renwick, 2002). The specialized cabbage aphid (Brevi- coryne brassicae L.) can however use the glucosinolates for its own protection against predators (Bridges et al., 2002; Kazana et al., 2007).

Indirect plant defenses

Attracting natural enemies, predators and parasitoids that prey upon her- bivorous insects, is an indirect defensive strategy employed by plants. Two particularly widespread indirect defense traits are extrafloral nectar and

23 volatile organic compounds (Thaler, 1999; Heil, 2008; Radhika et al., 2008).

Host plant-specific and herbivore-specific release of herbivore induced plant volatiles comprises the cue for predators and parasitoids used to locate prey

(de Vos and Jander, 2010). Volatile organic compounds released by plants can guide foraging behavior, especially significant for specialist insects, which have to invest more time in searching for host and prey than generalists. It has been shown that plant volatiles such as terpenoids and green leaf volatiles

(GLVs) induce electroantennographic responses of hoverfly (Episyrphus bal- teatus De Geer). Larvae of E. balteatus are insectivores and prey on aphids. The oviposition site is selected by the female and has important impact on off- spring performance. It has been demonstrated that (Z)-3-hexenol increases the motility of females, plant acceptance and oviposition activity even in the ab- sence of prey. The same effect on E. balteatus behavior has been observed for the aphid alarm pheromone, (E)-β-farnesene (Eβf) (Verheggen et al., 2008).

Seven-spot ladybird (Coccinella septempunctata L.) is considered an im- portant beneficial insect, an aphid predator. The ladybird is known to make use of volatile chemical cues in locating prey (Pettersson et al., 2008). It re- sponds positively to volatiles from aphid-infested and previously attacked barley plants, but not to uninfested plants and undisturbed nonfeeding aphids (Ninkovic et al., 2001). Moreover, the ladybird C. septempunctata has been shown to be able to learn to associate the odor of aphid-infested plants with the presence of prey (Glinwood et al., 2011a).

Triggering of plant defense response

Plant perception of pathogens and hemipteran insects share some common features. There are two types of receptors involved in triggering plant defense response, which have distinct cellular localization: plasma membrane pattern recognition receptors (PRRs) which induce pattern-triggered immunity (PTI)

24 and intracellular receptors, recognizing effectors delivered inside the plant cell, which results in effector triggered immunity (ETI) (Figure 4) (Smith and

Clement, 2012; Kaloshian and Walling, 2016). PTI usually activates basal re- sistance mechanisms, whereas ETI is the main mechanism for host resistance gene (R-gene) mediated resistance (Lee et al., 2016).

Figure 4. Schematic representation of PTI and ETI in plant–aphid interac- tions. (a) HAMPs are recognized by membrane-bound PRRs to trigger PTI. (b) Delivery of effectors that interfere with PTI leading to increased suscep- tibility.(c) Effectors may be recognized by intracellular R-proteins (encoded for example by Mi or Vat genes) resulting in activation of ETI (figure from Hogenhout and Bos, 2011; reproduced with the permission from the publisher).

Pattern-triggered immunity (basal defense)

The plant-attacker interaction begins after recognition of molecular pat- terns associated with the microbes (microbe associated molecular patterns,

MAMPs) or upon perception of HAMPs, by plant PRRs (Kaloshian and

Walling, 2016). HAMPs may function in concert with herbivory-induced mol- ecules originated from plants themselves, DAMPs. PTI is an important

25 component of plant innate immunity to insects. No aphid-associated molecu- lar pattern has yet been identified, however it was shown that extract of the

GPA as well as treatments with GPA saliva triggered responses characteristic of PTI in Arabidopsis (de Vos and Jander, 2009; Prince et al., 2014). PTI was also shown to be activated by patterns associated with the bacterial endosym- bionts. Buchnera’s most abundant protein GroEL has been identified in aphid saliva and honeydew (Baumann et al., 1996; Sabri et al., 2013; Chaudhary et al., 2014; Vandermoten et al., 2014). The perception of aphid effectors involves

BAK1 (BRASSINOSTERIOD INSENSITIVE 1-ASSOCIATED RECEPTOR KI-

NASE 1), which is known to interact with PRRs to induce defense responses.

HAMPs induce influx of Ca2+ ions as part of the signal transduction chain (Wu and Baldwin, 2010). The subsequent accumulation of superoxide anion (O2−), hydrogen peroxide (H2O2), singlet oxygen (1O2) and hydroxyl radical (·OH), collectively called ROS, marks the early post-invasive plant defense to biotic stress (Wojtaszek, 1997; Jones and Dangl, 2006; Wu and Baldwin, 2010;

Jaouannet et al., 2014; Prince et al., 2014). ROS are produced in mitochondria, chloroplasts, and peroxisomes, as well as on the plasma membrane’s external surface. The herbivore-induced production of ROS is dependent on plasma membrane-bound NADPH oxidases (Wu and Baldwin, 2010). ROS produc- tion was shown to differ in Arabidopsis depending on the attacking aphid species. ROS amounts were lower in compatible Arabidopsis-GPA interac- tions, than in the case of poor- or non-host interactions with black cherry aphid (Myzus cerasi Fabricius) or BCA, respectively (Jaouannet et al., 2015). In addition, Arabidopsis NADPH oxidase knock-out lines with hindered pro- duction of ROS showed increased populations of GPA and M. cerasi, as well as increased survival of BCA (Jaouannet et al., 2015).

With regard to plant damage, aphids which display characteristic feeding behavior with active suppression of plant defense responses do not cause

26 plant tissue damage comparable to damage caused by chewing insects. Com- ponents of aphid saliva and the process of probing may potentially lead to formation of DAMPs (Miles, 1999). The largest and best characterized class of plant DAMPs are polypeptides/peptides produced from larger precursor pro- teins, such as the systemin of Solanaceous species (Choi and Klessig, 2016).

Different DAMPs are most likely recognized by yet unknown PRRs, which triggers the synthesis of jasmonic acid (JA) in the chloroplast and activates JA- signaling (discussed in detail in the section “JA and the LOX pathway in plant response to aphids”). Alternatively, Ca2+-influxes which can, among others, be triggered by the damage-associated perception of extracellular ATP, initiate the formation of ROS by NADPH oxidase and downstream activation of de- fense genes (Wu and Baldwin, 2010; Heil and Land, 2014).

Effector-triggered immunity (active defense)

Some pathogens or insect pests have evolved to suppress the basal defence by deploying effectors into plant cells. In turn, some plant hosts have evolved a second line of defence, ETI, in response to such effectors. This type of plant response requires resistance (R) genes in plants which can interact with insect effector molecules (avirulence factors, Avr). Most plant R genes providing re- sistance to pathogens and pests encode nucleotide-binding leucine rich repeat proteins (NB-LRR) (Jones and Dangl, 2006; Hogenhout and Bos, 2011). Three

R genes coding for proteins with NB-LRR structure and conferring resistance to aphids have been cloned and characterized: the Mi-1.2 gene conferring re- sistance to potato aphid in tomato (Rossi et al., 1998), the melon Vat gene providing resistance to melon/cotton aphid (Dogimont et al., 2014; Smith and

Chuang, 2014), and Ra gene from lettuce providing resistance to lettuce root aphid (Pemphigus bursarius L.) (Wroblewski et al., 2007). Those mentioned R-

27 gene encoded proteins lack organellar localization signals and thus are pre- sumed to be confined to the cytosol (Kaloshian and Walling, 2016).

Aphid-associated effectors

A number of potential effectors in aphid saliva and salivary glands have been identified by proteomics (Harmel et al., 2008; Carolan et al., 2009; Cooper et al., 2010; Carolan et al., 2011; Cooper et al., 2011; Nicholson et al., 2012; Rao et al., 2013; Vandermoten et al., 2014) and functional genomics studies (Mutti et al., 2006; Mutti et al., 2008; Bos et al., 2010; Pitino et al., 2011; Pitino and

Hogenhout, 2013; Elzinga et al., 2014; Wang et al., 2015). These aphid-specific proteins can inhibit plant defences, prevent phloem sieve-element occlusion, and promote phloem feeding by aphids (Elzinga and Jander, 2013). For exam- ple, a salivary protein Mp55 from GPA increased aphid reproduction when overexpressed in cultivated tobacco (Nicotiana tabacum L.) (Elzinga et al.,

2014). A C002 protein from salivary gland of A. pisum was shown to be crucial in the feeding establishment of this aphid on fava bean (Vicia faba L.) (Mutti et al., 2008). RNAi-based knockdown of A. pisum and GPA C002 transcripts sig- nificantly reduced the life span of pea aphids on fava bean leaves and fecun- dity of GPA (Mutti et al., 2006; Pitino et al., 2011).

It is known that chitin is a fungal PAMP; it is also an important component of the aphid stylets and exoskeleton structures (Uzest et al., 2010) and thus aphid saliva may contain chitin fragments detached during stylet movement

(van Bel and Will, 2016). It has been suggested that plants detect aphid chitin as non-self molecules to strengthen the plant-defense response, however this aspect remains to be further investigated (Jaouannet et al., 2014; van Bel and

Will, 2016).

28 Defense signaling pathways

The defense response in plants involves phytohormone signaling. Phloem feeders have been shown to modulate plant responses regulated by three key compounds of the plant defense system: salicylic acid (SA), JA and ethylene

(ET) (Thompson and Goggin, 2006; Kuśnierczyk et al., 2011; Morkunas et al.,

2011).

SA in plant response to aphids

SA is a product of the shikimic acid biosynthesis pathway important for the hypersensitive response (Waling 2000) and known to promote a broad- range resistance to pathogens (Reymond and Farmer, 1998; Morkunas et al.,

2011). SA-dependent responses use SA and MeSA to activate a number of de- fense-related genes, among them pathogenesis related (PR) genes, whose products localize in the apoplast (Moran and Thompson, 2001). Signaling downstream of SA is largely dependent on the regulatory protein NPR1 (non- expressor of PR genes 1) (Figure 5). SA presence in induced cells (e.g. after pathogen attack) leads to monomerization of NPR1 and its translocation to the nucleus. In the nucleus, NPR1 monomers interact with TGA transcription factors, which bind to the promoters of SA-responsive genes, such as PR-1, resulting in their activation. In addition, many WRKY transcription factor genes are induced, which can activate or repress SA responses. During this process, NPR1 becomes phosphorylated, ubiquitinylated and degraded in the proteasome (Figure 5) (Kesarwani et al., 2007; Pieterse et al., 2012).

29 related genes in barley (Beβer et al., 2000; Forslund et al., 2000; Li et al., 2005;

Glinwood et al., 2007).

MeSA is also known as an aphid repellent. It was shown to inhibit BCA attraction to aphid-infested oat (Pettersson et al., 1994), reduce BCA settling on barley, both in greenhouse and field conditions (Ninkovic et al., 2003;

Glinwood et al., 2007), as well as decrease infestation on wheat by S. avenae and rose-grain aphid (Metopolophium dirhodum Walker) by 40 to 50%

(Pettersson et al., 1994).

JA and the LOX pathway in plant response to aphids

JA is a low molecular mass regulator of inducible defense gene expression and plays an important role in a range of developmental processes (seed ger- mination, tuber formation, senescence, flower development) (Wasternack,

2007). JA and its derivatives are products of the allene oxide synthase (AOS) branch of the LOX pathway (Wasternack, 2007; Mosblech et al., 2009; Lyons et al., 2013; Wasternack and Hause, 2013; Borrego and Kolomiets, 2016). Lev- els of JA rise after wounding and herbivore attack, which triggers the plant defense responses. A similar effect can be obtained by exogenous application of JA or MeJA to plants (Glinwood et al., 2007; Morkunas et al., 2011). JA can be readily metabolized to MeJA or conjugated to amino acids such as isoleu- cine via the JA conjugate synthase JAR1. Binding of JA-Ile to COI1 (corona- tine-insensitive 1), a component of the ubiquitin ligase in induced cells (e.g. after herbivore attack) (Figure 6) leads to ubiquitylation and subsequent deg- radation of JAZ (jasmonate ZIM domain) transcriptional repressor proteins via the proteasome. This disables the physical interaction between JAZ pro- teins and transcriptional activators (Figure 6) (Thines et al., 2007). In Ara- bidopsis, two major branches of the JA signaling pathway are recognized: the

MYC branch associated with the wound response and defense against insect

31 herbivores, and the ERF branch associated with enhanced resistance to necrot- rophic pathogens (Howe and Jander, 2008; Pieterse et al., 2012). In Arabidop- sis, VEGETATIVE STORAGE PROTEIN2 (VSP2) and PLANT DEFENSIN1.2

(PDF1.2) are the two marker genes for the two respective branches, MYC and

ERF (Figure 6) (Pieterse et al., 2012).

Figure 6. Schematic overview of the JA signaling pathways in induced cells, as summarized in the text. Abbreviations: SCF COI1, multi-protein E3- ubiquitin ligase named for the SKP/cullin/F-box proteins of the complex where COI1 is the F-box component; Ub, ubiquitinylated protein; (figure adapted from Pieterse et al., 2012).

JA and its derivatives were shown to induce expression of defense-related genes (Hause et al., 1999; Beβer et al., 2000; Bachmann et al., 2002; Glinwood et al., 2007; Walia et al., 2007). It has been suggested that JA regulates effective defenses against phloem feeders in dicot plants (Ellis et al., 2002; Gao et al.,

2007; Brunissen et al., 2010). Studies on monocot plants suggest that JA-

32 mediated pathway might affect aphid acceptance or feeding, without strong antibiosis effects (Slesak et al., 2001; Glinwood et al., 2007; Cao et al., 2014).

The hydroperoxide lyase (HPL) branch of the LOX pathway directs the for- mation of C6-volatiles (Bate and Rothstein, 1998; Matsui, 2006), which play a role in plant defense against pathogens and herbivores (Scala et al., 2013;

Naeem ul Hassan et al., 2015; Matsui and Koeduka, 2016). The involvement of C6 volatiles in defense against aphids has been shown in the case of tobacco aphid (Myzus nicotianae Blackman) and GPA (Hildebrand et al., 1993;

Vancanneyt et al., 2001).

ET in plant responses to aphids

ET and JA can act synergistically and activate defense responses in plants, which are different and often antagonized by responses induced by SA (Smith and Boyko, 2007). Both ET and JA are involved in induced systemic resistance activated in plants in response to non-pathogenic bacteria (van Loon et al.,

1998). ET signaling is in many cases dependent on JA induction (Von Dahl and Baldwin, 2007). It was shown that numbers of GPA were greatly reduced on Arabidopsis mutant cev1 with constitutive expression of JA and ET re- sponses (Ellis et al., 2002). Aphid feeding was also shown to significantly in- crease ET production in an aphid-resistant barley cultivar, as compared to susceptible plants. This has been shown for S. graminum and BCA

(Aragandona et al., 2001) as well as for D. noxia (Miller et al., 1994). In addi- tion, it was shown that feeding by potato aphid (Macrosiphum euphorbiae

Thomas) induces expression of ET biosynthetic genes both in susceptible

(mi/mi) and resistant (Mi-1/Mi-1) tomato cultivars, which correlated with a burst of ET after aphid feeding (Mantelin et al., 2009).

33 The cross-talk of SA and JA pathways

In summary, the above-mentioned signaling pathways do not act inde- pendently of each other, but rather are part of a complex network with posi- tive and negative regulatory interactions, to fine tune the defense responses.

The SA response pathway is typically recognized as effective against micro- bial biotrophic pathogens, whereas JA and its derivatives, as well as ET, act against necrotrophic pathogens and herbivores (Glazebrook, 2005). It has been suggested that SA inhibits the biosynthesis of JA and subsequent induc- tion of gene expression; similarly, that JA inhibits the effects of SA (Kunkel and Brooks, 2002; Morkunas et al., 2011; Thaler et al., 2012; Vidhyasekaran,

2015). Down-regulation of the JA-pathway was shown for Arabidopsis mu- tants that accumulate high levels of SA (Kachroo et al., 2003a; 2003b).

The NPR1 has been shown to be involved in the suppressive action of SA against JA signaling. In the cytosol, the SA-activated NPR1 pool may be in- volved in facilitating the delivery of negative regulators of JA-responsive genes to the nucleus or by inhibiting positive regulators of JA-responsive gene expression (Spoel, 2003). In Arabidopsis, the transcription factor WRKY70 is known to be involved in determining the balance between SA and JA signal- ing pathways (Li et al., 2004; Li et al., 2006). It was shown to inhibit expression of JA-responsive genes (Li et al., 2004). JA is known to regulate WRKY70 tran- script levels negatively, whereas SA acts as positive regulator (Li et al., 2004).

Gene expression profiling analysis has shown that SA and JA act coordi- nately to induce defense in response to aphid feeding (Moran and Thompson,

2001). Up-regulation of both SA and JA/ET-responsive genes was shown in

Arabidopsis after feeding by B. brassicae and A. gossypii (Moran and

Thompson, 2001). It was proposed, however, that while the activation of the two signaling pathways is harmonized and temporarily controlled

(Kuśnierczyk et al., 2008), aphids can manipulate plant-induced resistance, to

34 suppress JA-regulated defenses by the induction of the less efficient SA-regu- lated responses (Kuśnierczyk et al., 2008; Giordanengo et al., 2010). Genes coding for JAZ proteins which act as negative feedback regulators of JA sig- naling were activated at 6 h after B. brassicae attack, with subsequent regres- sion of JA-connected responses at 12 h (Kuśnierczyk et al., 2008). This “decoy” hypothesis was supported also by results from studies on Arabidopsis mu- tants, where GPA resistance was shown independent of high levels of SA

(Pegadaraju et al., 2005). Similarly, no differences were observed in GPA re- production between wild-type plants and mutants compromised in SA sig- naling (Moran and Thompson, 2001).

Aphid resistance genes

Previous studies have shown that gene up-regulation caused by aphid feeding provides resistance against aphids. Examples are the phytoalexin de- ficient 4 (PAD4) gene, the trehalose phosphate synthase 11 (TPS11) gene and

Myzus persicae-induced lipase 1 (MPL1) gene, all shown to control antibiosis and antixenosis against the GPA (Louis et al., 2010; Singh et al., 2011; Louis and Shah, 2015). PAD4 knockout and TPS11 or MPL1 overexpression were shown to support respectively higher and lower aphid population growth on

Arabidopsis (Louis et al., 2010; Singh et al., 2011; Louis and Shah, 2015). Sim- ilarly, a tomato α-dioxygenase, induced by potato aphids was shown to con- tribute to plant defense against this aphid and silencing of this gene resulted in higher aphid numbers (Avila et al., 2013).

In barley, BCA-induced gene expression was compared between two moderately aphid-resistant and two susceptible barley genotypes in a micro- array-based study (Delp et al., 2009). One of the resistant genotypes used in that study was an accession of wild barley H. vulgare L. ssp. spontaneum

(Hsp5). The other, double haploid breeding line DH28:4 (full number 5172-

35 28:4), was an offspring from a cross of Hsp5 with the susceptible cultivar Lina and backcrossed to Lina (Åhman et al., 2000). The resistance in Hsp5 and

DH28:4 was demonstrated as decreased (up to 40% and 15%, respectively) nymphal growth (Åhman et al., 2000). Results from this comparative tran- scriptome profiling revealed that four genes were up-regulated by BCA only in the resistant genotypes and not in the susceptible barley. Those genes were encoding for: a protease inhibitor (CI2c, known also as BCI-7), a lipoxygenase

(LOX2.2), a calcium-binding EF-hand protein (BCI-4) and a Ser/Thr kinase

(Delp et al., 2009). Those four genes were considered candidate aphid-re- sistance factors and two of them, the CI2c and LOX2.2, were in the main focus of this thesis.

Aphid management in agriculture

Aphid management is commonly based on pesticide applications, how- ever they have adverse environmental effects (Dewar, 2007; Hillocks, 2012).

The active compounds of insecticides, the organophosphates and carbamates, inhibitors of acetylcholine esterase, tend to have high toxicity to beneficial in- sects and mammals. As a consequence, many have been withdrawn from use

(Dewar, 2007). Control of viruses has also proven to be difficult using chemi- cal applications. Insecticides are usually applied after aphid infestation has reached a certain threshold and at this point virus transmission has already occurred. In addition, a number of aphid species have been reported to have developed insecticide resistance, some of them belonging to the most prob- lematic pest species (Foster et al., 2007; Bass et al., 2014; Bass et al., 2015;

Bettaibi et al., 2016).

Breeding for constitutive and inducible aphid resistance have been sug- gested as an alternative for more sustainable aphid control (Åhman, 2009).

With regard to inducible resistance, a variety of biological and chemical

36 agents have been successfully used, in the form of foliar sprays or seed treat- ments, under controlled and field conditions, to handle pathogen infections and insect pests populations, including aphids (Pushpalatha et al., 2011;

Worrall et al., 2011; Walters et al., 2013; Hamada et al., 2017). Volatile treat- ments show similar potential (Shulaev et al., 1997; Kishimoto et al., 2005;

Ninkovic and Åhman, 2009). Plants release volatile organic compounds from reproductive and vegetative parts and the amounts released increase upon mechanical damage or herbivore infestation (Bernasconi et al., 1998; Turlings et al., 1998; Paré and Tumlinson, 1999). However, there exist chemical inter- actions also between undamaged plants (Glinwood et al., 2011b). The term

“allelobiosis” was introduced to describe chemical interaction between un- damaged plants, which is beneficial for the receiving plant and which in- volves organisms from other trophic levels (Ninkovic et al., 2006). It has been shown that mixing different plant species or different genotypes of the same species can reduce incidences of herbivory (Andow, 1991; Ninkovic et al.,

2002; Ninkovic and Åhman, 2009). Volatile-induced resistance was shown to be a heritable trait and presents the option to breed for allelobiotic effects

(Ninkovic and Åhman, 2009). Mixed cultivation is however problematic in agriculture and there exists a risk that allelobiosis will not be easily trans- ferred into practice (Ninkovic and Åhman, 2009).

The potential of Eβf, an aphid alarm pheromone, in the control of pest aphids has also been recognized. It was shown that Arabidopsis plants ex- pressing Eβf synthase gene cloned from Mentha × piperita and emitting Eβf, caused alarm and repellent responses of GPA; additionally, transgenic plants became attractive to aphid parasitoid Diaeretiella rapae (M'Intosh). Similarly,

Eβf was shown to be an effective kairomone for the two-spot ladybird (Adalia bipunctata L) (Francis et al., 2004).

37 Breeding for aphid resistance has resulted in cultivars resistant to D. noxia,

S. graminum (Berzonsky et al., 2002), lettuce aphid (Nasonovia ribisnigri Mos- ley) (Liu and McCreight, 2006) and soybean aphid (Aphis glycines Matsumura)

(Li et al., 2007). Moreover, wheat cultivars which reduce BCA populations or tolerate the effects of aphid feeding have also been identified (Girvin et al.,

2017). The selection of a resistant cultivar was based on observed plant symp- toms after aphid infestation. This can be difficult in a situation when the aphid, such as BCA, does not cause visible symptoms. In barley and wheat, quantitative trait loci for BCA resistance have been reported (Cheung et al.,

2010; Crespo-Herrera et al., 2014) and a number of gene sequences proposed as candidate resistance factors (Delp et al., 2009; Mehrabi et al., 2014).

Identification of genes related to aphid resistance would be of high value in breeding programs. Conventional breeding develops new plant varieties expressing desirable traits, by crossing between parental lines, followed by selection of progeny. This process is time and labor consuming, since it might need several rounds of backcrossing, inbreeding and selection. Introducing resistance genes into cultivars could be facilitated with the use of transgenic approaches (Gepts, 2002). Additionally, identification of resistance genes would enable marker-assisted selection, instead of time consuming and labor- intensive insect bioassays.

More recent approaches in biotechnology aiming to produce aphid-re- sistant plants include the plant-mediated RNAi strategy, where transgenic plants expressing double stranded RNA (dsRNA) designed to knock-down insect target genes show increased protection against insect herbivory (Yu et al., 2016). A decline in GPA fecundity was reported due to plant-mediated

RNAi of salivary effectors MpC002, MpPIntO1 and MpPIntO2 (Pitino et al.,

2011; Pitino and Hogenhout, 2013). GPA showed also impaired reproduction on Arabidopsis plants expressing dsRNA of the GPA serine protease gene

38 MySP (Bhatia et al., 2012) and on tobacco plants expressing dsRNA against the gap gene hunchback (hb), which regulates insect axial patterning (Mao and

Zeng, 2014). Promising results are observed when approaches based on pes- ticide usage are combined with RNAi knock-down of aphid genes involved in insecticide detoxification. An example of this concept is the plant-mediated

RNAi silencing of the carboxylesterase CbEE4, which reduces the tolerance of

S. avenae to Phoxim insecticides (Xu et al., 2014). The new methods of plant- mediated RNAi insect control need however to face a few challenges such as the persistence of the RNAi effect, delivery and production of dsRNA by transgenic plants, potential off-target RNAi effect and potential development of aphid resistance towards RNAi plants (Yu et al., 2016)

The advantage of the transgenic breeding is that a particular desirable trait may be introduced, or the level of expression of such traits can be modified.

Application of such methods is however hindered in the European Union, by the Directive 2001/18/EC of the European Parliament and European Council on non-contained production and propagation of genetically modified organ- isms (GMOs). The new plant breeding techniques (NPBTs) however, such as the clustered regularly interspaced short palindrome repeats (CRISPR)/Cas9 nuclease technique or RNA-based DNA methylation, which differ from clas- sical breeding and conventional transgenic approaches, give opportunities for crop genetic improvement. Those methods escape one of the main arguments concerning hazard connected with GMOs, which is the uncontrolled integra- tion of recombinant DNA into the genome (Hartung and Schiemann, 2014).

39

40 Aim of the thesis

The aim of this thesis was to evaluate the candidate aphid-resistance genes,

CI2c and LOX2.2 in barley. To achieve this goal, the following investigations in the four papers were made:

I. A barley cDNA encoding the protease inhibitor CI2c was expressed in

Arabidopsis and the performance of the generalist GPA was studied in

control and transgenic plants, with constitutive or phloem-specific ex-

pression of the transgene.

II. Barley CI2c was overexpressed in barley and bioassays were carried

out with the specialist BCA and the generalist GPA. Aphid tests were

followed by analysis of aphid-induced plant responses related to de-

fense in plants.

III. The effects of the JA-responsive barley LOX2.2 gene overexpression

and silencing were studied with regard to consequences on expression

of other JA-responsive genes, changes in volatile emissions from trans-

genic lines, as well as BCA and GPA performance.

IV. The effects of volatile treatments on different cultivars of barley were

studied with regard to activation of plant defense responses on molec-

ular level, followed by bioassays with BCA on the treated plants.

BCA and GPA represent aphids with different host ranges and were exam- ined in detail in Papers II and III, to provide information regarding involve- ment of the studied resistance factors in host and non-host resistance.

41

42 Main approaches

Hypotheses for the thesis work

The work in Papers I, II and III is based on the hypothesis, that genes in- duced by aphids confer increased aphid resistance. It has been previously sug- gested that genes specifically induced in moderately aphid-resistant plants, and not in susceptible cultivars, may function as resistance factors in barley

(Delp et al., 2009).

For paper IV we hypothesized that treatments with volatiles would affect gene expression in barley in a way that aphid population growth would be reduced. The volatiles of special interest were those with known functions in plant defense: MeSA, in plants synthetized in the shikimic acid biosynthesis pathway, MeJA and (Z)-3-hexen-1-ol which both are produced via the LOX pathway. The genes of interest were those with suggested function in plant defense against aphids, of special interest were barley LOX genes, LOX2.1 and

LOX2.2, induced by BCA (Delp et al., 2009) and involved in production of JA or GLVs.

Putative resistance factors against aphids

With regard to the first hypothesis, the supporting evidence comes from the previously mentioned microarray study, where two moderately aphid- resistant and two susceptible barley genotypes were compared regarding

BCA-induced gene expression (Delp et al., 2009). One of the four genes in- duced by BCA in resistant genotypes and studied in Papers I and II was CI2c, which encodes a protease inhibitor. It belongs to a family of six chymotrypsin

43 inhibitor 2 (CI2) genes localized in the Mla (powdery mildew) resistance locus

(Wei et al., 2002). PIs are small molecular weight proteins involved in plant defense responses (Solomon et al., 1999; Hartl et al., 2011; Fluhr et al., 2012;

Jamal et al., 2012). They are mainly located in seeds or tubers and can be in- duced in leaves and roots (Habib and Fazili, 2007). They can be involved in regulation of endogenous plant protease activities or inhibition of exogenous pathogen- or insect-derived proteases during plant defense responses (Ryan,

1990; Carrillo et al., 2011a).

The second sequence suggested as a putative aphid resistance factor and studied in Paper III was a LOX2.2 gene. Plant LOXs are non-heme iron-con- taining enzymes which catalyse the addition of molecular oxygen to polyun- saturated fatty acids thus generating hydroperoxy fatty acids. This step initi- ates the synthesis of group of active compounds collectively called oxylipins.

Molecular oxygen can be incorporated to either the 9 or the 13 position of the

C18 chain of linoleic and linolenic acid by 9-LOXs and 13-LOXs respectively.

This results in synthesis of 9- or 13-hydroperoxy fatty acids, which are further channelled to different branches of the LOX pathway (Hildebrand, 1989;

Howe and Schilmiller, 2002; Porta and Rocha-Sosa, 2002). Another base of classification of LOXs is the localization. Type 1-LOXs do not have the chlo- roplast transit peptide and show high (>75%) sequence similarity and type 2-

LOXs have a chloroplast transit peptide and are characterized by lower

(>35%) sequence similarity to other LOXs (Shibata et al., 1994; Liavonchanka and Feussner, 2006; Andreou and Feussner, 2009). In general type 2-LOXs have been shown to exhibit 13-LOX activity and type 1-LOXs show 9-LOX activity, however there exist enzymes with no or dual position specificity

(Wang et al., 2008; Andreou and Feussner, 2009). Downstream of plant 13-

LOXs are two biosynthetic pathways: the AOS branch leading to the synthesis of JA and the HPL branch resulting in formation of GLVs (Matsui, 2006;

44 Mwenda and Matsui, 2014). End products from both pathways are important in plant defense against insects (Matsui, 2006; Scala et al., 2013; Wang and Wu,

2013; Naeem ul Hassan et al., 2015). Barley leaves contain at least three distinct forms of JA-responsive 13-LOX localized in chloroplasts: LOX2.1, LOX2.2 and

LOX2.3 (Vörös et al., 1998; Hause et al., 1999; Bachmann et al., 2002). LOX2.1 and LOX2.2 were shown to be induced by aphid infestation in barley (Delp et al., 2009). It has been shown that treatment of floating barley leaves with JA or JA-isoleucine conjugate caused accumulation of barley LOX transcripts, with LOX2.2 being induced prior to up-regulation of the LOX2.1 and LOX2.3 genes (Bachmann et al., 2002).

Induction of resistance

Regarding the second hypothesis, it has been previously shown for SA, JA, their methyl esters and GLVs in various grass species, that plant volatiles af- fect expression of defense genes in exposed plants (Vörös et al., 1998; Hause et al., 1999; Bachmann et al., 2002; Farag et al., 2005; Li et al., 2005; Xu et al.,

2005; Glinwood et al., 2007; Kakei et al., 2015; Yang et al., 2015). MeSA was of special interest in our study since it has been shown previously to negatively affect BCA preference and reduce BCA settling on barley, as well as infesta- tion on wheat by S. avenae and M. dirhodum (Pettersson et al., 1994). Results from studies on plants after treatment with MeJA or JA suggest that JA regu- lates effective defenses against phloem feeders in dicot plants (Gao et al., 2007;

Brunissen et al., 2010). For monocots, MeJA treatment was shown to reduce settling and negatively affect feeding behavior of BCA in wheat (Slesak et al.,

2001) and acceptance in no-choice tests in barley after exposures during day- light (Glinwood et al., 2007). With regard to (Z)-3-hexen-1-ol, the treatment has been shown to up-regulate genes related to the JA pathway in maize

(Farag et al., 2005; Xu et al., 2005; Engelberth et al., 2013). Fecundity of tobacco

45 aphid M. nicotianae was reduced on tobacco leaves exposed to C6 volatiles

(Hildebrand et al., 1993). More evidence for the role of GLVs in plant defense comes from a study where potato plants were depleted in HPL. Silenced lines showed lower emission levels of hexanal and 3-hexenal, which caused an in- crease in GPA fecundity when compared to wild type plants (Vancanneyt et al., 2001).

46 Comments on methodology

Plant transformation

The work in Papers I, II and III is based on the use of transformed Ara- bidopsis or barley plants. The protocol for Arabidopsis transformation using the floral dip method is well established and the transformation process sim- ple, with no need for tissue culture (Zhang et al., 2006). Barley transformation was carried out using plant immature embryos, which produces plants with low transgene copy numbers and supports stable expression (Bartlett et al.,

2008). In the above cases, the transformation was mediated by Agrobacterium tumefaciens, a gram-negative soil bacterium, which can incorporate a fragment of its Ti plasmid (T-DNA) into the genome of the plant host (Klee et al., 1987;

Ziemienowicz, 2014). The possibility to engineer plants to overexpress or down-regulate expression of selected genes is a tool of great value to study gene functions and gives possibilities for applications, i.e. to improve crop re- sistance to abiotic and biotic stresses.

A barley CI2c gene sequence was cloned and transformed into Arabidopsis ecotype Columbia (Col-0) under the control of the constitutive cauliflower mosaic virus (CaMV) 35S promoter (Paper I) and to barley under the control of the maize Ubi-1 promoter (Paper II). For monocots the CaMV 35S promoter is usable, although its activity has been reported as relatively low (Battraw and Hall, 1990). The maize Ubi-1 was shown to be comparatively stronger in monocots (Christensen et al., 1992). Additionally to constitutive transgene ex- pression, CI2c was transformed to Arabidopsis and to barley under the con- trol of rolC promoter from A. rhizogenes, with phloem-restricted expression reported in both dicot and monocot plants (Paper I and II) (Schmülling et al.,

47 1989; Saha et al., 2007; Srivastava et al., 2009). The use of a phloem-specific promoter was considered highly relevant for our studies. Expression of re- sistance factors in phloem tissues would provide valuable knowledge about the effects on aphids which successfully established feeding, narrowing the putative negative effects to antibiosis-based resistance.

For LOX2.2 overexpression and silencing, the Ubi-1 promoter was used

(Paper III). The down-regulation of LOX2.2 was accomplished by antisense transformation. Expression of transgenic antisense RNA was the first method of initiating post-transcriptional gene silencing (PTGS) in plants (Bourque,

1995). Antisense constructs consist of an inverted gene coding sequence. The homologous pairing of sense and antisense RNAs produces dsRNA, which was shown to be the initiator of silencing via production of small interfering

RNA (siRNA) (Di Serio et al., 2001). Antisense approach was used in Paper III for the reason that the sequence of the target gene to be silenced was known.

Other methods commonly used for PTGS are the heritable RNAi and

CRISPR/Cas9 techniques. Transient methods of gene silencing include virus- induced gene silencing (VIGS) or the use of synthetic antisense oligodeoxynu- cleotides (ODN). The above-mentioned methods differ regarding simplicity of application and level and specificity of silencing. Antisense RNA, RNAi,

VIGS and ODN are known to result only in partial silencing of the target gene.

The ODN method is not designed for intact plants and thus not suitable to study plant-aphid interactions (Gilchrist and Haughn, 2010; Dinc et al., 2011).

Moreover, siRNA-based approach can affect RNAs that are not perfectly com- plementary and considered off-targets. To secure a small number of targets with high sequence similarity, approaches based on micro RNA, short non- coding RNA, constitute alternatives (Schwab et al., 2006). In the CRISPR/Cas9 method the inactivation of genes is done by genome editing, which results in complete knock-out of the target gene. This method however needs

48 development for each species and is thereby more labor and time consuming than the above-mentioned techniques (Lawrenson et al., 2015; Lowder et al.,

2015). Another possible approach for gene silencing and used to investigate gene function is the targeting induced local lesions in genomes (TILLING) procedure. An advantage of TILLING is that it is a non-transgenic reverse ge- netic approach. It requires the production of mutants, amplification of a target gene and the recognition of the plants carrying a mutant allele (Talamè et al.,

2008).

PTGS has become a powerful method for studying gene function in plant interactions with phloem-feeding insects. For example, VIGS was used in wheat to silence resistance and susceptibility genes and study their function against D. noxia (van Eck et al., 2010; Anderson et al., 2014). Antisense-medi- ated transformation was used to silence the HPL gene in potato plants

(Vancanneyt et al., 2001) or OsHI-LOX gene in rice (Zhou et al., 2009), to study their involvement in resistance against aphids and the phloem-feeding brown plant hopper, respectively.

Selection and controls

Choosing a suitable control for studying the effects of transformation or treatment is important in order to avoid producing non-conclusive evidence.

For transgenic plants, few alternatives are possible. Wild type (WT) plants are used often (Vancanneyt et al., 2001; Aharoni et al., 2003; Ribeiro et al., 2006;

Wang et al., 2008; Zhou et al., 2009; Carrillo et al., 2011b), however those plants are not the most suitable controls to transgenic plants, since they do not expe- rience the same treatments (transformation process) and growth conditions.

Alternatives to WT plants are plants transformed with empty vector or azy- gous lines. Empty vector controls will show any possible effects of the vector itself on gene expression in the target organism. Azygous lines are those

49 which segregated as non-transformants and thus experienced the exact same conditions of transformation and selections as transformants, securing the same “seed history” effects, i.e., the same induced defenses carried from pre- vious generations (Pastor et al., 2013).

With the selection process, we aimed to identify homozygous transgenic plants with a single insertion of the transgene and a corresponding azygous control, coming from the same transformation event. This was accomplished by repeated selections on medium containing kanamycin (Paper I) or hygro- mycin (Papers II and III). For experiments with aphids we aimed to select ho- mozygous transgenic plants with varying level of transgene expression. Due to this process, the number of available lines for aphid bioassays was limited to five CI2c-overexpressing Arabidopsis lines in Paper I (two with constitutive and three with phloem-specific transgene expression), four CI2- transgenic lines in Paper II (two with constitutive and two with phloem-specific transgene expression), and two LOX2.2-transgenic lines in Paper III (one with overexpression and one with down-regulation of the gene). More transgenic lines were identified, but experimental restrictions led us to focus on a limited number.

Confirmation of transformation

To confirm transformation in Arabidopsis (Paper I) and barley (Paper II and III) we analyzed accumulation of the transgene using RT- qPCR. CI2c and

LOX2.2 gene expression was analyzed also in azygous lines, to detect levels of native expression of those sequences. In Paper I and II we confirmed trans- lation of the overexpressed CI2c into a functional protein by studying the in- hibitory effects of CI2c against added chymotrypsin in enzyme extracts in vitro. Similar protocols using fluorogenic protease substrates have been pre- viously used to identify plant protease inhibitors (Carrillo et al., 2011a;

50 Carrillo et al., 2011b). More direct methods to detect proteins are the com- monly used Western-blotting or ELISA techniques, however they require spe- cific antibodies (Bass et al., 2017). In Paper III we attempted to directly confirm the presence of LOX2.2 protein by Western-blotting. We tested an antibody directed against the wheat LOX2 protein (anti-LOX2, AS08 340, Agrisera), with 42% similarity to barley LOX2.2, since antibodies for barley LOX2.2 were not commercially available. Unfortunately, no barley LOX protein could be detected with this method, which could be explained firstly by low specificity of the wheat anti-LOX2 antibody towards antigens present on barley LOX and secondly due to low concentrations of LOX protein in protein extracts from barley. This problem could be solved by recombinant expression of the target protein, followed by purification and antibody preparation.

Two indirect methods were employed additionally in Paper III to confirm transformation. Firstly, we carried out the analysis of differential expression of JA-regulated genes in control, overexpressing and antisense LOX2.2 lines.

Secondly, the analysis of volatiles released from the control and overexpress- ing LOX2.2 lines, without and with BCA infestation was performed. The method of collecting plant volatiles by air-entrainment followed by GC/MS has been previously shown as suitable for barley (Glinwood and Pettersson,

2000b; Ninkovic et al., 2001; Glinwood et al., 2007). Our analysis was carried out after five days of aphid infestation. At this time point we observed effects of LOX2.2 overexpression on BCA and GPA fecundity (Figure 3 in Paper III), suggesting that during five days the activity of LOX2.2 was sufficient to affect aphids. It was also shown before, that collecting volatiles after

72 h of aphid infestation on barley was effective to induce responses in C. septempunctata, suggesting that changes in emitted volatiles occurred

(Ninkovic et al., 2001).

51 Plant phenotype

While assessing the effects of transformation and treatment on aphids, it is important to consider possible effects on plant phenotype. Adverse pheno- type affecting the nutritional value of plants could influence the interaction between plants and aphids and introduce false information into the research.

For example, plant PIs which constitute an important component of the pro- tein turnover machinery, can also be a part of the plant defense system against insects. It was shown that overexpression of the rice PI protein OCPI2 in Ara- bidopsis affected the plant’s phenotype by improving biomass accumulation and reproductive potential, features important when studying interaction be- tween plants and aphids (Tiwari et al., 2015). A negative effect on GPA in choice tests was observed in Arabidopsis plants expressing the FaNES1 gene from cultivated strawberry (Fragaria × anannasa cv Elsanta) encoding S-linal- ool/(3S)-E-nerolidol synthase. Those plants however showed retarded growth

(to a varying degree) when compared to control plants, possibly due to the reduction in the availability of substrates for other metabolites produced in the plastids, or the toxic effect of the high levels of linalool produced, compro- mising the results from aphid bioassays (Aharoni et al., 2003). A careful inves- tigation of the effects of transformation or treatments was carried out in our studies, with regards to effects on seed germination, flowering and leaf area

(Paper I), as well as plant growth and biomass accumulation (Papers II, III and

IV). It is also known that the interaction between plants and aphids differs depending on the growth stage of plants (de Vos et al., 2007; Louis et al., 2012), and thus in Paper I we used both bolting and non-bolting plants for tests with aphids.

52 Aphid performance

Antixenosis

Antixenosis describes the inability of a plant to serve as a host, and antixenosis-based resistance pressures the insect pest to select an alternate host plant (Smith, 2005a). To measure aphid host acceptance in barley, tests based on observations of aphid settling in no-choice and choice conditions have been carried out previously (Pettersson et al., 1996; Åhman et al., 2000;

Glinwood et al., 2007; Ninkovic and Åhman, 2009). Aphid settling is usually measured in time scales of hours; a period of 2 h has been reported to be suf- ficient for aphids to accept the host and reach the phloem (Prado and

Tjallingii, 1997). The preference behavior can be also measured in an olfac- tometer, where aphids are presented with different olfactory cues (Pettersson et al., 1994; Slesak et al., 2001; Glinwood et al., 2004), or with the EPG tech- nique, enabling monitoring aberrances in aphid feeding behavior, by record- ing salivation, probing and phloem ingestion activities (Tjallingii and Esch,

1993).

In our work we carried out tests monitoring aphid behavior in no-choice and choice conditions. Those tests provided us with information about GPA settling on control and transgenic Arabidopsis plants overexpressing barley

CI2c gene (Paper I). Aphid acceptance in no-choice conditions was measured after 2, 4 and 6 h (Figure 2 in Paper I). Aphid preference for control or trans- genic was measured after 24 h (Figure 3 in Paper I). Tests with GPA were ad- justed, since barley is a poor host for this species and GPA showed general avoidance towards barley (Papers II and III).

53 Antibiosis

Antibiosis resistance occurs when the negative effects of a resistant plant affect the biology of an insect pest attempting to use that plant as a host; the effects might be mild to lethal, chronic or acute (Smith, 2005b). Antibiosis re- sistance may affect different aspects of aphid biology such as growth, fecun- dity or longevity.

Analysis of aphid growth (i.e. weight) was previously used to identify aphid-resistant barley genotypes (Åhman et al., 2000; Ninkovic and Åhman,

2009). We measured the effects on fecundity in short-term tests, which pro- vided us with information on possible resistance factors in the peripheral plant tissues encountered by aphids at the time of penetration and probing

(Powell et al., 2006). Analyzing aphid numbers instead of weight of individual insects is an easier approach, introducing less error into the results. Another approach could be comparing the total weight of counted aphids between ex- perimental plants.

In our tests, we used both synchronous (Figure 5 in Paper I) and mixed populations of adult aphids (other tests in Papers I, II, III and IV). Using the uniform, calibrated populations in bioassays requires more time and labor in- vestment, they are however more suitable for time-scale experiments, since possible effects on aphids would be pronounced at the same time in all repli- cates, allowing to determine the moment of emergence and duration of those effects.

In the long term life span test, the reproduction of individual nymphs was monitored during their total life and since it follows aphids that have estab- lished feeding, it mainly gives indication of resistance factors present in the phloem (Nam et al., 2013; Will et al., 2013).

Both antixenosis- and antibiosis-based resistance can be measured not only on intact plants but also on excised plant leaves, leaf discs or with the use of

54 artificial diets (Forslund et al., 1998; Slesak et al., 2001; Bos et al., 2010; Pitino et al., 2011). For example, a functional leaf disc-based assay was used to de- termine whether aphid virulence was affected by overexpression of aphid candidate effectors (Bos et al., 2010). The results from assays performed on excised leaves should however be interpreted with caution as they may differ significantly from those obtained from intact plants (Schmelz et al., 2001). The same caution applies for tests using artificial diets, where in general, aphid performance is decreased when compared to plant-reared aphids (Douglas and van Emden, 2007). It should also be noted that results obtained under controlled laboratory conditions may not be observed at lager scales, i.e. in the field, where additional factors and interactions apply.

Another important aspect is the selection of rearing plants for aphid colo- nies, which should preferably be a different species than the plants used for bioassays with aphids. Insects may become adapted to a specific species or cultivar prior to experiment, which may affect the results. The aspect of rear- ing plants was relevant for consideration in our studies with BCA and GPA and addressed in Papers II and IV.

55

56 Main results

Effects of transformation or plant treatments with volatiles

Arabidopsis transformation with the barley CI2c sequence was confirmed at transcript and enzyme activity level (Figure 1 in Paper I). Transcript accu- mulation was much higher in lines with constitutive transgene expression driven by CaMV 35S promoter, than in phloem specific (ps) lines with expres- sion restricted to phloem tissue, controlled by rolC promoter (Figure 1 in Pa- per I). We did not observe any adverse effects on plant shoot phenotype at the age of the aphid tests (Table S1, Figures S2 and S3 in Paper I).

CI2c expression in barley was confirmed at gene expression and enzyme activity level (Figure 1 in Paper II), with no effects on plant phenotype (Figure

S2 in Paper II). CI2c accumulation was also studied in lines with expression restricted to phloem tissues (under control of rolC promoter). As in the case of

Arabidopsis transformants, transcript accumulation was to minimal levels when compared to lines with expression under Ubi-1 promoter and did not differ significantly from the expression in control lines (data not shown).

Thus, the transgenic status of barley plants expressing CI2c under the rolC promoter could not be confirmed. This issue could be solved by carrying out transformation using constructs containing a marker gene, encoding for ex- ample for green fluorescent protein (Stewart Jr, 2001).

In Paper III, LOX2.2 overexpression (OeLOX2.2 lines) was confirmed at transcript level as a significantly higher expression of LOX.2.2 and a number of other JA-regulated genes (Figure 7; Figures 4 and 5 in Paper III). LOX2.2 was expressed at low levels both in the control and antisense (antiLOX2.2)

57 lines. The antisense status was confirmed indirectly, by showing a lower ex- pression of three other JA-regulated genes, i.e. LOX2.1, THIO1567 and CI2c

(Figure 7), and no significant upregulation of LOX2.2 by aphids (Figures 4 and

5 in Paper III) in antiLOX2.2 line.

Figure 7. Relative transcript abundance of LOX2.2 and a selection of other jasmonic acid (JA)-regulated genes in OeLOX2.2 and antiLOX2.2 lines and their respective controls; bars indicate SE. Letters indicate significant differ- ence between lines for a particular gene (Kruskal-Wallis test at p < 0.05). Samples were from leaves of 14-day old plants (second leaf). Reference genes: Hsp70 and Tubulin.

No barley LOX protein could be detected with Western-blotting, for possible reasons discussed in section “Confirmation of transformation”. Another pos- sible way to indirectly detect active LOX2.2 protein would be to determine the concentration of JA and its conjugates in control and transgenic plants, assum- ing LOX2.2 involvement in JA biosynthesis (Zhou et al., 2009; Zhou et al.,

2014).

Similarly to LOX2.2, HPL was shown previously to be localized in plastids

(Froehlich, 2001; Koeduka et al., 2003), and thus possibly supplied by prod- ucts from the overexpressed LOX2.2 to produce GLVs. Analysis of volatiles

58 showed however no differences in the amounts or composition between un- infested or BCA-infested control and OeLOX2.2 line. LOX-derived volatiles however, especially (Z)-3-hexen-1-ol, were reduced after aphid infestation in the control line but not in the overexpressing line (Table 1 in Paper III). This is possibly due to LOX2.2 not being involved in production of GLVs, or to the fact that the amounts of other volatiles were at low concentrations and not detected in our analysis. We cannot exclude the possibility that we missed time points with differences in volatile composition. For a more comprehen- sive study, it would be interesting to include also the antisense lines and more time points, possibly within hours of the start of aphid infestation. It was shown in maize that the most pronounced changes in transcriptome and metabolome occur in the first few hours of R. maidis feeding. After four days, gene expression and metabolomic changes caused by aphid infestation were again similar to those of control plants (Tzin et al., 2015). In view of the fact that both MeJA and MeSA treatments were effective against aphids during daylight (Glinwood et al., 2007) and that HPL expression shows upregulation during the day in barley (unpublished results), it would also be of interest to use more plants and compare samples collected during the light period only.

In Paper IV, we studied the effects of volatile treatments on plants. There were no effects of MeSA, (Z)-3-hexen-1-ol or low concentrations (45 μM in

70% ethanol, 25 μl on filter paper in each cage) of MeJA on the plant pheno- type, however high concentrations (440 μM in 99% ethanol, 20 μl on filter pa- per in each cage) of MeJA during life span tests with daily treatments caused obvious changes in plant metabolism resulting in stunted growth (Figure 5 in

Paper IV). Reduction in plant fitness could be expected however, since con- stant induction of defense is a cost for the plant. MeJA treatments of tomato

(Solanum lycopersicum cv. Micro-Tom) seeds were shown previously to induce defenses against tomato fruit worm (Helicoverpa zea Boddie) larvae, however

59 a reduction in seedling germination was observed with higher (0.2, 0.4, 1 mM) doses of MeJA (Paudel et al., 2014).

Effects of transformation or treatments on aphid performance

CI2c overexpression

Effects on GPA

The no-choice test showed no difference between control and transgenic

Arabidopsis in the proportion of aphids settled (Figure 2 in Paper I). Results from choice tests showed a significant difference in aphid settling in one com- bination: with control and CI2c p:7 bolting plants (Figure 3 in Paper I). The results from short term fecundity tests showed lower numbers of aphids on both CI2c transgenic lines and on psCI2c p:3, as compared to on control plants

(Figure 4 in Paper I). More detailed investigation of the time scale of this effect was studied over a period of eight days with control and CI2c p:8 line. The difference in number of nymphs between the two lines was significant at day

5-7 (Figure 8).

60

Figure 9. Numbers of GPA adults and nymphs with time in choice tests. White bars represent control plants and black bars transgenic plants. Ten adult apterous GPAs were placed within small cages on control and CI2c 6- 4 plants in the same pot. After 24 h, the cages were opened and the aphids were free to move. Adults and nymphs were counted on a subset of plants each day after the release. Bars show average numbers ± SE. Dotted lines indicate trends. Asterisks indicate significant differences between control and the overexpressing CI2c line for a certain day at p≤0.05 (Wilcoxon matched pair test). n=6, except for day 5 where n=9.

Effects on BCA

The results from a fecundity test, where aphids were feeding on either con- trol or CI2c-overexpressing barley growing in separate pots showed no differ- ences in aphid numbers. Similarly, there were no negative effects of CI2c over- expression on aphid settling in the prolonged five-day choice test (Figure 2 in

Paper II) or life span (Figure S3 in Paper II).

With regard to defense against aphids, it has been shown that PIs can in- hibit proteases in aphid saliva or gut during probing, feeding establishment and digestion (Carolan et al., 2009; Carolan et al., 2011; Pyati et al., 2011; Furch et al., 2015; van Bel and Will, 2016), causing reduction in their growth and reproduction (Ryan, 1990). The negative effect on aphid biology was tested both in transgenic plants (Rahbé et al., 2003; Ribeiro et al., 2006; Alvarez-

Alfageme et al., 2011; Carrillo et al., 2011b) and in artificial diets (Tran et al.,

1997; Azzouz et al., 2005; Pyati et al., 2011). Whether the PIs affect aphids by

62 binding to digestive proteases is still under investigation, but evidence exists for the presence of digestive proteases in aphid’s gut, of the cysteine and ser- ine type, which are inhibited by plant PIs (Pyati et al., 2011; Carrillo et al.,

2011b). It is unlikely that CI2c targets aphid’s digestive proteases, since we did not observe any long-term negative effects on aphid biology. It is possible that the inhibitor act on functions related to aphid reproduction or interaction with bacterial symbionts. This idea was suggested by Rahbé et al. (2003), who demonstrated that protease inhibitor oryzacystatin OC-I expressed in oilseed rape (Brassica napus L.) and fed to GPA was localized within the bacteriocytes of the aphids (Rahbé et al., 2003). Another possibility is that CI2c regulates function of plant endogenous proteases involved in plant stress responses

(Solomon et al., 1999; Luo et al., 2009).

LOX2.2 overexpression and down-regulation

LOX2.2-overexpressing plants supported significantly lower BCA and

GPA numbers, and antiLOX2.2 plants higher aphid numbers, when compared to respective controls (Figure 10).

Figure 10. Aphid numbers on control and transgenic lines, five days after infestation with twenty adult apterous aphids. (a) Numbers of bird cherry- oat aphid (BCA), (b) numbers of green peach aphid (GPA). White and black bars indicate, respectively, control and transgenic lines. Bars indicate the average (±SE). n = 12 for BCA and n = 22 for GPA. Asterisks indicate signif- icant differences after aphid infestation between control and transgenic plants (t-test, * p ≤ 0.05; ** p ≤ 0.01)

63 Overexpression or down-regulation of LOX2.2 did not affect BCA or GPA settling in no-choice conditions (Figure S3 in Paper III), or the life span param- eters, when control and transgenic plants were compared (Table 2, Paper III).

The results point to a role of LOX2.2 in basic plant defense. The observed ef- fects on aphid fecundity may be caused by proteins, such as thionins, or pro- tein inhibitors, present at higher or lower levels respectively in the over-ex- pressing and antisense LOX2.2 transformants, and targeting aphid metabo- lism and/or reproductive functions (Delp et al., 2009; Mehrabi et al., 2014).

Volatile treatments

The results from aphid tests showed that treatment with MeJA had a neg- ative effect on BCA fecundity (Figure 11), whereas exposure to MeSA pro- longed the pre-reproductive period, as well as decreased the intrinsic popu- lation increase rate, without any effect on aphid reproduction (Table 2 in Pa- per IV, Figure 11).

64

Figure 11. Aphid numbers on control and volatile-treated barley cultivars seven days after infestation with ten adult apterous BCAs. White and black bars indicate respectively control and volatile-exposed plants. Volatiles used were: MeSA (three repetitions with 5-12 plants in each), MeJA (one experiment with indicated umber of plants for each cultivar) and (Z)-3- hexen-1-ol (three repetitions with 7-10 plants in each). Bars indicate means (±SE). Aphid numbers were normalized to average aphid numbers on con- trol plants as 100 (these numbers were for cv. Scandium 82 and 41, for cv. Barke 101 and 183 and for cv. Lina 102 and 100 in the experiment with MeSa and MeJA, respectively and 76 for cv. Lina in the experiment with (Z)-3- hexen-1-ol). Asterisks indicate significant differences after aphid infestation between control and volatile-treated plants (t-test, *p ≤ 0.05)

Treatment with (Z)-3-hexen-1-ol did not affect aphid performance on ex- posed plants, when compared to controls (Figure 11). Treatment with MeJA affected aphid fecundity on barley cultivars Scandium and Barke, but not on

Lina, which was used as a rearing plant and thus aphids became adapted (Fig- ure 11).

For volatile treatment, single compounds or complex blends may cause different responses in plants and aphids (Glinwood et al., 2007; Webster et al.,

2010; Webster, 2012). Important aspects to consider are the timing and length of the exposure, as well as the concentration of the compound. It was shown

65 that aphid acceptance was negatively affected by exposures to volatiles from undamaged plants of barley cv. Alva or thistle (Cirsium vulgare (Savi) Ten.), only when plants were exposed during darkness, whereas acceptance of plants treated with MeSA or MeJA was significantly reduced only after expo- sures during daylight. Interestingly, marker genes for treatment with MeSA and MeJA, BCI-4 and AOS2, respectively, were induced both during daylight and darkness (Glinwood et al., 2007). The effect of treatment on pathogen re- sistance was shown to depend on volatile dose and exposure time. Low con- centration of MeSA over 6 h did not reduce Pseudomonas syringae infections in lima bean (Phaseolus lunatus L.), whereas exposure to the same concentration over 24 h, or higher concentration for shorter period, significantly enhanced resistance (Girón-Calva et al., 2012). Analyzing the effects of volatile expo- sures on aphids, it is important to consider whether the observed effect comes from the volatile itself or from the changes in host plant status. For aphid be- havior, this aspect can be tested in an olfactometer, where aphid responses are measured after presenting the insects with cues coming either from inducing volatiles or exposed plants (Glinwood et al., 2004). Further, the issue can be solved by separating the treatments from the bioassays, both spatially and temporary. In our 7-day fecundity tests treated plants were kept in a growth chamber separate from controls. In the case of MeSA and MeJA, the inducing agent was removed, and the plants left overnight before introducing aphids.

After exposure with (Z)-3-hexen-1-ol, the aphids were introduced immedi- ately, since this treatment was shorter than with the above-mentioned com- pounds, and the effects observed one gene expression were mainly within the first few hours after the beginning of the exposure (Figure 2 in Paper IV).

66 Other effects of transformation or treatments

In Paper II we showed that a number of defense-related genes (for example

CI2c, TPPE and hordolisin), which were upregulated by aphids in control and moderately resistant barley genotypes (Hsp5 and DH28:4), were expressed at the same or lower levels in CI2c-overexpressing plants after aphid infestation

(Figure 12).

67 Figure 12. Transcript abun- dance in barley leaves with and without BCA. Primary leaves were infested during 48 h with twenty adult ap- terous BCA. White bars represent uninfested plants and black bars aphid-in- fested plants (±SE). The transcript abundance was calculated relative to the reference genes: Hsp70 and SF427 and normalized to uninfested control line set as 1.00. Different letters in- dicate significant differ- ences between the lines (Kruskal-Wallis test, p ≤ 0.05), asterisks indicate sig- nificant difference in one genotype with or without aphids (Mann-Whitney test, *p ≤ 0.05, **p ≤ 0.01). Six biological replicates (with two plants each) and three technical replicates.

68 Although CI2c was suppressed in aphid infested samples from transgenic plants, the chymotrypsin inhibitory activity was higher after aphid infestation when compared to un-infested plants, with a significant difference for line

CI2c 6-4 (Fig 5A in Paper II). In addition, the transgenic lines showed higher endogenous proteolytic activity than control or moderately-resistant lines

(Figure 5B in Paper II), possibly due to induction of other, inhibitor-insensitive proteases in transgenic plants. The indirect method to detect CI2c employed in our studies has one weakness however. In the tests we measured the com- bination of the inhibitory activity of CI2c as well as the activity of added chy- motrypsin and plant endogenous proteases. This makes it difficult to exactly identify the level of CI2c under conditions where both endogenous protease and CI2c were induced. Thus, although we could clearly confirm the overex- pression of CI2c in Arabidopsis and barley, the contribution of CI2c upon aphid infestation in barley was more difficult to compare between the geno- types (Figure 5A in Paper II). These problems can be solved by applying direct methods of protein detection such as Western-blotting or ELISA techniques

(Bass et al., 2017)

In Paper III, LOX2.2 overexpressing lines showed up-regulation of several

JA-regulated genes, such as LOX2.1, AOS1, LOX1 and CI2c (Figure 7). Anti- sense lines showed down-regulation of such genes (LOX2.1, THIO1567 and

CI2c) (Figure 7). Aphid infestation led to further up-regulation of the studied

JA-regulated genes (i.e. LOX2.1, AOS1, AOS2, CI2c), the expression pattern was shown similar for both studied aphid species (Figure 4 and 5 in Paper III).

This suggests that LOX2.2 plays a role in the activation of JA-mediated de- fense responses in plants.

In the study presented in Paper IV we found genes encoding CI2c, LOX2.2 and BCI-4 induced by exposure to MeSA. The effect was transient, observed at 6, 12 and 24 h after start of the treatment and decreasing with time (Figure

69 1 in Paper IV). Both MeJA and (Z)-3-hexen-1-ol treatments up-regulated

LOX2.1 and AOS2, the pattern of up-regulation by MeJA varied however from the induction observed for (Z)-3-hexen-1-ol, probably due to the different ex- perimental setup (Figures 2 and 3 in Paper IV). The analysis of induction of gene expression by the three volatiles was based on few selected genes, espe- cially in the case of MeJA treatment of Lina, and it would be of importance to complement the study by analyzing additional gene sequences and cultivars, as well as to carry out a careful investigation with regard to exposure time and volatile concentrations.

Differences in aphid performance

Between aphid species: BCA and GPA on barley

BCA and GPA performance on barley was studied in Papers II and III.

BCA, which is a specialist on grasses, did not show any type of avoidance or restlessness while settling on barley in no-choice (Figure S3 in Paper III) and choice conditions (Figure 2 in Paper II). In the choice tests, BCAs were released on a filter paper placed between two plants (Figure S1 in Paper II). GPA did not settle on barley when non-confined and thus aphids were added within a small cage mounted on the second leaf of each of a control and transgenic plant in the same pot (Figure S1 in Paper II). Only when forced to stay on plants, did the aphids stay, establish feeding and produced nymphs, with numbers in the same order of magnitude as in the case of BCA (60 vs 100)

(Paper II). However, when confined in small cages mounted around the base part of the plant, GPA did not survive on barley, whereas BCA performed equally well on the base as on the leaf blade (Figure 13). This indicates that compounds or macromolecules with antibiosis activities that provide re- sistance against GPA are present at higher relative levels in the base of barley

70 plants. In the choice tests with aphids, additional tests with alataes would be of value since it is known that the winged migrants respond differently to olfaction cues than do apterous forms (Webster, 2012), and are responsible for deciding whether to leave or land on plants, whereas wingless females are mainly responsible for sustaining colony growth.

50

40

30

20 Number of aphids Number of 10

0 GPA BCA

Figure 13. Numbers of GPAs and BCAs five days after application of ten adult aphids in the cage mounted at the base part of the plant. Error bars indicate SE, n=8.

During long-term life span tests, GPA showed a delayed onset of repro- duction and lower fecundity, compared to BCA (Table 2 in Papers II and III), however other aspects of aphid life history (length of reproduction and life span) were comparable between the two species (Table 2 and Figure S3 in Pa- per II, Table 2 in Paper III).

The production of nymphs per day had a broad optimum and declined with time for BCA, whereas for GPA it was low throughout the reproductive life, which suggests that barley, although it secures survival, is a poor host for

GPA (Figure S3 in Paper II).

71 Between host plants: GPA on Arabidopsis and barley

GPA performed better on Arabidopsis when compared to barley. The pre- reproductive period on Arabidopsis was shorter, 8.7 ± 0.2 (SE) on control plants (Table 1 in Paper I), when compared to 12.0 ± 0.6 (SE) on control barley

(Table 2 in Paper III). The number of nymphs per individual or per reproduc- tive day, as well as the intrinsic rate of population increase were higher on

Arabidopsis, when compared to aphids feeding on barley. The reproductive life and life span were also longer for GPA on Arabidopsis than barley (Table

1 in Paper I vs Table 2 in Paper II and III). Some similarities between the two hosts were however observed in 5-day fecundity tests. In non-confined con- ditions on Arabidopsis aphid numbers reached 55 ± 4.1 (Figure 4 in Paper I).

For barley, in experiments starting with twenty adult aphids, total aphid num- bers after five days in confined and non-confined conditions were 60 ± 5.2 (SE)

(Paper II) and 32.2 ± 4.3 (SE), respectively (Figure 10). Those results suggest that GPA avoids using barley as a host when an opportunity to escape exists.

Aphids survive and reproduce when forced to stay on the plants, which indi- cates that factors encountered during penetration rather than the phloem composition are responsible for the observed avoidance (Figure 2 in Paper II;

Figure 9). Another aspect when analyzing aphid responses to host plants is the selection of suitable rearing plant for aphid colonies. It was shown that

GPAs reared on potato did not survive to reproduce when added to barley, but survived when reared on barley (Davis and Radcliffe, 2008), or as in our studies, on kohlrabi (Paper II and III).

72 Conclusions

This thesis provides new knowledge important to explain plant-aphid in- teractions, by characterisation of plant resistance genes against aphids, as well as explanation of induced plant defense responses. The main results can be summarized as follows:

1. The barley gene CI2c when expressed in Arabidopsis resulted in a

gene product that was taken up by GPA from the tissue during

probing and transiently reduced GPA fecundity on transgenic Ar-

abidopsis. CI2c overexpression in barley did not affect the special-

ist BCA, but decreased avoidance of the specialist GPA due to sec-

ondary effects on defense induction.

2. LOX2.2 was shown to be involved in activation of other JA-regu-

lated genes related to defense. Overexpression of LOX2.2 de-

creased BCA and GPA performance on barley, suggesting a role of

the gene in basic plant defense, most likely due to increased syn-

thesis of JA or GLVs.

3. MeSA, MeJA and (Z)-3-hexen-1-ol treatment caused up-regulation

of defense-related genes. From the three volatiles tested, only the

exposure to MeJA reduced aphid populations. Exposure to high

concentrations of MeJA had negative consequences for plant

growth, but even the stunted plants served as good hosts for the

aphids in life span tests, i.e. the induced defense was only active

during probing and penetration and the phloem composition had

no negative effect on aphid performance.

73 The results presented in the thesis show that products of certain genes, such as CI2c and LOX2.2, can have a negative effect on aphid biology (Papers

I and III), as well as act indirectly by influencing other genes involved in plant defense responses (Papers II and III). The resistant plants used in our studies were shown in the background study to have a moderate resistance (up to

40% reduction in growth), when compared to susceptible cultivars. Knowing this, we could not expect each resistant gene identified in the resistant geno- types to provide a large difference, but rather contribute to the observed ef- fect. Testing gene induction in the resistant plants as well as volatile treat- ments proved to be a useful approach in finding answers regarding compati- ble and non-compatible plant-aphid interactions and provided new research questions.

74 Future perspectives

 The studies of CI2c indicated that the CI2c protein had a negative in-

fluence on GPA. This aspect could be further explored by purifying

the protein and adding it to artificial diets. There is a limited value of

this in barley breeding, since GPA is not an important pest on this

crop, but it may be of interest for improving other host crops by trans-

genic approaches. The NPBTs would be the methods of choice. It is

not however clear yet whether those techniques or organisms pro-

duced with the use of NPBTs are subjects of the current GMO legisla-

tion, and thus whether they will avoid the problems of GMO ap-

proval.

 More detailed analysis of the gene up-regulation by aphids in Papers

II and III including early time points would help to gain knowledge

about time-scale of aphid-induced responses. Knowing which re-

sponses are activated and when would help to detect additional fac-

tors acting in plant-aphid interactions and be useful in finding new

targets for inducing resistance.

 The effects of overexpression and silencing of LOX2.2 were analysed

with regard to gene-expression, volatile emissions and aphid perfor-

mance. We have shown that JA/GLV-related induction helps to pro-

tect against both BCA and GPA, however whether this level of re-

sistance will be sufficient to protect against aphids in the field would

need further investigation. Assuming that lower numbers of aphids

would be observed also in the field, one could expect decreasing num-

ber of offspring in each generation and fewer occasions where

75 thresholds for insecticide treatments would be reached. The next step

would thus be the evaluation of the LOX2.2-transgenic plants in field

trials.

 Analysis of volatiles would benefit from testing plants infested with

aphids for different time periods as well as including the LOX2.2 an-

tisense line into the investigation. An important question left unan-

swered in is whether GLVs or JA are inducing the other JA-regulated

genes. An experiment designed to answer this question would be to

carry out tests on transgenic plants with knocked-out HPL gene and

hindered production of GLVs.

 The effects of volatile treatments were studied on only a few selected

barley cultivars. To get a more detailed information of induced plant

defense responses, it would be of importance to include more de-

fense-related gene sequences in the analysis, extend the study by in-

cluding more cultivars and carry out a careful investigation with re-

gard to exposure time and volatile concentrations. This would help to

characterize the inducible factors acting in plant-aphid interaction

and assess how they depend on the strength (concentration, time of

exposure) of the inducing agent. Applications of volatiles could be a

way forward to less insecticide use, however there is still a gap in our

understanding of their action on a large scale in the field, where many

other factors are present, such as additional biotic and abiotic interac-

tions.

 Our suggestion of barley as a non-host for GPA needs further inves-

tigation. To provide more knowledge about this interaction, one

could study plant responses in compatible and incompatible interac-

tions on molecular level on a broader scale (microarray, RNA se-

quencing) and with regard to the responses activated in different

76 plant tissues. Non-host resistance is often described as multi-layered

and it would be of interest to answer questions on the localisation of

the resistance, for example by monitoring aphid feeding behaviour

using EPG.

 In Paper IV it has been suggested that the results from aphid experi-

ments depend on the possibility for aphid adaptation. This aspect is

not well studied in plant-aphid interactions and it would be interest-

ing to test for such possibilities. This could be done for example by

rearing aphids on different cultivars prior to experiments. Addition-

ally, this approach could provide additional insights to the barley-

GPA interaction, by analysing the possibility of GPA to adapt to feed-

ing on barley cultivars over several generations.

77

78 Acknowledgements

I would like to take this opportunity and thank everyone who made this

PhD thesis possible.

Firstly, I would like to express my endless gratitude to my supervisor Pro- fessor Lisbeth Jonsson. Thank you for giving me the opportunity to work with you and for sharing your knowledge. I am grateful for your guidance, pa- tience and constant support. You have helped me to develop as a scientist and as a person.

I am thankful to my co-supervisors, Docents Ulla Rasmussen and Robert

Glinwood for their assistance and encouragement during my studies and sup- port with writing the thesis. I am obliged to Professor Katharina Pawlowski and Docent Edouard Pesquet for careful revision of my thesis, valuable advice and constructive criticism.

I would like to express my appreciation to Professor Inger Åhman for help- ing with barley cultivation and collection of seeds used in my experiments.

Many thanks to the previous group members, to Lisa Beste for her scientific input, to Sara Mehrabi and Professor Afaf Hamada for sharing their experi- ence and making the “lab time” a “nice time”. I am thankful to Anneli Björk for all her ideas and incredible involvement in the experimental work during her Master’s thesis project. I wish to acknowledge all students involved in de- veloping and carrying out tests with aphids, especially François, Thomas, Ma- rie and Cyrielle. It was a pleasure to work with you in the lab.

I want to say thank you to my Uni friends, present and past. To Marco,

Eric, Sandra, Lina, Denis, Paul and Van. To my office neighbours from

79 Ecology, especially to Pil and Beate, the pub-party girls. Thank you all for your friendship and kindness.

I am grateful to my parents for all the effort they put in my development and education, to my sisters, Kasia, Marta and Magdalena, for their interest in my work, and to Elżbieta and Jens Losvik, for always believing in me.

Dziękuję!

And lastly, I would like to thank my dearest husband Johan. Without you,

I would never accomplish this much. Thank you for making me brave.

To all people who made it possible for me to be where I am in my life,

Thank you!

80 References

Aharoni A, Giri AP, Deuerlein S, Griepnik F, de Kogel W-J, Verstappen FWA, Verhoeven HA, Jogsma MA, Schwab W, Bouwmeester HJ (2003) Terpenoid metabolism in wild-type and transgenic Arabidopsis plants. Plant Cell 15: 2866– 2884 Ahmad S, Veyrat N, Gordon-Weeks R, Zhang Y, Martin J, Smart L, Glauser G, Erb M, Flors V, Frey M, et al (2011) Benzoxazinoid metabolites regulate innate immunity against aphids and fungi in maize. Plant Physiol 157: 317–327 Alvarez-Alfageme F, Maharramov J, Carrillo L, Vandenabeele S, Vercammen D, Van Breusegem F, Smagghe G (2011) Potential use of a serpin from Arabidopsis for pest control. PLoS One 6: e20278 Anderson VA, Haley SD, Peairs FB, van Eck L, Leach JE, Lapitan NL V. (2014) Virus- induced gene silencing suggests (1,3;1,4)-β-glucanase is a susceptibility factor in the compatible Russian wheat aphid–wheat interaction. Mol Plant Microbe Interact 27: 913–922 Andow DA (1991) Vegetational diversity and population response. Annu Rev Entomol 36: 561–586 Andreou A, Feussner I (2009) Lipoxygenases - Structure and reaction mechanism. Phytochemistry 70: 1504–1510 Aragandona VH, Chaman M, Cardemil L, Munoz O, Zúñiga GE, Corcurea LJ (2001) Ethylene production and peroxidase activity in aphid-infested barley. J Chem Ecol 27: 53–68 Avila CA, Arevalo-Soliz LM, Lorence A, Goggin FL (2013) Expression of α- DIOXYGENASE 1 in tomato and Arabidopsis contributes to plant defenses against aphids. Mol Plant Microbe Interact 26: 977–986 Awmack CS, Leather SR (2007) Growth and development. In HF van Emden, R Harrington, eds, Aphids as Crop pests. CABI Publishing, Wallingford, UK, pp 135–151 Azzouz H, Cherqui A, Campan EDM, Rahbé Y, Duport G, Jouanin L, Kaiser L, Giordanengo P (2005) Effects of plant protease inhibitors, oryzacystatin I and soybean Bowman-Birk inhibitor, on the aphid Macrosiphum euphorbiae (Homoptera, ) and its parasitoid Aphelinus abdominalis (Hymenoptera, Aphelinidae). J Insect Physiol 51: 75–86 Bachmann A, Hause B, Maucher H, Garbe E, Vörös K, Weichert H, Wasternack C, Feussner I (2002) Jasmonate-induced lipid peroxidation in barley leaves initiated by distinct 13-LOX forms of chloroplasts. Biol Chem 383: 1645–1657 Bartlett JG, Alves SC, Smedley M, Snape JW, Harwood WA (2008) High-throughput Agrobacterium-mediated barley transformation. Plant Methods 4: 22 Bass C, Denholm I, Williamson MS, Nauen R (2015) The global status of insect resistance to neonicotinoid insecticides. Pestic Biochem Physiol 121: 78–87 Bass C, Puinean AM, Zimmer CT, Denholm I, Field LM, Foster SP, Gutbrod O,

81 Nauen R, Slater R, Williamson MS (2014) The evolution of insecticide resistance in the peach potato aphid, Myzus persicae. Insect Biochem Mol Biol 51: 41–51 Bass JJ, Wilkinson DJ, Rankin D, Phillips BE, Szewczyk NJ, Smith K, Atherton PJ (2017) An overview of technical considerations for Western blotting applications to physiological research. Scand J Med Sci Sport 27: 4–25 Bate NJ, Rothstein SJ (1998) C6-volatiles derived from the lipoxygenase pathway induce a subset of defense-related genes. Plant J 16: 561–569 Battraw MJ, Hall TC (1990) Histochemical analysis of CaMV 35S promoter-β- glucuronidase gene expression in transgenic rice plants. Plant Mol Biol 15: 527– 538 Baumann P, Baumann L, Clark MA (1996) Levels of Buchnera aphidicola chaperonin GroEL during growth of the aphid Schizaphis graminum. Curr Microbiol 32: 279– 285 Baumann P, Baumann L, Lai C, Rouhbakhsh D, Moran NA, Clark MA (1995) Genetics, physiology, and evolutionary relationships of the genus Buchnera: intracellular symbionts of aphids. Annu Rewiev Microbiol 49: 55–94 van Bel AJE, Will T (2016) Functional evaluation of proteins in watery and gel saliva of aphids. Front Plant Sci 7: 1840 Bernasconi ML, Turlings TCJ, Ambrosetti L, Bassetti P, Dorn S (1998) Herbivore- induced emissions of maize volatiles repel the corn leaf aphid, Rhopalosiphum maidis. Entomol Exp Appl 87: 133–142 Berzonsky WA, Ding H, Haley SD, Harris MO, Lamb RJ, McKenzie RIH, Ohm HW, Patterson FL, Peairs FB, Porter DR, et al (2002) Breeding wheat for resistance to insects. In J Janick, ed, Plant Breeding Reviews. John Wiley & Sons, Inc, Oxford, UK, pp 221–296 Bettaibi K, Mezghani-Khémakhem M, Bouktila D, Charaabi K, Raboudi F, Makni H, Makni M (2016) A novel method for molecular targeting of insecticide resistance in Rhopalosiphum padi L. (Homoptera: Aphididae). Int J Pest Manag 62: 284–287 Beβer K, Jarosch B, Langen G, Kogel K-H (2000) Expression analysis of genes induced in barley after chemical activation reveals distinct disease resistance pathways. Mol Plant Pathol 1: 277–286 Bhatia V, Bhattacharya R, Uniyal PL, Singh R, Niranjan RS (2012) Host generated siRNAs attenuate expression of serine protease gene in Myzus persicae. PLoS One 7: e46343 Blackman RL, Eastop VF (2007) Taxonomix issues. In HF van Emden, R Harrington, eds, Aphids as Crop Pests. CABI, Wallingford, pp 1–29 Blackman RL, Eastop VF (2000) Aphids on the World’s Crops: An Identification and Information Guide, 2nd ed. Wiley & Sons, Chichester, UK Bohlmann H, Clausen S, Giese H, Hiller C, Schrader G, Apel K, Kiel BIDC (1988) Leaf-specific thionins of barley-a novel class of cell wall proteins toxic to plant- pathogenic fungi and possibly involved in the defence mechanism of plants. EMBO J 7: 1559–1565 Borrego E, Kolomiets M (2016) Synthesis and functions of jasmonates in maize. Plants 5: 41 Bos JIB, Prince D, Pitino M, Maffei ME, Win J, Hogenhout SA (2010) A functional genomics approach identifies candidate effectors from the aphid species Myzus

82 persicae (green peach aphid). PLoS Genet 6: e1001216 Bourque JE (1995) Antisense strategies for genetic manipulations in plants. Plant Sci 105: 125–149 Bridges M, Jones AME, Bones AM, Hodgson C, Cole R, Bartlet E, Wallsgrove R, Karapapa VK, Watts N, Rossiter JT (2002) Spatial organization of the glucosinolate – myrosinase system in brassica specialist aphids is similar to that of the host plant. Proceeding R Soc London B 269: 187–191 Brunissen L, Vincent C, Le Roux V, Giordanengo P (2010) Effects of systemic potato response to wounding and jasmonate on the aphid Macrosiphum euphorbiae (Sternorryncha: Aphididae). J Appl Entomol 134: 562–571 Bunge S, Wolters J, Apel K (1992) A comparison of leaf thionin sequences of barley cultivars and wild barley species. Mol Gen Genet 231: 460–468 Cao H-H, Wang S-H, Liu T-X (2014) Jasmonate- and salicylate-induced defenses in wheat affect host preference and probing behavior but not performance of the grain aphid, Sitobion avenae. Insect Sci 21: 47–55 Carolan JC, Caragea D, Reardon KT, Mutti NS, Dittmer N, Pappan K, Cui F, Castaneto M, Poulain J, Dossat C, et al (2011) Predicted effector molecules in the salivary secretome of the pea aphid (Acyrthosiphon pisum): a dual transcriptomic/proteomic approach. J Proteome Res 10: 1505–1518 Carolan JC, Fitzroy CIJ, Ashton PD, Douglas AE, Wilkinsonl TL (2009) The secreted salivary proteome of the pea aphid Acyrthosiphon pisum characterised by mass spectrometry. Proteomics 9: 2457–2467 Carrillo L, Herrero I, Cambra I, Sánchez-Monge R, Diaz I, Martinez M (2011a) Differential in vitro and in vivo effect of barley cysteine and serine protease inhibitors on phytopathogenic microorganisms. Plant Physiol Biochem 49: 1191– 1200 Carrillo L, Martinez M, Alvarez-Alfageme F, Castañera P, Smagghe G, Diaz I, Ortego F (2011b) A barley cysteine-proteinase inhibitor reduces the performance of two aphid species in artificial diets and transgenic Arabidopsis plants. Transgenic Res 20: 305–319 Chaman ME, Copaja S V., Argandoña VH (2003) Relationships between salicylic acid content, phenylalanine ammonia-lyase (PAL) activity, and resistance of barley to aphid infestation. J Agric Food Chem 51: 2227–2231 Charity J, Hughes P, Anderson MA, Bittisnich D, Whitecross M, Higgins T (2005) Pest and disease protection conferred by expression of barley β-hordothionin and Nicotiana alata proteinase inhibitor genes in transgenic tobacco. Funct Plant Biol 32: 35–44 Chaudhary R, Atamian HS, Shen Z, Briggs SP, Kaloshian I (2014) GroEL from the endosymbiont Buchnera aphidicola betrays the aphid by triggering plant defense. Proc Natl Acad Sci 111: 8919–8924 Cherqui A, Tjallingii WF (2000) Salivary proteins of aphids, a pilot study on identification, separation and immunolocalisation. J Insect Physiol 46: 1177–1186 Cheung WY, Di Giorgio L, Åhman I (2010) Mapping resistance to the bird cherry-oat aphid (Rhopalosiphum padi) in barley. Plant Breed 129: 637–646 Choi HW, Klessig DF (2016) DAMPs, MAMPs, and NAMPs in plant innate immunity. BMC Plant Biol 16: 232 Christensen AH, Sharrock RA, Quail PH, Plant UU, Expression G, Street B (1992)

83 Maize polyubiquitin genes: structure, thermal perturbation of expression and transcript splicing, and promoter activity following transfer to protoplasts by electroporation. Plant Mol Biol 18: 675–689 Cooper WR, Dillwith JW, Puterka GJ (2011) Comparisons of salivary proteins from five aphid (: Aphididae) species. Environ Entomol 40: 151–156 Cooper WR, Dillwith JW, Puterka GJ (2010) Salivary proteins of Russian wheat aphid (Hemiptera: Aphididae). Environ Entomol 39: 223–231 Crespo-Herrera LA, Akhunov E, Garkava-Gustavsson L, Jordan KW, Smith CM, Singh RP, Åhman I (2014) Mapping resistance to the bird cherry-oat aphid and the greenbug in wheat using sequence-based genotyping. Theor Appl Genet 127: 1963–1973 Von Dahl CC, Baldwin IT (2007) Deciphering the role of ethylene in plant-herbivore interactions. J Plant Growth Regul 26: 201–209 Davis J, Radcliffe EB (2008) Reproduction and feeding behavior of Myzus persicae on four cereals. J Econ Entomol 101: 9–16 Delp G, Gradin T, Åhman I, Jonsson L (2009) Microarray analysis of the interaction between the aphid Rhopalosiphum padi and host plants reveals both differences and similarities between susceptible and partially resistant barley lines. Mol Genet Genomics 281: 233–248 Dewar AM (2007) Chemical control. In HF van Emden, R Harrington, eds, Aphids as Crop Pests. CABI Publishing, Wallingford, UK, pp 391–422 Dinant S, Bonnemain J-L, Girousse C, Kehr J (2010) Phloem sap intricacy and interplay with aphid feeding. C R Biol 333: 504–515 Dinc E, Toth SZ, Schansker G, Ayaydin F, Kovacs L, Dudits D, Garab G, Bottka S (2011) Synthetic antisense oligodeoxynucleotides to transiently suppress different nucleus- and chloroplast-encoded proteins of higher plant chloroplasts. Plant Physiol 157: 1628–1641 Dixon AFG (1985a) Introduction. In Aphid Ecology. An optimization approach. Springer, Dordrecht, pp 6–7 Dixon AFG (1985b) Host specificity and speciacion. In Aphid Ecology. An optimization approach. Springer, Dordrecht, pp 27–58 Dixon AFG (1971) The life-cycle and host preferences of the bird cherry-oat aphid, Rhopalosiphum padi L., and their bearing on the theories of host alternation in aphids. Ann Appl Biol 68: 135–147 Dixon AFG (1985c) Structure of aphid populations. Annu Rev Entomol 30: 155–174 Dixon AFG, Glen DM (1971) Morph determination in the bird cherry–oat aphid, Rhopalosiphum padi L. Ann Appl Biol 68: 11–21 Dixon AFG, Kindlmann P, Leps J, Holman, J (1987) Why there are so few species of aphids, especially in the tropics. Am Nat 129: 580–592 Dogimont C, Chovelon V, Pauquet J, Boualem A, Bendahmane A (2014) The Vat locus encodes for a CC-NBS-LRR protein that confers resistance to Aphis gossypii infestation and A. gossypii-mediated virus resistance. Plant J 80: 993–1004 Douglas AE (2006) Phloem-sap feeding by : problems and solutions. J Exp Bot 57: 747–754 Douglas AE (1998) Nutritional interactions in insect-microbial symbioses: aphids and their symbiotic bacteria Buchnera. Annu Rev Entomol 43: 17–37 Douglas AE, van Emden HF (2007) Nutrition and symbiosis. In HF van Emden, R

84 Harrington, eds, Aphids as Crop Pests. CABI Publishing, Wallingford, UK, pp 115–134 Dreyer DL, Reese JC, Jones KC (1981) Aphid feeding deterrents in sorghum. Bioassay, isolation, and characterization. J Chem Ecol 7: 273–284 van Eck L, Schultz T, Leach JE, Scofield SR, Peairs FB, Botha AM, Lapitan NLV (2010) Virus-induced gene silencing of WRKY53 and an inducible phenylalanine ammonia-lyase in wheat reduces aphid resistance. Plant Biotechnol J 8: 1023–1032 Ellis C, Karafyllidis I, Turner JG (2002) Constitutive activation of jasmonate signaling in an Arabidopsis mutant correlates with enhanced resistance to Erysiphe cichoracearum, Pseudomonas syringae, and Myzus persicae. Mol Plant Microbe Interact 15: 1025–1030 Elzinga D a, Jander G (2013) The role of protein effectors in plant-aphid interactions. Curr Opin Plant Biol 16: 451–456 Elzinga DA, Vos M De, Jander G (2014) Suppression of plant defenses by a Myzus persicae (green peach aphid) salivary effector protein. Mol Plant Microbe Interact 27: 747–756 van Emden HF (1972) Aphids as phytochemists. In JB Harborne, ed, Phytochemical Ecology. Academi Press INC, London, pp 25–43 Escudero-Martinez C, Morris J, Hedley P, Bos J (2017) Barley transcriptome analyses upon interaction with different aphid species identify thionins contributing to resistance. Plant Cell Environ 2628–2643 FAOSTAT (2014) Food and Agriculture Organization of the United Nations (online). Available at: http://www.fao.org (accessed 22 August 2017). Finlay KJ, Luck JE (2011) Response of the bird cherry-oat aphid (Rhopalosiphum padi) to climate change in relation to its pest status, vectoring potential and function in a crop–vector–virus pathosystem. Agric Ecosyst Environ 144: 405–421 Fluhr R, Lampl N, Roberts TH (2012) Serpin protease inhibitors in plant biology. Physiol Plant 145: 95–102 Forslund K, Pettersson J, Ahmed E, Jonsson L (1998) Settling behaviour of Rhopalosiphum padi (L.) in relation to cyanogenic glycosides and gramine contents in barley. Acta Agric Scand Sect B - Soil Plant Sci 48: 107–112 Forslund K, Pettersson J, Bryngelsson T, Jonsson L (2000) Aphid infestation induces PR-proteins differently in barley susceptible or resistant to the bird cherry-oat aphid (Rhopalosiphum padi). Physiol Plant 110: 496–502 Foster SP, Devine G, Devonshire AL (2007) Insecticide resistance. In HF van Emden, R Harrington, eds, Aphids as Crop Pests. CABI Publishing, Wallingford, UK, pp 261–285 Francis F, Lognay G, Haubruge E (2004) Olfactory responses to aphid and host plant volatile releases: (E)-β-farnesene an effective kairomone for the predator Adalia bipunctata. J Chem Ecol 30: 741–755 Froehlich JE (2001) Tomato allene oxide synthase and fatty acid hydroperoxide lyase, two cytochrome P450s involved in oxylipin metabolism, are targeted to different membranes of chloroplast envelope. Plant Physiol 125: 306–317 Furch ACU, van Bel AJE, Will T (2015) Aphid salivary proteases are capable of degrading sieve-tube proteins. J Exp Bot 66: 533–539 Gao L-L, Anderson JP, Klingler JP, Nair RM, Edwards OR, Singh KB (2007) Involvement of the octadecanoid pathway in bluegreen aphid resistance in

85 Medicago truncatula. Mol Plant Microbe Interact 20: 82–93 Gepts P (2002) A comparison between crop domestication, classical plant breeding, and genetic engineering. Crop Sci 42: 1780–1790 Gilchrist E, Haughn G (2010) Reverse genetics techniques: engineering loss and gain of gene function in plants. Briefings Funct Genomics Proteomics 9: 103–110 Giordanengo P, Brunissen L, Rusterucci C, Vincent C, van Bel A, Dinant S, Girousse C, Faucher M, Bonnemain J-L (2010) Compatible plant-aphid interactions: how aphids manipulate plant responses. C R Biol 333: 516–523 Girón-Calva PS, Molina-Torres J, Heil M (2012) Volatile dose and exposure time impact perception in neighboring plants. J Chem Ecol 38: 226–228 Girvin J, Whitworth RJ, Rojas LMA, Smith CM (2017) Resistance of select winter wheat (Triticum aestivum) cultivars to Rhopalosiphum padi (Hemiptera: Aphididae). J Econ Entomol 110: 1886–1889 Glazebrook J (2005) Contrasting mechanisms of defense against biotrophic and necrotrophic pathogens. Annu Rev Phytopathol 43: 205–227 Glinwood R, Ahmed E, Qvarfordt E, Ninkovic V (2011a) Olfactory learning of plant genotypes by a polyphagous insect predator. Oecologia 166: 637–647 Glinwood R, Gradin T, Karpinska B, Ahmed E, Jonsson L, Ninkovic V (2007) Aphid acceptance of barley exposed to volatile phytochemicals differs between plants exposed in daylight and darkness. Plant Signal Behav 2: 321–326 Glinwood R, Ninkovic V, Pettersson J (2011b) Chemical interaction between undamaged plants-effects on herbivores and natural enemies. Phytochemistry 72: 1683–1689 Glinwood R, Ninkovic V, Pettersson J, Ahmed E (2004) Barley exposed to aerial allelopathy from thistles (Cirsium spp.) becomes less acceptable to aphids. Ecol Entomol 29: 188–195 Glinwood R, Pettersson J (2000a) Change in response of Rhopalosiphum padi spring migrants to the repellent winter host component methyl salicylate. Entomol Exp Appl 94: 325–330 Glinwood R, Pettersson J (2000b) Host choice and host leaving in Rhopalosiphum padi (Hemiptera: Aphididae) emigrants and repellency of aphid colonies on the winter host. Bull Entomol Res 90: 57–61 Goggin FL (2007) Plant-aphid interactions: molecular and ecological perspectives. Curr Opin Plant Biol 10: 399–408 Goławska S, Łukasik I, Leszczyński B (2008) Effect of alfalfa saponins and flavonoids on pea aphid. Entomol Exp Appl 128: 147–153 Gray S, Gildow FE (2003) Luteovirus-aphid interactions. Annu Rev Phytopathol 41: 539–566 Habib H, Fazili KM (2007) Plant protease inhibitors: a defense strategy in plants. Biotech Mol Biol Rev 2: 68–85 Hamada A, Fatehi J, Jonsson L (2017) Seed treatments with thiamine reduce the performance of generalist and specialist aphids on crop plants. Bull Entomol Res 5: 1–9 Harmel N, Létocart E, Cherqui A, Giordanengo P, Mazzucchelli G, Guillonneau F, De Pauw E, Haubruge E, Francis F (2008) Identification of aphid salivary proteins: A proteomic investigation of Myzus persicae. Insect Mol Biol 17: 165– 174

86 Hartl M, Giri AP, Kaur H, Baldwin IT (2011) The multiple functions of plant serine protease inhibitors. Plant Signal Behav 6: 1009–1011 Hartung F, Schiemann J (2014) Precise plant breeding using new genome editing techniques: Opportunities, safety and regulation in the EU. Plant J 78: 742–752 Hause B, Vörös K, Kogel K-H, Besser K, Wasternack C (1999) A jasmonate-responsive lipoxygenase of barley leaves is induced by plant activators but not by pathogens. J Plant Physiol 154: 459–462 Heil M (2008) Indirect defence via tritrophic interactions. New Phytol 178: 41–61 Heil M, Land WG (2014) Danger signals - damaged-self recognition across the tree of life. Front Plant Sci 5: 578 Hildebrand DF (1989) Lipoxygenases. Physiol Plant 76: 249–253 Hildebrand DF, Brown GC, Jackson DM, Hamilton-Kemp TR (1993) Effects of some leaf-emitted volatile compounds on aphid population increase. J Chem Ecol 19: 1875–1887 Hillocks RJ (2012) Farming with fewer pesticides: EU pesticide review and resulting challenges for UK agriculture. Crop Prot 31: 85–93 Hogenhout SA, Bos JIB (2011) Effector proteins that modulate plant-insect interactions. Curr Opin Plant Biol 14: 422–428 Howe G a, Jander G (2008) Plant immunity to insect herbivores. Annu Rev Plant Biol 59: 41–66 Howe GA, Schilmiller AL (2002) Oxylipin metabolism in response to stress. Curr Opin Plant Biol 5: 230–236 Irwin ME, Thresh J (1990) Epidemiology of barley yellow dwarf virus: a study in ecological complexity. Annu Rev Phytopathol 28: 393–424 Jamal F, Pandey PK, Singh D, Khan MY (2012) Serine protease inhibitors in plants: nature’s arsenal crafted for insect predators. Phytochem Rev 12: 1–34 Jaouannet M, Morris JA, Hedley PE, Bos JIB (2015) Characterization of Arabidopsis transcriptional responses to different aphid species reveals genes that contribute to host susceptibility and non-host resistance. PLoS Pathog 11: e1004918 Jaouannet M, Rodriguez PA, Thorpe P, Lenoir CJG, MacLeod R, Escudero-Martinez C, Bos JIB (2014) Plant immunity in plant-aphid interactions. Front Plant Sci 5: 663 Jones B, Lookhart G, Johnson D (1985) Improved separation and toxicity analysis methods for purothionins. Cereal Chem 62: 327–331 Jones JDG, Dangl JL (2006) The plant immune system. Nature 444: 323–329 Kachroo A, Lapchyk L, Fukushige H, Hildebrand D, Klessig D (2003a) Plastidial fatty acid signaling modulates salicylic acid – and jasmonic acid – mediated defense pathways in the Arabidopsis ssi2 mutant. Plant Cell 15: 2952–2965 Kachroo P, Kachroo A, Lapchyk L, Hildebrand D, Klessig DF (2003b) Restoration of defective cross talk in ssi2 mutants: role of salicylic acid, jasmonic acid, and fatty acids in SSI2-mediated signaling. Mol Plant Microbe Interact 16: 1022–1029 Kaloshian I, Walling LL (2016) Hemipteran and dipteran pests: Effectors and plant host immune regulators. J Integr Plant Biol 58: 350–361 Karley AJ, Mitchell C, Brookes C, Mcnicol J, Neill TO, Roberts H, Graham J, Johnson SN (2016) Exploiting physical defence traits for crop protection: leaf trichomes of Rubus idaeus have deterrent effects on spider mites but not aphids. Ann Appl Biol 168: 159–172

87 Kazana E, Pope TW, Tibbles L, Bridges M, Pickett JA, Bones AM, Powell G, Rossiter JT (2007) The cabbage aphid: a walking mustard oil bomb. Proceeding R Soc B 274: 2271–2277 Kelly PJ, Bones A, Rossiter JT (1998) Sub-cellular immunolocalization of the glucosinolate sinigrin in seedlings of Brassica juncea. Planta 206: 370–377 Kesarwani M, Yoo J, Dong X (2007) Genetic interactions of TGA transcription factors in the regulation of pathogenesis-related genes and disease resistance in Arabidopsis. Plant Physiol 144: 336–346 Kishimoto K, Matsui K, Ozawa R, Takabayashi J (2005) Volatile C6-aldehydes and Allo-ocimene activate defense genes and induce resistance against Botrytis cinerea in Arabidopsis thaliana. Plant Cell Physiol 46: 1093–1102 Klee H, Horsch R, Rogers S (1987) Agrobacterium-mediated plant transformation and its further applications to plant biology. Annu Rev Plant Physiol 38: 467–486 Koeduka T, Stumpe M, Matsui K, Kajiwara T, Feussner I (2003) Kinetics of barley FA hydroperoxide lyase are modulated by salts and detergents. Lipids 38: 1167– 1172 Kunkel BN, Brooks DM (2002) Cross talk between signaling pathways in pathogen defense. Curr Opin Plant Biol 5: 325–331 Kuśnierczyk A, Tran DH, Winge P, Jørstad TS, Reese JC, Troczyńska J, Bones AM (2011) Testing the importance of jasmonate signalling in induction of plant defences upon cabbage aphid (Brevicoryne brassicae) attack. BMC Genomics 12: 423 Kuśnierczyk A, Winge P, Jørstad TS, Troczyńska J, Rossiter JT, Bones AM (2008) Towards global understanding of plant defence against aphids-timing and dynamics of early Arabidopsis defence responses to cabbage aphid (Brevicoryne brassicae) attack. Plant Cell Environ 31: 1097–1115 Larsson KAE, Saheed S a, Gradin T, Delp G, Karpinska B, Botha CEJ, Jonsson L (2011) Differential regulation of 3-aminomethylindole/N-methyl-3- aminomethylindole N-methyltransferase and gramine in barley by both biotic and abiotic stress conditions. Plant Physiol Biochem 49: 96–102 Larsson KAE, Zetterlund I, Delp G, Jonsson L (2006) N-Methyltransferase involved in gramine biosynthesis in barley: Cloning and characterization. Phytochemistry 67: 2002–2008 Lawrenson T, Shorinola O, Stacey N, Li C, Østergaard L, Patron N, Uauy C, Harwood W (2015) Induction of targeted, heritable mutations in barley and Brassica oleracea using RNA-guided Cas9 nuclease. Genome Biol 16: 258 Lee S, Whitaker VM, Hutton SF (2016) Mini review: potential applications of non-host resistance for crop improvement. Front Plant Sci 7: 997 Leszczynski B, Dixon AFG (1990) Resistance of cereals to aphids: Interaction between hydroxamic acids and the aphid Sitobion avenae (Homoptera: Aphididae). J Appl Entomol 117: 21–30 Li J, Brader G, Kariola T, Tapio Palva E (2006) WRKY70 modulates the selection of signaling pathways in plant defense. Plant J 46: 477–491 Li J, Brader G, Palva T (2004) The WRKY70 transcription factor: a node of convergence for jasmonate-mediated and salicylate-mediated signals in plant defense. Plant Cell 16: 319–331 Li Y-F, Zhu R, Xu P (2005) Activation of the gene promoter of barley beta-1,3-glucanase

88 isoenzyme GIII is salicylic acid (SA)-dependent in transgenic rice plants. J Plant Res 118: 215–221 Li Y, Hill CB, Carlson SR, Diers BW, Hartman GL (2007) Soybean aphid resistance genes in the soybean cultivars Dowling and Jackson map to linkage group M. Mol Breed 19: 25–34 Liavonchanka A, Feussner I (2006) Lipoxygenases: Occurrence, functions and catalysis. J Plant Physiol 163: 348–357 Liu Y-B, McCreight JD (2006) Responses of Nasonovia ribisnigri (Homoptera: Aphididae) to susceptible and resistant lettuce. J Econ Entomol 99: 972–978 van Loon LC, Bakker PAHM, Pieterse CM (1998) Systemic resistance induced by rhizosphere bacteria. Annu Rev Phytopathol 36: 453–483 Louis J, Lorenc-Kukula K, Singh V, Reese J, Jander G, Shah J (2010) Antibiosis against the green peach aphid requires the Arabidopsis thaliana MYZUS PERSICAE-INDUCED LIPASE1 gene. Plant J 64: 800–811 Louis J, Shah J (2013) Arabidopsis thaliana-Myzus persicae interaction: shaping the understanding of plant defense against phloem-feeding aphids. Front Plant Sci 4: 213 Louis J, Shah J (2015) Plant defence against aphids: the PAD4 signalling nexus. J Exp Bot 66: 449–454 Louis J, Singh V, Shah J (2012) Arabidopsis thaliana — aphid interaction. Arab B 10: e0159 Lowder LG, Zhang D, Baltes NJ, Paul JW, Tang X, Zheng X, Voytas DF, Hsieh T-F, Zhang Y, Qi Y (2015) A CRISPR/Cas9 toolbox for multiplexed plant genome editing and transcriptional regulation. Plant Physiol 169: 971–985 Luo M, Wang Z, Li H, Xia KF, Cai Y, Xu ZF (2009) Overexpression of a weed (Solanum americanum) proteinase inhibitor in transgenic tobacco results in increased glandular trichome density and enhanced resistance to Helicoverpa armigera and Spodoptera litura. Int J Mol Sci 10: 1896–1910 Lyons R, Manners JM, Kazan K (2013) Jasmonate biosynthesis and signaling in monocots: a comparative overview. Plant Cell Rep 32: 815–827 Mantelin S, Bhattarai KK, Kaloshian I (2009) Ethylene contributes to potato aphid susceptibility in a compatible tomato host. New Phytol 183: 444–456 Mao J, Zeng F (2014) Plant-mediated RNAi of a gap gene-enhanced tobacco tolerance against the Myzus persicae. Transgenic Res 23: 145–152 Mathre DE (1997) Compendium of barley diseases. American Phytopathological Society, St. Paul, Minnesota Matsui K (2006) Green leaf volatiles: hydroperoxide lyase pathway of oxylipin metabolism. Curr Opin Plant Biol 9: 274–280 Matsui K, Koeduka T (2016) Green leaf volatiles in plant and response. Subcell Biochem 86: 427–443 Mayer KFX, Waugh R, Brown JWS, Schulman A, Langridge P, Platzer M, Fincher GB, Muehlbauer GJ, Sato K, Close TJ, et al (2012) A physical, genetic and functional sequence assembly of the barley genome. Nature 491: 711–716 Mehrabi S, Åhman I, Jonsson L (2014) Transcript abundance of resistance- and susceptibility-related genes in a barley breeding pedigree with partial resistance to the bird cherry-oat aphid (Rhopalosiphum padi L.). Euphytica 198: 211–222 Meihls LN, Handrick V, Glauser G, Barbier H, Kaur H, Haribal MM, Lipka AE,

89 Gershenzon J, Buckler ES, Erb M, et al (2013) Natural variation in maize aphid resistance is associated with 2,4-dihydroxy-7-methoxy-1,4-benzoxazin-3-one glucoside methyltransferase activity. Plant Cell 25: 2341–2355 Miles PW (1999) Aphid saliva. Biol Rev 74: 41–85 Miller HL, Neese PA, Ketring DL, Dillwith JW (1994) Involvement of ethylene in aphid infestation of barley. J Plant Growth Regul 13: 167–171 Miller WA, Rasochová L (1997) Barley yellow dwarf viruses. Annu Rev Phytopathol 35: 167–190 Mithöfer A, Boland W (2008) Recognition of Herbivory-Associated Molecular Patterns. Plant Physiol 146: 825–831 Mithöfer A, Boland W (2012) Plant defense against herbivores: chemical aspects. Annu Rev Plant Biol 63: 431–450 Moran PJ, Cheng Y, Cassell JL, Thompson GA (2002) Gene expression profiling of Arabidopsis thaliana in compatible plant-aphid interactions. Arch Insect Biochem Physiol 51: 182–203 Moran PJ, Thompson GA (2001) Molecular responses to aphid feeding in Arabidopsis in relation to plant defense pathways. Plant Physiol 125: 1074–1085 Moreno A, Garzo E, Fernandez-Mata G, Kassem M, Aranda MA, Fereres A (2011) Aphids secrete watery saliva into plant tissues from the onset of stylet penetration. Entomol Exp Appl 139: 145–153 Morkunas I, Mai VC, Gabryś B (2011) Phytohormonal signaling in plant responses to aphid feeding. Acta Physiol Plant 33: 2057–2073 Mosblech A, Feussner I, Heilmann I (2009) Oxylipins: Structurally diverse metabolites from fatty acid oxidation. Plant Physiol Biochem 47: 511–517 Mutti NS, Louis J, Pappan LK, Pappan K, Begum K, Chen M-S, Park Y, Dittmer N, Marshall J, Reese JC, et al (2008) A protein from the salivary glands of the pea aphid, Acyrthosiphon pisum, is essential in feeding on a host plant. Proc Natl Acad Sci 105: 9965–9969 Mutti NS, Park Y, Reese JC, Reeck GR (2006) RNAi knockdown of a salivary transcript leading to lethality in the pea aphid, Acyrthosiphon pisum. J Insect Sci 6: 38 Naeem ul Hassan M, Zainal Z, Ismail I, ul Hassan MN, Zainal Z, Ismail I (2015) Green leaf volatiles: biosynthesis, biological functions and their applications in biotechnology. Plant Biotechnol J 13: 727–739 Nam KJ, Powell G, Hardie J (2013) Does phloem-based resistance to aphid feeding affect host-plant acceptance for reproduction? Parturition of the pea aphid, Acyrthosiphon pisum, on two near-isogenic lines of Medicago truncatula. Bull Entomol Res 103: 683–689 Nicholson SJ, Hartson SD, Puterka GJ (2012) Proteomic analysis of secreted saliva from Russian Wheat Aphid (Diuraphis noxia Kurd.) biotypes that differ in virulence to wheat. J Proteomics 75: 2252–2268 Niemeyer HM (1988) Hydroxamic acids (4-hydroxy-1,4-benzoxazin-3-ones), defence chemicals in the Gramineae. Phytochemistry 27: 3349–3358 Ninkovic V, Al Abassi S, Pettersson J (2001) The influence of aphid-induced plant volatiles on ladybird beetle searching behavior. Biol Control 21: 191–195 Ninkovic V, Ahmed E, Glinwood R, Pettersson J (2003) Effects of two types of semiochemical on population development of the bird cherry oat aphid

90 Rhopalosiphum padi in a barley crop. Agric For Entomol 5: 27–33 Ninkovic V, Glinwood R, Pettersson J (2006) Communication between undamaged plants by volatiles: the role of allelobiosis. In F Baluška, S Mancuso, D Volkmann, eds, Communication in Plants: Neuronal Aspects of Plant Life. Springer, Berlin, pp 421–434 Ninkovic V, Olsson U, Pettersson J (2002) Mixing barley cultivars affects aphid host plant acceptance in field experiments. Entomol Exp Appl 102: 177–182 Ninkovic V, Åhman I (2009) Aphid acceptance of Hordeum genotypes is affected by plant volatile exposure and is correlated with aphid growth. Euphytica 169: 177– 185 Ordon F, Perovic D (2013) Virus resistance in barley. In R Varshney, R Tuberosa, eds, Translational Genomics for Crop Breeding: Volume 1 - Biotic Stress. Wiley-Blackwell, pp 63–75 Paré PW, Tumlinson JH (1999) Plant volatiles as a defense against insect herbivores. Plant Physiol 121: 325–331 Paudel S, Rajotte EG, Felton GW (2014) Benefits and costs of tomato seed treatment with plant defense elicitors for insect resistance. Arthropod Plant Interact 8: 539– 545 Paulitz TC, Steffenson BJ (2011) Biotic stress in barley: disease problems and solutions. In SE Ullrich, ed, Barley: production, improvement, and uses. Wiley-Blackwell, Ames, Iowa, pp 307–354 Pegadaraju V, Knepper C, Reese J, Shah J (2005) Premature leaf senescence modulated by the Arabidopsis PHYTOALEXIN DEFICIENT4 gene is associated with defense against the phloem-feeding green peach aphid. Plant Physiol 139: 1927–1934 Pelegrini B, Franco L (2005) Plant γ-thionins: Novel insights on the mechanism of action of a multi-functional class of defense proteins. Int J Biochem Cell Biol 37: 2239–2253 Pettersson J, Ninkovic V, Glinwood R, Abassi S Al, Birkett M, Pickett J, Wadhams L (2008) Chemical stimuli supporting foraging behaviour of Coccinella septempunctata L. (Coleoptera: Coccinellidae): volatiles and allelobiosis. Appl Entomol Zool 43: 315–321 Pettersson J, Pickett J, Pye B, Quiroz A, Smart L, Wadhams LJ, Woodcock C (1994) Winter host component reduces colonization by bird-cherry-oat aphid, Rhopalosiphum padi (L.) (Homoptera, Aphididae), and other aphids in cereal fields. J Chem Ecol 20: 2565–2574 Pettersson J, Quiroz A, Fahad AE (1996) Aphid antixenosis mediated by volatiles in cereals. Acta Agric Scand 46: 135–140 Pettersson J, Tjallingii WF, Hardie J (2007) Host-plant selection and feeding. In HF van Emden, R Harrington, eds, Aphids as Crop pests. CABI Publishing, Wallingford, UK, UK, pp 87–113 Pieterse CMJ, Van der Does D, Zamioudis C, Leon-Reyes A, Van Wees SCM (2012) Hormonal modulation of plant immunity. Annu Rev Cell Dev Biol 28: 489–521 Pitino M, Coleman A, Maffei M, Ridout CJ, Hogenhout SA (2011) Silencing of aphid genes by dsRNA feeding from plants. PLoS One 6: e25709 Pitino M, Hogenhout SA (2013) Aphid protein effectors promote aphid colonization

91 in a plant species-specific manner. Mol Plant Microbe Interact 26: 130–139 Plumb RT (2002) Viruses of Poaceae: a case history in plant pathology. Plant Pathol 51: 673–682 Porta H, Rocha-Sosa M (2002) Plant lipoxygenases. Physiological and molecular Features. Plant Physiol 130: 15–21 Powell G, Tosh CR, Hardie J (2006) Host plant selection by aphids: behavioral, evolutionary, and applied perspectives. Annu Rev Entomol 51: 309–330 Prado E, Tjallingii WF (1997) Effects of previous plant infestation on sieve element acceptance by two aphids. Entomol Exp Appl 82: 189–200 Prince DC, Drurey C, Zipfel C, Hogenhout SA (2014) The leucine-rich repeat receptor-like kinase BRASSINOSTEROID INSENSITIVE1-ASSOCIATED KINASE1 and the cytochrome P450 PHYTOALEXIN DEFICIENT3 contribute to innate immunity to aphids in Arabidopsis. Plant Physiol 164: 2207–2219 Pržulj N, Dragović S, Malešević M, Momčilovič V, Mladenov N (1998) Comparative performance of winter and spring malting barleys in semiarid growing conditions. Euphytica 101: 377–382 Pushpalatha HG, Sudisha J, Geetha NP, Amruthesh KN, Shekar Shetty H (2011) Thiamine seed treatment enhances LOX expression, promotes growth and induces downy mildew disease resistance in pearl millet. Biol Plant 55: 522–527 Pyati P, Bandani AR, Fitches E, Gatehouse JA (2011) Protein digestion in cereal aphids (Sitobion avenae) as a target for plant defence by endogenous proteinase inhibitors. J Insect Physiol 57: 881–891 Quisenberry S, Ni X (2007) Feeding injury. In HF van Emden, R Harrington, eds, Aphids as Crop Pests. CABI Publishing, Wallingford, UK, pp 331–352 Radhika V, Kost C, Bartram S, Heil M, Boland W (2008) Testing the optimal defence hypothesis for two indirect defences: extrafloral nectar and volatile organic compounds. Planta 228: 449–457 Rahbé Y, Deraison C, Bonadé-Bottino M, Girard C, Nardon C, Jouanin L (2003) Effects of the cysteine protease inhibitor oryzacystatin (OC-I) on different aphids and reduced performance of Myzus persicae on OC-I expressing transgenic oilseed rape. Plant Sci 164: 441–450 Rao SAK, Carolan JC, Wilkinson TL (2013) Proteomic profiling of cereal aphid saliva reveals both ubiquitous and adaptive secreted proteins. PLoS One 8: e57413 Renwick JAA (2002) The chemical world of crucivores: lures, treats and traps. Entomol Exp Appl 104: 35–42 Reymond P, Farmer EE (1998) Jasmonate and salicylate as global signals for defense gene expression. Curr Opin Plant Biol 1: 404–411 Ribeiro APO, Pereira EJG, Galvan TL, Picanco MC, Picoli EAT, Silva DJH, Fari MG, Otoni WC (2006) Effect of eggplant transformed with oryzacystatin gene on Myzus persicae and Macrosiphum euphorbiae. J Appl Entomol 130: 84–90 Rossi M, Goggin FL, Milligan SB, Kaloshian I, Ullman DE, Williamson VM (1998) The nematode resistance gene Mi of tomato confers resistance against the potato aphid. Proc Natl Acad Sci U S A 95: 9750–9754 Ryan CA (1990) Protease inhibitors in plants: genes for improving defenses against insects and pathogens. Annu Rev Phytopathol 28: 425–449 Sabri A, Vandermoten S, Leroy PD, Haubruge E, Hance T, Thonart P, De Pauw E, Francis F (2013) Proteomic investigation of aphid honeydew reveals an

92 unexpected diversity of proteins. PLoS One 8: e74656 Saha P, Chakraborti D, Sarkar A, Dutta I, Basu D, Das S (2007) Characterization of vascular-specific RSs1 and rolC promoters for their utilization in engineering plants to develop resistance against hemipteran insect pests. Planta 226: 429–442 Sato Y, Kudoh H (2015) Tests of associational defence provided by hairy plants for glabrous plants of Arabidopsis halleri subsp. gemmifera against insect herbivores. Ecol Entomol 40: 269–279 Scala A, Allmann S, Mirabella R, Haring MA, Schuurink RC (2013) Green leaf volatiles: a plant’s multifunctional weapon against herbivores and pathogens. Int J Mol Sci 14: 17781–17811 Schmelz EA, Alborn HT, Tumlinson JH (2001) The influence of intact-plant and excised-leaf bioassay designs on volicitin- and jasmonic acid-induced sesquiterpene volatile release in Zea mays. Planta 214: 171–179 Schmülling T, Schell J, Spena A (1989) Promoters of the rolA, B, and C genes of Agrobacterium rhizogenes are differentially regulated in transgenic plants. Plant Cell 1: 665–670 Schwab R, Ossowski S, Riester M, Warthmann N, Weigel D (2006) Highly specific gene silencing by artificial microRNAs in Arabidopsis. Plant Cell 18: 1121–1133 Di Serio F, Schö H, Iglesias A, Tarina C, Bouldoires E, Meins F, Phillips RL, Schob H, Iglesias A, Tarina C, et al (2001) Sense- and antisense-mediated gene silencing in tobacco is inhibited by the same viral suppressors and is associated with accumulation of small RNAs. Proc Natl Acad Sci 98: 6506–6510 Shibata D, Slusarenko A, Casey R, Bell E (1994) Lipoxygenases. Plant Mol Biol Report 12: 41–42 Shulaev V, Silverman P, Raskin I (1997) Airborne signalling by methyl salicylate in plant pathogen resistance. Nature 385: 718–721 Simon J-C, Martinez-Torres D, Latorre A, Moya A, Hebert PDN (1996) Molecular characterization of cyclic and obligate parthenogens in the aphid Rhopalosiphum padi (L.). Proc R Soc B 263: 481–486 Singh V, Louis J, Ayre BG, Reese JC, Shah J (2011) TREHALOSE PHOSPHATE SYNTHASE11-dependent trehalose metabolism promotes Arabidopsis thaliana defense against the phloem-feeding insect Myzus persicae. Plant J 67: 94–104 Slesak E, Slesak M, Gabrys B (2001) Effect of methyl jasmonate on hydroxamic acid content, protease activity, and bird cherry-oat aphid Rhopalosiphum padi (L.) probing behavior. J Chem Ecol 27: 2529–2543 Smith CM (2005a) Antixenosis - adverse effects of resistance on anthropod behavior. In Plant Resistance to . Springer, Dordrecht, The Netherlands., pp 19–63 Smith CM (2005b) Antibiosis - adverse effects of resistance on arthropod biology. In Plant Resistance to Arthropods. Springer, Dordrecht, The Netherlands., pp 65–99 Smith CM, Boyko E V (2007) The molecular bases of plant resistance and defense responses to aphid feeding: current status. Entomol Exp Appl 122: 1–16 Smith CM, Chuang W-P (2014) Plant resistance to aphid feeding: behavioral, physiological, genetic and molecular cues regulate aphid host selection and feeding. Pest Manag Sci 70: 528–540

93 Smith CM, Clement SL (2012) Molecular bases of plant resistance to arthropods. Annu Rev Entomol 57: 309–328 Solomon M, Belenghi B, Delledonne M, Menachem E, Levine A (1999) The involvement of cysteine proteases and protease inhibitor genes in the regulation of programmed cell death in plants. Plant Cell 11: 431–443 Spoel SH (2003) NPR1 modulates cross-talk between salicylate- and jasmonate- dependent defense pathways through a novel function in the cytosol. Plant Cell 15: 760–770 Srivastava AC, Ganesan S, Ismail IO, Ayre BG (2009) Effective carbon partitioning driven by exotic phloem-specific regulatory elements fused to the Arabidopsis thaliana AtSUC2. BMC Plant Biol 9: 7 Stec B (2006) Plant thionins – the structural perspective. Cell Mol Life Sci 63: 1370–1385 Stewart Jr CN (2001) The utility of green fluorescent protein in transgenic plants. Plant Cell Rep 20: 376–382 Strażyński P, Ruszkowska M (2015) The life cycle functional response of Rhopalosiphum padi (L.) to higher temperature: territorial expansion of permanent parthenogenetic development as a result of warmer weather conditions. J Plant Prot Res 55: 162–165 Talamè V, Bovina R, Sanguineti MC, Tuberosa R, Lundqvist U, Salvi S (2008) TILLMore, a resource for the discovery of chemically induced mutants in barley. Plant Biotechnol J 6: 477–485 Thaler JS (1999) Jasmonate-inducible plant defences cause increased parasitism of herbivores. Nature 399: 686–688 Thaler JS, Humphrey PT, Whiteman NK (2012) Evolution of jasmonate and salicylate signal crosstalk. Trends Plant Sci 17: 260–270 Thines B, Katsir L, Melotto M, Niu Y, Mandaokar A, Liu G, Nomura K, He SY, Howe GA, Browse J (2007) JAZ repressor proteins are targets of the SCFCOI1complex during jasmonate signalling. Nature 448: 661–665 Thompson GA, Goggin FL (2006) Transcriptomics and functional genomics of plant defence induction by phloem-feeding insects. J Exp Bot 57: 755–766 Tiwari LD, Mittal D, Chandra Mishra R, Grover A (2015) Constitutive over- expression of rice chymotrypsin protease inhibitor gene OCPI2 results in enhanced growth, salinity and osmotic stress tolerance of the transgenic Arabidopsis plants. Plant Physiol Biochem 92: 48–55 Tjallingii WF (2006) Salivary secretions by aphids interacting with proteins of phloem wound responses. J Exp Bot 57: 739–745 Tjallingii WF, Esch TH (1993) Fine structure of aphid stylet routes in plant tissues in correlation with EPG signals. Physiol Entomol 18: 317–328 Tran P, Cheesbrough T, Keickhefer R (1997) Plant proteinase inhibitors are potential anticereal aphid compounds. J Ecol Entomol 90: 1672–1677 Turlings TCJ, Bernasconi M, Bertossa R, Bigler F, Caloz G, Dorn S (1998) The induction of volatile emissions in maize by three herbivore species with different feeding habits: possible consequences for their natural enemies. Biol Control 11: 122–129 Tzin V, Fernandez-Pozo N, Richter A, Schmelz EA, Schoettner M, Schäfer M, Ahern KR, Meihls LN, Kaur H, Huffaker A, et al (2015) Dynamic maize responses to aphid feeding are revealed by a time series of transcriptomic and metabolomic

94 assays. Plant Physiol 169: 1727–1743 Ullrich SE (2011) Significance, adaptation, production, and trade of barley. In SE Ullrich, ed, Barley: production, improvement, and uses. Wiley-Blackwell, Ames, Iowa, pp 3–13 Urbanska A, Freddy Tjallingii W, Dixon AFG, Leszczynski B (1998) Phenol oxidising enzymes in the grain aphid’s saliva. Entomol Exp Appl 86: 197–203 Uzest M, Gargani D, Dombrovsky A, Cazevieille C, Cot D, Blanc S (2010) The “acrostyle”: A newly described anatomical structure in aphid stylets. Arthropod Struct Dev 39: 221–229 Vancanneyt G, Sanz C, Farmaki T, Paneque M, Ortego F, Castañera P, Sánchez- Serrano JJ (2001) Hydroperoxide lyase depletion in transgenic potato plants leads to an increase in aphid performance. Proc Natl Acad Sci 98: 8139–8144 Vandermoten S, Harmel N, Mazzucchelli G, De Pauw E, Haubruge E, Francis F (2014) Comparative analyses of salivary proteins from three aphid species. Insect Mol Biol 23: 67–77 Vanholme R, Demedts B, Morreel K, Ralph J, Boerjan W (2010) Lignin biosynthesis and structure. Plant Physiol 153: 895–905 Verheggen FJ, Arnaud L, Bartram S, Gohy M, Haubruge E (2008) Aphid and plant volatiles induce oviposition in an aphidophagous hoverfly. J Chem Ecol 34: 301– 307 Vidhyasekaran P (2015) Salicylic acid signaling in plant innate immunity. Plant Hor- mone Signaling Systems in Plant Innate Immunity. Signaling and Com- munication in Plants. Springer, Dordrecht, pp 27–122 Villada ES, González EG, López-Sesé AI, Castiel AF, Gómez-Guillamón ML (2009) Hypersensitive response to Aphis gossypii Glover in melon genotypes carrying the Vat gene. J Exp Bot 60: 3269–3277 de Vos M, Jander G (2010) Volatile communication in plant-aphid interactions. Curr Opin Plant Biol 13: 366–371 de Vos M, Jander G (2009) Myzus persicae (green peach aphid) salivary components induce defence responses in Arabidopsis thaliana. Plant, Cell Environ 32: 1548– 1560 de Vos M, Kim JH, Jander G (2007) Biochemistry and molecular biology of Arabidopsis-aphid interactions. BioEssays 29: 871–883 de Vos M, Oosten VR Van, Poecke RMP Van, van Pelt JA, Pozo MJ, Mueller MJ, Buchala AJ, Métraux J-P, Loon LC Van, Dicke M, et al (2005) Signal signature and transcriptome changes of Arabidopsis during pathogen and insect attack. Mol Plant Microbe Interact 18: 923–937 Vörös K, Feussner I, Kühn H, Lee J, Graner A, Löbler M, Parthier B, Wasternack C (1998) Characterization of a methyljasmonate-inducible lipoxygenase from barley (Hordeum vulgare cv. Salome) leaves. Eur J Biochem 251: 36–44 Walia H, Wilson C, Condamine P, Liu X, Ismail AM, Close TJ (2007) Large-scale expression profiling and physiological characterization of jasmonic acid- mediated adaptation of barley to salinity stress. Plant Cell Environ 30: 410–421 Walters DR, Ratsep J, Havis ND (2013) Controlling crop diseases using induced resistance: challenges for the future. J Exp Bot 64: 1263–1280 Wang R, Shen W, Liu L, Jiang L, Liu Y, Su N, Wan J (2008) A novel lipoxygenase gene

95 from developing rice seeds confers dual position specificity and responds to wounding and insect attack. Plant Mol Biol 66: 401–414 Wang W, Dai H, Zhang Y, Chandrasekar R, Luo L, Hiromasa Y, Sheng C, Peng G, Chen S, Tomich JM, et al (2015) Armet is an effector protein mediating aphid- plant interactions. FASEB J 29: 2032–2045 Wang Y, Sheng L, Zhang H, Du X, An C, Xia X, Chen F, Jiang J, Chen S (2017) CmMYB19 over-expression improves aphid tolerance in Chrysanthemum by promoting lignin synthesis. Int J Mol Sci 18: 619 War AR, Paulraj MG, Ahmad T, Ahad A, Hussain B, Ignacimuthu S, Sharma HC, War AR, Paulraj MG, Ahmad T, et al (2012) Mechanisms of plant defense against insect herbivores. Plant Signal Behav 7: 1306–1320 Wasternack C (2007) Jasmonates: an update on biosynthesis, signal transduction and action in plant stress response, growth and development. Ann Bot 100: 681–697 Wasternack C, Hause B (2013) Jasmonates: biosynthesis, perception, signal transduction and action in plant stress response, growth and development. An update to the 2007 review in Annals of Botany. Ann Bot 111: 1021–1058 Webster B (2012) The role of olfaction in aphid host location. Physiol Entomol 37: 10– 18 Webster B, Bruce T, Pickett J, Hardie J (2010) Volatiles functioning as host cues in a blend become nonhost cues when presented alone to the black bean aphid. Anim Behav 79: 451–457 Wei F, Wing RA, Wise RP (2002) Genome dynamics and evolution of the Mla (powdery mildew) resistance locus in barley. Plant Cell 14: 1903–1917 Westermann J-C, Craik DJ (2010) Plant peptide toxins from nonmarine environments. Compr Nat Prod II 5: 257–285 Wiktelius S, Weibull J, Pettersson J (1990) Aphid host plant ecology: the bird cherry– oat aphid as a model. In RK Campbell, RD Eikenbary, eds, Aphid-Plant Gen- otype Interactions. Elsevier, Amsterdam, pp 21–36 Will T, Furch ACU, Zimmermann MR (2013) How phloem-feeding insects face the challenge of phloem-located defenses. Front Plant Sci 4: 336 Will T, Kornemann SR, Furch ACU, Tjallingii WF, van Bel AJE (2009) Aphid watery saliva counteracts sieve-tube occlusion: a universal phenomenon? J Exp Biol 212: 3305–3312 Will T, Steckbauer K, Hardt M, van Bel AJE (2012) Aphid gel saliva: sheath structure, protein composition and secretory dependence on stylet-tip milieu. PLoS One 7: e46903 Williams IS, Dixon AFG (2007) Life cycle and polymorphism. In HF van Emden, R Harrington, eds, Aphids as Crop pests. CABI, Wallingford, UK, pp 69–85 Wójcicka A (2015) Surface waxes as a plant defense barrier towards grain aphid. Acta Biol Cracoviensia Ser Bot 57: 95–103 Wojtaszek P (1997) Mechanisms for the generation of reactive oxygen species in plant defence response. Acta Physiol Plant 19: 581–589 Woodhead S, Padgham DE, Bernays EA (1980) Insect feeding on different sorghum cultivars in relation to cyanide and phenolic acid content. Ann Appl Biol 95: 151– 157 Worrall D, Holroyd GH, Moore JP, Glowacz M, Croft P, Taylor JE, Paul ND, Roberts

96 MR (2011) Treating seeds with activators of plant defence generates long-lasting priming of resistance to pests and pathogens. New Phytol 193: 770–778 Wroblewski T, Piskurewicz U, Tomczak A, Ochoa O, Michelmore RW (2007) Silencing of the major family of NBS-LRR-encoding genes in lettuce results in the loss of multiple resistance specificities. Plant J 51: 803–818 Wu J, Baldwin IT (2010) New insights into plant responses to the attack from insect herbivores. Annu Rev Genet 44: 1–24 Xu L, Duan X, Lv Y, Zhang X, Nie Z, Xie C, Ni Z, Liang R (2014) Silencing of an aphid carboxylesterase gene by use of plant-mediated RNAi impairs Sitobion avenae tolerance of Phoxim insecticides. Transgenic Res 23: 389–396 Yu X-D, Liu Z-C, Huang S-L, Chen Z-Q, Sun Y-W, Duan P-F, Ma Y-Z, Xia L-Q (2016) RNAi-mediated plant protection against aphids. Pest Manag Sci 72: 1090–1098 Zhang X, Henriques R, Lin S-S, Niu Q-W, Chua N-H (2006) Agrobacterium-mediated transformation of Arabidopsis thaliana using the floral dip method. Nat Protoc 1: 641–646 Zhou G, Qi J, Ren N, Cheng J, Erb M, Mao B, Lou Y (2009) Silencing OsHI-LOX makes rice more susceptible to chewing herbivores, but enhances resistance to a phloem feeder. Plant J 60: 638–648 Zhou G, Ren N, Qi J, Lu J, Xiang C, Ju H, Cheng J, Lou Y (2014) The 9-lipoxygenase Osr9-LOX1 interacts with the 13-lipoxygenase-mediated pathway to regulate resistance to chewing and piercing-sucking herbivores in rice. Physiol Plant 152: 59–69 Ziemienowicz A (2014) Agrobacterium-mediated plant transformation: Factors, applications and recent advances. Biocatal Agric Biotechnol 3: 95–102 Zohary D, Hopf M (2000) Domestication of plants in the Old World: The origin and spread of cultivated plants in West Asia, Europe, and the Nile Valley, 3rd ed. Oxford University Press Zúñiga GE, Argandona V, Niemeyer H, Corcuera L (1983) Hydroxamic acid content in wild and cultivated Gramineae. Phytochemistry 22: 2665–2668 Zúñiga GE, Corcuera LJ (1986) Effect of gramine in the resistance of barley seedlings to the aphid Rhopalosiphum padi. Entomol Exp Appl 40: 259–262 Zúñiga GE, Salgado M, Corcuera LJ (1985) Role of indole alkaloid in the resistance of barley seedlings to aphids. Phytochemistry 24: 945–947 Zúñiga GE, Varanda EM, Corcuera LJ (1988) Effect of gramine on the feeding behavior of the aphids Schizaphis graminum and Rhopalosiphum padi. Entomol Exp Appl 47: 161–165 Åhman I (2009) Breeding for inducible resistance against insects – applied plant breeding aspects. In A Schmitt, B Mauch-Mani, N Birch, M Dicke, eds, Proceed- ings Working Group “Induced Resistance in Plants Against Insects and Diseases”. IOBC/WPRS, Heraklion, pp 121–130 Åhman I, Glinwood R, Ninkovic V (2010) The potential for modifying plant volatile composition to enhance resistance to arthropod pests. CAB Rev Perspect Agric Vet Sci Nutr Nat Resour 5: 1–10 Åhman I, Tuvesson S, Johansson M (2000) Does indole alkaloid gramine confer resistance in barley to aphid Rhopalosiphum padi? J Chem Ecol 26: 233–255

97